DOI:
10.1039/C5RA11202H
(Paper)
RSC Adv., 2015,
5, 73742-73751
Heteroepitaxial TiO2@W-doped VO2 core/shell nanocrystal films: preparation, characterization, and application as bifunctional window coatings
Received
12th June 2015
, Accepted 17th August 2015
First published on 17th August 2015
Abstract
In this study, we have employed rutile TiO2 nanoparticles (TiNPs) as cores for coating with vanadium sols of various concentrations and a tungsten doping of 2 at%. We grew the W-doped VO2(M) nanocrystals (WVNCs) heteroepitaxially as shells onto the TiNP cores after a sintering process. Needle-like structures gradually appeared for the WVNCs on the TiNP surfaces upon increasing the concentration of the vanadium sol coating. These needle-like structures decreased the surface contact area sufficiently to result in superhydrophobicity. Falling water droplets rebounded completely from the TiNPs@WVNC films, thereby potentially preventing fouling of such materials. The superhydrophobicity of a TiNPs@WVNC film remained stable after 90 min of UV irradiation. The WVNC shells grew epitaxially on the TiNP seeds and enhanced the visible transmittance and near-infrared switching efficiency due to the needle-like structures; this was similar to the behavior of an anti-reflection coating. In addition, the TiNP cores in the TiNPs@WVNC films completely retained their photocatalytic properties. Such bifunctional (self-cleaning and thermochromic) nanomaterials might have applications in energy-saving smart windows.
1. Introduction
Increasingly sophisticated intelligent window coatings respond to an external stimulus on an “as-needed” basis. Such materials include thermochromic coatings, which change their reflectance–transmission properties in response to temperature. Monoclinic-phase vanadium dioxide [VO2(M)] is a promising material for use in thermochromic smart windows because its reversible insulator–metal phase transition (IMPT) between monoclinic and rutile phases [VO2(M) ↔ VO2(R)], at a phase transition temperature (Tp) of approximately 338 K, dramatically affects its electrical and optical properties.1 VO2(M) is a metallic material, which exhibits high infrared reflection when the temperature is above Tp, but becomes a semiconductor with a reasonable infrared transmission when the temperature is below Tp.2,3 The modulation of Tp to room temperature is essential for many practical applications, and not just limited to smart windows. A number of approaches have been reported for decreasing Tp; for example, doping VO2(M) with metal ions (e.g., W6+, Mo6+, Ta5+, Nb5+, and Ru4+) is a commonly used strategy, but it often decreases the thermochromic performance.4
Rutile TiO2 has the same structure as VO2(R) and considerably similar lattice parameters. Because the crystal structure of VO2 is sensitive to its under- and/or over coating materials, rutile TiO2 nanocrystals might be useful as seeds for assisting the nucleation and growth of nanostructured rutile VO2. In the field of VO2-based thermochromic smart windows, TiO2 is also an important composite layer because of its good compatibility and crystalline structure similar to that of VO2.5 It has been reported that TiO2 films are effective underneath layers or cover layers for inducing the crystallization of VO2(M/R),6 constructing heterostructures for photoemission,7,8 enhancing the oxidation durability,9,10 regulating the phase transition properties,11 and improving the optical performance of VO2.1,12 Moreover, VO2(M)/TiO2 composites have their own advantages, not only for increasing visible transmittance but also for modifying the infrared modulation ability.13,14 Compared to TiO2–VO2 multilayered structures, single-layered composite films are more attractive because of their ease of preparation, mixing homogeneity of the two components, and, most importantly, their display of near-infrared (NIR) modulation behavior at wavelengths that are significantly shorter than those for typical thin films of pure VO2, thus providing advantages when applied in systems requiring efficient visible transmittance and NIR switching.15
Thin films of both TiO2 and VO2 can be prepared using a range of methods, primarily chemical vapor deposition (CVD),16,17 sol–gel processing,17–19 and physical vapor deposition (PVD).20,21 CVD methods have the advantage of being more easily integrated into float-glass production lines. Although these methods can provide single-layered TiO2(A)–VO2 through the simultaneous deposition of TiO2 and VO2 at a critical temperature,22,23 the diffusion of Ti atoms into VO2 crystal lattices weakens the thermochromic performance of VO2.24 The reported methods for fabricating VO2(M)/TiO2 composites are limited to using PVD, which cannot meet the requirements for the production of large-area VO2 thermochromic windows because of technical and cost problems. Because the crystallization temperature of TiO2(A) is considerably lower than that of VO2(R),25 a “two-step method” method has been developed for producing TiO2(A)/VO2 single-layered composite films, in which VO2 nanoparticles (NPs) are first prepared and dispersed in a titanium sol to form a composite film and then annealed for TiO2 crystallization. This method largely restricts Ti diffusion and maintains the thermochromic performance of VO2. Thus, the resulting materials become more cost-effective for large-scale coatings of glass. It has been demonstrated recently that combining TiO2 and VO2 within a phase composite produces a material that possesses both self-cleaning and thermochromic properties.26,27 Furthermore, the films exhibit a decrease of 17 °C in the thermochromic switching temperature (Tp) to 51 °C. If this Tp value could be lowered further, such coatings would be strong candidates for application in self-cleaning solar control glazing. In this study, we report a “two-step method” for the preparation of TiO2/VO2 composites, in which rutile TiO2 nanoparticles (TiNPs) were first dispersed in a W-doped vanadium sol to obtain a composite film, which was then annealed to ensure the crystallization of the W-doped VO2. We found that the specific needle-like structures of the W-doped VO2 nanocrystals (WVNCs) could be regulated by varying the ratio of TiO2 cores to VO2 shells. The films of the as-prepared composite particles on glass exhibited remarkable superhydrophobicity and hysteresis widths that were considerably narrower than that found for pure VO2. Furthermore, this specific morphology increased the visible transmittance of the film and enhanced its regulation ability. Such films performed three functions: thermochromism (from the VO2(M) layer) for solar energy modulation, photocatalysis (due to the TiNPs), and superhydrophobicity (for anti-fouling and self-cleaning effects). The as-obtained nanocomposite coatings appear to have applicability in smart windows.
