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 31 毫秒
1.
Impedance measurements are made on purified, vacancy-doped and interstitial-doped silver bromide crystals of thicknesses from 0.061 cm to 0.260 cm using constant ionic strength solution contacts at temperatures from 20°C to 35°C. Results characterize the macroscopic transport properties from 316 kHz to 0.01 Hz and at d.c. In all cases the high frequency limit resistive component of the impedance, R, is equal to the d.c. resistance R0. This result indicates rapid ion exchange at the surfaces and neither Warburg diffusion nor surface kinetic-limitation of transport. As expected, R is a function of charge carrier concentration, which depends on extrinsic dopant concentration over the temperature range measured. For pure crystals the resistance, R, is linearly dependent on crystal thickness, δ, while thickness correlation could not be tested for the doped crystals as dopant level was not uniform. The geometric capacitance for all crystals, Cg, is linear with δ?1. The dimensionless dielectric constant, κ, calculated from Cg and crystal dimentsions, is 13.9±0.7. The electric relaxation time, τel, defined as the product RCg, at 25°C is found to be 50±3 μs for pure crystals, 1 to 4 μs for the Cd2+-doped crystals and 12 to 66 μs for the S2?-doped crystals. Temperature dependance of R allows determination of crystal transport activation energies. For pure and Cd2+-doped crystals a value of 0.33±0.02 eV is found. For the S2+-doped crystals values from 0.48 to 0.28 eV are found. The parameter α, in the non-ideal Cole-Cole representation of impedance plane arcs, is 0.96±0.01 for purified crystals, 0.93±0.03 for Cd2+-doped, and 0.92±0.02 for S2?-doped crystals.  相似文献   

2.
Dielectric polarization of solutions of un-ionized linear poly(methacrylic acid) in polar associated liquids is studied in the temperature range 20–50°C. The solutions are in methanol, with the molar fraction of polymer units x 2 = (3 × 10?3)?(1.5 × 10?2), and in water, with x 2 = (4 × 10?5)?(4 × 10?3). The permittivity ε12 of the polyacid solutions in methanol is shown to be lower than the permittivity of the pure solvent ε1; the permittivity of the polyacid solutions in water exceeds ε1 of water in the concentration range x 2 = (4 × 10?5)?(2.13 × 10?4) and becomes lower than ε1 as the polymer concentration in the solution increases further. A procedure for estimating the dipole moment μ2 per monomer unit of the polymer macromolecule in solution is proposed. The estimation is based on Buckingham’s statistical polarization theory for a two-component mixture of polar molecules under the conditions of infinite dilution. The μ2 values amount to 2.76–2.14 D (x 2 < 1.5 × 10?2) in methanol at 20–50°C and to 11.4?3.8 D (x 2 < 2.13 × 10?4) in water at 20–40°C. The difference in the dipole moments of the polyacid and in the patterns of their temperature dependences in methanol and in water is due to the effects of the polyacid-solvent hydrogen bonding, to intramacromolecular hydrogen bonds, and to specificity of the local structure of the solvent. It is shown that the μ2 value corresponds to the dipole moment of the solvates and decreases with temperature owing to changes in the stoichiometry of the solvates, to the formation of cyclic associates in the macromolecule, and to conformational changes in the chain.  相似文献   

3.
We demonstrate that an ordered 2D perovskite can significantly boost the photoelectric performance of 2D/3D perovskite heterostructures. Using selective fluorination of phenyl-ethyl ammonium (PEA) lead iodide to passivate 3D FA0.8Cs0.2PbI3, we find that the 2D/3D perovskite heterostructures passivated by a higher ordered 2D perovskite have lower Urbach energy, yielding a remarkable increase in photoluminescence (PL) intensity, PL lifetime, charge-carrier mobilities (ϕμ), and carrier diffusion length (LD) for a certain 2D perovskite content. High performance with an ultralong PL lifetime of ≈1.3 μs, high ϕμ of ≈18.56 cm2 V−1 s−1, and long LD of ≈7.85 μm is achieved in the 2D/3D films when passivated by 16.67 % para-fluoro-PEA2PbI4. This carrier diffusion length is comparable to that of some perovskite single crystals (>5 μm). These findings provide key missing information on how the organic cations of 2D perovskites influence the performance of 2D/3D perovskite heterostructures.  相似文献   

