Next Article in Journal
High-Speed GaN-Based Superluminescent Diode for 4.57 Gbps Visible Light Communication
Previous Article in Journal
Remediation of Cd (II) Ion from an Aqueous Solution by a Starch-Based Activated Carbon: Experimental and Density Functional Theory (DFT) Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controllable Electrodeposition Adjusts the Electrochromic Properties of Co and Mo Co-Modified WO3 Films

1
National Innovation Center of High Speed Train (Qingdao), Qingdao 266108, China
2
College of Materials Science and Engineering, Chongqing University, Chongqing 400044, China
3
National Innovation (Qingdao) High Speed Train Material Research Institute Co. Ltd., Qingdao 266109, China
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(2), 190; https://doi.org/10.3390/cryst12020190
Submission received: 31 December 2021 / Revised: 25 January 2022 / Accepted: 26 January 2022 / Published: 27 January 2022
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
Metal ion modification is considered to be an effective way to construct metal oxides with specific physical and chemical properties. In this paper, we prepare a tungsten oxide (WO3) film co-modified by Co-ion and Mo-ion to serve as the electrochromic material through a one-step electrodeposition method. The effect of electrodeposition time on film thickness, surface morphology and electrochromic properties is systematically studied as well. The results show that, compared with pure WO3 film, the surface morphology of the tungsten oxide film modified by Co-ion and Mo-ion (WO3: Co, Mo) is significantly different. The Co and Mo co-modified film possesses a higher transmission modulation (58.5% at 600 nm) and rapid switching speed (coloring and bleaching time are 2.7 s and 5.6 s, respectively), low impedance value and excellent cycle stability. The performance enhancement is mainly attributed to the coral-like structure of the membrane, which provides a larger specific surface area, more ion adsorption sites and faster ion diffusion. Therefore, this work provides a fast and low-cost method to prepare tungsten oxide electrochromic films co-modified with cobalt and molybdenum ions. At the same time, it also provides an idea to obtain films with different electrochromic properties by adjusting the film thickness.

1. Introduction

The ever-increasing energy demand has become the most severe challenge brought about by the rapid development of world economy. The current energy structure is still dominated by fossil energies (such as coal, oil, and natural gas). These traditional fossil energy reserves are limited and non-renewable, and their excessive consumption causes global environmental problems, such as climate warming [1,2,3,4,5]. Therefore, there is an urgent need to reasonably improve energy efficiency and maintain a sustainable economic development [6,7]. According to statistics, building energy consumption mainly includes heating in winter and cooling in summer, which account for approximately 40% of the total energy consumption of the whole society, so building energy savings are of great significance [8].
Electrochromic technology provides a new solution for energy savings in modern buildings, which can adjust the visible light transmittance of glass and solar radiation energy according to the human will. S. Selkowitz, a senior consultant in architectural science at Lawrence Berkeley National Laboratory, once pointed out that “electrochromic technology is the most promising dimming technology in architectural applications” [9]. Within a certain voltage range, the electrochromic device can obtain different transmittance states by applying different voltages and has a color memory effect. It can still maintain its transmittance state when the voltage is applied under open circuit conditions. This has a wide range of potential applications that are very attractive, such as energy-saving smart windows, large-area information displays and anti-glare car rearview mirrors.
Tungsten oxide-based electrochromic smart glass can not only adjust the transmittance of the visible light region, but also have a blocking effect on infrared light in the colored state [10]. Therefore, electrochromic smart glass can actively block the heat loss caused by the temperature difference between indoor and outdoor environments by controlling the voltage and reduce the large amount of energy consumption caused by building cooling and heating.
In response to the above requirements, scholars have conducted extensive research on the transition metal oxide WO3. However, the traditional WO3 material has poor diffusion coefficient, so the light modulation is not ideal, and the area capacitance is poor. The properties of materials are closely related to their structure. Therefore, if we want to change the electrochromic properties of WO3, we should start with its structure. Metal ion modification is a well-studied strategy that can enhance the electrochromic properties of WO3 [11]. By introducing appropriate metal ions, the film structure on the surface of the material is changed, and faster reaction kinetics, higher optical modulation efficiency and higher coloring efficiency can be achieved [11,12,13,14,15]. Green et al. used a low-concentration Ni-ion to modify WO3 films and found that this method can improve the coloring efficiency of the films [16,17]. Zhou et al. modified WO3 nanofibers with trace Mo-ion, which not only improved the electrochromic performance of the film, but also increased the capacitance of the film [18]. Xu et al. used 1.5% Co-ion to change the growth mode of WO3 fibers, thereby increasing the specific surface area of the WO3 film and making the film exhibit a better electrochromic performance in an oily electrolyte solution of lithium salt [19]. However, most of the preparation processes reported above are designed with strict heating, pressurizing and vacuuming processes, which will not be conducive to industrialization. Therefore, the synthesis of high-performance amorphous films on the surface of a substrate through a simple, fast, and low-cost method should be highly valued and further developed. The electrodeposition method is a very demanding method [20]. Because the electrodeposition process does not require high pressure and high temperature, the prepared samples do not require annealing treatment. In addition, considering that single metal ion modification can improve the electrochromic properties of the film, we carry out an exploration of the double metal ion modification film in this paper to further improve the properties.
In this study, we successfully synthesize amorphous WO3 films modified by Co-ion and Mo-ion on the surface of indium-tin-oxide (ITO) glass by the one-step electrodeposition method. The electrodeposition time was adjusted to control the thickness and morphology of the film, in order to obtain improved electrochromic properties. In addition, compared with the unmodified WO3 film, the Co-ion and Mo-ion co-modified film with the same electrodeposition time has a lower charge transfer impedance and higher optical modulation rate.

2. Experimental Procedure

2.1. Chemicals

All chemicals were of analytical grade and used as received without further purification. Ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O), cobalt nitrate hexahydrate (Co(NO3)2·6H2O), sodium tungstate dehydrate (Na2WO4·2H2O), 30% hydrogen peroxide (H2O2), 0.5 M sulfuric acid (H2SO4) and ethanol were purchased from Aladdin Chemical Reagent Co., Ltd. (Shanghai, China). All aqueous solutions were freshly prepared with deionized (DI) water.

2.2. Preparation of Co and Mo Modified WO3 Films

Briefly, 0.02 mol Na2WO4·2H2O, 0.15 mmol Co(NO3)2·6H2O and 0.15 mmol (NH4)6Mo7O24·4H2O were added to 100 mL deionized water, dissolved and mixed by ultrasonication and magnetic stirring. Then, 0.01 mol H2O2 was added to the above solution. At this point, many bubbles were generated, so the solution was continuously stirred to make the bubbles escape. After stirring and dissolving, H2SO4 was slowly added dropwise to the above solution, and the pH was measured with an accurate pH meter until the pH = 1.0. The purpose of adding acid is to convert Na2WO4 into H2WO4, waiting for further oxidation. The electrodeposition procedure was carried out through a CHI660B electrochemical workstation. ITO glass was used directly as the working electrode, and an Ag/AgCl electrode and a Pt plate served as the reference electrode and the counter electrode, respectively. The electrodeposition current density is 0.5 mA cm−2, and the deposition durations were 60, 180 and 300 s. Finally, the prepared film was flushed several times with deionized water and air dried. For comparison, we also prepared a pure WO3 film.

2.3. Characterization

To analyze the sample’s composition and microstructure, X-ray diffraction (XRD) was conducted on a PANalytical X’Pert Powder (Panalytical B.V., Shanghai, China). XRD data were processed using Jade6.5 software. Field-emission scanning electron microscopy (FESEM) analysis was performed on the JEOL JSM-7800F (JEOL, Tokyo, Japan) to characterize the surface morphologies. High-resolution transmission electron microscopy (HRTEM), high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM), and energy-dispersive spectroscopy (EDS) mapping were performed on the Talos F200S (Thermofishier, Waltham, MA, USA). X-ray photoelectron spectroscopy (XPS) experiments were performed on the ESCALAB250Xi (Thermofishier, Waltham, MA, USA), the numerical fitting of XPS was performed using Advantage software, and all of the XPS binding energies were calibrated using the contaminant carbon (C1S) as a reference. All electrochemical measurements were performed on CHI660E electrochemical workstation (Shanghai Chenhua Instrument Co., Ltd., Shanghai, China) by using the typical three-electrode set-up at 25 °C in 0.5 M H2SO4. The optical properties of the films were studied by a Shimadzu UV-vis–NIR spectrometer. The CV measurements were conducted in a voltage window between −0.4–0.6 V (vs. Ag/AgCl) at 20 mV/s. The electrochemical impedance spectroscopy (EIS) was performed in the frequency range from 100 kHz to 10 mHz at open circuit voltage by applying a 5 mV signal. All these electrochemical measurements were carried out at room temperature.

3. Results and Discussion

As shown in Figure 1a, a certain amount of ammonium molybdate tetrahydrate, cobalt nitrate hexahydrate and sodium tungstate dehydrate were dissolved in deionized water to form a wine-red transparent solution, which was thoroughly mixed by ultrasonic and magnetic stirring. Then, H2O2 and H2SO4 were added dropwise to the above solution to make the solution Ph = 1, and at this time the solution gradually became a bright yellow transparent solution. When a constant current electrodeposition started, a blue film was quickly deposited on the surface of the ITO glass. Because the ITO glass was the cathode, the film appeared blue. The action of the electric field caused the H+ in the electrolyte to be injected into the formed oxide film, causing it to undergo a coloring reaction. When the deposition time increased, metal oxide continued to grow to thicken the film, so that the number of sites in contact with hydrogen ions increased, and the color became darker. During the electrodeposition process under the action of the electric field, the particles in the electrolyte moved in a directional motion and inelastically collided with the substrate, and then the particles stopped moving and were adsorbed on the surface of the substrate. After a certain period of accumulation, a series of groups of a certain size were formed, which are called nuclei. After agglomeration and nucleation, as the particles continued to deposit, the volume of the nuclei continued to increase [21]. After preparing the Co and Mo co-modified WO3 film, we used XRD characterization to analyze the crystal structure of the film. As shown in Figure 1b, after excluding the diffraction peaks of ITO glass, there is a broad peak near 2θ ≈ 26° representing amorphous materials, indicating the presence of amorphous oxides. This result is consistent with previous reports [22,23]. We also proved that through STEM. The peaks of In2O3 in ITO glass are 30°, 35.3°, 37.6°, 50.9°, 55.8° and 60.5°.
SEM analysis was carried out to detect the influence of deposition time on the surface morphology and thickness of the film. Based on Figure 2a–c, we can clearly see that, as the deposition time increases, the surface morphology of the film changes dramatically. As shown in Figure 2a, when the electrodeposition time is 60 s, a dense film is formed on the surface of the ITO glass, and linear objects are formed at the cracks, similar to plants just sprouting on the ground. At this point, the thickness of the oxide film is only 0.42 μm (Figure 2d). When the electrodeposition time increases to 180 s, the cracks in the film increase and split into independent masses, similar to a dry riverbed, as shown in Figure 2b and Figure S1a. At the same time, there are many coral-like pillars in the cracks, which greatly increase the specific surface area of the film in contact with the electrolyte solution. The thickness of the film increases to 1.26 μm (Figure 2e). As the electrodeposition time continues to increase, the lumps on the surface of the ITO glass are gradually covered by coral-like pillars, as shown in Figure 2c and Figure S2. Correspondingly, the thickness of the film is increased to 1.78 μm (Figure 2f). Based on this, we can conclude that the deposition time has a significant effect on the morphology of the film. In addition, to explore the effect of Co-ion and Mo-ion modification on the morphology of the film, we also prepared a pure WO3 film with an electrodeposition time of 180 s. Additionally, the appearance of surface cracks can be attributed to the tensile stresses induced during drying. Coupling with the rapidly and continuous deposition of particles and uneven nucleation, the irregular formation of the film creates a coral-like surface morphology speeding everywhere as shown in Figure S2. This phenomenon is similar to what was reported in previous studies in [24,25,26]. Comparing Figure 2b,e with Figure S3, it can be clearly seen that, for the same electrodeposition time, the metal ion modification has no significant effect on the film thickness, but has a greater effect on the film’s morphology. In Figure S3, there are almost no coral-like pillars in the cracks, which reduces the specific surface area of the film in contact with the electrolyte solution to a certain extent. As shown in Figure S4, according to previous studies, the Mo ions modified film is microscopically irregular porous [20], while the Co ions modified film is needle-like [19]. In our work, the films co-modified by cobalt and molybdenum ions are coral-like at the microscopic level, which results from the synergistic effect of cobalt and molybdenum ions. In addition, we also used EDS to determine the distribution of elements and atomic ratios (Table S1). Figure 2g shows that the Co, Mo and W elements are distributed very uniformly. In the nano-scale STEM-EDS (Figure S5a), the above three elements are still uniformly distributed. To determine the crystal form of the film and explore the nanoscale morphology of the film, we scraped the film on the surface of the ITO glass and dispersed it in an ethanol solution for TEM detection. The high-resolution TEM image (Figure 2h) and the HAADF-STEM image (Figure S5b) show that the WO3: Co, Mo film is a stacked layered structure. The selected area electronic diffraction (SAED) (Figure 2h inset) proves that the WO3: Co, Mo film prepared by electrodeposition has an amorphous structure, which corresponds to the previous XRD test. However, when the scale is enlarged to 20 nm (Figure 2i), we can observe a few nanometer-scale low-range ordered regions (the blue part in the figure). This indicates that some crystalline nanocrystals are randomly dispersed in the amorphous matrix. According to previous studies, the electrochromic properties of inorganic materials have a great relationship with their crystallinity [27,28,29]. Crystalline WO3 deflects incident light due to the existence of its crystal lattice, so the film often shows a certain haze in the transparent state. At the same time, the ion intercalation channel of crystalline WO3 is narrow, which is limited to cation intercalation with a smaller radius. Therefore, crystalline WO3 exhibits poor electrochromic performance. However, the cycle stability of crystalline WO3 is better than that of amorphous WO3 [28,29]. Therefore, the formation of WO3 nanocrystals is expected to be a comprehensive method to improve the electrochromic properties of WO3 films.
To determine the chemical composition and valence state of metal ion of WO3: Co, Mo film, XPS was performed. Fitting the XPS peaks of the high-resolution Ti 2p spectrum in the sample (Figure 3a), the peaks at 35.7, 37.8 and 41.2 eV correspond to W 4f 7/2, W 4f 5/2 and W 4f 3/2, respectively [19,30]. For W 4f 7/2 and W 4f 5/2, the higher intensity peak is the W6+ oxidation state in the film, while the other lower intensity peaks can be assigned to W5+ [31,32]. In the Mo 3d region (Figure 3b), the two peaks can be assigned to Mo6+ 3d5/2 (232.7 eV) and Mo6+ 3d3/2 (235.8 eV) [20]. This shows that Mo atoms mainly exist in +6 oxidation states. Figure 3c shows the high-resolution Co 2p spectrum, and the peaks at 781.4 and 797.5 eV belong to the Co 2p3/2 and Co 2p1/2 characteristic peaks of cobalt in the +2 valence state, respectively. In addition, the satellite peaks appearing in the cobalt spectrum belong to their corresponding characteristic peaks of oxidized substances [19]. The difference between WO3: Co, Mo film and pure WO3 film in the W 4f spectra is shown in Figure 3d. When the Co-ion and Mo-ion are added to the film, the peaks shift slightly to lower binding energy, illustrating that the W5+/W6+ proportion increases and the formation of more oxygen vacancy defects [33,34]. This is an important mechanism that electrochromic function can be achieved, while pure WO3 is a stoichiometric compound whose electronic movement is almost inhibited and thus has no electrochromic function.
The CV test is an effective method to evaluate the ability of the film to insert and extract H-ion. The CV measurements were conducted in a voltage window between −0.4–0.6 V (vs. Ag/AgCl) at 20 mV/s in 0.5 M H2SO4 electrolyte. During each scan, when hydrogen ions are inserted, the film appears blue, and when hydrogen ions are extracted, the film becomes transparent. As shown in Figure 4a, the Co-ion and Mo-ion co-modified WO3 films show higher anode and cathode current densities than pure WO3 films. In addition, the increase in the CV area means that the ion storage capacity of the film becomes larger, which indicates that the addition of Co and Mo ions improves the electrochemical activity of the film. This is related to the modified nanostructure, which provides more electron transfer sites on the surface of the film. In addition, the Co-ion and Mo-ion co-modified WO3 films prepared at different electrodeposition time have different thicknesses, so the ion storage capacity has a large difference. As the thickness of the film increases, the ion storage capacity of the film increases, which also leads to differences in optical properties. The shift of the oxidation peak to a high potential indicates that the oxidation reaction on the surface of the film is more intense, which also explains the lower transmittance of the film in the colored state. Additionally, the charge storage capacity is an important indicator to evaluate the electrochromic performance. Therefore, we calculated the capacitance of the film from the CV plot obtained by cyclic voltammetry. Compared with the WO3 -180s film (119.9 mF), the capacitance of the WO3: Co, Mo -180s film (131.7 mF) is larger, indicating that it has a greater charge storage capacity, which also explains the lower transmittance of the WO3: Co, Mo -180s film in the colored state. In addition, the thickness of the film also has a great influence on the capacitance of the film, and the charge storage capacity of the WO3: Co, Mo -300s film (237.95 mF) is stronger than that of the WO3: Co, Mo -60s film (38.96 mF), which corresponds to the lower transmittance in the colored state.
By comparing the transmittance modulation in the visible and near-infrared regions, the effects of ion modification and deposition time on the optical properties of WO3 films were investigated, as shown in Figure 4b. For the Co-ion and Mo-ion co-modified WO3 films, the optical transmittance of the WO3: Co, Mo -60s film in the bleached state (Tb) at a wavelength of 600 nm is 83.1%, and the optical transmittance of the colored state (Tc) is 18.8%, which contributes to a transmittance modulation (ΔT = Tb − Tc) of 64.3%. Similarly, at a wavelength of 600 nm, the optical transmittance of the WO3: Co, Mo -180s film in the bleached state (Tb) is 68.2%, and the colored state (Tc) is 9.7%. Therefore, the transmittance modulation of the WO3: Co, Mo -180s film is 58.5%, which is larger than that of the pure WO3-180s film (67.8% of the bleached state and 24.1% of the colored state). The increase in transmittance modulation is attributed to the diffusion nanostructure formed after the addition of cobalt and molybdenum ions, which makes ion diffusion easier and maximizes the use of active areas for charge transfer reactions [19]. For the WO3: Co, Mo -300s film, the sharp drop in Tb may be due to the over larger thickness of the film, which limits the escape of H-ion. Based on the above discussion, Co-ion and Mo-ion modification can reduce the optical transmittance of the colored state (Tc), and the electrodeposition time also has a great influence on the transmittance modulation of the film.
The electrochromic time is also an important indicator representing the performance of the film. As can be seen from Figure 4c and Figure S6, the coloring time (tc) of the WO3: Co, Mo -180s film is 2.7 s, and the bleaching time (tb) is 5.6 s. These values are slightly higher than those of the pure WO3 film (tc = 2.5 s, tb = 5.3 s). This is because, after the specific surface area increases, the contact area between the film and the electrolyte increases, thereby exposing more active sites. Therefore, more hydrogen ions need to be adsorbed to achieve a lower colored state transmittance. As the film thickness increases, the tc and tb of the WO3: Co, Mo -300s film (tc = 8.5 s, tb = 8.6 s) also increase.
We used cyclic voltammetry to test the cyclic stability of the film, as shown in Figure 4d and Figure S6. After 1000 cycles, the WO3: Co, Mo -180s film still maintains 73.1% of ion storage capacity (96.3 mF), while the pure WO3-180s film only maintains 62.7% of ion storage capacity (75.2 mF). This shows that cobalt and molybdenum ions can improve the adhesion of the film to the glass substrate. For the WO3: Co, Mo -60s film, even though it shows the highest transmittance modulation, the thin film is corroded by sulfuric acid during the cycle, which causes its ion storage capacity to drop quickly. For the WO3: Co, Mo -300s film, due to the thick film, the extraction of H-ion is likely to be restricted, and some H-ion is trapped inside the film in the later stage, resulting in the performance degradation. It can be seen from the performance comparison chart (Figure 4e) that WO3: Co, Mo -180s film shows the best overall performance. In addition, we found that metal ion modification can reduce the charge transfer barrier of the film, as shown in Figure 4f. This is because the coral-like structure provides a larger specific surface area, more intercalation sites and effective ion transmission. The resistance of the WO3: Co, Mo -180s film (7.871 Ω) is lower than that of the pure WO3-180s film (8.282 Ω), which explains why its CV area is larger. Additionally, the thickness of the film also has a significant effect on the impedance; as the thickness of the film increases, its resistance also increases. The resistance of the WO3: Co, Mo -60s film (4.059 Ω) is significantly lower than that of the WO3: Co, Mo -300s film (10.331 Ω). The WO3: Co, Mo -180s film still exhibits good performance compared to other works in Table S2.

4. Conclusions

In summary, we used a one-step electrodeposition method to quickly deposit metal ion-modified WO3 films with different thicknesses on the surface of the ITO film. The films prepared by the electrodeposition time of 60 s, 180 s and 300 s have thicknesses of 0.42, 1.26 and 1.78 microns, respectively, and completely different morphologies. Both Co and Mo ion modification and electrodeposition time lead to different electrochromic properties of prepared films. Compared with pure tungsten oxide film, the WO3: Co, Mo -180s film has significantly enhanced electrochromic performance and higher transmission modulation (58.5% at 600 nm) and rapid switching speed (coloring and bleaching time are 2.7 s and 5.6 s, respectively), low impedance value and excellent cycle stability. This is because the coral-like structure formed after the modification of metal ions provides a larger specific surface area, more intercalation sites and effective ion transmission.

Supplementary Materials

The following supporting information can be downloaded at: www.mdpi.com/article/10.3390/cryst12020190/s1, Figure S1: SEM images of WO3: Co, Mo film prepared by electrodeposition for 180s. (a) The scale is 10 μm. (b) The scale is 1 μm; Figure S2: SEM image of WO3: Co, Mo film prepared by electrodeposition for 300s; Figure S3: (a) SEM image of pure WO3 film prepared by electrodeposition for 180s. (b) Cross-sectional SEM images of pure WO3 film prepared by electrodeposition for 180s; Figure S4: (a)STEM-EDS image of WO3: Co, Mo film prepared by electrodeposition for 180s. (b) HAADF-STEM image of WO3: Co, Mo film prepared by electrodeposition for 180s; Figure S5: The switching time of (a) pure WO3-180s film, (b) WO3: Co, Mo -60s film and (c) WO3: Co, Mo -300s film at the wavelength of 600 nm; Figure S6. The stability of (a) pure WO3-180s film, (b) WO3: Co, Mo -60s film and (c) WO3: Co, Mo -300s film; Table S1: The ratio of cobalt and molybdenum ions in different films; Table S2: Comparison of the electrochromic performance of WO3: Co, Mo -180s with other reports.

Author Contributions

L.J. and W.M. contributed equally to this work. Conceptualization, L.J. and W.M.; methodology, L.J. and W.M.; software, L.J. and W.M.; validation, L.J., W.M. and J.D.; formal analysis, L.J. and W.M.; investigation, L.J. and W.M.; resources, L.J. and W.M.; data curation, L.J. and W.M.; writing—original draft preparation, L.J. and W.M.; writing—review and editing, J.D.; visualization, Q.Z.; supervision, Y.Z.; project administration, J.D.; funding acquisition, J.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Independent research project of State Key Laboratory of Mechanical Transmission of China, grant number SKLMT-ZZKT-2021M10.

Data Availability Statement

Some or all data that support the findings of thisstudy are available from the corresponding author upon reasonable request.

Acknowledgments

Thanks are owed to the financial supports from the Qingdao science and technology bureau, and the management committee of the Qingdao rail transit industry demonstration zone.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Engelhardt, V.; Kuhri, S.; Fleischhauer, J.; García-Iglesias, M.; González-Rodríguez, D.; Bottari, G.; Torres, T.; Guldi, D.M.; Faust, R. Light-harvesting with panchromatically absorbing BODIPY–porphyrazine conjugates to power electron transfer in supramolecular donor–acceptor ensembles. Chem. Sci. 2013, 4, 3888–3893. [Google Scholar] [CrossRef]
  3. Lv, Z.; Ma, W.; Dang, J.; Wang, M.; Jian, K.; Liu, D.; Huang, D. Induction of Co2P Growth on a MXene (Ti3C2Tx)-Modified Self-Supporting Electrode for Efficient Overall Water Splitting. J. Phys. Chem. Lett. 2021, 12, 4841–4848. [Google Scholar] [CrossRef] [PubMed]
  4. Lv, Z.; Ma, W.; Wang, M.; Dang, J.; Jian, K.; Liu, D.; Huang, D. Co-Constructing Interfaces of Multiheterostructure on MXene (Ti3C2Tx)-Modified 3D Self-Supporting Electrode for Ultraefficient Electrocatalytic HER in Alkaline Media. Adv. Funct. Mat. 2021, 31, 2102576. [Google Scholar] [CrossRef]
  5. Lv, Z.; Wang, M.; Liu, D.; Jian, K.; Zhang, R.; Dang, J. Synergetic Effect of Ni2P and MXene Enhances Catalytic Activity in the Hydrogen Evolution Reaction. Inorg. Chem. 2021, 60, 1604–1611. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, M.; Ma, W.; Lv, Z.; Liu, D.; Jian, K.; Dang, J. Co-Doped Ni3N Nanosheets with Electron Redistribution as Bifunctional Electrocatalysts for Efficient Water Splitting. J. Phys. Chem. Lett. 2021, 12, 1581–1587. [Google Scholar] [CrossRef]
  7. Jian, K.; Ma, W.; Lv, Z.; Wang, M.; Lv, X.; Li, Q.; Dang, J. Tuning the Electronic Structure of the CoP/Ni2P Nanostructure by Nitrogen Doping for an Efficient Hydrogen Evolution Reaction in Alkaline Media. Inorg. Chem. 2021, 60, 18544–18552. [Google Scholar] [CrossRef]
  8. Granqvist, C.G. Oxide electrochromics: An introduction to devices and materials. Sol. Energy Mater. Sol. Cells 2012, 99, 1–13. [Google Scholar] [CrossRef]
  9. Granqvist, C.G. Electrochromics for smart windows: Oxide-based thin films and devices. Thin Solid Films 2014, 564, 1–38. [Google Scholar] [CrossRef]
  10. Sorar, I.; Welearegay, T.G.; Primetzhofer, D.; Österlund, L.; Granqvist, C.G.; Niklasson, G.A. Electrochromism in Ni Oxide Thin Films Made by Advanced Gas Deposition and Sputtering: A Comparative Study Demonstrating the Significance of Surface Effects. J. Electrochem. Soc. 2020, 167, 116519. [Google Scholar] [CrossRef]
  11. Zhang, L.; Man, Y.; Zhu, Y. Effects of Mo Replacement on the Structure and Visible-Light-Induced Photocatalytic Performances of Bi2WO6 Photocatalyst. ACS Catal. 2011, 1, 841–848. [Google Scholar] [CrossRef]
  12. Xue, J.; Ma, S.; Bi, Q.; Zhang, X.; Guan, W.; Gao, Y. Revealing the modification mechanism of La-doped Ti/SnO2 electrodes related to the microelectronic structure by first-principles calculations. J. Alloys Compd. 2018, 747, 423–430. [Google Scholar] [CrossRef]
  13. Tan, J.; Han, Y.; He, L.; Dong, Y.; Xu, X.; Liu, D.; Yan, H.; Yu, Q.; Huang, C.; Mai, L. In situ nitrogen-doped mesoporous carbon nanofibers as flexible freestanding electrodes for high-performance supercapacitors. J. Mater. Chem. A 2017, 5, 23620–23627. [Google Scholar] [CrossRef]
  14. Wu, W.; Fang, H.; Ma, H.; Wu, L.; Zhang, W.; Wang, H. Boosting Transport Kinetics of Ions and Electrons Simultaneously by Ti3C2Tx (MXene) Addition for Enhanced Electrochromic Performance. Nano-Micro Lett. 2020, 13, 20. [Google Scholar] [CrossRef] [PubMed]
  15. Li, W.; Zhang, J.; Zheng, Y.; Cui, Y. High performance electrochromic energy storage devices based on Mo-doped crystalline/amorphous WO3 core-shell structures. Sol. Energy Mater. Sol. Cells 2022, 235, 111488. [Google Scholar] [CrossRef]
  16. Green, S.V.; Pehlivan, E.; Granqvist, C.G.; Niklasson, G.A. Electrochromism in sputter deposited nickel-containing tungsten oxide films. Sol. Energy Mater. Sol. Cells 2012, 99, 339–344. [Google Scholar] [CrossRef]
  17. Green, S.V.; Kuzmin, A.; Purans, J.; Granqvist, C.G.; Niklasson, G.A. Structure and composition of sputter-deposited nickel-tungsten oxide films. Thin Solid Film. 2011, 519, 2062–2066. [Google Scholar] [CrossRef] [Green Version]
  18. Zhou, D.; Shi, F.; Xie, D.; Wang, D.H.; Xia, X.H.; Wang, X.L.; Gu, C.D.; Tu, J.P. Bi-functional Mo-doped WO3 nanowire array electrochromism-plus electrochemical energy storage. J. Colloid Interf. Sci. 2016, 465, 112–120. [Google Scholar] [CrossRef] [PubMed]
  19. Shen, K.; Sheng, K.; Wang, Z.; Zheng, J.; Xu, C. Cobalt ions doped tungsten oxide nanowires achieved vertically aligned nanostructure with enhanced electrochromic properties. Appl. Surf. Sci. 2020, 501, 144003. [Google Scholar] [CrossRef]
  20. Xie, S.; Bi, Z.; Chen, Y.; He, X.; Guo, X.; Gao, X.; Li, X. Electrodeposited Mo-doped WO3 film with large optical modulation and high areal capacitance toward electrochromic energy-storage applications. Appl. Surf. Sci. 2018, 459, 774–781. [Google Scholar] [CrossRef]
  21. Tesler, A.B.; Kim, P.; Kolle, S.; Howell, C.; Ahanotu, O.; Aizenberg, J. Extremely durable biofouling-resistant metallic surfaces based on electrodeposited nanoporous tungstite films on steel. Nat. Commun. 2015, 6, 8649. [Google Scholar] [CrossRef] [PubMed]
  22. Cai, G.F.; Zhou, D.; Xiong, Q.Q.; Zhang, J.H.; Wang, X.L.; Gu, C.D.; Tu, J.P. Efficient electrochromic materials based on TiO2@WO3 core/shell nanorod arrays. Sol. Energy Mater. Sol. Cells 2013, 117, 231–238. [Google Scholar] [CrossRef]
  23. Zhang, J.; Tu, J.P.; Cai, G.F.; Du, G.H.; Wang, X.L.; Liu, P.C. Enhanced electrochromic performance of highly ordered, macroporous WO3 arrays electrodeposited using polystyrene colloidal crystals as template. Electrochim. Acta 2013, 99, 1–8. [Google Scholar] [CrossRef]
  24. Bi, Z.; Li, X.; Chen, Y.; He, X.; Xu, X.; Gao, X. Large-Scale Multifunctional Electrochromic-Energy Storage Device Based on Tungsten Trioxide Monohydrate Nanosheets and Prussian White. ACS Appl. Mater. Inter. 2017, 9, 29872–29880. [Google Scholar] [CrossRef]
  25. Cai, G.F.; Gu, C.D.; Zhang, J.; Liu, P.C.; Wang, X.L.; You, Y.H.; Tu, J.P. Ultra fast electrochromic switching of nanostructured NiO films electrodeposited from choline chloride-based ionic liquid. Electrochim. Acta 2013, 87, 341–347. [Google Scholar] [CrossRef]
  26. Au, B.W.C.; Tamang, A.; Knipp, D.; Chan, K.Y. Post-annealing effect on the electrochromic properties of WO3 films. Opt. Mater. 2020, 108, 110426. [Google Scholar] [CrossRef]
  27. Cai, G.; Wang, X.; Zhou, D.; Zhang, J.; Xiong, Q.; Gu, C.; Tu, J. Hierarchical structure Ti-doped WO3 film with improved electrochromism in visible-infrared region. RSC Adv. 2013, 3, 6896–6905. [Google Scholar] [CrossRef]
  28. Llordés, A.; Garcia, G.; Gazquez, J.; Milliron, D.J. Tunable near-infrared and visible-light transmittance in nanocrystal-in-glass composites. Nature 2013, 500, 323–326. [Google Scholar] [CrossRef]
  29. Lee, S.H.; Deshpande, R.; Parilla, P.A.; Jones, K.M.; To, B.; Mahan, A.H.; Dillon, A.C. Crystalline WO3 Nanoparticles for Highly Improved Electrochromic Applications. Adv. Mater. 2006, 18, 763–766. [Google Scholar] [CrossRef]
  30. Cai, G.; Cui, M.; Kumar, V.; Darmawan, P.; Wang, J.; Wang, X.; Lee-Sie Eh, A.; Qian, K.; Lee, P.S. Ultra-large optical modulation of electrochromic porous WO3 film and the local monitoring of redox activity. Chem. Sci. 2016, 7, 1373–1382. [Google Scholar] [CrossRef] [Green Version]
  31. Yang, P.; Sun, P.; Chai, Z.; Huang, L.; Cai, X.; Tan, S.; Song, J.; Mai, W. Large-Scale Fabrication of Pseudocapacitive Glass Windows that Combine Electrochromism and Energy Storage. Angew. Chem. Int. Edit. 2014, 53, 11935–11939. [Google Scholar] [CrossRef] [PubMed]
  32. Cong, S.; Tian, Y.; Li, Q.; Zhao, Z.; Geng, F. Single-Crystalline Tungsten Oxide Quantum Dots for Fast Pseudocapacitor and Electrochromic Applications. Adv. Mater. 2014, 26, 4260–4267. [Google Scholar] [CrossRef] [PubMed]
  33. Koo, B.R.; Kim, K.H.; Ahn, H.J. Switching electrochromic performance improvement enabled by highly developed mesopores and oxygen vacancy defects of Fe-doped WO3 films. Appl. Surf. Sci. 2018, 453, 238–244. [Google Scholar] [CrossRef]
  34. Bon-Ryul, K.; Kim, K.H.; Ahn, H.J. Novel tunneled phosphorus-doped WO3 films achieved using ignited red phosphorus for stable and fast switching electrochromic performances. Nanoscale 2019, 11, 3318–3325. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Schematic illustration of the fabrication of Co and Mo co-modified WO3 film. (b) XRD patterns of the prepared films.
Figure 1. (a) Schematic illustration of the fabrication of Co and Mo co-modified WO3 film. (b) XRD patterns of the prepared films.
Crystals 12 00190 g001
Figure 2. Scanning electron micrographs of WO3: Co, Mo films prepared with different electrodeposition time: (a) 60 s; (b) 180 s; (c) 300 s. Cross-sectional SEM images of WO3: Co, Mo films prepared with different electrodeposition time: (d) 60 s; (e) 180 s; (f) 300 s. (g) SEM-EDS spectrum of WO3: Co, Mo film prepared by electrodeposition for 180 s. (h) Transmission electron micrograph of WO3: Co, Mo film prepared by electrodeposition for 180 s and its 20 nanometer scale (i). The inset displays the SAED pattern with a 10 nm−1 scale bar.
Figure 2. Scanning electron micrographs of WO3: Co, Mo films prepared with different electrodeposition time: (a) 60 s; (b) 180 s; (c) 300 s. Cross-sectional SEM images of WO3: Co, Mo films prepared with different electrodeposition time: (d) 60 s; (e) 180 s; (f) 300 s. (g) SEM-EDS spectrum of WO3: Co, Mo film prepared by electrodeposition for 180 s. (h) Transmission electron micrograph of WO3: Co, Mo film prepared by electrodeposition for 180 s and its 20 nanometer scale (i). The inset displays the SAED pattern with a 10 nm−1 scale bar.
Crystals 12 00190 g002
Figure 3. XPS spectra of WO3: Co, Mo film prepared by electrodeposition for 180 s. High-resolution of (a) W 4f, (b) Mo 3d, (c) Co 2p. (d) The difference between WO3: Co, Mo film and pure WO3 film in the W 4f spectra.
Figure 3. XPS spectra of WO3: Co, Mo film prepared by electrodeposition for 180 s. High-resolution of (a) W 4f, (b) Mo 3d, (c) Co 2p. (d) The difference between WO3: Co, Mo film and pure WO3 film in the W 4f spectra.
Crystals 12 00190 g003
Figure 4. (a) CV curves of different films. (b) Transmittance spectra of different films. The solid line represents the bleached state, and the dotted line represents the colored state. (c) The switching time of WO3: Co, Mo -180s film at a wavelength of 600 nm, the illustration shows the colored and bleached state of the film. (d) The stability of WO3: Co, Mo -180s film. (e) The comparison between different films in transmittance modulation and cycle stability. (f) EIS of the different films.
Figure 4. (a) CV curves of different films. (b) Transmittance spectra of different films. The solid line represents the bleached state, and the dotted line represents the colored state. (c) The switching time of WO3: Co, Mo -180s film at a wavelength of 600 nm, the illustration shows the colored and bleached state of the film. (d) The stability of WO3: Co, Mo -180s film. (e) The comparison between different films in transmittance modulation and cycle stability. (f) EIS of the different films.
Crystals 12 00190 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jia, L.; Ma, W.; Zhuang, Q.; Zhang, Y.; Dang, J. Controllable Electrodeposition Adjusts the Electrochromic Properties of Co and Mo Co-Modified WO3 Films. Crystals 2022, 12, 190. https://doi.org/10.3390/cryst12020190

AMA Style

Jia L, Ma W, Zhuang Q, Zhang Y, Dang J. Controllable Electrodeposition Adjusts the Electrochromic Properties of Co and Mo Co-Modified WO3 Films. Crystals. 2022; 12(2):190. https://doi.org/10.3390/cryst12020190

Chicago/Turabian Style

Jia, Lihong, Wansen Ma, Qianyu Zhuang, Yani Zhang, and Jie Dang. 2022. "Controllable Electrodeposition Adjusts the Electrochromic Properties of Co and Mo Co-Modified WO3 Films" Crystals 12, no. 2: 190. https://doi.org/10.3390/cryst12020190

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