2. Experimental section
2.1 Materials
Vanadium pentoxide (V2O5, Acros Organics, >99.6%) and tungsten(VI) oxide (WO3, Hayashi Pure Chemical Industries, >99%) were employed as starting materials to prepare VO2+ solutions. Rutile TiNPs (sizes: 11–20 nm; Acros Organics; CAS no. 13463-67-7) were used as received. Methylene blue (C16H18ClN3S, MB, 98%) was purchased from Aldrich. All other chemicals and solvents were of reagent grade and purchased from Aldrich. All the reagents were used without further purification.
2.2 TiNPs@WVNCs films
TiNPs (ca. 0.02–0.05 g) were ultrasonically dispersed in anhydrous EtOH (50 mL) for 15 min. An aqueous solution of the Triton X-100 surfactant (Sigma Aldrich, 800 μL, 0.1 M) was added into the suspension, which was then stirred for 2 h. Another stock aqueous sol was produced using a previously reported method as follows:28 V2O5 powder (15 g) was heated in a crucible until it was molten and then poured into distilled water at room temperature. After vigorous stirring, the sol was filtered, and solutions of various concentrations (ca. 1, 2, 3, and 4 at%) were obtained. Tungsten(VI) oxide powder was dissolved in a few drops of distilled water and added to the sol in quantities calculated to achieve a doping level (assuming 100% substitution) of 2 at%. The stock solution was subsequently added dropwise into the TiNPs suspension, followed by sonication for 30 min prior to use. Soda lime glass slides were ultrasonically cleaned for 10 min in 20% EtOH, 20% acetone, and deionized water. After drying the glass slides by purging with N2, they were subjected to argon/oxygen plasma (AOP) treatment using a TCP 9400SE instrument (Lam Research).29 The surface was chemically modified (strongly hydrophilic or polar) due to the AOP treatment. The introduction of these polar groups provided a more polar surface for the formation of the film. For each coating, the sol was spin-coated on the AOP-treated glass slides, and then dried in a clean air cabinet for 10–15 min. The TiNPs@WVNC films on the AOP-treated glass substrates were obtained after spin-coating at a speed of 3000 rpm. To obtain a crystalline coating, the as-obtained films were sintered in a vacuum quartz tube furnace under a low pressure of a reducing gas mixture (50% CO and 50% CO2). Typically, at a pressure of approximately 3 Torr, the sample was heated from ambient temperature to 500 °C at a heating rate of approximately 20 °C min−1 and then held at 500 °C for 2 h, before allowing it to cool naturally to room temperature. The annealing program was optimized to transform the amorphous shells into crystallized VO2(M) while keeping the TiO2 cores chemically intact. The samples of TiNPs that had been coated with the 1, 2, 3, and 4 at% vanadium sols and then sintered are denoted herein as TV1, TV2, TV3, and TV4, respectively.
2.3 Characterization of heteroepitaxial TiNPs@WVNCs films
The valence states of vanadium, tungsten, and titanium were studied using X-ray photoelectron spectroscopy (XPS; Scientific Theta Probe, UK). The structures of the samples were determined using X-ray diffraction (XRD; Rigaku D/max-RC). The phase transition behavior was analyzed through differential scanning calorimetry (DSC; Netzsch DSC-4000) at a temperature ramp rate of 10 °C min−1 within the temperature range from −20 to +100 °C under a flowing N2 atmosphere. The chemical compositions of the as-obtained samples were investigated using energy-dispersive X-ray spectrometry (EDX) combined with high-resolution scanning electron microscopy (HR-SEM; JEOL JSM-6500F, Japan). Transmission electron microscopy (TEM) images were obtained using a field emission transmission electron microscope (Philips Tecnai G2 F20) operated at an accelerating voltage of 200 kV. Static water contact angles (SWCAs) were measured using a FDSA MagicDroplet-100 contact angle goniometer. Each SWCA was determined, under normal laboratory ambient conditions at room temperature under 40% humidity, by fitting a Young–Laplace curve around the drop. Each mean value was calculated from at least 10 individual measurements; the measurement error was less than 3°. As-prepared superhydrophobic surfaces were mounted on a rotatable stage. The stage was tilted until a droplet on the stage slid, thus characterizing the roll-off angle of the surface. The tilt angle of the surface was fixed at a certain angle. The changes in the SWCAs after UV irradiation were monitored at specific time intervals.
