Hoon Kanga,
Gyung Hyun Kima,
Sungho Kanga,
Tae Hoon Parka,
Kwanchul Kima,
Hwan Kyu Kim
*b and
Yekyung Kim
*a
aAdvanced Institute of Convergence Technology, Seoul National University, Suwon-si, Gyeonggi-do 16229, Republic of Korea. E-mail: yekyung@snu.ac.kr
bDepartment of Advanced Materials Chemistry, Korea University, Sejong 30019, Republic of Korea. E-mail: hkk777@korea.ac.kr
First published on 7th August 2025
Electrochromic tungsten oxide (WO3) is one of the most commonly used materials for energy-efficient smart windows owing to its reversible optical changes driven by cation insertion/extraction. However, optimizing the fabrication process to balance ion mobility, structural stability, and long-term durability remains a critical challenge. In this study, WO3 thin films were fabricated via spin-coating followed by thermal treatment at various temperatures. The structural, optical, and electrochemical properties of the films were investigated to determine the optimal conditions for electrochromic performance. Characterization techniques such as differential scanning calorimetry, thermogravimetric analysis, X-ray diffraction, X-ray photoelectron spectroscopy, focused ion beam scanning electron microscopy, and electrochemical analysis revealed the critical role of thermal treatment in the fabrication of WO3 films. Amorphous WO3 formed at an optimal temperature of 300 °C exhibited the best Li+ insertion/extraction dynamics, which resulted in the highest coloration efficiency and stable electrochromic performance over 500 cycles. In contrast, crystalline WO3 fabricated at 350 °C and above showed reduced stability caused by restricted ion mobility in the dense structure, while films treated at insufficiently low temperatures (80 and 190 °C) exhibited significant performance decline attributed to structural weaknesses. This study underscores the importance of optimizing the thermal processing conditions for enhancing the durability and electrochromic performance of WO3 films, while emphasizing the necessity of assessing stability through long-term cycle testing.
WO3 (transparent) + xM+ + xe− ↔ MxWO3 (deep blue) | (1) |
A reduction reaction occurs when a negative voltage is applied, turning the film blue. Conversely, applying a positive voltage induces oxidation, which restores the film to its transparent state. This reversible reaction in WO3 is driven by electron transfer between the non-equilibrium states of W5+ and W6+. In amorphous WO3, this behavior is attributed to localized electrons causing polaronic transitions, whereas in crystalline WO3, it follows the Drude model.15–20 Structural variations in WO3 lead to differences in electron transport mechanisms, which influence the mobility of electrons and ions and ultimately affect the electrochromic performance. Thermal treatment is essential for optimizing the properties of WO3 thin films by adjusting their crystallinity and microstructure. Untreated WO3 thin films often suffer from structural instability, which limits their performance over extended cycling. In contrast, WO3 thin films that have been properly annealed demonstrate enhanced electrochemical stability due to reduced defects and improved amorphous structure. This leads to better electrochromic performance. For example, the transition from amorphous to crystalline WO3 alters ion diffusion pathways and electrochemical reactivity, which impacts the electrochromic performance. Amorphous WO3 provides high ion mobility and fast response times; however, it suffers from challenges in terms of maintaining structural stability during long-term operation. Conversely, crystalline WO3 offers better structural stability but restricts ion mobility because of its dense and orderly lattice structure. The electrochromic performance of WO3 thin films is significantly affected by its crystalline properties and by its structural, chemical, and morphological properties, which are highly dependent on the fabrication process. Despite the extensive studies on WO3, the relationship between thermal treatment conditions, structural changes, and electrochromic performance remains a subject of ongoing investigation.
Key challenges, such as ensuring reversibility and long-term electrochemical stability, need to be addressed to enable the widespread adoption of electrochromic devices across various applications. However, many studies focused on initial electrochromic performance, with less emphasis on long-term cycling stability critical for practical applications.
