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 0 毫秒
1.
《Synthetic Metals》2006,156(7-8):610-613
Polypyrrole (PPy) nanowire modified electrodes were prepared electrochemically by template-free method based on graphite electrodes. The freshly prepared electrodes were dipped in 10% HClO4 solution at least 24 h for removal of carbonate ions. The modified electrodes toward ascorbic acid were characterized by potentiostatic method. The experiment's results show that the PPy modified electrodes have obvious electrocatalytic effect toward ascorbic acid oxidation. The oxidation current density has a good linearity in the concentration range of 5.0 × 10−4 and 2.0 × 10−2 mol L−1 of ascorbic acid. The determination sensitivity may be significantly affected by the thickness of PPy film and pH of the test solution. The method has promising application in determination of ascorbic acid in the real samples.  相似文献   

2.
《Synthetic Metals》2004,143(1):119-128
By combining the piezoelectric quartz crystal impedance (PQCI) and electrochemical impedance spectroscopy (EIS) measurements, we conducted a comparative study on polyaniline (PANI) degradation in different media, HClO4 and H2SO4, and on different piezoelectric quartz crystal (PQC) electrodes, Pt and Au. It is concluded that (1) the PANI film grown on an Au electrode is more stable than that on a Pt electrode; (2) the PANI degradation reaction abides by the zero-order kinetic law in HClO4 with rate constants from 0.17 to 2.25 Hz s−1 on Pt and 0.16 to 0.83 Hz s−1 on Au, but the first-order kinetic law in H2SO4 with rate constants from 2.61×10−3 to 7.00×10−3 s−1 on Pt and 1.00×10−3 to 5.00×10−3 s−1 on Au at different ion concentrations. The contrary effects of ClO4 and SO42− on PANI degradation may be understood from the Hofmeister series of anions; (3) the dissolution of the PANI film bulk, the increase of film porosity and the film attenuation occurred simultaneously during degradation via systematic analyses of the motional resistance (R1)and electrochemical impedance spectra responses, etc.  相似文献   

3.
Polypyrrole (PPy) and silver (Ag) nanorods are synthesized in cetyl trimethylammonium bromide–lauric acid (CTAB–LA) complex coacervate gel template. When PPy–CTAB–LA system is polymerized with AgNO3, Ag nanorods are produced while use of ammonium persulphate (APS) as initiator yields PPy nanorods. Ag-nanorods are produced from the initial stage while PPy nanorods take a longer time. The average diameter of Ag nanorods varies from 60 to 145 nm by increasing AgNO3 concentration from 0.27 M to 1.08 M and that of PPy varies from 145 nm to 345 nm by changing pyrrole concentration from 1 × 10?4 to 2 × 10?4 M, respectively. Fourier transformed infrared (FTIR) spectra indicate stabilization of Ag nanorods through complexation of PPy with adsorbed Ag+ ions. PPy nanoparticles are stabilized by adsorbed sulphate ions and lauric acid, both are acting as dopant to it. FFT pattern and EDX spectra clearly indicate the presence of Ag nanocrystals and PPy on the surface of Ag nanorods, respectively. The mechanism of nanorod formation is attributed from UV–Vis spectra showing a red shift of surface plasmon band of Ag and π–π* transition band of PPy with time. The highest dc conductivity of PPy–Ag composite is found to be 414.2 S/cm, 7 orders higher than that of PPy nanorods (9.3 × 10?4 S/cm). PPy–Ag systems show Ohmic behavior while PPy nanorods exhibit semi-conducting behavior. The preferential formation of Ag nanorod in AgNO3 initiated polymerization is attributed to the higher cohesive force of Ag than that of PPy. With two times higher LA and CTAB concentration in the gel the Ag nanorod diameter decreases only 12% while that of PPy nanorod decreases by 50%. Possible reasons are discussed from the hard and soft nature of the two nanorods and from the elasticity of the gel template.  相似文献   

4.
《Synthetic Metals》2002,128(3):283-287
We have studied photovoltaic cells using sexithiophene in (ITO/6T/Al) structure, and we have measured the action spectrum obtained by illumination through the aluminum side which is compared to the measured absorption spectrum. Two models were used to interpret this experiment: the Ghosh model using the carriers diffusion property in the bulk, and the kinetic model describing the dynamic behavior of the charge carriers in the device. The confrontation of these models with the experimental spectra allows us to reach the diffusion length of electron–hole pair L=2×10−6 cm. We also give the mobility of carriers (μn≈1.5×10−4 cm2/V s, and μp≈1.5×10−5 cm2/V s) and their diffusion coefficients (Dn≈4×10−6 cm2/s, and Dp≈4×10−7 cm2/s).  相似文献   

