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 421 毫秒
1.
The few clusters [B?nA+n+1]+ (n = 0,1) with resolvable mobilities formed in electrosprays of large salts have been used for nanoparticle instrument testing and calibration at sizes smaller than 2 nm. Extensions of this modest size range by charge reduction with uncontrolled gas phase ions has resulted in impure singly charged clusters. Here, we combine two oppositely charged electrosprays of solutions of the same salt B?A+, including: (CnH2n+1)4N+Br? (n = 4,7,12,16), the large phosphonium cation (C6H13)3(C16H33)P+ paired with the anions Im? [(CF3SO2)2N?] or FAP? [(C2F5)3PF3?], and the asymmetric pair [1-methyl-3-pentylimidazolium+FAP?]. Both polarities are simultaneously produced by this source in comparable abundances, primarily as singly charged A+nB?n±1, with tiny contributions from higher charge states. Some but not all of these clusters produce narrow mobility peaks typical of pure ions, even beyond n = 43. Excellent independent stable control of the positive and the negative sprays brought very close to each other is achieved by isolating them electrostatically with a symmetrically interposed metallic screen. Two nanoDMAs covering the size range up to 30 nm (Halfmini and Herrmann DMAs, with classification lengths of 2 and 10 cm) are characterized with these standards, revealing resolving powers considerably higher than previously seen with unipolar electrospray sources. The bipolar source of pure and chemically homogeneous clusters described permits studying size and charge effects in a variety of aerosol instruments in the 1–4 nm size range.

Copyright © 2017 American Association for Aerosol Research  相似文献   

2.
Abstract

N,N′‐dimethyl‐N,N′‐dioctylhexylethoxymalonamide, DMDOHEMA, and di‐n‐hexylphosphoric acid, HDHP, are the extractants of reference for the French DIAMEX–SANEX process for the separation of trivalent actinide ions from the lanthanide ions. In this work, the extraction of Eu3+ and Am3+ by the two extractants, alone or in mixtures, has been investigated under a variety of experimental conditions. The two cations are extracted by HDHP as the M(DHP · HDHP)3 complexes with an Eu/Am separation factor of ~10. With DMDOHEMA, Eu3+ and Am3+ are extracted as the M(NO3)3(DMDOHEMA)2 disolvate species with an Am/Eu separation factor of ~2. The metal distribution ratios measured with a mixture of the two reagents indicated that almost all lanthanides are extracted equally well. The extraction of Eu3+ and Am3+ by HDHP‐DMDOHEMA mixtures exhibits a change of extraction mechanism and a reversal of selectivity taking place at ~1 M HNO3 in the aqueous phase. Below this aqueous acidity, HDHP dominates the metal extraction by the mixture, whereas DMDOHEMA is the predominant extractant at higher aqueous acidities. Some measurements indicated apparent modest antagonism between the two extractants in the extraction of Eu3+ and synergism in the extraction of Am3+. These data were interpreted as resulting from the formation in the organic phase of mixed HDHP‐DMDOHEMA species containing two HDHP and five DMDOHEMA molecules.  相似文献   

3.
Abstract  Two Ag+ complexes [Ag(HL)2(PF6)] (1) and [(AgL) n  · n(CH2Cl2) · n(0.5H2O)] (2) (HL = 5-methyl-2-phenyl-4-[(2-o-tolylamino)-phenylmethylene]pyrazol-3(2H)-one) were synthesized and structurally characterized by EA analysis, IR spectra and X-ray crystallography. The result shows that two expected coordination modes (Modes I and III in Scheme 1) of the HL ligand, can be observed in its Ag+ complexes, while not in other transition metal ions (Ni2+, Co2+ or Cu2+) complexes whether deprotonation or not for the HL ligand. Graphical Abstract  Three possible coordination modes (Modes I, II or III in Scheme 1) of the selected HL (HL = 5-methyl-2-phenyl-4-[(2-o-tolylamino)-phenylmethylene]pyrazol-3(2H)-one) ligand, can be adopted, in which Modes I and III can be observed in its two Ag+ complexes [Ag(HL)2(PF6)](1) and [(AgL) n  · n(CH2Cl2) · n(0.5H2O)] (2), while Mode II just observed in its transition metal ions (Cu2+, Ni2+, or Co2+) complexes, resulting from the deprotonatd form of the HL ligand and the coordination characters of transition metal ions.   相似文献   

