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 187 毫秒
1.
Neodymium(III) peroxotitanate is used as a precursor for obtaining Nd2TiO5. The last one possesses numerous valuable electrophysical properties. TiCl4, Nd(NO3)3·6H2O and H2O2 in mol ratio 1:2:10 were used as starting materials. The reaction ambience was alkalized to pH = 9 with a solution of NH3. The obtained neodymium(III) peroxotitanate and intermediate compounds of the isothermal heating were proved by the help of quantitative analysis and infrared spectroscopy (IRS). It has Nd4[Ti2(O2)4(OH)12]·7H2O composition. The absorption band observed in IRS at 831 cm?1 relates to a triangular bonding of the peroxo group of Ti, at 1062 cm?1—terminal groups Ti–OH and at 1491 and 1384 cm?1—the bridging OH?-groups Ti–O(H)–Ti. Nd2TiO5 was obtained by thermal decomposition of neodymium(III) peroxotitanate. The isothermal conditions for decomposition were determined on the base of differential thermal analysis, thermogravimetric and differential scanning calorimetry results in the temperature range of 20–1000 °C. The mechanism of thermal decomposition of Nd4[Ti2(O2)4(OH)12]·7H2O to Nd2TiO5 was studied. In the temperature range of 20–208 °C, a simultaneous decomposition of the peroxo groups by the separation of oxygen and hydrate water is conducted and Nd4[Ti2O4(OH)12] is obtained. From 208 to 390 °C, the terminal OH?-groups are separated and Nd4[Ti2O7(OH)6] is formed. In the range of 390–824 °C, the bridging OH?-groups are completely decomposed to Nd2TiO5. The optimal conditions for obtaining nanocrystalline Nd2TiO5 are 900 °C for 6 h and 20–80 nm.  相似文献   

2.
In this work, the capping layer atop anodic TiO2 nanotube arrays (NTAs), which hinders filling of other guest materials and transport of charge carriers, is discerned to be TiO2 nanotapes. Then, it is completely removed by a novel sonication-polishing (SP) treatment, after which Sb2S3 is subsequently introduced to fill the nanotubes by chemical bath deposition. The morphological, structural, and optical properties of the SP-treated TiO2 NTAs and TiO2 NTAs/Sb2S3 heterogeneous structures are characterized systematically. The results indicate that SP treatment opens the tops of nanotubes with diameters of ~120 nm, which endure a phase conversion from amorphous to anatase after calcination at 450 °C; besides, stibnite Sb2S3 with a band gap of ~1.75 eV inside the TiO2 networks is formed upon heat treatment at 330 °C in Ar, which enhances the absorption in visible light range. The photoelectrochemical (PEC) and photovoltaic properties for the SP-treated TiO2 NTAs are investigated. Results shows that the photoresponse of TiO2 NTAs is improved by the SP treatment, and the photocurrent for the TiO2 NTAs/Sb2S3 electrode is substantially enhanced as compared to the bare TiO2 one. A high efficiency of 6.28 % is achieved in a TiO2 NTAs/Sb2S3 PEC cell. In addition, charge recombination in the photoanode of dye-sensitized solar cells (DSSCs) is observed to be greatly retarded by using the SP-treated TiO2 NTAs as compared to TiO2 nanoparticles (NPs). Thus, the SP anodic TiO2 NTAs are promising in applications in various PEC areas such as photocatalysis and sensitized solar cells.  相似文献   

3.
A new photoelectrochemical starch-O2 biofuel cell has been developed, consisting of chlorin-e6 (Chl-e6) adsorbed on a TiO2 layer onto an optical transparent conductive glass electrode as a photoanode, bilirubin oxidase (BOD)-modified electrode as a cathode, and a solution containing starch, glucoamylase, glucose dehydrogenase and NAD+ as fuel. The short-circuit photocurrent and the open-circuit photovoltage of this cell are 9.0 μA cm?2 and 530 mV, respectively. The maximum power, FF and \(\eta\) values are estimated to be 1.7 μW cm?2, 0.36 and 0.0017 %, respectively. Thus, this new type of the photochemical starch-O2 biofuel cell has been developed by using the visible light photosensitization of Chl-e6 on a TiO2 film photoanode.  相似文献   

