Next Article in Journal
Crystallographic and TEM Features of a TBC/Ti2AlC MAX Phase Interface after 1300 °C Burner Rig Oxidation
Previous Article in Journal
High-Performance Nanoplasmonic Enhanced Indium Oxide—UV Photodetectors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Step Solvothermal Process for Nanoarchitectonics of Metastable Hexagonal WO3 Nanoplates

1
Department of Materials Science and Engineering, The Ohio State University, Columbus, OH 43210, USA
2
National Centre for Nano Fabrication and Characterization, Denmark Technical University, DK-2800 Kongens Lyngby, Denmark
3
Department of Mechanical and Aerospace Engineering, The Ohio State University, Columbus, OH 43202, USA
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(4), 690; https://doi.org/10.3390/cryst13040690
Submission received: 26 March 2023 / Revised: 5 April 2023 / Accepted: 14 April 2023 / Published: 17 April 2023

Abstract

:
Hexagonal tungsten trioxide (h-WO3) has shown great potential for application in electrochromic devices, gas sensors, battery electrodes, and as photo-catalysts. The h-WO3 structure features a unique large network of open hexagonal channels that allow for intercalation. The hydrothermal synthesis of h-WO3 using sodium tungstate dihydrate as a precursor is widely explored, however, the residual alkaline ions are difficult to eliminate during the synthesis. The solvothermal synthesis using tungsten hexachloride as starting materials largely avoids the use of alkaline ions, but the effect of various synthesis parameters is not well-understood yet. To resolve these ambiguities, this study provides a reliable route to obtain h-WO3 via solvothermal synthesis and dehydration annealing. The effects of precursor concentration, water content, synthesis temperature, and synthesis time, as well as dehydration temperature, on the as-synthesized crystal structure and crystal morphology are studied.

1. Introduction

Tungsten trioxide (WO3) is a polymorph material, and the various crystalline polymorphs of WO3 are either stable or metastable. According to previous studies [1,2,3], at least five traditional polymorphs for bulk WO3 have been reported: ε-WO3 (space group: Pc), δ-WO3 (P1), γ-WO3 (P21/n), β-WO3 (Pbcn), and α-WO3 (P4/ncc). Further, a metastable hexagonal polymorph (h-WO3, P6/mmm) that exhibits a large network of open hexagonal channels along the c-axis has been reported [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31]. It has been shown that this characteristic of h-WO3 opens a wide range of applications in electrochemical, as well as in electrochromic appliances [6,7], as gas sensors [7,8,9,10] and in photo-catalysis [11,12].
The h-WO3 was first synthesized using an acid precipitation method of sodium tungstate dihydrate [4]. Over the years, researchers have shown that h-WO3 can be obtained by a variety of methods: by hydrothermal treatment of alkaline tungstate hydrates [13,14,15,16,17,18,19], by hydrothermal treatment of peroxopolytungstic acid [20], by oxidation of ammonium tungsten bronze [4,21,22], by sol–gel methods [23,24,25,26], by dehydration of WO3∙0.33(H2O) [27,28,29], by spray pyrolysis [32], or by solvothermal synthesis using tungsten hexachloride (WCl6) as precursor [30,31]. Among these methods, hydrothermal synthesis with alkaline tungstate hydrates as precursor is most commonly used because of its tunability to control particle size and morphology ranging from nanoparticle, nanowire, and nanosheet, to nanotube. The crystal structure of the product powder during hydrothermal synthesis is significantly influenced by the pH value during synthesis [13,14,15,16,17,18,19]. Several previous studies have shown that h-WO3 can be synthesized at pH values between 1.5 to 2.3 [13,14,15,16,18,19]. Ahmadian et al. [17] has shown that pH 2 and pH 4 can stabilize h-WO3, while at pH 3, WO3∙0.5(H2O) is dominant. In addition to pH, the presence of alkaline ions or ammonium ions is believed to play a role in stabilizing h-WO3 [4,33]. As for the particle morphology of h-WO3, several articles have shown that this is largely controlled by the templates, capping agent, or structure-directing agents (SDA), such as sodium sulfate, potassium sulfate, oxalic acid, ferrous ammonium sulfate, sodium chloride, and cobalt chloride [11,18,34,35,36,37].
It is also feasible to obtain h-WO3 by first obtaining WO3∙0.33(H2O) from hydrothermal synthesis and then dehydration by subsequent annealing [4,22,27,28,29]. Marques et al. [18] suggested that WO3∙0.33(H2O) can be synthesized at pH 0.4–1.8, depending on the types of SDA. Seguin et al. [28] studied the dehydration process of WO3∙0.33(H2O) and found that WO3∙0.33(H2O) first transforms to a super-metastable WO3 phase at about 350 °C and subsequently transforms to h-WO3 at around 380 °C. However, when the temperature reaches ~500 °C, h-WO3 transforms into thermostable γ-WO3 [28]. The exact temperature for this transition depends on the composition of h-WO3 (e.g., sodium content, ammonium ions content, and so on) [22].
A major challenge in hydrothermal treatment is the difficulty in controlling the amount of alkaline ions or ammonium ions in h-WO3 [4]. These ions originate from precursors or SDA and are believed to play a role in stabilizing h-WO3 [4,20,33]. In other words, it seems impossible to obtain pure h-WO3 from hydrothermal synthesis. Solvothermal synthesis using WCl6 as precursor can obtain h-WO3 free from alkaline ions [8,30,31]. Choi et al. [30] has shown that h-WO3 can be directly synthesized by solvothermal synthesis with WCl6 as precursor and a water/ethanol mixture solution as solvent with a water content between 2 vol% and 10 vol%. However, Chacón et al. [31] point out that the solvothermal synthesis product is WO3∙0.33(H2O) rather than h-WO3 under the condition mentioned in Choi et al. (i.e., [WCl6] = 0.0125 M–0.014 M, [H2O] = 10 vol%) [30]. Also, the effects of precursor concentration, reaction temperature, and reaction time are still not well-understood yet.
To address these open questions, we here present a systematical study on solvothermal synthesis using WCl6 as precursor and a mixed water/ethanol solution as the solvent. The effects of parameters such as water concentration, precursor concentration, synthesis temperature, synthesis time, as well as annealing temperature, are studied systematically. The results show that the solvothermal synthesis product can be either W18O49, WO3∙0.33(H2O) or γ-WO3 depending on precursor concentration and water content. h-WO3 is not observed in the powder produced from solvothermal synthesis, but it is obtained by subsequently annealing the WO3∙0.33(H2O) at 400–450 °C.