4.
《分离科学与技术》2012,47(12):1895-1902
Extraction of uranium (UO22+) and thorium (Th4+) from a nitric acid solution into an imidazolium-type ionic liquids (ILs) of 1-alkyl-3-methylimidazolium hexafluorophosphate ([Cnmim][PF6], n = 6 or 8) was carried out using N,N,N′,N′-tetraoctyl-3-oxapentanediamide (TODGA) as an extractant. It was found that the extraction efficiencies of UO22+ and Th4+ ions are higher in comparison with that done in n-dodecane. The extraction mechanism was deduced by the slope analysis and extraction experiment. Transfer of both ions is assumed to proceed predominantly through the neutral solvation mechanism from nitric acid solution into ILs. The UO22+ ion forms a 1:2 complex with TODGA in ILs at lower acidity, and a 1:1 complex in ILs and in n-dodecane at higher acidity. The Th4+ ion forms a 1:2 complex with TODGA in C6mimPF6 IL or a 1:1 complex in C8mimPF6 IL at lower acidity and a 1:1 complex in both ILs, and n-dodecane at higher acidity. Stripping studies were conducted using sodium salt of EDTA as a stripping ligand. The thermodynamics of extracting UO22+ ions and Th4+ ions from a 3 M HNO3 solution was also studied. The results indicated that the extraction reactions are spontaneous and go through an exothermic process.  相似文献   

5.
Ionic liquids have been projected as the best solvent for extraction and separation of bioactive compounds from various origins. This review offers a collection of the published results, using ionic liquids for the extraction and purification of biomolecules. Ionic liquids have been studied as solvents, co-solvents and supported materials for separation of bioactive compounds. The ionic liquids-based extraction procedures were previously reported, such as ionic liquids-based solid-liquid extraction, liquid-liquid extraction and ionic liquids-modified materials are reviewed and compared to their performance. In this review, the main activities and future challenges are discussed, with major gaps identified using ionic liquids in extraction procedures and by advancing few steps to overcome these drawbacks.

