Skip to content
BY 4.0 license Open Access Published by De Gruyter November 20, 2020

Nanostructured inorganic electrochromic materials for light applications

  • Wu Zhang ORCID logo , Haizeng Li ORCID logo EMAIL logo , Eric Hopmann and Abdulhakem Y. Elezzabi EMAIL logo
From the journal Nanophotonics

Abstract

Electrochromism, an emerging energy conversion technology, has attracted immense interest due to its various applications including bistable displays, optical filters, variable optical attenuators, optical switches, and energy-efficient smart windows. Currently, the major drawback for the development of electrochromism is the slow switching speed, especially in inorganic electrochromic materials. The slow switching speed is mainly attributed to slow reaction kinetics of the dense inorganic electrochromic films. As such, an efficient design of nanostructured electrochromic materials is a key strategy to attain a rapid switching speed for their real-world applications. In this review article, we summarize the classifications of electrochromic materials, including inorganic materials (e.g., transition metal oxides, Prussian blue, and polyoxometalates), organic materials (e.g., polymers, covalent organic frameworks, and viologens), inorganic-organic hybrids, and plasmonic materials. We also discuss the electrochromic properties and synthesis methods for various nanostructured inorganic electrochromic materials depending on structure/morphology engineering, doping techniques, and crystal phase design. Finally, we outline the major challenges to be solved and discuss the outlooks and our perspectives for the development of high-performance nanostructured electrochromic materials.

1 Introduction

Electrochromism is the phenomenon where the optical transmittance or absorbance of a material changes under an applied electric potential [1]. When a voltage is applied to an electrochromic material, the optical properties of the material are altered in a reversible fashion. Owing to the optical switching behavior, electrochromic materials provide great opportunities for a variety of energy-saving and color-tuning applications, including smart windows, multicolor bistable displays, and color-tunable optical elements [2], [3], [4], [5], [6], [7], [8].

Generally, electrochromic materials can be either organic or inorganic materials [9]. Among the organic electrochromic materials, organic molecules [10], polymers [11], [12], [13], and covalent organic frameworks [14] are widely investigated. These classes of organic electrochromic materials are characterized by their rich color palette, fast response, and ease of fabrication via solution-processing [15]. However, in comparison with the inorganic electrochromic counterparts, the organic electrochromic materials exhibit inferior thermal and chemical stabilities [16], [ 17], thus limiting their practical use. Recently, inorganic electrochromic materials have emerged at the forefront for real-world applications and potential commercialization. This class of materials offers key advantages, such as long-term durability, wide range of working temperature, and good chemical stability [18], [19], [20]. However, despite these advantages, they suffer from the shortcoming of poor switching speed [21], [ 22]. This is due to the low electrical conductivity of the inorganic material, the high-volume expansion during cycling, and the slow ionic transporting efficiency within the bulk material [23], [24], [25]. Reducing the size of the material is regarded to be an efficient strategy to enhance the dynamics of the inorganic material’s response and improve its switching dynamics.

In recent years, nanostructured electrochromic materials have been investigated as a means to improve the switching speed because the nanostructured surface provides numerous active sites and shortens the ion’s diffusion paths [26]. Moreover, the volume expansion of nanostructured electrochromic materials during the switching process is insignificant as the numerous surface active sites increase the capacitive contribution during the redox reactions [27]. In other words, the ionic diffusion pathways inside the nanostructured materials are greatly shortened, which, in turn, improve the electrical conductivity [28]. Within this context, a nanostructured electrochromic material is expected to attain a rapid switching speed.

Although many literature reviews focus more on multifunctional electrochromic devices [29], [30], [31], [32], [33], a comprehensive review of recent achievements on nanostructured electrochromic materials is still lacking. In this article, we systematically discuss recent progress on designing nanostructured inorganic materials for high-performance electrochromic applications. This article first introduces the general classification of electrochromic materials and the key metrics used in evaluating electrochromic properties. This is followed by in-depth design strategies of nanostructured inorganic electrochromic materials and correlations between the electrochromic properties and synthetic methods for different nanostructured electrochromic materials. Next, the article highlights the challenges that need to be addressed in the future design of photonic devices. Finally, the authors’ perspective and outlook to stimulate the development of electrochromic technology are presented.

2 Classification of electrochromic materials

As the core component in an electrochromic device, electrochromic materials, having different optical properties, have been utilized for various electrochromic applications. Correspondingly, the electrochromic materials can be classified into a variety of types, including inorganic electrochromic materials, organic electrochromic materials, inorganic-organic hybrids, and plasmonic materials.

2.1 Inorganic electrochromic materials

Inorganic electrochromic materials mainly consist of the oxides of transition metals, which exhibit several different valence states on reduction. Electron delocalization between the mixed-valence states results in different bandgaps of inorganic materials, thus leading to different absorption of the light and inducing distinct color changes of the material [34]. There are two principally different types of electrochromic transition metal oxides which are classified in accordance with the coloration mechanisms: cathodic electrochromic materials, which are colored under guest ion insertion (reduction process); and anodic electrochromic materials, which are colored under guest cation extraction (oxidation process) [33]. The transition metals whose oxides possessing electrochromic properties are shown in the periodic table of elements (Figure 1).

Figure 1: 
Electrochromic transition metal oxides showing both cathodic and anodic coloration. Reproduced with permission from Ref. [35]. Copyright 2017, Elsevier Ltd.
Figure 1:

Electrochromic transition metal oxides showing both cathodic and anodic coloration. Reproduced with permission from Ref. [35]. Copyright 2017, Elsevier Ltd.

As can be inferred from Figure 1, WO3, TiO2, Nb2O5, MoO3, and Ta2O5 exhibit the cathodic coloration effect, which means that these oxides can be tinted via a reduction or cation intercalation [36], [37], [38], [39], [40], [41]. These oxides commonly consist of highly distorted MO6 octahedra crystalline structure [42], [ 43], where M represents the transition metal atom. The layered structure, formed by the edge- and corner-sharing MO6 octahedra, facilitates ion transport through the conduits or chains of interstitial sites [44]. Notably, these oxides exhibit similar electronic band structures. Their intrinsically empty d-bands become populated on cathodic charge injection, and thus cause a color change through new intraband transitions [45], [ 46]. Among this class of oxides, WO3 is the most widely studied cathodic electrochromic material, especially used for electrochromic smart windows platforms. The WO3 layer is transparent in the oxidation state, whereas it can be tinted to blue as cations injected into the WO3 structures via the following reaction:

(1) WO 3 ( bleached ) + xe + xM + M x WO 3 ( coloured )

where M+ can be either H+ or an alkali metal ion, such as Li+, Na+, K+, Al3+, or the recently reported Zn2+ [3]. Compared with used monovalent ions (e.g., H+, Li+, K+ and Na+), multivalent ions (e.g., Mg2+, Al3+, Zn2+) are expected to provide multiple charges to accelerate the redox reactions for fast electrochromic switching times [47], [ 48]. Another key advantage of multivalent ions, especially the Zn2+, is their compatibility with aqueous electrolytes, which facilitates the formation of newly developed Zn-based electrochromic devices [2].

On the other hand, the anodic coloration effect is observed in NiO, IrO2, Cr2O5, MnO2, FeO2, and Co3O4 [49], [50], [51], [52], [53], [54], [55]. Such anodic electrochromic oxides usually serve as a counter electrode which is often used in conjunction with a cathodic electrochromic electrode in a complementary device [56], [ 57]. The most widely studied anodic electrochromic material is NiO, which is transparent in its reduced state and colors gray in its oxidized state [58]. The schematic reaction of NiOx cycled in the Li+-ion electrolytes can be summarized as follows [59], [ 60]:

(2) NiO x + α e + α Li + Li α NiO x

(3) Li α NiO x ( bleached ) γ Li + + γ e + Li β NiO x ( colored ) ( α = β + γ ) .

The irreversible reaction (i.e., Eq. 2) refers to an activation process in the initial cycling of NiOx. The reversible reaction, represented by Eq. (3), indicates the color switching processes accompanying the ion insertion/extraction. The insertion/extraction of OH can also lead to the oxidation/reduction of nickel oxide owing to the transformation of nickel ions between Ni2+ and Ni3+ [61]. This process can be summarized as follows [62]:

(4) NiO ( bleched ) e + OH NiOOH ( colored )

In addition to the aforementioned cathodic and anodic electrochromic materials, vanadium oxides (e.g., V2O3, V3O7, and V2O5) are considered to be a type of bifunctional electrochromic material. This is due to their multicolor change properties under both anodically and cathodically potentials [63].

Along with the aforementioned transition metal oxides, Prussian blue (PB), represented as (Fe4[Fe(CN)6]3·nH2O) and its analogs (e.g., KFe[Fe(CN)6] and Na2Fe[Fe(CN)6]) is another important type of inorganic electrochromic material. PB is a type of polynuclear transition metal hexacyanometallates. Similar to the anodic electrochromic transition metal oxides, PB exhibits a blue color in its oxidized state and can be reduced to Prussian white (colorless) via electrochemical redox reactions [64], [65], [66], [67]. Recent investigations of PB-based electrochromic devices focused on the Zn-PB electrochromic devices because they show excellent energy retrieval functionalities [68], [69], [70]. Here, a Zn anode was used to reduce PB in a K+-Zn2+ dual-ion electrolyte and the constructed device was capable of self-bleaching via lighting an LED. Kang et al. deposited quasi-cubic PB particles with diameters of ≈450 nm on top of a fluorine-doped tin oxide (FTO)-coated glass substrate (Figure 2A) [68]. The Zn-PB electrochromic device achieved a high optical modulation of 84.9% at 633 nm during the voltage variation from 1.9 to 0.5 V (Figure 2B and C). This Zn-PB device also exhibited excellent cycling stability for up to 2000 cycles. Another example of the Zn-PB electrochromic device was recently reported by Li and Elezzabi [69]. Figure 2D and E depict an ideal unit cell of PB with guest ions and the working principle of Zn-PB electrochromic device. The Zn-PB electrochromic device exhibited an optical contrast of 78.6% between coloration/bleaching states (Figure 2F). Moreover, the Zn-PB electrochromic device possessed an open-circuit potential of 1.46 V (Figure 2F, inset) in the fully charged state, thus enabling self-bleaching behavior when lighting an LED. This open-circuit potential arises from the redox potential difference between the Zn foil and the PB electrode, thus providing the driving force that activates the oxidation of Zn and the reduction of the PB film. The built-in voltage allows the device to switch its color from blue to transparent while powering an LED for more than 60 min (Figure 2G). The feature of PB that stands out, in comparison with transition metal oxide–based electrochromic materials, is its fast color response that resulting from the PB’s open framework structures [71], [ 72].

Figure 2: 
Structure, working principle, and application of Zn-PB electrochromic device.
(A) Scanning electron microscope (SEM) image of the PB thin films. (B) Transmittance spectra of the Zn-PB device at different discharge voltages from 0.5 to 1.9 V. (C) Photographs of the Zn-PB electrochromic device at different discharge voltages. (A–C) Reproduced with permission from Ref. [68]. Copyright 2020, Wiley. (D) A schematic diagram of an ideal PB unit cell. Mx+ represents guest cations (e.g., Li+, Zn2+, K+, Al3+, Na+ and the corresponding hydrated ions). (E) A schematic of the working principle of the Zn-PB electrochromic device. (F) Visible-near infrared transmittance spectrum of the Zn-PB electrochromic device under different states. The inset shows the open-circuit voltage (OCV) of the device. (G) Photographs of a 0.5 V LED powered by the Zn-PB electrochromic device along with an obvious self-bleaching phenomenon. (D–G) Reproduced with permission from Ref. [69]. Copyright 2020, Royal Society of Chemistry.
Figure 2:

Structure, working principle, and application of Zn-PB electrochromic device.

(A) Scanning electron microscope (SEM) image of the PB thin films. (B) Transmittance spectra of the Zn-PB device at different discharge voltages from 0.5 to 1.9 V. (C) Photographs of the Zn-PB electrochromic device at different discharge voltages. (A–C) Reproduced with permission from Ref. [68]. Copyright 2020, Wiley. (D) A schematic diagram of an ideal PB unit cell. Mx+ represents guest cations (e.g., Li+, Zn2+, K+, Al3+, Na+ and the corresponding hydrated ions). (E) A schematic of the working principle of the Zn-PB electrochromic device. (F) Visible-near infrared transmittance spectrum of the Zn-PB electrochromic device under different states. The inset shows the open-circuit voltage (OCV) of the device. (G) Photographs of a 0.5 V LED powered by the Zn-PB electrochromic device along with an obvious self-bleaching phenomenon. (D–G) Reproduced with permission from Ref. [69]. Copyright 2020, Royal Society of Chemistry.

Another class of electrochromic materials that attracts recent attention are based on MXene. MXenes represent a large family of 2D transition metal carbides and nitrides which have been utilized for energy storage and various optoelectronic applications [73]. For example, titanium carbide (Ti3C2Tx) has been shown to exhibit electrochromic properties cycling in 1 M phosphoric acid polyvinyl alcohol gel electrolyte (H3PO4/PVA gel) [74].

2.2 Organic materials

Electrochromic characteristics are shown by a wide range of organic materials, including organic molecules [75], [76], [77] and conjugated polymers [78], [79], [80], [81], [82], [83], [84], [85], [86], [87]. Typically, organic electrochromic materials offer multiple colors under different applied biases, which are highly desired for multicolor displays. Viologens, 1,1′-disubstituted-4,4′-bipyridinium salts, are by far the most intensively investigated small molecule–based electrochromic materials [75]. Viologens exhibit three reversible redox states: a dication (V2+), a radical cation (V+), and a neutral form (V), thus yielding differently colored species (Figure 3A). The intense color is produced upon reduction of the viologen dication, leading to a strong absorption at its radical cation state. The substitution groups on the nitrogen of the bipyridinium salt mainly control the colors of their reduction states. As illustrated in Figure 3B, an ethyl viologen–based electrochromic device exhibits a blue color in its reduced state and a light yellow color in its oxidized state [77]. The major advantages of viologens as electrochromic materials reside in their high optical contrast, excellent coloration efficiency (CE), ease in molecular design, and the feasibility for large-area assembly. A successful application is their use by Gentex Corporation as a smart window in Boeing 787 aircraft [30].

Figure 3: 
Characterization and demonstration of organic electrochromic materials.
(A) Different redox states of viologen. (B) The photographs of ethyl viologen–based electrochromic device at bleached and colored states. (A–B) Reproduced with permission from Ref. [75]. Copyright 2019, Multidisciplinary Digital Publishing Institute. (C) The photographs of the flexible polyaniline (PANI) film electrode in its colored state at the different potentials. (D) The corresponding optical transmittance spectra of the flexible PANI film electrode at different potentials. (C–D) Reproduced with permission from Ref. [86]. Copyright 2018, Elsevier Ltd. (E) Photographs of the electrochromic symmetrical displays built from PEDOT: PSS films with different thicknesses, showing an oxidized state for a potential of +1.6 V and a reduced (colored) state for a potential of −1.6 V. The corresponding measured CIE L*a*b* parameters are labeled. Reproduced with permission from Ref. [84]. Copyright 2019, Multidisciplinary Digital Publishing Institute.
Figure 3:

Characterization and demonstration of organic electrochromic materials.

(A) Different redox states of viologen. (B) The photographs of ethyl viologen–based electrochromic device at bleached and colored states. (A–B) Reproduced with permission from Ref. [75]. Copyright 2019, Multidisciplinary Digital Publishing Institute. (C) The photographs of the flexible polyaniline (PANI) film electrode in its colored state at the different potentials. (D) The corresponding optical transmittance spectra of the flexible PANI film electrode at different potentials. (C–D) Reproduced with permission from Ref. [86]. Copyright 2018, Elsevier Ltd. (E) Photographs of the electrochromic symmetrical displays built from PEDOT: PSS films with different thicknesses, showing an oxidized state for a potential of +1.6 V and a reduced (colored) state for a potential of −1.6 V. The corresponding measured CIE L*a*b* parameters are labeled. Reproduced with permission from Ref. [84]. Copyright 2019, Multidisciplinary Digital Publishing Institute.

Some long-chain polymers, such as polypyrroles [78], [79], [80], polythiophenes [81], [82], [83], and polyanilines [85], [ 86], are found to possess electrochromic properties by altering the structures of their π-conjugated systems. Polyaniline (PANI) films exhibit reversible multicolors: transparent ↔ pale yellow ↔ green ↔ blue, as the applied potential is varied [88]. As shown in Figure 3C, PANI films can be prepared on flexible indium tin oxide (ITO)/polyethylene terephthalate substrates [86]. The optical transmission spectra of PANI films at different potentials and color states are shown in Figure 3D. The transmittance variation between colorless and blue states can be as high as 49% at 630 nm as the applied potential increased from −0.2 to 1.0 V. Another interesting polymer is poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (abbreviated as PEDOT:PSS). PEDOT:PSS is a conductive polymer which exhibits a blue coloration on electrochemical reduction. Figure 3E depicts a series of images of the electrochromic symmetrical displays built from PEDOT:PSS films of various thickness [84]. A clear color gradient was observed with the increase of thickness in reduced (colored) state as distinguished by the CIE L*a*b* color space parameters. Clearly, utilizing PEDOT:PSS as a transparent conducting electrode is promising in applications requiring a stretchable electrochromic platform [89], [ 90].