2. Materials and Methods

2.1. Materials Synthesis

The materials synthesis procedure is shown in Figure 1. The starting solution was prepared by dissolving 4.9575 g WCl6 (Sigma-Aldrich, St. Louis, MO, USA) into 125 mL of anhydrous ethanol (Neta scientific Inc, Hainesport, NJ, USA) in an argon-filled glove box to prevent hydration of the WCl6 ([WCl6] = 0.1 M). The solution immediately turned yellow as HCl vapor escaped from the container. Eventually, the solution turned blue, ultimately reaching a deep green color after one week. Based on previous studies [38], the final (green) solution is a W2Cl4(OC2H5)6 solution. For convenience, we still use [WCl6] to denote the precursor concentration in this study, although WCl6 has been transformed to W2Cl4(OC2H5)6.
The final solution was prepared by mixing W2Cl4(OC2H5)6 solution (starting solution), anhydrous ethanol, and distilled water under magnetic stirring for 3 h. In order to prepared the final solution with different precursor concentration ([WCl6] = 0.01 M or 0.05 M) and different water content (0 vol%, 5 vol%, 10 vol%, 20 vol%, or 50 vol%), the volumes of the W2Cl4(OC2H5)6 solution (starting solution), the final solution, and of the distilled water are listed in Table 1. Then, the final solution was sealed, removed from the argon-filled glovebox and was transferred into a 100 mL Teflon-lined hydrothermal pressure vessel. The solvothermal synthesis was conducted at 200 °C for 12 h in an electric oven. After that, the as-synthesized material was centrifuged and washed with ethanol several times. The powder was then allowed to dry at room temperature in a fume hood.
To study the effect of synthesis temperature and time on the synthesis of WO3∙0.33(H2O), 50 mL of final solution containing 0.01 M of [WCl6] and a water content of 10 vol% was prepared. Then, the solvothermal synthesis was conducted at temperatures between 160 °C and 200 °C and synthesis times between 3 h and 12 h, as listed in Table 2.
To study the dehydration process of WO3∙0.33(H2O), the WO3∙0.33(H2O) powders (sample no.3) were characterized by thermal gravimetric analysis (TGA) in an argon gas environment. The TGA temperature ranged from room temperature to 750 °C, using a heating rate of 5 °C/min. The annealing experiments were conducted in an electric oven using sample no.3 as a starting material. The annealing temperatures were 300 °C, 350 °C, 400 °C, 450 °C, 500 °C, and 550 °C. Annealing time was set to 2 h.

2.2. Materials Characterization

The X-ray powder diffraction (XRD) data were collected using a Rigaku SmartLab diffractometer with a Cu Kα radiation source (λ = 1.5418 Å) at a scan rate of 1° min1 in the 2θ range from 10° to 40°. The morphologies of the samples were investigated using scanning electron microscopy (SEM, Thermo scientific Quattro ESEM) and transmission electron microscopy (TEM, FEI Tecnai G2 at 200 kV). To identify the structure of each powder, selected area electron diffraction (SAED) and high-resolution TEM (HRTEM) characterization was performed.