Abbreviation: [(HSO3)C4MIM]+: 1-(4-sulfonylbutyl)-3-methylimidazolium; [(C6H3OCH2)2im]+: 1,3-dihexyloxymethylimidazolium; [CnC1MIM]+: 1-alkyl-2,3-dimethylimidazolium; [CnMIM]+; [Cn, 2, 3, 4, 6, 8, 10, 12]: 1-alkyl-3-methylimidazolium; [CnC1pyr]+: 1-alkyl-3-methylpyridinium; [Cnim]+: 1-alkylimidazolium; [Cnpyr]+: 1-alkylpyridinium; [aCnim]+: 1-allyl-3-alkylimidazolium; [C7H7MIM]+: 1-benzyl-3-methylimidazolium; [C4(C1C1C1Si)im]+: 1-butyl-3-trimethylsilylimidazolium; [(HOOC)C2MIM]+: 1-carboxyethyl-3-methylimidazolium; [(OH)CnMIM]+: 1-hydroxyalkyl-3-methylimidazolium; [(C2H5O)3SiC3MIM]+: 1-methyl-3-(triethoxy)silypropyl imidazolium; [(NH2)C3MIM]+: 1-propylamine-3-methylimidazolium; [CwHxNyOz]+: Chirally functionalized methylimidazolium; [P10(3OH)(3OH)(3OH)]+: Decyltris(3-hydrox- ypropyl) phosphonium; [N111(2OH)]+: N,N,N-trimethyl-N-(2-hydroxyethyl) ammonium (cholinium); [N00nn]+: N,N-dialkylammonium; [N0nn(2OH)]+: N,N-dialkyl-N-(2-hydroxyethyl) ammonium; [C10C10C1gluc]+: N,N-didecyl-N-methyl-d-glucaminium; [N11(2(O)1)0]+: N,N-dimethyl(2-methoxyethyl) ammonium; [N11(2OH)(C7H7)]+: N-benzyl-N,N-dimethyl-N-(2-hydroxyethyl) ammonium; [P66614]+: Trihexyltetradecylph- osphonium; [Pi(444)1]+: Triisobutyl (methyl) phosphonium; P.minus: Polygonum minus; NPs: Nanoparticle; ZnO : Zinc oxide nanoparticles ; Ni NPs: Nickel nanoparticles; MO: Methyl orange; UAE: Ultrasonic-assisted extraction; LLE: Liquid-liquid extraction; ABS: Aqueous biphasic system ; [Ace]?: Acesulfamate; [Ala]?: alalinate; [TMPP]?: bis(2,4,4-trimethylpentyl)phosphinate; : ; [NTf2]?: bis(trifluoromethylsulfonyl)imide; [[Br]–]: [Br]omide; [Calc]: calkanoate; [Cl]: chloride; [Bz]?: benzoate; [PF6]?: hexafluorophosphate; [HSO4]?: hydrogenosulfate; [OH]?: hydroxide; I: iodide; [Lac]?: lactate; [NO3]?: nitrate; [[Cl]O4]?: perchlorate; [Phe]?: phenilalaninate; [BF4]?: tetrafluoroborate; [SCN]?: thiocyanate; [C(CN)3]?: tricyanomethanide; [CF3CO2]?: trifluoroacetate; [CF3SO3]?: trifluoromethanesulfonate; [FAP]?: tris(pentafluoroethyl)trifluorophosphate; ILs: Ionic liquids; Ag NPs: Silver nanoparticle; Cu NPs: Copper nanoparticle; MB: Methylene blue; MR: Methyl red ; MAE: Microwave-assisted extraction; SLE: solid-liquid extraction.  相似文献   


6.
The effect of tri-n-butyl phosphate (TBP) and iso-decanol as phase modifiers on the complexation of Eu3+ with octyl-phenyl N,N-diisobutyl carbamoyl methyl phosphine oxide (CMPO) was studied. On an increase of iso-decanol the extraction efficiency of Eu3+ decreases. The metal–ligand stoichiometry was found to be 1:3. The Eu complex of CMPO–TBP was more symmetric compared to CMPO–iso-decanol. The Eu3+ complex exhibited D3h and C3h symmetry for 30% and 5% iso-decanol, while that for TBP was C6v. The radiative, non-radiative life time, electric and magnetic dipole transition probabilities, branching ratio, quantum efficiency etc. were evaluated. The covalency between Eu3+ and CMPO is as follows: CMPO–TBP < CMPO-5% iso-decanol < CMPO-30% iso-decanol.  相似文献   

7.
Extraction of uranium with bis-(2-ethylhexyl) phosphonic acid (PC-88A) and bis(2,4,4-trimethylpentyl) phosphinic acid (Cyanex-272) was studied in several imidazolium-based room-temperature ionic liquids (RTILs), Cnmim·X (where n = 4, 6, 8 and X = PF6 and Tf2N). The extraction kinetics was slow and about 0.5–1 h equilibration time was required for most of the extraction systems, except in C8mim·PF6, where 2 h and 4 h were required to reach the equilibrium values for PC-88A and Cyanex-272, respectively. The extraction of UO22+ ion by the two ligands was significantly affected by the nature and the composition of the RTILs.  相似文献   

8.

