Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Mar 27;123(8):4353–4415. doi: 10.1021/acs.chemrev.2c00546

Vanadium Oxide: Phase Diagrams, Structures, Synthesis, and Applications

Peng Hu ‡,†,*, Ping Hu ¶,, Tuan Duc Vu , Ming Li , Shancheng Wang , Yujie Ke †,, Xianting Zeng §, Liqiang Mai ¶,⊥,*, Yi Long †,○,*
PMCID: PMC10141335  PMID: 36972332

Abstract

graphic file with name cr2c00546_0052.jpg

Vanadium oxides with multioxidation states and various crystalline structures offer unique electrical, optical, optoelectronic and magnetic properties, which could be manipulated for various applications. For the past 30 years, significant efforts have been made to study the fundamental science and explore the potential for vanadium oxide materials in ion batteries, water splitting, smart windows, supercapacitors, sensors, and so on. This review focuses on the most recent progress in synthesis methods and applications of some thermodynamically stable and metastable vanadium oxides, including but not limited to V2O3, V3O5, VO2, V3O7, V2O5, V2O2, V6O13, and V4O9. We begin with a tutorial on the phase diagram of the V–O system. The second part is a detailed review covering the crystal structure, the synthesis protocols, and the applications of each vanadium oxide, especially in batteries, catalysts, smart windows, and supercapacitors. We conclude with a brief perspective on how material and device improvements can address current deficiencies. This comprehensive review could accelerate the development of novel vanadium oxide structures in related applications.

1. Introduction

1.1. Vanadium

Vanadium was first discovered by Andrés Manuel del Rio in Mexico City from Pb5(VO4)3Cl in 1801.1 However, it was wrongly identified as a form of chromium by Hippolyte Victor Collet-Descotils in 1805.2 Until 1831, Swedish chemist Nil Gabriel Self-ström in Stockholm named the element vanadium, which is from the Norse Goddess Vanadis and means beauty and fertility.3 In the Earth’s crust, vanadium is the 20th most abundant element and the sixth most abundant element among the transition metals.35 However, some literature indicates that vanadium is the fourth most abundant transition metal after iron, titanium, and manganese.58 High purity vanadium (about 99.7%) was first produced in 1925 by reducing vanadium pentoxide (V2O5) with calcium metal.1 Pure vanadium exhibits a transition metal feature, which shows a high melting point and good corrosion resistance at low temperatures. Vanadium can be dissolved in nitric and sulfuric acids but is insoluble in hydrochloric acid.9 In nature, vanadium is difficult to exist in metal form because it easily reacts with oxygen, even nitrogen and carbon at elevated temperatures.2,10 Vanadium is an important component of specific steel alloys, which provides additional tensile strength and extra protection against rust and corrosion of these materials.

1.2. Vanadium Oxides

Vanadium has the electronic configuration [Ar]4s23d3. Therefore, the oxidation state of vanadium can range from +5 to −3, and the valences of +5, +4, +3, and +2 are most commonly observed.11,12 Four vanadium oxides feature single oxidation states (+2 for VO, +3 for V2O3, +4 for VO2, and +5 for V2O5), and others have mixed oxidation states. Different oxidation states exhibit various colors: +5 (orange to yellow), +4 (blue), and +3 (green).9 The vanadium oxides exhibit crystalline structures with different oxygen coordinations, which result in the formation of octahedral, pentagonal bipyramids, square pyramids, and tetrahedral sharing corners, edges, or faces.12 The oxidation state of the vanadium cations dramatically affects the physicochemical properties of the vanadium oxides with different phases.

Due to the multioxidation states and various crystalline structures, the vanadium oxides exhibit excellent intercalation properties to host–guest molecules or ions,5 giving excellent catalytic activities,4 strong electron–electron correlations,13 outstanding phase transitions (metal–insulator transition),14 and high electrical conductivity.5 Furthermore, the abundant nanostructures of vanadium oxides can be achieved by different preparation methods, which not only shorten the transportation distance of ions or electrons and yield a faster solid-state diffusion in electrochemical energy conversion systems,15,16 but also provide more active positions for the interaction with other molecules or ions and more exposed active crystal facets for catalysis applications.17,18 Thus, the vanadium oxides provide promising applications in energy conversion/saving fields,19,20 such as ion batteries,2125 water splitting,26 smart windows,27,28 supercapacitors,29,30 sensors,31 and so on.

A series of vanadium oxides with strong electron–electron correlations exhibit metal–insulator transition (MIT). The V2O3, VO2, and V2O5 with single oxidation undergo MIT at 160 K,32 340 K,33 and 530 K,34 respectively. These phase transitions are reversible and accompanied by a change of crystallographic, magnetic, optical, and electrical properties. The mixed-valence vanadium oxides belong to either Magnéli series (VnO2n-1) or Wadsley series (VnO2n+1). For the Wadsley series, V3O7 and V6O13 exhibit the phase transition at 5.2 and 155 K, respectively.35,36 Except for V7O13 (metallic), all the Magnéli series show a transition from a paramagnetic to an antiferromagnetic state and consequently exhibit an antiferromagnetic ground state at low temperatures, including V3O5 (430 K), V4O7 (250 K), V5O9 (135 K), V6O11 (170 K), and V8O15 (70 K) with different phase transition temperatures, respectively.37

1.3. Scope of the Review

Vanadium oxides have a long history and rapid development in recent years as they are one of the most promising candidates in versatile applications in batteries, energy-saving smart windows, sensors, catalysts, optoelectronic devices, and so on. Therefore, many research papers and reviews have been published. Pioneering reviews on the chemistry of oxovanadium were published in 1965.38 The synthesis of vanadium oxides through hydrothermal and gas phase was reviewed by Whittingham,39 Livage,40 and Bahlawane.41 The atomic layer deposition of vanadium oxides was summarized by Papakonstantinou.12 The synthesis, properties, and applications of vanadium oxide nanotube were described by Kianfar.6 The catalytic applications of vanadium oxides have been described by Delferro,3 Hess,42 Carrero,43 and Granozzi.44 The Raman spectroscopy of vanadium oxides was recently reviewed by Shvets.45 The sensing properties of vanadium oxide nanostructures were described by Sheikhi46 and Madanagurusamy.31 The energy-related applications of vanadium oxides were reviewed by Xie,5,20 Jiang,4 and Streb.47 A large number of reviews described the progress in metal ions batteries, including those by Chen,25 O’Dwyer,22 Mai,21,24,48 Lowe,23 Whittingham,49 Rashada,50 Lee,51 Yang,52 Zheng,53 Kim,54 Liang,55 and Cao.56 The vanadium oxides in supercapacitors were reviewed by Chen,30 Dutta,57 and Li.29 Furthermore, our group also reviewed the multistimuli responsive properties and energy applications of some vanadium oxides.19,27,58,59

The great potential of vanadium oxides for new applications and accelerated industrialization led to the dramatically increase of the importance of vanadium oxides over the last 5–10 years. Furthermore, due to the complexity of various oxidation states of vanadium, vanadium oxides show a large variety of stable and metastable structures, which pose an inevitable challenge to synthesize vanadium oxides with high purity, well controlled stoichiometry, and meticulously designed nanostructures, a must for high performance devices. Even though lots of reviews have been published, most of them focus on a specific kind or application of vanadium oxides, such as batteries, spectra, supercapacitors, and so on. No comprehensive reviews illustrate the different vanadium oxides with different applications. In this review, we focus on the most recent progress on the structure, synthesis, and applications of five thermodynamically stable vanadium oxides (V2O3, V3O5, VO2, V3O7, V2O5) and some metastable vanadium oxides (V2O2, V6O13, V4O9, etc.), which can provide a better understanding of a specific vanadium oxide phase and their process-structure–property interrelationships. Figure 1 summarizes the main contents of this review. This review begins with the phase diagram of the V–O system to show the different vanadium oxide phases, followed by the vanadium oxides with different stoichiometries. For each vanadium oxide, the structures, synthesis methods, and applications in several fields will be covered. The last section presents the future prospects and a summary of this review.

Figure 1.

Figure 1

Schematic illustration of the thermodynamically stable vanadium oxides and their applications.

2. Phase Diagram of the V–O System

The V–O binary phase diagram was compiled according to previously reported experimental data in 1989.60 The oxygen-rich phases are well-defined, which reveal more than 20 compounds.41 However, the vanadium-rich phases exhibit broad homogeneity ranges and high nonstoichiometry.61,62

Several groups have calculated the V–O binary phase diagram.6164Figure 2a shows a calculated phase diagram of the V–O system in the entire composition range at 1 atm. In the V-rich range, four types of solid solutions exist. The α and β solid solutions are formed by a certain amount of oxygen dissolved in the vanadium. The maximum solubilities of oxygen in α-V and β-V phase are up to 17.9 atom % and 27.4 atom %, respectively. The β-phase exhibits a wide range of homogeneities. With the increase of oxygen content, the γ and δ solid solutions phase can be formed. The γ-phase is monoclinic and δ-phase has the stoichiometry of VO with NaCl-type structure.

Figure 2.

Figure 2

(a) Calculated phase diagram of the V–O system in the entire composition range at 1 atm. (b) Enlarged phase diagram of the V2O3–V2O5 system at 1 atm. Reproduced with permission from ref (61). Copyright 2015 Elsevier.

For the stoichiometric phases, only five phases are thermodynamically stable compositions as pure compounds, which include divanadium trioxide (V2O3, cubic, Ia3), trivanadium pentoxide (V3O5, monoclinic, P2/c), vanadium dioxide (VO2, tetragonal, P42/mnm), trivanadium heptaoxide (V3O7, monoclinic, C2/c), and divanadium pentoxide (V2O5, orthorhombic, Pmnm). Other metastable phases or unstable phases are likely to decompose to stable phases. All the stable and metastable phases are listed in the enlarged phase diagram of the V2O3–V2O5 system, which is shown in Figure 2b.

There are two types of vanadium oxides with a mixed valence of vanadium. One is the Magnéli series, which is defined by the general stoichiometric formula:

2.

This type of homologous series has been reported for molybdenum oxides for the first time by Magnéli.65 The other homologous series is the Wadsley series, which has the general formula of VnO2n+1 (n ≥ 3). All the Magnéli phases maintain a triclinic symmetry (P1) and are metastable. They are expected to yield VO2 and V3O5 after lowering the system entropy. For the Wadsley series, V3O7 exhibits lower formation entropy compared with other Wadsley phases, which indicates the phase is more stable. Other Wadsley phases, such as V4O9 and V6O13, can be decomposed into a mixture of VO2 and V3O7.41 V4O9 and V6O13 show multiple metastable crystalline structures due to the close formation energies. Therefore, these compounds are candidates for polymorphism, where three crystalline structures were identified for V6O13 and two for V4O9.60

3. V2O3

3.1. Structures and Synthesis

V2O3 has a typical corundum-type hexagonal structure (space group: Rc) with lattice parameters of a = b = 4.9492(2) Å, c = 13.988(1) Å at room temperature.66,67 The crystal structures of V2O3 viewed in different directions are illustrated in Figure 3. It is interesting that after being calcined at 600 °C for several hours, vanadium vacancies (red circle in Figure 3a) formed, which is suitable for aqueous zinc metal batteries.67 From another viewing direction, V2O3 possesses an open tunnel structure consisting of a 3D V–V framework (Figure 3b).68 Such tunnel structures could efficiently facilitate the insertion of alkali metal ions, which provide potential application in metal ion batteries.

Figure 3.

Figure 3

Perspective view of V2O3: (a) along the (001) direction; (b) along the (110) direction.

V2O3 with different morphologies can be obtained by several methods, including reduction, oxidation, and hydrothermal approaches. For the reduction pathway, V2O5 was employed as initial materials, and hydrogen or ammonia gas was used as reducing agents, whereas vanadium metal was used as a starting material for the oxidation method. In the hydrothermal route, vanadium alkoxides or sulfides with some small organic molecules (e.g., thiourea, benzyl alcohol, etc.) were added together to the autoclave to grow V2O3.

Li et al.69 designed a plasma hydrogen reduction system to synthesize V2O3 nanocrystals via a single precursor of V2O5 powders. The coarse-grained V2O5 powders are injected into the hydrogen plasma by a powder feeder, reducing the V2O5 powders into V2O3 by hydrogen. Such single crystalline V2O3 nanocrystals have a spherical shape with average sizes in the range of 35–50 nm (Figure 4a). The morphology of the raw materials determines the V2O3 morphology. Seshadri et al.70 first synthesized the V2O5 nanorods via a hydrothermal reaction followed by reducing in 5% H2:95% N2 (reduction time = 3 h and reduction temperature = 600 °C) to obtain V2O3 nanorods (Figure 4b). The ammonia gas (NH3) is another agent to reduce V2O5 to V2O3.71 V2O3 shows a much larger size with several micrometers, and the morphologies of the V2O3 particles were micrometer layered structures that were assembled by nanometer or micrometer sheets. Tao et al.72 reported that the V2O3 nanoparticles had been synthesized by supercritical ethanol fluid reduction of VOC2O4 with an average size of 50 μm. Madanagurusamy et al. employed the oxidation way to obtain the V2O3 nanosheets.73 High-density vertically aligned V2O3 nanosheets on glass substrates were obtained via a simple one-step sputtering technique. Vanadium metal was first deposited on the well-cleaned glass substrates followed by oxidation in the argon and oxygen mixture gas with a ratio of 3:1. Well-ordered, ultrathin, vertically aligned V2O3 nanosheets with voids were obtained (Figure 4c).

Figure 4.

Figure 4

(a) Typical TEM image of V2O3 nanocrystals. Reproduced with permission from ref (69). Copyright 2012 Elsevier. (b) SEM image of V2O3 nanorods after reduction (time = 3 h and temperature = 600 °C). Reproduced with permission from ref (70). Copyright 2008 American Chemical Society. (c) SEM images of V2O3 nanosheets. Reproduced with permission from ref (73). Copyright 2018 Royal Society of Chemistry. (d) TEM of V2O3 nanocrystals. Reproduced with permission from ref (74). Copyright 2004 Elsevier.

Hydrothermal is one of the most popular methods for crystal growth, which is conducted under moderate temperature and a high vapor pressure environment in sealed containers with the unique advantages of being easy to handle and environmentally friendly. Compared with the reduction and oxidation methods, the hydrothermal reaction is more convenient for synthesis the V2O3 with various morphologies. Niederberger et al.74 adopted vanadium alkoxides and benzyl alcohol as precursors to synthesize V2O3 nanocrystals through hydrothermal reaction sizes ranging from 20 to 50 nm with good yields (Figure 4d). Without using any surfactant and template, Su et al.75 successfully synthesized dandelion-like V2O3 microspheres with core–shell structures. With increasing reaction time, the morphology of V2O3 can be tuned from a solid sphere to dandelion-like. If the reaction time is further increased, some broken V2O3 microspheres with core–shell structures could be observed with an average diameter of the core–shell microspheres of 2 μm.

3.2. Applications

3.2.1. Batteries

As the low valence of V3+ in vanadium oxide, corundum-type V2O3 with a metallic behavior shows that the electrons in the V-3d orbital travel along the V–V chains. The 3D V–V framework provides an intrinsic tunnel structure, which is suitable for ion transport and intercalation. Mai et al. reported uniform nitrogen-doped carbon-confined V2O3 (V2O3@NC) hollow spheres (Figure 5a). The in situ carbonization hollow structure can provide high ion/electron conductivity, short diffusion distance, and excellent structure adaptability, which is beneficial for lithium-ion storage (Figure 5b). The V2O3@NC delivers an average discharge capacity of 785, 599, 528, and 361 mAh g–1 at a current density of 100, 500, 1000, and 5000 mA g–1, respectively (Figure 5c). Furthermore, 811 mAh g–1 can be maintained after 120 cycles at a current density of 200 mA g–1 (Figure 5d).76 In addition, V2O3 can also be used as other ions (Na+/K+/Zn2+) storage material. Jiao’s group68 fabricated a flexible and self-standing electrode of V2O3 nanoparticles embedded in porous N-doped carbon nanofibers (V2O3@PNCNFs) through electrospinning assisted high-temperature sintering method, which can directly be used as a KIB anode material (Figure 5e). V2O3@PNCNFs deliver a capacity of 240 and 134 mAh g–1 at a current density of 50 and 1000 mA g–1, respectively. High-capacity retention of 94.5% can be maintained after 500 cycles (Figure 5f). The density functional theory (DFT) results demonstrate that 1 mol K+ can insert into the V2O3 crystal forming KV2O3, with the K+ occupying the 6e sites of the KV2O3 crystal. The main capacity of V2O3 in KIB is main contribution from capacitance (surface or near-surface redox reactions), which is very different from the conversion reaction of conventional transition metal oxides for ion storage. When K+ inserts into the tunnels of V2O3@PNCNFs, the material exhibits no structural phase transitions, which indicates that V2O3@PNCNFs can exhibit excellent K+ rate/cycling performance. This work suggests that the pseudocapacitive electrode materials could be suitable for a large-scale energy storage system. Generally, the conventional low valent V2O3 cannot effectively accommodate Zn2+ intercalation during the discharging process due to its inherently unsuitable structure and inferior physicochemical properties. Luo et al.77 attempted to utilize V2O3 in aqueous ZIBs by the in situ anodic oxidation strategy, and the hierarchical microcuboid structure V2O3 can accommodate nearly 2 electrons intercalation. Moreover, they found H2O is a reactant that participates in the first charge oxidation process of V2O3. Furthermore, the porous structure with a higher specific surface area of the V2O3 leads to more reaction sites and a faster phase transition from V2O3 to V2O5–x·nH2O (Figure 5h). Meanwhile, the high surface and the small size of V2O3 nanoparticles benefit the first charge oxidation reaction process. The V2O3 delivers a Zn2+ discharging capacity of 625 mAh g–1 at 0.1 A g–1, corresponding to 1.75-electron intercalation (Figure 5g). Specifically, the capacities can maintain 87% and 78% when the current increases to 10 and 20 A g–1, respectively. The V2O3 can maintain 100% after 10000 cycles at 10 A g–1, which is better than some previously reported ZIB cathode materials.

Figure 5.

Figure 5

XRD pattern (a), SEM image (b), charge/discharge profiles (c), and cycling performance (d) of V2O3@NC hollow spheres. Reproduced with permission from ref (76). Copyright 2018 Royal Society of Chemistry. The digital photos of the as-spun membrane and the self-standing (e), and rate performance (f) of V2O3@PNCNFs. Reproduced with permission from ref (68). Copyright 2018 Elsevier. The charge/discharge profiles at different voltage window (g) and schematic illustration of the oxidation at full charge state (h) of V2O3. Reproduced with permission from ref (77). Copyright 2020 American Chemical Society.

3.2.2. Catalysts

V2O3 and its composites have been widely used as catalysts for chemical looping dry reforming of methane,78 ammonium perchlorate decomposition,79 the hydrogen evolution reaction (HER),30,80 the oxygen evolution reaction (OER),80 water splitting,80 etc.

3.2.2.1. Propane Dehydrogenation and Ammonium Perchlorate Decomposition

Very recently, Zhu et al.81 examined the catalytic properties of propane dehydrogenation through single transition metal atom doping of a V2O3 (0001) surface by self-consistent DFT calculation. The results indicated that the single atoms act as promoters and active sites, and Mn–V2O3 is a good candidate as a catalyst for propane dehydrogenation. Huang et al.79 synthesized V2O3 and V2O3/carbon composites by a facile hydrothermal route, which exhibited excellent performance for ammonium perchlorate decomposition. The decomposition temperature decreased by 49 and 73 K for V2O3 and V2O3/carbon composites, respectively.

3.2.2.2. HER, OER, and Water Splitting

Electrocatalytic water splitting consisted of two half-reactions: HER and OER. Electrocatalysis is used to accelerate the rate of a chemical reaction through lowering the activation energy to reduce the electrochemical overpotentials.

V2O3 composited with other functional materials can exhibit efficient electrocatalysis performance. Zhang et al.82 reported a self-supported MoSx/V2O3 heterostructure for the HER. Two steps were involved in the synthesis process (Figure 6a). V2O3/carbon cloth (CC) was first obtained by a hydrothermal method, and the dense V2O3 was uniformly distributed on CC fibers. The as-prepared V2O3/CC was immersed in an ammonium thiomolybdate solution and dried under a vacuum. Then, followed by a thermal decomposition process, the final MoSx/V2O3/CC was successfully achieved with the same morphology as V2O3/CC. The XRD pattern (Figure 6b) shows the composite contains the MoSx and hexagonal V2O3. MoSx/V2O3/CC displayed an overpotential of 146 mV to achieve a 10 mA cm–2 HER current density (Figure 6c), which is lower than that of MoSx/CC (221 mV). The Tafel slope of MoSx/V2O3/CC is around 45 mV dec–1, suggesting the HER process obeys the Volmer-Heyrovsky mechanism. Meanwhile, the MoSx/V2O3/CC presents a higher double-layer capacitance (Cdl: 85 mF cm–2) than MoSx/CC (4 mF cm–2), indicating V2O3 can create more active sites for the HER (Figure 6d). Furthermore, the MoSx/V2O3/CC electrocatalyst has great stability in the acid electrolyte for the HER. Two reasons for the MoSx/V2O3/CC electrocatalyst were given to understand the improved HER activity: (1) V2O3 enhanced the electrochemically active surface area with more active sites for the HER; (2) better electron transfer between MoSx and V2O3. Qiu et al.80 synthesized V2O3 nanosheets anchored with NiFe nanoparticles as a bifunctional electrode for overall water splitting. A self-templated strategy was employed to synthesize the NiFe@V2O3 (Figure 6e). By a calcined reduction process, ZnO@NiFe@V2O3 nanosheets can be obtained. In-situ alkaline media corrosion was performed to dissolve the ZnO NPs and produce the porous NiFe@V2O3 nanosheets, which exhibit a clear V2O3 porous matrix and NiFe nanoparticles (Figure 6f). The porous NiFe@V2O3 exhibits good OER performance in an alkaline medium with an overpotential of 255 mV at 10 mA cm–2 and good stability (Figure 6g,h). Meanwhile, the NiFe@V2O3 also gave good HER performance in the same alkaline medium. It shows an overpotential of 84 mV at 10 mA cm–2 and good stability (Figure 6i,j). Furthermore, the porous NiFe@V2O3 delivers a small Tafel slope of 51 mV dec–1 for OER and a Tafel slope of 85.4 mV dec–1 for the HER, respectively. Considering the excellent HER and OER performance, two identical NiFe@V2O3 electrodes are integrated into a two-electrode cell to investigate the water splitting performance. The catalyst shows a cell potential of 1.56 V to reach 10 mA cm–2 with an ignorable cell voltage increase during 20 h measurements (Figure 6k).

Figure 6.

Figure 6

(a) Schematic diagram of the main steps in the synthesis of the MoSx/V2O3/CC. (b) XRD patterns of MoSx/V2O3/CC, V2O3/CC, and MoSx/CC. (c) Polarization curves for MoSx/V2O3/CC. (d) Double-layer capacitance for MoSx/V2O3/CC. Reproduced with permission from ref (82). Copyright 2020 Elsevier. (e) Synthetic strategy for preparing 3D hierarchical nanoporous NiFe@V2O3. (f) High-magnification TEM image of NiFe@V2O3. (g) OER polarization curves of NiFe@V2O3. (h) Long-term durability test conducted at 10 mA cm–2 of NiFe@V2O3. (i) HER polarization curves of NiFe@V2O3. (j) Long-term durability test conducted at 10 mA cm–2 of NiFe@V2O3. (k) Long-term potential-time curve set at 10 mA cm–2 of NiFe@V2O3||NiFe@V2O3 in 1.0 M KOH. Reproduced with permission from ref (80). Copyright 2019 American Chemical Society.

The stability of the catalyst is a key parameter for practical applications, which is crucial to determine whether the catalyst can be commercialized.83 Several factors limit the catalytic stability, such as poor chemical/electrochemical stability of the catalysts under operation, abandoned gas evolution leading to the physical detachment of catalysts, dissolution of the catalysts in the electrolytes with different pHs, poor mechanical stability, and so on.8486 For example, Schmidt et al.87 proved that the metal atoms are thermodynamic unstable during the OER process due to metal oxide released lattice oxygen, which leads to the low stability of the catalysts. From the theoretical perspective, there are three critical criterion of the catalysts that should be considered: excellent water dissociation performance, suitable Gibbs adsorption energy of H* (ΔGH*), and faster H2 desorption.83 Such properties enable the fast release of the active site without destroying the active site and further improve the stability of the catalysts.

3.2.3. Supercapacitors and Electromagnetic Wave Absorber

Supercapacitors (SCs) have attracted attention in recent years to bridge between a classic electrolytic capacitor and rechargeable battery, which are characterized by high power density and good cyclic stability. SCs are capable of storing electrical energy via two mechanisms: the electrochemical double layer capacitors having a nonfaradaic charge character and the pseudocapacitors based on faradaic electrochemical redox reactions.88 The theoretical specific capacitance of a pseudocapacitive electrode is proportional to the number of electrons involved in a specific redox reaction, and vanadium oxides possess four readily accessible valence states, making vanadium oxides especially promising for high pseudocapacitance.89 The conductivity of V2O3 (∼103 Ω–1 cm–1) is higher than monoclinic VO2 (∼4 Ω–1 cm–1) and comparable with Ru (∼104 Ω–1 cm–1).90 Meanwhile, V2O3 is stable in both acid and base mediums.91 Thus, V2O3 is a suitable material for energy storage, especially for SCs. However, the reported specific capacitances of V2O3-based materials are not good enough, suggesting that design and synthesis of new structured V2O3-based materials with high performance are required.92 Cao et al.93 synthesized V2O3 nanoparticles highly dispersed in amorphous carbon composites through the calcination of the (NH4)2V3O8/C precursor, which was fabricated through the hydrothermal reaction by using commercial NH4VO3 and glucose (Figure 7a). The as-prepared V2O3-based composites exhibit a specific capacitance of the electrode of 458.6 F g–1 at a current density of 0.5 A g–1 (Figure 7b), which is higher than other reported V2O3-based composites.94 Meanwhile, the asymmetric supercapacitors assembled by the as-prepared V2O3-based composites display good flexibility properties (Figure 7c). The capacitances are almost constant with the bending from 180° to 30°. The V2O3-based composites are one kind of high-efficiency electromagnetic wave absorber. Yan et al.95 produced the hierarchical Co/C@V2O3 hollow spheres by hydrothermal, aging, and annealing methods (Figure 7d). The VO2 spheres were first fabricated through a hydrothermal method followed by coating with ZIF-67 by precipitation. After that, the ZIF-67@VO2 composite was calcined in H2/Ar to form the hierarchical Co/C@V2O3 hollow spheres. The Co/C@V2O3 hollow spheres exhibited excellent electromagnetic wave absorption performance with a reflection loss of @40.1 dB and a broad bandwidth of 4.64 GHz at a small thickness of only 1.5 mm (Figure 7e). The good performance is mainly due to the impedance matching and low density, which come from the combination of hollow V2O3 spheres and porous Co/C.

Figure 7.

Figure 7

(a) A schematic illustration of the synthesis of highly dispersed VO-C. (b) GCD curves of VO-C collected at different current densities. (c) Capacitance retention of the VO-C||AC device measured under different bend conditions. Reproduced with permission from ref (93). Copyright 2019 Elsevier. (d) The synthesis process of hierarchical Co/C@V2O3 hollow spheres. (e) Reflection loss as a function of the frequency. Reproduced with permission from ref (95). Copyright 2019 John Wiley and Sons.

4. V3O5

4.1. Structures and Synthesis

V3O5 crystal structure was first reported in 1954,96 which possesses a monoclinic symmetry (space group P2/c).97 The crystal structure is shown in Figure 8. The oxygen atoms are occupied in the octahedral sites, and vanadium atoms are located in the center of octahedron. There are two types of octahedra chains along the c axis: face-shared octahedra via edge-sharing and corner-sharing octahedra. Such chains formed a framework with many large open spaces, which is capable of accommodating lithium ions.98

Figure 8.

Figure 8

Crystal structure of V3O5.

It is difficult to stabilize the V3O5 phase by using solid-state chemistry and to control the stoichiometry between oxygen and vanadium by using a solution method, which makes V3O5 an uncommon phase used for electrochemical and other applications.98,99 Therefore, the reported synthetic method for V3O5 is very limited. Reduction of V2O5 by different reductants is commonly used to obtain the V3O5 polycrystalline powders. Reisner et al.100 selected vanadium metal as a reductant to reduce V2O5, and the V3O5 polycrystalline powders were synthesized according to the following chemical reaction: V + V2O5 → V3O5. Vanadium powder and V2O5 powder were mixed together and pressed into bars, which were sealed in a quartz ampule and heated for 24 h at 870 K and for 100 h at 1220 K. The V3O5 polycrystalline powders exhibit a size around 10 μm. Alternatively, Yu et al.98 used sulfur powders as a reductant. The mixture of sulfur and V2O5 powders were vigorously grounded and calcined at 1023 K in a tube furnace under a vacuum for 2 h. The V3O5 microcrystals can be obtained by washing the calcined powders with nitrogen tetrasulfide several times. The obtained V3O5 microcrystals range from 1 to 3 μm. The large size single crystal of V3O5 was grown by chemical vapor transport in the 1970s.101,102 V3O5 powders or V2O3–VO2 mixed powders were used as starting materials, and TeCl4 was used as a transport agent. Both starting materials and transport agent were sealed in a quartz tube with low pressure (∼1 × 10–5 mbar), and then the tube was put into a two-zone furnace. Generally, the source zone has higher temperature, while the crystallization zone has a lower temperature, and the temperature gradient is around 100 to 150 K. The size of final V3O5 single crystal is as large as 1 cm.

4.2. Applications

4.2.1. Batteries

V3O5 is relatively less-studied though it exhibits a three-dimensional open-framework structure, which is due to the strict synthesis condition. Yu’s group98 successfully synthesized V3O5 microcrystals via vacuum calcination and first employed it as an LIB anode material. The oxygen atoms are closely arranged in a hexagonal shape, and the vanadium atoms take up three-fifths of the octahedral interstices (Figure 9a). The 3D open-framework structure of V3O5 is formed with chain connections, which contain distorted, sharing corners, edges, and faces of VO6 octahedral, respectively (Figure 9b). The 3D framework of V3O5 with much large open space endows the capacity of Li+ intercalation/deintercalation. No impurity peaks from the XRD pattern are observed (Figure 9c), which indicates that the V3O5 powder shows a single-phase nature with a monoclinic structure within the P2/c space group (JCPDS card 72-0977). It delivers a high capacity of 628 mAh g–1 at 100 mA g–1, a good rate (125 mAh g–1 at 50 A g–1), and a long stable cycling performance (117 mAh g–1 after 2000 cycles) (Figure 9d,e). There is no obvious crystal structure change of the V3O5 with Li+ intercalation, which is the main reason for the good rate and cycling performance (Figure 9f).

Figure 9.

Figure 9

[101] projection illustrating the hexagonal close packing of oxygen atoms (a), the connection of VO6 octahedral (b), the XRD pattern (c), charge/discharge profiles (inset: corresponding cycling performance at 100 mA g–1) (d), rate performance (e) and ex-situ XRD patterns (f) of V3O5. Reproduced with permission from ref (98). Copyright 2019 under CC BY license.

4.2.2. Other Applications

V3O5 thin films exhibit a photoinduced insulator-to-metal phase transition, which results in a strong nonlinear optical response.103105 Fernández et al.104 deposited the V3O5 directly on the SiO2 substrates by DC magnetron sputtering to form thin films. The ultrafast nonlinear optical response was probed by using a pump–probe scattering technique. A reduction in the transient relative scattered light signal was observed, which showed an ∼10% decrease within 800 fs. Such a response is due to the changes in the material’s optical constants and very likely related to the photoinduced insulator-to-metal phase transition.106 The photoinduced screening of electron correlations followed by melting of polaronic Wigner crystal and coalescence of V–O octahedra is the main reason for the order–disorder structural transition.103

5. V3O7

5.1. Structures and Synthesis

Figure 10a shows the crystal structure of V3O7. The unit cell contains 36 vanadium atoms (12 vanadium atoms are inside the octahedra and 24 vanadium atoms are five-coordinated).107 The V3O7 consists of VO6 octahedra and VO5 polyhedra, which are linked by corners and edges to form a three-dimensional framework. The crystal structure of V3O7·H2O shows a two-dimensional structure compared with V3O7 (Figure 10b). Each V3O8 layer consists of corner- or edge-shared VO6 octahedra and VO5 polyhedra.108 The water molecules are located at the sides of the V3O8 layer, where the hydrogen atoms are directly bonded to the vanadium atom of a VO5 polyhedron and hydrogen bonds connect two neighboring V3O8 layers.7,109

Figure 10.

Figure 10

(a) Crystal structure of V3O7 with VO6 octahedra (yellow) and VO5 (blue). (b) Crystal structure of V3O7·H2O (H2V3O8), with VO6 octahedra (yellow) and VO5 (green) and VO5 (blue) polyhedra. Hydrogen atoms are bonded to the oxygen atoms (purple), which are not shown. Reproduced with permission from ref (108). Copyright 2018 American Chemical Society.

Bulk V3O7 single crystal can be grown by a typical chemical vapor transport method using V3O7 polycrystalline powders as starting materials and NH4Cl as a transport agent.110 The mixed V3O7 polycrystalline powders and NH4Cl were pressed into a pellet and heated at 823 K in an evacuated silica tube for 7 days. The black needle-like single crystals of V3O7 were formed with a length of 2 mm (Figure 11a). Nanostructured V3O7 was generally obtained by a typical hydrothermal method. The precursor solution dramatically affects the morphology of the V3O7 nanostructures. Wen et al. combined a soft chemical topotactic synthesis and hydrothermal process to prepare V3O7 nanobelts.111 In the beginning, layered structured KV3O8 plate-like particles were first prepared as the precursor by a hydrothermal method of V2O5 and KOH. The H+-form vanadate (HVO) nanobelt colloidal solution was subsequently obtained by reacting KV3O8 plate-like particles with HNO3. Finally, the one-dimensional (1D) pure single crystal V3O7 nanobelts were successfully prepared after hydrothermally treating the colloidal solution at 180 °C for 12 h (Figure 11b). By using NH4VO3 and HCl as starting materials, the nest-like V3O7 self-assembled by porous V3O7 nanowires on Ti foil was also synthesized through a hydrothermal method.112 In the beginning, a layer of V3O7 nanosheets was deposited on the Ti substrate, which was subsequently placed in the NH4VO3–HCl solution and kept at 160 °C for 10 h for a hydrothermal method. With the increase of hydrothermal time, the nest-like V3O7 is self-assembled by nanowires. The schematic synthesis process and SEM images for nest-like V3O7 are shown in Figure 11c–f. The V3O7 fibers can be obtained by thermal treatment of the electrospun NH4VO3/PVP nanofibers in the presence of reductant.113 The thermal treatment condition dramatically affects the final products, and V2O5 may be obtained together with V3O7.

Figure 11.

Figure 11

(a) Optical images of bulk V3O7 single crystals. Reproduced with permission from ref (110). Copyright 2009 Elsevier. (b) The SEM image of V3O7 nanobelts obtained through a soft chemical topotactic synthesis and hydrothermal process. Reproduced with permission from ref (111). Copyright 2016 Royal Society of Chemistry. (c) Schematic synthesis process for nest-like V3O7. (d–f) SEM images of V3O7 prepared with the hydrothermal times of 1, 5, and 10 h. Reproduced with permission from ref (112). Copyright 2018 Royal Society of Chemistry.

The hydrothermal reaction is also a facile way to synthesize V3O7·H2O nanostructures. The V3O7·H2O nanobelts can be achieved by using V2O5, phenolphthalein and distilled water as starting materials through a hydrothermal method for 4 days of reaction at 180 °C.114 By changing phenolphthalein to ethanol or glucose, V3O7·H2O nanobelts were also successfully synthesized at 180 °C, while the reaction time was shortened to 12 h.115,116 In above-mentioned the methods, phenolphthalein, ethanol, or glucose is used as the reductant. Without the reductant, the V3O7·H2O nanobelts or nanowires can be obtained by a hydrothermal reaction of only V2O5 or NH4VO3.117,118 The key point is the pH of the precursor solution. After adjusting the pH to 3 by adding concentrated HCl, V3O7·H2O nanobelts and nanowires were obtained by a reaction at 190 °C for 24 h and 160 °C for 3 h, respectively. Furthermore, ultrafine V3O7·H2O nanogrids can be obtained through electrochemical oxidation.119 In the beginning, the nsutite-type VO2 black powder was synthesized by a hydrothermal method. Then, a three electrode system was employed to electrochemically transform the VO2 precursor to V3O7·H2O. The slurry, consisting of VO2 nanoplates, was coated on the working electrode. With a constant current density of 50 mA cm–2 and a cutoff potential of 1.7 V, V3O7·H2O nanogrids were obtained (Figure 12).

Figure 12.

Figure 12

Schematic illustration of the preparation of the nanogrid-shaped V3O7·H2O. Reproduced with permission from ref (119). Copyright 2019 Royal Society of Chemistry.

5.2. Applications

5.2.1. Batteries

V3O7 is a mixed-valence vanadium oxide for metal-ion storage. Yan et al.120 designed and synthesized a V3O7 nanowire templated graphene scroll (VGS) via an “oriented assembly” and “self-scroll” strategy. They used joint experimental-MD simulation to investigate the construction and formation mechanisms of VGS. The systemic energy, the curvature of nanowires, and the reaction time determined the length and formation process of the semihollow bicontinuous structure. Through this strategy, the VGS with a length up to 30 μm has interior cavities between the nanowire and scroll (Figure 13a). The unique structure of VGS with the nanowire templated graphene scroll offers a continuous Li+/ion transfer channel and free volume expansion space during Li+ de/intercalation (Figure 13b). The VGSs exhibit a high capacity of 321 mAh g–1 and good cycle stability (87.3% after 400 cycles), which is better than the pure V3O7 nanowire and V3O7 nanowire/graphene structure (Figure 13c).

Figure 13.

Figure 13

TEM image of VGS (the inset gives an HRTEM image of a V3O7 nanowire in graphene scrolls) (a), schematic of the VGS nanoarchitecture with continuous electron and Li ion transfer channels (b), and cycling performances (c) of VGS, V3O7 nanowire/graphene, and V3O7 nanowire. Reproduced with permission from ref (120). Copyright 2013 American Chemical Society. The schematic illustration of reaction mechanism (d), charge/discharge profiles in aqueous (e), and nonaqueous (f) electrolytes of Zn//V3O7·H2O batteries. Reproduced with permission from ref (121). Copyright 2018 Royal Society of Chemistry. The powder X-ray Rietveld refinement profile and crystal structure (g), charge/discharge profiles (h), and cycling performance (i) of V3O7·H2O nanowires in MIBs. Reproduced with permission from ref (108). Copyright 2018, American Chemical Society.

In addition, Nazar et al.121 synthesized layered V3O7·H2O nanobelts with single crystalline via a microwave solvothermal method and applied it to a ZIB cathode in nonaqueous and aqueous electrolytes. The electrochemical studies and in situ XRD results demonstrate the different electrochemical behaviors of layered V3O7·H2O in nonaqueous and aqueous electrolytes. Combining the DFT calculations, ∼2 mol Zn2+ can insert into V3O7·H2O per the formula in the ZnSO4/H2O aqueous electrolyte (Figure 13d). The V3O7·H2O delivers a capacity of 400 mAh g–1 (>2 mol Zn2+ insertion) with an average voltage of ∼0.65 V at the first discharge process, and maintains 375 mAh g–1 at the subsequent charge process (Figure 13e), while in the Zn(CF3SO3)2/acetonitrile nonaqueous electrolyte, the V3O7·H2O exhibits a poor Zn2+ storage performance (59 and 175 mAh g–1 for the first and 50th cycles at 5 mA g–1, respectively) (Figure 13f).

Magnesium-ion battery (MIB) as another multivalent ion battery has been attracting more attention due to its high abundance in the Earth and low redox potential (−2.37 V vs. SHE). V3O7·H2O with high electronic conductivity (V+4.67) has been widely used as a cathode material in LIB/NIB and hybrid Li+/Mg2+ batteries. Hong et al.108 synthesized V3O7·H2O nanowires via a one-step hydrothermal method and applied them to a high-energy MIB cathode (Figure 13g). The electrochemical tests and structural characterization results demonstrate that the structured water in V3O7·H2O will remain stable during the cycling. 0.97 mol Mg2+ inserts into V3O7·H2O, accompanying the formation of Mg0.97H2V3O8 at the first discharged state. V3O7·H2O exhibits an initial discharge capacity of 231 mAh g–1 at 10 mA g–1 with an average discharge voltage of ∼1.9 V, and the energy density can reach 440 Wh kg–1 (Figure 13h). Meanwhile, V3O7·H2O delivers a 171 mAh g–1 and maintains 132 mAh g–1 (77%) after 100 cycles at 40 mA g–1 (Figure 13i). The excellent Mg2+ storage performance is attributed to the unique crystal structure with direct bonding. This strategy of applying water-metal bonding and hydrogen bonding provides a new idea to search for new oxide-based MIB materials with stable and high energy density.

5.2.2. Ammonium Perchlorate Decomposition

Ammonium perchlorate, a common oxidizer, plays a key role in the combustion of composite solid propellants. Furthermore, the catalyst greatly affected the performance of composite solid propellants by the thermal decomposition of ammonium perchlorate.122,123 Huang et al.124 found that the thermal decomposition temperatures of ammonium perchlorate in the presence of V3O7·H2O nanobelts and V3O7·H2O@C core–shell structures can be dramatically reduced. Both V3O7·H2O nanobelts and V3O7·H2O@C core–shell structures were synthesized by a hydrothermal method. Especially, the core–shell structures are synthesized by using V3O7·H2O nanobelts as the cores and glucose as the source of carbon. A well-defined nanobelt morphology with a length up to several micrometers can be observed (Figure 14a), which consists of a V3O7·H2O core and carbon shell (Figure 14b). The thermogravimetric analysis (TGA) indicated that the addition of V3O7·H2O or V3O7·H2O@C in ammonium perchlorate exhibited a significant reduction in the decomposition temperature of ammonium perchlorate (Figure 14c). The thermal decomposition temperature was lowered by 70 and 89 °C by adding V3O7·H2O or V3O7·H2O@C, respectively. The V3O7·H2O@C core–shell structures exhibited a higher catalytic activity than V3O7·H2O, with two possible mechanisms proposed. First, the partially filled 3d orbit in the vanadium atom promoted the electrotransfer process by accepting electrons from ammonium perchlorate and further accelerated the thermal decomposition of ammonium perchlorate. Second, the amorphous carbon shell possessed lots of active groups (such as C=C, C=O), which could facilitate the thermal decomposition of ammonium perchlorate.

Figure 14.

Figure 14

(a) and (b) TEM images of V3O7·H2O@C core–shell structures, (c) TGA curves of pure ammonium perchlorate, V3O7·H2O with ammonium perchlorate and V3O7·H2O@C with ammonium perchlorate. Reproduced with permission from ref (124). Copyright 2011 Elsevier.

5.2.3. Supercapacitors

V3O7 and V3O7·H2O are promising supercapacitor materials due to their layered structure and mixed oxidation states of +4 and +5. V3O7 can be converted to V6O13 at the lowest potential of −0.6 V and V2O5 at the highest potential of 0.2 V. Thus, the working potential window is in the range of −0.6 and 0.2 V. Huang et al.112 fabricated porous V3O7 nanowire self-assembled nest-like V3O7 and investigated the supercapacitor properties. The pure nest-like V3O7 exhibits worse supercapacitor performance compared with N-doped carbon coated nest-like V3O7 composites. The N-doped carbon coated nest-like V3O7 electrode showed a higher specific capacity of 660.63 F g–1 at 0.5 A g–1 compared to V3O7 (362.63 F g–1). Even at a higher current density of 50 A g–1, the N-doped carbon - V3O7 electrode still exhibits a better performance (187.72 F g–1) than V3O7 (33.18 F g–1), as shown in Figure 15a. Furthermore, the N-doped carbon coated nest-like V3O7 electrode also delivers better stability (80.47% capacitance retention after 4000 cycles) compared with pure V3O7 (23.16% capacitance retention after 4000 cycles). The superior performance of N-doped carbon coated nest-like V3O7 is mainly due to the unique three-layer structure: V3O7 core/carbon/nitrogen doped carbon (Figure 15b). Such a unique three-layer structure can not only stabilize V3O7, but also provide high-speed ionic and electronic transmission channels, which is responsible for the good supercapacitor performance of N-doped carbon coated nest-like V3O7. Yu et al.125 reported the growth of V3O7 nanowires on a carbon fiber cloth through a hydrothermal method. The obtained V3O7/carbon fiber cloth composites show a spider web-like morphology, which exhibits robust adhesion. The composite electrode gives a maximum specific capacitance of 151 F g–1 at a current density of 1 A g–1 with ultrahigh cycling stability of 97% (after 100000 cycles) in a full cell configuration (Figure 15c). Meanwhile, the V3O7/carbon fiber cloth composites reveal maximum power and energy densities of 5.128 kW kg–1 and 24.7 Wh kg–1, respectively by using 1-ethyl-3-methylimidazolium trifluoromethanesulfonate as the electrolyte. Furthermore, coin cell-type configuration with the V3O7- carbon fiber cloth composites electrode was assembled. The symmetric supercapacitors successfully and effectively power light-emitting diodes to produce blue light (Figure 15d). Huang et al.115 reported that V3O7·H2O nanobelts exhibited a capacitance of 447.6 F g–1. However, the cycling performance is limited by the poor conductivity and high solubility in an aqueous electrolyte. Therefore, composing with another conductive phase could be an alternative way to fabricate high-performance V3O7·H2O based materials. When the V3O7·H2O nanobelt is incorporated with carbon nanotube and reduced graphene, the formed 3D hierarchical porous composites exhibit outstanding electrochemical performance with a high specific capacitance (685 F g–1 at 0.5 A g–1) and excellent cycle stability (99.7% after 10,000 cycles) (Figure 15e,f).126 Meanwhile, the composites also give relatively high energy densities and power densities of 34.3 Wh kg–1 and 150 W kg–1, respectively. The better electrochemical performance can be attributed not only to the highly conductive carbon materials, but also to the 3D hierarchical porous structure. The carbon materials offer the transport pathway bridges, leading to the rapid transfer of charges. Meanwhile, the 3D porous structure minimizes the diffusion distance and supplies a large surface area with abundant active sites.

Figure 15.

Figure 15

(a) Rate capability of V3O7 and N-doped carbon coated nest-like V3O7 calculated from the charge/discharge curves as a function of current density, (b) schematic diagram of the crystal structure and bonding in N-doped carbon coated nest-like V3O7. Reproduced with permission from ref (112). Copyright 2018 Royal Society of Chemistry. (c) Cycling stability of V3O7- carbon fiber cloth symmetric supercapacitors devices at a constant current density of 10 A g–1 for 100 000 cycles, (d) series connection of three-coin cells for practical applications, charging process of serially connected coin cells, and a demonstration of LEDs lit with charged coin cells. Reproduced with permission from ref (125). Copyright 2018 Royal Society of Chemistry. (e) Comparison of specific capacitance at different current densities from GCD curves, (f) cycling performance at 100 mV s–1 of V2O5 nanosheets, VOx/CNT, and V3O7·H2O/CNT/rGO. Reproduced with permission from ref (126). Copyright 2018 Elsevier.

In general, vanadium oxides have received massive interest as supercapacitor electrodes that exhibit high theoretical specific capacity than most of the other transition metal oxides due to their variable valence state from V2+ to V5+. In addition, the layered structure of vanadium oxides facilitates the intercalation/deintercalation of electrolyte ions during the charging/discharging process. However, vanadium oxides-based supercapacitor electrode materials still suffer from poor long-term cycling stability, which is usually caused by the collapse of a layered crystal structure, severe agglomeration of particles, and low electrical conductivity. The electrochemical stability of vanadium oxide-based supercapacitors can be improved by material modification, optimization of the structure, or combining with other materials with excellent electrical conductivity. Developing vanadium oxide nanomaterials with suitable micro-/nanostructures is an important factor to improve the cycling stability. 3D vanadium oxides with micro-/nanostructures are often employed, including microspheres and hollow spheres, which can inherit the superior high surface area characteristics of nanobuilding blocks, and simultaneously possess a decent structural stability.127130 Furthermore, the 3D structure can effectively reduce the agglomeration of particles, which is beneficial for the cycling and rate performance.131 Integrating vanadium oxides with carbon materials has been demonstrated to be effective to suppress the structural degradation upon cycling. The carbon materials, such as porous carbons and graphene, can serve as elastic buffering layers to release the strain within metal oxides during cycling.132 To some extent, carbon materials can avoid loose attachment between the electrode material and current collector, which helps improve both the conductivity and the stability of the supercapacitor.

5.2.4. Electrochromism

Nanostructured V3O7 thin films showed electrochromic properties by using lithium perchlorate as the electrolyte, which was prepared by a nebulizer spray pyrolysis technique.133 The color of the films is changed from yellow to pale blue by applying an external potential of 1.5 V for intercalation of Li+ ions, while the color is reversed by applying an external potential of −1.5 V for deintercalation of Li+ ions. Such results indicate that the nanostructured V3O7 thin films could be effectively used for smart window applications. In general, several models are proposed to explain the electrochromic phenomenon, such as the color center model (Deb model), electrochemical redox model, and so on.134 The “Deb’s color center model” is in nature related to the defects (such as oxygen vacancies) induced change of the visible light absorption, which is generally independent of the external electric field,135 while in the “electrochemical redox model” it is believed that the injection and trapping of a large density of electron and hole lead to coloration.136 The mechanism of electrochromism for V3O7 thin films is the variation of the band structure caused by intercalation/deintercalation of Li+ ions, which belongs to the electrochemical redox model. During the application of negative bias, Li+ ions are absorbed onto the surface of V3O7 and diffuse into the lattice of V3O7. The intercalated Li+ ions react with O2– ions to introduce oxygen vacancies in the lattice and reduce the V5+ in the mixed state V3O7 to V4+. As the result, the optical transparency of the film changes.137 Moreover, the intercalation of Li+ ions upshifts the Fermi level close to the conduction band, which leads to an increased transmittance of V3O7. The process can be reversed by deintercalating of Li+ ions through applying positive bias. Furthermore, the V3O7·H2O thick films also show good electrochromic properties (good reversibility and good color switching between reduced and oxidized states).138 By using lithium bis-trifluoromethanesulfonimide (LiTFSI) as the electrolyte, the color of the V3O7·H2O films changes from green to orange by applying a positive potential of 1.9 V, while it changes from green to blue by applying a negative potential of −1.2 V (Figure 16a). Such film presented a maximum optical reflectance modulation of 29% at 590 nm (Figure 16b). The V3O7·H2O films still exhibited color changes by using a Na-based electrolyte (Figure 16c) and a maximum optical reflectance modulation of 10% at 590 nm (Figure 16d). The maximum optical reflectance modulation of Na-based electrolyte is lower than that in the Li-based electrolytes, which is mainly due to a larger faradaic contribution resulting from the larger cation size of Na+ ions.

Figure 16.

Figure 16

(a) The visual appearance of V3O7·H2O films in the reduced and oxidized states in LiTFSI, (b) diffuse reflectance spectra for the V3O7·H2O films cycled in Li-electrolyte/Pt vs SCE, in reduction at −1.2 V and reoxidized at +1.9 V, (c) the visual appearance of V3O7·H2O films in the reduced and oxidized states in NaTFSI, (d) diffuse reflectance spectra for the V3O7·H2O films cycled in Na-electrolyte/Pt vs SCE, in reduction at −1 V and reoxidized at +1.9 V. Reproduced with permission from ref (138). Copyright 2020 Royal Society of Chemistry.

6. VO2

6.1. Structures and Synthesis

Vanadium dioxide (VO2) can exist in various polymorphic phases, including but not limited to VO2 (B), VO2 (A), VO2 (M), VO2 (R), VO2 (D), and VO2 (P).139 Some of these phases and their corresponding lattice parameters are shown in Figure 17. The discussion will focus more on VO2 (R) and its reversible MIT to VO2 (M) as well as on VO2 (B) due to its application as cathode materials in electrochemical devices.

Figure 17.

Figure 17

Different phases of VO2 with respective lattice parameters. Reproduced with permission from ref (147). Copyright 2016 under CC BY 4.0 license.

VO2 was first demonstrated to undergo MIT in 1959 by Morin. At the critical temperature (τC) of about 340 K, VO2 transforms from high-temperature, conducting rutile VO2 (R) to low-temperature, insulating monoclinic VO2 (M). There are two mechanisms that have been used to describe this ultrafast phase transition phenomenon in VO2, namely the Peierls model and Mott-Hubbard model. The Peierls model describes the nature of MIT as the change from a shared d-orbital between all vanadium atoms to localized d-orbital in the V–V dimer, which is a result of the change in the V–V distance from 2.88 Å in VO2 (R) to 2.65 and 3.12 Å in VO2 (M).20 Therefore, the Peierls model stated that the structural distortion is the cause of MIT. Wentzcovitch et al. applied the local-density approximation to study the electronic and structural change of VO2 during the MIT.140 From the view of band theory, a monoclinic distorted state is in good agreement with the experiment result. Meanwhile, the structural distortion enhances the bonding between neighboring V atoms, which is expected in the Peierls model. Moreover, Baum et al. utilized four-dimensional (4D) femtosecond electron diffraction to visualize the phase transformation process of VO2.141 They pointed out that during the MIT, the displacement of atoms happened within picoseconds, and followed by the sound wave shear motion of the crystal in the time scale of nanoseconds. The observation indicates the occurrence of fast structural distortion during the MIT.

On the other hand, the Mott-Hubbard model states that MIT would occur when the electron density (ne) and Bohr radius (aH) satisfy ne1/3aH ≈ 0.2.142 Compared with the Peierls model, the Mott-Hubbard model has the advantage in explaining the phenomenon such as the anomalously low conductivity in the metallic phase.143 Whittaker et al. summarized multiple experiment cases and pointed out that the metallic phase of VO2 might be introduced without the structural phase transformation if the excitation of carriers reaches a threshold density.144 Their observation provides key evidence in revealing the nature of MIT. While there remains a debate on which mechanism best describes the MIT, the usage of both models is strongly encouraged due to the transition kinetics145 as well as the stimuli involved during the transition. Shao et al.146 reported recent progress in understanding the mechanism and kinetics of MIT, including the lattice distortion and electron correlations (Peierls phase transition, Mott phase transition) and modulation methods (elemental doping, external electric field, light irradiation, and strain engineering).

Upon application of suitable external stimuli (i.e., photons, heat, electric, magnetic, electrochemical, and stress) to initiate the MIT, physical properties of VO2, such as electrical resistance, optical transmittance, and thermal conductivity, are reversibly and drastically changed. Long et al.27 have summarized the connections between the stimuli and responses of VO2-based devices in detail. This flexibility in external stimuli and corresponding responses is the main reason why VO2 is a material of interest in multiple novel devices spanning from thermally,139,148 electrically,149151 to optically152,153 activated devices. The functional performance of these devices is thus highly dependent on, but not limited to, the physical attributes, such as dimension, morphology, doping level, and crystallinity, of the fabricated VO2. Multiple techniques have been used to fabricate functional VO2 devices, each with their own strengths and weaknesses. Some of these fabrication processes are reviewed with polymer-assisted deposition methods,154 hydrothermal,139 sol–gel,155 chemical vapor deposition (CVD),19 and physical vapor deposition (PVD)156 methods discussed in detail regarding the controlled synthesis of VO2 for thermochromic application.

6.2. Applications

In recent years, multiple omnibus reviews157159 have been done in attempts to give the most encompassed view of VO2 research progress, often including a combination of the following topics: MIT mechanism, kinetics, fabrication techniques, and applications. While these reviews could offer wide coverage of VO2 research progress in multiple topics, in-depth discussion of each topic, applications in this case, is in much needed demand. In subsequent sections, reported applications of VO2 in functional devices both based and not based on MIT in recent years, especially in the last three years, are classified and discussed in different categories: optical, electrical, and mechanical applications.

6.2.1. Optical Applications

The operations of these VO2-based optical devices are often based on the changes in refractive index, n, and extinction coefficient, k of VO2 upon crossing the MIT. The changes to n and k at different temperatures are shown in Figure 18. These optical functional devices are further divided into two main groups: infrared regulators and optical switches.

Figure 18.

Figure 18

Optical constants n and k of VO2 film with respect to the wavelength and the temperature changes. Reproduced with permission from ref (160). Copyright 2019 John Wiley and Sons.

6.2.1.1. Infrared Regulators

Since its first introduction in 1985,161 the smart window has attracted much attention due to rising energy consumption in the commercial building sector, which contributes up to 40% of total consumption globally and leads to 30% of global greenhouse gas emissions today.162 VO2 is the prime candidate for smart window materials due to its ability to seamlessly and rapidly regulate the amount of infrared (IR) across MIT with miniscule side effects to the visible transmission. However, intrinsic limitations (high τC ≈ 68 °C, low luminous transmittance (Tlum) < 40%, poor solar modulation (ΔTsol) < 10%, and poor durability) prevent pristine VO2 from meeting the requirements for commercial smart window applications. The most common way to reduce τC is by elemental (i.e., W, Mo, Ti, F, Mg, etc.) doping as summarized by Cui et al.163 Different approaches exist to enhance Tlum and ΔTsol via advanced device morphological engineering such as multilayered VO2,164 biomimetic structure,165168 nanothermochromism,169 porous,170 and gridded structure.171173 Zhou et al. recently reported a new customized VO2 composite structure in which a new factor, the incident angle, was considered in the development of smart window devices. Figure 19a shows a schematic of how different incident angles in the summer and winter can be taken advantage of with the reported customizable composite structure. The VO2 composite structure was fabricated through a 3D printing procedure, enabling the flexibility of device dimensions. Several samples of different thicknesses, widths, and spacing are demonstrated in Figure 19b. The incident-angle dependency properties of this structure are shown in Figure 19c,d in which modulating performance of the 3D printed device improves significantly, and ΔTsol improves from 9.7% to 25.8% with respect to an increase in the incident angle from 0° to 45°.173

Figure 19.

Figure 19

(a) Schematic design of the 3D printed smart windows design. (b) Optical microscopic pictures of the printed composite structures with identical 0° tilted angles with a range of thickness, spacing, and width. (c) UV–vis–NIR transmittance spectrum of the printed composite structure with respect to different incident angles and temperature. (d) Tlum, ΔTsol, and ΔTIR diagram of the printed composite structures with different incident angles. Reproduced with permission from ref (173). Copyright 2020 John Wiley and Sons.

Aside from NIR transmittance regulation in smart windows, VO2 also has an unusual ability to change its long-wave infrared emissivity (εLWIR) upon crossing the MIT. An ideal smart window should have high transparency in the visible region (380–780 nm), while having a transparent state in the winter and an opaque state in the summer (Figure 20a). Moreover, the ideal smart window should have a high εLWIR at a high temperature to promote radiative cooling (RC) and a low εLWIR at a low temperature to suppress RC. Based on this concept, Long et al.174 fabricated a VO2-based multilayer structure which was able to regulate NIR transmittance and εLWIR spontaneously (Figure 20b). Through forming a Fabry–Perot resonator, the passive RC regulating thermochromic (RCRT) smart window possessed an εLWIR of 0.21 at 20 °C, while the εLWIR increased to 0.61 above τc. In addition, the RCRT window kept a promising Tlum of 27.8% and a ΔTsol of 9.3% (Figure 20c). With the actual building energy consumption simulation conducted with a 12-story building, the RCRT window yielded a higher energy savings compared with a commercial low-E window across different climate zones (Figure 20d). Meantime, Wu et al.175 designed a flexible temperature-adaptive radiative coating (TARC) through embedding lithographically patterned W-doped VO2 in a dielectric BaF2 layer on top of a reflective gold layer (Figure 20e). The TARC film had a low εLWIR (∼0.2) in the insulation state and a high εLWIR (∼0.9) at the metallic state (Figure 20f), and the observation agreed with the simulation (Figure 20g). Long et al.176 further expanded the concept of RC regulation from window to wall by preparing a switchable interwoven structure. As shown in Figure 20h, through pulling the block of interwoven structure, the original exposed block on the top side becomes concealed, and the underneath block becomes exposed. As a result, the structure switches its phase from phase 0 to phase 1. Taking into account different requirements of windows and walls, Long’s group designed on-demand interwoven structures. Figure 20i shows the interwoven structure for window and wall applications. As discussed in Figure 20a, a window requires high visible transmittance and dual-band regulation for NIR and LWIR ranges. An ITO/VO2/PVC combination was employed for windows. In this structure, VO2 was used to regulate NIR, while the εLWIR was regulated by alternatively exposing low-E ITO and high-E PVC. On the other hand, an ideal wall has high solar absorption and low εLWIR in the winter, and low solar absorption and high εLWIR in the summer. An ITO/black paint/PVDF-HFP combination was utilized to cater to this demand. On cold days, visible transparent ITO is exposed, and sunlight will be absorbed by black paint underneath, while on hot days, the highly solar reflective high-E PVDF-HFP is exposed to prompt RC. Hence, compared with a conventional performance index Tlum and ΔTsol, the newly proposed εLWIR needs to be included to gauge the real energy-saving performance.177 Moreover, VO2 is notoriously known for poor durability, which is the bottleneck for the applications in smart windows. There are two recent reports to embed VO2 in the V2O5 matrix, and such a strategy could increase the lifetime up to 33 years.178,179

Figure 20.

Figure 20

(a) Concept of the ideal energy-saving smart window. The red and blue lines represent the spectra for an ideal energy-saving smart window in the summer and winter. (b) Schematic structure of the RCRT window. (c) Spectra of the sample with RC regulation in the visible-NIR and LWIR range at 20 °C (blue line) and 90 °C (red line) against a normalized AM1.5 global solar spectrum (yellow shadow) and LWIR atmospheric transmittance window (blue shadow). (d) Estimated world heating and cooling energy-savings of a W-doped max Δε sample against a commercial low-E glass as the baseline. Reproduced with permission from ref (174). Copyright 2021 American Association for the Advancement of Science. (e) Schematics of materials composition and working mechanism of the TARC. (f) IR images of TARC compared with those of two conventional materials (references) with constantly low or high thermal emittance showing the temperature-adaptive switching in the thermal emittance of TARC. (g) Solar spectral absorptance and part of the thermal spectral emittance of TARC at a low temperature and a high temperature. Measurements (solid curves) show consistency with theoretical predictions (dashed curves). Reproduced with permission from ref (175). Copyright 2021 American Association for the Advancement of Science. (h) Demonstration of the surface transition in a meter-scale Al-paper sample. (i) Photographs and corresponding thermal images of the ITO/VO2/PVC sample (“window” in the figure) and the ITO/black paint/PVDF-HFP sample (“wall” in the figure) on the two phases. Effective areas are marked by the dashed lines. Reproduced with permission from ref (176). Copyright 2022 American Chemical Society.

Besides the application in building, the unique property of emissivity switching makes VO2 the material of interest for IR camouflage and passive radiator for military and aerospace applications because VO2-based devices can function entirely on the thermal trigger with no additional sources, electrical or otherwise, required. VO2-based camouflage devices work by reducing the amount of IR emitted into the environment, shrouding the user from being detected with an IR detector such as most night-vision technologies. Examples include VO2/graphene/CNT heterostructure by Xiao et al.,180 VO2/carbon hybrid by Wang et al.,181 and VO2/ZnS core–shell structure (Figure 21a) by Ji et al.182 As seen in Figure 21b, under a similar IR detector and temperature, the VO2/ZnS core–shell pallets exhibit the ability to control their IR radiation intensity and lower their detected temperature as compared to V2O5 pallets with constant emissivity.

Figure 21.

Figure 21

(a) Schematic of the VO2/ZnS core–shell nanopowder design. (b) The infrared thermal images of same actual temperature for VO2/ZnS core–shell nanopowder and reference V2O5 pellets at 45 °C (upper) and 90 °C (lower). Reproduced with permission from ref (182). Copyright 2018 Elsevier.

Different from a camouflage device, a VO2-based passive radiator requires modification to the device structure to counter the lower emissivity at the higher temperature. This intrinsic problem could be overcome by depositing VO2 on a highly reflective metal substrate with183,184 or without185 a spacer layer. Figure 22a is a multilayer Si/VO2/BaF2/Au structure reported by Kim et al.,184 which was designed and tested specifically for simulated space (vacuum) applications. Figure 22b demonstrates the radiated thermal power of the multilayer device. Experimental data were compared with simulated ones for both high and low-temperature operations. The measured radiated power at 300 and 373 K was 72 W/m2 and 552 W/m2, respectively, showing a massive jump in emitted radiation upon crossing the MIT threshold.184 Due to the typical multilayer design of a VO2-based passive radiator, factors such as functional emissivity difference between low and high temperature as well as the wavelength of emitted light can be further fine-tuned by adjusting the substrate/spacer/film combination. The electrochromic setup has also been shown to also result in IR regulating behavior in VO2-based devices.186Figure 22c is the schematic of a three-terminal thin-film-transistor-type electrochromic device by Katase et al.186 Upon application of external voltage (+12 V according to literature), the VO2 channel undergoes protonation, and the device becomes IR opaque, similar to smart window applications. When a reversed voltage is applied, deprotonation happens, and the device becomes IR transparent once again. Figure 22d shows optical transmittance spectra measured during this transition with +12 V stimulus. The optical transmittance modulation ratio at λ of 3000 nm was 49%.186 Electrical input into a VO2-based device can also be utilized as a Joule heating source for MIT. To realize this, VO2 can be combined with transparent conductive electrode materials such as ITO, Ag NWs, CNT, etc.187,188 to become an electro-optic modulator. An example of such a modulator is a VO2+Au/GaN/Al2O3 device by Fan et al.,189 which has the ability to change the transmission step-by-step according to the applied voltage.

Figure 22.

Figure 22

(a) Schematic of the Si/VO2/BaF2/Au passive radiator design. (b) Radiated thermal power of the Si/VO2/BaF2/Au device, with heating and cooling indicated with arrows; the dotted line is the simulation data for VO2 (M) and VO2 (R) respectively. Reproduced with permission from ref (184). Copyright 2019 under CC BY 4.0 license. (c) Schematic of the three-terminal transistor design. (d) Optical transmittance spectra measured before (red line) and after (blue line) applying V = +12 V. Reproduced with permission from ref (186). Copyright 2017 under CC BY license.

6.2.1.2. Optical Switches

Based on the sudden change in the optical constants n and k of VO2 across MIT, radio frequency (RF) switches or waveguides can be fabricated to control the flow of electromagnetic waves (i.e., microwave and radiowave).190 Even though the design of each device is largely dependent on whether it is thermally191 or electrically activated,120 the mechanism for turning from the ON to OFF state is still entirely based on the transition from VO2(M) to VO2(R) respectively. Examples of RF-switches are the thermally configurable hybrid Al nanoholes/VO2 photonic switch by Sun et al.,192 metamaterial design by Ding et al. which can act as an absorber from 0.562 to 1.232 THz at room temperature and a high-efficiency halfwave plate at high temperature,193 and temperature controlled asymmetric optical switch by Liu et al.194Figure 23a shows the working principal of this design in which different output of the same polarized electromagnetic wave input can be achieved at low temperature by physically reverting the device by 180° while achieving similar output at high temperature. This asymmetrical mechanism is demonstrated in Figure 23b,c. A large asymmetry exceeding 20% was detected at 23 °C, while it disappeared almost entirely at 87 °C (Figure 23b). With incident x-polarized waves, the device gave y-polarized waves output at 23 °C, and this can be considered the ON state. Upon heating to 87 °C, output waves returned to approximately x-polarized, similar to input waves, turning the switch OFF (Figure 23c).194 While the thermal- and electrical-activated optical switches mainly depend on the change of optical constant upon MIT, the optical-activated switches of VO2 focus on the speed of the transition as the defining factor. However, the details of the ultrafast induced phase transition of VO2 are not the focus of this review; it can be found in a summary and discussion by Wegkamp et al.195 VO2 ability to switch between the insulator and metal state within picoseconds is promising for the field of nanophotonics as well as all-optical integrated circuits (i.e., switches, modulators, and data-storage devices). VO2 integrated metamaterials have been reported to exhibit nonlinear transmittance by Liu et al.153 in the THz range as well as broadband responses spanning from the visible to mid-infrared range by Guo et al.196 VO2/Au nanoplate memory device was also reported by Lei et al.,197 giving a stepwise tuning ability with the use of successive laser pulses.

Figure 23.

Figure 23

(a) Schematic of the temperature-controlled VO2 metamaterial asymmetric optical switch. (b) Frequency dependence of the asymmetric transmission parameter for linearly polarized waves, and (c) transmitted polarization state at 1.15 THz for illumination with x-polarized waves. Reproduced with permission from ref (194). Copyright 2019 under CC BY 4.0 license.

6.2.1.3. Plasmonic Applications

The reversible crystal phase transition makes VO2 very unique among plasmonic materials. It undergoes a crystal phase transition from the monoclinic semiconductor state to rutile metallic state with significantly promoted conductivity and free carrier density,27 leading to a significant difference in its plasmonic property. Recently, VO2 nanoparticles (NPs) have been reported with thermal-responsive localized surface plasmonic resonance (LSPR) in the NIR region.168,198 Based on the colloidal lithography method,59 Long’s group successfully produced the hexagonally patterned VO2 NPs on quartz with controllable average diameters from ∼70 to ∼280 nm.198 It was observed that the LSPR position of metallic VO2 shifts to the longer wavelength on the larger NPs (∼1120 to ∼1220 nm) or under the increasing reflective index of the surrounding medium (∼1120 to ∼1360 nm) (Figure 24a). Besides, the NIR LSPR is temperature-responsive that is quenched on a low-temperature semiconductor state and can be gradually switched on from 20 to 100 °C (Figure 24b). They further investigated the LSPR-induced absorbance and scattering effects of VO2 plasmonics through a finite-difference time-domain method.168 On a single VO2 NP, it is revealed that both the absorbance and scattering are low at the semiconductor state (monoclinic, M), while a strong absorbance emerges in the metallic state (rutile, R) (Figure 24c). This result suggests that the LSPR in metallic VO2 is characterized as a strong absorbance enhancement and a relatively weak scattering effect. Moreover, they reported the dispersity- and strain-induced LSPR response on VO2 NPs in the polydimethylsiloxane (PDMS) elastomer matrix (Figure 24d,e).28 The dispersity-induced LSPR position can be attributed to the changes in average gaps among VO2 NPs in the matrix, which is consistent with the simulation result (Figure 24d), while the strain-dependent LSPR position change can be explained by the local reflective index change induced by the delamination between the NP and matrix under applied strains as being demonstrated by the finite element method (Figure 24e,f). A more recent report used a similar approach to tailor the VO2 surface plasmon by manipulating its atomic defects and establishing a universal quantitative understanding.199 Record high tunability is achieved for LSPR energy from 0.66 to 1.16 eV and a transition temperature range from 40 to 100 °C. The Drude model and DFT calculation reveal that the charge of cations plays a dominant role in the numbers of valence electrons to determine the free electron concentration. It is believed the investigation of VO2 LSPR is still in its early stage. The reversible crystal transition makes VO2 an intrinsic active plasmonic material, which is unique among the plasmonic field. It is expected for researchers to further understand the VO2 plasmonic and to explore its potential applications.

Figure 24.

Figure 24

(a) Effects of the particle diameter and medium reflective index to the LSPR position. Insets are the SEM images of VO2 NPs with corresponding diameters, and the array structures are highlighted as yellow hexagons. (b) Extinction spectrum of VO2 NP under different temperatures from 20 to 100 °C. Reproduced with permission from ref (198). Copyright 2017 American Chemical Society. (c) Simulated absorbance and scattering intensity crossing a VO2 NP embedded in PDMS matrix. Reproduced with permission from ref (168). Copyright 2020 Elsevier. (d) Experimental LSPR position of VO2 NPs with different dispersion degrees in PDMS, and the simulation result of two adjacent VO2 NPs with a different gap. (e) Experimental and simulation results of the LSPR position of VO2–PDMS composites under applied strains from 0% to 100%. (f) Simulated stress contours of the representative VO2–PDMS composite under applied strain from 0% to 100%. Reproduced with permission from ref (28). Copyright 2019, Cell Press.

6.2.2. Electrical Applications

Not only optical constants, but the electrical conductivity of VO2 is also altered dramatically upon transitioning from insulating VO2 (M) to metallic VO2 (R). This measurable electrical response to various external stimuli makes VO2 the prime candidate for electrical applications such as sensors or transistors. The following section discusses the electrical applications of VO2 and their corresponding devices.

6.2.2.1. Sensors

A sensor is defined by its ability to measure physical input and translate these measurements into interpretable data. Based on the significant conductivity changing of VO2 across the MIT, it is possible to convert physical environmental input into a readable electrical signal. Some examples of VO2-based sensors include temperature sensors, photodetectors, flexible strain sensors, and gas sensors. Intrinsically, VO2 is not suitable for temperature sensing applications because the change in its electrical conductivity only happens at 68 °C, even though Kim et al.200 managed to fabricate a programmable VO2 critical temperature sensor. VO2 was deposited on an Al2O3 (1010) substrate and between two nickel (Ni) electrodes. A voltage can be applied across these electrodes to cause the τC of VO2 to decrease, causing the VO2 (M) film to go into an intermediate phase before fully transitioning into VO2 (R). During this intermediate phase, the measured current through the device was found to be linear with the change in temperature. At a voltage of 20 V, the τC is found to be ∼20 °C, enabling full range sensing capabilities from 20 to 68 °C. Another approach, which is based on the sensing ability resulting from an abrupt change of the dielectrically constant of VO2 during MIT instead of conductivity, was done by Yang et al.201 As mentioned in the previous section, optically stimulated applications of VO2 are promising due to the ultrafast transition mechanics, stability, as well as the broadband optical response of VO2-based devices. Hou et al.202 demonstrated the device stability and speed of response using a VO2 (M) nanowire on Au electrode setup (Figure 25a). It was reported that the device needs less than 1.6 s to detect IR (980 nm) and <1.0 s to recover (Figure 25b). The device was also reported to maintain responsivity for more than 500 cycles. Another design from Takeya et al.203 combined the photoresponsivity of VO2 film and the localized surface plasmon resonance of silver nanorods. The results indicated a correlation between the incident light transmission and resistivity within a wavelength of 400–900 nm. While the VO2 acted as a photosensitive component, the nanorod array introduced a wavelength and polarization sensitivity to the photodetector. Because the MIT of VO2 results in the change in the lattice structure and constant, it is also possible to induce MIT by causing changes to the lattice through mechanical force, which serves as the basis for VO2 flexible strain sensor application. Hu et al.204 showed that a VO2 strain sensor device could be fabricated by bonding one single VO2 nanobeam to a polystyrene (PS) substrate with silver paste and measuring the resistivity of the nanobeam as tensile and compressive stress is applied along the length of the nanobeam. In this study, VO2 (R) was not formed, and only the VO2 (M2) phase was formed due to the constraint of force applied (only 0.25% as compared to the required 2% at room temperature). However, the device showed remarkable potential when exhibiting a stepwise response to as small as 0.05% tensile or compressive strain.

Figure 25.

Figure 25

(a) Schematic of the VO2 nanowire photodetector. (b) Time-dependent photodetection capabilities of the VO2 nanowire device. Reproduced with permission from ref (202). Copyright 2018 Elsevier. (c) Schematic of the VO2/CNT gas sensor. (d) Time-dependent humidity sensing capabilities of the VO2/SWCNT and n-doped variant at different humidity levels. Reproduced with permission from ref (206). Copyright 2018 Elsevier.

A typical gas sensor design is demonstrated in Figure 25c. For gas sensor application, the semimetallic VO2 (B) phase is more commonly used than the insulating VO2 (M) phase to maintain sensing capability at room temperature. VO2, regardless of the phase, responds to humidity, ammonia (NH3), and nitrogen dioxide (NO2).205,206 Compositing VO2 with carbon species such as single- or multiwalled carbon nanotubes (SWCNTs or MWCNTs) was reported by Evans et al.206 The setup was effective in creating a stable, responsive VO2–CNT gas sensor. Figure 25d shows excellent response and good recovery of VO2–SWCNT to different humidity levels. The resistive response was increased dramatically from 0.5 for pure VO2 (B) to 2.7 for VO2–SWCNT and 7.1 for VO2-MWCNT at 50% humidity. This p-type gas sensing response was also reported for NH3 in the same study despite the longer and lower recovery level recorded. Depending on the applications, property change across the MIT is not the only viable option to use VO2 in a functional device.

6.2.2.2. Electrical Switches, FETs, Oscillators, and Memristors

Different from sensing applications, the MIT of VO2 can be deliberately triggered with programmable duration and patterns to great advantage in electrical switching, FET, oscillator, and memory devices. Similar to optical switches in the previous section, by toggling VO2 across the MIT, it is possible to create an ON/OFF switching mechanism based on the difference in electrical resistance of VO2 (M) and VO2 (R). It has been demonstrated by Zhou et al.207 that a two-terminal VO2-based switching device can have ultrafast, reliable 2 orders of magnitude ON/OFF toggling ability within 2 ns. While the MIT in this report was induced by an applied current, an electrical switch activated by Joule heating was also reported in a separate study by Li et al.208 Mott FET is a gated FET device in which the conventional semiconductor channel is swapped with a Mott insulator, a material with the ability to switch from insulator to metal through external voltage to the gate. VO2, as a Mott insulator, is the prime material for novel Mott FET studies. An example of a typical Mott FET setup was reported by Yajima et al.209 in which a large current modulation can be observed at 315 K, indicating a positive-bias gate-controlled MIT near τC of VO2. Another novel Mott FET design was also fabricated by Shukla et al.210 with a VO2 as a source terminal. This design functioned well at room temperature (300 K) with reversible MIT triggered by the critical applied current. Taking advantage of the VO2-based Mott FET designs and combining it with a ferroelectric material, Zhang et al.211 fabricated a nonvolatile ferroelectric FET (FeFET) device with VO2 nanowires as the channel and Pb(Zr0.52Ti0.48)O3 (PZT) thin film as the dielectric gate (Figure 26a). The novel FeFET device was reported to achieve up to 85% resistance change under a gate voltage of 18 V (Figure 26b). Interestingly, the presence of the ferroelectric materials created a polarization effect after the applied voltage was removed, in which the channel resistance could attain up to 50%. Through this mechanism, it is possible to achieve multiple resistive states by the sweeping suitable gate voltage. To overcome the disadvantage of solid-gate oxide dielectric FET, such as current leakage, which might interfere with the MIT of VO2, ionic liquid (IL) and solid-state electrolyte gating have been the research interest for VO2 FET devices in recent years.212 However, the mechanism in which IL drives the MIT of VO2 is still a debate between different studies. Nanako et al.151 suggested that the underlying mechanism is the bulk carrier delocalization caused by the electrostatic effect. On a different train of thought, Jeong et al.213 attributed the transition to the field-induced creation of oxygen vacancies, rather than the purely electrostatic effect. Ji et al.214 and Shibuya et al.,215 however, suggested that electrochemical protonation was the origin of the modulation of electrical property in VO2, similar to what was observed in the electrochromic setup in the previous section. An electronic oscillator is a common component in modern electronic circuitry which can produce periodic signals such as a square wave or a sine wave. Due to the periodicity of the output, it is often used to convert a direct current (DC) input into an alternate current output. The two main types of electronic oscillators are the linear (harmonic) and nonlinear (relaxation) oscillator. Because of the ability to undergo a nonlinear MIT, VO2 can be used as the basis for a nonlinear oscillator with a relaxation behavior stimulated by external electricity input. A VO2-based oscillator design by Leroy et al.216 and its IV characteristic curve is shown in Figure 26c. The inset shows the schematic of the oscillator circuit in which a resistor (Rs) is connected to the VO2 device to produce a current-controlled negative differential resistance (NDR). The NDR portion happens when VO2 enters the transitive state between VO2 (M) and VO2 (R). It was reported that by controlling Rs, the VO2-based oscillator circuit can become self-sustaining, and the frequency can range from kHz to 1 MHz.152 Aside from the standard setup as shown in Figure 26c, studies have also been done in which two oscillators are coupled with a resistor, a capacitor, or FET in between.217 A memristor is a nonvolatile electronic memory device which has a programmable resistance. The resistance of a memristor is retained even after removal of the power and is dependent on the original applied voltage. It is crucial that the resistance can be reversed or reprogrammed. Thus, a two-terminal VO2 electrical device with the nonvolatile switching of resistance across the MIT can also be adapted into memristors.149 Bae et al.150 demonstrated a two-terminal memristor based on a single VO2 nanobeam. The nanobeam undergoes MIT when a bias of 3 and 5 V was applied for 0.25 s; the resistance of the device goes from an initial 1011 Ω to 109 and 108 Ω respectively. The resistance change can be reset with a zero-voltage bias. VO2 has also been utilized in other memory devices including a multistate free-standing VO2/TiO2 cantilever,218 resistive random-access memory (ReRAM) devices.219 and 3D memory array.189 Not included in the above discussion is the minor application of VO2 in field emitter and spintronic devices which are based on the abrupt drop in resistance across thermal- and magnetical-activated MIT, respectively. Studies on VO2/ZnO core–shell nanotetrapod thermal-activated field emitters were reported by Yin et al. in 2014.220 On the other hand, VO2-based spintronic devices and the behavior of the magnetoresistance of VO2 were reported in detail by Li et al.,221 Choi et al.,222 and Singh et al.223

Figure 26.

Figure 26

(a) Schematic of the VO2–NW-FeFET design. (b) Resistance change of the VO2–NW-FeFET ranging from 0 to 18 V. Reproduced with permission from ref (211). Copyright 2020 Royal Society of Chemistry. (c) IV characteristic in voltage- and current-mode of a device incorporating a VO2 pattern; inset is the schematic test circuit. Reproduced with permission from ref (216). Copyright 2012 Cambridge University Press and the European Microwave Association.

6.2.3. Mechanical Applications

The actuator is typically a component in a machine or a system which converts provided energy into mechanical motion. The concept of the actuator has been widely adapted into novel scientific research, especially in the field of microrobotics or micro-/nanoelectromechanics.224 VO2, which has a high theoretical work density (≈ 7 J cm–3) and fast response rate to external stimuli, is suitable for actuator applications. It offers the ability to offset disadvantageous low work density and the slow response rate of common actuator materials such as piezoelectric ceramics or polymers and CNT, respectively.225 In device fabrication, single crystal VO2 or composite bimorph of VO2 can be designed to respond to specific external stimuli such as light, heat, or electrical current. An example of a photodriven VO2 bimorph design was reported by Ma et al.226 in 2018. The VO2/CNC device was conceived by combining the carbon nanocoil (nanosprings twisted by hollow carbon nanofibers) core with a VO2 shell. When exposed to 980 nm radiation, the temperature of the spring increases unevenly, forming a temperature gradient from tip to end. This results in a transition gradient in which the tip becomes VO2(R) first and shrinks, creating the curvature. The VO2/CNC actuator delivers a large displacement-to-length ratio (∼0.4), fast response rate (9400 Hz), and long durability (>107 cycles). More recently, Shi et al.227 fabricated thermal-activated single-crystalline VO2 actuators (SCVAs) which were designed so that the τC of a single VO2 nanobeam is a gradient along the radial direction. When exposed to heat, one side of the fabricated W-doped VO2 nanobeam with lower τC would undergo MIT first and shrink, creating a bending as seen in Figure 27a. It was reported that this SCVAs design performed competitively with other reported VO2 bimorph actuator designs with an extremely high displacement-to-length ratio (∼1), high energy efficiency (∼0.83%), fast response rate in the order of kHz, and long durability (>107 cycles) (Figure 27b). VO2 electrothermal devices with joule heating activation for oscillator228 and microelectromechanical systems (MEMS)229 have also been fabricated with variable degrees of success. A resonator is a device that exhibits resonance at its eigenfrequency. With VO2, the eigenfrequency of a resonator can be dynamically controlled by thermally triggering the MIT. This specific frequency is positively related to Young’s modulus of VO2, which is widely different between the monoclinic phase (151 ± 2 GPa) and the rutile phase (218 ± 3 GPa).230 Studies have been made to determine the effects of elemental doping on the frequency modulating ability of VO2-based resonators. Rúa et al.231 compared the Cr-doped VO2 resonator with the undoped one and concluded that the doped sample had a higher frequency change due to a lower Young’s modulus. Other strategies to improve performance such as changing the shape from a simple cantilever have also been done by Manca et al.232

Figure 27.

Figure 27

(a) Optical images of the VO2 nanobeam undergoing MIT. (b) Amplitude versus cycle number plots of the nanobeam subjected to a chopped laser (100 Hz in a vacuum and 2000 Hz in air). Reproduced with permission from ref (227). Copyright 2019 John Wiley and Sons.

6.2.4. Supercapacitors

VO2 (B) with its layered structure and multioxidation states is ideal as an electrode material with charge storage through insertion and fast surface Faradaic reaction.233 However, being in common with all metastable VO2 phases, VO2 (B) structural instability is not suitable for cyclic stability of SC application. Multiple studies have been done to combine VO2 (B) with various carbon composites to create stable electrode materials.234 An example of a VO2 and reduced graphene oxide (VO2 (B)/rGO) device by Liu et al.235 is shown in Figure 28. The schematic diagram of the all-solid-state sandwich-structured supercapacitor design is shown in Figure 28a, where symmetrical VO2 (B)/rGO and PVA/LiCl gel are used as electrodes and electrolytes, respectively. The performance of this design was reported to have a superior specific capacitance of 353 F g–1 at 1 A g–1 and a maximum power density of 7152 W kg–1 at an energy density of 3.13 Wh kg–1. By compositing VO2 (B) with rGO, 78% capacitance was retained after 10000 cycles, improving the stability of the device immensely (Figure 28b).

Figure 28.

Figure 28

(a) Schematic illustration of the all-solid-state supercapacitor. (b) Cyclic stability test of the VO2/rGO supercapacitor device. Reproduced with permission from ref (235). Copyright 2019 under CC BY 4.0 license.

6.2.5. Magnetic Refrigeration

Magnetic refrigeration is a cooling technique that is based on the magnetocaloric effect (MCE). The MCE describes the phenomenon in which a suitable material can be heated up or cooled down when exposed to a changing magnetic field. Due to the changing magnetization when crossing the MIT, VO2 was first shown to be suitable for magnetic refrigeration application by Wu et al.236 in 2011 with a single crystalline nanorod fabrication technique. Although the potential was shown for VO2 in magnetic refrigeration applications, studies to further improve this are still limited.

6.2.6. Batteries

VO2 formed by edge-sharing VO6 octahedra with a unique bilayer structure exhibits a large lattice spacing that can accommodate Li+ (0.76 Å), Na+ (1.02 Å), K+ (1.38 Å), and Zn2+ (0.74 Å) insertion/extraction. Fan et al.237 designed and synthesized a binder-free VO2 cathode via biface VO2 arrays directly growing on a graphene foam (GF) network (Figure 29a). They constructed a geometric model of bilayered VO2 nanobelts through the growth direction and lattice spacings (Figure 29b). The relatively high stacking rate of the “steplike” VO6 octahedra along the [010] direction determines the preferred growth direction. As a result, the (001) facet of the VO2 nanobelt is the thinnest, and the interlayers between the (200) crystal planes provide a facile channel for Li+ and Na+ diffusion. Meanwhile, the graphene quantum dots (GQDs) coating on the VO2 surfaces can act as highly efficient surface protection to further enhance the Li+/Na+ storage. When as-prepared GF-supported GQD-anchored VO2 arrays (GVGs) are directly used as a LIBs/NIBs cathode, it exhibits two advantages: high ion diffusion sensitization and charge transport kinetics are beneficial to obtain high-rate capacities, and the homogeneous GQDs suppress VO2 dissolution which is in favor of retaining long-term cycles. The GVG electrode delivers a high specific capacity of 421 mAh g–1 at 1/3 C for Li+, which is much higher than that of the uncoated GF@VO2 electrode (391 mAh g–1). It can maintain 151 mAh g–1 even at 120 C,and 94% of the initial capacity can be retained after 1500 cycles (Figure 29c). When it is used as an NIB cathode, it exhibits a specific capacity of 306 mAh g–1 at 1/3 C and good capacity retention (Figure 29d). These results demonstrate that the GVG is an excellent electrode material for Li+/Na+ storage.

Figure 29.

Figure 29

SEM image (a), geometrical model (b), LIB rate performance (c), and SIB charge/discharge profiles (d) of bilayered VO2 nanobelt. Reproduced with permission from ref (237). Copyright 2015 American Chemical Society. The GITT curves (e), rate performances (f) of SA-VO2 and VO2, and the schematic of the enhanced K ion storage ability (g) of SA-VO2. Reproduced with permission from ref (238). Copyright 2020 John Wiley & Sons. The Rietveld refinement result from the XRD data and crystallographic structure (h), GITT curve (i), and most significant changes of lattice parameter in each stage (j) of VO2. Reproduced with permission from ref (239). Copyright 2019 American Chemical Society.

Zhang et al.238 first designed and synthesized a surface amorphized VO2 (B) nanorod (SA-VO2) with a crystalline core and a surface-amorphized shell heterostructure by an interfacial engineering strategy. The crystalline/amorphous heterointerface in SA-VO2 substantially narrows the bandgap, lowers the surface energy, and reduces the K+ diffusion barrier of VO2 (B) via DFT calculations. Therefore, the as-obtained SA-VO2 electrode exhibits a higher reversible capacity of 288.3 mAh g–1 (at 50 mA g–1), superior rate capacity (141.4 mAh g–1), and long-term cyclability (86% after 500 cycles at 500 mA g–1) (Figure 29e), while the VO2 only delivers a specific capacity of 147.2 mAh g–1 at 50 mA g–1 and maintains 16.5% capacity after 200 cycles at 500 mA g–1 (Figure 29f). Compared with oxygen-rich defect amorphous shell VO2, the crystalline/amorphous heterointerface SA-VO2 enhances the K+ storage capacity and enables rapid K+/electron transfer, which results in large capacity and outstanding rate capability (Figure 29g).

Mai et al.240 reported VO2 hollow microspheres with a high surface area and excellent structural stability via a facile and controllable ion-modulating approach. VO2 hollow microspheres deliver the best Li+ storage performance compared to six-armed microspindles and random nanowires. The highest surface area of VO2 hollow microspheres can provide efficient self-expansion, self-shrinkage buffering, and self-aggregation during lithiation/delithiation, which delivers 3 times higher capacity than that of random nanowires. In addition, they also synthesized highly homogeneous VO2 nanorods by a rapid and simple hydrothermal method for aqueous ZIBs cathode material (Figure 29h).239 The in situ XRD and ex-situ XPS/TEM results demonstrate that the VO2 undergoes a single-phase reaction during the discharge process, accompanying a phase transition process of VO2–Zn0.07VO2–Zn0.29VO2–Zn0.54VO2 with a unit cell volume expansion of 6.69%. On the contrary, the evolution of Zn0.54VO2–Zn0.25VO2–Zn0.09VO2–Zn0.04VO2 occurs during the Zn2+ deintercalation from the Zn0.54VO2. Meanwhile, detailed qualitative analysis verified that the VO2 unit cell expands in the a, b, and c directions sequentially during the discharge/charge processes. Satisfactorily, the VO2 nanorods deliver a high specific capacity of 325.6 mAh g–1 and excellent long cycle performance (86% after 3000 cycles), which is outstanding performance among the reported cathode materials of the aqueous ZIBs (Figure 29i,j).

6.2.7. HER, OER, and Water Splitting

VO2 is a well-known semiconductor material with a band gap of 0.7 eV, which is seldom considered as a candidate material as a catalyst or photocatalyst for the production of hydrogen.241 VO2 can be used as a photocatalyst for hydrogen evolution through phase engineering. Ajayan et al.242 synthesized the body-centered-cubic nanostructured VO2, which shows excellent photocatalytic activity with a hydrogen production rate up to 800 mmol m–2 h–1 from a mixture of water and ethanol under UV light at a power density of ∼27 mW cm–2. