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.
The reaction in dichloromethane at room temperature of the polyspirophosphazene copolymer {[NP(O2C12H8)]0.7[NP(OC5H4N)2]0.3} n (1) (O2C12H8=2,2′dioxybiphenyl) with [Ru2(η 6-p-cymene)2Cl4] gives the polymer supported organometallic complex {[NP(O2C12H8)]0.7[NP(OC5H4N-Ru(η 6-p-cymene)Cl2)2]0.3} n (2). Similarly the chiral phosphazene {[NP(O2C20H12)]0.9[NP(OC5H4N)2]0.1} n (3) (O2C20H12=R-2,2′-dioxy,1′1′-binaphthyl) reacts with the appropriate amount of the Ru precursor to give the related polymeric complex {[NP(O2C20H12)]0.9[NP(OC5H4N)(OC5H4N-Ru(η 6-p-cymene)Cl2]0.1} n (4) as an orange solid. Carrying out the reaction of 3 with the Ru precursor in acetone at room temperature using 0.35 equivalents of Ru per pyridine site, gives the crosslinked yellow material with formula {[NP(O2C20H12)]0.9[NP(OC5H5N)2]0.1[Ru(η 6-p-cymene)Cl2]0.07]} n (5), that may contain cationic [Ru(η 6-p-cymene)Cl]+ units trapped in the rigid interior of a chiral network.  相似文献   

2.
The correlation between the surface energy and thermal stability of polymers plays an important role in engineering of plastic materials. In this work, microstructural characteristics of copolymers of poly(dimethyl siloxane) with benzyl methacrylate, ethyl methacrylate, and methyl acrylate are correlated with their surface energy and thermal stability. The poly(dimethyl siloxane) segments in the copolymer chains affected the hydrophobic behavior. The surface energy of the synthesized copolymers decreased by increasing segments of alkyl methacrylates. The thermal stability of copolymers suggesting that heat resistance of poly(dimethyl siloxane) copolymers used in this correlation can be improved by adjusting the units of alkyl methacrylates in copolymers.  相似文献   

3.
Dissociation of C60 from Ir(CO)(PPh3)2(Cl)(η2-C60) in a binary mixture of solvents (solvent1 and solvent2) produced non-equilibrium mixtures of Ir(CO)(PPh3)2(Cl)(solvent1) and Ir(CO)(PPh3)2(Cl)(solvent2). Once the solvated species were produced, they underwent a relative fast solvent exchange between them to produce an equilibrium mixture.  相似文献   

4.
Fulvic acid–poly(methylmethacrylate) graft copolymers were synthesized by surface-initiated atom transfer radical polymerization with fulvic acid. The result demonstrated that the hydrophobicity of fulvic acid–poly(methylmethacrylate) was improved after modification by surface-initiated atom transfer radical polymerization. Furthermore, poly(lactic acid)/fulvic acid–poly(methylmethacrylate) composites were prepared to improve the performances of poly(lactic acid) by blend melting. Compared to poly(lactic acid) with Xc of 5.38%, the Xc of poly(lactic acid)/fulvic acid–poly(methylmethacrylate) composites was 19.94%. Moreover, the impact strength of poly(lactic acid)/fulvic acid–poly(methylmethacrylate) composites was increased by 5.19% compared to poly(lactic acid). In all, this study provided an effective and feasible method for optimizing interface performance and enhancing the thermal stability of poly(lactic acid).  相似文献   

5.
To investigate the effect of metal ion type on the crystal structure and optical and thermal behaviors of coordination compounds, two homometal and one heterometal 2,2′-bipyridine complexes of Pb(II) and Cd(II) have been synthesized and characterized by elemental analysis, PXRD, FT-IR and single crystal X-ray diffraction. Crystal structure analysis of heterometal coordination polymer, [Pb2Cd(2,2′-bipy)4(NO3)6]n, displays the attendance of a centrosymmetric 1D coordination polymer that crystallizes in the triclinic system with the space group of \({\text{p}}_{1}^{ - }\). Thermal behavior of prepared coordination compounds was examined under air atmosphere by thermogravimetric analysis. The study of optical properties of compounds showed that metal ion type of coordination compounds is influential on their photophysical properties. Moreover, heterometal coordination polymer was doped into a PVK:PBD blend in two different concentrations as a light emitting material in the fabrication of two organic light-emitting diodes.  相似文献   