In this study, WO3 thin films were fabricated via spin-coating followed by thermal treatment at various temperatures, and the influence of their characteristics on the long-term stability of electrochromic performance was investigated. Although a simple and cost-effective deposition technique, spin-coating generally yields lower film uniformity and somewhat inferior electrochemical performance compared to methods such as sputtering. Nevertheless, spin-coating remains an attractive option owing to its versatility and ease of process optimization. Careful tuning of fabrication parameters such as precursor composition, coating speed, and post-treatment conditions can significantly improve the performance of spin-coated films, offering wide potential for applications. Through comprehensive analysis of the thermal, structural, chemical, and electrochemical properties of the prepared films, we identified optimal thermal treatment conditions for balancing ion mobility, structural stability, and long-term cycling performance. Moreover, the operational stability of the electrochromic WO3 layer was carefully examined for all films prepared at various annealing temperatures from low to high. The long-term operational behavior of WO3 was studied in terms of its crystallinity, chemical composition, and structural characteristics. By comparing the WO3 films, this study provides insights and foundational understanding for enhancing the long-term stability of these systems in future research.
WCl6 + C3H7OH → W(OC3H7)Cl5 + HCl | (2) |
W(OC3H7)Cl5 → WOCl4 + C3H7Cl | (3) |
WOCl4 + 2H2O → WO3 + 4HCl | (4) |
Among various alcohols, the moderately low vapor pressure and boiling point of IPA lead to stable evaporation, making it more effective for producing uniform thin films than other alcohols such as methanol or ethanol. The aforementioned sequential reactions indicate the conversion of WCl6 into WO3 through intermediate compounds involving hydrolysis and oxidation. The precursor solution was prepared by dissolving WCl6 in IPA, followed by stirring for an extended period, enabling reactions (2) and (3) to proceed naturally. The formation of WO3 via reaction (4) was driven by the interaction between WOCl4 and atmospheric moisture during the spin-coating and thermal treatment processes. In this study, ambient moisture was not restricted, and it affected the film formation process.
The effect of thermal treatment on WO3 formation was indirectly examined through DSC and TGA analyses. Fig. 1 shows the DSC and TGA curves obtained from the WO3 powder, prepared by drying the precursor solution at 50 °C. Upon drying, a hydrolysis reaction with ambient moisture was completed, facilitating the evaluation of the thermal behavior of WO3. The DSC curve exhibits broad endothermic peaks from RT to 200 °C, attributed to the evaporation of water, residual solvent, and by-products of the reaction. The first endothermic peak, observed as overlapping peaks at 91.46 and 123.60 °C (onset point of 30 °C and end point of 141 °C), was attributed to the evaporation of the residual solvent, reaction by-products, and water adsorbed on the surface. The second distinct peak at 157.62 °C (onset point of 141 °C and end point of 240 °C) was likely caused by the evaporation of water molecules chemically bonded to WO3 or incorporated within its structure. Finally, a weak exothermic reaction appeared beyond 350 °C, which can be attributed to the crystallization of WO3.21,27–29 This interpretation was supported by the TGA curve, which showed a significant mass loss up to 240 °C. The first endothermic peak region in the DSC curve, where two peaks overlap, corresponded to the range up to 141 °C, during which the TGA curve exhibited a weight loss of 8.38%. In the subsequent range up to 240 °C, which corresponded to the second peak in the DSC curve, an additional weight loss of 5.02% was recorded. This substantial weight loss indicated the release of moisture. Moreover, additional weight loss from 240 to 400 °C was only 2.56%. The weak exothermic response observed in the DSC curve above 350 °C, without a corresponding sharp mass loss, likely indicated WO3 crystallization. The thermal analysis