4.
Perovskite single crystals and polycrystalline films have complementary merits and deficiencies in X-ray detection and imaging. Herein, we report preparation of dense and smooth perovskite microcrystalline films with both merits of single crystals and polycrystalline films through polycrystal-induced growth and hot-pressing treatment (HPT). Utilizing polycrystalline films as seeds, multi-inch-sized microcrystalline films can be in situ grown on diverse substrates with maximum grain size reaching 100 μm, which endows the microcrystalline films with comparable carrier mobility-lifetime (μτ) product as single crystals. As a result, self-powered X-ray detectors with impressive sensitivity of 6.1×104 μC Gyair−1 cm−2 and low detection limit of 1.5 nGyair s−1 are achieved, leading to high-contrast X-ray imaging at an ultra-low dose rate of 67 nGyair s−1. Combining with the fast response speed (186 μs), this work may contribute to the development of perovskite-based low-dose X-ray imaging.  相似文献   

5.
Nonlinear optical susceptibility χ(3) of polydiacetylenes has been evaluated by third harmonic generation. First, in order to obtain the χ(3) tensor component along the polymer main chain, thin single crystals of two representative polydiacetylenes, poly-PTS and poly-DCHD were made by utilization of sampling technique in microscopy, i.e., the microtome cutting of single crystal embedded in plastic resin. The THG intensity observed was proportional to cos8θ, where θ was the angle between the polymer main chain and the polarization of laser light. The χ(3) value of poly-DCHD was found to be rather higher than that of poly-PTS. At resonant wavelength of 1.97 μm, the χ(3) of poly-DCHD attained 8 x 10?10 esu. Furthermore, it was confirmed that when geometrical correction were properly made, the χ(3) obtained from polycrystalline thin film of poly-PTS agreed well with that from thin single crystals. Secondary, regarding poly-diphenyldiacetylene derivatives, it was found that the π-conjugation between the polymer main chain and aromatic substituents was effective on the improvement of χ(3) values. The χ(3) magnitudes of poly-BTFP and poly-DFMP reflect well the dihedral angles between polymer main chain and the phenyl substituents (58° for poly-BTFP and 67° for poly-DFMP) as a measure of π-conjugation. Especially, at nonresonant region of 2.1 μm the χ(3) of poly-BTFP is about 5 times greater than that of poly-PTS.  相似文献   

6.
Two model polymers, containing fluorene as an electron‐donating moiety and benzothiadiazole (BT) as an electron‐accepting moiety, have been synthesized by Suzuki coupling reaction. Both polymers are composed of the same chemical composition, but the BT acceptor can be either at a side‐chain (i.e., S‐polymer) or along the polymer main chain (i.e., M‐polymer). Their optical, electrochemical, and photovoltaic properties, together with the field‐effect transistor (FET) characteristics, have been investigated experimentally and theoretically. The FET carrier mobilities were estimated to be 5.20 × 10?5 and 3.12 × 10?4 cm2 V?1 s?1 for the S‐polymer and M‐polymer, respectively. Furthermore, polymeric solar cells (PSCs) with the ITO/PEDOT:PSS/S‐polymer or M‐polymer:PC71BM(1:4)/Al structure were constructed and demonstrated to show a power conversion efficiency of 0.82 and 1.24% for the S‐polymer and M‐polymer, respectively. The observed superior device performances for the M‐polymer in both FET and PSCs are attributable to its relatively low band‐gap and close molecular packing for efficient solar light harvesting and charge transport. This study provides important insights into the design of ideal structure–property relationships for conjugate polymers in FETs and PSCs. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

7.
The electrochemical deposition of copper in reduced polyaniline (PAN) films is studied at different polymer layer thicknesses. A saturation in the number density, n0, of active sites for metal deposition is found at 0.5 μm PAN thickness (n0=108 cm−2). The surface spatial distribution of the deposited copper crystals is analysed in two cases. It is found that nucleation exclusion zones play an appreciable role in the course of the deposition process in thin PAN films. In contrast, a random distribution and a larger number of deposited crystals are obtained in thicker PAN layers.  相似文献   

8.
Tanja Djekic 《Adsorption》2007,13(3-4):231-237
The scope of this work is to determine the effective intraparticle diffusion coefficient of CoCl2 over mesoporous functionalized silica. Silica is selected as a carrier of the functionalized groups for its rigid structure which excludes troublesome swelling, often found in polymeric adsorbents. 2-(2-pyridyl)ethyl-functionalized silica is selected as a promising affinity adsorbent for the reversible adsorption of CoCl2. The adsorption kinetics is investigated with the Zero Length Column (ZLC) method. Initially, experiments were performed at different flow rates to eliminate the effect of external mass transfer. The effect of pore size (60 Å and 90 Å), particle size (40?10?6 m–1000?10?6 m) and initial CoCl2 concentration (1 mol/m3–2.0 mol/m3) on the mass transfer was investigated. A model was developed to determine the pore diffusion coefficient of CoCl2 by fitting the experimental data to the model. The pore diffusion coefficients determined for two different pore sizes of silica are D p (60 Å) =1.95?10?10 [m2/s] and D p (90 Å) =5.8?10?10 [m2/s]. The particle size and the initial CoCl2 concentration do not have an influence on the value of diffusion coefficient. However, particle size has an influence on the diffusion time constant. In comparison with polymer adsorbents, silica based adsorbents have higher values of diffusion coefficients, as well as a more uniform and stable pore structure.  相似文献   

9.
To obtain novel low‐bandgap materials with tailored hole‐transport properties and extended absorption, electron rich 3,4‐ethylenedioxythiophene is introduced as a comonomer in diketopyrrolo[3,4‐c]pyrrole copolymers with different aryl flanking units. The polymers are characterized by absorption and photoluminescence spectroscopy, dynamic scanning calorimetry, cyclic voltammetry, and X‐ray diffraction. The charge transport properties of these new materials are studied carefully using an organic field effect transistor geometry where the charge carriers are transported over a narrow channel at the semiconductor/dielectric interface. These results are compared to bulk charge carrier mobilities using space‐charge limited current (SCLC) measurements, in which the charge carrier is transported through the complete film thickness of several hundred nanometers. Finally, charge carrier mobilities are correlated with the electronic structure of the compounds. We find that in particular the thiophene‐flanked copolymer PDPP[T]2‐EDOT is a very promising candidate for organic photovoltaics, showing an absorption response in the near infrared region with an optical bandgap of 1.15 eV and a very high bulk hole mobility of 2.9 × 10?4 cm2 V?1 s?1 as measured by SCLC. This value is two orders of magnitudes higher than SCLC mobilities reported for other polydiketopyrrolopyrroles and is in the range of the well‐known hole transporting polymer poly(3‐hexylthiophene). © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2016 , 54, 639–648  相似文献   

10.
Poly((2‐Alkylbenzo[1,2,3]triazole‐4,7‐diyl)vinylene)s (pBTzVs) synthesized by Stille coupling show different absorption spectra, solid‐state morphology, and photovoltaic performance, depending on straight‐chain versus branched‐chain (pBTzV12 and pBTzV20) pendant substitution. Periodic boundary condition density functional computations show limited alkyl pendant effects on isolated chain electronic properties; however, pendants could influence polymer backbone conjugative planarity and polymer solid film packing. The polymers are electronically ambipolar, with best performance by pBTzV12 with hole and electron transport mobilities of 4.86 × 10?6 and 1.96 × 10?6 cm2 V?1 s?1, respectively. pBTzV12 gives a smooth film morphology, whereas pBTzV20 gives a very different fibrillar morphology. For ITO/PEDOT:PSS/(1:1 w/w polymer:PC71BM)/LiF/Al devices, pBTzV12 gives power conversion efficiency (PCE) up to 2.87%, and pBTzV20 gives up to PCE = 1.40%; both have open‐circuit voltages of VOC = 0.6–0.7 V. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2015 , 53, 1539–1545  相似文献   

