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 40 毫秒
1.
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   

2.
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   

3.
Ozone reacts slowly with Ag+ (circumneutral pH, k = (11 ± 3) × 10?2 M?1 s?1). After some time, ozone decay kinetics may suddenly become faster with the concomitant formation of silver sol. As primary process, an O-transfer from O3 to Ag(I) is suggested, whereby Ag(III) is formed [Ag+ + O3 + 2 H2O → Ag(OH)3 + O2 + H+]. This conproportionates with Ag(I), which is in large excess, leading to Ag(II) [Ag+ + Ag(OH)3 ? 2 Ag(OH)+ + HO?]. Further, Ag(II) reacts with ozone in a high exergonic reaction [Ag(OH)+ + O3 → Ag + 2 O2 + H+], where ozone acts as a reducing agent. Thereby, a single silver atom, Ag, is formed that can be oxidized by O2 and O3 or can aggregate to a silver sol. Aggregation slows down the rate of oxidation. When Ag+ is complexed by acetate ions, ozone decay and silver sol formation are speeded up by enhancing Ag(II) formation [Ag(I)acetate + O3 → Ag(III)acetate → Ag(II) + CO2 + ?CH3]. In the presence of oxalate, the formed complex reacts faster with ozone than Ag+, and Ag(III)oxalate decarboxylates rapidly [Ag(I)oxalate + O3 → Ag(III)oxalate → Ag+ + 2 CO2]. This enhances ozone decay but prevents silver sol formation. Quantum chemical calculations have been carried out for substantiating mechanistic suggestions.  相似文献   

4.
The universal quasichemical functional‐group activity coefficients (UNIFAC) model for ionic liquids (ILs) has become notably popular because of its simplicity and availability via modern process simulation softwares. In this work, new group binary interaction parameters (αmn and αnm) between CO (H2) and IL groups were obtained by correlating the solubility data in pure ILs at high temperatures (above 273.2 K) collected from the literature. the solubility of CO in [BMIM]+[BF4]?, [OMIM]+[BF4]?, [OMIM]+[Tf2N]?, and their mixtures, as well as that of H2 in [EMIM]+[BF4]?, [BMIM]+[BF4]?, [OMIM]+[Tf2N]?, and their mixtures, at temperatures from 243.2 to 333.2 K and pressures up to 6.0 MPa were measured. The UNIFAC model was observed to well predict the solubility in pure and mixed ILs at both high (above 273.2 K) and low (below 273.2 K) temperatures. Moreover, the selectivity of CO (or H2) to CO2 in ILs increases with decreasing temperature, indicating that low temperatures favor for gas separation. © 2014 American Institute of Chemical Engineers AIChE J 60: 4222–4231, 2014  相似文献   

5.
Two type of one-dimensional compounds, K2Ag2GeSe4(2) and K3AgSn3Se8(4), were synthesized with thiophenol as a mineralizer, whilst two oligomers, Cs4Ge2Se6 (1) and K4Sn3Se8(3), were obtained in absence of thiophenol. Compounds 1 and 3 contain dimeric [Ge2Se6]4? and trimeric [Sn3Se8]4? anions, respectively. Compound 2 contains isolated GeSe4 tetrahedra connected by linear-coordinated Ag+ ions to form an infinite anionic chains [Ag2GeSe4]2?. The building blocks [Sn3Se8]4? are linked by tetrahedral coordinated Ag+ ions to generate infinite chain [AgSn3Se8]3? of 4.  相似文献   

6.
The reaction of thioimidazolylborate-zinc(II)-perchlorate complex [Ttxyly·Zn? OClO3] 1 (Ttxyly=hydrotris[N-xylyl-thioimidazolyl]borate) with cysteine and its derivative N-acetyl cysteine and S-methyl cysteine leads to the formation of three new monomeric and dimeric thiolate complexes: [Ttxyly·Zn? Cys? Zn·Ttxyly] 2, [Ttxyly·Zn? Cys(NAc)? Zn·Ttxyly] 3, and [Ttxyly·Zn? Cys(SMe)] 4. The attachment of the cysteine derivatives to the Tt·Zn unit serves as structural models for the active site of methionine synthase. Methylation of the coordinated thiolate in the dinuclear zinc(II) complex 2 with methyl iodide appears to occur intramolecularly at the zinc-bound thiolates, forming methyl thioether-containing zinc(II) complex [Ttxyly·Zn? Cys(SMe)] 4 and iodo complex [Ttxyly·Zn? I] 5 with a clean second-order reaction of k=1.0×10?5 M?1 s?1.  相似文献   

7.
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   


8.
《分离科学与技术》2012,47(4):617-625
Sorption of Ag+ by natural mordenite and its Na-exchanged form was investigated by the batch method. Maximum Ag+ uptake was observed at initial pH > 3.0 and contact time 90 min. The kinetic data fitted very well to the pseudo-second-order rate model with values of k 2 from 37.31 to 0.487 and 48.32 to 0.491 g/meq min for the natural and Na-exchanged mordenite, respectively. The Langmuir model is in good correlation with the isotherm data up to initial concentration of 500 mg Ag+/L (q m = 57.41 (natural sample) and 87.72 mg/g (Na-exchanged sample). The obtained data are promising for clean-up of polluted water.  相似文献   

9.
Abstract

(±)?Syn?dibenzo[a,l]pyrene diol epoxide (DB[a,l]PDE) and (±)?anti?DB[a,l]PDE were reacted with deoxyadenosine (dA) or deoxyguanosine (dG) in dimethylformamide at 100 °C for 30 min. The crude products were purified by reverse phase HPLC under gradient and isocratic conditions. The structure of each adduct was assigned by 1D and 2D NMR spectra and by fast atom bombardment mass spectrometry. Five adducts were isolated from the reaction of (±)?syn?DB[a,l]PDE and dA: syn?DB[a,l]PDE?N6dA?1, syn?DB[a,l]PDE?N6dA?2, syn?DB[a,l]PDE?N6dA?3, syn?DB[a,l]PDE?N6dA?4 and syn?DB[a,l]PDE?N7Ade. Four adducts were isolated from the reaction of (±)?anti?DB[a,l]PDE and dA: anti?DB[a,l]PDE?N6dA?1, anti?DB[a,l]PDE?N6dA?2, anti?DB[a,l]PDE?N6dA?3 and anti?DB[a,l]PDE?N6dA?4. Two adducts were isolated from the reaction of (±)?syn?DB[a,l]PDE and dG: (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N2dG and (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N7Gua. Two adducts were isolated from the reaction of (±)?anti?DB[a,l]PDE and dG: (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N2dG and (±)?11,12,13?trihydroxy?tetrahydroDB[a,l]P?14?N7Gua.  相似文献   

10.
Effect of temperature (5°–65°C) on the separation of 11 inorganic anions by ion interaction chromatography (IIC) was studied employing RP C18 and C PhenylHexyl columns and aqueous mobile phase: 2.8 mM NaHCO3 + 0.7 mM TBAOH (tetra-n-butylammonium hydroxide). The apparent enthalpy changes, ΔH for hydrophobic ions like I?, SCN?, and ClO4 ? largely exceeded 3 kcal/mole suggesting that added to ion exchange they are retained by hydrophobic adsorption. Unlike conventional strongly basic anion exchangers, our system can be used at elevated temperatures with alkaline eluents without irreversible damaging the column.  相似文献   