highlighted the critical role of thermal processing and ambient moisture in the formation and properties of WO3 thin films. Moreover, the DSC measurements on WO3 powders dried at different temperatures revealed that peaks observed in Fig. 1 gradually diminished with an increase in drying temperature (Fig. S2). In the DSC curve of WO3 dried at 80 °C, the portion of the first endothermic peak at 93.60 °C was no longer observed. For powders dried at 300 and 400 °C, both endothermic peaks were nearly absent. These findings suggested that the WO3 prepared at higher temperatures contained a significantly lower amount of chemically or physically bound water.28
The crystalline state of electrochromic WO3 films was analyzed using XRD, and the diffraction patterns are presented in Fig. 2. The XRD pattern of the ITO substrate reveals several peaks in the 2θ range of 20–60°. A decrease in the intensity of the substrate peaks was observed in all thin film samples, with the formation of WO3 films attributed to spin-coating and the subsequent thermal treatment. In addition to the substrate peaks, no additional peaks were detected at thermal treatment temperatures up to 300 °C, indicating that films treated below this temperature exhibited no specific crystalline structure and were thus amorphous. Additional distinct peaks were observed for films with thermal treatment at 350 and 400 °C. These peaks signify the crystallization of the amorphous WO3. The crystalline WO3 formed at 350 °C and above exhibits an orthorhombic phase,21 with diffraction peaks at 2θ values of 23.15, 24.10, 33.85, 49.65, and 55.55°, corresponding to the (001), (200), (220), (400), and (420) planes, respectively (JCPDS No. 20-1324). These results confirm that the spin-coated WO3 precursor solution forms amorphous WO3 at thermal treatment temperatures up to 300 °C, whereas crystallization occurs at higher temperatures. This finding aligns with the thermal analysis results shown in Fig. 1, wherein the exothermic reaction starting at 350 °C corresponds to the onset of crystallization.
![]() | ||
Fig. 2 XRD diffraction patterns of WO3 thin films thermally treated at different temperatures (80, 190, 250, 300, 350, and 400 °C) and ITO substrate. |
XPS was performed for further investigation of the oxidation state of the films. The chemical states of W and O on the thin film surface were analyzed. High-resolution XPS spectra of W 4f and O 1s are presented in Fig. 3, with deconvolution performed using Gaussian–Lorentzian fitting. The deconvoluted W 4f spectrum (Fig. 3) reveals the presence of two tungsten species, W6+ and W5+. Prominent peaks at 35.89 ± 0.16 and 38.04 ± 0.17 eV correspond to W6+ 4f7/2 and W6+ 4f5/2, respectively, while subpeaks at 34.81 ± 0.20 and 36.83 ± 0.15 eV correspond to W5+ 4f7/2 and W5+ 4f5/2, respectively.31–36 Further, the O 1s spectrum comprises three species of O2−, OH−, and interstitial H2O, located at 530.67 ± 0.22, 532.03 ± 0.29, and 533.25 ± 0.09 eV, respectively.27,30 These results indicate that thin films formed via the wet chemical process are not composed solely of stoichiometric WO3 but also contain a small amount of W5+. The presence of W5+ in WO3 films is attributed to oxygen vacancies—associated with the presence of OH− species. Previous studies reported that these defective oxygen species enhance electrochromic performance.27,31 The analysis of the O 1s spectra supports this finding because the peak corresponding to OH− is the most prominent in WO3 films treated at 300 °C. To further quantify the degree of oxygen deficiency, the W5+ to W6+ ratio was estimated from the deconvoluted W 4f spectra. Using this ratio, the stoichiometric coefficient (x) in WOx was calculated by applying a chemical balance between the oxidation states, considering the total amount of tungsten detected (Table 1). These results indicate that WO300 possesses the highest density of oxygen vacancies among the thermally treated WOx thin films—excluding WO80, which was annealed at a temperature lower than the dominant moisture evaporation range. The chemical formula of WO300 was determined to be WO2.93, representing the most reduced state among the samples (excluding WO80).