5.
Electrical characteristics of polypyrrole films electrodeposited in different aqueous electrolyte solutions including p-toluenesulfonate, naphtalenesulfonate, nitrate, tetrafluoroborate, and perchlorate anions were investigated using the Van der Pauw procedure. The polymer films were synthesized by electrochemical oxidation at a fixed potential. Experimental parameters including the pyrrole concentration, electrolyte, applied potential and substrate were shown to affect the electrical conductivity σ of polypyrrole films. Since the substrate contributes significantly to the overall conductivity of polypyrrole-coated electrodes, the results obtained with free standing polymer films appeared more reliable. The results indicated that the p-toluenesulfonate doped PPy film showed the highest average conductivity (σ293 K = 4.5 × 105 S m?1) whereas the perchlorate doped one produced the lowest of all the films prepared (σ293 K = 2 × 104 S m?1).  相似文献   

6.
《Synthetic Metals》2004,140(1):9-13
Two new charge-transfer (CT) salts: (ET)2·SO3CF3 and (ET)·ClO4 were prepared by chemical oxidation of ET with AgSO3CF3 and AgClO4, respectively. Their crystal structures were determined: monoclinic system, Cc, a=35.239(7) Å, b=6.5440(13) Å, c=14.646(3) Å, β=110.26(3)°, V=3168.5(11) Å3 for (ET)2·SO3CF3; monoclinic system, C2/c, a=15.981(3) Å, b=10.627(2) Å, c=11.495(2) Å, β=120.61(3)°, V=1680.2(6) Å3 for (ET)·ClO4. Electrical conductivity measurements indicate that (ET)2·SO3CF3 shows semi-conducting behaviour with room temperature conductivity of 0.29 S cm−1, while (ET)·ClO4 is an insulator.  相似文献   

7.
《Synthetic Metals》2006,156(2-4):351-355
A free-standing polypyrrole (PPy) film actuator, prepared electrochemically from a methyl benzoate solution of 1,2-dimethyl-3-propylimidazolium tris(trifluoromethylsulfonyl)methide (DMPIMe), exhibited up to 36.7% electrochemical strain in a propylene carbonate (PC)/water solution of lithium bis(nonafluorobutylsulfonyl)imide, Li(C4F9SO2)2N (LiNFSI). The maximum electrochemical strain of Me-doped PPy film depended on the electrolyte used for driving the Me-doped PPy actuator. When a PC/water mixed solution of lithium bis(trifluoromethylsulfonyl)imide (LiTFSI) was used as the driving electrolyte, the maximum electrochemical strain, measured by cycling between −0.9 and +0.7 V versus Ag/Ag+ at 2 mV s−1, was 24.2%, smaller than that (30.0%) driven with LiNFSI. When a PC/water suspension of DMPIMe was used as the driving electrolyte, the maximum electrochemical strain was 31.9%. However, the response speed of Me-doped PPy actuator driven with DMPIMe was slower than those driven with Li(CnF2n+1SO2)2N, due presumably to the size and shape of the anions. The addition of CF3COOH in electrolytic solutions for electropolymerization increased the maximum electrochemical strains (36.7% and 36.6%) of Me-doped PPy actuator driven by using a PC/water solution of LiNFSI and a PC solution of DMPIMe, respectively.  相似文献   

8.
《Synthetic Metals》2005,148(2):169-173
The field-effect mobility of holes in regioregular poly(3-alkylthiophene)s was determined for a series of 5 alkyl chain lengths from C4 (n-butyl) to C12 (n-dodecyl). Contrary to a previous report, a non-monotonic dependence of field-effect mobility on alkyl chain length was found. The average hole mobility varied from 1.2 × 10−3 cm2/Vs in poly(3-butylthiophene) and 1 × 10−2 cm2/Vs in poly(3-hexylthiophene) to 2.4 × 10−5 cm2/Vs for poly(3-dodecylthiophene). We believe that the hexyl side chain is optimum for charge transport because of better self-organization in poly(3-hexylthiophene) compared to other polymers in the series. The present results provide an important structure–carrier mobility relationship for the regioregular poly(3-alkylthiophene)s which are of wide interest for thin film transistors and photovoltaic cells.  相似文献   