2.3 Inorganic–organic hybrids

Different electrochromic materials have their respective advantages and disadvantages, thus instead of using a single class of electrochromic materials, organic–inorganic electrochromic hybrids can combine the advantages of each platform, thus offering enhanced electrochromic properties. Considerable efforts have been devoted toward the development of organic–inorganic electrochromic nanocomposites, wherein the assembly of the organic polymers into an inorganic porous framework is shown to lead to enhanced electrochromic performance and chemical stability [91], [92], [93]. For example, the organic polymers PEDOT:PSS can also be introduced into porous nanorod layers of tungsten molybdenum oxide (W0.71Mo0.29O3) [92]. Figure 4A and B schematically illustrate the spray layer-by-layer assembly process used to prepare the W0.71Mo0.29O3/PEDOT:PSS hybrid electrochromic films. The PEDOT:PSS chains were observed to permeate into the porous nanorod layers of the W0.71Mo0.29O3, forming an interconnected conductive network (Figure 4C and D). The W0.71Mo0.29O3/PEDOT:PSS hybrid electrochromic films realized an improved electrochromic performance that is superior to both pure PEDOT:PSS and W0.71Mo0.29O3 films. In the organic–inorganic nanocomposites, the interfacial interactions between the organic and the inorganic phases have a large impact on the properties of the nanocomposites. A good distribution of the introduced materials within the matrix material and their strong interfacial interaction is shown to improve the mechanical and electrochromic properties of structural nanocomposites because of the synergic effect of these materials [94], [ 95]. A recent study demonstrated a class of hybrid electrochromic materials utilizing polythiophenes and tin-doped indium oxide nanoparticles that enabled the independently modulating visible and near-infrared light [96]. This inorganic–organic hybrid electrochromic film combines polythiophenes to modulate visible light with tin-doped indium oxide nanoparticles that modulate near-infrared light. Therefore, such hybrid materials can be operated in three distinct voltage regimes, each of which uniquely modulates visible and near-infrared light. Besides simple physical contact, the strong interfacial interactions between the organic and inorganic components (e.g., covalent bond, coordination bond, electrostatic interaction, hydrogen bond, and π–π stacking interaction) are conducive for avoiding structural damage caused by the volume change during the switching process and thus improving the structural stability of the hybrids [97], [98], [99], [100]. For example, Yang et al. studied covalently bonded WO3/polyvinylimidazole core–shell microspheres [98]. It is found that the chemical and thermal stabilities of covalently bonded WO3/polyvinylimidazole are higher than those of pure WO3 nanoparticles and noncovalently bonded WO3/polyvinylimidazole. The covalent bond (∼100 to 500 kJ mol−1) can supply solid strength for the organic/inorganic electrochromic nanocomposites, leading to enhanced electrochemical and electrochromic performances [97].

Figure 4: 
Preparation and characterization of organic-inorganic hybrid films.
(A) Schematic illustration of the spray layer-by-layer assembly process used to prepare the W0.71Mo0.29O3/PEDOT:PSS hybrid electrochromic films. (B) Schematic illustration of the evolution of the W0.71Mo0.29O3/PEDOT:PSS hybrid electrochromic films during the initial four cycles. Layers 1 and 3 depict the porous structure of W0.71Mo0.29O3 material, whereas Layers 2 and 4 depict the PEDOT:PSS material. (C) Field emission scanning electron microscope (FESEM) image of the spray-coated pure W0.71Mo0.29O3 films (D) FESEM image of the spray-coated W0.71Mo0.29O3/PEDOT:PSS hybrid films (A–D) Reproduced with permission from Ref. [92]. Copyright 2018, American Chemical Society. (E) Optical absorbance of FeL2-MEPEs at different potentials with the corresponding photographs of the thin films. Reproduced with permission from Ref. [104]. Copyright 2017, American Chemical Society. (F) Schematic illustration of linear OsII‐based MEPEs electrochromic device. Reproduced with permission from Ref. [105]. Copyright 2018, Wiley.
Figure 4:

Preparation and characterization of organic-inorganic hybrid films.

(A) Schematic illustration of the spray layer-by-layer assembly process used to prepare the W0.71Mo0.29O3/PEDOT:PSS hybrid electrochromic films. (B) Schematic illustration of the evolution of the W0.71Mo0.29O3/PEDOT:PSS hybrid electrochromic films during the initial four cycles. Layers 1 and 3 depict the porous structure of W0.71Mo0.29O3 material, whereas Layers 2 and 4 depict the PEDOT:PSS material. (C) Field emission scanning electron microscope (FESEM) image of the spray-coated pure W0.71Mo0.29O3 films (D) FESEM image of the spray-coated W0.71Mo0.29O3/PEDOT:PSS hybrid films (A–D) Reproduced with permission from Ref. [92]. Copyright 2018, American Chemical Society. (E) Optical absorbance of FeL2-MEPEs at different potentials with the corresponding photographs of the thin films. Reproduced with permission from Ref. [104]. Copyright 2017, American Chemical Society. (F) Schematic illustration of linear OsII‐based MEPEs electrochromic device. Reproduced with permission from Ref. [105]. Copyright 2018, Wiley.

Another important example of inorganic–organic hybrids is metallo-supramolecular polyelectrolytes (MEPEs). This hybrid electrochromic material is based on the metal-to-ligand charge transfer and the intervalence charge transfer [101], [102], [103]. Through altering the metal ion center or the surrounding organic ligands, MEPE can switch between various colors. As shown in Figure 4E, Kurth et al. reported a reversible redox behavior of FeII/FeIII in FeL2-MEPEs device where a color-switching between the red and the green colors was demonstrated [104]. Furthermore, electrochromic performance can be observed in OsII-based MEPEs [105]. Figure 4F illustrates an electrochromic device using linear OsII‐based MEPEs, where a distinct color change from yellow-green to pale-pink was observed on reduction reaction.

2.4 Nanoparticles and nanostructured plasmonic material

Plasmonic devices can experience an unconventional electrochromic process, where the color modulation of the electrochromic device is primarily based on dynamically changing the resonance of plasmonic nanostructures. Nanoparticles plasmonic materials are identified as either colloidal metal nanoparticles, which allow for reversible metal deposition, or highly doped metal oxide nanocrystals that can be electrochemically charged in a capacitive manner due to their extremely small size and high surface area [45]. Light couples to the colloidal plasmonic material via the resonant excitation of free charge carriers as localized surface plasmon resonances (LSPRs). Here, LSPR enables a wide range of spectral tunability by changing the nanocrystal size, altering the host environment, or via doping or injection of charges. When the incident electromagnetic radiation frequency matches the natural resonance frequency of collective electron oscillations in the conduction band (or holes in the valence band) of the nanocrystals (Figure 5A), light absorption and scattering of nanocrystals are notably enhanced. Although the LSPR frequency of nanocrystals is slightly tunable by their shapes, the doping level, and the nanostructure size, the free electron density plays a key role because it determines the starting position, width, and intensity of the LSPR absorption peak [106], [107], [108]. Typically, metal-based plasmonic nanostructures exhibit intrinsic free carrier density on an order of 1022 cm−3 [109], which leads to high energy electronic oscillations resulting in plasmonic resonance frequencies in the visible spectral range. Metal-based electrochromic devices have been demonstrated via reversible electrochemical deposition. Although metals lack the oxidation states necessary for an inherent electrochromism, unconventional electrochromic devices have been demonstrated through the deposition of Ag nanoparticles on an ITO electrode [110], [111], [112]. Here, the redox reaction occurs when Ag+-ions are reduced at the electrode and thereby agglomerate into elemental metal nanoparticles. Oppositely, an inverse charge can lead to the dissolution of Ag nanoparticles. As such, the deposition and dissolution of Ag nanoparticles are reversible, and thus leading to the reversible optical evolution. Based on the size and shape of the deposited particles, which are strongly dependent on the deposition time, bias voltage, and chemical makeup of the solvent, the resonance frequency can be modulated over a wide range. A recent study by Kobayashi et al. demonstrated a multicolor electrochromic device based on the Ag deposition mechanism [111]. Figure 5B illustrates the morphology of Ag deposited on the ITO particle-modified electrode having a diameter size of 21 nm. With the deposition of Ag on the ITO electrode, the LSPRs of Ag nanoparticles result in displaying multicolors (Figure 5C).

Figure 5: 
Depiction of plasmonic electrochromic nanocrystals and nanostructures.
(A) Illustration of a localized surface plasmon in a nanocrystal. Incident electromagnetic radiation displaces the free electron cloud, which then experiences a restoring force exerted by the positively charged ion cores. Reproduced with permission from Ref. [45]. Copyright 2014, Royal Society of Chemistry. (B) Scanning electron microscope (SEM) image of the Ag nanoparticles deposited on the ITO particle-modified electrode. (C) Photographs of the working electrodes with various ITO film thicknesses during the Ag deposition process. (B–C) Reproduced with permission from Ref. [111]. Copyright 2014, American Chemical Society. (D) The mechanism of capacitive electrochromism based on nanocrystal films. In the OFF state, a positive potential is applied to the nanocrystals, no electrons and ions gather at the surface of nanocrystals. In the ON state, a negative potential is applied to the nanocrystals, ions are absorbed to the surface of nanocrystals in order to neutralize the injected electrons. Reproduced with permission from Ref. [45]. Copyright 2014, Royal Society of Chemistry. (E) TEM image of ITO nanocrystals. (F) Potentials-dependent optical density of a film composed of 4.1 nm diameter, 16.8% Sn ITO nanocrystals. (E–F) Reproduced with permission from Ref. [114]. Copyright 2011, American Chemical Society. (G) Schematic of a plasmochromic device based on an electrochromic WO3 layer and a gold resonator. Reproduced with permission from Ref. [118]. Copyright 2020, American Chemical Society.
Figure 5:

Depiction of plasmonic electrochromic nanocrystals and nanostructures.

(A) Illustration of a localized surface plasmon in a nanocrystal. Incident electromagnetic radiation displaces the free electron cloud, which then experiences a restoring force exerted by the positively charged ion cores. Reproduced with permission from Ref. [45]. Copyright 2014, Royal Society of Chemistry. (B) Scanning electron microscope (SEM) image of the Ag nanoparticles deposited on the ITO particle-modified electrode. (C) Photographs of the working electrodes with various ITO film thicknesses during the Ag deposition process. (B–C) Reproduced with permission from Ref. [111]. Copyright 2014, American Chemical Society. (D) The mechanism of capacitive electrochromism based on nanocrystal films. In the OFF state, a positive potential is applied to the nanocrystals, no electrons and ions gather at the surface of nanocrystals. In the ON state, a negative potential is applied to the nanocrystals, ions are absorbed to the surface of nanocrystals in order to neutralize the injected electrons. Reproduced with permission from Ref. [45]. Copyright 2014, Royal Society of Chemistry. (E) TEM image of ITO nanocrystals. (F) Potentials-dependent optical density of a film composed of 4.1 nm diameter, 16.8% Sn ITO nanocrystals. (E–F) Reproduced with permission from Ref. [114]. Copyright 2011, American Chemical Society. (G) Schematic of a plasmochromic device based on an electrochromic WO3 layer and a gold resonator. Reproduced with permission from Ref. [118]. Copyright 2020, American Chemical Society.

Highly doped metal oxide nanocrystals exhibit charge carrier densities in the range of 1020–1022 cm−3, which leads to pronounced surface plasmon resonances in the near-infrared spectral range. The resonance of free carriers can be modulated through an electrochemical gating process [113]. The change in carrier concentration leads to a shift in the LSPR frequency and a change in absorption, giving rise to plasmonic electrochromism [45]. Even though this plasmonic electrochromism does not lead to a change in the oxidation state of the material, devices based on metal oxide plasmonic nanostructures show reversible optical transmission changes through electrochemical charge modulation. Figure 5D depicts an illustration of the capacitive nature of the plasmonic electrochromic mechanism. The modulation of the electron/hole density is electrochemically controlled by the application of negative and positive potentials to the material. Milliron et al. demonstrated a surface plasmon effect in tin-doped indium oxide (ITO) nanocrystalline film [114]. Figure 5E illustrates a transmission electron microscopy (TEM) image of the ITO nanocrystals having varying sizes of 4.1 ± 0.6 nm. When different potentials are applied, the surface plasmon resonance peak of the ITO nanocrystalline film (4.1 nm diameter, 16.8% Sn) was shown to be enhanced and shifts to higher energy (Figure 5F).

In addition, conventional electrochromic materials, such as tungsten oxide or polyaniline, can be used to modulate the LSPR frequency of metallic plasmonic nanostructures. Chumanov et al. utilized a WO3-sol on an array of Ag nanoparticles to reversibly modulate the resonance frequency of the Ag particles by changing their dielectric surroundings [115]. Recently, several studies have shown the potential of electrochromic modulation of the dielectric environment of metallic nanostructures for a full-color generation [116], [117], [118]. In general, these devices make use of the inherent change of the refractive index of an electrochromic material during the intercalation and the deintercalation processes. This change is utilized to shift the resonance frequency of plasmonic nanostructures, such as colloids and nanohole arrays. Hopmann and Elezzabi reported a plasmonic metal-insulator–nanohole cavity based on an electrochromic WO3 insulator layer. Through intercalation and deintercalation of lithium ions (Li+) into the WO3 film, a broad peak resonance wavelength shift of 64 nm was realized, leading to distinct color reflection (Figure 5G) [118].

3 Key metrics for electrochromic materials

An electrochromic material alters its optical and electrical properties in response to an external voltage bias. A high-performance electrochromic material must possess fast color switching, reversible light modulations, high color contrast, low energy consumption, and long-term chemical and thermal stabilities. Key performance metrics are introduced and highlighted below.

3.1 Color and optical contrast

Color is an important aspect in assessing the performance of electrochromic materials as their applications rely on switching between different colors. The colorimetric analysis is usually performed based on the protocols set out by the International Commission on Illumination (CIE). The CIE 1931 xyY color space is widely used to specify color in practice. As shown in Figure 6A, an electrochromic film configured in ultracompact Fabry–Perot nanocavities displays seven distinguished colors [17]. The Fabry–Perot nanocavities utilize partially reflective tungsten metal as the current collector and the reflector layer. As the thickness of the WO3 thin film changes, various structural colors are generated. These colors cause a remarkable shift of the CIE color coordinates (Figure 6B), forming a circular area with different tints. More interestingly, the Fabry–Perot nanocavity-type electrochromic films exhibit a large variety of multicolor states under different applied potentials (Figure 6C). Although the 163-nm-thick WO3/W electrode produces a gradual color change from red to green at different biases (Figure 6D), the reflection spectrum displays a drastic shift from 760 to 517 nm in its peak position (Figure 6E).

Figure 6: 
Demonstration of electrochromic materials having different colors and light spectra.
(A) Photographs of the WO3/W electrodes having different thicknesses of the WO3 layer. (B) CIE color coordinates of the WO3/W electrodes having different thicknesses of the WO3 layer. (C) Color gallery obtained from the WO3/W electrodes under different voltage biases. (D) CIE color coordinates of the 163/200-nm-thick WO3/W electrode under different voltage biases. (E) Reflectance spectra and photographs of the 163-nm-thick WO3/W electrode under different voltage biases. (A–E) Reproduced with permission from Ref. [17]. Copyright 2020, Nature Publishing Group. (F) Transmittance spectra and photographs of MoO2+x-WO3 electrochromic device. Reproduced with permission from Ref. [22]. Copyright 2019, American Chemical Society.
Figure 6:

Demonstration of electrochromic materials having different colors and light spectra.

(A) Photographs of the WO3/W electrodes having different thicknesses of the WO3 layer. (B) CIE color coordinates of the WO3/W electrodes having different thicknesses of the WO3 layer. (C) Color gallery obtained from the WO3/W electrodes under different voltage biases. (D) CIE color coordinates of the 163/200-nm-thick WO3/W electrode under different voltage biases. (E) Reflectance spectra and photographs of the 163-nm-thick WO3/W electrode under different voltage biases. (A–E) Reproduced with permission from Ref. [17]. Copyright 2020, Nature Publishing Group. (F) Transmittance spectra and photographs of MoO2+x-WO3 electrochromic device. Reproduced with permission from Ref. [22]. Copyright 2019, American Chemical Society.

The intensity difference between the colors is an important factor in determining the quality of the electrochromic coloration process. As such, the optical contrast, referring to the degree of optical change during the switching process (intensity change in transmittance [ΔT %], absorbance [ΔA %] or reflectance [ΔR %]), is another metric used in evaluating the electrochromic material performance. However, when describing the optical contrast of electrochromic materials or devices, it is customary to report this parameter at a single wavelength. This wavelength is selected to be either the wavelength at which a maximum absorbance occurs or the wavelength of maximum human visual sensitivity (i.e., 550 nm) [44]. Figure 6F depicts a WO3-based electrochromic device, where a 63% change in transmittance at 633 nm was achieved at an applied voltage of −2.5 V [22]. The choice of the 633 nm wavelength used to calculate the optical contrast is dictated by the maximum absorbance observed at this wavelength. In general, the optical contrast of a material mainly depends on the thickness of the electrochromic layer [119]. As the film thickness increasing, the transmittance of both the colored and bleached states will decrease. Typically, the maximum optical contrast is reported at an attained intermediate thickness of a few hundred nanometers [120], [ 121]. The design of nanostructured electrochromic materials is expected to enhance the electrochromic performance (e.g., optical contrast, switching times, CE, and cycling stability) [122], [123], [124]. A recent study by Scherer et al. demonstrated that the gyroid‐structured vanadium pentoxide films surpass previous inorganic electrochromic materials in all relevant parameters: the switching speed, coloration contrast, and composite CE [125].

3.2 Coloring and bleaching switching times

Switching times, also known as the response times, refer to the times required to switch between the colored/bleached and the bleached/colored states of an electrochromic material or a device. The coloration and bleaching times are defined as the time required for 90% change in the entire optical modulation [126]. As a key parameter when considering the suitability of an electrochromic device for light modulation applications, fast switching between the different color states is highly desirable. Figure 7A illustrates an Al anode–based electrochromic energy storage smart window switching between 4 V (transparent state), 2.2 V (cool state), and 1.6 V (dark state) [4]. The switching times of this device were calculated to be 71 s for bleaching and 19 s for coloration at 633 nm, and 56 s for bleaching and 18 s for coloration at 1200 nm (Figure 7B and C). Generally, the switching times of an electrochromic layer will increase as the film thickness increasing. Thicker films with a fixed porosity would lead to longer ion-transport pathways to insert into the inner active materials, which results in longer switching times [127], [128], [129]. The nanostructured electrochromic materials with large surface active area are promising for fast switching times as the active materials are easy to access for guest ions [130], [ 131]. For example, rapid switching times of 0.5/0.9 s for the coloration and bleaching processes were reported for nanostructured electrochromic materials: core–shell halloysite nanotube (HNT)@inherently conducting polymer nanocomposites [130]. Other Parameters, such as the electrolyte properties, counter electrode material composition, substrate sheet resistance and device size, all have a significant effect on the switching times.

Figure 7: 
Al anode–based electrochromic energy storage windows.
(A) Photographs of an Al anode–based electrochromic energy storage window at different applied potentials. (B) The dynamic test of the device between 4 and 2.2 V at 1200 nm. (C) The dynamic test of the device between 4 and 1.6 V at 633 nm. (D) Coloration efficiency of the device switching at 633 and 1200 nm. (A–D) Reproduced with permission from Ref. [4]. Copyright 2019, Wiley.
Figure 7:

Al anode–based electrochromic energy storage windows.

(A) Photographs of an Al anode–based electrochromic energy storage window at different applied potentials. (B) The dynamic test of the device between 4 and 2.2 V at 1200 nm. (C) The dynamic test of the device between 4 and 1.6 V at 633 nm. (D) Coloration efficiency of the device switching at 633 and 1200 nm. (A–D) Reproduced with permission from Ref. [4]. Copyright 2019, Wiley.

3.3 Coloration efficiency

An energy-efficient electrochromic material requires a high optical contrast while consuming fewer charges in the process. In this regard, CE, representing the change in optical density (ΔOD) per unit of charge intercalated into the electrochromic layer, is proposed to evaluate the energy-related efficiency accordingly,

(4) Δ OD = log ( T c / T b ) = η Q / A ,

where η is CE, Q is the charge (C), A is the active area of the electrochromic material (cm2), Tc and Tb are the transmittances of electrochromic devices at coloration and bleaching states, respectively. An example of CE value calculation is shown in Figure 7D. Here, the Al anode–based electrochromic energy storage window exhibited CE values of 165 cm2 C−1 at 1200 nm and 85 cm2 C−1 at 633 nm, respectively. The CE will increase as the thickness increasing till a maximum point, then the CE will decrease as the thickness keeps increasing [132], [ 133]. The maximum CE value is obtained at an intermediate thickness (∼ a few hundred nanometers), which requires rational design and carefully screening several thicknesses of electrochromic films. This phenomenon probably relates to the different ionic/electric resistance of the electrochromic films at various thicknesses [134]. Because the CE is an intrinsic parameter of electrochromic materials and mainly relies on the morphology of the materials [135], the design of nanostructured electrochromic materials having large surface active area is required to achieve high CE values. As reported by Ma et al., the CE value of WO3 nanoflake films film is much higher than the WO3 nanoparticle/nanorod films [136].

3.4 Cycling stability

Cycling stability is the loss of the optical contrast on electrochemical cycling and is another important parameter used to evaluate the long-term stability of the electrochromic device [137]. For practical applications, electrochromic materials are expected to exhibit a cycle life of more than 105 cycles. However, in laboratory exploration, cycling stability is often reported for a small area of electrochromic film which is not sufficient for real-world applications. Both physical and chemical stabilities are crucial to the cycling stability of an electrochromic thin film. Generally, as the electrochromic film’s thickness increasing, the electrochromic film would suffer a poor adhesion to the substrate and thus leading to the significant degradation of the cycle performance due to the peeling of active materials [138]. A strong interfacial interaction between the film and the substrate will certainly enhance the cycling stability [139]. The design of nanostructures is expected to improve both physical and chemical stability of thin films. Nanostructured electrochromic materials can be designed to optimize the contacting area between the film and the substrate and hence providing a strong interfacial adhesion [140]. In addition, the volume expansion of nanostructured electrochromic materials during the switching process is supposed to be significantly eliminated as the numerous surface active sites increase the capacitive contribution during the redox reactions [141]. In other words, nanostructured electrochromic materials possess excellent chemical stability on cycling. Besides the stability of thin films, the cycling stability of an electrochromic device is influenced by several factors, including the electrolyte properties, the counter electrode material, and the device dimensions.

4 Design of nanostructured inorganic electrochromic materials

To develop electrochromic materials having high performance, including the aforementioned metrics for potential real-world applications, a list of recent achievements on the fabrication of promising inorganic electrochromic materials is presented. It is hoped that universal strategies for fabricating nanostructured inorganic electrochromic materials can be realized to advance the future development of the electrochromism.

4.1 Tungsten oxides

The most commonly used cathodic electrochromic materials are based on tungsten oxides (e.g., tungsten oxide, hydrated tungsten oxide, and doped tungsten oxide). Naturally, bulk tungsten oxide exhibits intrinsic oxygen deficiencies [142], making the material appear yellowish. By contrast, in a thin-film configuration, stoichiometric tungsten oxide is transparent and becomes blue on applying a cathodic potential [143]. Tungsten oxide has a nearly cubic structure formed by WO6 octahedra that share corners [144]. The void space inside the cubes provides the accessibility of numerous interstitial sites where guest ions (e.g., protons or alkali cations) can be inserted. The first study on the electrochromic effect in tungsten trioxide was reported in 1969 [145]. Over the past decades, the most promising application of the WO3 electrochromic coating is as a switchable glazing for smart windows. The color change in WO3 electrochromic smart windows offers an important energy-efficient module for designing environmentally friendly buildings by reducing the cost of air conditioning, controlling direct sun illumination, and simultaneously improving the indoor living comfort for the occupants [25].

Recently, nanostructured WO3 films have been realized for practical applications because of their excellent electrochromic performance. A variety of crystallographic and morphological structures of WO3 films have been reported to exhibit shorter switching time, higher coloration efficiencies, and higher cycling stabilities compared with their bulk counterparts. The nanostructured WO3, including nanoparticles [62], [ 146], nanowires [147], nanosheets [148], nanofibers [149], and hierarchical structures [150], possess a large interfacial contact area with the electrolyte, short ions diffusion pathway, and good structure stability associated with the intercalation/deintercalation of guest ions. These advantages of nanostructured WO3 give rise to highly efficient and stable electrochromic performance. Shan et al. reported the first demonstration of tungsten oxide quantum dots (QDs) for electrochromic applications [151]. As shown in Figure 8A, because the QDs size is very small (∼1.6 nm), this yielded a fast electron/ion transport during the coloration/bleaching process. Similarly, Yao et al. reported WO3 QDs with an average diameter of 1.2 nm (Figure 8B) [152]. The cycle stability of WO3 QDs electrochromic films can reach 20,000 cycles on the intercalation/deintercalation of Al3+. Moreover, larger WO3 QD nanocrystals and their inks (∼4 nm in diameter, Figure 8C and D) were investigated by Zhang et al. for inkjet-printed flexible electrochromic devices [146]. The printed tungsten oxide thin films are assembled into a novel Zn-based electrochromic pseudocapacitive device, which exhibits a relatively high capacity (≈260C g−1 at 1 A g−1). In addition to the aforementioned QD WO3 nanocrystalline structures, Zhang et al. reported oxygen-deficient monoclinic tungsten oxide nanowire (m-WO3−x NW) (Figure 8E and F). With efficient Al3+ intercalation/deintercalation, the m-WO3−x NWs-based electrochromic device delivered not only efficient and independent control of near-infrared and visible light transmittance but also exhibited high CE [47]. High optical contrast of 93.2%, high CE of 121 cm2 C−1, fast switching times of 16/13 s, and good cycling stability (5.5% capacity loss after 2000 cycles) were reported. An interesting form of nanostructured WO3 films is the hierarchical nest-like WO3·0.33H2O film (Figure 8G). Li et al. reported a self-seeded hydrothermal process that eliminates the grain boundaries existing in the nanocrystalline seed layer [153]. This process accelerates electron transport to the FTO glass substrate and promotes electron transfer efficiency. As shown in Figure 8H, the prepared WO3·0.33H2O films have an optical contrast of 40.8 and 57.6% after applying voltages of −0.5 and −1.0 V, respectively. The response times of WO3·0.33H2O films are found to be 26 and 5.5 s under ±3.0 V bias. Owing to the large active surface area, this self-seeded film also exhibits a high CE value of 126.34 cm2 C−1.

Figure 8: 
Structure and characterization of WO3-based electrochromic films.
(A) Transmission electron microscopy (TEM) image of tungsten oxide quantum dots (QDs) with average sizes of 1.6 nm. Reproduced with permission from Ref. [151]. Copyright 2014, Wiley. (B) High-resolution TEM image of as-prepared tungsten oxide QDs. Reproduced with permission from Ref. [152]. Copyright 2020, Elsevier. (C) TEM image of WO3−x nanocrystals. Inset shows the nanocrystals diameter statistics. (D) WO3−x inks based on four different solvents. (C–D) Reproduced with permission from Ref. [146]. Copyright 2020, Wiley. (E) XRD patterns and crystal structure (inset) of m-WO3−x NWs. (F) Surface and cross-sectional (inset) SEM images of m-WO3−x NW films. (E–F) Reproduced with permission from Ref. [47]. Copyright 2018, Royal Society of Chemistry. (G) Scanning electron microscope (SEM) image of self-seeded grown WO3·0.33H2O electrode with less amount of ethylene glycol. (H) Optical transmittance spectra and photographs of the self-seeded grown WO3·0.33H2O film measured at different biases. (G–H) Reproduced with permission from Ref. [153]. Copyright 2014, Royal Society of Chemistry. FESEM images of WO3 films doped with different Ni content: (I) 0%, (J) 0.5%, (K) 1.5% and (L) 2.5%. (I–L) Reproduced with permission from Ref. [56]. Copyright 2016, Royal Society of Chemistry.
Figure 8:

Structure and characterization of WO3-based electrochromic films.

(A) Transmission electron microscopy (TEM) image of tungsten oxide quantum dots (QDs) with average sizes of 1.6 nm. Reproduced with permission from Ref. [151]. Copyright 2014, Wiley. (B) High-resolution TEM image of as-prepared tungsten oxide QDs. Reproduced with permission from Ref. [152]. Copyright 2020, Elsevier. (C) TEM image of WO3−x nanocrystals. Inset shows the nanocrystals diameter statistics. (D) WO3−x inks based on four different solvents. (C–D) Reproduced with permission from Ref. [146]. Copyright 2020, Wiley. (E) XRD patterns and crystal structure (inset) of m-WO3−x NWs. (F) Surface and cross-sectional (inset) SEM images of m-WO3−x NW films. (E–F) Reproduced with permission from Ref. [47]. Copyright 2018, Royal Society of Chemistry. (G) Scanning electron microscope (SEM) image of self-seeded grown WO3·0.33H2O electrode with less amount of ethylene glycol. (H) Optical transmittance spectra and photographs of the self-seeded grown WO3·0.33H2O film measured at different biases. (G–H) Reproduced with permission from Ref. [153]. Copyright 2014, Royal Society of Chemistry. FESEM images of WO3 films doped with different Ni content: (I) 0%, (J) 0.5%, (K) 1.5% and (L) 2.5%. (I–L) Reproduced with permission from Ref. [56]. Copyright 2016, Royal Society of Chemistry.

Doping WO3 with other elements has been realized as an efficient method to tune the nanostructures of WO3 [56]. Li et al. reported that the molybdenum doping of WO3 significantly reduces the particle size of the WO3. The solution-based fabrication process was simple and inexpensive to allow for large-scale fabrication of electrochromic films [126]. Furthermore, a slight increase of the Ni content (from 0 to 2.5%) alters the surface morphology of the WO3 film (Figure 8I–L) [56]. For the undoped WO3 film (Figure 8I), the SEM image shows a porous network structure. Whereas by doping the film with only 0.5% Ni, the Ni-WO3 grows into vertically aligned nanorod array structures (Figure 8J). As the doping concentration of Ni is increased to 1.5%, the nanorods become randomly distributed nanorods (Figure 8K). Once the Ni doping content was increased to 2.5%, nanorods disappeared and nanoflowers composed of clusters of nanoplates appear (Figure 8L). For the WO3 films with 0–1.5% Ni doping, the electrochromic performance is found to be enhanced with the increase of Ni concentration. The optical contrast in both the visible and near-infrared region increases with increasing Ni doping. At a light wavelength of 600 nm, the optical contrast enhanced from 50.9 to 86.0% at 0 and 1.5% Ni concentrations, respectively. Similarly, the CE is also higher at 60.5 cm2 C−1 for 1.5% Ni content compared with 20.8 cm2 C−1 for the undoped WO3 film. Notably, other doping techniques with different elements, such as Nb [154], Mo [155], Fe [156], and Ti [157], were shown to greatly improve the electrochromic performance of WO3 films. The most efficient strategy to develop the high-performance electrochromic WO3 films is the doping method [158], which requires further investigation from the community. To distinguish the electrochromic properties of different nanostructured WO3, Table 1 summarizes and compares the key electrochromic metrics.

Table 1:

Summary of the electrochromic properties of nanostructured WO3 with different morphologies.

Ref. Nanostructures Optical contrast (∆T) Switching time (s) Coloration efficiency (cm2C−1) Cycling numbers
Yao et al. [152] WO3

Quantum dots
97.8% at 633 nm Coloration: 4.5

Bleaching: 4
76.8 20,000
Cong et al. [151] WO3

Quantum dots
94% at 633 nm Coloration: 0.9

Bleaching: 1.0
154 50
Zhang et al. [146] WO3−x nanocrystals 76% at 633 nm Coloration: 4.5

Bleaching: 3.7
97.7 1000
Zhang at al. [47] m-WO3−x nanowires 93.2% at 633 nm Coloration: 16

Bleaching: 13
121 2000
Azam et al. [148] WO3 nanosheets 62.6% at 700 nm Coloration: 10.7

Bleaching: 6.9
1000
Adhikari et al. [149] WO3

Nanofibers
43% at 550 nm Coloration: 6.5

Bleaching: 4.0
63.15 500
Li et al. [150] WO3

Nanotrees
74.7% at 630 nm Coloration: 7.3

Bleaching: 2.6
75.35 250
Li et al. [153] WO3·0.33H2O

Nanonests
57.6% at 633 nm Coloration: 26

Bleaching: 5.5
126.34 1000
Zhou et al. [56] Ni-doped WO3

Nanorods
86% at 600 nm Coloration: 19

Bleaching: 6.2
60.5 5500
Zhan et al. [157] Ti-doped WO3

Nanoplates
67.6% at 633 nm Coloration: ∼15

Bleaching: ∼5
106.6 3000
Li et al. [126] W0.71Mo0.29O3−x nanoparticles 42.9% at 633 nm Coloration: 10

Bleaching: 7.5
36.3 2000
Li et al. [3] Mo/Ti:WO3 nanowires 76% at 633 nm Coloration: 14

Bleaching: ∼10
100

4.2 Molybdenum oxides

Similar to tungsten oxides, molybdenum oxides (e.g., molybdenum oxide, hydrated molybdenum oxide, and doped molybdenum oxide) are transparent in their oxidized state and become colored gray-blue on reduction. The layered crystalline structure of MoO3 allows it to exist in a quasi-two-dimensional state [159]. The ion intercalation capabilities, bandgap tunability, and various oxidation states allow MoO3 to be a promising electrochromic material [160]. Naturally, MoO3 exists in three phases: orthorhombic α-MoO3, monoclinic β-MoO3, and hexagonal h-MoO3 [161]. The orthorhombic α-MoO3 is the most stable phase of MoO3. This phase is constituted by MoO6 octahedra that share edges and corners. The structure’s unique layers parallel to the (010) plane provide open channels for guest ion intercalation (Figure 9A and B). However, molybdenum oxides suffer inferior cycling stability due to the large-volume expansion during the intercalation/deintercalation processes. As such, nanostructured molybdenum oxides in the form of nanoparticles [25], nanobelts [161], and nanosheets [162] have been shown to have improved electrochromic and electrochemical performances. Xie et al. reported on the fabrication of metastable hexagonal MoO3 nanobelts for electrochromic device applications (Figure 9C) [161]. The open channel structure of the hexagonal MoO3 nanobelts favors the efficient insertion of the guest ions (Figure 9D). As such, the MoO3 nanobelt films exhibit a more stable cycling performance than orthorhombic α-MoO3 films. Moreover, the nanohybridization of molybdenum oxide nanoparticles with tungsten molybdenum oxide nanowires offers an interesting electrochromic material platform as it circumvents many of the issues inherent to MoO3 (e.g., slow ionic transportation efficiency and volume expansion during cycling) [25]. Through a solution-processed method, MoO3 nanocrystals were embedded in the nanopores of W0.71Mo0.29O3 nanowire film to form a MoO3-W0.71Mo0.29O3 nanocomposite film (Figure 9E). The W0.71Mo0.29O3 nanowires not only serve as a buffer matrix which eliminates the large-volume expansion of MoO3 during the intercalation/deintercalation of guest ions, but also provides numerous charge transport pathways to enhance transport kinetics. Notably, the MoO3-W0.71Mo0.29O3 nanohybrid films show improved cycling stability compared with the intrinsic MoO3 films.

Figure 9: 
Structure and characterization of MoO3-based electrochromic films.
(A–B) Schematic representation of orthorhombic α-MoO3. (C) SEM image (inset is a cross-section view) of hexagonal MoO3 nanobelts. (D) Schematic representation of possible intercalation sites in hexagonal MoO3. (A–D) Reproduced with permission from Ref. [161]. Copyright 2009, American Chemical Society. (E) Schematic illustration of the matrix effect of the W0.71Mo0.29O3 nanowires. Reproduced with permission from Ref. [25]. Copyright 2018, Elsevier.
Figure 9:

Structure and characterization of MoO3-based electrochromic films.

(A–B) Schematic representation of orthorhombic α-MoO3. (C) SEM image (inset is a cross-section view) of hexagonal MoO3 nanobelts. (D) Schematic representation of possible intercalation sites in hexagonal MoO3. (A–D) Reproduced with permission from Ref. [161]. Copyright 2009, American Chemical Society. (E) Schematic illustration of the matrix effect of the W0.71Mo0.29O3 nanowires. Reproduced with permission from Ref. [25]. Copyright 2018, Elsevier.

4.3 Titanium oxides

Titanium oxides (e.g., titanium oxide, hydrated titanium oxide, and doped titanium oxide) have been recognized as one of the most promising electrode materials for semiconductor electrochemistry [163]. Naturally, bulk titanium oxide exists in three crystalline phases: rutile, anatase, and brookite [164]. Refer to the crystal structure of titanium oxide, the Ti4+ ions are surrounded by six O2− ions creating an [TiO6] octahedron, forming large vacant sites to accommodate guest ions. With the intercalation of guest ions, the transmittance of titanium oxide changes from a transparent state to a blue color. However, the electrochromic performance of intrinsic TiO2 is different from other open structures like WO3. A recent report by Tong et al. demonstrated that a TiO2 mesoporous nanotube film exhibited an optical contrast of 33.5% at 700 nm, which is higher than the bulk TiO2 films due to the enhanced Li-ion insertion kinetics [165]. Although the optical contrast is enhanced via designing nanotube structures, the electrochromic effect is still lower than other electrochromic materials. This is attributed to the intrinsic moderate electrochromic effect of pure TiO2 [166]. As such, the aliovalent doping strategy is supposed the most convenient method to attain superior electrochromic effect of TiO2-based electrochromic materials [167]. Hence, the focus of titanium oxide research has been shifted toward the synthesis of nanostructured doped-TiO2 films. Recently, Cao and et al. synthesized Ta-doped TiO2 nanocrystals [168] (Figure 10A and B). The fabricated Ta-doped TiO2 thin films showed excellent electrochromic performance in terms of high optical contrast for both the visible and the near-infrared light spectra regions (i.e., 86.3% at 550 nm and 81.4% at 1600 nm). After 2000 cycles, the optical contrasts of Ta-doped TiO2 thin films slightly decreased by 1.3 and 6.7% at 550 and 1600 nm, respectively, indicating good electrochemical stability. This Ta-doped TiO2 nanocrystal represents a promising new electrode material for smart windows applications [169].

Figure 10: 
Structure and characterization of TiO2-based electrochromic films.
(A) Photograph of Ta-doped TiO2 nanocrystals dispersion in tetrachloroethylene. (B) Transmittance and corresponding photographs of Ta-doped TiO2 thin films electrode at different states. (A–B) Reproduced with permission from Ref. [168]. Copyright 2018, American Chemical Society. (C) Spectral variation of the optical transmittance at −1.6 V for four batches of Nb-doped TiO2 nanocrystal films with different doping levels (from 0 to 15% of niobium content). (D) Transmittance spectra of a device embodying 10%Nb-doped TiO2 nanocrystals electrode at different bias potentials. Reproduced with permission from Ref. [36]. Copyright 2017, American Chemical Society.
Figure 10:

Structure and characterization of TiO2-based electrochromic films.

(A) Photograph of Ta-doped TiO2 nanocrystals dispersion in tetrachloroethylene. (B) Transmittance and corresponding photographs of Ta-doped TiO2 thin films electrode at different states. (A–B) Reproduced with permission from Ref. [168]. Copyright 2018, American Chemical Society. (C) Spectral variation of the optical transmittance at −1.6 V for four batches of Nb-doped TiO2 nanocrystal films with different doping levels (from 0 to 15% of niobium content). (D) Transmittance spectra of a device embodying 10%Nb-doped TiO2 nanocrystals electrode at different bias potentials. Reproduced with permission from Ref. [36]. Copyright 2017, American Chemical Society.

Another dopant of TiO2 reported by Barawi et al. [36] to enhance the electrochromic performance of TiO2 is Nb. Four batches of Nb-doped TiO2 nanocrystals having different doping levels (from 0 to 15% of niobium content) were used to prepare highly transparent mesoporous electrodes. As shown in Figure 10C, the 10% Nb-doped TiO2 nanocrystals electrode exhibit the best electrochromic performance in both the visible and near-infrared light spectral regions. A dual-band (i.e., capable of independent control of visible light and near-infrared transmittance) electrochromic device was constructed using a 10% Nb-doped TiO2 nanocrystals electrode. This dual-band electrochromic device is capable of independent control of visible light and near-infrared transmittance, showing a 67% optical contrast at 2000 nm (Figure 10D).

4.4 Vanadium oxides

By virtue of their multicolor behaviors, vanadium oxides are regarded as the most favorable inorganic materials for electrochromic displays [125], [170], [171]. Vanadium cation exists in V2+, V3+, V4+, and V5+ states in vanadium compounds, thus inducing the associated multicolor characteristics. While VO2 is known to be a thermochromic material [172], V2O3 [173], V3O7 [174], and V2O5 [175] are found to have attractive electrochromic properties. The most widely investigated vanadium oxide is V2O5. With the intercalation/deintercalation of guest ions, this oxide offers reversible three-color states (i.e., yellow ⇄ green ⇄ blue) [23]. However, owing to their low electrical conductivity, significant volume expansion during cycling, and the slow reaction kinetics of bulk, intrinsic unstructured vanadium oxides have not been widely used in electrochromic devices [24]. In recent years, nanostructured vanadium oxides have been studied to mediate these drawbacks because the nanostructures shorten the diffusion paths of ions and provide abundant active sites on the surface. Steiner et al. demonstrated a self-supporting double-gyroid (DG) structure of V2O5 (Figure 11A) [171]. The DG nanostructured V2O5 possessed a highly ordered structure having 11 nm wide struts and a high specific surface-to-bulk volume ratio of 161.4 μm−1, leading to a fast and efficient lithium-ion intercalation/extraction process. The assembled electrochromic supercapacitor based on two opposite DG nanostructured V2O5 electrodes is shown in Figure 11B, and its three-color displays (i.e., yellow, green, and blue) are presented in Figure 11C.

Figure 11: 
Structure and characterization of vanadium oxide-based electrochromic displays.
(A) Scanning electron microscope (SEM) of a mesoporous V2O5 double-gyroid film on an FTO substrate (B) Schematic illustration of the electrochromic supercapacitor design based on two laterally offset double-gyroid structured electrodes. (C) Photographs of an electrochromic supercapacitor device displaying color change on charge and discharge. The letters “NO” form the double-gyroid structured top electrode and “KIA” form the bottom electrode. (A–C) Reproduced with permission from Ref. [171]. Copyright 2012, American Chemical Society. (D) Crystal structure of the SVO nanorods. (E) Schematic illustration of the Zn–SVO electrochromic display. (F) Schematic illustration of the color overlay effect via the combination of orange, yellow, and green colors. (G) Digital photographs of the Zn–SVO display showing six colors obtained through the color overlay effect. Reproduced with permission from Ref. [16]. Copyright 2020, Nature Publishing Group.
Figure 11:

Structure and characterization of vanadium oxide-based electrochromic displays.

(A) Scanning electron microscope (SEM) of a mesoporous V2O5 double-gyroid film on an FTO substrate (B) Schematic illustration of the electrochromic supercapacitor design based on two laterally offset double-gyroid structured electrodes. (C) Photographs of an electrochromic supercapacitor device displaying color change on charge and discharge. The letters “NO” form the double-gyroid structured top electrode and “KIA” form the bottom electrode. (A–C) Reproduced with permission from Ref. [171]. Copyright 2012, American Chemical Society. (D) Crystal structure of the SVO nanorods. (E) Schematic illustration of the Zn–SVO electrochromic display. (F) Schematic illustration of the color overlay effect via the combination of orange, yellow, and green colors. (G) Digital photographs of the Zn–SVO display showing six colors obtained through the color overlay effect. Reproduced with permission from Ref. [16]. Copyright 2020, Nature Publishing Group.

Extending the color palette for electrochromic devices is one of the major technological challenges in the field [5]. In 2020, Zhang et al. reported a transparent multicolor display enabled by Zn-based electrochromic devices [16]. The authors developed a method for synthesizing NaV3O8·1.5H2O (SVO) nanorods as an electrochromic material. The hydrated sodium ions, inserted between the V3O8 layers, act as pillars to stabilize the layered structure (Figure 11D). The SVO film exhibits a reversible multicolor switch (orange ⇄ yellow ⇄ green) during Zn2+ insertion and extraction. The presence of sodium in SVO induces the orange color. A Zn-SVO device was assembled via sandwiching a Zn frame anode and a gel electrolyte (polyvinyl alcohol–ZnSO4) between two SVO electrodes (Figure 11E). Through the combination of two SVO electrode segments, the color overlay effect broadened the resultant color palettes of the Zn-SVO display. Figure 11F illustrates the color overlay effect obtained by superimposing the orange, yellow, and green colors. Because the two SVO electrode segments can be colored and bleached independently, the Zn-SVO device can display six colors (i.e., orange, amber, yellow, brown, chartreuse, and green) (Figure 11G).

4.5 Nickel oxides

As typical anodic electrochromic materials, serving as complementary material to tungsten oxides when assembling a device, nickel oxides (e.g., nickel oxide, hydrated nickel oxide, and doped nickel oxide) are known for their high coloration efficiency and optical contrast [59]. In contrast to the cathodic electrochromic materials such as WO3, the colored state of NiO is caused by electrochemical oxidation, with anion insertion (OH) or cation extraction (Li+) [58]. On the other hand, the NiO films become bleached by electrochemical reduction. The NiO film exhibits a reversible color change that switches between brown color and transparent state [176]. Similar to other intrinsic oxide electrochromic materials, critical issues, such as slow switching speed, low color contrast, and poor cycle durability of NiO limit its practical utility in electrochromic devices [177], [ 178]. Enhanced electrochromic performances of NiO thin films have been recently achieved through nanostructured NiO films. Boo et al. reported a nanoporous NiO thin film fabricated by chemical bath deposition [55]. The nanoporous NiO thin films having high porosity showed excellent electrochromic properties, with a transmittance change of 77% at 550 nm and a coloration efficiency of 99.5 cm2 C−1. Liou et al. described the growth of NiO nanorods via hot-filament metal-oxide vapor deposition where large-area arrays of one-dimensional NiO nanorods were grown on a conducting ITO film [179]. The NiO nanorod films showed excellent electrochromic performance, including high optical transmittance difference (60%), large diffusion coefficient (∼6.33 × 10−8 cm2 s−1), and very fast coloration and bleaching times (1.55 and 1.22 s).

An interesting platform of nanostructured NiO thin films is the one prepared by inkjet printing of NiO nanoparticles [181]. Inkjet printing is a low-cost and efficient technology for patterning and deposition of multilayered nanostructured materials on various substrates. The printed NiO film, with nine printed layers, exhibited an optical contrast of 64.2% at 550 nm and a coloration efficiency of 136.7 cm2 C−1. Moreover, by doping NiO with other elements, such as Cu [182], Al [176], and Co [180], the material electrochromic performance was greatly enhanced. As shown in Figure 12A–F, with an increase in Co doping (from 0 to 3%), the thickness of the formed NiO film becomes thin and the pore diameter of the nanoflakes decreases [180]. However, nanoflake-shaped NiO films cannot form when the amount of Co doping reaches 3%. This suggests that an optimum Co doping concentration is required for electrochromic applications. In this regard, the 1% Co-doped NiO nanoflake array film exhibited an outstanding optical contrast of 88.3% at 550 nm. Most recently, Liang et al. developed a self-templating method to build up metal-organic framework–derived hierarchical-porous carbon-embedded nickel oxide nanoparticle (NiO@C) films for high-performance electrochromism [62]. Figure 12G illustrates the preparation procedures of the hierarchical-porous NiO@C electrodes. As a result, the hierarchical-porous NiO@C electrodes with both good ion diffusion and electrical conductivity were achieved. This electrode showed a fast switching speed (0.46/0.25 s for coloring/bleaching), high coloration efficiency (113.5 cm2 C−1), and excellent cycling stability (90.1% after 20,000 cycles).

Figure 12: 
Characterization of NiO-based electrochromic films.
Scanning electron microscope (SEM) images of the films: (A and B) undoped NiO, (C) NiO–Co 0.3%, (D) NiO–Co 0.5%, (E) NiO–Co 1%, and (F) NiO–Co 3% (cross-sectional view is presented in the inset). Reproduced with permission from Ref. [180]. Copyright 2014, Royal Society of Chemistry. (G) Preparation procedures of the hierarchical-porous NiO@C electrode. Reproduced with permission from Ref. [62]. Copyright 2019, Royal Society of Chemistry.
Figure 12:

Characterization of NiO-based electrochromic films.

Scanning electron microscope (SEM) images of the films: (A and B) undoped NiO, (C) NiO–Co 0.3%, (D) NiO–Co 0.5%, (E) NiO–Co 1%, and (F) NiO–Co 3% (cross-sectional view is presented in the inset). Reproduced with permission from Ref. [180]. Copyright 2014, Royal Society of Chemistry. (G) Preparation procedures of the hierarchical-porous NiO@C electrode. Reproduced with permission from Ref. [62]. Copyright 2019, Royal Society of Chemistry.

4.6 Cobalt oxides

Cobalt oxide normally exists in three different polymorphs: CoO, Co3O4, and Co2O3 [183]. Among them, the crystalline Co3O4 phase has been widely studied because of its thermodynamic stability and electrochemical properties [184]. Co3O4 exhibits a reversible color change between a yellow-brown and dark brown under potential cycling [185]. Similar to other metal oxides suffering the large-volume expansion during the coloration and bleaching processes, Co3O4 suffers from structural decomposition, poor stability, and cation kinetics. To minimize such shortcomings, several reports described the growth of nanostructured Co3O4 thin films for electrochromic applications. Xia et al. demonstrated a facile method for synthesizing ordered bowl-like Co3O4 arrays [186]. The as-prepared Co3O4 films had a hierarchical porous structure (Figure 13A) when annealed at 200 °C, and it exhibits reversible color switchings between dark gray and pale yellow (Figure 13B). The coloration efficiency was calculated to be 29 cm2 C−1 at 633 nm, with an optical contrast of up to 33%. Notably, owing to the microporous structures, the bowl-like Co3O4 arrays are found to have good electrochemical stability.

Figure 13: 
Characterization of Co3O4-based electrochromic films.
(A) Scanning electron microscope (SEM) images of the Co3O4 film annealed at 200 °C for 1 h. (B) Photographs of the Co3O4 electrode with a size of 1.5 × 1.5 cm2 in the colored and bleached states. Reproduced with permission from Ref. [186]. Copyright 2010, American Chemical Society.
Figure 13:

Characterization of Co3O4-based electrochromic films.

(A) Scanning electron microscope (SEM) images of the Co3O4 film annealed at 200 °C for 1 h. (B) Photographs of the Co3O4 electrode with a size of 1.5 × 1.5 cm2 in the colored and bleached states. Reproduced with permission from Ref. [186]. Copyright 2010, American Chemical Society.

The aforementioned strategies for developing nanostructured inorganic electrochromic materials are mainly focused on the solution process techniques. This nanofabrication process has several advantages, such as low-cost, ease of designing nanostructures, and fast switching times which allow the fabrication of electrochromic material having kinetics compared to the vacuum deposition methods. To further enhance their electrochemical stability, future research needs to be focused on the interface engineering of the solution-processed nanostructured films, via strengthening the adhesion between the nanostructural electrochromic layers and the substrates. Notwithstanding, there is still a need for innovative designs of high-performance nanostructural inorganic electrochromic films to make the electrochromic device more practical for light control applications.

5 Outlook and future challenges

This review summarizes the recent progress of nanostructured inorganic electrochromic materials, which we hope to be helpful to those scientists who have recently entered this filed. The classification of common electrochromic materials is outlined as inorganic, organic, inorganic–organic hybrids, and plasmonic materials. The evaluation metrics for electrochromic materials, including color and optical contrast, coloration efficiency, switching time, and cycling stability are discussed. Electrochromic properties and synthetic methods for different nanostructured electrochromic materials are evaluated based on structure/morphology engineering, doping techniques, and crystal phase designs. Although the field of nanostructured inorganic electrochromic materials has witnessed rapid development and remarkable achievements, there are still some major challenges that hinder the high performance of electrochromic materials and their real-world applications: 1. Most of the electrochromic performance tests were conducted based on a single electrode but not a device. However, there are enormous differences between a single electrode and a device in terms of electrochromic performance. Because a device is required for practical use, future research of electrochromism should be more focused on the assembly of an electrochromic device, especially large area devices as the size exerts a significant effect on both the switching times and color uniformity. 2. The practical use of electrochromic materials requires long-term durability favoring innovative designs of robust nanostructured inorganic electrochromic layers. As such, a better understanding of the interface engineering between the electrochromic layers and the substrates is critical to promote performance and to enhance stability. 3. The multifunctional electrochromic devices (e.g., electrochromic supercapacitors [5], [ 85], electrochromic batteries [2], [ 3], and electrochromic fibers [187], [ 188]) are poised for future development of novel electrochromic devices. Exploring nanostructured material functionalities will provide additional avenues for multifunctional electrochromic devices.

Such key issues illustrate the importance of understanding electrochromic materials on the nanoscale and processing compatibility of electrochromic devices on a large scale. With the recent achievement regarding the designs of nanostructured inorganic electrochromic materials, future efforts should demonstrate the electrochromic devices to meet real-world applications, such as durable smart windows, multicolor inorganic displays, and color-tunable optical devices.


Corresponding authors: Haizeng Li and Abdulhakem Y. Elezzabi, Department of Electrical and Computer Engineering, Ultrafast Optics and Nanophotonics Laboratory, University of Alberta, Edmonton, Alberta, T6G 2V4, Canada, E-mail: (H. Li), (A. Y. Elezzabi)

Funding source: Natural Sciences and Engineering Research

Award Identifier / Grant number: CRDPJ 509210-17

Funding source: Alberta Innovates

Funding source: All Weather Windows Ltd

Acknowledgments

This work was supported by the Natural Sciences and Engineering Research Council of Canada (Grant File No: CRDPJ 509210-17), Alberta Innovates, and All Weather Windows Ltd.

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This work was supported by the Natural Sciences and Engineering Research Council of Canada (Grant File No: CRDPJ 509210-17), Alberta Innovates, and All Weather Windows Ltd.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

[1] B. A. Korgel, “Composite for smarter windows,” Nature, vol. 500, pp. 278–279, 2013, https://doi.org/10.1038/500278a.Search in Google Scholar PubMed

[2] H. Z. Li, C. J. Firby, and A. Y. Elezzabi, “Rechargeable aqueous hybrid Zn2+/Al3+ electrochromic batteries,” Joule, vol. 3, pp. 2268–2278, 2019, https://doi.org/10.1016/j.joule.2019.06.021.Search in Google Scholar

[3] H. Z. Li, L. McRae, C. J. Firby, and A. Y. Elezzabi, “Rechargeable aqueous electrochromic batteries utilizing Ti-substituted tungsten molybdenum oxide based Zn2+ ion intercalation cathodes,” Adv. Mater., vol. 31, p. 1807065, 2019, https://doi.org/10.1002/adma.201807065.Search in Google Scholar PubMed

[4] S. L. Zhang, S. Cao, T. R. Zhang, et al.., “Overcoming the technical challenges in Al anode-based electrochromic energy storage windows,” Small Methods, vol. 4, p. 1900545, 2020, https://doi.org/10.1002/smtd.201900545.Search in Google Scholar

[5] J. Chen, Z. Wang, Z. G. Chen, S. Cong, and Z. G. Zhao, “Fabry-Perot cavity-type electrochromic supercapacitors with exceptionally versatile color tunability,” Nano Lett., vol. 20, pp. 1915–1922, 2020, https://doi.org/10.1021/acs.nanolett.9b05152.Search in Google Scholar PubMed

[6] C. J. Barile, D. J. Slotcavage, J. Y. Hou, M. T. Strand, T. S. Hernandez, and M. D. McGehee, “Dynamic windows with neutral color, high contrast, and excellent durability using reversible metal electrodeposition,” Joule, vol. 1, pp. 133–145, 2017, https://doi.org/10.1016/j.joule.2017.06.001.Search in Google Scholar

[7] D. D. Miller, J. Y. Li, S. M. Islam, J. F. Jeanetta, and C. J. Barile, “Aqueous alkaline electrolytes for dynamic windows based on reversible metal electrodeposition with improved durability,” J. Mater. Chem. C, vol. 8, pp. 1826–1834, 2020, https://doi.org/10.1039/c9tc05546k.Search in Google Scholar

[8] S. Zhang, S. Cao, T. Zhang, and J. Y. Lee, “Plasmonic Oxygen-deficient TiO2-x nanocrystals for dual-band electrochromic smart windows with efficient energy recycling,” Adv. Mater., p. 2004686, 2020, https://doi.org/10.1002/adma.202004686.Search in Google Scholar PubMed

[9] K. Bange and T. Gambke, “Electrochromic materials for optical switching devices,” Adv. Mater., vol. 2, pp. 10–16, 1990, https://doi.org/10.1002/adma.19900020103.Search in Google Scholar

[10] Y. Y. Wang, S. Wang, X. J. Wang, et al.., “A multicolour bistable electronic shelf label based on intramolecular proton-coupled electron transfer,” Nat. Mater., vol. 18, pp. 1335–1342, 2019, https://doi.org/10.1038/s41563-019-0471-8.Search in Google Scholar PubMed

[11] C. M. Amb, A. L. Dyer, and J. R. Reynolds, “Navigating the color palette of solution-processable electrochromic polymers,” Chem. Mater., vol. 23, pp. 397–415, 2011, https://doi.org/10.1021/cm1021245.Search in Google Scholar

[12] P. J. Shi, C. M. Amb, E. P. Knott, et al.., “Broadly absorbing black to transmissive switching electrochromic polymers,” Adv. Mater., vol. 22, pp. 4949–4953, 2010, https://doi.org/10.1002/adma.201002234.Search in Google Scholar PubMed

[13] X. Wang, K. Chen, L. S. de Vasconcelos, et al.., “Mechanical breathing in organic electrochromics,” Nat. Commun., vol. 11, p. 211, 2020, https://doi.org/10.1038/s41467-019-14047-8.Search in Google Scholar PubMed PubMed Central

[14] S. Xiong, Y. Wang, X. Wang, et al.., “Schiff base type conjugated organic framework nanofibers: solvothermal synthesis and electrochromic properties,” Sol. Energy Mater. Sol. Cells, vol. 209, p. 110438, 2020, https://doi.org/10.1016/j.solmat.2020.110438.Search in Google Scholar

[15] F. S. Han, M. Higuchi, and D. G. Kurth, “Metallo-supramolecular polymers based on functionalized bis-terpyridines as novel electrochromic materials,” Adv. Mater., vol. 19, pp. 3928–3931, 2007, https://doi.org/10.1002/adma.200700931.Search in Google Scholar

[16] W. Zhang, H. Li, W. W. Yu, and A. Y. Elezzabi, “Transparent inorganic multicolour displays enabled by zinc-based electrochromic devices,” Light Sci. Appl., vol. 9, p. 121, 2020, https://doi.org/10.1038/s41377-020-00366-9.Search in Google Scholar PubMed PubMed Central

[17] Z. Wang, X. Y. Wang, S. Cong, et al.., “Towards full-colour tunability of inorganic electrochromic devices using ultracompact fabry-perot nanocavities,” Nat. Commun., vol. 11, p. 302, 2020, https://doi.org/10.1038/s41467-019-14194-y.Search in Google Scholar PubMed PubMed Central

[18] H. Eric, H. Z. Li, and A. Y. Elezzabi, “Rechargeable Zn2+/Al3+ dual-ion electrochromic device with long life time utilizing dimethyl sulfoxide (DMSO)-nanocluster modified hydrogel electrolytes,” RSC Adv., vol. 9, pp. 32047–32057, 2019, https://doi.org/10.1039/c9ra06785j.Search in Google Scholar PubMed PubMed Central

[19] L. L. Xie, S. W. Zhao, Y. Zhu, et al.., “High performance and excellent stability of all-solid-state electrochromic devices based on a Li1.85AlOZ ion conducting layer,” ACS Sustain. Chem. Eng., vol. 7, pp. 17390–17396, 2019, https://doi.org/10.1021/acssuschemeng.9b04501.Search in Google Scholar

[20] H. Yu, J. J. Guo, C. Wang, et al.., “High performance in electrochromic amorphous WOx film with long-term stability and tunable switching times via Al/Li-ions intercalation/deintercalation,” Electrochim. Acta, vol. 318, pp. 644–650, 2019, https://doi.org/10.1016/j.electacta.2019.06.129.Search in Google Scholar

[21] S. L. Zhang, Y. Li, T. R. Zhang, et al.., “Dual-band electrochromic devices with a transparent conductive capacitive charge-balancing anode,” ACS Appl. Mater. Interfaces, vol. 11, pp. 48062–48070, 2019, https://doi.org/10.1021/acsami.9b17678.Search in Google Scholar PubMed

[22] W. Zhang, H. Z. Li, C. J. Firby, M. Al-Hussein, and A. Y. Elezzabi, “Oxygen-vacancy-tunable electrochemical properties of electrodeposited molybdenum oxide films,” ACS Appl. Mater. Interfaces, vol. 11, pp. 20378–20385, 2019, https://doi.org/10.1021/acsami.9b04386.Search in Google Scholar PubMed

[23] Z. Q. Tong, N. Li, H. M. Lv, et al.., “Annealing synthesis of coralline V2O5 nanorod architecture for multicolor energy-efficient electrochromic device,” Sol. Energy Mater. Sol. Cells, vol. 146, pp. 135–143, 2016, https://doi.org/10.1016/j.solmat.2015.11.008.Search in Google Scholar

[24] K. Zhang, N. Li, X. X. Ma, et al.., “Building ultrathin polyaniline encapsulated V2O5 heterogeneous nanowires and its electrochromic performance,” J. Electroanal. Chem., vol. 825, pp. 16–21, 2018, https://doi.org/10.1016/j.jelechem.2018.08.001.Search in Google Scholar

[25] H. Z. Li, L. McRae, C. J. Firby, M. Al-Hussein, and A. Y. Elezzabi, “Nanohybridization of molybdenum oxide with tungsten molybdenum oxide nanowires for solution-processed fully reversible switching of energy storing smart windows,” Nano Energy, vol. 47, pp. 130–139, 2018, https://doi.org/10.1016/j.nanoen.2018.02.043.Search in Google Scholar

[26] G. F. Cai, P. Darmawan, X. Cheng, and P. S. Lee, “Inkjet printed large area multifunctional smart windows,” Adv. Energy Mater., vol. 7, p. 1602598, 2017, https://doi.org/10.1002/aenm.201602598.Search in Google Scholar

[27] D. Y. Ma and J. M. Wang, “Inorganic electrochromic materials based on tungsten oxide and nickel oxide nanostructures,” Sci. China Chem., vol. 60, pp. 54–62, 2017, https://doi.org/10.1007/s11426-016-0307-x.Search in Google Scholar

[28] Y. Y. Song, Z. D. Gao, J. H. Wang, X. H. Xia, and R. Lynch, “Multistage coloring electrochromic device based on TiO2 nanotube arrays modified with WO3 nanoparticles,” Adv. Funct. Mater., vol. 21, pp. 1941–1946, 2011, https://doi.org/10.1002/adfm.201002258.Search in Google Scholar

[29] G. F. Cai, J. X. Wang, and P. S. Lee, “Next-generation multifunctional electrochromic devices,” Acc. Chem. Res., vol. 49, pp. 1469–1476, 2016, https://doi.org/10.1021/acs.accounts.6b00183.Search in Google Scholar PubMed

[30] Z. Wang, X. Y. Wang, S. Cong, F. X. Geng, and Z. G. Zhao, “Fusing electrochromic technology with other advanced technologies: a new roadmap for future development,” Mater. Sci. Eng. R, vol. 140, p. 100524, 2020, https://doi.org/10.1016/j.mser.2019.100524.Search in Google Scholar

[31] Y. Ke, J. Chen, G. Lin, et al.., “Smart windows: electro‐, thermo‐, mechano‐, photochromics, and beyond,” Adv. Energy Mater., vol. 9, p. 1902066, 2019, https://doi.org/10.1002/aenm.201902066.Search in Google Scholar

[32] J. Kim, M. Rémond, D. Kim, H. Jang, and E. Kim, “Electrochromic conjugated polymers for multifunctional smart windows with integrative functionalities,” Adv. Mater. Technol., vol. 5, p. 1900890, 2020, https://doi.org/10.1002/admt.201900890.Search in Google Scholar

[33] P. H. Yang, P. Sun, and W. J. Mai, “Electrochromic energy storage devices,” Mater. Today, vol. 19, pp. 394–402, 2016, https://doi.org/10.1016/j.mattod.2015.11.007.Search in Google Scholar

[34] K. J. Patel, G. G. Bhatt, J. R. Ray, P. Suryavanshi, and C. J. Panchal, “All-inorganic solid-state electrochromic devices: a review,” J. Solid State Electrochem., vol. 21, pp. 337–347, 2017, https://doi.org/10.1007/s10008-016-3408-z.Search in Google Scholar

[35] P. Jittiarporn, S. Badilescu, M. N. Al Sawafta, L. Sikong, and V. V. Truong, “Electrochromic properties of sol-gel prepared hybrid transition metal oxides – a short review,” J. Sci. Adv. Mater. Devices, vol. 2, pp. 286–300, 2017, https://doi.org/10.1016/j.jsamd.2017.08.005.Search in Google Scholar

[36] M. Barawi, L. De Trizio, R. Giannuzzi, G. Veramonti, L. Manna, and M. Manca, “Dual band electrochromic devices based on Nb-doped TiO2 nanocrystalline electrodes,” ACS Nano, vol. 11, pp. 3576–3584, 2017, https://doi.org/10.1021/acsnano.6b06664.Search in Google Scholar PubMed

[37] X. Q. Che, Z. H. Wu, G. B. Dong, et al.., “Properties of all-thin-film glass/ITO/WO3:H/Ta2O5/NiOx/ITO electrochromic devices prepared by magnetron sputtering,” Thin Solid Films, vol. 662, pp. 6–12, 2018, https://doi.org/10.1016/j.tsf.2018.07.005.Search in Google Scholar

[38] W. J. Dong, Y. Lv, L. L. Xiao, Y. Fan, N. Zhang, and X. Y. Liu, “Bifunctional MoO3-WO3/Ag/MoO3-WO3 Films for efficient ITO-free electrochromic devices,” ACS Appl. Mater. Interfaces, vol. 8, pp. 33842–33847, 2016, https://doi.org/10.1021/acsami.6b12346.Search in Google Scholar PubMed

[39] G. K. Ong, C. A. S. Cabezas, M. N. Dominguez, S. L. Skjaervo, S. Heo, and D. J. Milliron, “Electrochromic niobium oxide nanorods,” Chem. Mater., vol. 32, pp. 468–475, 2020, https://doi.org/10.1021/acs.chemmater.9b04061.Search in Google Scholar

[40] X. Zhang, Y. L. Tian, W. J. Li, et al.., “Preparation and performances of all-solid-state variable infrared emittance devices based on amorphous and crystalline WO3 electrochromic thin films,” Sol. Energy Mater. Sol. Cells, vol. 200, p. 109916, 2019, https://doi.org/10.1016/j.solmat.2019.109916.Search in Google Scholar

[41] Q. Zhao, Y. S. Fang, K. Qiao, W. Wei, Y. J. Yao, and Y. F. Gao, “Printing of WO3/ITO nanocomposite electrochromic smart windows,” Sol. Energy Mater. Sol. Cells, vol. 194, pp. 95–102, 2019, https://doi.org/10.1016/j.solmat.2019.02.002.Search in Google Scholar

[42] C. G. Granqvist, “Electrochromic materials: microstructure, electronic bands, and optical properties,” Appl. Phys. A, vol. 57, pp. 3–12, 1993, https://doi.org/10.1007/bf00331209.Search in Google Scholar

[43] S. Cong, F. X. Geng, and Z. G. Zhao, “Tungsten oxide materials for optoelectronic applications,” Adv. Mater., vol. 28, pp. 10518–10528, 2016, https://doi.org/10.1002/adma.201601109.Search in Google Scholar PubMed

[44] X. F. Li, K. Perera, J. Z. He, A. Gumyusenge, and J. G. Mei, “Solution-processable electrochromic materials and devices: roadblocks and strategies towards large-scale applications,” J. Mater. Chem. C, vol. 7, pp. 12761–12789, 2019, https://doi.org/10.1039/c9tc02861g.Search in Google Scholar

[45] E. L. Runnerstrom, A. Llordes, S. D. Lounis, and D. J. Milliron, “Nanostructured electrochromic smart windows: traditional materials and NIR-selective plasmonic nanocrystals,” Chem. Commun., vol. 50, pp. 10555–10572, 2014, https://doi.org/10.1039/c4cc03109a.Search in Google Scholar PubMed

[46] C. G. Granqvist, “Electrochromics for smart windows: oxide-based thin films and devices,” Thin Solid Films, vol. 564, pp. 1–38, 2014, https://doi.org/10.1016/j.tsf.2014.02.002.Search in Google Scholar

[47] S. L. Zhang, S. Cao, T. R. Zhang, A. Fisher, and J. Y. Lee, “Al3+ intercalation/de-intercalation-enabled dual-band electrochromic smart windows with a high optical modulation, quick response and long cycle life,” Energy Environ. Sci., vol. 11, pp. 2884–2892, 2018, https://doi.org/10.1039/c8ee01718b.Search in Google Scholar

[48] W. Zhang, H. Z. Li, M. Al-Hussein, and A. Y. Elezzabi, “Electrochromic battery displays with energy retrieval functions using solution-processable colloidal vanadium oxide nanoparticles,” Adv. Opt. Mater., vol. 8, p. 1901224, 2020, https://doi.org/10.1002/adom.201901224.Search in Google Scholar

[49] S. Bulja, R. Kopf, K. Nolan, et al.., “Tuneable dielectric and optical characteristics of tailor-made inorganic electro-chromic materials,” Sci. Rep., vol. 7, p. 13484, 2017, https://doi.org/10.1038/s41598-017-13941-9.Search in Google Scholar PubMed PubMed Central

[50] D. T. Gillaspie, R. C. Tenent, and A. C. Dillon, “Metal-oxide films for electrochromic applications: present technology and future directions,” J. Mater. Chem., vol. 20, pp. 9585–9592, 2010, https://doi.org/10.1039/c0jm00604a.Search in Google Scholar

[51] S. A. Majeed, B. Ghazal, D. E. Nevonen, et al.., “Evaluation of the intramolecular charge-transfer properties in solvatochromic and electrochromic zinc octa(carbazolyl)phthalocyanines,” Inorg. Chem., vol. 56, pp. 11640–11653, 2017, https://doi.org/10.1021/acs.inorgchem.7b01570.Search in Google Scholar PubMed

[52] Y. S. Nam, H. Park, A. P. Magyar, D. S. Yun, T. S. Pollom, and A. M. Belcher, “Virus-templated iridium oxide-gold hybrid nanowires for electrochromic application,” Nanoscale, vol. 4, pp. 3405–3409, 2012, https://doi.org/10.1039/c2nr30115f.Search in Google Scholar PubMed

[53] Y. Ren, X. G. Zhou, H. Zhang, L. Lei, and G. Y. Zhao, “Preparation of a porous NiO array-patterned film and its enhanced electrochromic performance,” J. Mater. Chem. C, vol. 6, pp. 4952–4958, 2018, https://doi.org/10.1039/c8tc00367j.Search in Google Scholar

[54] N. Sakai, Y. Ebina, K. Takada, and T. Sasaki, “Electrochromic films composed of MnO2 nanosheets with controlled optical density and high coloration efficiency,” J. Electrochem. Soc., vol. 152, pp. E384–E389, 2005, https://doi.org/10.1149/1.2104227.Search in Google Scholar

[55] H. Yang, J. H. Yu, H. J. Seo, R. H. Jeong, and J. H. Boo, “Improved electrochromic properties of nanoporous NiO film by NiO flake with thickness controlled by aluminum,” Appl. Surf. Sci., vol. 461, pp. 88–92, 2018, https://doi.org/10.1016/j.apsusc.2018.05.231.Search in Google Scholar

[56] J. Zhou, Y. Wei, G. Luo, J. Zheng, and C. Xu, “Electrochromic properties of vertically aligned Ni-doped WO3 nanostructure films and their application in complementary electrochromic devices,” J. Mater. Chem. C, vol. 4, pp. 1613–1622, 2016, https://doi.org/10.1039/c5tc03750f.Search in Google Scholar

[57] L. Liu, K. Du, Z. He, et al.., “High-temperature adaptive and robust ultra-thin inorganic all-solid-state smart electrochromic energy storage devices,” Nano Energy, vol. 62, pp. 46–54, 2019, https://doi.org/10.1016/j.nanoen.2019.04.079.Search in Google Scholar

[58] J. Guo, M. Wang, G. Dong, et al.., “Mechanistic insights into the coloration, evolution, and degradation of NiOx electrochromic anodes,” Inorg. Chem., vol. 57, pp. 8874–8880, 2018, https://doi.org/10.1021/acs.inorgchem.8b00793.Search in Google Scholar PubMed

[59] Q. R. Liu, Q. Q. Chen, Q. Q. Zhang, et al.., “In situ electrochromic efficiency of a nickel oxide thin film: origin of electrochemical process and electrochromic degradation,” J. Mater. Chem. C, vol. 6, pp. 646–653, 2018, https://doi.org/10.1039/c7tc04696k.Search in Google Scholar

[60] D. Ma, G. Shi, H. Wang, Q. Zhang, and Y. Li, “Hierarchical NiO microflake films with high coloration efficiency, cyclic stability and low power consumption for applications in a complementary electrochromic device,” Nanoscale, vol. 5, pp. 4808–4815, 2013, https://doi.org/10.1039/c3nr00887h.Search in Google Scholar PubMed

[61] I. Bouessaya, A. Rougier, P. Poizot, J. Moscovici, A. Michalowicz, and J. M. Tarascon, “Electrochromic degradation in nickel oxide thin film: a self-discharge and dissolution phenomenon,” Electrochim. Acta, vol. 50, pp. 3737–3745, 2005, https://doi.org/10.1016/j.electacta.2005.01.020.Search in Google Scholar

[62] H. Liang, R. Li, C. Li, et al.., “Regulation of carbon content in MOF-derived hierarchical-porous NiO@C films for high-performance electrochromism,” Mater. Horiz., vol. 6, pp. 571–579, 2019, https://doi.org/10.1039/c8mh01091a.Search in Google Scholar

[63] W. Kang, C. Yan, X. Wang, et al.., “Green synthesis of nanobelt-membrane hybrid structured vanadium oxide with high electrochromic contrast,” J. Mater. Chem. C, vol. 2, pp. 4727–4732, 2014, https://doi.org/10.1039/c4tc00158c.Search in Google Scholar

[64] J. M. Wang, L. Zhang, L. Yu, et al.., “A bi-functional device for self-powered electrochromic window and self-rechargeable transparent battery applications,” Nat. Commun., vol. 5, p. 4921, 2014, https://doi.org/10.1038/ncomms5921.Search in Google Scholar PubMed

[65] Y. Yue, H. Li, K. Li, et al.., “High-performance complementary electrochromic device based on WO3·0.33H2O/PEDOT and prussian blue electrodes,” J. Phys. Chem. Solids, vol. 110, pp. 284–289, 2017, https://doi.org/10.1016/j.jpcs.2017.06.022.Search in Google Scholar

[66] J. H. Qian, D. Y. Ma, Z. P. Xu, D. Li, and J. M. Wang, “Electrochromic properties of hydrothermally grown Prussian blue film and device,” Sol. Energy Mater. Sol. Cells, vol. 177, pp. 9–14, 2018, https://doi.org/10.1016/j.solmat.2017.08.016.Search in Google Scholar

[67] Z. Wang, Q. Z. Zhang, S. Cong, et al.., “Using intrinsic intracrystalline tunnels for near-infrared and visible-light selective electrochromic modulation,” Adv. Opt. Mater., vol. 5, p. 1700194, 2017, https://doi.org/10.1002/adom.201700194.Search in Google Scholar

[68] B. Wang, M. W. Cui, Y. F. Gao, et al.., “A long-life battery-type electrochromic window with remarkable energy storage ability,” Solar RRL, vol. 4, p. 1900425, 2020, https://doi.org/10.1002/solr.201900425.Search in Google Scholar

[69] H. Z. Li and A. Y. Elezzabi, “Simultaneously enabling dynamic transparency control and electrical energy storage via electrochromism,” Nanoscale Horiz, vol. 5, pp. 691–695, 2020, https://doi.org/10.1039/c9nh00751b.Search in Google Scholar PubMed

[70] H. Li, W. Zhang, and A. Y. Elezzabi, “Transparent zinc-mesh electrodes for solar-charging electrochromic windows,” Adv. Mater., p. 2003574, 2020, https://doi.org/10.1002/adma.202003574.Search in Google Scholar PubMed

[71] C. D. Wessells, R. A. Huggins, and Y. Cui, “Copper hexacyanoferrate battery electrodes with long cycle life and high power,” Nat. Commun., vol. 2, p. 550, 2011, https://doi.org/10.1038/ncomms1563.Search in Google Scholar PubMed

[72] C. D. Wessells, S. V. Peddada, M. T. McDowell, R. A. Huggins, and Y. Cui, “The effect of insertion species on nanostructured open framework hexacyanoferrate battery electrodes,” J. Electrochem. Soc., vol. 159, pp. A98–A103, 2012.10.1149/2.060202jesSearch in Google Scholar

[73] B. Anasori, M. R. Lukatskaya, and Y. Gogotsi, “2D metal carbides and nitrides (MXenes) for energy storage,” Nat. Rev. Mater., vol. 2, p. 16098, 2017, https://doi.org/10.1038/natrevmats.2016.98.Search in Google Scholar

[74] P. Salles, D. Pinto, K. Hantanasirisakul, K. Maleski, C. E. Shuck, and Y. Gogotsi, “Electrochromic effect in titanium carbide MXene thin films produced by dip-coating,” Adv. Funct. Mater., vol. 29, p. 1809223, 2019, https://doi.org/10.1002/adfm.201809223.Search in Google Scholar

[75] K. W. Shah, S. X. Wang, D. X. Y. Soo, and J. W. Xu, “Viologen-based electrochromic materials: from small molecules, polymers and composites to their applications,” Polymers, vol. 11, p. 1839, 2019, https://doi.org/10.3390/polym11111839.Search in Google Scholar PubMed PubMed Central

[76] E. Dmitrieva, M. Rosenkranz, Y. Alesanco, and A. Vinuales, “The reduction mechanism of p-cyanophenylviologen in PVA-borax gel polyelectrolyte-based bicolor electrochromic devices,” Electrochim. Acta, vol. 292, pp. 81–87, 2018, https://doi.org/10.1016/j.electacta.2018.09.137.Search in Google Scholar

[77] B. Gelinas, D. Das, and D. Rochefort, “Air-stable, self-bleaching electrochromic device based on viologen-and ferrocene-containing triflimide redox ionic liquids,” ACS Appl. Mater. Interfaces, vol. 9, pp. 28726–28736, 2017, https://doi.org/10.1021/acsami.7b04427.Search in Google Scholar PubMed

[78] P. Camurlu, “Polypyrrole derivatives for electrochromic applications,” RSC Adv., vol. 4, pp. 55832–55845, 2014, https://doi.org/10.1039/c4ra11827h.Search in Google Scholar

[79] M. Ghoorchian, F. Tavoli, and N. Alizadeh, “Long-term stability of nanostructured polypyrrole electrochromic devices by using deep eutectic solvents,” J. Electroanal. Chem., vol. 807, pp. 70–75, 2017, https://doi.org/10.1016/j.jelechem.2017.11.026.Search in Google Scholar

[80] W. T. Neo, Q. Ye, S. J. Chua, and J. W. Xu, “Conjugated polymer-based electrochromics: materials, device fabrication and application prospects,” J. Mater. Chem. C, vol. 4, pp. 7364–7376, 2016, https://doi.org/10.1039/c6tc01150k.Search in Google Scholar

[81] L. Gomes, A. Branco, T. Moreira, F. Feliciano, C. Pinheiro, and C. Costa, “Increasing the electrical conductivity of electrochromic PEDOT: PSS films – a comparative study,” Sol. Energy Mater. Sol. Cells, vol. 144, pp. 631–640, 2016, https://doi.org/10.1016/j.solmat.2015.10.001.Search in Google Scholar

[82] F. Hu, H. C. Peng, S. H. Zhang, Y. C. Gu, B. Yan, and S. Chen, “PEDOT nanoparticles fully covered on natural tubular clay for hierarchically porous electrochromic film,” Sol. Energy Mater. Sol. Cells, vol. 199, pp. 59–65, 2019, https://doi.org/10.1016/j.solmat.2019.04.017.Search in Google Scholar

[83] T. G. Yun, M. Park, D. H. Kim, et al.., “All-transparent stretchable electrochromic supercapacitor wearable patch device,” ACS Nano, vol. 13, pp. 3141–3150, 2019, https://doi.org/10.1021/acsnano.8b08560.Search in Google Scholar PubMed

[84] D. Levasseur, I. Mjejri, T. Rolland, and A. Rougier, “Color tuning by oxide addition in PEDOT: PSS-based electrochromic devices,” Polymers, vol. 11, p. 179, 2019, https://doi.org/10.3390/polym11010179.Search in Google Scholar PubMed PubMed Central

[85] X. L. Chen, H. J. Lin, J. Deng, et al.., “Electrochromic fiber-shaped supercapacitors,” Adv. Mater., vol. 26, pp. 8126–8132, 2014, https://doi.org/10.1002/adma.201403243.Search in Google Scholar PubMed

[86] K. L. Zhou, H. Wang, J. T. Jiu, J. B. Liu, H. Yan, and K. Suganuma, “Polyaniline films with modified nanostructure for bifunctional flexible multicolor electrochromic and supercapacitor applications,” Chem. Eng. J., vol. 345, pp. 290–299, 2018, https://doi.org/10.1016/j.cej.2018.03.175.Search in Google Scholar

[87] Q. Lu, X. Zhang, W. Cai, et al.., “Donor−acceptor conjugated polymers containing isoindigo block for novel multifunctional materials for electrochromic, resistance memory, and detector devices,” Sol. Energy Mater. Sol. Cells, vol. 200, p. 109979, 2019, https://doi.org/10.1016/j.solmat.2019.109979.Search in Google Scholar

[88] L. Zhang, B. Wang, X. Li, et al.., “Further understanding of the mechanisms of electrochromic devices with variable infrared emissivity based on polyaniline conducting polymers,” J. Mater. Chem. C, vol. 7, pp. 9878–9891, 2019, https://doi.org/10.1039/c9tc02126d.Search in Google Scholar

[89] J. Jensen, M. Hosel, I. Kim, J. S. Yu, J. Jo, and F. C. Krebs, “Fast switching ITO free electrochromic devices,” Adv. Funct. Mater., vol. 24, pp. 1228–1233, 2014, https://doi.org/10.1002/adfm.201302320.Search in Google Scholar

[90] R. Singh, J. Tharion, S. Murugan, and A. Kumar, “ITO-free solution-processed flexible electrochromic devices based on PEDOT:PSS as transparent conducting electrode,” ACS Appl. Mater. Interfaces, vol. 9, pp. 19427–19435, 2017, https://doi.org/10.1021/acsami.6b09476.Search in Google Scholar PubMed

[91] W. B. Kang, C. Y. Yan, C. Y. Foo, and P. S. Lee, “Foldable electrochromics enabled by nanopaper transfer method,” Adv. Funct. Mater., vol. 25, pp. 4203–4210, 2015, https://doi.org/10.1002/adfm.201500527.Search in Google Scholar

[92] H. Z. Li, L. McRae, and A. Y. Elezzabi, “Solution-processed interfacial PEDOT:PSS assembly into porous tungsten molybdenum oxide nanocomposite films for electrochromic applications,” ACS Appl. Mater. Interfaces, vol. 10, pp. 10520–10527, 2018, https://doi.org/10.1021/acsami.7b18310.Search in Google Scholar PubMed

[93] H. Ling, L. Liu, P. S. Lee, D. Mandler, and X. H. Lu, “Layer-by-layer assembly of PEDOT:PSS and WO3 nanoparticles: enhanced electrochromic coloration efficiency and mechanism studies by scanning electrochemical microscopy,” Electrochim. Acta, vol. 174, pp. 57–65, 2015, https://doi.org/10.1016/j.electacta.2015.05.147.Search in Google Scholar

[94] H. Ling, J. L. Lu, S. Phua, et al.., “One-pot sequential electrochemical deposition of multilayer poly(3,4-ethylenedioxythiophene):poly(4-styrenesulfonic acid)/tungsten trioxide hybrid films and their enhanced electrochromic properties,” J. Mater. Chem., vol. 2, pp. 2708–2717, 2014, https://doi.org/10.1039/c3ta14781a.Search in Google Scholar

[95] D. Y. Ma, G. Y. Shi, H. Z. Wang, Q. H. Zhang, and Y. G. Li, “Controllable growth of high-quality metal oxide/conducting polymer hierarchical nanoarrays with outstanding electrochromic properties and solar-heat shielding ability,” J. Mater. Chem., vol. 2, pp. 13541–13549, 2014, https://doi.org/10.1039/c4ta01722f.Search in Google Scholar

[96] C. J. Barile, D. J. Slotcavage, and M. D. McGehee, “Polymer–nanoparticle electrochromic materials that selectively modulate visible and near-infrared light,” Chem. Mater., vol. 28, pp. 1439–1445, 2016, https://doi.org/10.1021/acs.chemmater.5b04811.Search in Google Scholar

[97] S. X. Xiong, S. Y. Yin, Y. Y. Wang, et al.., “Organic/inorganic electrochromic nanocomposites with various interfacial interactions: a review,” Mater. Sci. Eng. B, vol. 221, pp. 41–53, 2017, https://doi.org/10.1016/j.mseb.2017.03.017.Search in Google Scholar

[98] Z. L. Yang, H. T. Pu, and J. L. Yin, “Preparation and electrochromic property of covalently bonded WO3/polyvinylimidazole core-shell microspheres,” J. Colloid Interface Sci., vol. 292, pp. 108–112, 2005, https://doi.org/10.1016/j.jcis.2005.08.017.Search in Google Scholar PubMed

[99] D. M. DeLongchamp and P. T. Hammond, “Multiple-color electrochromism from layer-by-layer-assembled polyaniline/Prussian Blue nanocomposite thin films,” Chem. Mater., vol. 16, pp. 4799–4805, 2004, https://doi.org/10.1021/cm0496624.Search in Google Scholar

[100] W. S. Liu, X. Y. Zhang, J. Q. Liu, et al.., “Electrochromic properties of organic-inorganic composite materials,” J. Alloys Compd., vol. 718, pp. 379–385, 2017, https://doi.org/10.1016/j.jallcom.2017.05.222.Search in Google Scholar

[101] M. Lahav and M. E. van der Boom, “Polypyridyl metallo-organic assemblies for electrochromic applications,” Adv. Mater., vol. 30, p. 1706641, 2018, https://doi.org/10.1002/adma.201706641.Search in Google Scholar PubMed

[102] A. Maier, A. R. Rabindranath, and B. Tieke, “Coordinative supramolecular assembly of electrochromic films based on metal ion complexes of polyiminofluorene with terpyridine substituent groups,” Chem. Mater., vol. 21, pp. 3668–3676, 2009, https://doi.org/10.1021/cm901158t.Search in Google Scholar

[103] L. Motiei, M. Lahav, D. Freeman, and M. E. van der Boom, “Electrochromic behavior of a self-propagating molecular-based assembly,” J. Am. Chem. Soc., vol. 131, pp. 3468–3469, 2009, https://doi.org/10.1021/ja900507c.Search in Google Scholar PubMed

[104] S. Pai, M. Moos, M. H. Schreck, C. Lambert, and D. G. Kurth, “Green-to-red electrochromic Fe(II) metallo-supramolecular polyelectrolytes self-assembled from fluorescent 2,6-bis(2-pyridyl)pyrinnidine bithiophene,” Inorg. Chem., vol. 56, pp. 1418–1432, 2017, https://doi.org/10.1021/acs.inorgchem.6b02496.Search in Google Scholar PubMed

[105] M. K. Bera, C. Chakraborty, U. Rana, and M. Higuchi, “Electrochromic Os(II)-based metallo-supramolecular polymers,” Macromol. Rapid Commun., vol. 39, p. 1800415, 2018, https://doi.org/10.1002/marc.201800415.Search in Google Scholar PubMed

[106] T. R. Gordon, T. Paik, D. R. Klein, et al.., “Shape-dependent plasmonic response and directed self-assembly in a new semiconductor building block, indium-doped cadmium oxide (ICO),” Nano Lett., vol. 13, pp. 2857–2863, 2013, https://doi.org/10.1021/nl4012003.Search in Google Scholar PubMed

[107] M. Kanehara, H. Koike, T. Yoshinaga, and T. Teranishi, “Indium tin oxide nanoparticles with compositionally tunable surface plasmon resonance frequencies in the near-IR region,” J. Am. Chem. Soc., vol. 131, pp. 17736–17737, 2009, https://doi.org/10.1021/ja9064415.Search in Google Scholar PubMed

[108] Y. Liu, M. Liu, and M. T. Swihart, “Plasmonic copper sulfide-based materials: a brief introduction to their synthesis, doping, alloying, and applications,” J. Phys. Chem. C, vol. 121, pp. 13435–13447, 2017, https://doi.org/10.1021/acs.jpcc.7b00894.Search in Google Scholar

[109] A. Agrawal, S. H. Cho, O. Zandi, S. Ghosh, R. W. Johns, and D. J. Milliron, “Localized surface plasmon resonance in semiconductor nanocrystals,” Chem. Rev., vol. 118, pp. 3121–3207, 2018, https://doi.org/10.1021/acs.chemrev.7b00613.Search in Google Scholar PubMed

[110] S. Araki, K. Nakamura, K. Kobayashi, A. Tsuboi, and N. Kobayashi, “Electrochemical optical-modulation device with reversible transformation between transparent, mirror, and black,” Adv. Mater., vol. 24, pp. OP122–OP126, 2012, https://doi.org/10.1002/adma.201200060.Search in Google Scholar PubMed

[111] A. Tsuboi, K. Nakamura, and N. Kobayashi, “Multicolor electrochromism showing three primary color states (cyan magenta yellow) based on size- and shape-controlled silver nanoparticles,” Chem. Mater., vol. 26, pp. 6477–6485, 2014, https://doi.org/10.1021/cm5039039.Search in Google Scholar

[112] N. Li, P. P. Wei, L. A. Yu, et al.., “Dynamically switchable multicolor electrochromic films,” Small, vol. 15, p. 1804974, 2019, https://doi.org/10.1002/smll.201804974.Search in Google Scholar PubMed

[113] B. Tandon, S. Ghosh, and D. J. Milliron, “Dopant selection strategy for high-quality factor localized surface plasmon resonance from doped metal oxide nanocrystals,” Chem. Mater., vol. 31, pp. 7752–7760, 2019, https://doi.org/10.1021/acs.chemmater.9b02917.Search in Google Scholar

[114] G. Garcia, R. Buonsanti, E. L. Runnerstrom, et al.., “Dynamically modulating the surface plasmon resonance of doped semiconductor nanocrystals,” Nano Lett., vol. 11, pp. 4415–4420, 2011, https://doi.org/10.1021/nl202597n.Search in Google Scholar PubMed

[115] Z. C. Wang and G. Chumanov, “WO3 sol-gel modified Ag nanoparticle arrays for electrochemical modulation of surface plasmon resonance,” Adv. Mater., vol. 15, pp. 1285–1289, 2003, https://doi.org/10.1002/adma.200304989.Search in Google Scholar

[116] J. L. Peng, H. H. Jeong, Q. Q. Lin, et al.., “Scalable electrochromic nanopixels using plasmonics,” Sci. Adv., vol. 5, p. eaaw2205, 2019, https://doi.org/10.1126/sciadv.aaw2205.Search in Google Scholar PubMed PubMed Central

[117] M. Gugole, O. Olsson, K. L. Xiong, et al.., “High-contrast switching of plasmonic structural colors: inorganic versus organic electrochromism,” ACS Photonics, vol. 7, pp. 1762–1772, 2020, https://doi.org/10.1021/acsphotonics.0c00394.Search in Google Scholar

[118] E. Hopmann and A. Y. Elezzabi, “Plasmochromic nanocavity dynamic light color switching,” Nano Lett., vol. 20, pp. 1876–1882, 2020, https://doi.org/10.1021/acs.nanolett.9b05088.Search in Google Scholar PubMed

[119] J. Y. Lim, H. C. Ko, and H. Lee, “Systematic prediction of maximum electrochromic contrast of an electrochromic material,” Synth. Met., vol. 155, pp. 595–598, 2005, https://doi.org/10.1016/j.synthmet.2005.09.040.Search in Google Scholar

[120] J. Padilla, A. M. Osterholm, A. L. Dyer, and J. R. Reynolds, “Process controlled performance for soluble electrochromic polymers,” Sol. Energy Mater. Sol. Cells, vol. 140, pp. 54–60, 2015, https://doi.org/10.1016/j.solmat.2015.03.018.Search in Google Scholar

[121] J. Padilla, V. Seshadri, G. A. Sotzing, and T. F. Otero, “Maximum contrast from an electrochromic material,” Electrochem. Commun., vol. 9, pp. 1931–1935, 2007, https://doi.org/10.1016/j.elecom.2007.05.004.Search in Google Scholar

[122] X. Q. Ju, F. Yang, X. Zhu, and X. L. Jia, “Zinc ion intercalation/deintercalation of metal organic framework-derived nanostructured NiO@C for low-transmittance and high-performance electrochromism,” ACS Sustain. Chem. Eng., vol. 8, pp. 12222–12229, 2020, https://doi.org/10.1021/acssuschemeng.0c03837.Search in Google Scholar

[123] M. H. Wang, Y. Chen, B. W. Gao, and H. Lei, “Electrochromic properties of nanostructured WO3 thin films deposited by glancing-angle magnetron sputtering,” Adv. Electron Mater., vol. 5, p. 1800713, 2019, https://doi.org/10.1002/aelm.201800713.Search in Google Scholar

[124] R. Ayranci, G. Baskaya, M. Guzel, et al.., “Enhanced optical and electrical properties of PEDOT via nanostructured carbon materials: a comparative investigation,” Nano Struct. Nano Objects, vol. 11, pp. 13–19, 2017, https://doi.org/10.1016/j.nanoso.2017.05.008.Search in Google Scholar

[125] M. R. J. Scherer, L. Li, P. M. S. Cunha, O. A. Scherman, and U. Steiner, “Enhanced electrochromism in gyroid-structured vanadium pentoxide,” Adv. Mater., vol. 24, pp. 1217–1221, 2012, https://doi.org/10.1002/adma.201104272.Search in Google Scholar PubMed

[126] H. Z. Li, J. W. Chen, M. Q. Cui, et al.., “Spray coated ultrathin films from aqueous tungsten molybdenum oxide nanoparticle ink for high contrast electrochromic applications,” J. Mater. Chem. C, vol. 4, pp. 33–38, 2016, https://doi.org/10.1039/c5tc02802g.Search in Google Scholar

[127] D. S. Dalavi, M. J. Suryavanshi, D. S. Patil, et al.., “Nanoporous nickel oxide thin films and its improved electrochromic performance: effect of thickness,” Appl. Surf. Sci., vol. 257, pp. 2647–2656, 2011, https://doi.org/10.1016/j.apsusc.2010.10.037.Search in Google Scholar

[128] G. Tahtali, Z. Has, C. Doyranli, C. Varlikli, and S. Koyuncu, “Solution processable neutral state colourless electrochromic devices: effect of the layer thickness on the electrochromic performance,” J. Mater. Chem. C, vol. 4, pp. 10090–10094, 2016, https://doi.org/10.1039/c6tc04112d.Search in Google Scholar

[129] B. W. C. Au, K. Y. Chan, and D. Knipp, “Effect of film thickness on electrochromic performance of sol-gel deposited tungsten oxide (WO3),” Opt. Mater., vol. 94, pp. 387–392, 2019.10.1016/j.optmat.2019.05.051Search in Google Scholar

[130] F. Hu, B. Yan, E. H. Ren, et al.., “Constructing spraying-processed complementary smart windows via electrochromic materials with hierarchical nanostructures,” J. Mater. Chem. C, vol. 7, pp. 14855–14860, 2019, https://doi.org/10.1039/c9tc04204k.Search in Google Scholar

[131] W. Zhang, X. Chen, X. Wang, S. Zhu, S. Wang, and Q. Wang, “Pulsed electrodeposition of nanostructured polythiothene film for high-performance electrochromic devices,” Sol. Energy Mater. Sol. Cells, vol. 219, p. 110775, 2021, https://doi.org/10.1016/j.solmat.2020.110775.Search in Google Scholar

[132] D. R. Sahu, T. J. Wu, S. C. Wang, and J. L. Huang, “Electrochromic behavior of NiO film prepared by e-beam evaporation,” J. Sci. Adv. Mater. Devices, vol. 2, pp. 225–232, 2017, https://doi.org/10.1016/j.jsamd.2017.05.001.Search in Google Scholar

[133] I. Sorar, E. Pehlivan, G. A. Niklasson, and C. G. Granqvist, “Electrochromism of DC magnetron-sputtered TiO2: role of film thickness,” Appl. Surf. Sci., vol. 318, pp. 24–27, 2014, https://doi.org/10.1016/j.apsusc.2013.12.015.Search in Google Scholar

[134] C. G. Wu, M. I. Lu, S. J. Chang, and C. S. Wei, “A solution-processable high-coloration-efficiency low-switching-voltage electrochromic polymer based on polycyclopentadithiophene,” Adv. Funct. Mater., vol. 17, pp. 1063–1070, 2007, https://doi.org/10.1002/adfm.200600381.Search in Google Scholar

[135] D. Ma, G. Shi, H. Wang, Q. Zhang, and Y. Li, “Morphology-tailored synthesis of vertically aligned 1D WO3 nano-structure films for highly enhanced electrochromic performance,” J. Mater. Chem., vol. 1, pp. 684–691, 2013, https://doi.org/10.1039/c2ta00090c.Search in Google Scholar

[136] D. Ma, H. Wang, Q. Zhang, and Y. Li, “Self-weaving WO3 nanoflake films with greatly enhanced electrochromic performance,” J. Mater. Chem., vol. 22, pp. 16633–16639, 2012, https://doi.org/10.1039/c2jm32784h.Search in Google Scholar

[137] M. A. Arvizu, H. Y. Qu, U. Cindemir, et al.., “Electrochromic WO3 thin films attain unprecedented durability by potentiostatic pretreatment,” J. Mater. Chem., vol. 7, pp. 2908–2918, 2019, https://doi.org/10.1039/c8ta09621j.Search in Google Scholar PubMed PubMed Central

[138] R. C. D. Peres, V. F. Juliano, M. A. Depaoli, S. Panero, and B. Scrosati, “Electrochromic properties of dodecyclbenzenesulfonate doped poly(pyrrole),” Electrochim. Acta, vol. 38, pp. 869–876, 1993, https://doi.org/10.1016/0013-4686(93)87003-v.Search in Google Scholar

[139] L. P. Hao, W. Wang, H. J. Niu, and Y. Zhou, “Grafting triphenylamine groups onto polysiloxanes to improve interaction between the electrochromic films and ITO,” Electrochim. Acta, vol. 246, pp. 259–268, 2017, https://doi.org/10.1016/j.electacta.2017.06.061.Search in Google Scholar

[140] Z. H. Jiao, X. W. Sun, J. M. Wang, L. Ke, and H. V. Demir, “Hydrothermally grown nanostructured WO3 films and their electrochromic characteristics,” J. Phys. D Appl. Phys., vol. 43, p. 285501, 2010, https://doi.org/10.1088/0022-3727/43/28/285501.Search in Google Scholar

[141] T. T. Ye, Y. Sun, X. Zhao, B. P. Lin, H. Yang, and X. Q. Zhang, “Long-term-stable, solution-processable, electrochromic carbon nanotubes/polymer composite for smart supercapacitor with wide working potential window,” J. Mater. Chem., vol. 6, pp. 18994–19003, 2018, https://doi.org/10.1039/c8ta04465a.Search in Google Scholar

[142] J. M. Berak and M. J. Sienko, “Effect of oxygen-deficiency on electrical transport properties of tungsten trioxide crystals,” J. Solid State Chem., vol. 2, pp. 109–133, 1970, https://doi.org/10.1016/0022-4596(70)90040-x.Search in Google Scholar

[143] L. Y. Su, L. G. Zhang, J. H. Fang, M. H. Xu, and Z. H. Lu, “Electrochromic and photoelectrochemical behavior of electrodeposited tungsten trioxide films,” Sol. Energy Mater. Sol. Cells, vol. 58, pp. 133–140, 1999, https://doi.org/10.1016/s0927-0248(98)00171-8.Search in Google Scholar

[144] P. R. Somani and S. Radhakrishnan, “Electrochromic materials and devices: present and future,” Mater. Chem. Phys., vol. 77, pp. 117–133, 2003, https://doi.org/10.1016/s0254-0584(01)00575-2.Search in Google Scholar

[145] S. K. Deb, “A novel electrophotographic system,” Appl. Opt., vol. 8, pp. 192–195, 1969, https://doi.org/10.1364/ao.8.s1.000192.Search in Google Scholar

[146] L. Zhang, D. L. Chao, P. H. Yang, et al.., “Flexible pseudocapacitive electrochromics via inkjet printing of additive-free tungsten oxide nanocrystal ink,” Adv. Energy Mater., vol. 10, p. 2000142, 2020, https://doi.org/10.1002/aenm.202000142.Search in Google Scholar

[147] J. Zhang, J. P. Tu, X. H. Xia, X. L. Wang, and C. D. Gu, “Hydrothermally synthesized WO3 nanowire arrays with highly improved electrochromic performance,” J. Mater. Chem., vol. 21, pp. 5492–5498, 2011, https://doi.org/10.1039/c0jm04361c.Search in Google Scholar

[148] A. Azam, J. Kim, J. Park, et al.., “Two-dimensional WO3 nanosheets chemically converted from layered WS2 for high-performance electrochromic devices,” Nano Lett., vol. 18, pp. 5646–5651, 2018, https://doi.org/10.1021/acs.nanolett.8b02150.Search in Google Scholar PubMed

[149] S. Adhikari and D. Sarkar, “High efficient electrochromic WO3 nanofibers,” Electrochim. Acta, vol. 138, pp. 115–123, 2014, https://doi.org/10.1016/j.electacta.2014.06.062.Search in Google Scholar

[150] Y. Q. Li, W. A. McMaster, H. Wei, D. H. Chen, and R. A. Caruso, “Enhanced electrochromic properties of WO3 nanotree-like structures synthesized via a two-step solvothermal process showing promise for electrochromic window application,” ACS Appl. Nano Mater., vol. 1, pp. 2552–2558, 2018, https://doi.org/10.1021/acsanm.8b00190.Search in Google Scholar

[151] S. Cong, Y. Y. Tian, Q. W. Li, Z. G. Zhao, and F. X. Geng, “Single-crystalline tungsten oxide quantum dots for fast pseudocapacitor and electrochromic applications,” Adv. Mater., vol. 26, pp. 4260–4267, 2014, https://doi.org/10.1002/adma.201400447.Search in Google Scholar PubMed

[152] Y. J. Yao, Q. Zhao, W. Wei, et al.., “WO3 quantum-dots electrochromism,” Nano Energy, vol. 68, p. 104350, 2020, https://doi.org/10.1016/j.nanoen.2019.104350.Search in Google Scholar

[153] H. Z. Li, G. Y. Shi, H. Z. Wang, Q. H. Zhang, and Y. G. Li, “Self-seeded growth of nest-like hydrated tungsten trioxide film directly on FTO substrate for highly enhanced electrochromic performance,” J. Mater. Chem., vol. 2, pp. 11305–11310, 2014, https://doi.org/10.1039/c4ta01803f.Search in Google Scholar

[154] S. R. Bathe and P. S. Patil, “Influence of Nb doping on the electrochromic properties of WO3 films,” J. Phys. D Appl. Phys., vol. 40, pp. 7423–7431, 2007, https://doi.org/10.1088/0022-3727/40/23/025.Search in Google Scholar

[155] S. J. Xie, Z. J. Bi, Y. B. Chen, et al.., “Electrodeposited Mo-doped WO3 film with large optical modulation and high areal capacitance toward electrochromic energy-storage applications,” Appl. Surf. Sci., vol. 459, pp. 774–781, 2018, https://doi.org/10.1016/j.apsusc.2018.08.045.Search in Google Scholar

[156] B. R. Koo, K. H. Kim, and H. J. Ahn, “Switching electrochromic performance improvement enabled by highly developed mesopores and oxygen vacancy defects of Fe-doped WO3 films,” Appl. Surf. Sci., vol. 453, pp. 238–244, 2018, https://doi.org/10.1016/j.apsusc.2018.05.094.Search in Google Scholar

[157] Y. Zhan, M. R. J. Tan, X. Cheng, et al.., “Ti-doped WO3 synthesized by a facile wet bath method for improved electrochromism,” J. Mater. Chem. C, vol. 5, pp. 9995–10000, 2017, https://doi.org/10.1039/c7tc02456h.Search in Google Scholar

[158] M. A. Arvizu, G. A. Niklasson, and C. G. Granqvist, “Electrochromic W1–x–yTixMoyO3 thin films made by sputter deposition: large optical modulation, good cycling durability, and approximate color neutrality,” Chem. Mater., vol. 29, pp. 2246–2253, 2017, https://doi.org/10.1021/acs.chemmater.6b05198.Search in Google Scholar

[159] A. Arash, S. A. Tawfik, M. J. S. Spencer, et al.., “Electrically activated UV-A filters based on electrochromic MoO3-x,” ACS Appl. Mater. Interfaces, vol. 12, pp. 16997–17003, 2020, https://doi.org/10.1021/acsami.9b20916.Search in Google Scholar PubMed

[160] S. Santhosh, M. Mathankumar, S. S. Chandrasekaran, A. K. N. Kumar, P. Murugan, and B. Subramanian, “Effect of ablation rate on the microstructure and electrochromic properties of pulsed-laser-deposited molybdenum oxide thin films,” Langmuir, vol. 33, pp. 19–33, 2017, https://doi.org/10.1021/acs.langmuir.6b02940.Search in Google Scholar PubMed

[161] L. Zheng, Y. Xu, D. Jin, and Y. Xie, “Novel metastable hexagonal MoO3 nanobelts: synthesis, photochromic, and electrochromic properties,” Chem. Mater., vol. 21, pp. 5681–5690, 2009, https://doi.org/10.1021/cm9023887.Search in Google Scholar

[162] T. G. Novak, J. Kim, A. P. Tiwari, et al.., “2D MoO3 nanosheets synthesized by exfoliation and oxidation of MoS2 for high contrast and fast response time electrochromic devices,” ACS Sustain. Chem. Eng., vol. 8, pp. 11276–11282, 2020, https://doi.org/10.1021/acssuschemeng.0c03191.Search in Google Scholar

[163] N. N. Dinh, N. T. T. Oanh, P. D. Long, M. C. Bernard, and A. H. L. Goff, “Electrochromic properties of TiO2 anatase thin films prepared by a dipping sol-gel method,” Thin Solid Films, vol. 423, pp. 70–76, 2003, https://doi.org/10.1016/s0040-6090(02)00948-3.Search in Google Scholar

[164] E. Silik, S. Pat, S. Ozen, et al., “Electrochromic properties of TiO2 thin films grown by thermionic vacuum arc method,” Thin Solid Films, vol. 640, pp. 27–32, 2017.10.1016/j.tsf.2017.07.073Search in Google Scholar

[165] Z. Tong, S. Liu, X. Li, L. Mai, J. Zhao, and Y. Li, “Achieving rapid Li-ion insertion kinetics in TiO2 mesoporous nanotube arrays for bifunctional high-rate energy storage smart windows,” Nanoscale, vol. 10, pp. 3254–3261, 2018, https://doi.org/10.1039/c7nr07703c.Search in Google Scholar PubMed

[166] J. Z. Chen, W. Y. Ko, Y. C. Yen, P. H. Chen, and K. J. Lin, “Hydrothermally processed TiO2 nanowire electrodes with antireflective and electrochromic properties,” ACS Nano, vol. 6, pp. 6633–6639, 2012, https://doi.org/10.1021/nn300787r.Search in Google Scholar PubMed

[167] L. De Trizio, R. Buonsanti, A. M. Schimpf, et al.., “Nb-doped colloidal TiO2 nanocrystals with tunable infrared absorption,” Chem. Mater., vol. 25, pp. 3383–3390, 2013, https://doi.org/10.1021/cm402396c.Search in Google Scholar

[168] S. Cao, S. L. Zhang, T. R. Zhang, and J. Y. Lee, “Fluoride-assisted synthesis of plasmonic colloidal Ta-doped TiO2 nanocrystals for near-infrared and visible-light selective electrochromic modulation,” Chem. Mater., vol. 30, pp. 4838–4846, 2018, https://doi.org/10.1021/acs.chemmater.8b02196.Search in Google Scholar

[169] S. Cao, S. L. Zhang, T. R. Zhang, Q. F. Yao, and J. Y. Lee, “A visible light-near-infrared dual-band smart window with internal energy storage,” Joule, vol. 3, pp. 1152–1162, 2019, https://doi.org/10.1016/j.joule.2018.12.010.Search in Google Scholar

[170] C. Costa, C. Pinheiro, I. Henriques, and C. A. T. Laia, “Electrochromic properties of inkjet printed vanadium oxide gel on flexible polyethylene terephthalate/indium tin oxide electrodes,” ACS Appl. Mater. Interfaces, vol. 4, pp. 5266–5275, 2012, https://doi.org/10.1021/am301213b.Search in Google Scholar PubMed

[171] D. Wei, M. R. J. Scherer, C. Bower, P. Andrew, T. Ryhanen, and U. Steiner, “A nanostructured electrochromic supercapacitor,” Nano Lett., vol. 12, pp. 1857–1862, 2012, https://doi.org/10.1021/nl2042112.Search in Google Scholar PubMed

[172] Y. Y. Cui, Y. J. Ke, C. Liu, et al.., “Thermochromic VO2 for energy-efficient smart windows,” Joule, vol. 2, pp. 1707–1746, 2018, https://doi.org/10.1016/j.joule.2018.06.018.Search in Google Scholar

[173] I. Mjejri, M. Gaudon, G. Song, C. Labrugere, and A. Rougier, “Crystallized V2O5 as oxidized phase for unexpected multicolor electrochromism in V2O3 thick film,” ACS Appl. Energy Mater., vol. 1, pp. 2721–2729, 2018, https://doi.org/10.1021/acsaem.8b00386.Search in Google Scholar

[174] I. Mjejri and A. Rougier, “Color switching in V3O7·H2O films cycled in Li and Na based electrolytes: novel vanadium oxide based electrochromic materials,” J. Mater. Chem. C, vol. 8, pp. 3631–3638, 2020, https://doi.org/10.1039/c9tc06753a.Search in Google Scholar

[175] M. M. Margoni, S. Mathuri, K. Ramamurthi, R. R. Babu, V. Ganesh, and K. Sethuraman, “Hydrothermally grown nano and microstructured V2O5 thin films for electrochromic application,” Appl. Surf. Sci., vol. 449, pp. 193–202, 2018, https://doi.org/10.1016/j.apsusc.2018.01.288.Search in Google Scholar

[176] F. Lin, D. Nordlund, T. C. Weng, et al.., “Hole doping in Al-containing nickel oxide materials to improve electrochromic performance,” ACS Appl. Mater. Interfaces, vol. 5, pp. 301–309, 2013, https://doi.org/10.1021/am302097b.Search in Google Scholar PubMed

[177] Q. R. Liu, G. B. Dong, Y. Xiao, M. P. Delplancke-Ogletree, F. Reniers, and X. G. Diao, “Electrolytes-relevant cyclic durability of nickel oxide thin films as an ion-storage layer in an all-solid-state complementary electrochromic device,” Sol. Energy Mater. Sol. Cells, vol. 157, pp. 844–852, 2016, https://doi.org/10.1016/j.solmat.2016.07.022.Search in Google Scholar

[178] K. K. Purushothaman and G. Muralidharan, “Nanostructured NiO based all solid state electrochromic device,” J. Sol. Gel Sci. Technol., vol. 46, pp. 190–194, 2008, https://doi.org/10.1007/s10971-007-1657-0.Search in Google Scholar

[179] R. A. Patil, R. S. Devan, J. H. Lin, Y. R. Ma, P. S. Patil, and Y. Liou, “Efficient electrochromic properties of high-density and large-area arrays of one-dimensional NiO nanorods,” Sol. Energy Mater. Sol. Cells, vol. 112, pp. 91–96, 2013, https://doi.org/10.1016/j.solmat.2013.01.003.Search in Google Scholar

[180] J. H. Zhang, G. F. Cai, D. Zhou, et al.., “Co-doped NiO nanoflake array films with enhanced electrochromic properties,” J. Mater. Chem. C, vol. 2, pp. 7013–7021, 2014, https://doi.org/10.1039/c4tc01033g.Search in Google Scholar

[181] G. F. Cai, P. Darmawan, M. Q. Cui, et al.., “Inkjet-printed all solid-state electrochromic devices based on NiO/WO3 nanoparticle complementary electrodes,” Nanoscale, vol. 8, pp. 348–357, 2016, https://doi.org/10.1039/c5nr06995e.Search in Google Scholar PubMed

[182] S. Moghe, A. D. Acharya, R. Panda, et al.., “Effect of copper doping on the change in the optical absorption behaviour in NiO thin films,” Renew. Energy, vol. 46, pp. 43–48, 2012, https://doi.org/10.1016/j.renene.2012.02.028.Search in Google Scholar

[183] C. R. Dhas, R. Venkatesh, R. Sivakumar, A. M. E. Raj, and C. Sanjeeviraja, “Fast electrochromic response of porous-structured cobalt oxide (Co3O4) thin films by novel nebulizer spray pyrolysis technique,” Ionics, vol. 22, pp. 1911–1926, 2016, https://doi.org/10.1007/s11581-016-1707-0.Search in Google Scholar

[184] H. S. Shim, V. R. Shinde, H. J. Kim, Y. E. Sung, and W. B. Kim, “Porous cobalt oxide thin films from low temperature solution phase synthesis for electrochromic electrode,” Thin Solid Films, vol. 516, pp. 8573–8578, 2008, https://doi.org/10.1016/j.tsf.2008.05.055.Search in Google Scholar

[185] X. H. Xia, J. P. Tu, J. Zhang, X. H. Huang, X. L. Wang, and X. B. Zhao, “Improved electrochromic performance of hierarchically porous Co3O4 array film through self-assembled colloidal crystal template,” Electrochim. Acta, vol. 55, pp. 989–994, 2010, https://doi.org/10.1016/j.electacta.2009.09.071.Search in Google Scholar

[186] X. H. Xia, J. P. Tu, J. Zhang, J. Y. Xiang, X. L. Wang, and X. B. Zhao, “Cobalt oxide ordered bowl-like array films prepared by electrodeposition through monolayer polystyrene sphere template and electrochromic properties,” ACS Appl. Mater. Interfaces, vol. 2, pp. 186–192, 2010, https://doi.org/10.1021/am900636g.Search in Google Scholar

[187] K. R. Li, Q. H. Zhang, H. Z. Wang, and Y. G. Li, “Red, Green, Blue (RGB) Electrochromic fibers for the new smart color change fabrics,” ACS Appl. Mater. Interfaces, vol. 6, pp. 13043–13050, 2014, https://doi.org/10.1021/am502929p.Search in Google Scholar PubMed

[188] H. W. Fan, K. R. Li, X. L. Liu, et al.., “Continuously processed, long electrochromic fibers with multi-environmental stability,” ACS Appl. Mater. Interfaces, vol. 12, pp. 28451–28460, 2020, https://doi.org/10.1021/acsami.0c09589.Search in Google Scholar PubMed

Received: 2020-08-13
Accepted: 2020-11-06
Published Online: 2020-11-20

© 2020 Wu Zhang et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 29.3.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2020-0474/html
Scroll to top button