4.
The degradation of ofloxacin (OFX) at low concentration in aqueous solution by UVA-LED/TiO2 nanotube arrays photocatalytic fuel cells (UVA-LED/TiO2 NTs PFCs) was investigated. TiO2 nanotube arrays (TiO2 NTs) photoanode prepared by anodization-constituted anatase–rutile bicrystalline framework. The results indicated that the degradation efficiency of OFX by UVA-LED/TiO2 NTs PFC was significantly enhanced by 14.3% compared with UVA-LED/TiO2 NTs photocatalysis. The pH affected the degradation efficiency markedly; the highest degradation efficiency (95.0%) and the pseudo-first-order reaction rate constant k value (0.049 min?1) were achieved in neutral condition (pH 7.0). The degradation efficiency increased with the increasing concentration of dissolved oxygen (DO) in the UVA-LED/TiO2 NTs PFC. The main reactive species of OFX degradation are positive holes (h+) and superoxide ion radicals (O 2 ·? ) in a DO sufficient condition. Furthermore, the possible pathways of OFX degradation were proposed.  相似文献   

5.
Homogeneous p-type cobalt (II) oxide (CoO) nanoparticles were successfully deposited on n-type three-dimensional branched TiO2 nanorod arrays (3D-TiO2) through photochemical deposition and thermal decomposition to form a novel CoO/3D-TiO2 p-n heterojunction nanocomposite. Due to the narrow band gap of CoO nanoparticles (~2.4 eV), the as-synthesized CoO/3D-TiO2 exhibited an excellent visible light absorption. The amounts of deposited CoO nanoparticles obviously influence the hydrogen production rate in the photoelectrochemical (PEC) water splitting. The as-synthesized CoO/3D-TiO2-5 obtains the highest PEC hydrogen production rate of 0.54 mL h?1 cm?2 after five-time CoO deposition cycles (at a potential of 0.0 V vs Ag/AgCl). The photocurrent density of CoO/3D-TiO2-5 is 1.68 mA cm?1, which is ca. 2.5 times greater than that of pure 3D-TiO2. The results showed that the formation of internal electrical-field between the CoO/3D-TiO2 heterojunction, which has a direction from n-type TiO2 to p-type CoO, facilitated the charge separation and transfer and resulted in a high efficiency and stable PEC activity.  相似文献   

6.
The compound [Ni(NH3)6][VO(O2)2(NH3)]2 was prepared and characterized by elemental analysis and vibrational spectra. The single crystal X-ray study revealed that the structure consists of [Ni(NH3)6]2+ and [VO(O2)2(NH3)] ions. As a result of weak interionic interactions V′···Op (Op-peroxo oxygen), ([VO(O2)2(NH3)])2 dimers are formed in the solid-state. The thermal decomposition of [Ni(NH3)6][VO(O2)2(NH3)]2 is a multi-step process with overlapped individual steps; no defined intermediates were obtained. The final solid products of thermal decomposition up to 600°C were Ni2V2O7 and V2O5.  相似文献   

7.
A photoconversion efficiency of 2.12% was obtained under visible light illumination by nanostructure-sensitized photoelectrochemical cells using Mn/CdS as sensitizer loaded on TiO2 nanotube arrays (NTAs) (Mn/CdS/TiO2). Sensitization of Mn on CdS nanoparticles pre-loaded on TiO2 NTAs was carried out by a two-step electrodeposition method. Compared with unsensitized TiO2 NTAs, the photocurrent had increased from 0.03 to 4.12 for Mn/CdS/TiO2 prepared at 1 min. The effects of deposited Mn on the physical, chemical, and photoelectrochemical properties of the CdS/TiO2 NTAs nanostructure were investigated by using UV–visible diffuse reflectance spectroscopy, X-ray diffractometry, and field-emission scanning electron microscopy coupled with energy dispersive X-ray spectroscopy. The photoelectrochemical analysis was examined in a three-electrode system under a halogen illumination by using the prepared film as the photo-anode.  相似文献   

8.
9.
The authors describe a highly efficient photoelectrochemical (PEC) scheme for the determination of hydrogen peroxide (H2O2). BiVO4 microrods were hydrothermally synthesized and deposited on fluorine - doped tin oxide (FTO) glass which acts as the working electrode. Scanning electron microscopy, X-ray powder diffraction and Raman spectroscopy were utilized for the characterization of the microrods. On irradiation with visible light, the holes generated in the microrods are capturing electrons from H2O2 to produce a photocurrent at an operating potential of 0 V vs. Ag/AgCl. Under optimal conditions, the photocurrent increases with the concentration of H2O2 in the range from 50 μmol·L?1 to 1.5 mmol·L?1, and the limit of detection is 8.5 μmol·L?1 (at 3σ). A repeatability and intermediate precision of ≤6.6% was accomplished at H2O2 levels of 0.1, 0.5 and 1.0 mmol·L?1. The method was applied to the determination of H2O2 in spiked sterilized milk samples and gave satisfactory results. As the method works at zero potential, the photocurrent can be measured with simple instrumentation such as digital multimeters, and this will enable expensive electrochemical workstations to be replaced in future.
Graphical abstract An enzyme-free photoelectrochemical sensing strategy is described for sensitive determination of hydrogen peroxide in foodstuff using fluorine-doped tin oxide electrode modified with BiVO4 microrods.
  相似文献   

10.
A GOx/Ag/TiO2 glucose biosensor was achieved by photoreducing Ag nanoparticles on TiO2 nanotube arrays (NTAs) following with adsorption of GOx. The morphology, structure, and element component of Ag/TiO2 NTAs were characterized by scanning electron microscope, transmission electron microscope, and X-ray diffraction. Ag nanoparticles were uniformly deposited on surface of TiO2 NTAs with average size of 15 nm and the size and distribution changed with the immersing time of TiO2 NTAs in AgNO3 aqueous solution. Electrochemical properties of Ag/TiO2 NTAs were characterized by cyclic voltammetry and amperometric detection of H2O2, revealing that TiO2 NTAs with immersing time of 30 min achieve the best electrochemical activity. The GOx/Ag/TiO2 NTAs biosensor with optimum conditions achieves a sensitivity of 0.39μA mM?1 cm?2 with liner range from 0.1 to 4 mM.  相似文献   

11.
Li4Ti5O12/Li2TiO3 composite nanofibers with the mean diameter of ca. 60 nm have been synthesized via facile electrospinning. When the molar ratio of Li to Ti is 4.8:5, the Li4Ti5O12/Li2TiO3 composite nanofibers exhibit initial discharge capacity of 216.07 mAh g?1 at 0.1 C, rate capability of 151 mAh g?1 after being cycled at 20 C, and cycling stability of 122.93 mAh g?1 after 1000 cycles at 20 C. Compared with pure Li4Ti5O12 nanofibers and Li2TiO3 nanofibers, Li4Ti5O12/Li2TiO3 composite nanofibers show better performance when used as anode materials for lithium ion batteries. The enhanced electrochemical performances are explained by the incorporation of appropriate Li2TiO3 which could strengthen the structure stability of the hosted materials and has fast Li+-conductor characteristics, and the nanostructure of nanofibers which could offer high specific area between the active materials and electrolyte and shorten diffusion paths for ionic transport and electronic conduction. Our new findings provide an effective synthetic way to produce high-performance Li4Ti5O12 anodes for lithium rechargeable batteries.  相似文献   

12.
Controllable CdS nanoparticles (NPs) decorated on TiO2 nanotube arrays (NTAs) were prepared via electrodeposition in DMSO solution at room temperature, aiming to improve the photoelectrochemical properties of TiO2 NTA electrode in visible-light region. By tuning the concentrations of sulfur and Cd2 + as well as the deposition time, CdS NPs with different sizes can be controllably synthesized at room temperature. Excellent photocurrent response and incident photo to current conversion efficiency were achieved with smaller CdS NPs with optimal reactant concentrations and deposition time, which can be attributed to highly efficient charge separation and high dispersion of CdS NPs on both inner and outer surfaces of TiO2 nanotubes.  相似文献   

13.
The kinetics and mechanism by which monochloramine is reduced by hydroxylamine in aqueous solution over the pH range of 5–8 are reported. The reaction proceeds via two different mechanisms depending upon whether the hydroxylamine is protonated or unprotonated. When the hydroxylamine is protonated, the reaction stoichiometry is 1:1. The reaction stoichiometry becomes 3:1 (hydroxylamine:monochloramine) when the hydroxylamine is unprotonated. The principle products under both conditions are Cl, NH+4, and N2O. The rate law is given by ?[d[NH2Cl]/dt] = k+[NH3OH+][NH2Cl] + k0[NH2OH][NH2Cl]. At an ionic strength of 1.2 M, at 25°C, and under pseudo‐first‐order conditions, k+= (1.03 ± 0.06) ×103 L · mol?1 · s?1 and k0=91 ± 15 L · mol?1 · s?1. Isotopic studies demonstrate that both nitrogen atoms in the N2O come from the NH2OH/NH3OH+. Activation parameters for the reaction determined at pH 5.1 and 8.0 at an ionic strength of 1.2 M were found to be ΔH? = 36 ± 3 kJ · mol–1 and Δ S? = ?66 ± 9 J · K?1 · mol?1, and Δ H? = 12 ± 2 kJ · mol?1 and Δ S? = ?168 ± 6 J · K?1 · mol?1, respectively, and confirm that the transition states are significantly different for the two reaction pathways. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 38: 124–135, 2006  相似文献   

14.
CuCr2O4 spinel powders were synthesized starting from different chromium sources, namely (i) chromium oxide (α-Cr2O3) and (ii) ammonium dichromate ((NH4)2Cr2O7). The copper source was a Cu(II) carboxylate-type complex. The Cu(II) carboxylate complex was obtained by the redox reaction between Cu(NO3)2·3H2O and 1,3-propanediol (1,3PG) at 130 °C. In the first case (i), we have started from a mixture of α-Cr2O3, Cu(NO3)2·3H2O and 1,3PG that upon heating formed the copper malonate complex, which decomposed around 220 °C forming an oxide mixture (CuO + α-Cr2O3). In the second case (ii), (NH4)2Cr2O7, Cu(NO3)2·3H2O and 1,3PG were homogenously mixed. Heating this mixture at 130 °C resulted, in situ, in the Cu(II) complex. On controlled temperature increase, the violent decomposition of (NH4)2Cr2O7 took place at 180 °C along with the decomposition of the Cu(II) complex, leading to an amorphous oxide mixture of Cr2O3+x and CuO. By annealing the samples in the temperature range 400–1000 °C, the spinel phase (CuCr2O4) was obtained in both cases: (i) at 800 °C and (ii) at 600 °C as a result of the interactions between the precursors used, when the oxide system was amorphous and highly reactive. The presence of CuCr2O4 was highlighted by XRD and FTIR analyses.  相似文献   

15.
Fluorine-doped tin dioxide (FTO) films were deposited on silicon wafers by inverted pyrosol technique using solutions with different doping concentration (F/Sn=0.00, 0.12, 0.75 and 2.50). The physical and electrical properties of the deposited films were analyzed by SEM, XRF, resistivity measurement by four-point-probe method and Hall coefficient measurement by van der Pauw method. The electrical properties showed that the FTO film deposited using the solution with F/Sn=0.75 gave a lowest resistivity of 3.2·10–4 ohm cm. The FTO films were analyzed by temperature programmed desorption (TPD). Evolved gases from the heated specimens were detected using a quadruple mass analyzer for mass fragments m/z, 1(H+), 2(H2 +), 12(C+), 14(N+), 15(CH3 +), 16(O+), 17(OH+ or NH3 +), 18(H2O+ or NH4 +), 19(F+), 20(HF+), 28(CO+ or N2 +), 32(O2 +), 37(NH4F+), 44(CO2 +), 120(Sn+), 136(SnO+) and 152(SnO2 +). The majority of evolved gases from all FTO films were water vapor, carbon monoxide and carbon dioxide. Fluorine (m/z 19) was detected only in doped films and its intensity was very strong for highly-doped films at temperature above 400°C.  相似文献   

16.
A new photoelectrochemical method for the determination of glucose based on the photoelectrochemical effect of poly(thionine) photoelectrode to hydrogen peroxide (H2O2) was reported. The H2O2‐sensitive photoelectrode was fabricated by electropolymerizing thionine on the surface of ITO electrode. And then glucose oxidase was immobilized on the photoelectrode via the aid of chitosan enwrapping, forming an enzyme‐modified photoelectrode. The photoelectrode was employed as an electron acceptor; H2O2 from the catalytic reaction of enzyme was employed as an electron donor, developing an analytical method of glucose without hydrogen peroxidase. In the paper, the photoelectrochemical effects of photoelectrode to H2O2 and glucose were studied. The effects of the bias voltage and the electrolyte pH on the photocurrent were investigated. The linear response of glucose concentrations ranged from 0.05 to 2.00 mmol/L was obtained with a detection limit of 22.0 µmol/L and sensitivity of 73.2 nA/(mmol·L?1). The applied feasibility of method was acknowledged through monitoring the glucose in practical samples.  相似文献   

17.
Solid polymer electrolyte membranes were prepared by complexing tetrapropylammoniumiodide (Pr4N+I?) salt with polyethylene oxide (PEO) plasticized with ethylene carbonate (EC), and these were used in photoelectrochemical (PEC) solar cells fabricated with the configuration glass/FTO/TiO2/dye/electrolyte/Pt/FTO/glass. The PEO/Pr4N+I?+I2?=?9:1 ratio gave the best room temperature conductivity for the electrolyte. For this composition, the plasticizer EC was added to increase the conductivity, and a further conductivity enhancement of four orders of magnitude was observed. An abrupt increase in conductivity occurs around 60–70 wt% EC; the room temperature conductivity was 5.4?×?10?7 S cm?1 for 60 wt% EC and 4.9?×?10?5 S cm?1 for the 70 wt% EC. For solar cells with electrolytes containing PEO/Pr4N+I?+I2?=?9:1 and EC, IV curves and photocurrent action spectra were obtained. The photocurrent also increased with increasing amounts of EC, up to three orders of magnitude. However, the energy conversion efficiency of this cell was rather low.  相似文献   

18.
Photocatalytic degradation of methyl orange by TiO2–SiO2–NiFe2O4 suspensions was investigated. Adsorption studies revealed photocatalytic degradation occurred mainly on the surface of the TiO2–SiO2–NiFe2O4. The disappearance of the compound followed the zero-order kinetics according to the Langmuir–Hinshelwood model and the rate constant was 0.0035 mg L?1 min?1. The rate constant depended on the amount of photocatalyst, initial pH, and the presence of additional scavengers. ?OH radicals and h+ had important roles in the photocatalytic degradation of methyl orange by TiO2–SiO2–NiFe2O4.  相似文献   

19.
One-dimensional (1D) Ag/AgBr/TiO2 nanofibres (NFs) have been successfully fabricated by the one-pot electrospinning method. In comparison with bare TiO2 NFs and Ag/AgBr/PVP (polyvinylpyrrolidone) NFs, the 1D Ag/AgBr/TiO2 NFs photocatalyst exhibits much higher photocatalytic activity in the degradation of a commonly used dye, methylene blue (MB), under visible light. The photocatalytic removal efficiency of MB over Ag/AgBr/TiO2 NFs achieves almost 100 % in 20 min. The photocatalytic reaction follows the first-order kinetics and the rate constant (k) for the degradation of MB by Ag/AgBr/TiO2 NFs is 5.2 times and 6.6 times that of Ag/AgBr/PVP NFs and TiO2 NFs, respectively. The enhanced photocatalytic activity is ascribed to the stronger visible light absorption, more effective separation of photogenerated electron-hole pairs, and faster charge transfer in the long nanofibrous structure. The Ag/AgBr/TiO2 NFs maintain a highly stable photocatalytic activity due to its good structural stability and the self-stability system of Ag/AgBr. The mechanisms for photocatalysis associated with Ag/AgBr/TiO2 NFs are proposed. The degradation of MB in the presence of scavengers reveals that h+ and ?O 2 ? significantly contribute to the degradation of MB.  相似文献   

20.
[Mn(NH3)6](NO3)2 crystallizes in the cubic, fluorite (C1) type crystal lattice structure (Fm \( \overline{3} \) m) with a = 11.0056 Å and Z = 4. Two phase transitions of the first-order type were detected. The first registered on DSC curves as a large anomaly at T C1 h  = 207.8 K and T C1 c  = 207.2 K, and the second registered as a smaller anomaly at T C2 h  = 184.4 K and T C2 c  = 160.8 K (where the upper indexes h and c denote heating and cooling of the sample, respectively). The temperature dependence of the full width at half maximum of the band associated with the δs(HNH)F1u mode suggests that the NH3 ligands in the high temperature and intermediate phase reorientate quickly with correlation times in the order of several picoseconds and with activation energy of 9.9 kJ mol?1. In the phase transition at T C2 c probably only a some of the NH3 ligands stop their reorientation, while the remainders continue to reorientate quickly with activation energy of 7.7 kJ mol?1. Thermal decomposition of the investigated compound starts at 305 K and continues up to 525 K in four main stages (I–IV). In stage I, 2/6 of all NH3 ligands were seceded. Stages II and III are connected with an abruption of the next 2/6 and 1/6 of total NH3, respectively, and [Mn(NH3)](NO3)2 is formed. The last molecule of NH3 per formula unit is freed at stage IV together with the simultaneous thermal decomposition of the resulting Mn(NO3)2 leading to the formation of gaseous products (O2, H2O, N2 and nitrogen oxides) and solid MnO2.  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