3. Results

3.1. Effects of Water Content and Precursor Concentration on Crystal Structure of As-Synthesized Powder

The results of the X-ray diffraction (XRD) experiments on samples 1–5 are shown in Figure 2. The XRD pattern of powder synthesized under pure ethanol condition (sample 1) has a strongest diffraction peak at 23.32° that can be assigned to W18O49 (JCPDS no. 36-0101). The other characteristic peaks of W18O49 are not observed due to the strong broadening effect, which is commonly observed in the XRD spectra of nanocrystalline materials. The XRD data in Figure 2 suggests that the addition of water into the solvent stabilizes WO3∙0.33(H2O) and γ-WO3. Specifically, when the water content is 5 vol% (sample 2) and 10 vol% (sample 3), respectively, the XRD patterns exhibit diffraction peaks at 13.96°, 18.1°, 22.98°, 24.26°, 27.04°, and 28.18°, respectively (JCPDS no. 87-1203). All these observed peaks correspond to diffraction peaks of WO3∙0.33(H2O). Some minor peaks of γ-WO3 phases are also observed (at 23.6°, 26.12°), but the intensity is extremely low, suggesting that WO3∙0.33(H2O) is the dominant phase in sample 2 and sample 3. Several other studies about solvothermal synthesis at similar water and precursor concentrations have suggested before that WO3∙0.33(H2O) is the product without any further heat treatment [31,39]. However, both our and these previous results are in conflict with the results of Choi et al. [30], in which pure hexagonal phase is directly synthesized under this condition ([WCl6] = 0.0125 M–0.014 M, [H2O] = 10 vol%).
One possible explanation for the discrepancy between our work and Choi et al. [30] is that the latter did not distinguish the XRD patterns of WO3∙0.33(H2O) and h-WO3. Specifically, only the XRD peaks at 18.1°, 22.78°, and 22.98° can be used to identify these two compounds based on XRD pattern analysis. Since the XRD pattern in Choi et al. [30] was acquired at a 2θ range from 20° to 45°, it does not include the characteristic peaks of WO3∙0.33(H2O) at 18.1°. Also, the 001 peak of “hexagonal WO3” in Choi et al. [30] is at 23° ± 0.1°, which should have been indexed as 002 peak of WO3∙0.33(H2O) rather than the 001 peak of h-WO3 (at 22.78°).
Further increase in water content to 20 vol% (sample 4) or 50 vol% (sample 5) results in the formation of γ-WO3, as suggested by the appearance of XRD peaks at 23.12°, 23.6°, 24.34°, 26.12°, 33.28°, and 34.14° in Figure 2 (JCPDS no. 32-1295). The powder synthesized using a water content of 20 vol% consists of WO3∙0.33(H2O) and a small amount of γ-WO3 as suggested by the strong WO3∙0.33(H2O) diffraction peaks and weak γ-WO3 peak at 23.1°. However, when water content is increased to 50 vol%, γ-WO3 becomes dominate, in agreement with results published by Chacón et al. [31].
When the precursor concentration is increased to 0.05 M, a totally different situation emerges. Figure 3 shows the XRD patterns of powders synthesized under precursor concentration of 0.05 M and water content (0 vol%, 5 vol%, 10 vol%, 20 vol%, or 50 vol%). When no water is added (sample 6), the as-synthesized powder consists of only W18O49, as suggested by the black spectra in Figure 3. In case of a water content of 5 vol% (sample 7), 10 vol% (sample 8), and 20 vol% (sample 9), as well as of 50 vol% (sample 10), the corresponding diffraction peaks can all be indexed as XRD peaks of γ-WO3 (see the spectra in Figure 3). This result suggests that the addition of water at this precursor concentration does not stabilize the WO3∙0.33(H2O). A possible explanation is that the γ-WO3 phase becomes much more stable under high precursor concentration. Based on these results, it can be concluded that the low precursor concentration as well as the appropriate water content are two important (key) parameters to stabilize WO3∙0.33(H2O) in solvothermal synthesis, as shown in Figure 4.
In addition to identifying the crystal structure by XRD, the particle morphology was explored using detailed SEM (Figure 5) and TEM (Figure 6) characterization. Figure 5a,f (SEM) and Figure 6a,d (TEM) indicate that the crystal morphology of W18O49 (i.e., samples 1 and 6) is an urchin-like structure assembled from a bunch of nanorods with an average size of ~10 nm. This then explains the presence of a strong broadening of XRD diffraction peaks observed from sample 1 (Figure 2) and sample 6 (Figure 3). The rate of crystal growth of W18O49 seems to be much slower compared to the rate for WO3∙0.33(H2O) and γ-WO3 crystals. Such a property has been observed before by others [40]. This characteristic might allow the control of the size of a W18O49 quantum dot structure as suggested in Ref. [40].
With respect to WO3∙0.33(H2O), Figure 5b,c and Figure 6b show that as-synthesized WO3∙0.33(H2O) powder consists of large faceted platelets with hexagonal shape and an average size of ~5 μm. Such a crystal morphology leads to the presence of dominant facets, i.e., preferential crystallographic orientation (PCO). According to the HRTEM images (Figure 6e) and SAED pattern (insets in Figure 6b), the orientation of the particle along the direction perpendicular to the hexagonal plate is determined to be 002. This can explain the 002 XRD peak of the WO3∙0.33(H2O) (see red or blue curves of Figure 2) having the highest intensity, while the 220 peak in the standard XRD pattern is the strongest peak. The appearance of PCO of WO3∙0.33(H2O) and of the hexagonal-faceted morphology is the result of the natural growth preference of the orthorhombic WO3∙0.33(H2O) structure along six symmetrical directions [41,42].
Figure 5d shows that two kinds of particle are present in sample 4. One type is the WO3∙0.33(H2O) hexagonal-shape particle. The other seems to be a porous sheet with a square shape. As the XRD results shown in Figure 2 suggest that sample 4 consists of both WO3∙0.33(H2O) and γ-WO3, we postulate that the squared sheet-like particle is a γ-WO3 particle.
SEM images and XRD results of sample 5 as well as of samples 7–10 confirm this hypothesis. Figure 5e,g–h show that these samples consist of porous squared-sheet particles with an average side length of ~6 µm. Based on XRD results from sample 5 and samples 7–10 indicating the presence of dominantly γ-WO3, porous square sheets can definitely be identified as γ-WO3 particles. Similar to the hexagonal-shaped particles, the squared-sheets particles also exhibit PCO. The PCO is determined to be 002 direction by HRTEM images (Figure 6f) and SAED patterns (insets in Figure 6c). This also explains that the 002 XRD peak of as-synthesized γ-WO3 is the dominant peak (see Figure 2) compared to any other diffraction peaks. According to the Gibbs–Wulff theorem about crystal morphology, the PCO could be attributed to the fact that the 001 facet of γ-WO3 in ethanol has a lower surface energy [43].

3.2. Effect of Synthesis Temperature and Time on the Synthesis of WO3∙0.33(H2O)

The above-presented results suggest that a low precursor concentration and appropriate content of water is critical to obtain WO3∙0.33(H2O) through solvothermal synthesis. Two other important parameters are the synthesis temperature as well as the synthesis time. In a systematic approach to identify the required temperature, the synthesis temperature was varied between 160 °C (sample 11), 180 °C (sample 12), and 200 °C (sample 3) while maintaining all other parameters at a constant, i.e., the water content at 10 vol%, the precursor concentration ([WCl6]) at 0.01 M, and the synthesis time at 12 h. The XRD results of the as-synthesized powder under these three temperatures are shown in Figure 7, and the results show that the powder synthesized at 160 °C is W18O49 and powders synthesized at 180 °C and 200 °C are WO3∙0.33(H2O). Thus, a temperature higher than 180 °C is required to stabilize WO3∙0.33(H2O).
To study the effect of synthesis time, powder was synthesized at 200 °C but for different times. Figure 8 shows the XRD pattern of powder synthesized at 200 °C for 3 h (sample 13), 6 h (sample 14), 9 h (sample 15), and 12 h (sample 3). W18O49 and WO3∙0.33(H2O) are both present when a synthesis time of 3 h is selected, as suggested by the black XRD spectra in Figure 8. Increasing the synthesis time to 6 h results in the increase in the fraction of WO3∙0.33(H2O). When the synthesis time is increased to 9 h and 12 h, the WO3∙0.33(H2O) phase becomes the dominant phase. The presence of the W18O49 phase at short synthesis time (3 h and 6 h) can be attributed to an incomplete reaction of water, oxygen, and W2Cl4(OC2H5)6 due to the insufficient time to fully react. The unreacted precursor finally transforms to W18O49 as indicted in XRD pattern when the water content is 0 vol% (samples 1 and 6). An increase in synthesis time results in a decreased fraction of the unreacted precursor, thereby decreasing the fraction of W18O49 in the as-synthesized powder.