6.
Poly(1,2-phenylenedithiocarbamate) (PPDTC) was prepared by the reaction of 2-aminothiophenol with carbon disulfide followed by condensation through the removal of H2S gas. PPDTC was used as a ligand to prepare four poly(1,2-phenylenedithiocarbamate)–metal complexes of iron(II), cobalt(II), copper(II), and lead(II), by refluxing with the metal salts. The polymer and its metal complexes were investigated by elemental analyses, UV–visible and IR spectroscopy, inherent viscosity, and magnetic susceptibility. The DC electrical conductivity variation with the temperature in the range 298–498 K of PPDTC and its polymeric copper complex was measured. Both polymer and polymer metal complexes showed an increase in electrical conductivity with an increase in temperature: typical semiconductor behavior. The proposed structure of the complexes is (MLX2·mH2O) n .  相似文献   

7.
Allyl molybdenum complex [(η3-C3H5)(η5-C5H5)Mo(CO)2] (Mo-A) was examined as a new drier for solvent-borne alkyd binder. The drying activity of pure Mo-A and the of combinations of Mo-A with cobalt(II) 2-ethylhexanoate (Co-Nuodex) was examined using standard methods for measurements of the drying time and hardness development of the alkyd film. The detailed study of the drying process was studied by time-resolved infrared spectroscopy on ethyl linoleate as well as on thin coating of alkyd binder.  相似文献   

8.
Two new organometallic compounds with the general formula [Ru(η6-anethole)(en)(X)]PF6, where X = Br (5) or I (6) and en = ethylenediamine, were synthesized and fully characterized using standard techniques (NMR, MS and elemental analysis). Displacement of the proton and carbon aromatic chemical shifts of anethole to the upper field in both compounds indicated organometallic Ru(η6-anethole) bond formation, and the characteristic “piano-stool” structure of these types of compounds was confirmed by the single crystal X-ray diffraction of compound 5. Compounds 5 and 6 exhibited interesting cytotoxicity, especially toward the human colon tumor cell line HT-29, similar to the previously studied compound [Ru(η6-anethole)(en)(Cl)]PF6 (4). A slight tendency toward cytotoxicity could be observed by varying the halogen leaving group (IC50 value): 6 (10 ± 2 μM) > 5 (18 ± 2 μM) = 4 (18 ± 3 μM). Compound 6 showed 3-fold more biological activity in HT-29 cells than toward the non-tumor colon cell line CCD-841, making it an interesting candidate for further studies and drug development.  相似文献   

9.
Visible light curing of diglycidyl ether of bisphenol-A (DGEBA) epoxy oligomer and acrylate monomers photoinitiated by (η6-benzophenone)(η5-cyclopentadienyl) iron hexafluorophosphate (Fc-BP) under a halogen lamp were studied by near infrared spectroscopy. Fc-BP exhibited high efficiency in the radical photopolymerization of acrylate monomers, even without the presence of tertiary amines. Under the same light source, however, benzophenone did not show any photoinitiating ability. Fc-BP could also be used to photoinitiate the cationic polymerization of DGEBA. There was an obvious increase in the photopolymerization rate of DGEBA and a decrease in the induction period when benzoyl peroxide was used as a photosensitizer. The induction period at the beginning of DGEBA cationic polymerization was eliminated by introducing a certain amount of cycloaliphatic epoxy monomer ERL4221 as an active diluent. However, the final epoxy conversion was decreased when ERL4221 was used.  相似文献   

10.
The conversion of methane to higher hydrocarbons on single crystal Ru catalysts has been investigated using combined elevated-pressure kinetic measurements/surface science studies. The reaction consists of activation of methane on Ru(0001) and Ru(11¯20) surfaces to produce carbonaceous intermediates at temperatures between 350 and 700 K and rehydrogenation of these species to ethane and propane at 370 K. It is found that under the reaction conditions employed, the maximum yield in ethane/propane production occurs at 500 K on both surfaces. Influence of the hydrogenation temperature on the production of ethane and propane is also examined. On Ru(0001), the yields of ethane and propane maximize at = 400 K, whereas no maximum yield was observed on Ru(11 0) in the 300–500 K temperature range. Under optimum reaction conditions, hydrocarbon products consist of 16% ethane and 2% propane. High-resolution electron energy-loss spectroscopy (HREELS) has been used to identify various forms of hydrocarbonaceous intermediates following methane decomposition. An effort is made to relate the hydrocarbon intermediates identified by HREELS to the gas phase products observed in the elevated pressure experiments.  相似文献   

11.
Topics in Catalysis - Selective epoxidation of the (R) and (S) isomers of limonene by dioxomolybdenum(VI) complexes anchored covalently on TiO2 nanotubes using UV–Vis light and O2 as the...  相似文献   

12.
Reaction of compound Ru(O2CCF3)(CHCHtBu)(CO)(PPh3)2 with CO gives the η1-alkeneacyl complex Ru(O2CCF3)(OCCHCHtBu)(CO)(PPh3)2, which is in equilibrium with the dicarbonyl Ru(O2CCF3)(CHCHtBu)(CO)2(PPh3)2 derivative in CH2Cl2 solution. The η1-acyl form involves an η1-coordination of the O2CCF3 ligand, whereas the dicarbonyl form contains the carboxylate ligand η2-coordinated to the metal. The same mixture of carbonylated compounds can be obtained from the reaction of Ru(CHCHtBu)Cl(CO)2(PPh3)2 with Na[O2CCF3] in a CH2Cl2/MeOH solution. These reactions reveal the significance of ancillary bidentate ligands for the η-nature of the acyl–metal bond. The molecular structure of the complex Ru(O2CCF3)(OCCHCHtBu)(CO)(PPh3)2 was established by X-ray diffraction study of a monocrystal obtained from a CH2Cl2/MeOH solution of the mixture of carbonylated compounds.  相似文献   

13.
To obtain flexile poly(lactic acid)-based melt-blown nonwoven filtration material, poly(lactic acid)/poly(?-caprolactone) melt-blown nonwoven with various components were melt-spun by melt-blown processing in the Melt-blown Experiment Line. The 3 wt.% tributyl citrate to poly(?-caprolactone) was added in the composites as compatibilizer. The effect of poly(?-caprolactone) on the structure, morphology, mechanical and filtration properties of poly(lactic acid)/poly(?-caprolactone) melt-blown nonwoven was reported. Scanning electron microscopy micrographs revealed good dispersion of the additive in the fiber webs. The crystallinity of melt-blown webs with poly(?-caprolactone) was more than that of poly(lactic acid) alone. The tensile strength, ductility and air permeability of poly(lactic acid) melt-blown nonwovens were enhanced significantly. The input of poly(?-caprolactone) increased the diameter of fibers and decreased the filtration efficiency of poly(lactic acid)/poly(?-caprolactone) melt-blown nonwoven.  相似文献   

14.
The thermal decomposition of Ru3(CO)10(dppe) in refluxing benzene gives, in contrast to the pyrolysis of the dppm analogue, the tetranuclear cluster Ru4(μ-CO)(CO)944-C6H4)(η214-PCH2CH2PPh2) (1) along with Ru3(CO)9212-C6H5)(η312-PPhCH2CH2PPh2) (2). The single-crystal structure analysis of 1 reveals a square-planar tetraruthenium skeleton containing a η44-benzyne ligand as well as a η214-phosphinidene–phosphine ligand.  相似文献   

15.
Hydride ruthenium complexes, RuHCl(CO)(PPh3)2(L–L) 3 (L–L=bidentate ligand having nitrogen and oxygen) react with allenes to give Ru(η3-allyl)(CO)(PPh3)(L–L) complexes 5 in good yields via hydrometalation reaction. The complexes 5 have planar chirality at the η3-allyl ligand and central chirality at the Ru metal, and consist of one pair of enantiomers. Ligand substitution reaction of Ru(η3-allyl)Cl(CO)(PPh3)2 complexes 6 with bidentate ligands (L–L) also afford the complexes 5 which have the same stereochemistry as those formed by the hydrometalation reaction. The planar chirality is controlled by the central chirality at the Ru metal in both the formations of the complexes 5. The structure of 5a (L–L=N–N bidentate ligand) was determined by the X-ray crystal structure analysis.  相似文献   

16.
In order to obtain a better anhydrous precursor for various applications in materials science and catalysis, thermal dehydration reactions of Y(TFA)3(H2O)3 (TFA = trifluoroacetate) (A) were investigated. Thermal treatment of A at different temperatures under vacuum (5 × 10?2 mm) for several hours failed to give totally anhydrous yttrium trifluoroacetate (as indicated by IR). Two different complexes, a partially dehydrated [Y(μ,η11-TFA)3(THF)(H2O)]1∞·THF (1) and a partially hydrolyzed [Y43-OH)4(μ,η11-TFA)61-TFA)(η2-TFA)(THF)3(DMSO)(H2O)] · 6THF (2), were obtained with good and moderate yield, respectively, by crystallization of two different thermally treated batches of A from THF (or THF + DMSO) at room temperature. More efficient dehydration of A could be achieved at 200 °C in a furnace, the obtained anhydrous yttrium tris-trifluoroacetate giving Y(TFA)3(THF)2 (3) on crystallization from THF. All the products were characterized by elemental analyses, FT-IR and 1H NMR spectroscopy as well as thermo-gravimetric analysis. In addition, single crystal X-ray structures are reported for 1 and 2, which show either a terminal (η1 and η2) or bridging (μ,η11) bonding behavior of the TFA ligand.  相似文献   

17.
Pincer type aminomethylphosphine–Pd(II) complexes supported on multi-walled carbon nanotube (MWCNT) have been synthesized and characterized using X-Ray diffraction spectrometry, scanning electron microscopy, thermal analysis, energy dispersive X-Ray, Fourier transform infrared spectrometry, transmission electron microscopy (TEM) and ultraviolet–visible spectrometry techniques. The novel complexes were tried as catalysts in Heck C–C coupling reactions. The crystallite size and lattice strain of the MWCNT based compounds were calculated by the Scherrer’s equation. The optical parameters of the MWCNT based structures were analyzed and the band gap enhanced from 4.42 to 4.98 eV. Different solvents (toluene, 1,4-diooxane, DMF and NMP) and bases (Et3N, Na2CO3, NaOAc and K2CO3) were tried at different temperatures (80, 100 and 110 °C) in the cross-coupling of bromobenzene with styrene. The optimum yield was found in the presence of K2CO3, 110 °C in 1,4-dioxane solvent system.  相似文献   

18.
Treatment of Pd(PPh3)4 with Me2NC(S)Cl in dichloromethane at −20 °C produces the complex [Pd(PPh3)21-SCNMe2)(Cl)], 2. Variable temperature 1H and 31P{H} NMR experiments of complex 2 shows the dissociation of either the chloride or the triphenylphosphine ligand to form complex [Pd(PPh3)22-SCNMe2)][Cl], 3 or the dipalladium complex [Pd(PPh3)Cl]2(μ,η2-SCNMe2)2, 4. The reaction of complex 2 with NaPF6 affords complex [Pd(PPh3)22-SCNMe2)][PF6], 5. Complexes 2, 4, and 5 are characterized by X-ray diffraction analyses.  相似文献   

19.
[Ni(fiprdtc)2] (1), [Ni(fiprdtc)(PPh3)(NCS)] (2), [Ni(fiprdtc)(PPh3)2]ClO4 (3), [Zn(fiprdtc)2] (4), [Zn(fiprdtc)2(1,10-phen)] (5) and [Zn(fiprdtc)2(2,2′-bipy)] (6) (f iprdtc=N-furfuryl-N-isopropyldithio- carbamate, 1,10-phen=1,10-phenanthroline and 2,2′-bipy=2,2′-bipyridine) complexes were prepared and characterized by elemental analysis, electronic, IR and NMR spectra and the structure of 2 was determined by single-crystal X-ray crystallography. UV–Vis spectral data of 13 are consistent with the formation of square planar complexes. IR spectra of the complexes show the contribution of the thioureide form to the structure. A single-crystal X-ray structural analysis of 2 proved four-coordinated nickel in a distorted square planar arrangement with a S2PN donor set. Significant asymmetry in Ni–S bond distances was observed in [Ni? S1=2.1655(8); Ni? S2=2.2120(8) Å]. This observation clearly supports the less effective trans of SCN? over PPh3. The observed shielding in N13CS2 chemical shifts of heteroleptic nickel complexes 2 and 3 when compared with homoleptic nickel complex 1 indicates the effect of PPh3 on the mesomeric drift of electron density toward nickel through the thioureide C? N bond. The N13CS2 chemical shift of 5 and 6 are additionally deshielded compared with 4 owing to the increase in coordination number. Complexes were screened for in vitro antibacterial activity and significant activity has been found.  相似文献   

20.
The reaction of [Os4(CO)9(RCCR)(η6-C6H6)] (R=Me, Ph) with Me3NO in the presence of 1,3-cyclohexadiene or 1,4-cyclohexadiene yields the clusters [{Os4(CO)8(RCCR)(η6-C6H6)}2222-C6H8-1,3)] and [{Os4(CO)8(RCCR)(η6-C6H6)}2222-C6H8-1,4)], respectively. The molecular structure of [{Os4(CO)8(MeCCMe)(η6-C6H6)}2222-C6H8-1,3)] has been determined by single crystal X-ray diffraction, and represents the first example of two osmium cluster fragments linked by 1,3-cyclohexadiene through a μ222-bridge.  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