9.
《Synthetic Metals》2002,130(3):221-227
The effect of protonation of the pyrrolidine ring nitrogen of 2-(n-alkyl)fulleropyrrolidines, C60pyr–Cm (m=4, 6, 8, 10 and 12), on the properties of the Langmuir and Langmuir–Blodgett (LB) films was investigated. The isotherms of both surface pressure (π) and surface potential change (ΔV) versus area per molecule (A) for the Langmuir films of C60pyr–Cm were determined simultaneously. It was found that the longer the alkyl chain of the fulleropyrrolidine the larger is the film compressibility, κ, i.e., κ(C60pyr–C4)=(2.1±0.4)×10−2 m mN−1, κ(C60pyr–C8)=(3.5±0.4)×10−2 m mN−1 and κ(C60pyr–C12)=(4.1±0.5)×10−2 m mN−1, as expected for the liquid surface films. The values of surface area at zero surface pressure (A1) differ in the range 0.6 nm2 molecule−1 (for m=4–8) to 1.4 nm2 molecule−1 (for m=10–12), indicating that all 2-(n-alkyl)fulleropyrrolidines form multilayer or aggregated films on neutral water subphase. However, acidification of the water subphase increases the A1 values for all investigated fulleropyrrolidines up to ca. 1.9 nm2 molecule−1, i.e., the value corresponding to maximum area occupied by a fulleropyrrolidine at horizontal orientation in a monolayer film. Apparently, 2-(n-alkyl)fulleropyrrolidinium cations formed at low pH are markedly de-aggregated in the films and their orientation is changed due to protonation of the pyrrolidine nitrogen. Similarly, however, less pronounced effects are observed if ionic strength of the subphase solution, I, is increased in the range 0≤I≤1.0 mol dm−3 NaCl. The Langmuir films formed on a water subphase were the most stable with respect to the LB transfer onto 5 MHz Au–quartz crystal vibrators. Simultaneous cyclic voltammetry and piezoelectric microgravimetry at an electrochemical quartz crystal microbalance of these films showed at least two electroreductions where the fulleropyrrolidine mono anions were stable with respect to dissolution.  相似文献   

10.
《Acta Materialia》2007,55(18):6182-6191
High-temperature oxidation and hot corrosion behaviors of Cr2AlC were investigated at 800–1300 °C in air. Thermogravimetric–differential scanning calorimetric test revealed that the starting oxidation temperature for Cr2AlC is about 800 °C, which is 400 °C higher than other ternary transition metal aluminum carbides. Thermogravimetric analyses demonstrated that Cr2AlC displayed excellent high-temperature oxidation resistance with parabolic rate constants of 1.08 × 10−12 and 2.96 × 10−9 kg2 m−4 s−1 at 800 and 1300 °C, respectively. Moreover, Cr2AlC exhibited exceptionally good hot corrosion resistance against molten Na2SO4 salt. The mechanism of the excellent high-temperature corrosion resistance for Cr2AlC can be attributed to the formation of a protective Al2O3-rich scale during both the high-temperature oxidation and hot corrosion processes.  相似文献   

11.
《Synthetic Metals》2006,156(5-6):444-453
Electropolymerization of aniline in sulfuric acid solution in the presence of o-phenylenediamine (oPD) of various concentrations was investigated via the electrochemical quartz crystal microbalance (EQCM) technique. It was found that the polymerization occurred more favorably at high aniline-to-oPD molar ratios (F1, 20 or above). The stabilities of the resultant copolymers against degradation were efficiently improved compared with that of polyaniline (PANI). The first-order kinetic constants for polymer degradation were estimated to be 2.07 × 10−3 s−1 for polyaniline, and 3.91 × 10−4 and 1.28 × 10−4 s−1 for copolymers with F1 values of 50 and 20, respectively. The degradation product, benzoquinone, was also detected at the tip electrode of a scanning electrochemical microscope (SECM).  相似文献   

12.
《Intermetallics》2007,15(3):241-244
The coefficients of thermal expansion (CTE) of the W5Si3 and T2 phases of the W–Si–B system were determined using high-temperature X-ray diffraction in the 298–1273 K temperature interval. Alloys with nominal compositions 62.5W37.5Si (at%) and 58W21Si21B (at%) were prepared from high-purity materials through arc melting followed by heat treatment at 2073 K for 12 h under argon atmosphere. The highly different thermal expansion coefficients of W5Si3 along the a (5.0 × 10−6 K−1) and c (16.3 × 10−6 K−1) axes lead to a high thermal expansion anisotropy (αc/αa  3.3). On the other hand, the T2-phase exhibits similar thermal expansion coefficients along the a (6.9 × 10−6 K−1) and c (7.6 × 10−6 K−1) axes, indicating a behavior close to isotropic (αc/αa  1.1).  相似文献   