3.3. Dehydration of WO3∙0.33(H2O)

After systematically investigating the effects of synthesis parameters such as precursor concentration, water content, synthesis temperature, and synthesis time for the synthesis of WO3∙0.33(H2O), the next focus is on the transformation of WO3∙0.33(H2O) to hexagonal h-WO3. This post-synthesis transformation can be achieved by annealing, since WO3∙0.33(H2O) is believed to dehydrate upon heating and the h-WO3 phase forms at ~350 °C [27,28,29]. However, the h-WO3 phase is a metastable phase and is believed to further transform to the thermodynamically stable γ-WO3 at a temperature between 475–532 °C. The exact temperature of the h-WO3 to γ-WO3 phase transformation depends on the purity of the h-WO3 (e.g., sodium content, ammonium ions content, and so on) [22]. Thus, to determine the exact temperature range to stabilize h-WO3 for this solvothermal method, thermal gravimetric analysis (TGA) characterization was carried out, and the result is shown in Figure 9. As the TGA curves suggest, the dehydration process starts at 144.3 °C and stops at 505.2 °C. The total weight loss is 2.657%, which is exactly the weight difference between WO3∙0.33(H2O) and WO3. The first derivative peak temperature (Tp) of dehydration, indicating the point of greatest rate of change on the weight loss curve, is at around 358.6 °C. The highest temperature observed in our study is higher than the temperature reported in previous studies [27,28,29]. This might be caused by the high ramping rates in our TGA experiments. To determine the best annealing temperature for dehydration, annealing experiments at different temperatures were conducted. Figure 10 shows the XRD pattern of the as-annealed powder. The XRD pattern of the powder annealed at 300 °C and at 350 °C, can be indexed as WO3∙0.33(H2O) plus a minor amount of gamma-WO3 (the intensity of gamma diffraction peak is too low to be quantified, the fraction is definitely less than 1%). Comparing the spectra at 300 °C (black) and at 350 °C (red), the XRD diffraction peak at 22.98° at 350 °C (red curve) seems to have an asymmetric shape (i.e., slight peak tail at lower 2θ). This indicates that there must be a minor peak present with a θ only slightly less 22.98°. The very low intensity of this peak makes it difficult to determine the exact peak position.
However, this is of less importance here because WO3∙0.33(H2O) is still the dominant phase after annealing at 350 °C for 2 h, which means that this annealing temperature cannot fully transform WO3∙0.33(H2O) to h-WO3. After the annealing temperature has been increased to 400 °C and to 450 °C, the characteristic XRD peaks of WO3∙0.33(H2O) at 2θ = 18.10° and 22.98° disappear and a new peak at 22.78° appears. This new pattern can be indexed as h-WO3 (JCPDS no. 75-2187). For the sample annealed at 500 °C, both h-WO3 and γ-WO3 XRD diffraction peaks are observed, suggesting that the powder consists of both phases. However, γ-WO3 dominates when the sample has been annealed at 550 °C. In conclusion, to obtain h-WO3, the annealing temperature should be kept between 400–450 °C. The morphologies of this annealed powder are also investigated by SEM, and the results are shown in Figure 11. The hexagonal morphology is clearly preserved during the dehydration process, as suggested by the images in Figure 11a–d. This characteristic means that h-WO3 hexagonal sheet can be obtained via the here presented solvothermal synthesis method. The h-WO3 hexagonal sheet still has a 002 (0002 for Miller–Bravais indices for hexagonal structure)-preferred orientation, as suggested by the observed XRD pattern. Once the annealing temperature is increased to 500 °C, a change in particle morphology due to the h-WO3 to γ-WO3 transformation is observed as shown in Figure 11e–f. The hexagonal sheet seems to decompose into several rods, which presumably are γ-WO3 phase particles.
Summarizing all the results presented here, we are now able to outline a two-step procedure to obtain metastable h-WO3 using a solvothermal route followed by a post-synthesis high temperature annealing:
  • First, WO3∙0.33(H2O) should be synthesized by a solvothermal process. To ensure no other phase is produced, the WCl6 concentration should be as low as 0.01 M, the water content should be between 5–10 vol%, the synthesis temperature should be higher than 180 °C, and the synthesis time should be longer than 9 h;
  • Then, in the subsequent annealing process, the WO3∙0.33(H2O) is transformed to h-WO3. In order to stabilize the metastable hexagonal structure, the annealing temperature needs to be carefully controlled between 400 °C and 450 °C.

