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.
《Ceramics International》2016,42(8):9712-9716
A novel mixed-potential type NO2 sensor was fabricated using La10Si6O27 electrolyte and WO3 sensing electrode (SE). The sinterability of La10Si6O27 was significantly improved by the introduction of Y2O3 as sintering aid. WO3 with different morphologies prepared by the citric acid (CA) assisted hydrothermal method was examined as the sensing electrodes of the mixed-potential type NO2 sensors based on La10Si6O27 electrolyte. The results showed that 6 wt% Y2O3 added La10Si6O27 electrolyte sample could get quite dense at a temperature as low as 1500 °C. The morphologies and phase constituents of WO3 were influenced by the CA content. The sensor showed good response–recovery characteristics. Compared with the sensor based on the irregular WO3 particles or nanorods, the sensor using WO3 nanosheets-SE with hexagonal structure exhibited much higher sensitivity (195.6 mV/decade) to NO2 at 550 °C. The response signals of the sensor were slightly affected by coexistent O2 varying from 5 to 20 vol%.  相似文献   

2.
《Journal of Catalysis》2007,245(1):1-10
The redox mechanism governing the selective catalytic reduction (SCR) of NO/NO2 by ammonia at low temperature was investigated by transient reactive experiments over a commercial V2O5/WO3/TiO2 catalyst for diesel exhaust aftertreatment. NO + NH3 temperature-programmed reaction runs over reduced catalyst samples pretreated with various oxidizing species showed that both NO2 and HNO3 were able to reoxidize the V catalyst at much lower temperature than gaseous O2: furthermore, they significantly enhanced the NO + NH3 reactivity below 250 °C via the buildup of adsorbed nitrates, which act as a surface pool of oxidizing agents but are decomposed above that temperature. Both such features, which were not observed in comparative experiments over a V-free WO3/TiO2 catalyst, point out a key catalytic role of the vanadium redox properties and can explain the greater deNOx efficiency of the “fast” SCR (NO + NH3 + NO2) compared with the “standard” SCR (NO + NH3 + O2) reaction. A unifying redox approach is proposed to interpret the overall NO/NO2–NH3 SCR chemistry over V-based catalysts, in which vanadium sites are reduced by the reaction between NO and NH3 and are reoxidized either by oxygen (standard SCR) or by nitrates (fast SCR), with the latter formed via NO2 disproportion over other nonreducible oxide catalyst components.  相似文献   

3.
《Ceramics International》2016,42(6):7309-7314
Metal oxide nanocomposite sensors based on γ-Fe2O3 and WO3 were investigated in acetone vapor of various concentrations (1–100 ppm) at operating temperatures between 250 and 350 °C. The composites were prepared by simple solid state mixing and porous thick-film gas sensors were fabricated on alumina substrates. The γ-Fe2O3:WO3 (50:50) nanocomposite showed a marked enhancement in sensing response down to 1 ppm acetone vapor detection at 300 °C. The response was ~2-fold better compared to pure WO3 or pure γ-Fe2O3 with a very fast response (1 s) and very short recovery time (3 s). No appreciable sensitivity was observed towards alcohol vapor (an interfacing agent for diabetics) and in moisture (present in breath). The enhanced performance was due to n–n heterojunction effect.  相似文献   

4.
《Ceramics International》2017,43(16):13185-13192
WO3 is one of the inspiring sensing materials that show high response to O3; an efficient fabrication of WO3 film with incorporation of complementary additives is essential for enhanced sensitivity. Here we report film deposition by liquid flame spraying, characterization of nanostructured WO3-reduced graphene oxide (rGO) composites and their gas-sensing activities to O3. The starting feedstock was prepared from WCl6 and rGO for pyrolysis synthesis by flame spraying. Nano-porous WO3-rGO films were successfully fabricated and characterized by transmission electron microscopy, field emission scanning electron microscopy, Raman spectrometry, thermal analyses and X-ray diffraction. Nanosized WO3 grains exhibited oriented nucleation on rGO flakes whereas rGO retained intact its nano-structural features after spraying. Constrained grain growth of WO3 of 60–70 nm in size was realized in the rGO-containing films with as compared to ~220 nm in the pure WO3 film. The WO3-rGO film sensors showed quicker response to O3 and faster recovery than rGO-free WO3 film sensors. Addition of rGO in 1.0 wt% or 3.0 wt% in the films caused a significantly reduced effective working temperature of the film sensors from ~ 250 °C to ~ 150 °C.  相似文献   