Abstract  

With the complex 1 or 2 ([Ag(3-pmpmd)]n·n(X) (X = BF4 , 1; X = PF6 , 2) from the semi-rigid 3-pmpmd (N,N′-bis(3-pyridylmethyl)-pyromellitic diimide) ligand and AgBF4 or AgPF6 as the precursor, two new coordination polymers [Ag2(3-pmpmd)2(dppe)(BF4)2]n·4nDMF (3) and [Ag2(3-pmpmd)2(dppe)(PF6)2]n·4nDMF (4) with the 2D cationic MOFs (metal-organic frameworks), have been obtained in the presence of the second dppe (Ph2P(CH2)2PPh2) ligand. In the 2D layer network, the 3-pmpmd ligands show the Z T -mode and the Z C -mode conformations, and the bridged dppe ligands have the same anti conformation. In the meantime, the functions of the two selected ligands, together with the supramolecular interactions from counter ions and solvates molecules within, should play a key role in the construction of the 2D noninterpenetrated network.  相似文献   

9.
Using the melt-quench technique, potassium zinc borophosphate (KZnBP) glasses incorporated with Dy3+, Eu3+, and Dy3+/Eu3+ ions individually and combinedly were prepared, and their photoluminescence (PL)-related features were investigated. The KZnBP glass containing an optimized content of Dy3+ (0.5 mol%) is co-doped with Eu3+ in various contents, and the energy transfer (ET) process between them was studied at λexci = 349, 364, 387 (Dy3+), and 394 nm (Eu3+). The Dy3+/Eu3+ co-doped system, when excited with Dy3+ excitations has resulted in a significant decrease in the intensity of Dy3+ peaks observed at 480 nm (4F9/26H15/2, blue) and 574 nm (4F9/26H13/2, yellow), with simultaneous enhancement of the intensity of Eu3+ peaks at 591 nm (5D07F1, orange) and 617 nm (5D07F2, red). This trend is due to the efficient energy transfer from Dy3+ to Eu3+, indicating that Eu3+ ions were sensitized by Dy3+ ions. Dexter's theory and the Inokuti–Hirayama (I–H) model revealed that the dipole–dipole interaction is accountable for the energy transfer from Dy3+ to Eu3+ through energy-transfer channels [4F9/2(Dy3+)+7F1,2(Eu3+)→6H15/2(Dy3+)+5D2(Eu3+)] and [4F9/2(Dy3+)+7F0(Eu3+)→6H13/2(Dy3+)+5D0(Eu3+)]. The color coordinates of the Dy3+/Eu3+ co-doped glasses under various excitations fall within the white light emission spectrum, indicating their potential application in warm white LEDs.  相似文献   

10.
Generation of monomobile molecular standards by electrospray (ES) followed by classification in a differential mobility analyzer (DMA) fails at diameters above ~2 nm because many clusters in different charge states z crowd in a narrow mobility range. Use of a second DMA (DMA2) in series (tandem) with DMA1 is very helpful because, unexpectedly, many multiply charged ions selected in DMA1 undergo spontaneous transitions, appearing as pure species at different mobilities in DMA2. Remarkably, for salt clusters of composition (CA) n (C+ ) z carrying z elementary charges and n neutral ion pairs, (i) ion evaporation (CA) n (C+ ) z →(CA) n –1(C+ ) z– 1+(CA)C+ and (ii) neutral evaporation transitions (CA) n (C+ ) z →(CA) n –1(C+ ) z+CA affect a substantial fraction of the clusters. Neutral evaporation (fueled by the Kelvin effect) is effective in isolating singly charged clusters, yielding mobility standards easily exceeding 2 nm. Ion evaporation (fueled by large electric fields) produces even larger well-resolved standards. Singly charged clusters of up to 2.5 nm rising in isolation result from metastable doubly charged parent ions (z = 2→1 transition). Isolated doubly charged ions of up to 3.5 nm arise from the z = 3→2 transition, but are harder to resolve from the products of higher initial charge states. We report tandem DMA measurements for electrosprayed nanodrops of two ionic liquids: EMI-Im and EMI-Methide, both based on the small cation EMI+ (1-Ethyl-3-methylimidazolium+) and two relatively large anions: Im? = (CF3SO2)2N?; Methide? = (CF3SO2)3C?. Some exploration on the effect of actively reducing the charge on the clusters as they pass between both analyzers is also included.

Copyright 2013 American Association for Aerosol Research  相似文献   

11.
《分离科学与技术》2012,47(8):2090-2102
Abstract

Liquid–liquid extraction with imidazolium based ionic liquids ([C4mim][PF6], [C6mim][PF6], and [C8mim][PF6]) is proposed for the separation of furfural or 5-methylfurfural from aqueous solution. Factors affecting the extraction of furfural or 5-methylfurfural have been studied. It was shown that the extraction equilibria can be achieved within 30 min and the process was less affected by the factors such as volume ratio and feed concentration. The partition coefficients of furfural and 5-methylfurfural decreased with increasing temperature. [C6mim][PF6] was found to have the best extraction ability among the three ionic liquids studied. Presence of small amount of NaCl or Na2SO4 in the aqueous phase results in the considerable increase in the partition coefficients of furfural because of the competitive hydration of the salts with furfural. A thermodynamic study revealed that the extraction process was driven mainly by hydrophobic interactions. Further experimental results indicated that furfural can be separated selectively from aqueous furfural/acetic acid mixtures.  相似文献   

12.
A glycolamide-functionalized ionic liquid (G-FIL) was synthesized for the first time and was evaluated for the extraction of actinide ions such as Am3+, Pu4+ and UO22+ and fission product element ions such as Eu3+, Sr2+ and Cs+. The extraction of the trivalent metal ions was found to be exceptionally high at low acid concentrations, which rapidly decreased with increasing acidity. In view of the high viscosity of the G-FIL, the studies were carried out using its diluted solution in a commercial ionic liquid, viz. 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C4mim][Tf2N]).  相似文献   

13.
ABSTRACT

During the extraction of lithium from high Mg-containing salt lake brines by tributyl phosphate (TBP) in the presence of Fe(III), H+ is used to stabilize Fe(III). However, the distribution ratio of H+ (DH) is 4–6 times higher than that of Li+ (DLi), which affects the extraction of Li+ significantly. In this study, the competition mechanism between H+ and Li+ was investigated by spectral analysis and thermodynamic equilibrium. The extracted species are determined as HFeCl4 · 2TBP and LiFeCl4 · 2TBP for H+ and Li+, respectively. The apparent equilibrium constants are KH = 799.8 and KLi = 120.6, respectively. Both equilibrium constants and the distribution ratios for H+ and Li+ extraction show that extraction of H+ is stronger than Li+.  相似文献   

14.
In this study, a series of red-emitting Ca3Sr3(VO4)4:Eu3+ phosphors co-doped with La3+ was prepared using the combustion method. The microstructures, morphologies, and photoluminescence properties of the phosphors were investigated. All Ca3Sr3(VO4)4:Eu3+, La3+ samples synthesized at temperatures greater than 700 ℃ exhibited the same standard rhombohedral structure of Ca3Sr3(VO4)4. Furthermore, the Ca3Sr3(VO4)4:Eu3+, La3+ phosphor was effectively excited by near-ultraviolet light of 393 nm and blue light of 464 nm. The strong excitation peak at 464 nm corresponded to the 7F05D2 electron transition of Eu3+. The strong emission peak observed at 619 nm corresponded to the 5D07F2 electron transition of Eu3+. Co-doping with La3+ significantly improved the emission intensity of Ca3Sr3(VO4)4:Eu3+ red phosphors. The optimum luminescence of the phosphor was observed at Eu3+ and La3+ concentrations of 5% and 6%, respectively. Moreover, co-doping with La3+ also improved the fluorescence lifetime and thermal stability of the Ca3Sr3(VO4)4:Eu3+ phosphor. The CIE chromaticity coordinate of Ca3Sr3(VO4)4:0.05Eu3+, 0.06La3+ was closer to the NTSC standard for red phosphors than those of other commercial phosphors; moreover, it had greater color purity than that of all the samples tested. The red emission intensity of Ca3Sr3(VO4)4:0.05Eu3+, 0.06La3+ at 619 nm was ~1.53 times that of Ca3Sr3(VO4)4:0.05Eu3+ and 2.63 times that of SrS:Eu2+. The introduction of charge compensators could further increase the emission intensity of Ca3Sr3(VO4)4:Eu3+, La3+ red phosphors. The phosphors synthesized herein are promising red-emitting phosphors for applications in white light-emitting diodes under irradiation by blue chips.  相似文献   

15.
A series of Eu3+-doped C12H18Ca3O18 phosphors were synthesized through a facile hydrothermal method and the properties of as-prepared phosphors were explored by X-ray diffractometer (XRD), scanning electron microscope (SEM), and photoluminescence (PL) spectrometer. The exploration results indicated that the C12H18Ca3O18:Eu3+ had been successfully synthesized. The morphology of C12H18Ca3O18:Eu3+ was a strip with the size of 100–4000 nm × 50–400 nm × 50–200 nm and the ratio of length to width of 2–80. The strongest emission peak of C12H18Ca3O18:Eu3+ around 620 nm was ascribed to 5Do7F2 transition of Eu3+, and the peaks centered at 590, 653 and 694 nm respectively corresponded to 5Do7F1, 7F3, and 7F4 transitions. C12H18Ca3O18: Eu3+ gave the red light emission, as indicated by color coordinate analysis. The photoluminescence intensity of the phosphors prepared under the Eu3+ concentration of 6% was the highest. The crystal structure of C12H18Ca3O18:Eu3+ was changed after europium ions occupied the lattice position of calcium ions. Europium ion could displace calcium arbitrarily. As a new kind of matrix, calcium citrate possesses the properties of both organic and inorganic compounds and the luminescent C12H18Ca3O18: x Eu3+ particles may be applied in biological fluorescent tags and luminescent materials.  相似文献   

16.
Rhabdophane-type Eu3+,Tb3+-codoped LaPO4·nH2O single-crystal nanorods with the compositions La0.99999-xEuxTb0.00001PO4·nH2O (x?=?0–0.03), La0.99999-yTbyEu0.00001PO4·n′H2O (y?=?0–0.010), and La0.99999-zTbzEu0.000007PO4·n′′H2O (z?=?0–0.012) were hydrothermally synthesized with microwaves. It is shown that the Eu3+,Tb3+ codoping does not affect the thermal stability of these nanorods, which is due to the formation of substitutional solid solutions with both Eu3+ and Tb3+ replacing La3+ in the crystal lattice. Moreover, it is also shown that monazite-type Eu3+,Tb3+-codoped LaPO4 single-crystal nanorods can be obtained by calcining their rhabdophane-type Eu3+,Tb3+-codoped LaPO4·(n,n′ or n′′)H2O counterparts at moderate temperature in air, and that they are thermally stable. It is also observed that, for the same Eu3+,Tb3+-codoping content, the monazite-type Eu3+,Tb3+-codoped LaPO4 nanorods exhibit higher photoluminescent efficiency than the rhabdophane-type Eu3+,Tb3+-codoped LaPO4· (n,n′ or n′′)H2O nanorods. Moreover, it is found that the highest photoluminescence emission corresponds to the monazite-type La0.96999Eu0.02Tb0.00001PO4 nanorods for the La0.99999-xEuxTb0.00001PO4 system. However, for those compositions energy transfer from Tb3+ to Eu3+ does not occur. In addition, for an efficient energy transfer to occur, a content of at least 1?mol% Tb3+ is needed in all the studied materials.  相似文献   

17.
The extraction of UO22+ ion was studied using six different solvent systems containing 2-thenoyltrifluoroacetone (HTTA) in room temperature ionic liquids such as [Cnmim][X] (where, n = 4, 6, or 8 and X? = PF6? or NTf2?) from low to moderate pH solutions for the first time. The extraction kinetics studies indicated rather slow attainment of equilibrium which in some cases improved if the solutions were pre-equilibrated with the aqueous phase prior to the actual experiments. The DU values were found to increase with increasing pH and leading to a plateau like profile at higher pH values. The D values were quite high as compared to that obtained with molecular diluents. The nature of the extracted species was ascertained by slope analysis method which suggested species of the type: UO2(TTA)+IL, UO2(TTA)2,IL, and UO2(TTA)2(HTTA)IL in different ionic liquid based solvents. Temperature variation studies on UO22+ ion extraction were also carried out and the thermodynamic parameters were calculated which indicated high endothermicity of the reactions with large positive entropy values.  相似文献   

18.
《分离科学与技术》2012,47(3):437-447
ABSTRACT

The solvent extraction of Nd3+ , Sm3+, Eu3+ , Gd3+, Dy3+ , Ho3+ , and Er3+ from a M ionic strength medium (CH3COONa), pH (CH3COOH) by 8-quinolinol (HOx) and by a mixture of HOx and trioctylphosphine oxide (TOPO) in CCl4, CHCl3, C6H6, C6H5—NO2, C6H5—CH3 C6H5Cl and C6H5—Br is studied using titrimetric methods. A synergistic factor of over 10,000 is obtained. The atomic number and diluent effects are evident in the calculated formation constants and the lanthanide separation factors. A combination of the data obtained and that of Th4+ indicates that an excellent separation of thorium from the lanthanides can be obtained with this extraction system.  相似文献   

19.
《分离科学与技术》2012,47(12):1906-1912
Preparative high-speed counter-current chromatography (HSCCC) was used to separate and purify bioactive constituents from the stems and leaves of Lophatherum gracile Brongn. Six flavone C-glycosides each at over 95% purity including two new compounds were obtained in one-step separation by HSCCC with an optimized two-phase solvent system composed of ethyl acetate-n-butanol-ethanol-water at volume ratio of 4:2:1.5:8.5 (v/v/v/v). The experiment yielded 19.9 mg of luteolin 6-C-β-D-galactopyranosiduronic acid (1→2)-β-D-glucopyranoside (1), 28.5 mg of luteolin 6-C-α-L-arabinopyranosyl-7-O-β-D-glucopyranoside (2), 31.5 mg of isoorientin (3), 44.8 mg of orientin (4), 25.3 mg of swertiajaponin (5) and 12.1 mg of apigenin 6-C-β-D-galactopyranosiduronic acid (1→2)-β-D-glucopyranoside (6) from 500 mg of crude extracts. The purity of these compounds was determined by high-performance liquid chromatography (HPLC). Their chemical structures were identified by electron spray ionization mass spectroscopy (ESI-MS), 1H and 13C nuclear magnetic resonance spectroscopy (NMR).  相似文献   

20.
Abstract

The kinetics of the forward and backward extraction of the title process have been investigated using a Lewis cell operated at 3 Hz and flux or (F) – method of data treatment. The dependences of (F) in the forward extraction on [Fe3+], [H2A2](o), pH, and [HSO4 ?] are 1, 0.5, 1, and ?1, respectively. The value of the forward extraction rate constant (k f ) has been estimated to be 10?7.37 kmol3/2 m?7/2 s?1. The analysis of the experimentally found flux equation gives the following simple equation: F f =100.13 [FeHSO4 2+] [A?], on considering the monomeric model of BTMPPA and the stability constants of Fe(III)‐HSO4 ? complexes. This indicates the following elementary reaction occurring in the aqueous film of the interface as rate determining: [FeHSO4]2++A?→[FeHSO4.A]+. The very high activation energy of 91 kJ mol?1 supports this chemical reaction step as rate-determining. The negative value of the entropy change of activation (?94 J mol?1 K?1) indicates that the slow chemical reaction step occurs via the SN2 mechanism.

The backward extraction rate can be expressed by the equation: F b =10?5.13 [[FeHSO4A2]](o) [H+] [H2A2](o) ?0.5. An analysis of this equation leads to the following chemical reaction step as rate-determining: [FeHSO4A2](int)→[FeHSO4A]+A(i) ?. However, the activation energy of 24 kJ mol?1 suggests that the backward extraction process is intermediate controlled with greater contribution of the diffusion of one or the other species as a slow process. The equilibrium constant obtained from the rate study matches well with that obtained from the equilibrium study.  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