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 703 毫秒
1.
The polymerization of styrene at 60°C initiated by 2,2′-azobisisobutyronitrile (AIBN) was studied in N,N-dimethylformamide (DMF) in the presence of tris-(bathophenanthroline)iron(III) complex, [Fe(bathophen)3]3+. The complex was prepared in situ by mixing hexakis(N,N-dimethylformamide)iron(III) perchlorate with bathophen-anthroline (systematic IUPAC nomenclature: 4,7-diphenyl-1,10-phenanthroline) in the molar ratio of 1 : 3. The equilibrium constant for was 3.12×103 L3 mol−3. The transfer constant for bathophenanthroline was found to be 0.38 ± 0.01 for the styrene/DMF system at 60°C. Mean velocity constant at 60°C for interaction of polystyryl radical with [Fe(bathophen)3]3+ was 3.73× 104 L mol−1 s−1. © 1996 John Wiley & Sons, Inc.  相似文献   

2.
Biomimetic [Fe(pyridoxinato)2OH·(H2O)3] non-hem complex designated as [Fe(L)2OH·(H2O)3] was synthesized with basic solution of pyridoxine (Vitamin B6) and FeCl3 in methanol under reflux condition. It was then immobilized within the Al-MCM-41. Characterizations were carried out by powder X-ray diffraction, nitrogen adsorption desorption, FTIR and UV–Vis spectroscopy. It was found that pore volume, surface area, and pore diameter of Al-MCM-41 decreases after immobilization of iron complex. Density functional theory studies confirmed the experimental results of [Fe(L)2OH·(H2O)3] complex. Fe-complex/Al-MCM-41 was found to successfully catalyze the oxidation of cyclohexane, cyclooctane and adamantane using H2O2 as oxidant with 45–90 % conversion toward the corresponding alcohols and ketones is considerable.  相似文献   

3.
Natural terpenes, (−)-limonene and (+)-carvone, can be epoxidized by peroxyacetic acid (PAA) at room temperature if a dinuclear manganese(IV) complex with 1,4,7-trimethyl-1,4,7-triazacyclononane (L), [Mn2L2O3] [PF6]2, is used as a catalyst. The total yield of the epoxides based on the consumed olefins are 97 and 95%, respectively. A kinetic study of the dec-1-ene and cyclohexane oxygenations including the investigation of their simultaneous competitive oxidation was carried out. The olefin epoxidation rate does not depend on dec-1-ene concentration and the dec-1-ene concentration does not affect the rate of cyclohexane oxidation. The cyclohexane oxidation rate is proportional to the alkane concentration. The kinetic analysis led to the conclusion that two species X 1 and X 2 are generated in the system, and there is no their mutual interconversion. The rate equation for the dec-1-ene epoxidation was proposed: W = k +1[cat][PAA], where cat is the initial manganese complex or its derivative, and the constant was determined: k +1 = 3.5 mol−1 dm3 s−1. Species X 1 is apparently an effectively epoxidizing manganese peroxo derivative whereas species X 2 is an alkane hydroxylating manganese oxo complex.  相似文献   

4.
The photopolymerization of methyl methacrylate (MMA) sensitized by tris(2,2′-bipyridine)iron(III ) complex, [Fe(bpy)3]3+, was studied at 35°C in the presence of an electron donor, triethylamine (TEA) with UV radiation of wavelength 254nm. The initial rate of polymerization, Rp, shows a linear dependence on [MMA] with an exponential value of 1·18±0·04. Rp varies linearly with the square root of the photosensitizer concentration up to 2·00×10-4moll-1, and above this concentration, Rp decreases with the increase of photosensitizer concentration. The rate of polymerization is not affected by the concentration of the co-initiator, [TEA]. A suitable mechanism for the reaction is proposed to explain the kinetics of the reaction. © 1997 SCI.  相似文献   

5.
Solvent extraction of zirconium(IV) from acidic chloride solutions has been carried out with the thiosubstituted organophosphorus acids Cyanex 301 and Cyanex 302. The extraction follows an ion exchange mechanism: MO2+(aq) + 2 HA(org) ? MOA2(org) + 2 H+(aq), where, M = Zr(IV); HA = Cyanex 301 or Cyanex 302. The plots of log D (distribution ratio) vs log [HA], are linear with slopes of 2, indicating the association of two moles of extractant with the extracted metal species. The plots of log D vs log [H+] gave straight lines with a negative slope of 1.7 for Cyanex 301 and 1.8 for Cyanex 302, indicating the exchange of two moles of hydrogen ions for every mole of Zr(IV). Addition of sodium salts enhanced the extraction of metal. The stripping behavior of metal from the loaded organic (LO) with HCl and H2SO4 was studied. Increase of temperature during the extraction and the stripping stage increases the metal transfer, showing the process is exothermic. Mixed extractants, the extraction behavior of associated elements such as Hf(IV), Ti(IV), Al(III), Fe(III) and the IR spectra of the metal complexes were studied. Copyright © 2004 Society of Chemical Industry  相似文献   