5.
MnOx–WOx–CeO2 catalysts synthesized using a sol–gel method were investigated for the low-temperature NH3-SCR reaction. Among them, W0.1Mn0.4Ce0.5 mixed oxides exhibited above 80% NOx conversion from 140 to 300 °C. In addition, this catalyst exhibited high stability and CO2 tolerance in a 50 h activity test at 150 °C. Substantially reduced N2O production and enhanced N2 selectivity were achieved by WO3 doping, which was due to the weakened reducibility and increased number of acid sites. The decreased SO2 oxidation activity as well as the reduced formation of ammonium and manganese sulfates resulted in a high SO2 resistance of this catalyst.  相似文献   

6.
《Ceramics International》2016,42(11):13128-13135
A facile and well-controllable reduced graphene oxide/tungsten trioxide (rGO/WO3) nanocomposite electrode was successfully synthesized via an electrostatic assembly route at 350 rpm for 24 h. In this study, hexagonal-phase WO3 (h-WO3) nanofiber was well distributed on rGO sheets by applying optimal processing parameters. The as-synthesized rGO/WO3 nanocomposite electrode was compared with pure h-WO3 electrode. A maximum specific capacitance of 85.7 F g−1 at a current density of 0.7 A g−1 was obtained for the rGO/WO3 nanocomposite electrode, which showed better electrochemical performance than the WO3 electrode. The incorporation of WO3 into rGO could prevent the restacking of rGO and provide favourable surface adsorption sites for intercalation/de-intercalation reactions. The impedance studies demonstrated that the rGO/WO3 nanocomposite electrode exhibited lower resistance because of the superior conductivity of rGO that improved ion diffusion into the electrode. To evaluate the contribution of WO3 to the rGO/WO3 nanocomposite, the influence of mass loading of WO3 on the capacitance was investigated.  相似文献   

7.
DTA, XRD and SEM investigations were conducted on the (1  x)TeO2xWO3 glasses (where x = 0.15, 0.25 and 0.3). Whereas the 0.75TeO2–0.25WO3 and 0.7TeO2–0.3WO3 glasses show no exothermic peaks, an indication of no crystallization in their glassy matrices, two crystallization peaks were observed on the DTA plot of the 0.85TeO2–0.15WO3 glass. On the basis of the XRD measurements of the 0.85TeO2–0.15WO3 glass samples heated to 510 °C and 550 °C (above the peak crystallization temperatures), α-TeO2 (paratellurite), γ-TeO2 and WO3 phases were detected in the sample heated to 510 °C and the α-TeO2 and WO3 phases were present in the sample heated to 550 °C. SEM micrographs taken from the 0.85TeO2–0.15WO3 glass heated to 510 °C showed that centrosymmetrical crystals were formed as a result of surface crystallization and were between 3 μm and 15 μm in width and 12 μm and 30 μm in length. On the other hand, SEM investigations of the 0.85TeO2–0.15WO3 glass heated to 550 °C revealed the evidence of bulk massive crystallization resulting in lamellar crystals between 1 μm and 3 μm in width and 5 μm and 30 μm in length. DTA analyses were carried out at different heating rates and the Avrami constants for the 0.85TeO2–0.15WO3 glass heated to 510 °C and 550 °C were calculated as 1.2 and 3.9, respectively. Using the modified Kissinger equation, activation energies for crystallization were determined as 265.5 kJ/mol and 258.6 kJ/mol for the 0.85TeO2–0.15WO3 glass heated to 510 °C and 550 °C, respectively.  相似文献   