2.4 TiNPs@WVNC films for applications in bifunctional smart windows
The TiNPs@WVNC films underwent a reversible IMPT, which was accompanied by remarkable changes in their optical properties. Optical transmittances were monitored using a UV-vis-near-infrared spectrophotometer (Shimadzu UV3600) equipped with a thermo-regulated environmental cell. The reversible IMPT temperatures of the thin films were measured by recording the transmittance spectra as a function of temperature. For all the samples, the integral visible transmittance (Tint) was obtained based on the measured spectra, using the following equation:30 |
 | (1) |
where X denotes the transmittance measured using a UV-vis-near-infrared spectrophotometer. The values of Tint were obtained using the formula ϕρ = Φlum, where Φlum was the standard luminous efficiency function, and Φlum is 0 beyond this range.
The photocatalytic performance of the synthesized TiNPs@WVNC films was determined by evaluating the rate of degradation of MB. A glass plate (1 × 1 cm) presenting a thin TiNPs@WVNC film was immersed in an aqueous MB solution (20 ppm, 20 mL). The sample was placed in the dark to allow the complete absorption of MB onto the surface; it was then irradiated using a UV lamp. The concentration of MB in the solution was measured as a function of the irradiation time, monitoring the wavelength of maximum absorbance (λmax) of MB at 664 nm. The photometric analysis of the TiNPs@WVNC films before and after irradiation was performed to measure the degradation efficiency of MB (D%), which is defined using the following expression:11
|
 | (2) |
where
C0 is the initial concentration of MB and
C is the concentration of MB after irradiation of the samples for a certain period of time. The absorbance of the samples was measured using a UV-vis spectrophotometer with a resolution of 1 nm (Perkin-Elmer, lambda 25, 190–1100 nm). The decrease in the absorbance at
λmax of the samples after irradiation for a desired period of time allowed the determination of the rate of decolorization and, thus, the MB photodegradation efficiency, which also represents the activity of the TiNPs@WVNC films. The following equation was used to measure the rate of MB degradation (
k) at a given time:
|
 | (3) |
For comparison, thin films of a commercially available pure TiO2 catalyst were also examined under similar conditions.
3. Results and discussion
3.1 Characterization and heteroepitaxy of TiNPs@WVNC films
The synthesis of the TiNPs@WVNC films in this study can be regarded as a process of generating “TiNP seeds” upon which the VO2 crystals grew. We used sol–gel deposition to produce uniform precursor shells with various thicknesses on the TiO2 cores. We used XPS to investigate the chemical states on the surfaces of the as-obtained TiNPs@WVNC films. Fig. 1a indicates that the sample consists of only six elements: carbon, vanadium, titanium, oxygen, silicon, and tungsten. The presence of signals for carbon and silicon atoms presumably reflected surface contamination and the glass slides, respectively. For the vanadium precursor film, we assigned the peaks centered at 517.15 and 524.2 eV to the V 2p3/2 and V 2p1/2 orbitals, respectively. The signal for the V 2p3/2 orbitals of the sample appeared at a binding energy of 517.8 eV, which can be attributed to V5+.31 In addition, two peaks were centered at 529.6 and 532.2 eV in the O 1s region (Fig. 1b). We assigned the major peak, positioned at the lower binding energy of 529.6 eV, to the O2− ions in the V–O bonds.32 We assigned the second peak, located at 532.2 eV, to OH− groups in the H2O molecules. The V 2p core-level XPS spectrum recorded after sintering the sample revealed (Fig. 1c) that the V 2p doublet binding energies were 516.85 and 524.32 eV, corresponding to spin–orbital splitting of the V 2p3/2 and V 2p1/2 components, respectively.33 The V 2p3/2 peak can be split into major and minor peaks at 516.85 and 515.7 eV, respectively. The major V 2p3/2 peak shifted from 517.15 to 516.85 eV after sintering, suggesting the generation of V4+ species. The binding energy for V 2p3/2 peak was slightly higher than that found for pure VO2, but was in agreement with the value (516.3 eV) for W-doped VO2,34 suggesting that the V 2p3/2 binding energy increased slightly after W-doping and that the vanadium atoms in the doped sample resided in the +4 oxidation state. The minor V 2p3/2 peak at 515.7 eV could be assigned to the V–Ti bonds.35 We attributed the binding energies at 458.15 and 464.25 eV to the Ti 2p3/2 and Ti 2p1/2 orbitals, respectively, of the rutile-phase TiNPs (Fig. 1d), corresponding to the position for Ti4+ species in TiNPs (the reported value for pure Ti is 458.0–458.5 eV (ref. 25)). Fig. 1e displays the signals for the W 4f species in the sample, with binding energies of W 4f5/2 and W 4f7/2 centered at 37.7 and 35.7 eV, respectively. According to the standard binding energies, there was only a small amount of tungsten in the sample, with the oxidation state of the tungsten ions in these films being solely W6+. Table 1 lists the ratios of titanium to vanadium (Ti/V) in the TV1, TV2, TV3, and TV4 samples. We observed that the Ti/V ratio decreased upon increasing the concentration of the vanadium sol coating on the TiNPs.
 |
| Fig. 1 (a) Wide-range survey XPS spectra of TiNPs@WVNC films on AOP-treated glass slides. (b–d) High-resolution XPS profiles of the V 2p core level spectra of the (b) vanadium sol and (c) the TV3 film; (d) Ti 2p and (d) W 4f core level spectra of the TV3 film. | |
Table 1 Influence of TiNP seeds on thermochromic and photocatalytic properties
Sample |
Ti/Va (atom%) |
Ti/Vb (atom%) |
Tpc (°C) |
ΔTHd (°C) |
Roughnesse (nm) |
Tint,20f (%) |
Tint,50g (%) |
Esh (%) |
Ki |
Obtained from XPS and EDX data, respectively. Obtained from XPS and EDX data, respectively. Obtained from DSC data. Hysteresis between the heating and cooling cycles. Obtained from AFM images. Calculated by eqn (1) from 700 to 2500 nm at 20 and 50 °C, respectively. Calculated by eqn (1) from 700 to 2500 nm at 20 and 50 °C, respectively. Changes in the values of Tint at 20 and 50 °C. Rate of MB degradation at 664 nm, as determined from the UV spectra. |
TV1 |
1.61 |
1.54 |
43.1 |
5.6 |
279 |
43.3 |
37.5 |
13.4 |
0.0013 |
TV2 |
0.83 |
0.89 |
37.8 |
4.8 |
308 |
49.5 |
39.7 |
19.8 |
0.0015 |
TV3 |
0.32 |
0.39 |
31.4 |
3.4 |
632 |
56.4 |
39.3 |
30.3 |
0.0017 |
TV4 |
0.25 |
0.28 |
29.8 |
2.7 |
643 |
58.3 |
40.1 |
31.2 |
0.0018 |
Fig. 2 displays the XRD patterns of the WVNC powders TV1, TV2, and TV4 and of the pure TiNPs. The observed diffraction peaks of the pure WVNC powders can be indexed to the monoclinic phase of VO2(M) (JCPDS card no. 43-1051). No noticeable changes in the positions of the diffraction peaks occurred for samples TV1, TV2, and TV4, but they had shifted to a slightly lower angle when compared with that of the sample prepared without the TiO2 cores; this shift could have resulted from the superposition of the diffraction peaks from rutile TiO2 and VO2(M), both near 27.8°. For TV1, TV2, and TV4, we observed the XRD peaks for both VO2(M) and rutile TiO2. The intensity of the VO2(M) peak increased upon increasing the thickness of the coating on the TiNPs, which was consistent with the formation of VO2(M) shells.
 |
| Fig. 2 XRD patterns for the pure TiNPs, TV1, TV2, TV3, TV4, and pure WVNC films. The diamonds and circles in the XRD patterns mark the main peaks of the M-phase VO2 and rutile TiO2, respectively. | |
Endothermic and exothermic peaks were evident in each DSC curve (Fig. 3), further confirming the formation of VO2(M). Table 1 lists the values of Tp in the heating cycle and hysteresis (ΔTH) between the heating and cooling cycles, recorded from the DSC curves for TV1, TV2, TV3, and TV4. The value of Tp increased upon increasing the TiNP seed content, despite having the same W doping content (e.g., Tp increased from 29.8 to 34.1 °C when the Ti/V molar ratio increased from 0.25 to 1.61). This result suggests that the addition of the TiNP seeds suppressed the W doping content in VO2(M). The latent heats calculated from the heating cycles, as shown in Fig. 3, were 38.5 and 15.7 J g−1 for the WVNCs prepared with and without TiNP seeds, respectively. The higher latent heat indicates that the TiNPs@WVNCs were highly crystalline and relatively perfect in their crystalline structure.36 The hysteresis between the heating and cooling cycles increased upon increasing the content of TiO2 seeds (from 2.7 to 5.6 °C), which is in contrast to the reported result that TiO2 additives can remarkably narrow the hysteresis loop width;37 we suspect that this was due to the effect of the morphology of the TiNPs@WVNCs.38
 |
| Fig. 3 DSC curves of WVNCs in the (a) absence and (b) presence of TiNP seeds (Ti/V molar ratio: 0.32). | |
We used SEM to characterize the surface morphologies and nanostructures of our samples. Fig. 4 displays the SEM images and EDX analyses of the TiNPs coated using the 1, 2, 3, and 4 at% vanadium sols. The vanadium sols generated homogeneous thin films on the TiNPs prior to sintering, resulting in similar morphologies for the four samples (Fig. 4a–d). The thicker shells tended to assemble because of the lower number of heterogeneous nucleation sites.39 Crystalline structures appeared for the films after sintering, due to the formation of VO2 nanocrystals. Needle-like structures appeared initially on the TiNP surfaces of TV1 (Fig. 5a); their lengths ranged from 100 to 500 nm, indicating irregular thicknesses of the vanadium sol coating. For TV2, the needle-like structures of the WVNCs covered the entire surfaces of the TiNPs with more regular lengths ranging from 70 to 150 nm (Fig. 5b). Further increasing the concentration of the precursor sol to 3 at% caused the needle-like structures to exhibit high regularity (lengths of ca. 300 nm) on the surfaces of the TiNPs (Fig. 5c). An urchin-like structure appeared for the WVNCs of TV4, indicating that the aggregation of the TiNP cores had occurred on the TiNPs at a vanadium sol coating of 4 at% (Fig. 5d). In the absence of TiO2 seeds, VO2(M) exhibited an irregular shape.40 The roughnesses measured using AFM, presented in Table 1, confirm that the WVNC coatings had significantly different morphologies and Ti/V ratios in the TV1, TV2, TV3, and TV4 samples. After annealing, the core/shell interface became obscure (Fig. 6a), indicating that the shell coating crystallized with a relatively rough morphology. The needle-like structures of the WVNCs on the surfaces of the TiNPs were clearly evident for TV3. The lattice-resolved HRTEM images of TV3 revealed distinguishable crystallographic lattice fringes, with the lattices coherently bound to the TiO2 side faces (Fig. 6b). We could distinguish two different lattices for VO2 and TiO2, with nearly orthogonal and oblique crossing patterns (represented schematically in the rectangular region), respectively. The interplanar distances of 0.332 nm in the yellow region and 0.324 nm in the green region match well with the (−111) and (110) crystal planes of VO2(M) and rutile TO2, respectively, suggesting that the VO2(M) crystals grew epitaxially on the rutile TiNPs; the appearance of a dislocation in the interface region, arising from the slightly different interplanar distances in VO2(M) and rutile TiNPs, confirmed this epitaxial growth. Thus, a coupled interface existed between the VO2 and TiO2 phases, with the surfaces of the TiO2 cores serving as propitious sites for heterogeneous nucleation of the VO2 phase. We performed EDS analysis of the shell coating, applying a beam spot size of less than 1 nm. Signals were detected for both V and Ti co-existing on the shell, indicating some degree of Ti–V diffusion at the core/shell interface, presumably because of the infinite solubility of Ti in VO2 lattices.35 This finding also suggests that rutile TiO2 seeds can induce the epitaxial growth of VO2(M) nanocrystals.
 |
| Fig. 4 Top-view SEM images of the TiNPs coated with (a) 1, (b) 2, (c) 3, and (d) 4 at% vanadium sol without sintering. | |
 |
| Fig. 5 Top-view SEM images of the TiNPs@WVNC films TV1, TV2, TV3, and TV4. | |
 |
| Fig. 6 (a) TEM and (b) HRTEM images of a TiNPs@WVNC film prepared at a Ti/V molar ratio of 0.32. | |
When a liquid droplet was placed in contact with a superhydrophobic surface that was capable of self-cleaning, a high contact angle was formed with a small interfacial area between the liquid and the surface. In an ideal situation, the liquid would not wet the surface and would be free to roll off from it. This surface property, caused by the surface energy of the solid being lower than that of the liquid, results from either the chemistry of the material or its physical roughness, including both nanometer- and micrometer-scale features. Fig. 7a presents the SWCAs and roll-off angles of the surfaces of pure TiNPs, TV1, TV2, TV3, and TV4. Water droplets on the pure TiO2 surface and TV1, TV2, TV3, and TV4 surfaces provided SWCAs of 85° ± 3°, 139° ± 3°, 151° ± 3°, 163° ± 3°, and 155° ± 3°, respectively. Thus, the SWCAs of the heteroepitaxial surfaces were larger than that found for the pure-TiNP surface. The SWCAs of the surfaces increased gradually upon increasing the fraction of needle-like WVNC structures up to that in TV3, but did not change significantly thereafter. Prior to UV irradiation, the water droplets on the pure TiNPs, TV1 and TV2 surfaces exhibited roll-off angles of 47° ± 3°, 18° ± 3°, and 11° ± 3°, respectively; increasing the density of needle-like structures even further, to obtain the TV3 and TV4 surfaces, caused the roll-off angles to decrease abruptly to approximately 1° ± 3°. The water droplets penetrated the TV1 and TV2 surfaces because of their lower densities of needle-like WVNC structures, resulting in relatively large roll-off angles. When water droplets resided on the TV3 and TV4 surfaces, with their greater fractional coverages of needle-like structures, air was presumably trapped in the cavities of these rough surfaces, resulting in composite solid–air–liquid interfaces and, therefore, relatively low roll-off angles. Fig. 7b and c display the SWCAs and roll-off angles, respectively, of the pure TiNPs surface and the TV1, TV2, TV3, and TV4 surfaces, plotted with respect to the UV irradiation time. VO2(M) is a well-known semiconductor material with a band gap of 0.7 eV, and is seldom used as a photocatalyst. UV irradiation may not always change the VO2(M), and VO2(M) is not always oxidized under UV irradiation. The SWCA of the pure TiNPs surface was affected significantly, forming a completely hydrophilic surface (SWCA: <5°) upon increasing the UV irradiation time to 100 min (Fig. 7b). The SWCAs of the TV1 and TV2 surfaces decreased gradually to 113° and 133°, respectively, within 180 min because of their lower densities of needle-like structures. The SWCAs of TV3 and TV4 did not change significantly within 180 min of UV irradiation, due to their lower surface ratios—the fraction of the surface area over the total geometric area of a sample; this indicates the grooving on the surface, which allows air pockets to be formed between the solid surface and the overlaying liquid droplet. Fig. 7c plots the roll-off angles with respect to the UV irradiation time for the pure TiNPs surface and the TV1, TV2, TV3, and TV4 surfaces; data points marked with a roll-off angle of 90° represent droplets that were pinned to the substrates even when tilted vertically or flipped upside down. As expected, the roll-off angles for the pure TiNPs surface and the TV1 and TV2 substrates increased upon increasing the UV irradiation time, eventually reaching a state where the water droplet remained stuck to the surface even when it was turned by 90°. The roll-off angles of TV3 and TV4 did not change significantly within 180 min of UV irradiation. Such highly stable superhydrophobicity after UV irradiation may extend the lifetimes of smart windows. Impact experiments revealed that the needle-like structures exhibited superior slippage and robustness of their superhydrophobicity. We allowed water droplets to fall onto 35°-tilted pure TiNP and TV3 surfaces to observe their stickiness and rebound ability. We released droplets with a radius of 1 mm at rest from a height of 4.5 mm. As shown in Fig. 7d and e, the substrates were tilted at an angle of 35° to monitor their rebound trajectories. For the pure TiNPs surface, the droplets stuck to the surface (Fig. 7d). When needle-like WVNC structures were present on the TV3 surface, the droplets bounced completely from the substrate (Fig. 7e). Thus, needle-like structures improved the slippage of droplets on the pure TiNPs surfaces, resulting in effective rebounding.
 |
| Fig. 7 (a) SWCAs and roll-off angles of pure TiNPs, TV1, TV2, TV3, and TV4 surfaces. (b) SWCAs and (c) roll-off angles of pure TiNPs, TV1, TV2, TV3 and TV4 surfaces, plotted with respect to the UV irradiation time. Real-time images of the water droplets impacting on (d) pure TiNPs and (e) TV3 surfaces inclined at an angle of 35°. | |
3.2 TiNP@WVNC films for applications in bifunctional smart windows
The values of Tint and the NIR switching efficiency (Es) are two main parameters used to characterize smart windows. Although these two parameters are strongly related, their changes can be conflicting.41 A large visible transmittance is usually accompanied by a small Es value, and vice versa. Developing methods to improve the Es value while simultaneously maintaining a comparatively high visible transmittance has been challenging. We studied the thermochromic properties of the TiNPs@WVNC films by measuring the optical transmittance spectra of TV1, TV2, TV3, and TV4. Fig. 8a reveals that all these films displayed excellent visible transmittance at 20 and 50 °C as well as excellent Es values. For TV1, TV2, TV3 and TV4 films that were coated with a thickness of approximately 87 nm, the values of Tint for the semi-conductive states reached 31%, 44%, 62%, and 64%, respectively, with Es values in the range of 20–45% (Table 1). The values of Tint for TV3 and TV4 are higher than those for single-layer VO2 coatings with similar film thicknesses (reported values for 50 nm-thick films include 40% and 45%).42 Thus, the needle-like structured surfaces can be regarded as anti-reflecting coatings (ARCs) with enhanced transmittance.43 Single-layer VO2 crystals rarely display optical performance that includes high Tint values as well as good Es values. Thus, our current solution-based method appears to have many advantages over other preparation methods.
 |
| Fig. 8 (a) Transmittance spectra of the TiNPs@WVNCs of TV3 at a thickness of approximately 87 nm. The solid and dashed lines represent the spectra obtained at 50 and 20 °C, respectively. | |
To test the applicability of the photocatalytic effects in smart windows, we immersed our TiNPs@WVNC films into a reaction solution containing MB for various lengths of time. Fig. 8b displays the absorption of a solution of MB after UV irradiation in the presence of the TV3 film (0 min: time at which the container was placed in the dark). The TiNP cores coated with the WVNC shells retained their photocatalytic properties, and electron/hole pairs were excited under irradiation of UV light. The rate of MB degradation (k) over thin films of commercial TiNPs (particle size: 20 nm; P25) was approximately 0.0013. Apart from TV1, the rates of MB degradation over the thin films of TiNPs@WVNC were all higher than that found over the commercial TiNPs (Table 1). We suspect that the heteroepitaxial crystallinity induced the photocatalytic ability of VO2, and thus enhanced the rate of MB degradation. Therefore, these heteroepitaxial TiNPs@WVNC systems exhibited bifunctional properties—superior thermochromicity and photoactivity—under UV irradiation, due to their large surface areas compared to those of the thin films of pure TiNPs.
4. Conclusion
TiNPs can be heteroepitaxially coated with WVNCs, resulting in needle-like structured nanocrystals on their surfaces. The morphology changed from the particle shape of rutile TiNPs in the absence of the vanadium sol coating to needle-like shapes for the TiNPs@WVNCs. Films of the needle-like structured TiNPs@WVNCs exhibit superhydrophobicity (SWCA: >150°) and can completely rebound water droplets. The needle-like structures of the heteroepitaxial nanocrystals substantially enhanced the modulation of the visible transmittance. In addition, the TiNP cores within the W-doped VO2 shells completely retained their photocatalytic properties. This study not only provides a simple method for synthesizing core/shell TiNPs@WVNC films for bi-functional (photocatalytic and thermochromic) window coatings but also demonstrates their stable superhydrophobicity after UV irradiation.
Acknowledgements
We thank the Ministry of Science and Technology of the Republic of China for supporting this research financially.
References
- J. H. Park, J. M. Coy, S. Kasirga, C. Huang, Z. Fei, S. Hunter and D. H. Cobden, Measurement of a solid-state triple point at the metal–insulator transition in VO2, Nature, 2013, 500, 431–434 CrossRef CAS PubMed.
- F. C. Case, Improved VO2 thin films for infrared switching, Appl. Opt., 1991, 30, 4119–4123 CrossRef PubMed.
- F. Beteille and J. Livage, Optical switching in VO2 thin films, J. Sol-Gel Sci. Technol., 1998, 13, 915–921 CrossRef CAS.
- J. B. Goodenough, The two components of the crystallographic transition in VO2, J. Solid State Chem., 1971, 3, 490–500 CrossRef CAS.
- M. Hervieu, The surface science of metal oxides, Adv. Mater., 1995, 7, 91–92 CrossRef PubMed.
- Y. Muraoka and Z. Hiroi, Metal–insulator transition of VO2 thin films grown on TiO2 (001) and (110) substrates, Appl. Phys. Lett., 2002, 80, 583–585 CrossRef CAS PubMed.
- K. Okazaki, H. Wadati, A. Fujimori, M. Onoda, Y. Muraoka and Z. Hiroi, Photoemission study of the metal–insulator transition in VO2/TiO2 (001): Evidence for strong electron–electron and electron–phonon interaction, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69, 165104 CrossRef.
- R. Eguchi, S. Shin, A. Fukushima, T. Kiss, T. Shimojima, Y. Muraoka and Z. Hiroi, Thermoelectric properties and figure of merit of a Te-doped InSb bulk single crystal, Appl. Phys. Lett., 2005, 87, 201902 CrossRef PubMed.
- Z. T. Zhang, Y. F. Gao, L. T. Kang, J. Du and H. J. Luo, Effects of a TiO2 buffer layer on solution-deposited VO2 films: Enhanced oxidization durability, J. Phys. Chem. C, 2010, 114, 22214–22220 CAS.
- G. Fu, A. Polity, N. Volbers and B. K. Meyer, Annealing effects on VO2 thin films deposited by reactive sputtering, Thin Solid Films, 2006, 515, 2519–2522 CrossRef CAS PubMed.
- J. Chen, Y. He and C. Li, Synthesis of doubly functional nanowhiskers from a colloid solution including a V–Ti complex precursor, J. Mater. Sci., 2012, 47, 5287–5297 CrossRef CAS.
- L. Kang, Y. Gao, H. Luo, Z. Chen, J. Du and Z. Zhang, Nanoporous thermochromic VO(2) films with low optical constants, enhanced luminous transmittance and thermochromic properties, ACS Appl. Mater. Interfaces, 2011, 3, 135–138 CAS.
- M. Wilkinson, A. Kafizas, S. M. Bawaked, A. Y. Obaid, S. A. Al-Thaiti, S. N. Basahel, C. J. Carmalt and I. P. Parkin, Combinatorial atmospheric pressure chemical vapor deposition of graded TiO2–VO2 mixed-phase composites and their dual functional property as self-cleaning and photochromic window coating, ACS Comb. Sci., 2013, 15, 309–319 CrossRef CAS PubMed.
- N. R. Mlyuka, G. A. Niklasson and C. G. Granqvist, Thermochromic multilayer films of VO2 and TiO2 with enhanced transmittance, Sol. Energy Mater. Sol. Cells, 2009, 93, 1685–1687 CrossRef CAS PubMed.
- S. Y. Li, G. A. Niklasson and C. G. Granqvist, Nanothermochromics: Calculations for VO2 nanoparticles in dielectric hosts show much improved luminous transmittance and solar energy transmittance modulation, J. Appl. Phys., 2010, 108, 063525 CrossRef PubMed.
- T. D. Manning, I. P. Parkin, C. S. Blackman and U. Qureshi, APCVD of thermochromic vanadium dioxide thin films–solid solutions V2−xMxO2 (M = Mo, Nb) or composites VO2–SnO2, J. Mater. Chem., 2005, 15, 4560–4566 RSC.
- S. A. O'Neill, R. J. H. Clark, I. P. Parkin, N. Elliott and A. Mills, Anatase thin films on glass from the chemical vapor deposition of titanium(IV) chloride and ethyl acetate, Chem. Mater., 2003, 15, 46–50 CrossRef.
- A. Kafizas, C. W. Dunnill and I. P. Parkin, The relationship between photocatalytic activity and photochromic state of nanoparticulate silver surface loaded titanium dioxide thin-films, Phys. Chem. Chem. Phys., 2011, 13, 13827–13838 RSC.
- D. P. Partlow, S. R. Gurkovich, K. C. Radford and L. J. Denes, Switchable vanadium oxide films by a sol–gel process, J. Appl. Phys., 1991, 70, 443 CrossRef CAS PubMed.
- N. Martin, C. Rousselot, D. Rondot, F. Palmino and R. Mercier, Microstructure modification of amorphous titanium oxide thin films during annealing treatment, Thin Solid Films, 1997, 300, 113–121 CrossRef CAS.
- M.-H. Lee and M.-G. Kim, RTA and stoichiometry effect on the thermochromism of VO2 thin films, Thin Solid Films, 1996, 286, 219–222 CrossRef CAS.
- H. Kakiuchida, P. Jin and M. Tazawa, Optical characterization of vanadium titanium oxide films, Thin Solid Films, 2008, 516, 4563–4567 CrossRef CAS PubMed.
- J. Du, Y. F. Gao, H. J. Luo, L. T. Kang, Z. T. Zhang, Z. Chen and C. X. Cao, Significant changes in phase-transition hysteresis for Ti-doped VO2 films prepared by polymer-assisted deposition, Sol. Energy Mater. Sol. Cells, 2011, 95, 469–475 CrossRef CAS PubMed.
- I. Balberg, B. Abeles and Y. Arie, Phase transition in reactively co-sputtered films of VO2 and TiO2, Thin Solid Films, 1974, 24, 307–310 CrossRef CAS.
- R. Li, S. D. Ji, Y. M. Li, Y. F. Gao, H. J. Luo and P. Jin, Synthesis and characterization of plate-like VO2(M)@SiO2 nanoparticles and their application to smart window, Mater. Lett., 2013, 110, 241–244 CrossRef CAS PubMed.
- U. Qureshi, T. D. Manning, C. Blackman and I. P. Parkin, Composite thermochromic thin films: (TiO2)–(VO2) prepared from titanium isopropoxide, VOCl3 and water, Polyhedron, 2006, 25, 334–338 CrossRef CAS PubMed.
- U. Qureshi, T. D. Manning and I. P. Parkin, Atmospheric pressure chemical vapour deposition of VO2 and VO2/TiO2 films from the reaction of VOCl3, TiCl4 and water, J. Mater. Chem., 2004, 14, 1190–1194 RSC.
- T. J. Hanlon, J. A. Coath and M. A. Richardson, Molybdenum-doped vanadium dioxide coatings on glass produced by the aqueous sol–gel method, Thin Solid Films, 2003, 436, 269–272 CrossRef CAS.
- J. K. Chen, C. Y. Hsieh, C. F. Huang, P. M. Li, S. W. Kuo and F. C. Chang, Macromolecules, 2008, 41, 8729–8736 CrossRef CAS.
- C. X. Cao, Y. F. Gao and H. J. Luo, Pure single-crystal rutile vanadium dioxide powders: Synthesis, mechanism and phase-transformation property, J. Phys. Chem. C, 2008, 112, 18810–18814 CAS.
- Y. Wang and Z. Zhang, Synthesis and transport properties of nanostructured VO2 by mechanochemical processing, Phys. E, 2009, 41, 548–551 CrossRef CAS PubMed.
- C. J. Cui, G. M. Wu, J. Shen, B. Zhou, Z. H. Zhang, H. Y. Yang and S. F. She, Synthesis and electrochemical performance of lithium vanadium oxide nanotubes as cathodes for rechargeable lithium-ion batteries, Electrochim. Acta, 2010, 55, 2536–2541 CrossRef CAS PubMed.
- Y. Zhang, X. Liu, G. Xie, L. Yu, S. Yi, M. Hu and C. Huang, Hydrothermal synthesis, characterization, formation mechanism and electrochemical property of V3O7·H2O single-crystal nanobelts, Mater. Sci. Eng., B, 2010, 175, 164–171 CrossRef CAS PubMed.
- F. Li, X. H. Wang, C. L. Shao, R. X. Tan and Y. C. Liu, Biodegradable polyester hybrid nanocomposites containing titanium dioxide network and poly(ε-caprolactone): Synthesis and characterization, Mater. Lett., 2007, 61, 1328–1332 CrossRef CAS PubMed.
- Y. Li, S. Ji, Y. Gao, H. Luo and P. Jin, Modification of Mott phase transition characteristics in VO2@TiO2 core/shell nanostructures by misfit-strained heteroepitaxy, ACS Appl. Mater. Interfaces, 2013, 5, 6603–6614 CAS.
- Z. Chen, Y. F. Gao, L. T. Kang, C. X. Cao, S. Chen and H. Luo, Fine crystalline VO2 nanoparticles: Synthesis, abnormal phase transition temperatures and excellent optical properties of a derived VO2 nanocomposite foil, J. Mater. Chem. A, 2014, 2, 2718–2727 CAS.
- Y. Li, S. Ji, Y. Gao, H. Luo and M. Kanehira, Core/shell VO2@TiO2 nanorods that combine thermochromic and photocatalytic properties for application as energy-saving smart coatings, Sci. Rep., 2013, 3, 1370–1373 Search PubMed.
- J. Y. Suh, R. Lopez, L. C. Feldman and R. F. Haglund, Semiconductor to metal phase transition in the nucleation and growth of VO2 nanoparticles and thin films, J. Appl. Phys., 2004, 96, 1209–1212 CrossRef CAS PubMed.
- A. F. Demirörs, A. van Blaaderen and A. Imhof, A general method to coat colloidal particles with titania, Langmuir, 2010, 26, 9297–9303 CrossRef PubMed.
- O. Monfort, T. Rochb, L. Satrapinskyy, M. Gregor, T. Plecenik, A. Plecenik and G. Plesch, Reduction of V2O5 thin films deposited by aqueous sol–gel method toVO2(B) and investigation of its photocatalytic activity, Appl. Surf. Sci., 2014, 322, 21–27 CrossRef CAS PubMed.
- Z. T. Zhang, Y. F. Gao, Z. Chen, J. Du, C. X. Cao, L. T. Kang and H. J. Luo, Thermochromic VO2 thin films: Solution-based processing, improved optical properties, and lowered phase transformation temperature, Langmuir, 2010, 26, 10738–10744 CrossRef CAS PubMed.
- Z. Chen, Y. Gao, L. Kang, C. Cao, S. Chen and H. Luo, Fine crystalline VO2 nanoparticles: Synthesis, abnormal phase transition temperatures and excellent optical properties of a derived VO2, J. Mater. Chem. A, 2014, 2, 2718–2727 CAS.
- G. Xu, P. Jin, M. Tazawa and K. Yoshimura, Optimization of antireflection coating for VO2-based energy efficient window, Sol. Energy Mater. Sol. Cells, 2004, 83, 29–37 CrossRef CAS PubMed.
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.