13.
《Synthetic Metals》2006,156(5-6):420-425
Chemically synthesized polypyrroles of low (σ < 75 S/cm), medium (75 < σ < 200 S/cm) and high (σ > 200 S/cm) electrical conductivity (σ) with the same dopant and degree of doping have been investigated by means of Wide Angle X-ray Scattering (WAXS), 13C Cross Polarized Magic Angle Spinning Nuclear Magnetic Resonance (13C CP/MAS NMR) spectroscopy and Fourier Transform Infrared (FTIR) Spectroscopy to establish structure–conductivity relationships useful for industrial applications. A similar amorphous structure was found by WAXS even for the higher conducting PPy (σ = 288 S/cm). WAXS spectra for polypyrroles of medium and high conductivity showed a weak peak at 2θ = 10–11° due to improved order of the counterions in these materials. The effect of the counterion size in the asymmetry of the PPy main WAXS peak was elucidated by performing ion exchange of the Cl dopant with counterions of larger size such as BF4 and ClO4. From 13C CP/MAS NMR measurements predominantly α–α′ bonding was found in these materials. The main 13C CP/MAS NMR resonance peak of PPy located at 126–128 ppm was broadened upon increasing conductivity. Interestingly, a linear relationship was observed between the half-width at half-height (HWHH) of the 13C CP/MAS NMR peak and conductivity where a doubling of the polypyrrole conductivity leads to an increase of HWHH by 6–7 ppm. FTIR data of these materials were analysed in the framework of the Baughman–Shacklette theory describing the dependence of conductivity on conjugation length. By comparison of model predictions and experimental results, the PPy samples were found to be in the regime of long conjugation lengths, L  K2/kBT, where K2 is a parameter related to the energy change on going from j  1 to j charges on a conjugated segment of conjugation length L, kB the Boltzman constant and T is the absolute temperature.  相似文献   

14.
《Acta Materialia》2008,56(8):1857-1867
Chromium, a p-type dopant, has been incorporated into silicon carbide by laser doping. Secondary ion mass spectrometric data revealed enhanced solid solubility (2.29 × 1019 cm−3 in 6H–SiC and 1.42 × 1919 cm−3 in 4H–SiC), exceeding the equilibrium limit (3 × 1017 cm−3 in 6H–SiC above 2500 °C). The roughness, surface chemistry and crystalline integrity of the doped sample were examined by optical interferometry, energy dispersive X-ray spectrometry and transmission electron microscopy, respectively, and showed no crystalline disorder due to laser heating. Deep-level transient spectroscopy confirmed Cr as a deep-level acceptor with activation energies Ev + 0.80 eV in 4H–SiC and Ev + 0.45 eV in 6H–SiC. The Hall effect measurements showed that the hole concentration (1.942 × 1019 cm−3) is almost twice the average Cr concentration (1 × 1019 cm−3), confirming that almost all of the Cr atoms were completely activated to the double acceptor state by the laser-doping process without requiring any additional annealing step.  相似文献   

15.
We report the role of anthraquinone sulfonate dopants in promoting performance of electro-synthesized polypyrrole (PPy) composites for use in electrochemical supercapacitors. The incorporation of anthraquinone sulfonate species into the polymer matrix can significantly improve the surface area of PPy composites that are composed of submicron-/nano-sized particles, as evidenced from scanning electron microscopy (SEM) results. Cyclic voltammetry and galvanostatic charge–discharge measurements in 1 M KCl solution reveal that these dopants result in an improved specific capacitance, a wide working potential range and enhanced long-cycle stability as compared to ClO4? dopant. Among the samples investigated, the resulting PPy/AQS (9,10-anthraquinone-2-sulfonic acid sodium salt) composite exhibits the highest specific capacitance of 608 F g?1 at a scan rate of 5 mV s?1 within a potential range between ?0.9 and 0.5 V (vs. saturated calomel electrode, SCE).  相似文献   

16.
《Intermetallics》2006,14(10-11):1339-1344
The effect of growth rate on microstructure and mechanical properties of directionally solidified (DS) multiphase intermetallic alloy with the chemical composition Ni–21.9Al–8.1Cr–4.2Ta–0.9Mo–0.3Zr (at.%) was studied. The DS ingots were prepared at constant growth rates V ranging from 5.56 × 10−6 to 1.18 × 10−4 ms−1 and at a constant temperature gradient at the solid–liquid interface of GL = 12 × 103 K m−1. Increasing growth rate increases volume fraction of dendrites and decreases primary dendritic arm spacing, mean diameter of α-Cr (Cr-based solid solution) and γ′(Ni3Al) precipitates within the dendrites. Room-temperature compressive yield strength, ultimate compressive strength, hardness and microhardness of dendrites increase with increasing growth rate. All room-temperature tensile specimens show brittle fracture without yielding. The brittle-to-ductile transition temperature for tensile specimens is determined to be about 1148 K. Minimum creep rate is found to depend strongly on the applied stress and temperature according to the power law with a stress exponent of n = 7 and apparent activation energy for creep of Qa = 401 kJ/mol.  相似文献   

