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 656 毫秒
1.
Heat capacities of crystalline 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide [C4mim][NTf2] and 1-hexyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide [C6mim][NTf2] in the range 80 K-Tfus were measured in an adiabatic calorimeter. Anomalies in the heat-capacity curves for the both compounds occurred near 240 K. Positions of the anomalies depended on thermal history of the samples. More stable crystals had higher heat capacities in the range 220-260 K. Below 200 K heat capacities of all the crystals of the same compound were indistinguishable.  相似文献   

2.
A new ionic liquid (IL) based solid-phase microextraction (SPME) fiber was investigated and used for headspace (HS) extraction of methyl tert-butyl ether (MTBE) in a gasoline sample. Using the new IL coated HS-SPME fiber with the combination of gas chromatography-flame ionization detection (GC-FID); sub-to-low μg L−1 concentrations of MTBE were detected. Four different ILs including 1-butyl-3-methylimidazolium tetraflouroborate ([C4C1IM] [BF4]), 1-octyl-3-methylimidazolium tetraflouroborate ([C8C1IM] [BF4]), 1-octyl-3-methylimidazolium hexaflourophosphate ([C8C1IM] [PF6]) and 1-ethyl-3-methylimidazolium ethylsulphate ([C2C1IM] [ETSO4]) were synthesized and examined for extraction, preconcentration and determination of MTBE. It was observed that [C8C1IM] [BF4] showed the highest extraction efficiency and possessed the best extractability for MTBE. The fiber coating takes up the compounds from the sample by absorption in the case of liquid coatings. The calibration graph was linear in a concentration range of 1-120 μg L−1 (R2 > 0.994) with the detection limit of 0.09 μg L−1 level. The new IL-coated fiber was applied successfully for the determination of MTBE in a gasoline sample with good recoveries between 90 and 95%.  相似文献   

3.
Several imidazolium-based ionic liquids (ILs) with varying cation alkyl chain length (C4–C10) and anion type (tetrafluoroborate ([BF4]), hexafluorophosphate ([PF6]) and bis(trifluoromethylsulfonyl)imide ([Tf2N])) were used as reaction media in the microwave polymerization of methacrylate-based stationary phases. Scanning electron micrographs and backpressures of poly(butyl methacrylate-ethylene dimethacrylate) (poly(BMA-EDMA)) monoliths synthesized in the presence of these ionic liquids demonstrated that porosity and permeability decreased when cation alkyl chain length and anion hydrophobicity were increased. Performance of these monoliths was assessed for their ability to separate parabens by capillary electrochromatography (CEC). Intra-batch precision (n = 3 columns) for retention time and peak area ranged was 0.80–1.13% and 3.71–4.58%, respectively. In addition, a good repeatability of RSDRetention time = <0.30% and ∼1.0%, RSDPeak area = <1.30% and <4.3%, and RSDEfficiency = <0.6% and <11.5% for intra-day and inter-day, respectively exemplify monolith performance reliability for poly(BMA-EDMA) fabricated using 1-hexyl-3-methylimidazolium tetrafluoroborate ([C6mim][BF4]) porogen. This monolith was also tested for its potential in nanoLC to separate protein digests in gradient mode. ILs as porogens also fabricated different alkyl methacrylate (AMA) (C4–C18) monoliths. Furthermore, employing binary IL porogen mixture such as 1-butyl-3-methylimidazolium tetrafluoroborate ([C4mim][BF4]) and 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C4mim][Tf2N]) successfully decreased the denseness of the monolith, than when using [C4mim][Tf2N] IL alone, enabling a chromatographic run to be performed with 1:1 ratio produced baseline separation for the analytes. The combination of ILs and microwave irradiation made polymer synthesis very fast (∼10 min), entirely green (organic solvent-free) and energy saving process.  相似文献   