11.
Reaction of [Ru(bpym)3]2+ (bpym = 2,2′-bipyridmidine) with hexacyanoruthenate under forcing conditions affords a mixture of the trinuclear species [(bpym)Ru{(µ-bpym)Ru(CN)4}2]2?, [1]2?, and the tetranuclear species [Ru{(µ-bpym)Ru(CN)4}3]4?, [2]4?, in which two or three (respectively) of the peripheral vacant bpym binding sites of [Ru(bpym)3]2+ are occupied by {Ru(CN)4}2? fragments. Thus, [1]2? and [2]4? have eight and twelve externally-directed cyanide groups respectively for use in forming high connectivity coordination networks. The crystal structure of HK[1]·2MeOH·6.5H2O reveals a one-dimensional ladder structure in which [1]2? anions are connected by (i) cyanide/K+ and (ii) bpym/K+ coordination interactions.  相似文献   

12.
Two silver(I) complexes, [Ag(dmpyz)2][Ag(barb)2] (1) and {[Ag(ppz)][Ag(barb)2]·H2O} n (2) (barb = 5,5-diethylbarbiturate, dmpyz = 2,5-dimethylpyrazine and ppz = piperazine), have been synthesized and characterized by elemental analyses, IR, thermal analysis (TG-DTA) and single-crystal X-ray diffraction. Complex 1 consists of [Ag(dmpyz)2]+ and [Ag(barb)2]? ions in which the silver(I) ions are linearly coordinated by two dmpyz or two barb ligands. These two ions are connected by strong Ag–Ag interactions (Ag–Ag = 2.896 (1) Å). Complex 2 is a 1D coordination polymer in which the silver(I) ions are bridged by the ppz ligands in a linear fashion, leading to a zigzag chain of [Ag(ppz)] n + , which interacts with the [Ag(barb)2]? units by Ag–Ag interactions of 3.183 (1) Å. The 1D chains are further assembled to form 3D networks by strong N–H···O and OW–H···O hydrogen bonds. IR spectra and TG-DTA data are in agreement with the crystal structures. The fluorescent properties of 1 were also evaluated.  相似文献   

13.
《Ceramics International》2023,49(7):10822-10828
Ion-conducting glasses of AgI-based Ge-Ga-S, Ge-Ga-Se and Ge-Ga-Te with different sulphur group elements were prepared by melt quenching technique. The electrical properties and thermodynamic stability were investigated and compared. Glass transition temperatures and crystallisation temperatures of the samples were measured using differential scanning calorimetry. Results showed that Ge-Ga-S-AgI had the best thermodynamic stability and glass-forming ability, highest ionic conductivity (1.33 × 10?7 S/cm) at room temperature and lowest conductivity activation energy of 0.41 eV. Raman spectroscopy analysis showed that the [GeInS4-n] (n = 1,2,3) structural units in the S series glasses provided a good environment for the effective transport of Ag+ and formed an active ion transport channel. By contrast, the structural units in the Se and Te series contributed less to the transport of Ag+. This work systematically explored the effects of three elements in the sulphur main group on the performance of Ag+ conductive glasses. This study provides a reference for the rational formulation of sulphur-based Ag+ conductive glass components with enhanced comprehensive properties.  相似文献   

14.
To extract iridium(III), various physicochemical parameters were studied. 2-Octylaminopyridine was used for the extraction of iridium(III) from acetate medium at 8.5 pH. Quantitative extraction of iridium(III) was achieved via ion-pair formation of cation [2-OAPH+] and anion [Ir(CH3COO)4]?. The stripping of iridium(III)-laden organic phase was carried out 2 M HCl (3 × 10 mL) . The stoichiometry of the extracted ion–pair complex was found to be 1:4:1 (metal: acetate: extractant). The extracted species [2-OAPH+. Ir(CH3COO)4?] is assumed to be an ion association product of [Ir(CH3COO)4] ? and [2-OAPH]+. The proposed method was successfully used in the separation of iridium(III) from binary and ternary mixtures. Analysis of various alloy samples was also carried out.  相似文献   

