Next Article in Journal
Synergistically Boosting Li Storage Performance of MnWO4 Nanorods Anode via Carbon Coating and Additives
Previous Article in Journal
Self-Healing Composites: A Path to Redefining Material Resilience—A Comprehensive Recent Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Excellent Electrochromic Properties of Ti4+-Induced Nanowires V2O5 Films

1
Department of Materials Chemistry, School of Materials Science and Engineering, Jingdezhen Ceramic University, Jingdezhen 333403, China
2
Laboratoire de Recherche en Matériaux et Micro-Spectroscopies Raman et FTIR, Université de Moncton-Campus de Shippagan, Shippagan, NB E8S 1P6, Canada
3
National Engineering Research Centre for Domestic & Building Ceramics, Jingdezhen Ceramic University, Jingdezhen 333001, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(19), 4680; https://doi.org/10.3390/ma17194680
Submission received: 25 July 2024 / Revised: 13 September 2024 / Accepted: 20 September 2024 / Published: 24 September 2024

Abstract

:
Ti4+-doped V2O5 films with nanowires on top and a dense, long nanorod layer on the bottom were successfully fabricated using the spin-coating route. During the electrochromic cycling, charge transfer resistance (Rct) decreases while ion-diffusion ability (KΩ) rapidly drops in the first ten cycles and then levels off. Low Rct and morphology of nanowires collaboratively improved the electrochromic behavior of Ti4+-doped V2O5 films by enhancing the charge transfer speed and minimizing polarization and dissolution. The obtained Ti4+-doped V2O5 film shows better electrochromic properties than the undoped V2O5 film, with a coloration efficiency (CE) of 34.15 cm2/C, coloration time of 9.00 s, and cyclic retention of 82.6% at cycle 100. In contrast, the corresponding values for the undoped V2O5 film were 23.57 cm2/C, 13.16 s, and 43.6%.

Graphical Abstract

1. Introduction

Transition metal oxides (TMOs) offer a wide range of applications for energy storage and energy conversion [1], waste treatment [2], and gas sensing, to name a few. For instance, Ishibe et al. showed the power enhancement of embedded-ZnO nanowire structures for energy conversion [3], while Maeng et al. presented a highly sensitive SnO2 nano slab for NO2 gas sensing [4].
Among TMOs, V2O5, owing to its layered structure and high lithium-ion intercalation capacity, has attracted considerable attention for promising applications in devices such as batteries [5,6], supercapacitors [7], and electrochromic (EC) smart windows [8]. However, the practical applications of V2O5 in these fields are limited due to its poor cycling stability [9,10], which is generally attributed to low electronic conductivity and/or ionic conductivity [11,12], slow ion diffusion [13], and inert fragile structure [14].
Aiming to increase conductivity, some strategies have been adopted, such as incorporating a highly conductive material, introducing low valence or non-stoichiometric vanadium oxide [8], introducing a nanostructure or porous structure design, doping, or adding a small amount of another oxide [15]. On one hand, conductive materials, such as carbonaceous materials [16], conductive polymers [17], and metal oxides [18], significantly improve cycling stability by reducing the coexistence of multiple phases and dissolution of soluble intermediates in these multiple phases through fast charge transfer speeds [19]. On the other hand, they often decrease the visible transmittance region, which is inappropriate for EC smart window applications. The same issue occurs when introducing low valence vanadium oxide. Apart from incorporating a highly electrically conductive material, nanostructured materials and porous structures have also interested researchers because of their large surface area and corresponding shortened diffusion pathways for Li+ [20]. Various nanostructures have been prepared, such as nanorods [20], nanobelts [21], and even hierarchical nanostructures [22]. Porous structures [23,24], with an interconnected network of nanometer-thick walls suitable for electrolyte penetration and giving effective charge/ions transport pathways, were also prepared and deeply explored. Porous and nanostructured V2O5 suffer from poor stability due to their intrinsic low conductivity. However, by compositing V2O5 with other metal oxides [15,25,26] or doping with heteroatoms such as Y [27], Cu [28], Sn [29], etc., capacity retention has been greatly improved. The improvement was attributed to low impedance and, therefore, smaller polarization and faster kinetics [27].
Other than cycling stability, switching time, coloration efficiency (CE), and optical contrast are also critical parameters in an EC device (ECD). Switching time depends on several factors, including electronic conductivity, ion diffusion ability, the magnitude of the applied potential, morphology, and so forth [30,31]. Coloration efficiency is contingent upon the specific type of EC material, the quantity of charge insertion, and the voltage program [31]. In contrast to its application in batteries, the coloration mechanism of V2O5 in EC devices is far less explored and yet to be clearly understood. Furthermore, the present works on V2O5 mainly focus on exploring the effect of initial impedance on cycling stability, presuming that the impedance is constant during cycling. However, this is not the case since chemical dissolution and structural fracture continuously occur during cycling.
Herein, we first prepared Ti4+-doped V2O5 films with nanowire morphology on top and a dense, long nanorod layer on the bottom using the spin-coating route followed by annealing in the air. To our knowledge, this is the first time that the collation between EC properties and impedance of Ti-induced nanowire film has been dynamically investigated upon cycling. This gives us a better understanding of decaying upon cycling and, therefore, benefits the potential practical application of V2O5-based films. The EC properties upon cycling, including coloration time and cycling stability, were evaluated.

2. Materials and Methods

2.1. Materials

Vanadium oxytripropoxide (OV(OC3H7)3, VTIP) (98%), titanium tetraisopropoxide (Ti(OC3H7)4, TTIP) (97%), and lithium perchlorate (LiClO4) (99%) were purchased from Sigma-Aldrich LLC (St. Louis, MO, USA). Isopropanol (C3H7OH) (99%), acetylacetone (C5H8O2) (99%), acetic acid (CH3-CO2H) (99%), and ethanol (C2H5OH) (99.7%) were purchased from Sinoreagent Co. Ltd. (Beijing, China). Propylene carbonate (C4H6O3) (99%) was purchased from Ourchem LLC. (Guangzhou, China), and Triton X-100 (C34H62O11) was purchased from Macklin LLC (Albany, NY, USA).

2.2. Synthesis of V2O5 Precursor Solution

The V2O5 precursor solution was obtained by dissolving 0.3 mL vanadium oxytripropoxide into 9.98 mL isopropanol, which was added into an ethanolic solution of acetic acid (0.012 mL acetic acid in 9.98 mL ethanol). Then, 1 mL Triton X-100 was added to the obtained solution and stirred for 1.5 h.

2.3. Synthesis of Ti4+-Doped V2O5 Precursor Solution

First, solution A was prepared by adding 0.6 mL titanium tetraisopropoxide into a 12 μL acetylacetone in a 7.9 mL isopropanol solution under continuous stirring until a clear transparent solution was formed. Second, solution B was obtained by introducing a 0.3 mL vanadium oxytripropoxide solution in 9.98 mL isopropanol into a mixture of 0.012 mL acetic acid with 9.98 mL ethanol. Then, 0.195 mL of solution A was added gradually, in drops, to solution B until the solution cleared. The final solution was obtained by adding 1 mL Triton X-100 to the mixture of solutions A and B and stirring for another 1.5 h.

2.4. Fabrication of Undoped V2O5 and Ti4+-Doped V2O5 Films

The films were formed on transparent indium tin oxide (ITO)-conducting glass substrates by using spin-coating. The precursor solutions (36 μL drops) were spread onto 2.5 cm × 2.5 cm ITO substrates and allowed to uniformly cover the substrates for 30 s before starting the spin-coating process for 30 s at a spin rate of 1000 revolutions per minute (rpm). The spin-coating process was repeated twice to obtain three layers on the ITO glass. A sufficient time interval (30 s) was provided for air-drying between successive coatings.
All films were further calcined at 450 °C for two hours. Thus, the films obtained were V2O5 film and Ti4+-doped V2O5 film, respectively.

2.5. Fabrication of EC Devices

The EC devices were constructed with the following configuration: ITO-coated-glass-1/Ti4+-doped V2O5 film (or V2O5 film)/electrolyte/ITO-coated-glass-2, where ITO-coated-glass-1 and ITO-coated-glass-2 are the two transparent electrodes (TEs) used to apply the electric field, and Ti4+-doped V2O5 film (or V2O5 film) is the EC layer. The electrolyte was 0.5 mol/L lithium perchlorate in propylene carbonate. After making the electrical connections, the EC device was ready for testing. The area of the EC device was 2.0 cm × 2.0 cm.

2.6. Material Characterization

The films’ morphology was characterized using a field emission scanning electron microscope (Hitachi S-4800 FE-SEM, Ibaraki-ken, Japan). The thickness of the films was measured from their cross-sectional SEM images using ImageJ 1.8.0 software. For phase analysis, Raman spectroscopy was recorded at room temperature in the wavenumber range from 50 to 1200 cm−1 at an excitation wavelength of 532 nm (Thermo Scientific, Waltham, MA, USA). The spectra were generated with ~0.45 mW, 633 nm He-Ne laser excitation at the sample surface. Transmission electron microscope (TEM) images were obtained on a JEOL-2010F electron microscope operated at 200 kV.

2.7. Electrochemical and EC Measurements

Electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV) were carried out on an electrochemical workstation CHI600E (Chinstruments, Shanghai, China) using a three-electrode cell at open circuit voltage. EIS and CV measurements used the Ti4+-doped V2O5 film (or V2O5 film) deposited on ITO substrates as working electrodes. At the same time, a platinum grid served as a counter electrode, and a commercial Ag/AgCl 1 M KCl electrode served as a reference. A 0.5 mol/L LiClO4/propylene carbonate solution was used as an electrolyte. The test parameters of EIS were measured in the frequency range from 0.01 Hz to 1000 kHz at the open circuit voltage (OCV) after discharge/charge cycles of 5 mV amplitude. CV measurements were performed in the voltage range from −1.5 to 1.5 V at a cycling speed of 100 mV/s.
EC measurements on the fabricated EC devices were conducted by combining optical transmittance spectra using a UV-3600 spectrophotometer in the wavelength range of 190–1100 nm (Shimadzu, Tokyo, Japan) with chronoamperometry technology (CA) at an applied voltage from −3.0 V to +3.0 V, in increasing potential steps of 0.5 V kept for 30 s.
The corresponding change in optical density (ΔOD) was defined as the following [32]:
Δ OD = log ( T b l e a c h e d / T c o l o r e d )
Coloration switching time (τc) is the time required to reach 90% of the film’s complete transmittance change. The films’ coloration switching behavior was measured at wavelengths of 400 nm (τc400), while cycling stabilities were measured at a wavelength of 400 nm. A square wave voltage of +3.0 V to −3.0 V was alternately applied for 30 s for each state and was kept for 100 cycles.
Cycling stability was characterized by retention. Retention was calculated as the following [33]:
R e t e n t i o n = Δ T n Δ T 1
where Δ T 1 is the optical contrast of the 1st cycle, and Δ T n is the optical contrast of the nth cycle.
Coloration efficiency (η) was extracted as the slope of the line fitted to the linear region of the curve of OD versus the extracted charge density at −3.0 V of coloring potential at 400 nm [9] and expressed as the following [34]:
η = Δ O D q

3. Results and Discussion

3.1. Microstructural Characterization

Figure 1a shows the Raman spectra of V2O5 and Ti4+-doped V2O5 films. The distinctive Raman modes of α-V2O5 emerge at ~100, 142, 196, 281, 301, 405, 482, 526, 700, and 992 cm−1. The low-frequency modes at ~100, 142, and 196 cm−1 correspond to the relative motions of V2O5 layers (external modes) in V2O5 [8]. The two peaks at 142 and 196 cm−1 are intimately linked with the layered structure, proving long-range structural order [34]. The intermediate frequency peaks at ~281, 301, 405, 482, 526, and 700 cm−1 are related to the bending and stretching vibrations (internal modes) of the vanadium-oxide bond in V2O5 [35]. The peak at 281 cm−1 is attributed to the V = O bending vibration, peaks at 301 and 700 cm−1 correspond to the V3–O (triply coordinated oxygen) stretching mode, and peaks at 482 and 526 cm−1 correspond to the bending vibrations of the V–O–V (bridging doubly coordinated oxygen) bending vibration. The highest frequency peak at ~992 cm−1 corresponds to the stretching mode of the terminal oxygen (vanadyl oxygen, V = Oν) [36].
Despite that, the Raman modes of the anatase TiO2 phase, which should appear at ~145 and 198 cm−1, are close to those of α-V2O5 at 142 and 196 cm−1; the other Raman modes of the anatase TiO2 phase, at ~396, 515, and 640 cm−1, are not observed in the spectrum of the Ti4+-doped V2O5 film, demonstrating that there is no isolated TiO2 phase in the Ti4+-doped film. The doping process could be represented by Equation (4).
T i O 2 V 2 O 5 2 T i V + V O · · + 2 O O
In Figure 2, the SEM image of the undoped V2O5 film displays many tiny rod-like particles with ~200 nm length and ~100 nm diameter, aggregating to form a dense porous film (Figure 2a,e). In contrast, in the SEM image of the Ti4+-doped V2O5 film (Figure 2b), large nanowires are seen on top, as illustrated by green arrows, and long nanorods at the bottom, red arrows. The nanowires (top layer) have nearly the same diameter (~60 nm), with a length ranging from several hundred nanometers to tens of micrometers. Different from the rod-like particles in the undoped V2O5 film, the long nanorods on the bottom of the Ti4+-doped V2O5 film interlace to form a dense layer. Furthermore, the nanowires on the top seem to grow up from the bottom part with one end inserted within the bottom layer, indicating good interconnection between the top nanowires and the bottom long nanorods layer. The cross-sectional SEM images demonstrate that several nanowires grow from the bottom of undoped V2O5 film (see inset of Figure 2a), with the thickness of the bottom part at around 110 nm. After doping with Ti4+, the nanowires became dense, with a thickness of around 130 nm (inset of Figure 2b). Therefore, it is reasonable to deduce that the formation of nanowires comes from introducing Ti ions. Corresponding EDS results (Figure 2g) show that the atomic ratio Ti:V for the Ti4+-doped V2O5 film is 0.03:1.
Figure 3a,b shows the TEM images of a single Ti4+-doped V2O5 nanowire, whereas Figure 3c is the HRTEM image taken from the square frame shown in Figure 2a along with its diffraction pattern as an inset. The (101) interplanar spacing of Ti4+-doped V2O5 (Figure 3c) is enlarged (3.47 Å) in contrast to that of undoped α-V2O5, (3.20 Å), based on PDF#53-0538 [37]. Such an enlarged lattice could be attributed to the larger size of Ti4+ ions (0.61 Å) compared to V5+ ions (0.54 Å). Therefore, by combining Raman and SEM results, it was deduced that Ti4+ ions had been successfully inserted into the V2O5 lattice. Figure 3d shows the EDS retention mapping of all atoms in the Ti4+-doped V2O5 nanowire, and Figure 3e–g shows the atomic distribution of V, Ti, and O, respectively. Furthermore, this EDS mapping demonstrates that the Ti ions are uniformly distributed inside the Ti4+-doped V2O5 nanowire (Figure 3f). Additionally, the Ti:V ratio from the EDS spectrum obtained from TEM (Figure 3h) is 0.03:1.
EIS was used to understand and analyze the kinetic behavior and calculate the values of the circuit components of Li+ ion intercalation/deintercalation during cycling. The depressed semicircle in Nyquist plots (Figure 4a,b) observed in the high-frequency region is due to the charge transfer resistance (Rct). In contrast, the inclined line in the low-frequency region reflects Warburg impedance [38]. The films show a capacitive arc in the high-frequency region and diffusion in the low-frequency domain. The Rct values were determined by plot fitting using ZView2 software with the equivalent circuit depicted in Figure 4a,b insets. The slope of the inclined line in the low-frequency region after the semi-circle is related to Li-ion diffusion and is named KΩ, which is the ion diffusion ability. Generally, a high KΩ corresponds to fast ion diffusion or speed [14,39].
As seen in Figure 4c, Rct values decrease almost linearly with cycling for both undoped V2O5 and Ti4+-doped V2O5 films. The decrease is similar for both films since the slopes of Rct vs. the number of cycles are almost the same. The reduction in Rct could originate from vanadates, such as H2VO4, HVO42−, HV2O5, and vanadium-bronze species formed during Li-ion intercalation/deintercalation [19]. The formed vanadates dissolve into the electrolyte, leading to a decrease in the amount of V2O5 and therefore to a decrease in Rct. Further, the formed vanadium-bronze remains within the V2O5 lattice, decreasing Rct [40]. In addition, throughout the cycling process, the Ti4+-doped V2O5 film showed much smaller Rct values than those of the undoped V2O5 film (Figure 4c), indicating faster charge transfer in the Ti4+-doped V2O5 film. Such an observation is consistent with results from M. Panagopoulou et al., in which the charge transfer resistance decreases after doping Mg2+ into V2O5 [15]. However, the KΩ values for both samples are nearly the same during the cycling (Figure 4d), indicating that doping V2O5 with Ti ions does not improve the ion diffusion speed. Furthermore, a vast drop occurs from cycle 0 to cycle 10 in both samples, after which KΩ values remain constant, indicating that a dramatic structure change happens from the beginning of Li-ion intercalation/deintercalation; afterward, the intercalation/deintercalation process becomes dynamically balanced without further change upon cycling.

3.2. EC Properties

Figure 5a,b show the transmittance spectra of the EC device constructed using undoped and Ti4+-doped V2O5 films, respectively. After coloration, the transmittance of both samples increases in the visible region while it decreases in the NIR region. For instance, for the Ti4+-doped V2O5 film, the transmittance at 400 nm in the colored state (Tc) at −3.0 V is 35.90%, which is higher than that of the as-prepared state at 0 V (T0 = 17.8%) (Table 1), while at 900 nm, it is 56.30%, which is lower than the 71.80% at 0 V. On the other hand, the Ti4+-doped V2O5 film has better reversibility than the undoped V2O5 film by showing a spectrum at the bleached state (+3 V) closer to that of the as-prepared state (0 V).
Here, reversibility was calculated as the following:
Reversibility = 100 % × ( 1 T 0 T b T 0 )
where T0 and Tb are the transmittances at 0 V (as prepared state) and +3 V (bleached state).
The calculated reversibility at 400 nm for Ti4+-doped V2O5 film is 100%, noticeably better than that of the undoped V2O5 film, which agrees with the smaller Rct of the Ti4+-doped V2O5 film.
The absolute value of optical density (| Δ OD|) for Ti4+-doped V2O5 film is much higher than that of the undoped V2O5 film in the range from 320 to 1100 nm (Figure 5c). For instance, | Δ OD| for the Ti4+-doped V2O5 film is 0.30 at a wavelength of 400 nm, which is almost twice that of the undoped V2O5 film (0.19) (Table 1). Considering that the largest | Δ OD| occurs at a wavelength of ~400 nm for both samples, the following tests for the EC properties, including switching time, coloration efficiency, and cycling stability, were conducted at 400 nm to achieve a strong response.
Coloration efficiency (CE), which is defined as the change in optical density (OD) per unit of charge (Q) intercalated into or extracted from the EC film (see Equation (4)) [41,42], was found by calculating the slopes of ΔOD versus ΔQ at 400 nm, as seen in Figure 5d,e. CE is a practical parameter to measure the power requirements for the color-switching process, and a higher CE means better electronic utilization efficiency [31]. Coloration efficiency depends on the kind of EC material, amount of charge insertion, and voltage programs [31]. However, the physical basis for coloration is yet to be understood entirely [43,44]. In this work, the undoped V2O5 film has a CE of 23.57 cm2/C, close to a value reported by Liu et al. (24.5 cm2/C) [44]. After doping, the Ti4+-doped V2O5 film shows an improved CE of 34.15 cm2/C. This value is even higher than those reported in the literature, in which it is 4.7 cm2/C for V2O5-TiO2 coating with V/Ti (70/30) [25], 18.6 cm2/C for RF magnetron sputtered Mg2+-doped V2O5 film [15], and 24.12 cm2/C for 3.5:1 (V/W) composite W/WO3-V2O5 films [26]. Considering that both undoped and Ti4+-doped V2O5 films have close electrochemical capacities, as discussed above, such an excellent CE value (34.15 cm2/C) of Ti4+-doped V2O5 film could be attributed to the lower Rct, which renders faster Li+ intercalation/deintercalation and also faster charge transfer speed, and, therefore, better charge utilization efficiency than the undoped V2O5 film [19].
For the Ti4+-doped V2O5 film, coloration time (τc) first increases rapidly with cycling, peaking at 16.10 s for cycle 20 (Figure 6), after which it decreases, reaching ~9.00 s at cycle 100. For the undoped V2O5 film, τc also rapidly increases with cycling at almost the same rate as the Ti4+-doped V2O5 film until it reaches around cycle 10. From then on, its quick increase stops, and τc continues to grow at a somewhat constant rate until it reaches 13.16 s at cycle 100. At cycle 60, both samples have a close τc (12.19 s for the undoped V2O5 film and 12.65 s for the Ti4+-doped V2O5 film). After cycle 60, the Ti4+-doped V2O5 film shows shorter coloration times than the undoped V2O5 film.
Many factors can affect switching speed, including electronic conductivity, electrode materials, the underlying conductive layers, the ionic conductivity of the electrolyte, the morphology of the EC layer, the associated changes in ion diffusion within this morphology of the EC layer, and ion insertion kinetics [30,39]. Previous studies have shown the positive direct effect of the electronic conductivity of the electrode materials on coloration time [23,31]. However, in the present work, other factors influence coloration time because the Ti4+-doped V2O5 film does not show a shorter coloration time than the undoped V2O5 film during the cycling process despite its lower Rct and their close KΩ values.
Undoped and Ti4+-doped V2O5 films have different morphologies, as is seen in SEM images (Figure 2a,b). Apart from nanowires on the top layer, the Ti4+-doped V2O5 film has a dense bottom layer, while the undoped V2O5 film features a developed porous structure. The developed porous structure in the undoped film and its corresponding rough surface provide the advantages of quickly absorbing the electrolyte and favoring Li+ ion intercalation compared to the dense Ti4+-doped V2O5 film. Therefore, in the first five cycles, as potential was just applied to the samples, the undoped film shows a close coloration time to the Ti4+-doped film despite its lower Rct. Such an increase in both undoped and Ti4+-doped V2O5 films indicates that at the beginning of the EC process, an activation is needed to adjust the interaction between the electrolyte and electrode. After five cycles, the amount of Li+ ions intercalated almost reached saturation, combining a stable structure after the primary activation, and the coloration time for the undoped film became nearly leveled. As for the Ti4+-doped film, such a saturation process occurs at cycle 20, at which point the coloration time peaks. After cycle 20, the nanowire morphology on the top of the Ti4+-doped film reveals its importance by sharing a relatively high-strength electric field at its sharp edges [23], rendering a weaker electric field on the dense bottom layer. The polarization is significantly reduced since the high electric field was removed, facilitating Li+ ion intercalation. In addition to its reduced Rct with cycling, coloration time quickly decreases from cycle 20 to 100.
Cycling stability was investigated at 400 nm, defined as the number of cycles required for retention to decline to 60% of its value at cycle 1. In the first five cycles, retention of the undoped V2O5 film underwent a nearly vertical drop to 79.7% (Figure 7c) and then decreased slowly at a constant rate, reaching 43.6% at cycle 100. As for the Ti4+-doped V2O5 film, retention initially suffered a fast decrease, though not as steep as for the undoped V2O5 film. However, a turning point occurs at cycle 20, reaching the minimum retention of 77.39%, which coincides with the cycle at which coloration time is at a maximum (Figure 7c).
After cycle 20, retention levels off (~81%) with slight variations until cycle 100, for which retention is 82.6%. The experimental results we have obtained are significant. Our tested material outperforms those from the literature, which report 65% after 600 cycles for a 4 at.% Ti4+-doped V2O5 planar film [9] and 50% after 100 cycles for an 8.5 at.% Ti4+-doped V2O5 film [45], 76% after 100 cycles for a 5% at.% Ti4+-doped V2O5 film [46], 18.6% for a nanobelt-membrane hybrid-structured vanadium oxide film [21], and 24% for a self-organized multifunctional V2O5 nanofiber-liquid crystal polymer hybrid film [47].
In the present work, the fast fading in the first several cycles for both undoped and doped samples indicates that an important irreversible structural change and/or chemical dissolution primarily occurred at the beginning of cycling, which aligns with the continuous drop in Rct. In addition, faster fading for the undoped film compared to the Ti4+-doped film is observed at the beginning of the cycling process. As discussed, regarding coloration time, a highly developed porous structure plays a double-edged role throughout process: On one hand, it facilitates the electrolyte absorption and Li+ ion intercalation; on the other hand, such a rapid electrochemical process accelerates irreversible structural change and/or chemical dissolution. In addition to a larger Rct value of the undoped film than that of the Ti4+-doped film, the faster chemical dissolution leads to its faster fading [17]. Apart from this, a dark gray layer remained in the bleached state after cycle 1, as seen in Figure 7a, indicating a failed delithiation and the remaining V4+ compounds. These V4+ compounds slowly dissolve into the electrolyte during the cycling process because, at cycle 100, the dark layer becomes paler (Figure 7a). In contrast, there is no formation of a dark layer in Ti4+-doped V2O5 film in the bleaching state after cycle 1 (Figure 7b, bleaching) due to its low charge transfer resistance and reduced V4+ compound formation. A strong fading occurs in the bleaching state after the first cycle in both samples since their color became paler than the as-prepared samples, as seen in Figure 7. After 100 cycles, the Ti4+-doped V2O5 film kept nearly the same color as in cycle 1, while the undoped V2O5 film became paler than in cycle 1, consistent with the retention results (Figure 7c).
Apart from the low charge transfer resistance, the nanowires in the Ti4+-doped V2O5 film play a primary role in cycling stability since their nano edges reduce polarization through a shared high-strength electric field and, therefore, minimize those dissolvable intermediate products. The SEM images in Figure 2c,d further illustrate this observation. After 100 cycles, the undoped V2O5 film is transformed entirely and mostly dissolved, leaving only some scattered nanoparticles on the surface, highlighted with yellow circles in Figure 2c. In contrast, the Ti4+-doped V2O5 film kept its original morphology with nanowires on top and nanorods on the bottom layer. The corresponding EDS spectrum (Figure 3h) shows that, after 100 cycles, the atomic ratio Ti:V in the Ti4+-doped V2O5 film is around 0.02:1, close to that of its as-prepared state (0.03:1). In contrast, for the undoped V2O5 film, the content of V ions is too low to be detected in the EDS spectrum (highlighted with red circles in Figure 2g), indicating that most of the V2O5 was lost, resulting in correspondingly high fading upon cycling. Raman results (Figure 1b) show that after 100 cycles, both V2O5 and other vanadium oxide phases are formed in the Ti4+-doped V2O5 film, while there is no V2O5 phase found in what was originally the undoped V2O5 film. Doping V2O5 with a small amount of Ti ions transformed the morphology of the film’s top surface by forming a layer of long nanowires, which enhanced the conductivity of the film and improved its cycling stability.

4. Conclusions

Ti4+-doped V2O5 films were successfully fabricated by spin-coating, featuring nanowires on top and a dense long nanorod layer on the bottom. The nanowires show a uniform diameter of around ~60 nm, with lengths from several hundred nanometers to tens of micrometers. After doping, the (110) interplanar lattice spacing of the undoped V2O5 film enlarged to 3.47 Å from 3.20 Å. Upon EC cycling, a dynamic EIS characterization was used to analyze the EC properties, including coloration time, and cycling stability. During the cycling process of both undoped and doped samples, Rct decreased with cycling while KΩ rapidly dropped in the first 10 cycles and then leveled off, independent of cycling. However, the Ti4+-doped V2O5 film showed a much lower Rct than the undoped film. The introduction of Ti ions significantly influences EC properties in two ways. First, it lowered the Rct value; and second, it triggered the formation of nanowires. A small Rct value, corresponding to fast charge transfer speeds, delivers high CE, fast coloration time, and, most importantly, better cycling stability. The nanowires, taking advantage of their sharp nano-edges to share a relatively high-strength electric field, improve coloration speed and cycling stability by minimizing polarization and the corresponding dissolution of the formed intermediate phases. With a CE of 34.15 cm2/C, a coloration time of 9.00 s at cycle 100, and a cyclic retention of 82.6%, the thus obtained Ti4+-doped V2O5 film showed better EC properties than the undoped V2O5 film, for which those values are 23.57 cm2/C, 13.16 s, and 43.6%. The paper’s significance lies in the new insight into the effect of morphological and structural properties.

Author Contributions

Conceptualization, H.L. and J.L. (Jun Liao); methodology, H.L. and Y.D. (Yufei Deng); validation, Y.D. (Yufei Deng) and M.H.; investigation, Y.D. (Yufei Deng), H.L., J.L. (Jian Liang), J.L. (Jun Liao), M.H., Y.L. and R.C.; data curation, Y.D. (Yufei Deng), H.L., J.L. (Jian Liang), J.L. (Jun Liao), Y.L., R.C., J.R. and Y.D. (Yahia Djaoued); writing—original draft preparation, Y.D. (Yahia Djaoued) and H.L.; writing—review and editing, J.L. (Jian Liang), J.R. and Y.D. (Yahia Djaoued); project administration, H.L. and Y.D. (Yahia Djaoued); funding acquisition, H.L., J.L. (Jian Liang), and Y.D. (Yahia Djaoued). All authors have read and agreed to the published version of the manuscript.

Funding

The financial support from the National Sciences and Engineering Research Council (NSERC) of Canada (grant #2023-05942), National Natural Science Foundation of China (Grant #22365020 and grant #52362004), the John R. Evans Leaders Fund (Canada Foundation for Innovation) (grant #27741), and Jingdezhen Science and Technology Development Funds (20202GYZD013-11) is gratefully acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yuana, S.; Duana, X.; Liua, J.; Yun Ye, Y.; Lva, F.; Liu, T.; Wang, Q.; Zhang, X. Recent progress on transition metal oxides as advanced materials for energy conversion and storage. Energy Storage Mater. 2021, 42, 317–369. [Google Scholar] [CrossRef]
  2. Okpara, C.; Olatunde, O.C.; Wojuola, O.B.; Onwudiwe, D.C. Applications of transition metal oxides and chalcogenides and their composites in water treatment: A review. Environ. Adv. 2023, 11, 100341. [Google Scholar] [CrossRef]
  3. Ishibe, T.; Tomeda, A.; Watanabe, K.; Kamakura, Y.; Mori, N.; Naruse, N.; Mera, Y.; Yamashita, Y.; Nakamura, Y. Methodology of thermoelectric power factor enhancement by controlling nanowire interface. ACS Appl. Mater. Interfaces 2018, 10, 37709–37716. [Google Scholar] [CrossRef]
  4. Maeng, S.; Kim, S.-W.; Lee, D.-H.; Moon, S.-E.; Kim, K.-C.; Maiti, A. SnO2 nanoslab as NO2 sensor: Identification of the NO2 sensing mechanism on a SnO2 surface. ACS Appl. Mater. Interfaces 2014, 6, 357–363. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, C.; Xie, Y. Promising vanadium oxide and hydroxide nanostructures: From energy storage to energy saving. Energy Environ. Sci. 2010, 3, 1191–1206. [Google Scholar] [CrossRef]
  6. Liu, P.; Zhu, K.; Gao, Y.; Luo, H.; Lu, L. Recent progress in the applications of vanadium-based oxides on energy storage: From low-dimensional nanomaterials synthesis to 3D micro/nano-structures and free-standing electrodes fabrication. Adv. Energy Mater. 2017, 7, 1700547. [Google Scholar] [CrossRef]
  7. Zhang, Y.; Zheng, J.; Zhao, Y.; Hu, T.; Gao, Z.; Meng, C. Fabrication of V2O5 with various morphologies for high-performance electrochemical capacitor. Appl. Surf. Sci. 2016, 377, 385–393. [Google Scholar] [CrossRef]
  8. Li, H.; Liu, Y.; Jiang, W.; Liang, J.; Tang, Z.; Hu, M.; Robichaud, J.; Djaoued, Y. 2D vanadium oxide inverse opal films: Cycling stability exploration as an electrochromic active electrode. J. Mater. Chem. C 2023, 11, 923–993. [Google Scholar] [CrossRef]
  9. Lu, Y.; Liu, L.; Mandler, D.; Lee, P.S. High switching speed and coloration efficiency of titanium-doped vanadium oxide thin film electrochromic devices. J. Mater. Chem. C 2013, 1, 7380–7386. [Google Scholar] [CrossRef]
  10. Qi, Y.; Qin, K.; Zou, Y.; Lin, L.; Jian, Z.; Chen, W. Flexible electrochromic thin films with ultrafast responsion based on exfoliated V2O5 nanosheets/graphene oxide via layer-by-layer assembly. Appl. Surf. Sci. 2020, 514, 145950. [Google Scholar] [CrossRef]
  11. Wu, Y.; Gao, G.; Yang, H.; Bi, W.; Liang, X.; Zhang, Y.; Zhang, G.; Wu, G. Controlled synthesis of V2O5/MWCNT core/shell hybrid aerogels through a mixed growth and self-assembly methodology for supercapacitors with high capacitance and ultralong cycle life. J. Mater. Chem. A 2015, 3, 15692–15699. [Google Scholar] [CrossRef]
  12. Takahashi, K.; Wang, Y.; Cao, G.Z. Ni−V2O5∙nH2O core−shell nanocable arrays for enhanced electrochemical intercalation. J. Phys. Chem. B 2005, 109, 48–51. [Google Scholar] [CrossRef] [PubMed]
  13. Lantelme, F.; Mantoux, A.; Groult, H.D. Electrochemical study of phase transition processes in lithium insertion in V2O5 electrodes. J. Electrochem. Soc. 2003, 150, A1202. [Google Scholar] [CrossRef]
  14. Yu, M.; Zeng, Y.; Han, Y.; Cheng, X.; Zhao, W.; Liang, C.; Tong, Y.; Tang, H.; Lu, X. Valence-optimized vanadium oxide supercapacitor electrodes exhibit ultrahigh capacitance and super-long cyclic durability of 100000 cycles. Adv. Funct. Mater. 2015, 25, 3534–3540. [Google Scholar] [CrossRef]
  15. Panagopoulou, M.; Vernardou, D.; Koudoumas, E. Tunable properties of Mg-doped V2O5 thin films for energy applications: Li-ion batteries and electrochromics. J. Phys. Chem. C 2017, 121, 70–79. [Google Scholar] [CrossRef]
  16. Wu, Y.J.; Gao, G.H.; Wu, G.M. Self-assembled three-dimensional hierarchical porous V2O5/graphene hybrid aerogels for supercapacitors with high energy density and long cycle life. J. Mater. Chem. A 2015, 3, 1828–1832. [Google Scholar] [CrossRef]
  17. Chao, D.; Xia, X.; Liu, J.; Fan, Z.; Ng, C.F.; Lin, J.; Zhang, H.; Shen, Z.X.; Fan, H.J. A V2O5/conductive-polymer core/shell nanobelt array on three-dimensional graphite foam: A high-rate, ultrastable, and freestanding cathode for lithium-ion batteries. Adv. Mater. 2014, 26, 5794–5800. [Google Scholar] [CrossRef]
  18. Xiong, C.; Aliev, A.E.; Gnade, B.; Balkus, K.J. Fabrication of silver vanadium oxide and V2O5 nanowires for electrochromics. ACS Nano 2008, 2, 293–301. [Google Scholar] [CrossRef]
  19. Luo, Y.; Bai, Y.; Mistry, A.; Zhang, Y.; Zhao, D.; Sarkar, S.; Hand, J.V.; Rezaei, S.; Chuang, A.C.; Carrillo, L.; et al. Effect of crystallite geometries on electrochemical performance of porous intercalation electrodes by multiscale operando investigation. Nat. Mater. 2022, 21, 217–227. [Google Scholar] [CrossRef]
  20. Tong, Z.; Li, N.; Lv, H.; Tian, Y.; Qu, H.; Zhang, X.; Li Zhao, J.Y. Annealing synthesis of coralline V2O5 nanorod architecture for multicolor energy-efficient electrochromic device. Sol. Energy Mater. Sol. Cells 2016, 146, 135–143. [Google Scholar] [CrossRef]
  21. Kang, W.; Yan, C.; Wang, X.; Foo, C.; Tan, A.; Chee, K.; Lee, P. Green synthesis of nanobelt-membrane hybrid structured vanadium oxide with high electrochromic contrast. J. Mater. Chem. C 2014, 2, 4727–4732. [Google Scholar] [CrossRef]
  22. Li, G.; Qiu, Y.; Hou, Y.; Li, H.; Zhou, L.; Deng, H.; Zhang, Y. Synthesis of V2O5 hierarchical structures for long cycle-life lithium-ion storage. J. Mater. Chem. A 2015, 3, 1103–1109. [Google Scholar] [CrossRef]
  23. Tong, Z.; Hao, J.; Zhang, K.; Zhao, J.; Su, B.-L.; Li, Y. Improved electrochromic performance and lithium diffusion coefficient in three-dimensionally ordered macroporous V2O5 films. J. Mater. Chem. C 2014, 2, 3651–3658. [Google Scholar] [CrossRef]
  24. Wei, D.; Scherer, M.R.J.; Bower, C.; Andrew, P.; Ryhanen, T.; Steiner, U. A nanostructured electrochromic supercapacitor. Nano Lett. 2012, 12, 1857–1862. [Google Scholar] [CrossRef] [PubMed]
  25. Kim, S.; Taya, M. Electrochromic windows based on V2O5–TiO2 and poly (3,3-dimethyl-3,4-dihydro-2H-thieno[3,4-b][1,4]dioxepine) coatings. Sol. Energy Mater. Sol. Cells 2012, 107, 225–229. [Google Scholar] [CrossRef]
  26. Panagopoulou, M.; Vernardou, D.; Koudoumas, E.; Tsoukalas, D.; Raptis, Y.S. Tungsten doping effect on V2O5 thin film electrochromic performance. Electrochim. Acta 2019, 321, 134743. [Google Scholar] [CrossRef]
  27. Yao, J.H.; Yin, Z.L.; Zou, Z.G.; Li, Y.W. Y-doped V2O5 with enhanced lithium storage performance. RSC Adv. 2017, 7, 32327–32335. [Google Scholar] [CrossRef]
  28. Yu, H.; Rui, X.; Tan, H.; Chen, J.; Huang, X.; Xu, C.; Liu, W.; Denis, Y.; Hng, H.H.; Hoster, H.E. Cu doped V2O5 flowers as cathode material for high-performance lithium ion batteries. Nanoscale 2013, 5, 4937–4943. [Google Scholar] [CrossRef]
  29. Li, Z.; Zhang, C.; Liu, C.; Fu, H.; Nan, X.; Wang, K.; Li, X.; Ma, W.; Lu, X.; Cao, G. Enhanced electrochemical properties of Sn-doped V2O5 as a cathode material for lithium ion batteries. Electrochim. Acta 2016, 222, 1831–1838. [Google Scholar] [CrossRef]
  30. Thakur, V.K.; Ding, G.; Ma, J.; Lee, P.S.; Lu, X. Hybrid materials and polymer electrolytes for electrochromic device applications. Adv. Mater. 2012, 24, 4071–4096. [Google Scholar] [CrossRef]
  31. Yang, G.; Zhang, Y.-M.; Cai, Y.; Yang, B.; Gu, C.; Zhang, S.X.-A. Advances in nanomaterials for electrochromic devices. Chem. Soc. Rev. 2020, 49, 8687–8720. [Google Scholar] [CrossRef] [PubMed]
  32. Gong, H.; Li, W.; Fu, G.; Zhang, Q.; Liu, J.; Jin, Y.; Wang, H. Recent progress and advances in electrochromic devices exhibiting infrared modulation. J. Mater. Chem. A 2022, 10, 6269–6290. [Google Scholar] [CrossRef]
  33. Li, H.; Liao, J.; Liu, Y.; Deng, Y.; Liang, J.; Tang, Z.; Liu, F.; Robichaud, J.; Djaoued, Y. Impact of impedance on electrochromic properties of W-doped V2O5 films. Next Mater. 2024, 2, 100149. [Google Scholar] [CrossRef]
  34. Li, H.; Tang, Z.; Liu, Y.; Robichaud, J.; Liang, J.; Jiang, W.; Djaoued, Y. Two-Dimensional V2O5 Inverse Opal: Fabrication and Electrochromic Application. Materials 2022, 15, 2904. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, S.; Zhu, H.; Zhang, B.; Li, J.; Zhu, H.; Ren, Y.; Geng, H.; Yang, Y.; Liu, Q.; Li, C. Tuning the kinetics of zinc-ion insertion/extraction in V2O5 by in situ polyaniline intercalation enables improved aqueous zinc-ion storage performance. Adv. Mater. 2020, 32, 2001113. [Google Scholar] [CrossRef]
  36. Su, D.; Zhao, Y.; Yan, D.; Ding, C.; Ning, M.; Zhang, J. Enhanced composites of V2O5 nanowires decorating on graphene layers as ideal cathode materials for lithium-ion batteries. J. Alloys Compd. 2017, 695, 2974–2980. [Google Scholar] [CrossRef]
  37. Li, Y.W.; Yao, J.H.; Liu, C.J.; Zhao, W.M.; Deng, W.X.; Zhong, S.K. Effect of interlayer anions on the electrochemical performance of Al-substituted α-type nickel hydroxide electrodes. Int. J. Hydrogen Energy 2010, 35, 2539–2545. [Google Scholar] [CrossRef]
  38. Xia, X.; Tu, J.; Mai, Y.; Chen, R.; Wang, X.; Gu, C.; Zhao, X. Graphene sheet/porous NiO hybrid film for supercapacitor applications. Chem. Eur. J. 2011, 17, 10898–10905. [Google Scholar] [CrossRef]
  39. 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. Mat. Sol. C 2013, 117, 231–238. [Google Scholar] [CrossRef]
  40. Runnerstrom, E.L.; Llordés, A.; Lounis, S.; Milliron, D.J. Nanostructured electrochromic smart windows: Traditional materials and NIR-selective plasmonic nanocrystals. Chem. Commun. 2014, 50, 10555–10572. [Google Scholar] [CrossRef]
  41. Gillaspie, D.T.; Tenent, R.C.; Dillon, A.C. Metal-oxide films for electrochromic applications: Present technology and future directions. J. Mater. Chem. 2010, 20, 9585–9592. [Google Scholar] [CrossRef]
  42. Deb, S.K. Opportunities and challenges in science and technology of WO3 for electrochromic and related applications. Sol. Energy Mater. Sol. Cells 2008, 92, 245–258. [Google Scholar] [CrossRef]
  43. Liu, Y.N.; Jia, C.Y.; Wan, Z.Q.; Weng, X.; Xie, J.; Deng, L. Electrochemical and electrochromic properties of novel nanoporous NiO/V2O5 hybrid film. Sol. Energy Mater. Sol. Cells 2015, 132, 467–475. [Google Scholar] [CrossRef]
  44. Tritschler, U.; Beck, F.; Schlaad, H.; Cölfen, H. Electrochromic properties of self-organized multifunctional V2O5–polymer hybrid films. J. Mater. Chem. C 2015, 3, 950–954. [Google Scholar] [CrossRef]
  45. Salek, G.; Bellanger, B.; Mjejri, I.; Gaudon, M.; Rougier, A. Polyol synthesis of Ti-V2O5 nanoparticles and their use as electrochromic films. Inorg. Chem. 2016, 55, 9838–9847. [Google Scholar] [CrossRef] [PubMed]
  46. Wu, I.-H.; Chandrasekar, A.; Arul, K.T.; Huang, Y.-C.; Nga, T.T.T.; Chen, C.-L.; Chen, J.-L.; Wei, D.-H.; Asokan, K.; Yeh, P.-H.; et al. Improving electrochromic properties of V2O5 smart film through Ti incorporation: Local atomic and electronic perspectives. Opt. Mater. X 2024, 22, 100301. [Google Scholar] [CrossRef]
  47. Lin, Y.-S.; Tsai, C.-W.; Chen, P.-W. Electrochromic properties of V2O5-z thin films sputtered onto flexible PET/ITO substrates. Solid State Ion. 2008, 179, 290–297. [Google Scholar] [CrossRef]
Figure 1. Raman spectra for V2O5 and Ti4+-doped V2O5 films: (a) before cycling and (b) after cycle 100.
Figure 1. Raman spectra for V2O5 and Ti4+-doped V2O5 films: (a) before cycling and (b) after cycle 100.
Materials 17 04680 g001
Figure 2. SEM images of: (a,c,e) V2O5 film; (b,d) Ti4+-doped V2O5 film (nanowires (top) indicated with green arrows and the nanorod layer (bottom) with red arrows). (f,g) Corresponding EDS results for the films in the as-prepared state (a,b,e) and after 100 EC cycles (c,d). Inset in (a,b) showing the corresponding thickness of the films. Scale bars in Figures (a,b) 1 μm, in Figures (c,d) 500 nm, in Figure (e) 200 nm, and insets in Figures (a,b) 500 nm. Yellow circles in Figure (c) indicate nanoparticles left after cycling, and red circles in Figures (f,g) highlight the content changes of V ions before (black spectrum) and after (blue spectrum) cycling.
Figure 2. SEM images of: (a,c,e) V2O5 film; (b,d) Ti4+-doped V2O5 film (nanowires (top) indicated with green arrows and the nanorod layer (bottom) with red arrows). (f,g) Corresponding EDS results for the films in the as-prepared state (a,b,e) and after 100 EC cycles (c,d). Inset in (a,b) showing the corresponding thickness of the films. Scale bars in Figures (a,b) 1 μm, in Figures (c,d) 500 nm, in Figure (e) 200 nm, and insets in Figures (a,b) 500 nm. Yellow circles in Figure (c) indicate nanoparticles left after cycling, and red circles in Figures (f,g) highlight the content changes of V ions before (black spectrum) and after (blue spectrum) cycling.
Materials 17 04680 g002
Figure 3. TEM images (a,b) of a single Ti4+-doped V2O5 nanowire. (c) HRTEM image taken from the square frame in (a) and its diffraction pattern (inset). (d) Retention mapping showing all atoms. Atomic distribution of (e) V, (f) Ti, and (g) O. (h) Corresponding EDS result for the films.
Figure 3. TEM images (a,b) of a single Ti4+-doped V2O5 nanowire. (c) HRTEM image taken from the square frame in (a) and its diffraction pattern (inset). (d) Retention mapping showing all atoms. Atomic distribution of (e) V, (f) Ti, and (g) O. (h) Corresponding EDS result for the films.
Materials 17 04680 g003
Figure 4. Nyquist plots of V2O5 film (a) and Ti4+-doped V2O5 film (b) at cycle 0 (black), cycle 20 (red), cycle 60 (green), and cycle 100 (blue), with equivalent circuits shown as insets. Evolution of Rct (c) and KΩ (d) as a function of cycling for V2O5 film (square, black) and Ti4+-doped V2O5 film (triangle, red).
Figure 4. Nyquist plots of V2O5 film (a) and Ti4+-doped V2O5 film (b) at cycle 0 (black), cycle 20 (red), cycle 60 (green), and cycle 100 (blue), with equivalent circuits shown as insets. Evolution of Rct (c) and KΩ (d) as a function of cycling for V2O5 film (square, black) and Ti4+-doped V2O5 film (triangle, red).
Materials 17 04680 g004
Figure 5. Transmittance spectra of the EC device constructed using (a) the undoped V2O5 film and (b) the Ti4+-doped V2O5 film in the as-prepared state 0 V (black, triangle), at a coloration potential of −3.0 V (red, sphere), and a bleaching potential of +3.0 V (green, square); (c) optical density; (d,e) coloration efficiency taken from the fifth cycle, in which the red line indicates the line fitted to the linear region of the curve (green) of OD versus ΔQ.
Figure 5. Transmittance spectra of the EC device constructed using (a) the undoped V2O5 film and (b) the Ti4+-doped V2O5 film in the as-prepared state 0 V (black, triangle), at a coloration potential of −3.0 V (red, sphere), and a bleaching potential of +3.0 V (green, square); (c) optical density; (d,e) coloration efficiency taken from the fifth cycle, in which the red line indicates the line fitted to the linear region of the curve (green) of OD versus ΔQ.
Materials 17 04680 g005
Figure 6. (a) Coloration switching time of V2O5 film (top) and Ti4+-doped V2O5 film (bottom) at a wavelength of 400 nm after 5, 20, 60, and 100 cycles; (b) Evolution of the coloration time as a function of cycling for undoped (black trace) and Ti4+-doped V2O5 films (red trace).
Figure 6. (a) Coloration switching time of V2O5 film (top) and Ti4+-doped V2O5 film (bottom) at a wavelength of 400 nm after 5, 20, 60, and 100 cycles; (b) Evolution of the coloration time as a function of cycling for undoped (black trace) and Ti4+-doped V2O5 films (red trace).
Materials 17 04680 g006
Figure 7. Transmittance measured at 400 nm as a function of time (a) for undoped V2O5; (b) for Ti4+-doped V2O5 films. Insets in (a,b): optical photos of the as-prepared state (top), cycle 1 (middle), and cycle 100 (bottom) for V2O5 and Ti4+-doped V2O5 films at coloration and bleaching states. (c) The evolution of retention as a function of cycling for undoped V2O5 (black squares) and Ti4+-doped V2O5 films (red circles).
Figure 7. Transmittance measured at 400 nm as a function of time (a) for undoped V2O5; (b) for Ti4+-doped V2O5 films. Insets in (a,b): optical photos of the as-prepared state (top), cycle 1 (middle), and cycle 100 (bottom) for V2O5 and Ti4+-doped V2O5 films at coloration and bleaching states. (c) The evolution of retention as a function of cycling for undoped V2O5 (black squares) and Ti4+-doped V2O5 films (red circles).
Materials 17 04680 g007
Table 1. EC optical modulation from the 5th cycle at different voltages of undoped and Ti4+-doped V2O5 films.
Table 1. EC optical modulation from the 5th cycle at different voltages of undoped and Ti4+-doped V2O5 films.
400 nm
SamplesT0 (%) ¥Tc (%) ¥Tb (%) ¥ΔT (%) * Δ O D  $Reversibility & (%)CE (cm2/C)
Undoped V2O5 film28.0045.5029.70−15.800.1993.9323.57
Ti4+-doped V2O5 film17.8035.9017.80−18.100.30100.0034.15
¥ T0, Tc, Tb: transmittance at 0 V, −3 V, and +3 V, respectively; * ΔT (%) = Tb − Tc; $  Δ O D : absolute value of Δ O D ; & reversibility (%) = 100 × (1− T 0 T b / T 0 ).
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

Deng, Y.; Li, H.; Liang, J.; Liao, J.; Huang, M.; Chen, R.; Long, Y.; Robichaud, J.; Djaoued, Y. Excellent Electrochromic Properties of Ti4+-Induced Nanowires V2O5 Films. Materials 2024, 17, 4680. https://doi.org/10.3390/ma17194680

AMA Style

Deng Y, Li H, Liang J, Liao J, Huang M, Chen R, Long Y, Robichaud J, Djaoued Y. Excellent Electrochromic Properties of Ti4+-Induced Nanowires V2O5 Films. Materials. 2024; 17(19):4680. https://doi.org/10.3390/ma17194680

Chicago/Turabian Style

Deng, Yufei, Hua Li, Jian Liang, Jun Liao, Min Huang, Rui Chen, Yinggui Long, Jacques Robichaud, and Yahia Djaoued. 2024. "Excellent Electrochromic Properties of Ti4+-Induced Nanowires V2O5 Films" Materials 17, no. 19: 4680. https://doi.org/10.3390/ma17194680

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