4.
A simple analytical method, based on the coupling of ionic liquid-based extraction with high performance liquid chromatography (HPLC), is developed for the determination of Sudan dyes (I, II, III and IV) and Para Red in chilli powder, chilli oil and food additive samples. Two ionic liquids (ILs), 1-butyl-3-methylimidazolium hexafluorophosphate ([C4mim][PF6]) and 1-octyl-3-methylimidazolium hexafluorophosphate ([C8mim][PF6]), were compared as extraction solvents; experiments indicated that the latter possesses higher recoveries for each analyte. Parameters related to extraction of Sudan dyes and Para Red were also optimized. Under the optimal conditions, good reproducibility of extraction performance was obtained, with the relative standard deviation (RSD) values ranging from 2.0% to 3.5%. The detection limits of Sudan dyes and Para Red (LOD, S/N = 3) were in the range of 7.0-8.2 μg kg−1 for chilli powder and 11.2-13.2 μg L−1 for chilli oil and food additive. The recoveries were in the range of 76.8-109.5% for chilli powder samples and 70.7-107.8% for chilli oil and food additive samples.  相似文献   

5.
The effect of volatile organic compounds (VOCs) such as acetone, methanol, ethanol, chloroform, carbon tetrachloride, dichloromethane, and hexane on electrical conductivity of thin films of bis[tetrakis(alkylthio)phthalocyaninato]lutetium(III) double decker complexes [(CnH2n+1S)4Pc]2Lu(III) was investigated. The [(CnH2n+1S)4Pc]2Lu(III) molecules substituted with different alkylthia chains (n = 6, 8, 10, 12, and 16) were coated on interdigital transducers using a jet spray technique. A change (increase or decrease) in the conductivity of the [(CnH2n+1S)4Pc]2Lu(III) films was observed depending on the concentration of the VOCs, which was ranging from 500 to 5000 ppm. The decrease in the conductivity of the sensors for the dissolvent of the compounds (chloroform, carbon tetrachloride, dichloromethane and hexane) could be related to swelling of the films. On the other hand, the increase in the conductivity of the sensors for the other VOCs (acetone, methanol and ethanol) could be resulted from that the VOCs act as electron donors and/or acceptors in the films. A linear relationship between the sensor response and concentration of the VOC vapors is obtained. The sensitivities of the [(CnH2n+1S)4Pc]2Lu(III) films were in the range of 2.10−4-3.10−3%/ppm.  相似文献   

6.
The effect of ionic liquids on the formation of a partial positive charge on the surface of silver nanoparticle and its subsequent effect on facilitated olefin transport were investigated. Three different ionic liquids of 1-butyl-3-methylimidazolium tetrafluoroborate (BMIM+BF4), 1-butyl-3-methylimidazolium triflate (BMIM+Tf), and 1-butyl-3-methylimidazolium nitrate (BMIM+NO3) were employed to control the positive charge density of the surface of silver nanoparticles. The positive charge density of the silver nanoparticles, as characterized by the binding energy of the silver atom, was in the following order: BMIM+BF4/Ag ? BMIM+Tf/Ag > BMIM+NO3/Ag. This order was consistent with the tendency of ionic liquids to form free ions. The best separation performance for the propylene/propane mixtures was a mixed gas selectivity of 17 and a permeance of 7 GPU through a composite membrane consisting of BMIM+BF4/Ag. A better separation performance for olefin/paraffin mixtures was observed with a higher positive charge density of the silver nanoparticles. It was therefore concluded that facilitated olefin transport was a direct consequence of the surface positive charge of the silver nanoparticles induced by ionic liquids.  相似文献   

7.
The glycolysis of poly(ethylene terephthalate) (PET) was studied using several ionic liquids and basic ionic liquids as catalysts. The basic ionic liquid, 1-butyl-3-methylimidazolium hydroxyl ([Bmim]OH), exhibits higher catalytic activity for the glycolysis of PET, compared with 1-butyl-3-methylimidazolium bicarbonate ([Bmim]HCO3), 1-butyl-3-methylimidazolium chloride ([Bmim]Cl) and 1-butyl-3-methylimidazolium bromide ([Bmim]Br). FT-IR, 1H NMR and DSC were used to confirm the main product of glycolysis was bis(2-hydroxyethyl) terephthalate (BHET) monomer. The influences of experimental parameters, such as the amount of catalyst, glycolysis time, reaction temperature, and dosages of ethylene glycol on the conversion of PET, yield of BHET were investigated. The results showed a strong influence of the mixture evolution of temperature and reaction time on depolymerization of PET. Under the optimum conditions of m(PET):m(EG): 1:10, dosage of [Bmim]OH at 0.1 g (5 wt%), reaction temperature 190 °C and time 2 h, the conversion of PET and the yield of BHET were 100% and 71.2% respectively. Balance between the polymerization of BHET and depolymerization of PET could be changed when the reaction time was more than 2 h and contents of catalyst and EG were changed.  相似文献   

8.
This study examined standard solutions to assess the influence of the gas flow rate and organic solvent type on losses caused by gas blowdown of polychlorinated dibenzo-p-dioxins and dibenzofurans (PCDD/DFs) and coplanar polychlorinated biphenyls (Co-PCBs). Results obtained here will contribute to maintaining analytical method performance and system quality for PCDD/DFs and Co-PCBs analyses. An organic solvent (with 0.5 ml each of acetone, dichloromethane, n-hexane, and toluene), PCDD/DFs or Co-PCBs, and their 13C12-labeled compounds were put separately into 10 ml pear-shaped flasks. The samples were blown to dryness at room temperature until the last trace of solvent disappeared. They were subsequently reconstituted in those flasks. Analyte recoveries were calculated by comparing blown samples to those that had not been blown. Recoveries of Co-PCBs were more affected than those of PCDD/DFs when the gas flow rates were set at 203, 261, 332, and 456 ml/min. Losses of Co-PCBs were least at 203-332 ml/min. Regarding losses of PCDD/DFs and Co-PCBs, the toluene solution showed the least variation in recovery. An actual soil sample extract was also examined using optimized conditions for the gas flow rate and solvent types obtained by experiments in standard solutions. Thereby, the blowdown conditions gave quantitative recoveries of 13C12-labeled compounds in the sample extract.  相似文献   

9.
Binodal curves of the aqueous 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim]BF4) + sodium citrate (Na3C6H5O7), [Bmim]BF4 + sodium tartrate (Na2C4H4O6) and [Bmim]BF4 + sodium acetate (NaC2H3O2) systems have been determined experimentally at 298.15 K. The Merchuk equation was used to correlate the binodal data. The effective excluded volume (EEV) values obtained from the binodal model for these three systems were determined. The binodal curves and EEV both indicate that the salting-out abilities of the three salts follow the order: Na3C6H5O7 > Na2C4H4O6 > NaC2H3O2. The liquid–liquid equilibrium (LLE) data were obtained by density determination and binodal curves correlation of these systems. Othmer–Tobias and Bancraft, and Setschenow equations were used for the correlation of the tie-line data. Good agreement was obtained with the experimental tie-line data with both models.  相似文献   

10.
Chloride based ionic liquids were used as chloride source in Meerwein reaction either in [bmim]X (bmim = 1-butyl-3-methylimidazolium, X = BF4, PF6) as solvents or in solventless conditions. Satisfactory yields (49-71%) with diversely substituted diazonium salts were achieved by using 1,3-dibutylimidazolium chloride in the presence of a bimetallic Zn/Cu catalyst.  相似文献   

11.
The reaction between uranyl nitrate hexahydrate and phenolic ligand precursor [(N,N-bis(2-hydroxy-3,5-dimethylbenzyl)-4-amino-1-butanol) · HCl], H3L1 · HCl, leads to a uranyl complex [UO2(H2L1)2] (1a) and [UO2(H2L1)2] · 2CH3CN (1b). The ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-4-amino-1-butanol)H3L2 · HCl], H3L2 · HCl, yields a uranyl complex with a formula [UO2(H2L2)2] · CH3CN (2). The ligand [(N,N-bis(2-hydroxy-3,5-dimethylbenzyl)-5-amino-1-pentanol) · HCl], H3L3 · HCl, produces a uranyl complex with a formula [UO2(H2L3)2] · 2CH3CN (3) and the ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-5-amino-1-pentanol) · HCl], H3L4 · HCl, leads to a uranyl complex with a formula [UO2(H2L4)2] · 2CH3CN (4). The ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-6-amino-1-hexanol) · HCl], H3L5 · HCl, leads to a uranyl complex with a formula [UO2(H2L5)2] · 4toluene (5). The complexes 15 are obtained using a molar ratio of 1:2 (U to L) in the presence of a base (triethylamine). The molecular structures of 1a, 1b, 3, 4 and 5 were verified by X-ray crystallography. All complexes are neutral zwitterions and have similar centrosymmetric, mononuclear, distorted octahedral uranyl structures with the four coordinating phenoxo ligands in an equatorial plane. In uranyl ion extraction studies from water to dichloromethane with ligands H3L1 · HCl–H3L5 · HCl, ligands H3L1 · HCl, H3L4 · HCl and H3L5 · HCl are the most effective ones.  相似文献   

12.
Oxidative addition of methyl iodide to Vaska’s complex in the ionic liquids 1-butyl-3-methylimidazolium triflate [C4mim][OTf], [C4mim] bis(trifluormethylsulfonyl)imide [Tf2N], and N-hexylpyridinium [C6pyr][Tf2N] occurred cleanly to give the expected Ir(III) oxidative addition product. Pseudo-first order rate constants were determined for the oxidative addition reaction in each solvent ([Vaska’s] = 0.25 mM, [CH3I] = 37.5 mM). The observed rate constants under these conditions were 5-10 times slower than the rate seen in DMF. At high methyl iodide concentrations (>23 mM), the expected first order dependence on methyl iodide was not observed. In each ionic liquid, there was no change in the reaction rates within experimental error over the methyl iodide concentration range of 23-75 mM. At lower methyl iodide concentration, a decrease in rate was observed in [C4mim][Tf2N] with decreasing concentration of methyl iodide.  相似文献   

13.
Condensation of (S)-2-amino-2′-hydroxy-1,1′-binaphthyl with 1 equiv. of pyrrole-2-carboxaldehyde in toluene in the presence of molecular sieves at 70 °C gives (S)-2-(pyrrol-2-ylmethyleneamino)-2′-hydroxy-1,1′-binaphthyl (1H2) in 90% yield. Deprotonation of 1H2 with NaH in THF, followed by reaction with LnCl3 in THF gives, after recrystallization from a toluene or benzene solution, dinuclear complexes (1)3Y2(thf)2 · 3C7H8 (3 · 3C7H8) and (1)3Yb2(thf)2 · 3C6H6 (4 · 3C6H6), respectively, in good yields. Treatment of 1H2 with Ln[N(SiMe3)2]3 in toluene under reflux, followed by recrystallization from a benzene solution gives the dimeric amido complexes {1-LnN(SiMe3)2}2 · 2C6H6 (Ln = Y (5 · 2C6H6), Yb (6 · 2C6H6)) in good yields. All compounds have been characterized by various spectroscopic techniques, elemental analyses and X-ray diffraction analyses. Complexes 5 and 6 are active catalysts for the polymerization of methyl methacrylate (MMA) in toluene, affording syn-rich poly-(MMA)s.  相似文献   

14.
Temperature-controlled ionic liquid dispersive liquid phase microextraction (TCIL-DLPME) combined with high performance liquid chromatography-diode array detection (HPLC-DAD) was applied for preconcentration and determination of chlorobenzenes in well water samples. The proposed method used 1-butyl-3-methylimidazolium hexafluorophosphate ([C4mim][PF6]) as the extraction solvent. The effect of different variables on extraction efficiency was studied simultaneously using an experimental design. The variables of interest in the TCIL-DLPME were extraction solvent volume, salt effect, solution temperature, extraction time, centrifugation time, and heating time. The Plackett-Burman design was employed for screening to determine the variables significantly affecting the extraction efficiency. Then, the significant factors were optimized by using a central composite design (CCD) and the response surface equations were developed. The optimal experimental conditions obtained from this statistical evaluation included: extraction solvent volume, 75 μL; extraction time, 20 min; centrifugation time, 25 min; heating time, 4 min; solution temperature, 50 °C; and no addition of salt. Under optimal conditions, the preconcentration factors were between 187 and 298. The limit of detections (LODs) ranged from 0.05 μg L−1 (for 1,2-dichlorobenzene) to 0.1 μg L−1 (for 1,2,3-trichlorobenzene). Linear dynamic ranges (LDRs) of 0.5-300 and 0.5-500 μg L−1 were obtained for dichloro- and trichlorobenzenes, respectively. The performance of the method was evaluated for extraction and determination of chlorobenzenes in well water samples in micrograms per liter and satisfactory results were obtained (RSDs < 9.2%).  相似文献   

15.
An efficient, one-pot Friedel-Crafts acylation/hydrolysis reaction promoted by the acidic ionic liquid 1-ethyl-3-methylimidazolium chloroaluminate (generated from 1-ethyl-3-methylimidazolium chloride (EmimCl) and aluminum chloride (X(AlCl3), mole fraction X = 0.75) for the formation of 3-glyoxylic acid derivatives of electron-deficient, substituted 4- and 6-azaindoles is described.  相似文献   

16.
Mononuclear neutral arene ruthenium(II) β-diketonato complexes of the general formula (η6-arene)Ru(LL)Cl [LL = 1-phenyl-3-methyl-4-benzoyl pyrazol-5-one (L1), arene = C6H6 (1), p-iPrC6H4Me (2), C6Me6 (3); arene = p-iPrC6H4Me, LL = 1-benzoylacetone (L3) (8); arene = p-iPrC6H4Me, LL = dibenzoylmethane (L4) (9)] have been synthesized and their subsequent substitution reactions with NaN3 in alcohol at room temperature yielded the corresponding neutral terminal azido complexes (η6-arene)Ru(LL)N3 [LL = 1-phenyl-3-methyl-4-benzoyl pyrazol-5-one (L1), arene = C6H6 (4), p-iPrC6H4Me (6), C6Me6 (7); arene = p-iPrC6H4Me, LL = dibenzoylmethane (L4) (10)] as well as a cationic complex [(η6-p-iPrC6H4Me)Ru(L4) (PPh3)]BF4 (12) with PPh3. The [3 + 2] cycloaddition reaction of selective azido complexes with the activated alkynes dimethyl and diethyl acetylenedicarboxylates produced the arene triazolato complexes [(η6-arene)Ru(LL){N3C2(CO2R)2}] [arene = p-iPrC6H4Me, LL = L1, R = Me (13); arene = C6Me6, LL = L1, R = Me (14); arene = C6Me6, LL = acetyl acetone (L2), R = Me (15); arene = C6Me6, LL = L3, R = Me (16); arene = p-iPrC6H4Me, LL = L1, R = Et (17); arene = C6Me6, LL = L1, R = Et (18); arene = C6Me6, LL = L2, R = Et (19); arene = C6Me6, LL = L3, R = Et (20)]. With fumaronitrile the reaction yielded the triazoles [(η6-arene)Ru(LL)(N3C2HCN)] [arene = p-iPrC6H4Me, LL = L1 (21), arene = C6Me6, LL = L1 (22), arene = C6Me6, LL = L2 (23), arene = C6Me6, LL = L3 (24)]. In the above triazolato complexes only N(2) isomer was obtained. The complexes were characterized on the basis of spectroscopic data. Crystal structure of representatives complexes were determined by single crystal X-ray diffraction.  相似文献   

17.
Gurban AM  Rotariu L  Baibarac M  Baltog I  Bala C 《Talanta》2011,85(4):2007-2013
Simple and low cost biosensor based on screen-printed electrode for sensitive detection of some alkylphenols was developed, by entrapment of HRP in a nanocomposite gel based on single-walled carbon nanotubes (SWCNTs) and 1-butyl-3-methylimidazolium hexafluorophosphate ([BMIM][PF6]) ionic liquid. Raman and FTIR spectroscopy, CV and EIS studies demonstrate the interaction between SWCNTs and ionic liquid. The nanocomposite gel, SWCNT-[BMIM][PF6] provides to the modified sensor a considerable enhanced electrocatalytic activity toward hydrogen peroxide reduction. The HRP based biosensor exhibits high sensitivity and good stability, allowing a detection of the alkylphenols at an applied potential of −0.2 V vs. Ag/AgCl, in linear range from 5.5 to 97.7 μM for 4-t-octylphenol and respectively, between 5.5 and 140 μM for 4-n-nonylphenol, with a response time of about 5 s. The detection limit was 1.1 μM for 4-t-octylphenol, and respectively 0.4 μM for 4-n-nonylphenol (S/N = 3).  相似文献   

18.
Osmotic coefficients of binary mixtures containing an ionic liquid, (1-butyl-3-methylimidazolium tetrafluoroborate, [BMIm]BF4, 1-ethyl-3-methylimidazolium ethyl sulfate, [EMIm]ES, and 1-butyl-3-methylimidazolium methyl sulfate, [BMIm]MS) with water were measured until about 3 molal concentrations using vapor pressure osmometry method (VPO) at temperature ranges 298.15–328.15 K and modeled using different electrolyte excess Gibbs free energy models including electrolyte non-random two liquids (NRTL), modified NRTL (MNRTL), mean spherical approximation NRTL (MSA-NRTL), non random factor (NRF), and extended Wilson models. The results show that osmotic coefficient data increase with increasing temperature. The calculated standard deviations of the studied systems show that the applicability of these models for the correlation of VLE properties of ionic liquid solutions. The average standard deviations for the models have the order σ(?) MNRTL < σ(?) Wilson < σ(?) NRTL < σ(?) MSA-NRTL < σ(?)NRF. The results show MNRTL model is able to reproduce experimental osmotic coefficients of aqueous solution of studied ionic liquids with good precision.  相似文献   

19.
A room-temperature ionic liquid (RTIL), 1-butyl-3-methylimidazolium hexafluorophosphate, was investigated as a selective deposition for quartz crystal microbalances. RTILs constitute versatile matrices of tuneable physicochemical properties providing a high solute mobility compared to polymer matrices, and hence in principle short response and desorption times. This paper explores the feasibility of using RTIL-based quartz crystal microbalances as vapour sensors with particular focus on the physicochemical interactions between the RTIL and a model solute, ethyl acetate. Preliminary experiments were conducted and proved an excellent baseline recovery of the sensor. The transient sensor response was modelled using a basic mass transport equation which proved the simplicity of the system when compared to polymer coatings. The diffusion coefficient of ethyl acetate was calculated to be 10.8·10− 11 m2·s− 1, which was one order of magnitude higher than in a common polymeric deposition material, polydimethylsiloxane. However, it is pointed out that care must be taken when interpreting the sensor signal upon dissolution of a complex vapour; absorption of solutes into the RTIL can give rise to viscosity changes which affect the sensor signal and hence overlap with the response to the solute mass absorbed. Nevertheless, for simple sample vapours we believe that quartz resonators can also be a complementary and cheap method for investigating basic mass transport phenomena.  相似文献   

20.
The syntheses of five new aminoalkylbis(phenolate) ligands (as hydrochlorides) and their uranyl complexes are described. The reaction between uranyl nitrate hexahydrate and phenolic ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-1-aminopropane) · HCl], H2L1 · HCl, forms a uranyl complex [UO2(HL1)2] · 2CH3CN (1). The ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-1-aminobutane) · HCl], H2L2 · HCl, forms a uranyl complex with a formula [UO2(HL2)2] · 2CH3CN (2). The ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methyl benzyl)-1-aminohexane) · HCl], H2L3 · HCl, yields a uranyl complex with a formula [UO2(HL3)2] · 2CH3CN (3) and the ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-cyclohexylamine) · HCl], H2L4 · HCl, yields a uranyl complex with a formula [UO2(HL4)2] (4). The ligand [(N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-benzylamine) · HCl], H2L5 · HCl, forms a uranyl complex with a formula [UO2(HL5)2] · 2MeOH (5). The molecular structures of 1, 2′ (2 without methanol), 3, 4 and 5 were verified by X-ray crystallography. The complexes 15 are neutral zwitterions which form in a molar ratio of 1:2 (U to L) in the presence of a base (triethylamine) and bear similar mononuclear, distorted octahedral uranyl structures with the four coordinating phenoxo ligands forming an equatorial plane and resulting in a centrosymmetric structure for the uranyl ion. In uranyl ion extraction studies from water to dichloromethane with ligands H2L1 · HCl–H2L5 · HCl, the ligands H2L2 · HCl and H2L4 · HCl are the most effective ones.  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