4. Conclusions

The metastable hexagonal tungsten trioxide phase (h-WO3) was successfully synthesized via a solvothermal synthesis followed by a controlled annealing step. The WO3∙0.33(H2O) is first synthesized in a solvothermal process and subsequent annealing triggers the dehydration of WO3∙0.33(H2O), thereby transforming WO3∙0.33(H2O) to h-WO3. Compared with previously published studies on solvothermal synthesis of h-WO3, our study has resolved the ambiguity of the product of solvothermal synthesis of h-WO3. The impact of several synthesis parameters are more evident now as they have been studied systematically.
  • Parameters of the solvothermal process: WCl6 precursor concentration as low as 0.01 M, water content 5–10 vol%, synthesis temperature higher than 180 °C, synthesis time longer than 9 h;
  • Parameters of the subsequent annealing process: annealing temperature between 400 °C and 450 °C.
In addition, the following important dependencies on synthesis parameters were found:
(i)
Higher WCl6 precursor concentration and higher water content in the solvent results in formation of thermostable γ-WO3 with porous squared-sheet morphology. The preferential crystallographic orientation of the γ-WO3 nanosheet is 002 orientation. Precursor concentration of 0.01 M and water content of 5–10 vol% produce pure 002-orientated WO3∙0.33(H2O) hexagonal nanosheets. The urchin-like W18O49 is only present when water is absent from the solvent;
(ii)
Synthesis temperature below 180 °C and synthesis time below 6 h might also result in the formation of W18O49 and, therefore, should be avoided. This is attributed to the insufficient reaction caused by low synthesis temperature and short synthesis time;
(iii)
Annealing at temperature higher than 400 °C is required to dehydrate the WO3∙0.33(H2O) and produce h-WO3. However, annealing time higher than 500 °C results in a h to γ transition. The dehydration process does not destroy the hexagonal morphology of the particle and, thus, the h-WO3 hexagonal sheet can be produced.
Further work is focused on studying the purity of the h-WO3 synthesized here, including the amount and level of structural defects, as well as on upscaling this two-step process to produce h-WO3 and exploring the wide range of applications, e.g., in gas sensors or photocatalysts.

Author Contributions

Conceptualization, Z.Q. and P.-I.G.; methodology, Z.Q.; software, Z.Q.; validation, Z.Q., J.R.J. and P.-I.G.; formal analysis, Z.Q., J.R.J. and P.-I.G.; investigation, Z.Q.; resources, J.R.J. and P.-I.G.; data curation, Z.Q.; writing—original draft preparation, Z.Q.; writing—review and editing, Z.Q., J.R.J. and P.-I.G.; visualization, Z.Q.; supervision, J.R.J. and P.-I.G.; project administration, J.R.J. and P.-I.G.; funding acquisition, P.-I.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research has received funding by the Orton Chair Funds and the National Science Foundation award CBET #2029847.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the Center for Electron Microscopy and Analysis (CEMAS) as the XRD and electron microscopy characterization were performed at the Center for Electron Microscopy and Analysis (CEMAS) at the Ohio State University.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could appear to influence the work reported in this paper.

References

  1. Christopher, J.H.; Vittorio, L.; Kevin, S.K. High-temperature phase transitions in tungsten trioxide—The last word? J. Phys. Condens. Matter 2002, 14, 377. [Google Scholar] [CrossRef]
  2. Han, W.; Shi, Q.; Hu, R. Advances in Electrochemical Energy Devices Constructed with Tungsten Oxide-Based Nanomaterials. Nanomaterials 2021, 11, 692. [Google Scholar] [CrossRef] [PubMed]
  3. Sood, S.; Gouma, P. Polymorphism in nanocrystalline binary metal oxides. Nanomater. Energy 2013, 2, 82–96. [Google Scholar] [CrossRef]
  4. Szilágyi, I.M.; Madarász, J.; Pokol, G.; Király, P.; Tárkányi, G.; Saukko, S.; Mizsei, J.; Tóth, A.L.; Szabó, A.; Varga-Josepovits, K. Stability and Controlled Composition of Hexagonal WO3. Chem. Mater. 2008, 20, 4116–4125. [Google Scholar] [CrossRef]
  5. Sun, W.; Yeung, M.T.; Lech, A.T.; Lin, C.W.; Lee, C.; Li, T.; Duan, X.; Zhou, J.; Kaner, R.B. High Surface Area Tunnels in Hexagonal WO3. Nano Lett. 2015, 15, 4834–4838. [Google Scholar] [CrossRef]
  6. Adhikari, S.; Sarkar, D. High Efficient Electrochromic WO3 Nanofibers. Electrochim. Acta 2014, 138, 115–123. [Google Scholar] [CrossRef]
  7. Balázsi, C.; Sedlácková, K.; Pfeifer, J.; Tóth, A.L.; Zayim, E.O.; Szilágyi, I.M.; Wang, L.; Kalyanasundaram, K.; Gouma, P.-I. Synthesis and Examination of Hexagonal Tungsten Oxide Nanocrystals for Electrochromic and Sensing Applications. In Sensors for Environment, Health and Security: Advanced Materials and Technologies; Springer: Dordrecht, The Netherlands, 2009; pp. 77–91. [Google Scholar]
  8. Abe, O.O.; Qiu, Z.; Jinschek, J.R.; Gouma, P.-I. Effect of (100) and (001) Hexagonal WO3 Faceting on Isoprene and Acetone Gas Selectivity. Sensors 2021, 21, 1690. [Google Scholar] [CrossRef] [PubMed]
  9. Gouma, P.I.; Wang, L.; Simon, S.R.; Stanacevic, M. Novel Isoprene Sensor for a Flu Virus Breath Monitor. Sensors 2017, 17, 199. [Google Scholar] [CrossRef]
  10. Zhang, D.; Fan, Y.; Li, G.; Ma, Z.; Wang, X.; Cheng, Z.; Xu, J. Highly sensitive BTEX sensors based on hexagonal WO3 nanosheets. Sens. Actuators B Chem. 2019, 293, 23–30. [Google Scholar] [CrossRef]
  11. Mohamed, M.M.; Salama, T.M.; Hegazy, M.A.; Abou Shahba, R.M.; Mohamed, S.H. Synthesis of hexagonal WO3 nanocrystals with various morphologies and their enhanced electrocatalytic activities toward hydrogen evolution. Int. J. Hydrogen Energy 2019, 44, 4724–4736. [Google Scholar] [CrossRef]
  12. Peng, T.; Ke, D.; Xiao, J.; Wang, L.; Hu, J.; Zan, L. Hexagonal phase WO3 nanorods: Hydrothermal preparation, formation mechanism and its photocatalytic O2 production under visible-light irradiation. J. Solid State Chem. 2012, 194, 250–256. [Google Scholar] [CrossRef]
  13. Pang, H.-F.; Xiang, X.; Li, Z.-J.; Fu, Y.-Q.; Zu, X.-T. Hydrothermal synthesis and optical properties of hexagonal tungsten oxide nanocrystals assisted by ammonium tartrate. Phys. Status Solidi A 2012, 209, 537–544. [Google Scholar] [CrossRef]
  14. Zakharova, G.S.; Podval’naya, N.y.V.; Gorbunova, T.y.I.; Pervova, M.G.; Murzakaev, A.M.; Enyashin, A.N. Morphology-controlling hydrothermal synthesis of h-WO3 for photocatalytic degradation of 1,2,4-trichlorobenzene. J. Alloys Compd. 2023, 938, 168620. [Google Scholar] [CrossRef]
  15. Bhojane, P.; Shirage, P.M. Facile preparation of hexagonal WO3 nanopillars and its reduced graphene oxide nanocomposites for high-performance supercapacitor. J. Energy Storage 2022, 55, 105649. [Google Scholar] [CrossRef]
  16. Nekita, S.; Nagashima, K.; Zhang, G.; Wang, Q.; Kanai, M.; Takahashi, T.; Hosomi, T.; Nakamura, K.; Okuyama, T.; Yanagida, T. Face-Selective Crystal Growth of Hydrothermal Tungsten Oxide Nanowires for Sensing Volatile Molecules. ACS Appl. Nano Mater. 2020, 3, 10252–10260. [Google Scholar] [CrossRef]
  17. Ahmadian, H.; Tehrani, F.S.; Aliannezhadi, M. Hydrothermal synthesis and characterization of WO3 nanostructures: Effects of capping agent and pH. Mater. Res. Express 2019, 6, 105024. [Google Scholar] [CrossRef]
  18. Marques, A.C.; Santos, L.; Costa, M.N.; Dantas, J.M.; Duarte, P.; Goncalves, A.; Martins, R.; Salgueiro, C.A.; Fortunato, E. Office paper platform for bioelectrochromic detection of electrochemically active bacteria using tungsten trioxide nanoprobes. Sci. Rep. 2015, 5, 9910. [Google Scholar] [CrossRef]
  19. Wenderich, K.; Noack, J.; Kärgel, A.; Trunschke, A.; Mul, G. Effect of Temperature and pH on Phase Transformations in Citric Acid Mediated Hydrothermal Growth of Tungsten Oxide. Eur. J. Inorg. Chem. 2018, 2018, 917–923. [Google Scholar] [CrossRef]
  20. Zheng, J.Y.; Haider, Z.; Van, T.K.; Pawar, A.U.; Kang, M.J.; Kim, C.W.; Kang, Y.S. Tuning of the crystal engineering and photoelectrochemical properties of crystalline tungsten oxide for optoelectronic device applications. CrystEngComm 2015, 17, 6070–6093. [Google Scholar] [CrossRef]
  21. Szilágyi, I.M.; Wang, L.; Gouma, P.-I.; Balázsi, C.; Madarász, J.; Pokol, G. Preparation of hexagonal WO3 from hexagonal ammonium tungsten bronze for sensing NH3. Mater. Res. Bull. 2009, 44, 505–508. [Google Scholar] [CrossRef]
  22. Szilágyi, I.M.; Pfeifer, J.; Balázsi, C.; Tóth, A.L.; Varga-Josepovits, K.; Madarász, J.; Pokol, G. Thermal stability of hexagonal tungsten trioxide in air. J. Therm. Anal. Calorim. 2008, 94, 499–505. [Google Scholar] [CrossRef]
  23. Abbaspoor, M.; Aliannezhadi, M.; Tehrani, F.S. Effect of solution pH on as-synthesized and calcined WO3 nanoparticles synthesized using sol-gel method. Opt. Mater. 2021, 121, 111552. [Google Scholar] [CrossRef]