11.
The solid-state synthesis and properties are reported for a new polydiacetylene: poly[1,6-di(N-carbazolyl)-2,4-hexadiyne]. The monomer crystals polymerize quantitatively with γ irradiation or thermal annealing. An Autocatalytic effect is observed in both γ-ray polymerization and thermal polymerization and is attributed to an increase in chain propagation length at about 5% conversion. The activation energy for thermal polymerization is about 25 kcal/mole, independent of the degree of conversion to polymer. The exceptional thermal stability of the polymer crystals allowed a thermomechanical analysis over a large temperature range, ?50 to 300°C. With increasing temperature, the polymer contracts in the chain direction linearly with temperature over the entire range, yielding a thermal expansion coefficient of (?2.32 ± 0.02) × 10?5°C?1. Photoconductivity action spectra are reported for the polymer crystals. The energies for the photoconductivity onset (ca. 2.3 eV) and for the lowest energy optical transition (1.89 eV) are the lowest reported for the polydiacetylenes. The photoconduction onset is blue-shifted with respect to optical absorption—a result which is consistent with the excitonic assignment for the lowest energy optical transition in the polydiacetylenes.  相似文献   

12.
A solid-state reactor for detection of hydrogen peroxide in aqueous samples by peroxyoxalate chemiluminescence is described. Bis(2,4,6-trichlorophenyl)oxalate in solid form is packed into a bed reactor, which eliminates mixing problems and facilitates the instrumental development. Perylene is added as a sensitizer to a water/acetonitrile (20:80) carrier stream into which the samples (200–600 μl) are injected. Detection limits of 6 × 10?9 M H2O2 (0.2 μg l?1) are obtained with both a commercial and a home-made luminescence detector. Calibration graphs are linear up to 10?5 M. The r.s.d. for 2 × 10?7 M (6.7 μg?1) hydrogen peroxide (n = 10) is 2.8%. Sample throughput is ca. 120 h?1.  相似文献   

13.
The crystal structure of trimethyltin glycinate has been determined by Patterson and Fourier techniques to a final “R”-factor of 0.066 for 1103 unique reflections. The crystals are tetragonal with space group P41with a = b = 7.839(9) and c = 14.659(11) Å, Z = 4, and are composed of stacks of linearly polymeric trimethyltin glycinate molecules bridled axially at tin through the amino nitrogen atoms of the amino acid. There is hydrogen bonding between carbonyl oxygen and amino group NH moieties along the chains and between the chains to produce a perpendicular weave of one-dimensional polymer threads. The axial NSnO connections make an angle approaching linearity, but the tin atom is distinctly displaced toward the oxygen to give non-planar SnC3 units which are eclipsed in the
chain. This bridging rather than chelated amino acid configuration is only found in glycinatosilver(I) hemihydrate. Tin-119m Mössbauer resonance area data have been collected in the temperature range 77 ? T ? 185 K, and a logarithmic plot of the normalized area vs. temperature is linear in this range and the slope of ?1.15 x 10?2 K?1 yields values of the logarithmic tempemture coefficient of the recoil-free fraction (which at 296 is less than a tenth of its value at 77 K). These data have been used to evaluate the isotropic mean-square amplitudes of vibration <χiso(T)2> of the tin atom, normalized to the value at 296 K available from the crystallographic study, in the same temperature range. Isotropic-mean-square values of the vibrational amplitude increase from 1.78 x 10?2 at 77 to 2.63 x 10?2 at 185 and 3.5 x 10?2 at 296 K. From the doublet line asymmetry data assuming that the electric field gradient tensor, Vzz, is positive in sign and lies along the polymer axis (oblate field about the cylindrical axis), it is concluded that the tin atom vibrates with greater amplitude normal to the propagating axis (<χ2> = 1.93 x 10?2 at 77 and 3.11 x 10?2 Å2 at 185 K) than along it (<χ2> = 1.48 x 10?2 at 77 and 1.67 x 10?2 Å2 at 185 K). The temperature coefficient of the motion normal to the axis is also greater over the range examined. The differrence in the mean-square amplitudes perpendicular and parallel to the axis 2.96 x 10?2 from the Mössbauer treatment and 4.24 x 10?2 Å2 from the X-ray anisotropic thermal ellipsoids at 296 K.  相似文献   

