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 15 毫秒
1.
The present work reports the formation of multilayers of self-organized zirconium titanate nanotubes by anodizing a Ti–35Zr alloy in 1 M (NH4)2SO4 + 0.5 wt% NH4F electrolytes. It was found that multilayers consisting of different diameter nanotubes can be produced by repeated anodization steps under different conditions. Formation of new nanotubes starts in the gaps between the existing tubes. The process allows the formation of multilayer stacks consisting of layers of several 100 nm in length and adjustable nanotube diameters in a range from 50 to 180 nm.  相似文献   

2.
0.3 wt % ammonium fluoride (NH4F) or ammonium chloride (NH4Cl) was added to ethylene glycol (EG) as an active ingredient for the formation of anodic oxide comprising of ZrO2 nanotubes (ZNTs) by anodic oxidation of zirconium (Zr) at 20 V for 10 min. It was observed that nanotubes were successfully grown in EG/NH4F/H2O with aspect ratio of 144.3. Shorter tubes were formed in EG/NH4F/H2O2. This could be due to higher excessive chemical etching at the tip of the tubes. When fluoride was replaced by chloride in both electrolytes, multilayered oxide resembling pyramids was observed. The pyramids have width at the bottom of 3-4 μm and the top is 1-2 μm with 10.7 μm height. Oxidation of Zr in EG/NH4Cl/H2O2 was rater rapid. The multilayered structure is thought to have formed due to the re-deposition of ZrO2 or hydrated ZrO2 on the foil inside pores formed within the oxide layer. XRD result revealed an amorphous structure for as-anodized samples regardless of the electrolytes used for this work.  相似文献   

3.
Bioactive zirconium oxide nanotubular arrays on zirconium alloys are prepared electrochemically in fluoride and phosphate containing electrolyte. Geometric factors of the ZrO2 nanotubular layers, particularly the pore diameter and thickness, are affected by the electrochemical conditions, including applied potential and anodization time. Under specific sets of conditions, highly ordered ZrO2 nanotubular arrays are formed with diameters varying from 30 nm to 75 nm and lengths varying from 2 μm to 12 μm. XPS shows that the nanotubular layer contains a significant amount of phosphate species distributed almost homogeneously over the entire tubular length. The ZrO2 nanotubular layer formed in fluoride and phosphate containing electrolyte highly enhances the formation of bioactive hydroxyapatite coating in simulated biological fluid (SBF).  相似文献   

4.
Ag and Au nanoparticles were found to significantly enhance the photocatalytic activity of self-organized TiO2 nanotubular structures. The catalyst systems are demonstrated to be highly efficient for the UV-light induced photocatalytic decomposition of a model organic pollutant – Acid Orange 7. The metallic nanoparticles with a diameter of ∼10 ± 2 nm (Ag) and ∼28 ± 3 nm (Au) were attached to a nanotubular TiO2 layer that consists of individual tubes of ∼100 nm of diameter, ∼2 μm in length and approx. 15 nm of wall thickness. Both metal particle catalyst systems enhance the photocatalytic decomposition significantly more on the nanotubes support than placed on a compact TiO2 surface.  相似文献   

5.
The TiO2 nanotube films were prepared by anodizing Ti plates in 0.2 M NH4F ethylene glycol/10% H2O under different formation voltages keeping fixed the time length, or by keeping fixed the formation voltage and varying the time length. The morphology of the TiO2 film obtained was observed by SEM images and different morphological parameters were derived from them. Furthermore, the optical and semiconducting properties of TiO2 films were also measured. The photoelectrochemical performance toward water oxidation of the TiO2 only showed to be dependent with the inner diameter of the nanotubes, that could be related to the interaction of the film with the light and the transport of species in the electrolyte inward or outward the film.  相似文献   

6.
Electrochromical properties of anodic self-assembled nanotubes were investigated. It was found that amorphous titania nanotubes were able to insert H+ ions in a highly reversible manner. Coloration of the TiO2 nanotubes occurred at potentials below ?0.5 V vs. Ag/AgCl in 1M (NH4)2SO4 aqueous solution. The proton insertion reaction probably leads to the formation of a Ti3+/Ti4+ solid solution in the amorphous titania electrode, as was shown by the analysis of the derivative curve. The nanotubular titania electrode shows reasonable color efficiency when compared with other electrochromic materials and it is a promising candidate for the fabrication of low-cost interdigitated electrochromic devices.  相似文献   

7.
TiO2@Sn core–shell nanotube material prepared by thermal decomposition of SnCl4 on TiO2 nanotubes at 300 °C has been demonstrated superior Li-ion storage capability of 176 mA h/g even at high current rate of 4000 mA/g (charge and discharge of all TiO2 within 5 min) in spite of using low carbon content (5 wt%). This value corresponds to volumetric energy densities of 317 mA h/cm3, and its value was 3.5-fold larger than that of the bare TiO2 nanotubes.  相似文献   

8.
We report ultraviolet (UV) light detection of thin wall TiO2 nanotubes (TNTs) with open diameter ~ 20 nm obtained by a two anodization procedure. This small diameter nanotubular geometry shows significant enhancement of the photoresponsivity and results in a large increase of photocurrent. The photocurrent is one order higher than that of classical nanotubes with diameter of 140 nm at − 1.0 V bias. We attribute this improvement to the modulation of hole carrier density as a result of field effects from the diameter-dependent population of the surface-trapped electrons. This finding demonstrates inherent size effects of internal gain in semiconductor nanotubes.  相似文献   

9.
Chemical vapor deposition (CVD) was firstly used to simultaneously codope fluorine and boron into TiO2 nanotubes anodized Ti in C2H2O4 · 2H2O + NH4F electrolyte. F–B-codoping was successfully carried out by annealing the anodized TiO2 nanotubes through CVD, as evidenced from XPS analysis. SEM images showed that the higher the annealing temperature, the greater structure damage of F–B-codoped sample. XRD results confirmed that annealing temperature had an influence on the phase structure and boron and fluorine impurities could retard anatase–rutile phase transition. F–B-codoped samples displayed remarkably strong absorption in both UV and visible range. Under visible-light irradiation, F–B-codoped samples showed the higher Iph and catalytic activity in methyl orange photoelectrodegradation than F-doped sample and B-doped sample. This showed a convincing evidence of F–B-codoping of TiO2 had an obvious synergistic effect on the enhancement of photocurrents and photoelectrocatalytic activity.  相似文献   

10.
Titania nanotube arrays were fabricated by anodic oxidation of titanium foil in different electrolytes. The morphology, crystallinity and composition of the as-prepared nanotube arrays were studied by XRD, SEM and EDX. Electrochemical impedance spectroscopy (EIS) was employed to investigate their electrical conductivity and capacitance. Titania nanotube arrays co-adsorbed with horseradish peroxidase (HRP) and thionine chloride (Th) were studied for their sensitivity to hydrogen peroxide by means of cyclic voltammetric and galvanostatic measurements. The experiments showed that TiO2 nanotube arrays possessed appreciably different sensitivities to H2O2 due to their different conductivity. Further experiments revealed that TiO2 nanotubes have noticeably different ability of adsorbing HRP and Th, and the best sensitivity was achieved when the density of HRP is the highest. The TiO2 nanotube arrays fabricated in potassium fluoride solution demonstrated the best sensitivity on hydrogen peroxide in the range of 10−5–3 × 10−3 M at pH 6.7 and at a potential of −600 mV (vs. Ag/AgCl).  相似文献   

11.
Ti films sputtered on transparent fluorine-doped tin oxide glass substrates were anodized in fluoride-containing organic electrolyte in the presence of H2O. In this work, anodic TiO2 nanotubes (ATNs) as long as 9.2 ± 0.3 μm were obtained with high growth rate of 0.64 ± 0.3 μm min?1. We demonstrated the optimum anodization conditions for ATN growth on foreign substrates, were within the range of 0.3–0.5% (wt) NH4F, with 3–5% (vol) H2O at 60 V. XPS and ICP-MS were utilized to elucidate the increase of thickness and volume expansion obtained from the sputtered Ti film to their ATN forms. The ATN films exhibited excellent uniformity and adhesion to the substrates.  相似文献   

12.
A galvanostatic anodization is used to prepare long TiO2 nanotube arrays (TNTAs). TNTAs of over 100 μm in length, with similar nanotube size and structural regularity to the classic TNTAs made from potentiostatic mode, are achieved at 10 mA cm 2. After a post-anodization in a H3PO4-based electrolyte, the TNTAs with long nanotubes exhibit good adhesion to Ti substrate. The as-prepared long TNTAs yield a larger areal capacitance of 128.4 mF cm 2. Further, the long TNTAs possess a higher surface area, making them suitable as support templates for other active materials.  相似文献   

13.
This work shows that highly ordered and mechanically stable micrometer-long Ta2O5 nanotube arrays can be fabricated by galvanostatic anodization in a few seconds. Typically, ~ 7.7 μm long nanotubes can be grown at 1.2 A cm 2 in only 2 s. Such nanotubes can be converted to Ta3N5 nanotube arrays by nitridation. Photoelectrochemical (PEC) water splitting using AM 1.5G illumination yields for the Ta3N5 nanotube photoanode modified with cobalt phosphate (Co–Pi) remarkable photocurrents of 5.9 mA cm −2 at 1.23 VRHE and 12.9 mA cm −2 at 1.59 VRHE and after Ba-doping a value of 7.5 mA cm −2 at 1.23 VRHE is obtained.  相似文献   

14.
Titania nanotubes and nanofibers were grown by anodization of titanium in fluoride-free electrolytes composed of NaCl and KBr dissolved in either water, ethylene glycol (EG), glycerol, or 50/50 mixtures of water and EG, and water and glycerol. The tubes and fibers grew out of pits in the titanium foil. The 15 nm diameter of the nanotubes was significantly smaller, and the growth rates were much faster than those of tubes developed in fluoride solutions. Nanotubes were nucleated in all electrolytes investigated, but the tubes’ lengths were limited to a few nms in EG and glycerol. Nanofibers produced in the aqueous solutions and nanotubes formed in the 50/50 aqueous mixtures grew to many tens of microns in less than 60 s.  相似文献   

15.
N-doped TiO2 nanotubes were produced by anodization of a TiN alloy. The alloy was prepared to contain approximately 5 at.% of N from high-purity Ti and TiN powders using an arc-melting and consisted of a two-phase structure with different N-contents. Anodization of the alloy in fluoride-containing electrolyte results, under optimized conditions, in the growth of an ordered TiO2 nanotube layer on both phases. On the N-rich phase significantly smaller nanotubes are grown while on the low N-concentration phase nanotubes with larger diameter were formed. However, XPS and photoelectrochemical measurements demonstrate successful nitrogen doping of the resulting nanotube layers, which leads to a significant visible photoresponse from this material.  相似文献   

16.
Solar cells were fabricated using novel bubble-like CdSe nanoclusters sensitized highly ordered titanium oxide nanotube (TiO2 NT) array, prepared by anodization technique. The CdSe sensitization of TiO2 NT arrays was carried out by a chemical bath deposition method with freshly prepared sodium selenosufite, ammonium hydroxide and cadmium acetate dehydrate at different deposition times: 20, 40 and 60 min. The adsorption of CdSe nanoclusters on the upper and inner surface of the TiO2 NT arrays has been confirmed by field emission scanning electron and transmission electron microscopes. The results show the variation in cell a performance with different deposition times (20, 40, and 60 min) of CdSe on TiO2 NT arrays. The solar cell with CdSe, deposited for 60 min, shows reasonably high photovoltaic property compared to the reported results of similar studies. This solar cell shows the maximum photoelectric conversion efficiency of 1.56% (photocurrent of 7.19 mA/cm2; photovoltage of 0.438 V; and fill factor of 49.5%) and average incident photon to current efficiency of 50.2%. The photocurrent, incident photon-current efficiency and electron lifetime have been improved due to the increase of covered area and size of bubble-like CdSe nanoclusters on TiO2 NT arrays with the increase of deposition time.  相似文献   

17.
The solubility of newly synthesized chelating agents, i.e., tetraethylene glycol bis (2-ethylhexyl) dimethyl diphosphate (EG4EH), tetraethylene glycol bis (n-octyl) dimethyl diphosphate (EG4Oct), and tetraethylene glycol bis (2-butoxyethyl) dimethyl diphosphate (EG4BOE) in supercritical carbon dioxide (scCO2) were determined at temperatures ranging from (318.15 to 333.15) K and pressures ranging from (12 to 21) MPa. Solubility increases in the order of EG4Oct (MW = 606.33) < EG4BOE (MW = 582.26) < EG4EH (MW = 606.33), indicating that branched side chains of the ligands play an important part in increasing solubility in scCO2. Semi empirical density-based models proposed by Bartle and Chrastil were used to correlate the experimental data, and AARD values were calculated to be (1.2 to 2.9)% and (0.40 to 0.93)% for Bartle and Chrastil model, respectively. Additionally, the partial molar volumes of those compounds were estimated following the theory developed by Kumar and Johnston.  相似文献   

18.
By using suspended platinized titanium dioxide (Pt-TiO2) as a photocatalyst in an NH3 aqueous solution, NH3 was photodecomposed into H2 and N2. The amount of loaded Pt was changed between 0 and 2.0 wt% and the reaction was conducted under irradiation in an Ar atmosphere. In the Pt loading between 0 and 0.5 wt% the H2 and N2 evolution increased linearly with the Pt amount and the H2/N2 molar ratio was about 3:1. The evolution of H2 and N2 reached a maximum at 0.5 wt% Pt, and then decreased probably due to a filter effect by the Pt. The effect of pH on the photodecomposition of NH3 in the presence of Pt-TiO2 under Ar was investigated at pH from 0.68 to 13.7. The evolution of H2 increased steeply at the pH from 9 to 10 showing that the dissociation of NH4+ to free NH3 is important for the photodecomposition (pKa of NH4+/NH3 is 9.24). The time-course of H2 and N2 evolution in a 59 mM NH3 aqueous solution during 53 h irradiation gave the photodecomposition yield of 21.6%. The effect of RuO2 loading as a co-catalyst for the TiO2 or Pt-TiO2 was investigated showing that the RuO2 does not work as a H+ reduction catalyst. The activity of SrTiO3 used instead of TiO2 was only small, and the H2/N2 ratio (0.64) was very small showing that the Pt-SrTiO3 is not a good photocatalyst to decompose NH3. A TiO2/FTO nanoporous film photoanode was used in combination with a Pt counter electrode instead of the Pt-TiO2 photocatalyst, leading also to photodecomposition of NH3 to produce H2 and N2 at 3:1 ratio.  相似文献   

19.
《Chemical physics letters》2003,367(1-2):109-115
Metal-catalysed reactions between atomic H and C nanopowders in a modified CVD reactor are used to generate ordered arrays of aligned nanotubes. The technique allows to produce either deposits consisting of aligned bundles of nanotubes laying along the substrate or dense arrays of nanotubes oriented at variable angles with respect to the substrate surface. Moreover, controlled growth of densely packed nanotube arrays can be obtained at given locations of substrates by means of a process sequence which exposes patterned layers of thermally grown SiO2 onto Si(1 0 0) wafers to the reactant flows. This procedure has been successfully used to realize specific architectures, such as the selected-area growth of rather long (30–50 μm) and vertically aligned nanotubes bundles along the crossing lines of regularly shaped features.  相似文献   

20.
A new set of molar heat capacity data for aqueous {2-amino-2-hydroxymethyl-1,3-propanediol (TRIS) + glycol} at (30 to 80) °C and different concentrations (4% to 16% by weight TRIS or 56% to 44% by weight water, in a fixed amount of glycol – 40% by weight) were gathered via reliable measurement method and are presented in this report. The glycols considered were diethylene glycol (DEG), triethylene glycol (TEG), tetraethylene glycol (T4EG), propylene glycol (PG), dipropylene glycol (DPG), and tripropylene glycol (TPG). The 198 data points gathered fit the equation, Cp = Cp,a + B1m + B2m2 + B3m3, where Cp and Cp,a are the molar heat capacities of the (TRIS + glycol + water) and (water + glycol) systems, respectively, Bi the temperature-dependent parameters, and m the mole TRIS per kilogram (glycol + water). The overall average absolute deviation (AAD) of the experimental data from the corresponding values calculated from the correlation equation was 0.07%.  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