  24. Gouma, P.I.; Kalyanasundaram, K. Novel synthesis of hexagonal WO3 nanostructures. J. Mater. Sci. 2015, 50, 3517–3522. [Google Scholar] [CrossRef]
  25. Kharade, R.R.; Patil, K.R.; Patil, P.S.; Bhosale, P.N. Novel microwave assisted sol–gel synthesis (MW-SGS) and electrochromic performance of petal like h-WO3 thin films. Mater. Res. Bull. 2012, 47, 1787–1793. [Google Scholar] [CrossRef]
  26. Cremonesi, A.; Bersani, D.; Lottici, P.P.; Djaoued, Y.; Ashrit, P.V. WO3 thin films by sol–gel for electrochromic applications. J. Non-Cryst. Solids 2004, 345–346, 500–504. [Google Scholar] [CrossRef]
  27. Perfecto, T.M.; Zito, C.A.; Volanti, D.P. Room-temperature volatile organic compounds sensing based on WO3·0.33H2O, hexagonal-WO3, and their reduced graphene oxide composites. RSC Adv. 2016, 6, 105171–105179. [Google Scholar] [CrossRef]
  28. Seguin, L.; Figlarz, M.; Pannetier, J. A novel supermetastable WO3 phase. Solid State Ion. 1993, 63–65, 437–441. [Google Scholar] [CrossRef]
  29. Figlarz, M. New oxides in the WO3-MoO3 system. Prog. Solid State Chem. 1989, 19, 1–46. [Google Scholar] [CrossRef]
  30. Choi, H.G.; Jung, Y.H.; Kim, D.K. Solvothermal Synthesis of Tungsten Oxide Nanorod/Nanowire/Nanosheet. J. Am. Ceram. Soc. 2005, 88, 1684–1686. [Google Scholar] [CrossRef]
  31. Chacón, C.; Rodríguez-Pérez, M.; Oskam, G.; Rodríguez-Gattorno, G. Synthesis and characterization of WO3 polymorphs: Monoclinic, orthorhombic and hexagonal structures. J. Mater. Sci. Mater. Electron. 2014, 26, 5526–5531. [Google Scholar] [CrossRef]
  32. Ortega, J.M.; Martínez, A.I.; Acosta, D.R.; Magaña, C.R. Structural and electrochemical studies of WO3 films deposited by pulsed spray pyrolysis. Sol. Energy Mater. Sol. Cells 2006, 90, 2471–2479. [Google Scholar] [CrossRef]
  33. Wu, P.M.; Ishii, S.; Tanabe, K.; Munakata, K.; Hammond, R.H.; Tokiwa, K.; Geballe, T.H.; Beasley, M.R. Synthesis and ionic liquid gating of hexagonal WO3 thin films. Appl. Phys. Lett. 2015, 106, 042602. [Google Scholar] [CrossRef]
  34. Huang, K.; Pan, Q.; Yang, F.; Ni, S.; Wei, X.; He, D. Controllable synthesis of hexagonal WO3 nanostructures and their application in lithium batteries. J. Phys. D Appl. Phys. 2008, 41, 155417. [Google Scholar] [CrossRef]
  35. Miao, B.; Zeng, W.; Hussain, S.; Mei, Q.; Xu, S.; Zhang, H.; Li, Y.; Li, T. Large scale hydrothermal synthesis of monodisperse hexagonal WO3 nanowire and the growth mechanism. Mater. Lett. 2015, 147, 12–15. [Google Scholar] [CrossRef]
  36. Ghosh, K.; Roy, A.; Tripathi, S.; Ghule, S.; Singh, A.K.; Ravishankar, N. Insights into nucleation, growth and phase selection of WO3: Morphology control and electrochromic properties. J. Mater. Chem. C 2017, 5, 7307–7316. [Google Scholar] [CrossRef]
  37. Lu, Y.; Zhang, J.; Wang, F.; Chen, X.; Feng, Z.; Li, C. K2SO4-Assisted Hexagonal/Monoclinic WO3 Phase Junction for Efficient Photocatalytic Degradation of RhB. ACS Appl. Energy Mater. 2018, 1, 2067–2077. [Google Scholar] [CrossRef]
  38. Li, C.-P.; Lin, F.; Richards, R.M.; Engtrakul, C.; Tenent, R.C.; Wolden, C.A. The influence of sol–gel processing on the electrochromic properties of mesoporous WO3 films produced by ultrasonic spray deposition. Sol. Energy Mater. Sol. Cells 2014, 121, 163–170. [Google Scholar] [CrossRef]
  39. Li, Y.; Tang, Z.; Zhang, J.; Zhang, Z. Exposed facet and crystal phase tuning of hierarchical tungsten oxide nanostructures and their enhanced visible-light-driven photocatalytic performance. CrystEngComm 2015, 17, 9102–9110. [Google Scholar] [CrossRef]
  40. Wang, Y.; Wang, X.; Xu, Y.; Chen, T.; Liu, M.; Niu, F.; Wei, S.; Liu, J. Simultaneous Synthesis of WO3−x Quantum Dots and Bundle-Like Nanowires Using a One-Pot Template-Free Solvothermal Strategy and Their Versatile Applications. Small 2017, 13, 1603689. [Google Scholar] [CrossRef]
  41. Li, J.; Huang, J.; Wu, J.; Cao, L.; Li, Q.; Yanagisawa, K. Microwave-assisted growth of WO3·0.33H2O micro/nanostructures with enhanced visible light photocatalytic properties. CrystEngComm 2013, 15, 7904–7913. [Google Scholar] [CrossRef]
  42. Li, J.; Huang, J.; Yu, C.; Wu, J.; Cao, L.; Yanagisawa, K. Hierarchically Structured Snowflakelike WO3·0.33H2O Particles Prepared by a Facile, Green, and Microwave-assisted Method. Chem. Lett. 2011, 40, 579–581. [Google Scholar] [CrossRef]
  43. Gadewar, S.B.; Hofmann, H.M.; Doherty, M.F. Evolution of Crystal Shape. Cryst. Growth Des. 2004, 4, 109–112. [Google Scholar] [CrossRef]
Figure 1. Schematic of solvothermal synthesis procedure.
Figure 1. Schematic of solvothermal synthesis procedure.
Crystals 13 00690 g001
Figure 2. XRD data of samples 1–5 (see details in Table 1). For comparison: the top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Figure 2. XRD data of samples 1–5 (see details in Table 1). For comparison: the top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Crystals 13 00690 g002
Figure 3. XRD data of samples 6–10 (see details in Table 1). For comparison: the top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Figure 3. XRD data of samples 6–10 (see details in Table 1). For comparison: the top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Crystals 13 00690 g003
Figure 4. The effect of water content and precursor concentration on the crystal structure of as-synthesized powder.
Figure 4. The effect of water content and precursor concentration on the crystal structure of as-synthesized powder.
Crystals 13 00690 g004
Figure 5. SEM images of samples 1–10 (aj).
Figure 5. SEM images of samples 1–10 (aj).
Crystals 13 00690 g005
Figure 6. Low-magnification TEM images of (a) W18O49 (sample 1 and sample 6), (b) WO3∙0.33(H2O) (sample 2 and sample 3), and (c) γ-WO3 (sample 5 and samples 7–10); high-resolution TEM images (d) W18O49 (sample 1 and sample 6), (e) WO3∙0.33(H2O) (sample 2 and sample 3), and (f) γ-WO3 (sample 5 and samples 7–10). The inset in (b,c) is the selected area diffraction pattern.
Figure 6. Low-magnification TEM images of (a) W18O49 (sample 1 and sample 6), (b) WO3∙0.33(H2O) (sample 2 and sample 3), and (c) γ-WO3 (sample 5 and samples 7–10); high-resolution TEM images (d) W18O49 (sample 1 and sample 6), (e) WO3∙0.33(H2O) (sample 2 and sample 3), and (f) γ-WO3 (sample 5 and samples 7–10). The inset in (b,c) is the selected area diffraction pattern.
Crystals 13 00690 g006
Figure 7. XRD spectra of samples that are synthesized at temperature of 160 °C (sample 11), 180 °C (sample 12), and 200 °C (sample 3), each for 12 h. The top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Figure 7. XRD spectra of samples that are synthesized at temperature of 160 °C (sample 11), 180 °C (sample 12), and 200 °C (sample 3), each for 12 h. The top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Crystals 13 00690 g007
Figure 8. XRD spectra of samples that are synthesized at a temperature of 200 °C, but for 3 h (sample 13), 6 h (sample 14), 9 h (sample 15), and 12 h (sample 3). The top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Figure 8. XRD spectra of samples that are synthesized at a temperature of 200 °C, but for 3 h (sample 13), 6 h (sample 14), 9 h (sample 15), and 12 h (sample 3). The top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Crystals 13 00690 g008
Figure 9. (a) TGA curve of WO3∙0.33(H2O); (b) the first derivative of TGA curve. The first derivative curve is smoothed by Savitzky-Golay method with window of 20 points.
Figure 9. (a) TGA curve of WO3∙0.33(H2O); (b) the first derivative of TGA curve. The first derivative curve is smoothed by Savitzky-Golay method with window of 20 points.
Crystals 13 00690 g009
Figure 10. XRD spectra of samples that are annealed at temperature of 300 °C, 350 °C, 400 °C, 450 °C, 500 °C, and 550 °C for 2 h. The starting material is sample no. 3, which only consists of WO3∙0.33(H2O). The top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Figure 10. XRD spectra of samples that are annealed at temperature of 300 °C, 350 °C, 400 °C, 450 °C, 500 °C, and 550 °C for 2 h. The starting material is sample no. 3, which only consists of WO3∙0.33(H2O). The top inset is the standard reference diffraction patterns of W18O49 (JCPDS no. 36-0101), h-WO3 (JCPDS no. 75-2187), WO3∙0.33(H2O) (JCPDS no. 87-1203), and γ-WO3 (JCPDS no. 32-1295).
Crystals 13 00690 g010
Figure 11. SEM images of powder after annealed at (a) 300 °C (b) 350 °C (c) 400 °C (d) 450 °C (e) 500 °C, and (f) 550 °C.
Figure 11. SEM images of powder after annealed at (a) 300 °C (b) 350 °C (c) 400 °C (d) 450 °C (e) 500 °C, and (f) 550 °C.
Crystals 13 00690 g011
Table 1. Details of samples 1–10: list of volumes of the W2Cl4(OC2H5)6 solution (starting solution), the final solution, and the distilled water.
Table 1. Details of samples 1–10: list of volumes of the W2Cl4(OC2H5)6 solution (starting solution), the final solution, and the distilled water.
Sample No.Vstarting solutionVfinal solutionVwater[WCl6] for Final SolutionWater Content Synthesis TemperatureSynthesis Time
15 mL50 mL0 mL0.01 M0 vol%200 °C12 h
25 mL50 mL2.5 mL0.01 M5 vol%200 °C12 h
35 mL50 mL5 mL0.01 M10 vol%200 °C12 h
45 mL50 mL10 mL0.01 M20 vol%200 °C12 h
55 mL50 mL25 mL0.01 M50 vol%200 °C12 h
625 mL50 mL0 mL0.05 M0 vol%200 °C12 h
725 mL50 mL2.5 mL0.05 M5 vol%200 °C12 h
825 mL50 mL5 mL0.05 M10 vol%200 °C12 h
925 mL50 mL10 mL0.05 M20 vol%200 °C12 h
1025 mL50 mL25 mL0.05 M50 vol%200 °C12 h
Table 2. Details of the solvothermal synthesis: list of applied synthesis temperatures and times. For convenience, we use [WCl6] to denote precursor concentration.
Table 2. Details of the solvothermal synthesis: list of applied synthesis temperatures and times. For convenience, we use [WCl6] to denote precursor concentration.
Sample No.Vstarting solutionVfinal solutionVwater [WCl6] for Final SolutionWater ContentSynthesis TemperatureSynthesis Time
115 mL50 mL5 mL0.01 M10 vol%160 °C12 h
125 mL50 mL5 mL0.01 M10 vol%180 °C12 h
135 mL50 mL5 mL0.01 M10 vol%200 °C3 h
145 mL50 mL5 mL0.01 M10 vol%200 °C6 h
155 mL50 mL5 mL0.01 M10 vol%200 °C9 h
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Qiu, Z.; Jinschek, J.R.; Gouma, P.-I. Two-Step Solvothermal Process for Nanoarchitectonics of Metastable Hexagonal WO3 Nanoplates. Crystals 2023, 13, 690. https://doi.org/10.3390/cryst13040690

AMA Style

Qiu Z, Jinschek JR, Gouma P-I. Two-Step Solvothermal Process for Nanoarchitectonics of Metastable Hexagonal WO3 Nanoplates. Crystals. 2023; 13(4):690. https://doi.org/10.3390/cryst13040690

Chicago/Turabian Style

Qiu, Zanlin, Joerg R. Jinschek, and Pelagia-Irene Gouma. 2023. "Two-Step Solvothermal Process for Nanoarchitectonics of Metastable Hexagonal WO3 Nanoplates" Crystals 13, no. 4: 690. https://doi.org/10.3390/cryst13040690

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop