首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 378 毫秒
1.
The formation of IrB2, IrB1.35, IrB1.1 and IrB monoboride phases in the Ir–B ceramic nanopowder was confirmed during mechanochemical reaction between metallic Ir and elemental B powders. The Ir–B phases were analysed after 90 h of high energy ball milling and after annealing of the powder for 72 h at 1050°C in vacuo. The iridium monoboride (IrB) orthorhombic phase was synthesised experimentally for the first time and identified by powder X-ray diffraction. Additionally, the ReB2 type IrB2 hexagonal phase was also produced for the first time and identified by high resolution transmission electron microscope. Ir segregation along disordered domains of the boron lattice was found to occur during high temperature annealing. These nanodomains may have useful catalytic properties.  相似文献   

2.
Supported mononuclear iridium complexes with ethene ligands were prepared by the reaction of Ir(C2H4)2(acac) (acac is CH3COCHCOCH3) with highly dehydroxylated MgO. Characterization of the supported species by extended X-ray absorption fine structure (EXAFS) and infrared (IR) spectroscopies showed that the resultant supported organometallic species were Ir(C2H4)2, formed by the dissociation of the acac ligand from Ir(C2H4)2(acac) and bonding of the Ir(C2H4)2 species to the MgO surface. Direct evidence of the site-isolation of these mononuclear complexes was obtained by aberration-corrected scanning transmission electron microscopy (STEM); the images demonstrate the presence of the iridium complexes in the absence of any clusters. When the iridium complexes were probed with CO, the resulting IR spectra demonstrated the formation of Ir(CO)2 complexes on the MgO surface. The breadth of the νCO bands demonstrates a substantial variation in the metal–support bonding, consistent with the heterogeneity of the MgO surface; the STEM images are not sufficient to characterize this heterogeneity. The supported iridium complexes catalyzed ethene hydrogenation at room temperature and atmospheric pressure in a flow reactor, and EXAFS spectra indicated that the mononuclear iridium species remained intact. STEM images of the used catalyst confirmed that almost all of the iridium complexes remained intact, but this method was sensitive enough to detect a small degree of aggregation of the iridium on the support.  相似文献   

3.
We review recent advances in the spectroscopic properties of heteroleptic Ir(N^C)2(LX)-type iridium complexes, which are known as color-tuning materials. Most Ir(N^C)2(LX)-type Ir complexes give single emission, in accordance with Kasha’s rule. Dual emission, however, has been observed from a single Ir(N^C)2(LX) complex, depending on the choice of the N^C moiety and LX ligands. For example, Ir(dfppy)2(pq), Ir(ppy)2(dpq-3F), Ir(ppy)2(pq), and Ir(pq)2(tpy) (dfppy=2-(2,4-difluorophenyl)pyridine, pq=2-phenylquinoline, ppy=2-phenylpyridine, dpq-3F=2-(3-fluorophenyl)-4-phenylquinoline, tpy=2-p-tolylpyridine). Recently, triple emission was observed from Ir(ppy)2(BTZ)-type iridium complexes with two ppy ligands as (N^C)2 and one 2-(2-hydroxyphenyl)benzothiazole (BTZ) ligand, while quadruple emission from Ir(ppy)2Q-type iridium complexes with two ppy ligands as (N^C)2 and one quinolinolato (Q) ligand. These multiple emissions cover a spectral range from blue to red, leading to white emission. Of the four emission bands from Ir(ppy)2Q, the UV and violet emissions are attributed to the emission from the singlet states of IrQ and Ir(ppy), respectively, while the green and red emissions are attributed to emission from the triplet states of Ir(ppy) and IrQ. The appearance of the emission from each of the Ir(ppy) and IrQ (or Ir(BTZ)) components is understood by reduced Förster energy transfer between IrQ (or Ir(BTZ)) and Ir(ppy) due to an orientation factor of nearly zero, that is, due to orthogonality between the two ligand planes, while the appearance of both the fluorescence and phosphorescence bands from each of the ligands is understood by inefficient intersystem crossing from the upper singlet state to the lower triplet state.  相似文献   

4.
Pretreatment conditions for the activation of Ir/WO3–SiO2 for the selective catalytic reduction of NO by CO in the presence of excess O2 were studied. Sequential treatment involving calcination in the presence of O2 and H2O followed by reduction and then re-oxidation under mild conditions was found to effectively activate Ir/WO3–SiO2. Temperature-programmed desorption during calcination, X-ray diffraction, and temperature-programmed reduction by H2 revealed that calcination was necessary for oxidative removal of the NH3 ligands from the iridium precursor, that reduction produced metallic iridium and partially reduced tungsten oxide, and that re-oxidation produced tungsten oxide with low reducibility. Transmission electron microscopy revealed that Ir was supported on finely dispersed tungsten oxide and that the iridium particle size after the sequential activation was 1–1.5 nm.  相似文献   

5.
A mesoporous, silica‐supported, chiral iridium catalyst with a highly ordered dimensional‐hexagonal mesostructure was prepared by postgrafting the organometallic complex (1‐diphenylphosphino‐2‐triethylsilylethane)[(R,R)‐1,2‐diphenylethylenediamine]iridium chloride {IrCl[PPh2(CH2)2Si(OEt)3]2[(R,R)‐DPEN] (DPEN=1,2‐diphenylethylenediamine)} on SBA‐15 silica. During the asymmetric hydrogenation of various aromatic ketones under 40 atm of hydrogen, the mesoporous, silica‐supported, chiral iridium catalyst exhibited high catalytic activity (more than 95% conversions) and excellent enantioselectivity (up to more than 99% ee). The catalyst could be recovered easily and used repetitively seven times without significantly affecting the catalytic activity and the enantioselectivity.  相似文献   

6.
A series of half-sandwich structural iridium(III) phenanthroline (Phen) complexes with halide ions (Cl, Br, I) and pyridine leaving groups ([(η5-CpX)Ir(Phen)Z](PF6)n, Cpx: electron-rich cyclopentadienyl group, Z: leaving group) have been prepared. Target complexes, especially the Cpxbiph (biphenyl-substituted cyclopentadienyl)-based one, showed favourable anticancer activity against human lung cancer (A549) cells; the best one ( Ir8 ) was almost five times that of cisplatin under the same conditions. Compared with complexes involving halide ion leaving groups, the pyridine-based one did not display hydrolysis but effectively caused lysosomal damage, leading to accumulation in the cytosol, inducing an increase in the level of intracellular reactive oxygen species and apoptosis; this indicated an anticancer mechanism of oxidation. Additionally, these complexes could bind to serum albumin through a static quenching mechanism. The data highlight the potential value of half-sandwich iridium(III) phenanthroline complexes as anticancer drugs.  相似文献   

7.
A catalytic method employing the cationic iridium‐(Sc,Rp)‐DuanPhos [(1R,1′R,2S,2′S)‐2,2′‐di‐tert‐butyl‐2,2′,3,3‐tetrahydro‐1H,1′H‐1,1′‐biisophosphindole] complex and BARF {tetrakis[3,5‐bis(trifluoromethyl)phenyl]borate} counterion effectively catalyzes the enantioselective hydrogenation of acyclic N‐arylimines with high turnover numbers (up to 10,000 TON) and excellent enantioselectivities (up to 98% ee), achieving the practical synthesis of chiral secondary amines.  相似文献   

8.
The water–gas shift reaction occurs competitively to the main reaction of the Ir-catalysed methanol carbonylation process. To study the effect of seven factors including temperature, pressure, iridium, ruthenium, methyl iodide, methyl acetate and water concentrations on the formation of hydrogen and carbon dioxide as a result of the water–gas shift reaction and other side reactions in the carbonylation of methanol to acetic acid, the experimental design method combined with response surface methodology (RSM) was utilised. Central composite design at five levels (with α=1.63) was used to design experiments. A quadratic model that included the main and interaction effects of variables for H2 and CO2 formation was developed. For two responses, R2 was in reasonable agreement “Adj-R2”. Furthermore, statistical tests confirmed the accuracy and the precision of models developed in this research. For CO2 formation, pressure, iridium and methyl iodide concentrations and for H2 formation, water and iridium concentrations had the most pronounced effects. Optimum conditions to minimise H2 and CO2 and CH4 formation were determined as follows: temperature of 189 °C, pressure of 32.0 bar, iridium content of 859 ppm, ruthenium concentration of 528 ppm, methyl iodide content of 8.68 wt%, methyl acetate concentration of 23.9 wt% and water content of 6.49 wt%. Ultimately, an experiment at optimum conditions revealed satisfactory agreement between the experimental and predicted data.  相似文献   

9.
Barium hexaferrite BaFe12O19 and iridium-containing barium hexaferrites have been prepared by the citrates gel method. Their catalytic activity in methane combustion has been evaluated. BaFe12O19 is an efficient catalyst for this reaction, and the introduction of iridium in the hexaferrite structure does not improve this activity. Mössbauer spectroscopy suggests that a part of the iridium ions are incorporated in the hexaferrite structure, however in crystallographic sites where they cannot interact with the gas phase. Infrared study of CO adsorption reveals the presence of two types of iridium particles in the surface: small Ir particles, in strong interaction with the hexaferrite structure, and some larger Ir particles which were not incorporated into the lattice.  相似文献   

10.
Iridium(III) complexes of the type [Ir(η5‐C5Me5)Cl2{Ph2PCH2CH2CH2S(O)xPh‐κP}] (x=0–2; 1 – 3 ) and [Ir(η5‐C5Me5)Cl{Ph2PCH2CH2CH2S(O)xPh‐κPS}][PF6] (x=0–1; 4 and 5 ) with 3‐(diphenylphosphino)propyl phenyl sulfide, sulfoxide, and sulfone ligands Ph2PCH2CH2CH2S(O)xPh were designed, synthesized, and characterized fully, including X‐ray diffraction analyses for complexes 3 and 4 . In vitro studies against human thyroid carcinoma (8505C), submandibular carcinoma (A253), breast adenocarcinoma (MCF‐7), colon adenocarcinoma (SW480), and melanoma (518A2) cell lines provided evidence for the high biological potential of the neutral and cationic iridium(III) complexes. Neutral iridium(III) complex 5 proved to be the most active, with IC50 values up to about 0.1 μM , representing activities of up to one order of magnitude higher than cisplatin. Using 8505C cells, apoptosis was shown to be the main mechanism through which complex 5 exerts its tumoricidal action. The described iridium(III) complexes represent potential leads in the search for novel metal‐based anticancer agents.  相似文献   

11.
Bivalent ligands are potential tools to investigate the dimerisation of G‐protein‐coupled receptors. Based on the (R)‐argininamide BIBP 3226, a potent and selective neuropeptide Y Y1 receptor (Y1R) antagonist, we prepared a series of bivalent Y1R ligands with a wide range of linker lengths (8–36 atoms). Exploiting the high eudismic ratio (>1000) of the parent compound, we synthesised sets of R,R‐, R,S‐ and S,S‐configured bivalent ligands to gain insight into the “bridging” of two Y1Rs by simultaneous interaction with both binding sites of a putative receptor dimer. Except for the S,S isomers, the bivalent ligands are high‐affinity Y1R antagonists, as determined by Ca2+ assays on HEL cells and radioligand competition assays on human Y1R‐expressing SK‐N‐MC and MCF‐7 cells. Whereas the R,R enantiomers are most potent, no marked differences were observed relative to the corresponding meso forms. The difference between R,R and R,S diastereomers was most pronounced (about sixfold) in the case of the Y1R antagonist containing a spacer of 20 atoms in length. Among the R,R enantiomers, linker length and structural diversity had little effect on Y1R affinity. Although the bivalent ligands preferentially bind to the Y1R, the selectivity toward human Y2, Y4, and Y5 receptors was markedly lower than that of the monovalent argininamides. The results of this study neither support the presence of Y1R dimers nor the simultaneous occupation of both binding pockets by the twin compounds. However, as the interaction with Y1R dimers cannot be unequivocally ruled out, the preparation of a bivalent radioligand is suggested to determine the ligand–receptor stoichiometry. Aiming at such radiolabelled pharmacological tools, prototype twin compounds were synthesised, containing an N‐propionylated amino‐functionalised branched linker (Ki≥18 nM ), a tritiated form of which can be easily prepared.  相似文献   

12.
A variety of the ternary Hf–Ir–B phases formed via the reaction between iridium and hafnium diboride at elevated temperatures was found. The data on the phase and elemental composition, as well as crystal structure, obtained by powder and single-crystal X-ray diffraction, scanning electron microscopy/energy-dispersive X-ray spectrometer, and time-of-flight neutron diffraction analysis unambiguously confirm that HfIr3Bx solid solution, two known ternary borides (HfIr3B4, Hf2Ir5B2), as well as two novel ternary HfIr2.1B1.3 and HfIr5.7B2.7 phases, are formed at elevated temperatures. This result is fundamentally different from that previously obtained by us for the Hf–Ir–C system in which only one binary intermetallic compound, HfIr3, was produced. The measured Vickers microhardness for all the aforementioned ternary borides (13–19 GPa) allows us to consider them hard. The coefficients of thermal expansion of ternary borides were measured by in situ high-temperature X-ray analysis.  相似文献   

13.
A catalyst made in situ from the (cyclooctadiene)iridium chloride dimer, [Ir(COD)Cl]2, and the monodentate phosphoramidite ligand (S)‐PipPhos was used in the enantioselective hydrogenation of 2‐ and 2,6‐substituted quinoxalines. In the presence of piperidine hydrochloride as additive full conversions and enantioselectivities of up to 96% are obtained.  相似文献   

14.
An efficient multi‐gram scale synthesis protocol of a variety of P,N ligands is described. The synthesis is achieved in a two‐step reaction. First, the amine is deprotonated and subsequently the chlorophosphine is added to yield the corresponding P,N ligand. Deprotonation of the amine is normally achieved with n‐BuLi at low temperature, but for the preparation of ligands with a 2,2′‐dipyridylamino backbone and phosphines with a high steric demand KH has to be employed in combination with reaction temperatures of 110 °C for the salt metathesis step. The reaction of two equivalents of a selected P,N ligand with one equivalent of the iridium complex [IrCl(cod)]2 (cod=1,5‐cyclooctadiene) affords P,N ligand‐coordinated iridium complexes in quantitative yield. X‐Ray single crystal structure analysis of one of these complexes reveals a monomeric five‐coordinated structure in the solid state. The iridium complexes were used to form catalysts for the N‐alkylation of aromatic amines with alcohols. The catalyst system was optimized by studying 8 different P,N ligands, 9 different solvents and 14 different bases. Systematic variation of the substrate to base and the amine to alcohol ratios as well as the catalyst loading led to optimized catalytic reaction conditions. The substrate scope of the developed catalytic protocol was shown by synthesizing 20 different amines of which 12 could be obtained in isolated yields higher than 90%. A new efficient catalyst system for the selective monoalkylation of primary aromatic and heteroaromatic amines with primary aromatic, heteroaromatic as well as aliphatic alcohols has been established. The reaction proceeds with rather moderate catalyst loadings.  相似文献   

15.
A novel bipolar polymer host PC10CzPO, carrying hole‐transporting carbazole and electron‐transporting triphenylphosphine oxide units in the oxygen‐interrupted main chain, was synthesized and characterized. In addition to its excellent thermal stability and miscibility with phosphors, PC10CzPO is also reported to have a triplet energy (ET) as high as 2.83 eV due to oxygen‐interrupted π‐conjugation, ensuring that PC10CzPO can be used as a suitable host material. The PC10CzPO‐based phosphorescent devices were investigated and compared, while doping with typical blue phosphor {iridium(III)[bis(4,6‐difluorophenyl)pyridinato‐N, C2]picolinate, FIrpic)}, green phosphor {tris[2‐(p‐tolyl)pyridine]iridium(III), Ir(mppy)3}, and red phosphor [bis(1‐phenyl‐isoquinoline‐C2,N)(acetylacetonato)iridium(III), Ir(piq)2acac]. As a result, the FIrpic‐based blue devices showed better device performances than those of red and green devices, which was ascribed to more effective energy transfer. This indicates that the choice of proper host and dopant emitters to fabricate phosphorescent polymer light emitting diodes (PhPLED) is a simple and effective approach to optimize device performances. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 44461.  相似文献   

16.
The use of a chiral iridium catalyst generated in situ from the (cyclooctadiene)iridium chloride dimer, [Ir(COD)Cl]2, the P‐Phos ligand [4,4′‐bis(diphenylphosphino)‐2,2′,6,6′‐tetramethoxy‐3,3′‐bipyridine] and iodine (I2) for the asymmetric hydrogenation of 2,6‐substituted quinolines and trisubstituted pyridines [2‐substituted 7,8‐dihydroquinolin‐5(6H)‐one derivatives] is reported. The catalyst worked efficiently to hydrogenate a series of quinoline derivatives to provide chiral 1,2,3,4‐tetrahydroquinolines in high yields and up to 96% ee. The hydrogenation was carried out at high S/C (substrate to catalyst) ratios of 2000–50000, reaching up to 4000 h−1 TOF (turnover frequency) and up to 43000 TON (turnover number). The catalytic activity is found to be additive‐controlled. At low catalyst loadings, decreasing the amount of additive I2 was necessary to maintain the good conversion. The same catalyst system could also enantioselectively hydrogenate trisubstituted pyridines, affording the chiral hexahydroquinolinone derivatives in nearly quantitative yields and up to 99% ee. Interestingly, increasing the amount of I2 favored high reactivity and enantioselectivity in this case. The high efficacy and enantioselectivity enable the present catalyst system of high practical potential.  相似文献   

17.
A high throughput catalyst screening is presented employing an evolutionary approach. The method comprises the optimization of initial leads by subjecting the catalysts to iterative rounds of optimization, including structural elaboration of the ligands by creating new focused libraries. Highly modular supramolecular ligands, robotized synthesis combined by high throughput experimentation creates a platform for fast catalyst development. An illustrative example for the asymmetric hydrogenation of cyclic 2,3,3‐trimethyl‐3H‐indole using iridium catalysts is presented. The kinetic investigation of the best catalyst yields an unusual second order in iridium, first order in hydrogen and zeroth order in substrate. Under optimized reaction conditions a TOF of 100 mol mol−1 h−1 with 96% ee could be obtained with the best catalyst. A full catalyst screening and kinetic study was conducted within a three‐week time‐frame.  相似文献   

18.
The current research investigated N2O decompositions over the catalysts Ir/Fe-USY, Fe-USY and Ir-USY under various conditions, and found that a trace amount of iridium (0.1 wt%) incorporated into Fe-USY significantly enhanced N2O decomposition activity. The decomposition of N2O over this catalyst (Ir/Fe-USY-0.1%) was also partly assisted by NO present in the gas mixture, in contrast to the negative effect of NO over noble metal catalysts. Moreover, Ir/Fe-USY-0.1% can decompose more than 90% at 400 °C (i.e. the normal exhaust temperature) under simulated conditions of a typical nitric acid plant, e.g. 5000 ppm N2O, 5% O2, 700 ppm NO and 2% H2O in balance He, and such an activity can be kept for over 110 h under these strict conditions. The excellent properties of bimetallic Ir/Fe-USY-0.1% catalyst are presumably related to the good dispersion of Fe and Ir on the zeolite framework, the formation of framework Al–O–Fe species and the electronic synergy between the Ir and Fe sites. The reaction mechanism for N2O decomposition has been further discussed on the temperature-programmed desorption profiles of O2, N2 and NO2.  相似文献   

19.
A catalyst with an ultra high iridium load was prepared using a method involving multiple impregnations. The obtained iridium catalyst contained between 29 and 35 wt% of 2 nm-sized nanoparticles dispersed on a support such as reinforced alumina, bauxite and precipitated alumina. XAFS suggested a possible structural model of Ir4 surrounded by oxygen. The decomposition of hydrazine hydrate to its elements was used as a probe reaction. The results showed that a catalyst support with a high mechanical strength such as reinforced alumina and bauxite is essential for sustaining the decomposition reaction of hydrazine hydrate where there is a high degree of mechanical and thermal shock. The decomposition reaction of hydrazine monohydrate (N2H4 · H2O) proceeded rapidly to generate a COx-free hydrogen-rich gas through contact with the iridium catalyst at room temperature.  相似文献   

20.
Four iridium complexes with achiral phosphino‐oxazoline (PHOX) ligands were readily prepared in two steps starting from commercially available phenyloxazolines. The air‐stable complexes with tetrakis[3,5‐bis(trifluoromethyl)phenyl]borate (BArF) as counterion showed high reactivity in the hydrogenation of a range of tri‐ and tetrasubstituted olefins. The best results were obtained with an iridium complex ( 11 ) derived from a dicyclohexylphosphino‐oxazoline ligand bearing no additional substituents in the oxazoline ring. With several substrates, which gave only low conversion with the Crabtree catalyst, [Ir(Py)(PCy3)(COD)]PF6, full conversion was observed. The productivity of the Crabtree catalyst could be strongly increased by replacing the hexafluorophosphate anion with tetrakis[3,5‐bis(trifluoromethyl)phenyl]borate. In one case, in the hydrogenation of a tetraalkyl‐substituted CC bond, [Ir(Py)(PCy3)(COD)]BArF gave higher conversion than catalyst 11 . However, with several other substrates complex 11 proved to be superior.  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-23

京公网安备 11010802026262号