17.
《Synthetic Metals》2007,157(22-23):940-944
Hole transport polymers consisting of 4-methoxytriphenylamine (P-MOTPA) and 4-n-butyltriphenylamine (P-BTPA) were synthesized by oxidative coupling reaction using FeCl3 as an oxidant. These polymers had good solubility and their thin films showed sufficient morphological stability. 1H NMR and 13C NMR revealed that the monomers were exclusively connected at the p-positions of unsubstituted phenyl groups. Polymers had UV absorption maximum around 370 nm. Cyclic voltammograms of polymers showed well-defined pairs of reduction and oxidation peaks at 1.1 V versus Ag/AgCl, indicating that the polymers are electrochemically active. P-MOTPA and P-BTPA had hole drift mobility of being 4.04 × 10−5 and 2.98 × 10−5 cm2/V s at the electric field of 50 V/μm, respectively.  相似文献   

18.
《Synthetic Metals》2007,157(2-3):80-90
The synthesis of a hybrid material obtained by electropolymerization of a solution of pyrrole and [NEt4]2[Ni(dmit)2] (dmit = 1,3-dithiole-2-thione-4,5-dithiolato) in acetonitrile solution is reported. The material was characterized by cyclic voltammetry, UV–vis and infrared spectroscopies, scanning electron microscopy (SEM), atomic force microscopy (AFM) and thermal gravimetric analysis (TGA). The FTIR spectroscopy showed that the [Ni(dmit)2]2− anion has been inserted in the polypyrrole framework and was not destroyed or modified during the polymerization process. The voltammetric analysis indicated that the material has electroactivity and undergoes redox processes associated with the conducting polymer and the counteranion. The cyclic voltammetry results also suggest that the counteranion is not trapped in the PPy matrix undergoing anion exchange during the redox cycle of PPy. The PPy/[Ni(dmit)2]2− exhibits good thermal stability and a intrinsic conductivity value in the range of semiconductors (10−3 S cm−1).  相似文献   

19.
《Synthetic Metals》2001,124(1):217-219
Poly(phenylacetylene) (PPA) and poly(p-methoxyphenylacetylene) (PMOPPA) were synthesized by catalytic polymerization of monosubstituted alkynes and were fully characterized by conventional techniques. Linear and third-order non-linear optical properties of polymers solutions were investigated at λ=780 nm using a Z-scan set-up equipped with a femtosecond laser source at different operating regimes. A high repetition rate (76 MHz) regime was used to measure the linear and non-linear absorption coefficients by means of thermo-optical effects. Values of the linear and non-linear absorption coefficients were α=4.9×10−3 cm−1 and β=1.1×10−11 cm W−1 for PPA and α=2×10−2 cm−1 and β=2.1×10−10 cm W−1 for PMOPPA. Very low repetition rate (14 Hz) excitation was used to evidence the purely-optical non-linear refractive indexes, that were γ=6×10−18 cm2 W−1 and γ=11×10−18 cm2 W−1, respectively, for PPA and PMOPPA.  相似文献   

20.
《Synthetic Metals》2005,155(3):648-651
The self-doping mechanism for charge transport is investigated in layer-by-layer (LBL) films from two conducting polymers, namely poly(o-methoxyaniline) (POMA) and poly(3-thiophene acetic acid) (PTAA). The efficiency of charge intercalation, defined as the ratio between the charge and the mass change, is twice for the POMA/PTAA LBL film in comparison with a cast POMA film. This is attributed to differences in the diffusion-controlled charge and mass transport, where distinct ionic species participate in the LBL films, as demonstrated with experiments using a quartz crystal microbalance. The doping efficiency for LBL film is the same, i.e., 3.93 × 10−4 and 3.56 × 10−4 g/C for the Li+ and (C2H5)4N+ doped films, and is different for the cast POMA film, i.e., 11.3 × 10−4 for Li+ and 6.45 × 10−4 g/C for (C2H5)4N+. Therefore, once no significant differences in the intercalation mechanism are observed when different cations, Li+ or (C2H5)4N+, are used with the LBL films, this indicates that the self-doping mechanism is controlled by the exchange of anions.  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