6.
The complexation equilibria of iron(III) with 3-hydroxypicolinic acid (hypa) have been studied spectrophotometrically in 40% (v/v) ethanol—water medium at 25°C and an ionic strength of 0.1 M NaClO4. The equilibria occurring in solution were established and the basic characteristics of the complexes formed were determined. The spectral study of the reaction solutions containing equimolar concentrations or an excess of one component in the pH range 1.5 to 10.5 gave the evidence for the formation of [FeLH]2+, [FeL]+ and [FeL2]? complex species, depending upon the pH of the medium (LH2 symbolising the molecular form of hypa). The possible complex transitions that occur in solution were demonstrated using graphical and logarithmic analysis of the absorbance—pH graphs. A simple, rapid, sensitive and selective method for spectrophotometric determination of trace levels of iron(III) was proposed based on the formation of Fe(hypa)2 complex at pH 5.5 (λ = 440 nm, ? = 1.5 × 104 dm3 mol?1 cm?1). The interference of a large number of foreign ions was investigated. The method has been applied successfully to the analysis of multivitamin and mineral preparations containing iron.  相似文献   

7.
The kinetics of 2,2′-azobisisobutyronitrile (AIBN) initiated polymerization of styrene in N,N-dimethylformamide (DMF) at 60°C were investigated in the presence of dibromo(N,N-dimethylformamide)copper(II) complex. The complex was prepared in situ by mixing tetrakis (N,N-dimethylformamide)copper(II) perchlorate with LiBr in the molar ratio of 1 : 2. The equilibrium constant for [Cu(DMF)4]2+ + 2Br? ? Cu(DMF)2Br2 + 2DMF was calculated by the limiting logarithmic method as 1.80 × 103 L2 mol?2. The velocity constant at 60°C for the interaction of polystyryl radical with Cu(DMF)2Br2 is 7.46 × 104 L mol?1 s?1.  相似文献   

8.
A series of unsaturated polyester (UPR)‐co‐rare‐earth complex (REX) photoluminescence materials with red and green luminescence were fabricated. REXs with double bonds, including complex of europium (Eu3+) (methacrylic acid)3 and 1,10‐phenanthroline (Phen) [Eu(MAA)3Phen], and complex of terbium (Tb3+)(methacrylic acid)3 and Phen [Tb(MAA)3Phen], and UPR acted as functional monomers and the polymer matrix, respectively. Fourier transform infrared and UV absorption spectroscopy confirmed the chemical structure of the resulting UPR‐co‐REX according to the free‐radical polymerization mechanism. The study of fluorescence distribution by means of laser scanning confocal microscopy indicated that the REX materials were uniformly dispersed in the UPR matrix. The effects of the type and dosage of REX on the fluorescence intensity and stability were examined via fluorescence spectrometry. We found that the optical/physical properties of the REX were improved by UPR molecular skeleton structures. The fluorescence intensity increased with increasing use of the REX and reached a maximum value when the REX content was 12 wt %. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 45253.  相似文献   

9.
BACKGROUND: Selenium removal from aqueous solutions can be a significant industrial problem, particularly in the metallurgical industry. In order to evaluate new reducing agents for this application, the reduction of selenious acid (H2SeO3) species with stannous ions (Sn2+) from weakly acidic sulfate solutions containing 300 mg L?1 of selenium at 23 °C was studied. RESULTS: At initial pH values < 1.3 and molar ratio ≥ 2, less than 0.5 µg L?1 of selenium(IV) remained in solution after reduction. The reductive precipitation reaction started as soon as the stannous ions were added to the selenium‐bearing solution and was completed in less than 5 min. The reaction products, characterized using X‐ray diffraction, electron microscopy, particle and surface area measurements, X‐ray photoelectron spectroscopy and chemical analysis, were composed of approximately equal amounts of tin selenide and tin dioxide. In addition to tin selenide a minor amount of selenium(IV) was found to be removed via adsorption on the tin dioxide formed in situ. Tests with a complex industrial solution also resulted in full and stable selenium precipitation. CONCLUSION: Stannous ions were found to be very effective in removing selenious ions from synthetic and industrial solutions, producing very stable precipitates. Copyright © 2012 Society of Chemical Industry  相似文献   

10.
The spectra for 1:1 complexes formed between triscarbonatouranium(VI) + H2O2 and triscarbanatoneptunium(VI) + H2O2 are presented. The respective rates of formation (25°C, 0.05 M NA2CO3) are 565 ± 41 M−1 s−1 and (2.19 ± .01) X 103 M−1 s−1. The corresponding activation parameters are ΔH* = 67.8 ± 3.2 kJ/m, 43.6 ± 2.0 kJ/m, ΔS* = 30 ± 11 J/m °K and −36 ± 7 J/m °K, respectively. The U(VI) complex appears to be stable over a period of months while the Np(VI) complex is formed as a transient species that disappears via a complex process.  相似文献   

11.
A series of 2-[1-(2,6-dibenzhydryl-4-chlorophenylimino)ethyl]-6-[1-(arylimino)ethyl]pyridine ligands (L1–L5) as well as the ligand 2,6-bis[1-(2,6-dibenzhydryl-4-chloro-phenylimino)ethyl]pyridine (L6) were synthesized and reacted with FeCl2·4H2O to afford the iron(II) dichloride complexes [LFeCl2] (Fe1–Fe6). All new compounds were fully characterized by elemental and spectroscopic analysis, and the molecular structures of the complexes Fe1, Fe2 and Fe4 were determined by single-crystal X-ray diffraction, which revealed a pseudo-square-pyramidal geometry at iron. Upon activation with either MAO or MMAO, all iron pre-catalysts exhibited very high activity in ethylene polymerization with good thermal stability. To the best of our knowledge, the current system showed the highest activity amongst iron bis(imino)pyridine pre-catalysts reported to-date. The polymerization parameters were explored to determine the optimum conditions for catalytic activity, which were typically found to be 2500 eq. Al to Fe at 60 °C in the presence of MMAO, and 80 °C in the presence of MAO. The resultant polyethylene possessed a narrow molecular polydispersity index (PDI) consistent with the formation of single-site active species.  相似文献   

12.
Fe-Pt films with an Fe/Pt ratio close to one can be electrodeposited from an FeSO4-H2PtCl6-Na2SO4electrolyte. At the deposition potential, the hydrogen evolution and the reduction of the Pt complex are diffusion limited, and Fe overpotential deposition has not yet set in. The sources of the Fe incorporation are iron hydroxide formation together with Fe underpotential deposition due to Fe-Pt alloy formation. Mössbauer measurements show that the iron in the iron hydroxide is predominantly Fe(III). For stoichiometry reasons, a Pt-rich Fe-Pt phase must be present in addition to the Fe(III)-hydroxide. The Fe3+ that takes part in the hydroxide formation is produced in the electrolyte by the oxidation of Fe2+ by the complexed Pt ion. This exchange reaction results in a significantly higher Fe3+ content in the FeSO4-H2PtCl6-Na2SO4 electrolyte in comparison to the same electrolyte without H2PtCl6. Fe(III)-hydroxide formation can be depressed by adding citric acid, that acts as buffering and complexing agent. This leads to a lower iron content of the deposits. The Fe/Pt ratio close to one that is needed for hard magnetic properties can, however, only be achieved with a significant incorporation of iron hydroxide.  相似文献   

13.
The reaction of cis-[ReVO2I(PPh3)2] with 2-(3,5-dimethylpyrazol-1-yl)benzothiazole (dbt) in ethanol led to the isolation of the trigonal-bipyramidal oxorhenium(IV) complex trans-[ReOI2(PPh3)2] (1). The complex is the first example of a monomeric oxo complex of rhenium(IV). Complex 1 was characterized by FTIR, 1H NMR, microanalysis and single crystal X-ray diffraction.  相似文献   

14.
Chiral bioinspired iron complexes of N4 ligands based on the ethylenediamine backbone display remarkable levels of enantioselectivity for the first time in the asymmetric epoxidation of α,β‐unsaturated ketones using hydrogen peroxide (up to 87% ee) or peracetic acid as oxidant, respectively. Notablely, isotopic labeling with H218O strongly demonstrated that there is a reversible water binding step prior to generation of the significant intermediate. Besides, the complex [L2Fe(III)2(μ‐O)(μ‐CH3CO2)]3+ usually derived from the decay of the LFe(IV)O species or thermodynamic sinks for a number of iron complexes was identified by HR‐MS. In addition, the possible mechanisms were proposed and LFe(V)O species may be the main active intermediate in the catalytic system.  相似文献   

15.
The reaction of Fe(CO)5 with CF3COOH/(CF3CO)2O, or with CF3COOH, in DMF at about 140 °C produces the colourless compound 1, [Fe(DMF)6][Fe33-O) (CF3COO)6(DMF)3]2, constituted by the [Fe(DMF)6]2+ cation and by two [Fe33-O)(μ2-CF3COO)6(DMF)3] anions, with a triangular disposition of the three iron atoms, the central substantially coplanar oxygen atom being formulated as a singly charged anion O(–I) stabilized by complexation to iron(II).  相似文献   

16.
A series of iron(II) bis(triflate) complexes [Fe(L)(OTf)2] containing linear tetradentate bis(pyridylmethyl)diamine ligands with a range of ligand backbones has been prepared. The backbone of the ligand series has been varied from a two‐carbon linkage [ethylene ( 1 ), 4,5‐dichlorophenylene ( 2 ) and cyclohexyl ( 3 )] to a three‐carbon [propyl ( 4 )) and a four‐carbon linkage (butyl ( 5 )]. The coordination geometries of these complexes have been investigated in the solid state by X‐ray crystallography and in solution by 1H and 19F NMR spectroscopy. Due to the labile nature of high‐spin iron(II) complexes in solution, dynamic equilibria of complexes with different coordination geometries (cis‐α, cis‐β and trans) are observed with ligands 2 – 5 . In these cases, the geometry observed in the solid state does not necessarily represent the only or even the major geometry present in solution. The ligand field strength in the various complexes has been investigated by variable temperature magnetic moment measurements and UV‐vis spectroscopy. The strongest ligand field is observed with the most rigid ligands 1 and 2 , which generate complexes [Fe(L)(OTf)2] with a cis‐α coordination geometry and the corresponding complexes [Fe(L)(CH3CN)2]2+ display spin crossover behaviour. The catalytic properties of the complexes for the oxidation of cyclohexane, using hydrogen peroxide as the oxidant, have been investigated. An increased flexibility in the ligand results in a weaker ligand field, which increases the lability of the complexes. The activity and selectivity of the catalysts appear to be related to the strength of the ligand field and the stability of the catalyst in the oxidising environment.  相似文献   

17.
Hydrosulfide oxidation and iron dissolution kinetics were studied at normal pressure, under inert (N2) atmosphere, in a liquid–solid mechanically-stirred slurry reactor. The kinetic variables undergoing variations were: hydrosulfide initial concentration (0.90–3.30 mmol/L), oxide initial surface area (16–143 m2/L) and pH (8.0–11.0). The hydrosulfide consumption and products (thiosulfate and polysulfide) formation were quantified by means of capillary electrophoresis, while iron dissolution was monitored through atomic absorption spectroscopy. Most of Fe(II) produced at pH = 9.5 remained associated with the oxide surface in the time-scale of the experiments. The hydrosulfide oxidation by the iron/cerium (hydr)oxide was found to be surface-controlled, with rates (Ri) of both sulfide oxidation and Fe(II) dissolution expressed in terms of an empirical rate equation: Ri = ki[HS]t=0−0.5[A]t=0[H+]t=0−0.5 , where ki represents the apparent rate constants for the oxidation of HS (kHS) or the dissolution of Fe(II) (kFe), [HS]t = 0 is the initial hydrosulfide concentration, [A]t = 0 is the initial Fe/Ce (hydr)oxide surface area and [H+]t = 0 is the initial proton concentration. The rate constant, kHS, for the oxidation of hydrosulfide at pH = 9.5 was (3.4219 ± 0.65) × 10−4 mol2 L−1 m−2 min−1, with the rate of hydrosulfide oxidation being ca. 10 times faster than the rate of Fe(II) dissolution (assuming a 1:2 stoichiometric ratio between HS oxidized and Fe(II) produced; kFe = (3.9116 ± 0.41) × 10−5 mol2 L−1 m−2 min−1).  相似文献   