15.
Visible absorption spectra and the molar conductance curve of NiCl2 dissolved in dimethyl-sulphoxide (DMSO) have been determined at 25°C. The results indicate the formation of the [NiCl(DMSO)5]+ inner-sphere complex followed by the formation of the {NiCl(DMSO)5]+Cl?} outersphere ion-pair as the concentration of NiCl2 increases. The [NiCl2(DMSO)4] inner-sphere complex in pure dimethyl sulphoxide at 25°C is formed to but a small extent, but its relative content increases upon addition of the non-coordinating diluent, chlorobenzene. Dilutions with chlorobenzene also enhance coordination disproportionation producing the [NiCl3 DMSO]? tetrahedral complex. The stability constant of the monochloro complex derived from the conductometric data has the value of 297 (±10)M?2 at 25°C, and the estimated value of the association constant of the [NiCl (DMSO)5]+Cl? complex electrolyte, dominating in moderately concentrated solutions at room temperature, is 7(±3)M?1.  相似文献   

16.
The reutilisation of Dunaliella salina carotenogenesis medium, after microalgal biomass separation by centrifugation, was assessed. The wastewater had an NaCl concentration between 174 g dm?3 and 254 g dm?3 and an average total organic matter concentration of 1540 mg dm?3 ash‐free dry weight, of which 41% (w/v) was glycerol. The biological treatment was established at laboratory scale and batch operations used halophilic bacteria from the wastewater itself. The wastewater was supplemented with NH4+,PO43?, K+ and Mg2+ ions to enhance growth. The effect of each ion added per se was initially investigated and a response surface methodology (RSM) used to identify the optimal conditions for maximisation of glycerol removal from the wastewater, which was considered to be the main objective. Addition of NH4+ ions alone achieved 79% glycerol removal compared with only 59% in the absence of supplement, after 8 days incubation. The combined addition of ions ([NaCl] = 214 g dm?3, [Mg2+] = 114 mg dm?3, [K+] = 131 mg dm?3, [NH4+] = 113 mg dm?3, [PO43?] = 40 mg dm?3) increased glycerol removal from the wastewater such that, after 2 days incubation, no residual glycerol was apparent in cultures. These ion combinations enabled the halophilic bacteria to efficiently remove glycerol from the wastewater and consequently reduce organic matter. This treated wastewater should be appropriate for reutilisation as a carotenogenesis medium for β‐carotene production from D salina. © 2001 Society of Chemical Industry  相似文献   

17.
《分离科学与技术》2012,47(15):3503-3515
Abstract

A simple and effective homogeneous liquid–liquid extraction method has been used for the simultaneous extraction and preconcentration of cobalt, copper, and nickel after the formation of complex with 4‐benzylpiperidinedithiocarbamate potassium salt (K‐4‐BPDC), and later they were determined by flame atomic absorption spectrometry (FAAS) using (water/tetrabutylammonium ion (TBA+)/chloroform) as a ternary component system. The phase separation phenomenon occurred by an ion‐pair formation of TBA+ and perchlorate ion. After the optimization of complexation and extraction conditions ([K‐4‐BPDC]=2.0×10?4 mol l?1, [TBA+]=2.0×10?2 mol l?1, [CHCl3]=60.0 µl, [ClO4 ?]=2.0 ×10?2 mol l ?1 and pH=6.0), a preconcentration factor of 200 was obtained for only 10 ml of the sample.

The analytical curves were linear in the range of 20–1500, 15–2000, 35–1600 µg l?1 and the limits of detection were 10, 5, and 15 µg l?1 for Co2+, Cu2+, and Ni2+, respectively. The proposed method was applied for the extraction and determination of Co2+, Cu2+, and Ni2+ in natural water samples with satisfactory results.  相似文献   

18.
Heavy‐metal contamination is one of the most important environmental problems faced in the world, particularly in developing countries. Metals such as silver and mercury from drinking water, food, and air sources can accumulate in living organisms and present significant health concerns. Meanwhile, the demand for these metals in many industries continues to increase. In the present study, thioether‐functionalized corn oil (TFCO) from a photoinitiated thiol‐ene synthesis was utilized to remove Ag+ and Hg2+ ions from an aqueous solution. An aqueous solution containing AgNO3 and Hg[NO3]2 was prepared and contacted directly with TFCO. After vortex mixing for 60 s, the experiment ran for 351 min with the aqueous phase being periodically sampled for the analysis of metal ions (M n+). Results showed that 88.9% of Ag+ and 99.6% of Hg2+ ions were removed from the aqueous phase by the TFCO. Mass balances indicated that the total M n+ concentration in the oil phase was 13.890 g kg?1 under the conditions studied. TFCO exhibited higher selectivity for removing Hg2+ than for Ag+ ions. Analysis of the adsorption kinetics showed that a pseudosecond‐order model may be used to determine the rate of Ag+ ion sorption by the oil phase. The presence of the Hg2+ ions interfered with the adsorption of Ag+ ions from the aqueous solution.  相似文献   

19.
BACKGROUND: Thermodynamics and kinetics data are both important to explain the extraction property. In order to develop a novel separation technology superior to current extraction systems, many promising extractants have been developed including calixarene carboxylic acids. The extraction thermodynamics behavior of calix[4]arene carboxylic acids has been reported extensively. In this study, the mass transfer kinetics of neodymium(III) and the interfacial behavior of calix[4]arene carboxylic acid were investigated. RESULTS: The rate constant (Kao) becomes constant when the stirring speed was controlled between 250 rpm and 400 rpm. The activation energy (Ea) was calculated to be 21·41 kJ mol?1 or 88·17 kJ mol?1 (dependent on temperature) from the slope of log Kao against 1000/T. The linear relationship between the specific area and the extraction rate is the characteristic of an interfacial reaction control. The minimum bulk concentration of the extractant necessary to saturate the interface (Cmin) is lower than 4·19 × 10?4 mol L?1. CONCLUSION: The effect of stirring speed, temperature, and species concentration on the extraction rate demonstrates that the extraction regime depends on the extraction conditions. The chemical reaction control governs the extraction regime at temperatures below 303 K and a mixed control regime occurs when the temperature is between 303 K and 318 K. The probable locale for the chemical reaction is at the liquid–liquid interface and the rate equation is deduced to be: ? d[Nd3+](a)/dt = kf[Nd3+](a)[H4A](o)0·727[H+](a)?0·978. The rate‐controlling step was suggested by the analysis of the experimental results. Copyright © 2008 Society of Chemical Industry  相似文献   

20.
N-Methylpyrrole (N-MPy), 2,2′-bithiophene (BTh), and 3-(Octylthiophene) (OTh) were electrocopolymerized in 0.2 M NaClO4/CH3CN on glassy carbon electrode (GCE). The resulting terpolymers of N-MPy, BTh and OTh in different initial monomer feed ratios such as [N-MPy]0/[BTh]0/[OTh]0 = 1/1/1 and 1/2/5 were characterized by cyclic voltammetry (CV), Fourier-transform infrared attenuated total reflectance spectroscopy (FTIR-ATR), scanning electron microscopy (SEM), energy-dispersive X-ray analysis (EDX), and electrochemical impedance spectroscopy (EIS). The capacitive behaviors of the modified electrodes were defined via Nyquist, Bode-magnitude, Bode-phase, and Admittance plots. The equivalent circuit model of Rs(C dl1 (R 1 (QR 2 )))(C dl2 R 3 ) was performed to fit the theoretical and experimental data. The low-frequency capacitance (CLF) were obtained from initial monomer concentrations of 50 mM as CLF = ~2.34 × 10?4 mFcm?2 for P(N-MPy), CLF = 5.06 × 10?4 mF cm?2 for P(BTh), CLF = 5.07 m F cm?2 for P(OTh), and CLF = ~3.78 m Fcm?2 for terpolymer for [N-MPy]0/[BTh]0/[OTh]0 = 1/1/1. The terpolymer may be used as energy storage devices.  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