14.
Perfluoro-1,3,5-tris(p-quaterphenyl)benzene (PF-13Y) and perfluoro-1,3,5-tris(p-quinquephenyl)benzene (PF-16Y) have been synthesized and characterized. They showed higher glass transition temperatures compared with perfluoro-1,3,5-tris(p-terphenyl)benzene (PF-10Y). Organic light-emitting diodes were fabricated using these materials as the electron-transport layers. PF-13Y and -16Y are better electron transporters than PF-10Y. The electron mobilities of PF-10Y and Alq3 were measured by the time-of-flight technique. PF-10Y showed higher electron mobilities (10−4 cm2/V s) and weaker electric field dependence compared with Alq3.  相似文献   

15.
The interaction of Fe2O3 nanoparticles emphasized between poly(propylene glycol) (PPG 4000) and silver triflate (AgCF3SO3) on the conformal changes of coordination sites and the electrochemical properties have been investigated. On the influence of Fe2O3 nanoparticles distribution, the interactions between the ether oxygen in C–O–C of the polymer chain with Ag+ ion as a result of bond strength of the C–O–C stretching vibration, the end group effect has been examined by Fourier transform infrared (FT-IR) spectroscopy. The formation of transient cross-links between polymer chains and filler particles appears to be a characteristic change in the glass transition temperature (T g) and enhance the effective number of cations as well. The strength of ion–polymer interactions was revealed by the transport of ions, t Ag+, and found to be in the range of 0.42–0.50, and the ionic conductivity was ascertained by complex impedance analysis with a maximum of 9.2?×?10?4 S cm?1 at 298 K with a corresponding concentration of 10 wt% Fe2O3 nanoparticles. The temperature dependence of conductivity has been examined based on the Vogel–Tammann–Fulcher (VTF) equation, thereby suggesting the segmental chain motion and free volume changes. From the impedance data, both the dielectric and modulus behaviours have been revealed and both were well correlated as a function of frequency.  相似文献   

16.
Hybrid bismuth halides perovskites have emerged as promising candidates for X-ray detection, due to the strong absorptivity of high-energy X-ray photons, high resistivity, large carrier diffusion length and low toxicity. However, the mostly investigated hybrid bismuth iodides single crystals are usually opaque and require a harsh synthesis process. Herein, novel one-dimensional (1D) pentamethylenediamine bismuth bromide (PDA)BiBr5 single crystals were synthesized via an antisolvent-assisted crystallization method at room temperature. Bulk (PDA)BiBr5 single crystals have sizes of 10×1.3×1.5 mm3 and high transparency. They are shown to have low density of defects of 2.0×1010 cm−3 and obvious photoconductivity. Moreover, they exhibit large bulk resistivity of 2.13×1011 Ω cm and good X-ray attenuation coefficient. Consequently, the vertical structured (PDA)BiBr5 single crystal X-ray photoconductor produces a sensitivity of 3.8 μC Gyair−1 cm−2. This study provides a facile strategy for synthesizing bulk hybrid bismuth bromides single crystals with potential X-ray detection application.  相似文献   

17.
Thermoelectric properties of polycrystalline La1−xSrxCoO3, where Sr2+ is substituted in La3+ site in perovskite-type LaCoO3, have been investigated. Sr-doping increases the electrical conductivity (σ) of La1−xSrxCoO3, and also decreases the Seebeck coefficient (S) for 0.01?x?0.40. A Hall coefficient measurement reveals that the increase in electrical conductivity arises from increases in both carrier concentration and the Hall mobility. The decrease in the Seebeck coefficient is caused by a decrease in carrier effective mass as well as increase in carrier concentration. The highest power factor (σS2) is 3.7×10−4 W m−1 K−2 at 250 K for x=0.10. The thermal conductivity (κ) is about 2 W m−1 K−1 at 300 K for 0?x?0.04, and increases for x?0.05 because of an increase in heat transport by conductive carrier. The thermoelectric properties of La1−xSrxCoO3 are improved by Sr-doping, and the figure of merit (Z=σS2 κ−1) reaches 1.6×10−4 K−1 for x=0.06 at 300 K (ZT=0.048). For heavily Sr-doped samples, the thermoelectric properties diminish mainly because of the decrease in the Seebeck coefficient and the increase in thermal conductivity.  相似文献   