18.
A polymer‐supported Fe(III) complex of 1,3‐bis(benzimidazolyl)benzene [PS–Fe(III)BBZNH] was used in the photodegradation of xylenol orange (XO) dye with H2O2 under UV irradiation. The catalyst was synthesized and characterized by elemental analysis, and Fourier transform infrared, far‐infrared, and UV–visible–diffuse reflectance spectroscopy, Scanning electron microscopy, Brunauer–Emmett–Teller surface area measurements, thermogravimetric analysis, and magnetic measurements. An octahedral coordination around Fe(III) was confirmed by electronic spectral data, and a decrease in the intensity of the νCH2Cl peak in PS–Fe(III)BBZNH was observed compared to the polymer support; this indicated the binding of the ligand to the support. An array of experiments were carried out to assess the influence of various reaction parameters on its photocatalytic performance to ensure maximum dye degradation. The maximum photocatalytic activity was observed at pH 8 with 125 mg of catalyst, 300 ppm of XO, and 200 ppm of H2O2 with complete mineralization after 90 min, as confirmed by chemical oxygen demand measurements. Furthermore, the reactions were repeated under sunlight and under dark conditions to check the photocatalytic efficiency of PS–Fe(III)BBZNH. It displayed better catalytic performance compared than the unsupported complex, PS–Cu(II)BBZNH [Cu(II) complex of 1,3‐bis(benzimidazolyl)benzene], and PS–VO(IV)BBZNH [VO(IV) complex of 1,3‐bis(benzimidazolyl)benzene]. PS–Fe(III)BBZNH could be recycled for up to seven runs. A tentative mechanism involving ·OH radical was proposed. © 2018 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2018 , 135, 46480.  相似文献   

19.
Iron(II) complexes with 18- and 20-membered hexaaza macrocyclic ligands were entrapped in the nanoreactors of zeolite-Y by a two-step process in the liquid phase: (i) adsorption of [bis(diamine)iron(II)] (diamine = 1,2-diaminoethane, 1,3-diaminopropane, 1,2-diaminobenzene, 1,3-diaminobenzene); [Fe(N–N)2]2+–NaY; in the nanoreactors of the zeolite, and (ii) template condensation of the iron(II) precursor complex with 2,6-diacetylpyridine.The mode of bonding and overall geometry of the complexes and new complex nanoparticles entrapped in the nanoreactor of zeolite-Y ([Fe([18 or 20]py2N4)]2+–NaY, [Fe(Bzo2[18 or 20]py2N4)]2+–NaY) has been inferred through FT-IR, TGA, XRD, XPS spectroscopic techniques and elemental analysis as well as nitrogen adsorption.  相似文献   

20.
The extraction behavior of rare earth (RE) elements using N,N,N′,N′-tetraoctyl diglycolamide (TODGA) in an ionic liquid (IL) system was investigated by slope analyses. Metallic salts of Pr(III), Nd(III), and Dy(III) with bis(trifluoromethylsulfonyl)amide (TFSA) were synthesized and studied for their extraction mechanism. The selected concentration of TODGA was diluted with triethylpentylphosphonium bis(trifluoromethylsulfonyl)amide ([P2225][TFSA]) to prepare an extracting phase for the slope analyses. The stoichiometry of RE(III) was determined in order to estimate the extracted species. Furthermore, the complexation state of the extracted species was evaluated by spectroscopic analyses, including Fourier-transform infrared (FT-IR) spectroscopy, Raman spectroscopy, and ultraviolet–visible (UV–Vis) spectroscopy. The FT-IR and Raman spectra were estimated using density functional theory (DFT) calculations. Thorough analysis of the FT-IR spectrum was carried out in order to assign the TODGA group that mainly coordinated the metal ion. The solvation of the [TFSA]? anion in the coordination sphere of [Nd(TODGA)(2–3)]3+ was investigated by Raman spectroscopic analysis. The coordination ability of TODGA was investigated from the peak shift of the hypersensitive transition (4I9/22G7/2) in UV–Vis spectroscopic measurements. From electrochemical analysis, the extracted [Nd(TODGA)3]3+ complex in [P2225][TFSA] was found to be reduced as per the following reaction: [Nd(TODGA)3]3+ + 3e? → Nd(0) + 3[TODGA] at ?3.0 V, and the diffusion coefficient of [Nd(TODGA)3]3+ was calculated to be 1.6 × 10?11 m2 s?1 at 373 K. The direct electrodeposition of the extracted [Nd(TODGA)3]3+ in [P2225][TFSA] at 373 K allowed us to conclude that the middle layer of Nd electrodeposits was the metallic state, while a part of the top surface was the oxidation state by XPS analysis.  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