水热法制备(NH4)xWO3-y和WO3•1/3H2O: 反应温度对 产物物相、微结构和光学性能的影响
王琨1, 康利涛1, 陈石2, 董丽1, 梁伟1, 高峰1,3
1. 太原理工大学 材料科学与工程学院, 无机纳米能源实验室, 太原030024
2. 中国科学院 上海硅酸盐研究所, 上海200050
3. 中国科学院 煤炭化学研究所, 煤转化国家重点实验室, 太原030024
摘要

本工作研究了水热法制备钨系氧化物过程中反应温度对产物物相、微结构和光学性能的影响。XRD结果表明100 ℃条件下制备的样品为纳米六方相铵钨青铜 (NH4)xWO3-y。随着水热温度升高(140和180 ℃), 部分六方相(NH4)xWO3-y逐步转变为正交相三氧化钨WO3·1/3H2O, 形成两相共存产物。SEM和TEM分析证实(NH4)xWO3-y和WO3·1/3H2O粒子均为沿[001]方向(c轴)生长的短棒状单晶, 随水热反应进行它们会通过取向融并(oriented attachment)方式进一步长大。结构分析同时证实, 棒状晶的物相与其尺寸存在明显联系, 纳米尺寸的棒状晶为六方(NH4)xWO3-y相, 而微米尺寸的棒状晶为正交WO3·1/3H2O相。依据实验结果, 提出了一种表面能控制的六方相(NH4)xWO3-y向正交相WO3·1/3H2O相转变机理。光学性能测量表明, 100℃制备的六方相(NH4)xWO3-y样品可以很好的遮蔽近红外光(遮蔽率: 74.5%), 同时保持高可见光透过率(67.6 %)。

关键词: (NH4)xWO3-y; WO3·1/3H2O; 反应温度; 近红外光遮蔽性能
中图分类号:TB321   文献标志码:A    文章编号:1000-324X(2014)05-0550-07
Effects of Reaction Temperature on Structure and Optical Properties of Hydrothermally Prepared (NH4)xWO3-y and WO3·1/3H2O
WANG Kun1, KANG Li-Tao1, CHEN Shi2, DONG Li1, LIANG Wei1, GAO Feng1,3
1. Nano-Energy Inorganic Materials Laboratory, College of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, China
2. Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, China
3. State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030024, China
Corresponding author: KANG Li-Tao, lecturer. E-mail:kangltxy@gmail.com; GAO Feng, professor. E-mail:sxgaof@163.com
Abstract

Effects of reaction temperature on phase, microstructure and optical property of hydrothermally prepared (NH4)xWO3-y and WO3·1/3H2O were studied in this work. XRD results show that the sample prepared at 100℃ is hexagonal (NH4)xWO3-y (h-(NH4)xWO3-y). At higher reaction temperatures of 140℃ and 180℃, the hexagonal (NH4)xWO3-y gradually and partially transforms into orthorhombic WO3·1/3H2O (o-WO3·1/3H2O). SEM and TEM images illustrate that both (NH4)xWO3-y and WO3·1/3H2O crystallites show a short-rod shape with a [001] growth direction (c axis). The phases of crystallites seem to be strongly related to crystallite sizes; nano- and micro-sized crystallites adopt h-(NH4)xWO3-y and o-WO3·1/3H2O, respectively. Based on XRD, SEM and TEM results, a mechanism is extracted to explain the transformation from h-(NH4)xWO3-y to o-WO3·1/3H2O. Optical measurements indicate that the h-(NH4)xWO3-y nano-rods prepared at 100℃ exhibit a low infrared light transmittance of 25.5%, and maintaining a high luminous transmittance (67.6%), mainly due to the fice crystallite sizes.

Keyword: (NH4)xWO3-y; WO3·1/3H2O; reaction temperature; NIR shielding ability
0 Intruduction

As a wide bandgap n-type semiconductor, tungsten oxide (WO3) shows wide applications in gas sensors[ 1], electro-/gaso-chromic windows[ 2], photocatalysts[ 3], organic solar cells[ 4]and high temperature superconductors[ 5]. Significantly, with oxygen deficiency[ 6]or ternary elements intercalation[ 5, 7], free electrons can be injected into the LUMO band of WO3, resulting in conductive reduced tungsten oxides (WO3- x, WV2 xWVI1-2 xO3- x) or tungsten bron--zes (M xWO3, M = Tl, Rb, Cs, K, Na, NH4, etc)[ 8]. Nano-sized WO3- x and M xWO3 exhibit excellent near infrared (NIR) shielding ability due to localized surface Plasmon resonance (LSPR) of free electrons, while maintaining a high luminous transmittance ( Tlum). Therefore, they could be used as solar control filter to improve energy efficiency and internal comfort of automotive and architecture in hot climate[ 8].

Direct deposition of WO3- x (or M xWO3) films onto glass (by thermal evaporation[ 9, 10] or spray pyrolysis[ 11]) always requires expensive equipment and complex operation steps. One of the low-cost alternatives utilizes colloidal dispersions of WO3- x (or M xWO3) to form flexible WO3- x-polymer (or M xWO3-polymer) composite films. More importantly, these composite films exhibit similar Tlum of ~65%, while reject much more NIR light compared with the directly deposited films (with infrared light transmittance, Tinfr, being 10%-20% vs 40%-80%, especially in 1500-2000 nm), mainly caused by finite crystallite sizes.

The key step of this approach is preparation and dispersion of WO3- x or M xWO3 nano-particles with feature sizes well below visible light wavelength. So far, solid state reaction[ 8, 12] and solvo-/hydro-thermal synthesis[ 13, 14, 15, 16] are main ways to prepare WO3- x or M xWO3 crystallites. Generally speaking, solid state reaction results in crystallites tightly aggregated into large particles, making their dispersion into transparent matrices very difficult. On the other hand, even though the growth behavior of crystallites could be well manipulated by experimental conditions variation, the strong condensation tendency of tungstic acid in aqueous medium always results in relatively large crystallite sizes (diameter: 100-400 nm[ 17, 18, 19], length: 1.5-10 μm[ 13, 17, 18, 19]) for hydrothermally prepared products.

As a result, expensive organic solvents (solvothermal) are always employed to depress the condensation rate and thus crystallite sizes of tungsten oxides.

In current work, a one-step hydrothermal route for the preparation of hexagonal (NH4) xWO3- y nanorods (~9 nm and 20-100 nm in diameter and length) was reported. The effects of hydrothermal reaction temperature on phase, morphology and optical performance have been systematically addressed. It is shown that reaction temperature plays a crucial role for the formation of nano-sized h-(NH4) xWO3- y crystallites. At 100℃, uniform h-(NH4) xWO3- y nano-rods were firstly obtained, and some of them began to orientedly attach into larger ones along the radial directions. With reaction temperature increasing, NH4+ in h-(NH4) xWO3- y was gradually replaced by H2O; consequently, the large, orientedly-attached h-(NH4) xWO3- y rods gradually transformed into o-WO3·1/3H2O. Optical characterizations demonstrated that the NIR shielding ability of products (100 - Tinfr) was seriously deteriorated by the reaction temperature elevation.

1 Experimental
1.1 Material preparation

In a typical synthesis, 1.2 g of (NH4)10W12O41·5H2O was dissolved in 40 mL deionized water containing 1.9 g of citric acid as an organic stabilizer for tungstates. Then 0.124 g N2H4·2HCl was added to manipulate oxygen deficiency of the products. The acid of solution was adjust to pH=0.2 by dropwise addition of concentrated H2SO4 under stirring. Finally, the solution was transferred into a Teflon- lined 50 mL autoclave after 72 h aging for hydrothermal reaction at 100℃ for 24 h. After being collected by centrifugation, the products were washed thoroughly with deionized water and ethanol, followed by drying at 60℃ for 1 h for further experiments and characterization. To investigate the influence of hydrothermal temperature, we varied the hydrothermal temperature from 100℃ to 180℃ with an interval of 40℃, while keeping other synthesis parameters unchanged.

1.2 Material characterization

The phase compositions of samples were determined by X-ray diffraction (XRD) on an XPERT-PRO X-ray diffractometer using Cu Kα radiation (λ =0.1542 nm). The microscopic morphology of the products were observed by a Mira3 field emission scanning electron microscope (FE-S-EM). The morphology and phase information of samples were analyzed by a JEM-2100F transmission electron microscope (TEM). Thermal gravimetric analysis was recorded on a NETZSCH TG 209 F3 instrument from 100℃ to 900℃ at a heating rate of 10 ℃/min in air after being pre-h-ea-ted at 100℃ for 30 min. Infrared spectra were rec-o--rd-ed by means of a DigilabFTS-3000 spectrometer working in a 4000-400 cm-1 spectral range under a classical transmission configuration. The KBr disk technique was adopted with KBr and samples being 200 and ~0.23 mg, respectively. The resolution was 4 cm-1, and 64 scans were accumulated.

1.3 Optical Measurements

The as-prepared powders were dispersed ultrasonically in de-ionized water for 10 min to form a dispersion with a solid content of 150 mg/mL. Then an alcohol solution of 10wt% PVP (K30, average molecular weight = 58,000, serves as film-forming agent) was introduced to the dispersion (volume ratio 1:1). To fabricate the (NH4) xWO3- y (or WO3- x) - PVP composite films, the mixture dispersion was uniformly cast on fused silica glass substrates with a thickness of 40 μm and dried at 120℃ for 10 min. Optical properties of the obtained films were monitored on a Hitachi U-4100 UV-visible-NIR spectrophotometer in the wavelength range of 350-2600 nm. Integral luminous ( Tlum), solar ( Tsol) and infrared ( Tinfr) transmittance were obtained from the following equation:

(1)

where  T( λ) denotes transmittance at wavelength  λ, subscript i denotes lum, sol or infr,  φlum is the standard luminous efficiency function for the photopic vision,  φsol and φinfr are respectively the solar irradiance spectrum for air mass 1.5 (corresponding to the sun standing 37° above the horizon) in wavelength ranges of 350~2600 nm and 780~ 2600 nm[ 20].

2 Results and discussion
2.1 Results

XRD analysis was employed to detect phase evolution of the products during the variation of hydrothermal temperatures (Fig. 1). It is shown that sample-100 (obtained at 100℃) can be well indexed as hexagonal (NH4) xWO3- y (h-(NH4) xWO3- y) with the lattice parameters of a = b = 0.7392 nm, c = 0.7512 nm, α = β = 90° and γ = 120° (JCP-DS Card 42-0452). This sample was observed to be blue, indicating the appearance of W5+. When the reaction temperature increased up to 120℃, a tiny XRD peak centered at 18.1° appeared (not shown). Further increasing reaction temperature up to 140℃ and 180℃, this XRD peak is strengthened in intensity. Simultaneously, the XRD peak centered at ~24° gradually divided into three peaks, characterizing the appearance of orthorhombic WO3·1/3H2O phase (o-WO3·1/3H2O, JCPDS Card 54-1012, a= 1.2547 nm, b= 0.7737 nm, c= 0.7345 nm, α = β = γ = 90°) in products. This XRD results indicated that reaction temperature played a crucial role in controlling the phase composition of the final products. In our system, h-(NH4) xWO3- y was firstly synthesized at 100℃, which partially transformed into o-WO3·1/3H2O phase at high reaction temperatures (≥ 120℃), resulting in samples with mixed phases. Additionally, it is worthy to point out that the XRD peaks of h-(NH4) xWO3- y are obviously broaden than those of o-WO3·1/3H2O, implying that the former crystallite should possess a much finer size.

Figure 2 shows typical SEM images for sample-100, -140, and -180. As shown in Fig. 2(a), the sample obtained at 100℃ (sample-100) consisted of micro-scopic particles, which were further built by uniform nano-rods with lengths of ~100 nm. No other morphology has been detected in this product. For samples obtained at 140℃ (sample-140), in a stark contrast, a few micro-rods with lengths of ~2 μm appeared besides the uniform nano-rods (Fig. 2(c)-(d)). High-magnification SEM images illustrated that the micro-rods possessed orthorhombic end faces (pointed by a red arrow in Fig. 2(d)). Moreover, these orthorhombic micro-rods became larger in both volume-percentage and sizes with a higher reaction temperature of 180℃ (sample-180, Fig. 2(e)-(f)). This morphology changes of the orthorhombic micro-rods can reasonably explain the XRD signal evolution of o-WO3·1/3H2O phase in Fig. 1. Therefore, we assigned the orthorhombic micro-rods and tiny nano-rods in our samples to o-WO3-- x·1/3H2O and h-(NH4) xWO3- y, respectively.

The morphological and phase features of all samples had been further confirmed by TEM observation (Fig. 3). Low-magnification TEM images clearly showed that sample-100 comprised of short nano-rods with a uniform diameter of ~9 nm and lengths of 20~100 nm. The continuous lattice fringes in TEM images suggested that each nano-rod is a single h-(NH4) xWO3- y crystallite with a [001] growth direction. Analogous to SEM results, orthorhombic micro-rods were detected in sample-140 and sample-180, besides the tiny nano-rods. High- magnifi-cation TEM images and selected area electron diffraction (SAED) patterns indicated that nano-rods in sample-140 and sample-180 were still [001]- direction grown h-(NH4) xWO3- y crystallites; On the other hand, the orthorhombic micro-rods were found to be o-WO3·1/3H2O crystallites grown along the [001] direction (Fig. 3(f)). This TEM results were in good line with the XRD and SEM data.

Fig. 1 XRD patterns of sample-100, -140 and -180

Fig. 2 SEM images of sample-100 (a, b), -140 (b, d) and -180 (e, f)

Fig. 3 TEM images of sample-100 (a, b), -140 (c, d) and -180 (e, f)

Figure 4 shows the thermo-gravimetry (TG) curves of sample-100, -140 and -180 recorded in air at a heating rate of 10 ℃/min. As shown in Fig. 4, all samples manifested three mass evolution processes located at 100-273, 273-480 and 480-900 ℃, respectively. The end product after TG measurement is confirmed to be monoclinic WO3 by XRD technique. These mass evolution could be ascribed to the desorption of adsorbed H2O, the release of NH4+/crystal H2O, and the elimination of oxygen deficiency[ 21]. It is shown that the volumes of absorbed H2O (0.94wt%-1.04wt%) are comparable in three samples. Whereas, the volume of NH4+/crystal H2O (from 1.31wt% to 2.07wt%), as well as oxygen deficiency (0.33wt% to 1.05wt%), increased obviously in samples with reaction temperature elevation. In our samples, the oxygen deficiency was introduced by the addition of reducing agent N2H4·2HCl in the precursor solution. Oxygen deficiency has been proved to be beneficial for the improvement of NIR shielding ability by different researchers[ 8, 22].

Figure 5 shows the FTIR spectra of sample-100, -140 and -180. These spectra were very similar in profiles, implying the appearance of analogous functional groups in these samples. Generally, the absorption bands centered at 1401 and 3200 cm-1 were indicative of bending ( δNH) and stretching ( υNH) vibration of NH4+ ions[ 21]. With reaction temperature elevated, absorption peaks of NH4+ ions became weak, most possibly due to the reduction of NH4+ ions in products. Combining with the TG, XRD and FTIR data, we believed that the h-(NH4) xWO3- y products became oxygen deficient with reaction temperature increase. Consequently, some of the intercalated NH4+ ions in h-(NH4) xWO3- y were gradually replaced by intercalated H2O in a charge balance way (pay attention that both oxygen vacancies and NH4+ were positively charged). Finally, the h-(NH4) xWO3- y transformed into oxygen deficient o-WO3·1/3H2O.

Fig. 4 TG plots of sample-100, -140 and -180 recorded in an air flow at a heating rate of 10 ℃/min

Fig. 5 FTIR spectra of sample-100, -140 and -180 ( υ represents stretching vibration and δ represents bending vibration)

2.2 Mechanism

On the basis of XRD, SEM, TEM, TG and FTIR results, a mechanism has been extracted to explain the transformation from h-(NH4) xWO3- y to o-WO3·1/3H2O, as shown in Fig. 6 as well as Equation (2):

(2)

Where h and o represented hexagonal and orthorhombic structure type phases. By heating up to 100℃, the tungsten precursor in hydrothermal solution decomposed instantaneously to form plenty of monomers, and thus leading to a highly supersaturated initial reaction system. Therefore, a nucleation process was followed and large number of nucleus appeared. The growth of nucleus consumed the monomers quickly, restraining the further growing of the particles. As a result, numerous tiny h-(NH4) xWO3- y nano-rods were formed (Fig. 6(a)). Consequently, some of the nano-rods began to grow by oriented attachment along the radial directions of nano-rods, under the driving forces of surface energy reduction (Fig. 6(b)). At this stage, no characteristic peak of o-WO3·1/3H2O phase was observed in the XRD pattern. When the reaction temperature increased up to 140℃ and 180℃, large, orientedly-attached h-(NH4) xWO3- y rods transformed into o-WO3·1/3H2O ones (Fig. 6(c)). The o-WO3·1/3H2O rods were observed to further attached into bigger ones during following reactions (Fig. 2(c)-(f), 3(c)-(e) and 6(d)-(e)), finally resulting in samples consisted of tiny h-(NH4) xWO3- y nano-rods and extremely large o-WO3·1/3H2O micro-rods (Fig. 3(e) and 6(e)).

In both crystal structures of h-(NH4) xWO3- y and o-WO3·1/3H2O, distorted [WO6] octahedra are connected in the a-, b-, and c-direction, thereby forming a three- dim-en-sional structure, in which, NH4+ and H2O randomly occupies the hexagonal channels of h-(NH4) xWO3-y and dodecahedron of o-WO3·1/3H2O[ 23]. According to these theoretical models, one can find that the hexagonal framework of [WO6] octahedra could be transformed into orthorhombic one by simply shifting the neighbored a- b planes by 1/2 a.

In this study, the direct TEM images, which were capable to exhibit the phase transition process from h-(NH4) xWO3- y to o-WO3·1/3H2O, were failed to obtain. Nevertheless, noticing the dramatically difference of sizes (and thus specific surface areas) between o-WO3·1/3H2O and h-(NH4) xWO3- y crystallites, we speculated that finite size effects may play a crucial role during the phase transformation. Generally, occurrence of phase transfor-ma-tions was triggered by Gibbs energy decrease. However, the surface energy variation between two phases should also be taken in account for systems with high specific surface area[ 24]. For our samples, the specific surface energy of h-(NH4) xWO3- y is possibly lower than that of o-WO3·1/3H2O. Therefore, two energy evolution, surface energy increase (Δ Es) and Gibbs energy decrease (Δ Eg), co-determined the occurrence of the phase transformation. With the decrease of crystallite size, Δ Es becomes the dominating fact and thus makes the transformation from h-(NH4) xWO3- y to o-WO3·1/3H2O energetically unfavorable, resulting in products comprised of small h-(NH4) xWO3- y and large o-WO3·1/3H2O crystallites. Of course, more detailed mechanism should be further investigated.

Fig. 6 Scheme illustration of the transformation mechanism from h-(NH4) xWO3- y to o-WO3·1/3H2O (f) and corresponding TEM proofs (a-e). Panels (b) and (d) are the enlarged images of (a) and (c), respectively, and they clearly reveal the undergoing oriented attachment process between primary rods. This process has also been schematically illustrated in the framed regions of red circles in (a) and (c)

Figure 7 shows the transmittance spectra of a bare fused silica glass substrate, sample-100, -140 and -180. The Tlum, Tsol, and Tinfr (summarized in Table 1) were calculated with transmittance spectra and equation (1). As expected, the fused silica glass substrate was highly transparent for both visible and near infrared light (transmittance ~91%). For all our samples, the Tlum were still noteworthy (56.1%- 70.3%). The relatively poor Tlum of sample-180 could be ascribed to the light scattering effects by its large crystallites. The transmittance of all samples decreased rapidly in wavelength ranges of 350-450 nm, due to the intrinsic absorption of tungsten oxides[ 9, 22, 25].

Significantly, sample-100, -140 and -180 exhibited dra-matic different Tinfr of 25.5%, 62.7% and 66.8%, respec-tively. It is widely reported that the conductive h-(NH4) xWO3- y nano-particles could introduce a selective absorption of NIR light by localized surface Plasmon resonance[ 8, 10]. The loss of NIR shielding ability with reaction temperature elevation could be attributed to the increase of the large o-WO3∙1/3H2O phase, which offer weak or no LSPR.

Fig. 7 Transmittance spectra of a bare fused silica glass substrate, sample-100, -140 and -180

Table 1 Optical parameters of a bare fused silica glass substrate, sample -100, -140 and -180
3 Conclusion

In summary, we preformed a single-factor experiment to explore the effects of reaction temperature on the structure and optical properties of hydrothermally prepared tungsten oxides. Experimental results confirmed the vital role of reaction temperature on determining the phase and morphology of products. Uniform and well dispersed h-(NH4) xWO3- y nano-rods could be facilely obtained at 100℃. Elevated reaction temperatures ( e.g., 140℃ and 180℃) favored the orientedly-attachment of h-(NH4) xWO3- y rods and the phase transformation of h-(NH4) xWO3- y into o-WO3·1/3H2O. Interestingly, the phase of crystallites seemed strongly related with their sizes; small and large crystallites in products adopted h-(NH4) xWO3- y and o-WO3·1/3H2O phases, respectively. As a result, a mechanism was concluded to explain this transformation. Optical tests indicated that the h-(NH4) xWO3- y nano-rods obtained at 100℃ exhibited the best performance with a Tlum of 67.9% and a NIR shielding ability of 74.5%.

参考文献
[1] ZOU X X, LI G D, WANG P P, et al. A precursor route to single- crystalline WO3 nanoplates with an uneven surface and enhanced sensing properties. Dalton Transactions, 2012, 41(32): 9773-9780. [本文引用:1] [JCR: 3.806]
[2] LEE S H, DESHPANDE R, PARILLA P A, et al. Crystalline WO3 nanoparticles for highly improved electrochromic applications. Advanced Materials, 2006, 18(6): 763-766. [本文引用:1] [JCR: 14.829]
[3] SZILÁGYI I M, FÓRIZS B, ROSSELER O, et al. WO3 photocatalysts: Influence of structure and composition. Journal of Catalysis, 2012, 294: 119-127. [本文引用:1] [JCR: 5.787]
[4] CHOI H, KIM B S, KO M J, et al. Solution processed WO3 layer for the replacement of PEDOT: PSS layer in organic photovoltaic cells. Organic Electronics, 2012, 13(6): 959-968. [本文引用:1] [JCR: 3.836]
[5] HIRAI D, CLIMENT-PASCUAL E, CAVA R J. Superconductivity in WO2. 6F0. 4 synthesized by reaction of WO3 with teflon. Physical Review B, 2011, 84(17): 174519-1-3 [本文引用:2]
[6] BERAK J M, SIENKO M J. Effect of oxygen-deficiency on electrical transport properties of tungsten trioxide crystals. Journal of Solid State Chemistry, 1970, 2(1): 109-133. [本文引用:1] [JCR: 2.04]
[7] SHANKS H R, SIDLES P H, DANIELSON G C. Electrical properties of the tungsten bronzes. Adv. Chem. Ser. , 1963, 39: 237-245. [本文引用:1]
[8] TAKEDA H, ADACHI K. Near infrared absorption of tungsten oxide nanoparticle dispersions. Journal of the American Ceramic Society, 2007, 90(12): 4059-4061. [本文引用:5] [JCR: 2.107]
[9] GREEN M, TRAVLOS A. Sodium-tungsten bronze thin films: I. Optical properties of dilute bronzes. Philosophical Magazine B, 1985, 51(5): 501-520. [本文引用:2] [JCR: 1.343]
[10] GREEN M, HUSSAIN Z. Optical properties of dilute hydrogen tungsten bronze thin films. Journal of Applied Physics, 1991, 69(11): 7788-7796. [本文引用:2] [JCR: 2.21]
[11] ORTEGA J M, MARTÍNEZ A I, ACOSTA D R, et al. Structural and electrochemical studies of WO3 films deposited by pulsed spray pyrolysis. Solar Energy Materials and Solar Cells, 2006, 90(15): 2471-2479. [本文引用:1] [JCR: 4.63]
[12] GUO J, DONG C, YANG L H, et al. A green route for microwave synthesis of sodium tungsten bronzes NaxWO3(0. Journal of Solid State Chemistry, 2005, 178(1): 58-63. [本文引用:1] [JCR: 2.04]
[13] GU Z J, ZHAI T Y, GAO B F, et al. Controllable assembly of WO3 nanorods/nanowires into hierarchical nanostructures. The Journal of Physical Chemistry B, 2006, 110(47): 23829-23836. [本文引用:2]
[14] SONG R R, SONG H H, ZHOU J S, et al. Hierarchical porous carbon nanosheets and their favorable high-rate performance in lithium ion batteries. Journal of Materials Chemistry, 2012, 22(24): 12369-12374. [本文引用:1] [JCR: 5.968]
[15] LE HOUX N, POURROY G, CAMEREL F, et al. WO3 nanoparticles in the 5- 30 nm range by solvothermal synthesis under microwave or resistive heating. The Journal of Physical Chemistry C, 2009, 114(1): 155-161. [本文引用:1]
[16] ZHANG J, TU J P, DU G H, et al. Ultra-thin WO3 nanorod embedded polyaniline composite thin film: synthesis and electrochromic characteristics. Solar Energy Materials and Solar Cells, 2013, 114: 31-37. [本文引用:1] [JCR: 4.63]
[17] BAI S J, ZHANG K W, LUO R X, et al. Low-temperature hydrothermal synthesis of WO3 nanorods and their sensing properties for NO2. Journal of Materials Chemistry, 2012, 22(25): 12643-12650. [本文引用:2] [JCR: 5.968]
[18] WANG J M, KHOO E, LEE P S, et al. Synthesis, assembly, and electrochromic properties of uniform crystalline WO3 nanorods. The Journal of Physical Chemistry C, 2008, 112(37): 14306-14312. [本文引用:2]
[19] GU Z J, MA Y, YANG W S, et al. Self-assembly of highly oriented one-dimensional h-WO3 nanostructures. Chemical Communications, 2005(28): 3597-3599. [本文引用:2] [JCR: 6.378]
[20] KANG L T, GAO Y F, LUO H J, et al. Nanoporous thermochromic VO2 films with low optical constants, enhanced luminous transmittance and thermochromic properties. ACS Applied Materials & Interfaces, 2011, 3(2): 135-138. [本文引用:1] [JCR: 5.008]
[21] FOUAD N E, NOHMAN A K H, MOHAMED M A, et al. Characterization of ammonium tungsten bronze [(NH4)0. 33WO3] in the thermal decomposition course of ammonium paratungstate. Journal of Analytical and Applied Pyrolysis, 2000, 56(1): 23-32. [本文引用:2] [JCR: 2.56]
[22] XI G C, YE J H, MA Q, et al. In situ growth of metal particles on 3D urchin-like WO3 nanostructures. Journal of the American Chemical Society, 2012, 134(15): 6508-6511. [本文引用:2] [JCR: 10.677]
[23] CHEVRIER G, GERAND B, SEGUIN L, et al. Comparative structure refinement of MO3•1/3H2O (M= Mo, W) from X-Ray and neutron powder diffraction data and dehydration process. Materials Science Forum, 1996, 228: 695-700. [本文引用:1] [JCR: 0.399]
[24] VALMALETTE J C, ISA M. Size effects on the stabilization of ultrafine zirconia nanoparticles. Chemistry of Materials, 2002, 14(12): 5098-5102. [本文引用:1] [JCR: 8.238]
[25] GUO C, YIN S, ZHANG P L, et al. Novel synthesis of homogenous CsxWO3 nanorods with excellent NIR shielding properties by a water controlled-release solvothermal process. Journal of Materials Chemistry, 2010, 20(38): 8227-8229. [本文引用:1] [JCR: 5.968]