18.
Poly-p-phenylene (PPP), synthesized chemically by the Kovacič method, was mixed with 7.5% carbon black and used as a pressed thin-layer electrode (100 μm) on Pt mesh. Its overoxidation as an anode in aqueous acid electrolytes (8–18 M H2SO4, 8–11.3 M HClO4, 8 M HBF4) was investigated systematically.Slow cyclic voltammetry reveals that the reversible redox peak 1, leading to an insertion compound up to 1 F per mole (C6H4)6, is following by a large second peak 2 at more positive potentials, corresponding to about 15 F with respect to the unit mentioned above. Peak separation increases with increasing acid concentration due to a negative shift of the insertion potential (peak 1) by 60 mV mol−1 dm3, but normal Nernst behaviour is observed in the case of peak 2. The strong decay of the current efficiency for reversible cycling of PPP at concentrations below 10 M can be understood readily from these findings. The overoxidation starts with the formation of the polymer radical cation, which is oxidized further to the quinone, then to the product of ring opening and finally to the product of cleavage of the polymer chain. The final product seems to be a derivative of maleic acid. The p-quinone path is preferred over the o-quinone path, quite contrary to the case with graphite. The practical consequences with respect to battery application are discussed in detail.  相似文献   

19.
The synthesis of the nickel dialkynyl complex Ni(C?C? C6H4? C?CH)2(PPh3)2 and of the corresponding polyyne polymer containing nickel in the main chain ? [Ni(PPh3)2? C?C? C6H4? C?C? ]n are described and discussed. A new mixed solvent system DMSO/HNEt2 and homogeneous step-wise condensation method used for their synthesis are presented for the first time. The Ni-polyyne polymer obtained is dark yellow powder and soluble in THF or CH2Cl2. Its M?w is about 104, and the MWD is less than 2. Both the prepared complex and polymer have been characterized by IR, UV, 1H-NMR, and DTA. Preliminary results on photoluminesence of nickel polyyne polymers are present. © 1995 John Wiley & Sons, Inc.  相似文献   

20.
New dithienothiophene‐containing conjugated polymers, such as poly(2,6‐bis(2‐thiophenyl‐3‐dodecylthiophene‐2‐yl)dithieno[3,2‐b;2′,3′‐d]thiophene, 4 and poly(2,6‐bis (2‐thiophenyl‐4‐dodecylthiophene‐2‐yl)dithieno[3,2‐b;2′,3′‐d]thiophene, 8 have been successfully synthesized via Stille coupling reactions using dodecyl‐substituted thiophene‐based monomers, bistributyltin dithienothiophene, and bistributyltin bithiophene; these polymers have been fully characterized. The main difference between the two polymers is the substitution position of the dodecyl side chains in the repeating group. Grazing‐incidence X‐ray diffraction (GI‐XRD) gave clear evidence of edge‐on orientation of polycrystallites to the substrate. The semiconducting properties of the two polymers have been evaluated in organic thin film transistors (OTFTs). The two conjugated polymers 4 and 8 exhibit fairly high hole carrier mobilities as high as μave = 0.05 cm2/Vs (ION/OFF = 3.42 × 104) and μave = 0.01 cm2/Vs, (ION/OFF = 1.3 × 105), respectively, after thermal annealing process. The solvent annealed films underwent reorganization of the molecules to induce higher crystallinity. Well‐defined atomic force microscopy (AFM) topography supported a significant improvement in TFT device performance. The hole carrier mobilities of the solvent annealed films are comparable to those obtained for a thermally annealed sample, and were one‐order higher than those obtained with a pristine sample. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010  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