Sample | W5+/W6+ ratio | Stoichiometric coefficient x in WOX |
---|---|---|
WO80 | 0.189 | 2.92 |
WO190 | 0.066 | 2.97 |
WO250 | 0.129 | 2.94 |
WO300 | 0.157 | 2.93 |
WO350 | 0.075 | 2.97 |
WO400 | 0.081 | 2.96 |
The surface morphology and cross-sectional structure of WO3 electrochromic films were investigated using FIB-SEM to examine the effects of thermal treatment (Fig. 4 and Fig. S3). Fig. 4 presents the surface and cross-sectional images of films treated at three representative temperatures (80, 300, and 400 °C) selected based on the characteristic features identified in the XRD and XPS analyses. Fig. S3 illustrates the changes in surface morphology across all thermal treatment conditions. As indicated in the magnified surface image (Fig. 4a-1), the WO3 film treated at 80 °C exhibited numerous cracks on its surface, as well as grains with nano-sized voids. These surface cracks and voids decreased significantly with an increase in the thermal treatment temperature—a trend clearly visible in the surface morphology images in Fig. S3. From 300 °C onward, the surface of the films became significantly smoother with a reduction in both cracks and intergranular voids. The cross-sectional images in Fig. 4 (labeled as Fig. 4a-2, 4b-2, and 4c-2) highlight the structural changes in the films with increasing temperature. At 80 °C, WO80 exhibited voids ranging from a few to a few hundred nanometers within its structure. The films showed progressively dense cross-sectional structures with an increase in the thermal treatment temperature, accompanied by a reduction in film thickness from 390 nm at 80 °C to 260 nm at 400 °C. This structural densification correlates with the formation rate of WO3 and the effect of ambient moisture. At lower thermal treatment temperatures, the slower formation rate of WO3 promotes granular aggregation, which leads to void formation between grains. At 80 °C, below the boiling point of water, the moisture within the film does not evaporate rapidly, resulting in a porous structure—consistent with the cracks and vacancies observed in the SEM images. In addition, the results of the thermal analysis support this observation as a large endothermic reaction corresponding to water evaporation was observed for the sample dried at 80 °C (Fig. S2). In contrast, films treated at higher temperatures (300 and 400 °C) exhibit smooth surfaces and dense cross-sections because of the rapid evaporation of internal moisture and accelerated formation of WO3. This is consistent with DSC results, where almost no water evaporation reaction was detected in the samples prepared at 300 and 400 °C (Fig. S2). These conditions suppress granule formation, which leads to a more uniform and compact structure. The differences in morphology and structure are likely governed by the thermal treatment process and its effect on the evaporation dynamics of water and nucleation and growth of WO3 (Table 2).
Samples | Surface morphology | Cross-sectional structure |
---|---|---|
WO80 | Highly porous with numerous irregular cracks | High porosity |
WO300 | Moderately porous with few visible cracks | Intermediate porosity |
WO400 | Densely packed and smooth surface without visible cracks | Dense |
The electrochemical redox reactions of the WO3 films were investigated. Fig. 5a illustrates the CV results for WO3 films thermally treated at 80, 300, and 400 °C, measured at a scan rate of 50 mV s−1. The CV results for all WO3 layers are presented in Fig. S4, and the scan rate-dependent behavior of the CV curves is shown in Fig. S5. When a voltage was applied in the negative direction, the WO3 exhibited a reduction current response caused by the insertion of Li+ ions from the electrolyte solution into the WO3 film. Conversely, the oxidation current response was due to the extraction reaction of Li+ ions from the film when a voltage was applied in the positive direction. The CV curves revealed that the oxidation peak potential shifted toward more negative voltages with an increase in the thermal treatment temperature. Furthermore, as the thermal treatment temperature increased, the current density initially rose to its highest value at 300 °C before decreasing for films prepared at 350 and 400 °C. This trend highlights the effect of thermal treatment on the electrochemical performance of the WO3 films. Amorphous WO3 structures are known to exhibit higher diffusion coefficients than crystalline structures, which makes them more favorable for electrochromic applications.32–35 In this study, WO3 films treated at 350 and 400 °C, where crystallization occurred, consistently showed significantly reduced electrochemical activity. This reduction was attributed to the dense crystalline structure of WO3, which impeded ion mobility and reduced electrochromic reactions, thereby resulting in lower current density. To confirm this, the diffusion coefficient was calculated from the CV data, and the interfacial resistance between the film and electrolyte was examined using impedance spectroscopy. Fig. S6 shows the diffusion coefficients (D) calculated using the Randles–Ševčík equation. The D values were 7.12 × 10−10 mA cm−2 for WO80, 1.76 × 10−9 mA cm−2 for WO300, and 1.84 × 10−10 mA cm−2 for WO400, indicating that the amorphous structure of WO300 has the highest ion exchange. The impedance spectra of WO80, WO300, and WO400 are presented in Fig. 5b. Electrochemical circle fitting was performed for the high-frequency region to extract Rs (ohmic resistance) and Rct (charge transfer resistance) for each film. The Rs values for the films treated at 80, 300, and 400 °C were 93.91, 88.08, and 210.57 Ω, respectively, while the corresponding Rct values were 25.63, 22.00, and 80.44 Ω. These results clearly demonstrate that the WO3 film prepared at 300 °C exhibited the lowest values for both ohmic resistance and Li+ charge transfer resistance.
![]() | ||
Fig. 5 (a) Cyclic voltammograms for WO80, WO300, and WO400 at a scan rate of 50 mV s−1. (b) Electrochemical impedance analysis for WO80, WO300, and WO400. |
Additionally, the spectra in the wavelength range of 200–800 nm for the as-deposited, bleached, and colored states are presented in Fig. S7. The bleached and colored states were measured after applying voltages of +1.0 and −0.8 V for 60 s, respectively. All WO3 films exhibited a significant decrease in transmittance within the visible light range in the colored state. The optical modulation (ΔT) at 550 nm and 630 nm is presented in Fig. S7, with the highest ΔT recorded for WO300. For amorphous WO3 (treated up to 300 °C), the transmittance spectra returned to their initial states after bleaching the colored film, which indicates high transparency reversibility. However, films treated at 350 and 400 °C (crystalline WO3) did not fully recover to their initial transmittance levels. This indicates that the Li+ ions inserted into WO3 were not fully extracted due to insufficient reaction time or slow electrochemical reaction kinetics during the oxidation potential application period, indicating that the insertion of Li+ ions was easier than their extraction from the crystalline structure of WO3. In addition, the optical bandgap (Eg) was analyzed using Tauc plot method (Fig. S8) based on the transmittance spectra measured in the range of 200–800 nm (as-deposited spectra shown in Fig. S7). Overall, Eg tended to decrease with increasing annealing temperature; however, the WO3 film annealed at 300 °C (WO300) exhibited an anomalously high Eg value of 3.79 eV. The formation of Eg is influenced by several factors, including intermolecular distance, grain size, and structural defects such as oxygen vacancies. Generally, a reduction in intermolecular distance and an increase in grain size lead to a narrower bandgap, whereas structural defects—particularly oxygen deficiencies—tend to widen the bandgap. In the WO3 films studied here, higher annealing temperatures promoted both grain growth and crystallinity, while simultaneously increasing the density of oxygen vacancies. These competing effects act in opposite directions on Eg. As a result of this interplay, WO300 exhibited a distinct increase in Eg, which can be attributed to the higher density of oxygen vacancies at this annealing temperature. These vacancies facilitate the partial reduction of W6+ to W5+, raising the Fermi level and widening the energy gap between the valence and conduction bands. In contrast, at temperatures above 300 °C, further crystallization leads to a decrease in the density of localized states and an enhancement of band alignment, resulting in a narrowing of the optical bandgap.36,37 Although a narrower bandgap is typically favorable for electron and ion transport, thus enhancing electrochemical reactivity, the results of this study suggest that increased crystallinity and reduced defect states may hinder charge mobility. It indicates that excessive structural ordering can negatively impact electrochemical performance.
To evaluate the effect of transparency reversibility and assess the long-term stability of the electrochromic WO3 films, cyclic stability tests were performed for 500 cycles of coloration (−0.8 V, 60 s) and bleaching (+1.0 V, 60 s). The transmittance during the cycle test was monitored at a wavelength of 550 nm, and the results are shown in Fig. 6 and summarized in Table 3. Differences in the initial transmittance (Fig. S7) and electrochemical properties (Fig. 5 and Fig. S4) were already observed among the WO3 fabricated under different thermal treatment conditions. Moreover, the cycle tests provided a clear distinction in the performance of the electrochromic WO3 films. As indicated in Fig. 6, WO300 exhibited the largest ΔT and the most stable transmittance behavior during the cycle test. In contrast, WO80 showed a rapid decrease in its performance with ΔT dropping below 10% after 45 cycles and below 5% after 75 cycles. Similarly, WO190 displayed a decline in ΔT to below 10% after 40 cycles and below 5% after 75 cycles. These results suggest that WO3 films treated at temperatures below 190 °C initially exhibit a high transmittance contrast; however, their coloration performance decreased significantly with operation. This dramatic performance degradation was correlated with the surface and cross-sectional structures observed in Fig. 4 and Fig. S3. The amorphous and porous structures of low-temperature annealed WO3 likely cannot withstand the structural stress caused by repeated Li+ ion insertion and extraction, thereby leading to rapid deterioration. Among the remaining films, WO300 demonstrated the most stable cycling performance, while WO250, WO350, and WO400 showed low stability, with ΔT decreasing to 13.09%, 19.03%, and 10.15%, respectively, after 500 cycles. However, these films did not exhibit the dramatic decline seen in WO80 and WO190. Additionally, the cycling behavior differed between the films prepared at 250 and 300 °C and those at 350 and 400 °C. The optical modulation of WO250 and WO300 decreased because of the increased transmittance in the colored state, with a minimal reduction in the bleached state. Conversely, WO350 and WO400 showed the opposite trend, with greater reductions in the transmittance of the bleached state while maintaining coloration. This behavior suggests that, in amorphous WO3, prolonged cycling restricts the insertion of Li+ ions into the film. However, in crystalline WO3 films, the dense structure impedes the mobility of Li+ ions, particularly during extraction, which leads to challenges in the bleaching process. In conclusion, the amorphous WO3 treated at the highest temperature before crystallization exhibited the best electrochromic performance, highlighting its structural advantage in facilitating Li+ insertion and extraction.
![]() | ||
Fig. 6 Transmittance behavior during the long-term cycle test at 550 nm for WO3 films treated at: (a) 80 °C, (b) 190 °C, (c) 250 °C, (d) 300 °C, (e) 350 °C, and (f) 400 °C. |
Samples | ΔT (%) | ΔT reduction ratio (%) | Response time (s) | CE (cm2 C−1) | ||||
---|---|---|---|---|---|---|---|---|
10th cycle (ΔT10) | 500th cycle (ΔT500) | 10th cycle | 500th cycle | |||||
tc | tb | tc | tb | |||||
WO80 | 33.18 | 0.22 | 99.34 | 8.71 | 3.18 | — | — | 126.41 |
WO190 | 17.74 | 2.09 | 88.23 | 8.41 | 1.98 | — | — | 43.84 |
WO250 | 41.50 | 13.09 | 68.47 | 13.61 | 2.46 | 13.62 | 4.63 | 127.11 |
WO300 | 58.84 | 38.65 | 34.32 | 11.03 | 2.93 | 10.20 | 2.94 | 217.39 |
WO350 | 29.35 | 19.03 | 35.16 | 33.27 | 23.68 | 21.22 | 34.56 | 78.61 |
WO400 | 23.03 | 10.15 | 55.93 | 21.53 | 34.47 | 17.70 | 42.18 | 65.04 |
The key metrics of electrochromic performance, such as response time and CE, were evaluated through the cycle test. Response time refers to the time required for a 90% change in ΔT during the transition from the bleached state to the colored state (coloration time, tc) or from the colored state to the bleached state (bleaching time, tb).38 Fig. 7a and b show transmittance changes recorded during the 10th and 500th cycles, respectively, while the initial and final response times are summarized in Table 3. For films annealed at 80 and 190 °C, the ΔT values after 500 cycles were reduced to 0.22 and 2.09%, respectively, making the calculation of response times unreliable because of the drastic decrease in transmittance. The WO3 film annealed at 250 °C exhibited a slight change in response time after 500 cycles, despite a 68.47% reduction in transmittance. This suggests that the amount of Li+ ions participating in the electrochromic reaction decreased while the insertion/extraction speed of Li+ ions remained largely unchanged. The WO3 film fabricated at 300 °C maintained consistent response times and exhibited the smallest reduction of transmittance between the 10th and 500th cycles, indicating high cycling stability. WO3 films with crystalline structures formed at 350 and 400 °C exhibited changes in response time between the initial and 500th cycles. Specifically, the coloration time decreased after 500 cycles, whereas the bleaching time increased; this trend was more pronounced in WO400 compared to WO350. These results indicate that repeated cycling facilitates Li+ insertion into the crystalline structure, thereby making it easier over time. However, the extraction of Li+ ions appears to be hindered by the dense crystalline structure, which leads to an increase in bleaching time. This observed variation in response time aligns with the findings in Fig. 6e and f, which demonstrate differences in transmittance behavior. The results highlight that, while crystalline WO3 films may exhibit improved Li+ insertion kinetics over the cycle test, the dense structure imposes significant limitations on ion extraction, thereby adversely affecting the bleaching process and overall performance.
![]() | ||
Fig. 7 Time-dependent transmittance behavior of WO3 films observed at 550 nm during the (a) 10th and (b) 500th cycles. |
CE, a key indicator of electrochromic performance, represents the amount of coloration achieved per unit charge injected into the material. It is defined as the ratio of the change in optical density to the charge density.
OD = log(tb/tc), | (5) |
CE = ΔOD/Δ(Q/A), | (6) |
![]() | ||
Fig. 8 Optical density vs. charge density plots measured at 550 nm under an applied voltage of −0.8 V for 60 s. |
To further investigate the long-term stability of the electrochromic thin film, long-term tests of up to 3000 cycles were performed. The preliminary results from 500 cycles showed that the color change occurred within 20 s (−0.8 V) and 10 s (+1.0 V) after voltage application. Therefore, the long-term stability test (3000 cycles) was performed for +1.0 and −0.8 V using 10 and 20 s durations, respectively. The transmittance degradation of WO300 was measured to be −23.21% at 500 cycles, −21.07% at 1000 cycles, and −36.79% at 3000 cycles, as shown in Fig. 9.
![]() | ||
Fig. 9 Transmittance changes during the 3000 cycles test at 550 nm for WO3 films treated at: (a) 80 °C, (b) 300 °C, and (c) 400 °C. |
Among the WO3 electrochromic layers thermally treated at various temperatures, WO300 exhibited the highest and most stable electrochromic performance, demonstrating the lowest degradation rate during the cycle test compared to that of other films. This superior performance was attributed to the structural properties of WO3, which were significantly influenced by the thermal treatment temperature.
This study highlighted the importance of optimizing thermal treatment conditions to enhance the electrochromic performance of WO3 layers. Proper thermal treatment can prevent excessive crystallization, facilitating Li+ ion mobility within the amorphous WO3 structure. In addition, structural stability in the film contributes to long-term operational durability. These factors collectively enable improved electrochromic performance and cycling stability. This study underscores the necessity of evaluating cycling stability over hundreds of cycles to fully determine the durability of electrochromic WO3 films. Performance assessments based solely on initial coloration and bleaching cycles or early electrochromic evaluations are insufficient to determine the long-term stability of films. Therefore, comprehensive cycling tests are critical to ensuring the practical applicability of WO3-based electrochromic devices. Furthermore, while spin-coating is a simple and cost-effective deposition technique, its electrochemical performance may be somewhat inferior compared to sputtering owing to lower film uniformity. Nevertheless, in this study, we systematically analyzed the electrochemical properties of spin-coated thin films at various annealing temperatures, providing valuable insights into their electrochromic durability. Although many studies have reported the electrochemical properties of various deposition processes, including spin-coating, relatively few have conducted long-term stability tests (Table 4). Therefore, our findings contribute to a broader understanding of the electrochromic durability of spin-coated films, while supporting the exploration of other deposition techniques to further improve film performance.
Material | Synthesis method | Electrolyte | ΔT [%] | ΔTx [%] | tc [s] | tb [s] | Voltage window | CE [cm2 C−1] |
---|---|---|---|---|---|---|---|---|
ΔTx: transmittance window at a specific cycle. | ||||||||
This study | Spin-coating | 0.1 M LiClO4-PC | 60.0 (550 nm) | 38.1% 3000th | 11.03 | 2.93 | −0.8 to +1.0 V | 217.39 |
WO3 film7 | Electrodepostion | 1 M LiClO4-PC | 80.25 (600 nm) | — | 3.5 | 6.0 | −1.0 to +1.0 V | 98.89 |
WO3 film41 | Electrodepostion | 0.5 M H2SO4 | 70.3% (750 nm) | 48.1% 3000th | 4.8 | 3.5 | −0.7 to +1.0 V | 43.2 |
Etched A-WO3 film42 | Hydrothermal method | 0.5 M LiClO4-PC | 49.3% (630 nm) | — | 1.98 | 2.5 | −0.7 to +0.7 V | 178.7 |
WO3 film43 | Spin-coating | 0.5 M LiClO4-PC | 54.7% | — | 6.7 | 2.6 | −0.6 to +1.0 V | 65.7 |
WO3 film44 | Spin-coating | 1 M PEG-LiI | 40.0% (633 nm) | — | 9.58 | 31.72 | −1.0 to +1.0 V | 34.8 |
WO3 film45 | Spin-coating | 0.5 M LiClO4-PC | 69.8% (500 nm) | — | — | — | −0.9 to +0.9 V | — |
WO3 film46 | Spin-coating | 1 M PEG-LiI | 40.0% (633 nm) | — | 31.72 | 9.58 | −1.0 to +1.0 V | 34.8 |
MXene/WO3 film47 | Spin-coating | 1.25 M LiClO4-PC![]() ![]() |
56.0% (660 nm) | 48.8% 200th | 13.0 | 6.0 | −3.0 to +3.0 V | 69.1 |
WO3 film48 | Spin-coating | 1 M LiClO4-PC | 54.8% (550 nm) | 53 | 36 | −1.0 to +1.0 V | 41.1 | |
WO3 nanosheet49 | Spin-coating | 1 M KCl | 41.78% (700 nm) | 35.91% 1000th | 10.5 | 9.2 | −3.0 to +3.0 V | — |
WO3 film41 | Spin-coating | 0.5 M H2SO4 | 31.7% (750 nm) | 4.85% 780th | 5.0 | 5.2 | −0.7 to +1.0 V | 29.5 |
WO3 film48 | Dip-coating | 1 M LiClO4-PC | 60.5% (550 nm) | 204 | 9.0 | −1.0 to +1.0 V | 41.8 | |
WO3 film50 | Magnetron-sputtering | 1 M LiClO4-PC | 74.4% (550 nm) | 69.5% 1000th | 11.3 | 7.0 | −0.9 to +0.9 V | 41.5 |
WO3 nanowire51 | Magnetron-sputtering | 1 M LiClO4-PC | 84.5% (633 nm) | 77.7% 3000th | 3.6 | 1.2 | −1.0 to +1.0 V | 83.6 |
WO3 film52 | CVD | 0.1 M H2SO4 | 55.6% (633 nm) | — | 3.8 | 9.7 | −1.0 to +1.0 V | 40.37 |
WO3 nanosheet53 | CVD | 1 M LiClO4-PC | 48.0% (798 nm) | — | 5.1 | 9.7 | −3.0 to +3.0 V | 120.9 |
The SI includes detailed data on the thermal, morphological, electrochemical, and electrochromic properties of the studied WO3 layer in this study. See DOI: https://doi.org/10.1039/d5nr02842f.
This journal is © The Royal Society of Chemistry 2025 |