8.
《Ceramics International》2017,43(16):13581-13591
The nanocomposites of WO3 nanoparticles and exfoliated graphitized C3N4 (g-C3N4) particles were prepared and their properties were studied. For this purpose, common methods used for characterization of solid samples were completed with dynamic light scattering (DLS) method and photocatalysis, which are suitable for study of aqueous dispersions.The WO3 nanoparticles of monoclinic structures were prepared by a hydrothermal method from sodium tungstate and g-C3N4 particles were prepared by calcination of melamine forming bulk g-C3N4, which was further thermally exfoliated. Its specific surface area (SSA) was 115 m2 g−1.The nanocomposites were prepared by mixing of WO3 nanoparticles and g-C3N4 structures in aqueous dispersions acidified by hydrochloric acid at pH = 2 followed by their separation and calcination at 450 °C. The real content of WO3 was determined at 19 wt%, 52 wt% and 63 wt%. It was found by the DLS analysis that the g-C3N4 particles were covered by the WO3 nanoparticles or their agglomerates creating the nanocomposites that were stable in aqueous dispersions even under intensive ultrasonic field. Using transmission electron microscopy (TEM) the average size of the pure WO3 nanoparticles and those in the nanocomposites was 73 nm and 72 nm, respectively.The formation of heterojunction between both components was investigated by UV–Vis diffuse reflectance (DRS) and photoluminescence (PL) spectroscopy, high-resolution transmission electron microscopy (HRTEM), X-ray photoelectron spectroscopy (XPS), photocatalysis and photocurrent measurements. The photocatalytic decomposition of phenol under the LED source of 416 nm identified the formation of Z-scheme heterojunction, which was confirmed by the photocurrents measurements. The photocatalytic activity of the nanocomposites decreased with the increasing content of WO3, which was explained by shielding of the g-C3N4 surface by bigger WO3 agglomerates. This study also demonstrates a unique combination of various characterization techniques working in solid and liquid phase.  相似文献   

9.
《Ceramics International》2015,41(6):7729-7734
We report bead-like ZnO nanostructures for gas sensing applications, synthesized using multi-walled carbon nanotube (MWCNT) templates. The ZnO nanostructures are grown following a two-step process: in the first, ZnO nanoparticles are synthesized on MWCNTs by thermal evaporation of a Zn powder; and in the second, the hybrid nanostructures are heat-treated at 800 °C. Scanning and transmission electron microscopy images indicate that the bead-like ZnO nanostructures have surface protuberances with nanoparticle sizes ranging from 20 to 60 nm, and a well-crystallized hexagonal structure. Gas sensors based on multiple-networked bead-like ZnO showed considerably enhanced electrical responses and better stability to both oxidizing (NO2) and reducing (CO) gases compared with previously reported nanostructured gas sensors, even if the response to CO gas was slow to increase. Both the NO2 and CO gas sensing properties increased dramatically when the working temperature was increased up to 300 °C. The response sensitivities measured were 2953%, 5079%, 9641%, 3568%, and 3777% to 20 ppm NO2 at 200, 250, 300, 350 and 400 °C, respectively. For CO gas on the other hand, the response sensitivities were 107%, 110%, 114%, 118%, and 122% at 5, 10, 20, 50, and 100 ppm concentrations, respectively. For concentrations between 5 and 20 ppm, the recovery time of the oxidizing gas was much shorter than the response time. The origin of the NO2/CO gas sensing mechanism of the bead-like ZnO nanostructures is discussed.  相似文献   

10.
The effect of dispersant on deposition mechanism of TiO2 nanoparticles at 1 Hz under non-uniform AC fields was investigated. It was found that by adding Dolapix to suspension, deposition pattern is drastically changed enabling particles to enter the gap leaving the electrodes intact. Using low frequency AC electrophoretic deposition technique in the presence of dispersant, we succeeded in fabricating gas sensor in less than 2 min. Gas sensing measurements were performed in the temperature range of 450–550 °C. The results explained that the sensor has good stability in time and repeatability performance toward high response. The maximum sensitivity of about 180 for the TiO2 nanoparticles sensor is observed with 47 ppm NO2 gas and the response and recovery times is about 60–150 s. The optimum temperature of the gas sensor was obtained in 450 °C where sensor showed a linear trend up to 50 ppm of NO2 gas. This sensing behavior in un-doped TiO2 as NO2 sensor can be mainly ascribed to the porous structure of the sensing film and its good contacts to the substrate and electrode assembly.  相似文献   

11.
Tungsten oxide, originally poor in capacitive performance, was made an excellent electrode material for supercapacitors, by dispersing it to carbon aerogels (CA), a conductive and mesoporous hosting template, that drastically improved the utilization of WO3 for capacitance generation. The WO3 was introduced to the CA, in a form of well-dispersed single crystalline nanoparticles of 15–40 nm in size, with a simple immersion-calcination process. A one order of magnitude improvement in specific capacitance was achieved with the present composition, from 54 F/g for WO3 nanoparticles to 700 F/g for WO3/CA composites (scaned at 25 mV/s in 0.5 M H2SO4 over a potential window of −0.3 to 0.5 V). The WO3/CA composites exhibited an excellent high rate capability with a 60% retention in specific capacitance at 500 mV/s, almost perfect cycle efficiency of 99%, and outstanding cycling stability of only 5% decay in specific capacitance after 4000 cycles.  相似文献   

12.
《Ceramics International》2016,42(14):15301-15310
Co-precipitated undoped and Cr-doped WO3 nanosheets have been investigated by X-ray diffraction (XRD), Raman spectroscopy, field emission scanning electron microscopy (FESEM), and transmission electron microscopy (TEM) in order to study the influence of Cr doping on their structural and morphological properties. XRD analyses confirm the monoclinic structure of nanocrystalline WO3, whereas the FESEM and TEM images exhibit nanosheet-like morphology of the as-synthesized WO3 materials. Among all the samples examined, the 1.5 at% Cr-doped WO3 nanosheets exhibit the selective maximum response (~82%) to formaldehyde over methanol, ethanol, propan-2-ol and acetone at the operating temperature of 200 °C for 50 ppm concentration in air. The sensing mechanism has been explained based on chemisorption of oxygen on the WO3 surface and the subsequent reaction between the adsorbed oxygen species and the formaldehyde molecules.  相似文献   

13.
WO3(0–6 mol%)-doped 0.94Bi0.5Na0.5TiO3–0.06BaTiO3 lead-free ceramics were synthesized by conventional solid-state reaction. The effect of WO3 addition on the structure and electrical properties were investigated. The result revealed that a small amount of WO3 (≤1 mol%) can diffuse into the lattice and does not significantly affect the phase structure, however, more addition will result in distortion and enlargement of the unit cells. The maximum permittivity temperature (Tm) is suppressed dramatically as the dopant increasing, while the depolarization temperature (Td) fall to the minimum with 1 mol% WO3 additive. The remanent polarization (Pr) was enhanced and coercive field (Ec) was reduced by doping with WO3. The strain shows the largest value for 1 mol% doped sample, which is due to a field-induced antiferroelectric–ferroelectric phase transition.  相似文献   

14.
The stability of δ-TeO2 phase was studied in binary TeO2–WO3, TeO2–CdO and ternary TeO2–WO3–CdO glasses. The samples were prepared by heating high purity powder mixtures of TeO2, WO3 and/or CdO to 800 °C in a platinum crucible with a closed lid, holding for 30 min and quenching in water bath. Differential thermal analysis (DTA), X-ray diffraction (XRD) and scanning electron microscopy (SEM) techniques were used to characterize the thermal, phase and microstructural properties of the δ-TeO2 phase. The addition of CdO into the tellurite glasses increased the stability range of the δ-TeO2 phase up to higher temperature values and expanded the compositional δ-TeO2 formation range. The formation of δ-TeO2 phase in the binary systems was observed for samples containing 5–10 mol% WO3 and 5–15 mol% CdO. However, for the ternary TeO2–WO3–CdO system the formation of δ-TeO2 phase was determined in a wider compositional range.  相似文献   

15.
The influence of the WO3 addition as sintering aids on the structural, microstructural and optical properties of (Pb,La)(Zr,Ti)O3—PLZT based ceramics was investigated. Ferroelectric (Pb0.9La0.1)(Zr0.65Ti0.35)0.975O3 + x wt% WO3 (x = 0.0, 0.5, 1.5 and 2.0) ceramics were densified by oxygen assisted uniaxial hot pressing. From the XRD results it was found that the hot pressed samples displayed single pseudocubic perovskite phase. EDS analysis evidenced the presence of WO3 rich phase at the grain boundaries. An inhibition of grain growth and an evolution from transgranular to intergranular fracture behavior was observed, as a consequence of the formation of a PbO–WO3 liquid phase, as the amount of WO3 addition was increased. The optical transmittance in the visible and infrared range was decreased due to the presence of the liquid phase in grain boundaries, for WO3 content lower than 2.0 wt%.  相似文献   

16.
《Fuel》2007,86(10-11):1577-1586
The NO2, NO (O2) adsorption and temperature programmed desorption (TPD) were studied systematically to probe into the selective catalytic reduction of NO by methane (CH4–SCR) over CoH-ZSM-5 (SiO2/Al2O3 = 25, Co/Al = 0.132–0.312). Adsorption conditions significantly affect the adsorption of NO, NO2 and NO + O2. Adsorbed NO species are unstable and desorbed below the reactive temperature 523 K. Increasing adsorption temperature results in the decrease of the adsorbed NO species amount. The amount of –NOy species formed from NO2 adsorption increases with the increase of NO2 concentration in the adsorption process, while decreases significantly with the increase of adsorption temperature. Though NO species are adsorbed weakly on CoH-ZSM-5, competitive adsorption between NO and –NOy species decreases the amount of adsorbed –NOy species. Similar desorption profiles of NO2 were obtained over CoH-ZSM-5 while they were contacted with NO2 or NO + O2 followed by TPD. If NO2 was essential to form adsorbed –NOy species, the amount of adsorbed –NOy species for NO + O2 adsorption should be the least among the adsorptions of NO2, NO + O2 and NO + NO2 because of the lowest NO2 concentration and highest NO concentration. In fact, the amount of adsorbed –NOy species is between the other two adsorption processes. These indicate that the formation of adsorbed –NOy species may not originate from NO2.  相似文献   

17.
《Ceramics International》2015,41(8):9426-9432
We demonstrate low-temperature formation of copper oxide (CuO) nanostructures as well as temperature-controlled variation of morphology by applying hydrothermal methods with copper(II) acetate Cu(CH3COO)2·H2O and 2-piperidinemethanol (2PPM) as starting materials. Monoclinic CuO nanostructures produced at 25 °C were of dendritic morphology with short nanorod-like substructures and exhibited a consequently large surface area (179 m2 g−1). Cyclic voltammetry measurements confirmed pseudocapacitive behavior of these dendritic CuO nanostructures giving specific capacitance ca. 28.2 F g−1 at a scan rate of 5 mV s−1. Oxide nanomaterials prepared in this investigation were characterized using powder X-ray diffraction, scanning and transmission electron microscopies, and nitrogen adsorption/desorption techniques. It is expected that these materials exhibit improved sensing and catalytic properties due to the increased availability of surface adsorption sites.  相似文献   

18.
A new supramolecular assembly of [(CuL)2(WO4)2{CuL(H2O)}2][W10O32] · 8H2O (1) (L = 4′-(2-pyridyl)-2,2′:6′,2′-terpyridine) containing a decatungstate polyanion and a novel cationic Cu/O/W heterometallic cluster was successfully synthesized and characterized by single crystal X-ray diffraction and IR. The tungsten atoms adopt different coordination geometry in anion and cation, octahedral geometry (WO6) in decatungstate cluster and rare tetrahedral geometry (WO4) in cation cluster. During the hydrothermal process, the decomposition and reassembly based on [W6O19]2  lead to the formation of [W10O32]4  and Cu/O/W heterometallic cluster.  相似文献   

19.
Catalytic dehydration of fructose and its conversion to 5-hydroxymethylfurfural was studied using tungstated zirconia oxides, with various tungsten oxide loadings (1–20 wt.%). The samples were prepared by incipient wetness impregnation and thoroughly characterized using a combination of different techniques: structural, thermal and calorimetric analyses. Zirconia was predominantly present in the investigated samples in the tetragonal phase when the WO3 loading was above 10 wt.%. The samples exhibited amphoteric characteristics, as they adsorbed both ammonia and sulfur dioxide on their surface. The number of surface acid sites increased with increasing WO3 content. Fructose dehydration tests evidenced the formation of 5-hydroxymethylfurfural and by-products (formic and levulinic acids). The results show that the ratio of basic to acidic sites of the solid catalysts is the key parameter for the selectivity in 5-HMF, while the global fructose conversion was mainly related to the presence of acid sites of a given strength with 150 > Qdiff > 100 kJ·molNH3 1.  相似文献   

20.
The Ba,K/CeO2 catalyst is active both for NOx trapping and soot combustion. In this work we report a Ba–K interaction that prevents K sulfation when NOx is present, thus preserving the activity of K towards soot combustion during the working period of the trap. This effect is originated in the K2SO4(s) + Ba(NO3)2(s)  2KNO3(s) + BaSO4(s) reaction, which is thermodynamically favored. In the absence of NOx, the soot combustion reaction is strongly depressed by SO2 whereas when NOx is present both the sulfated and the non-sulfated catalysts present similar TPO patterns.  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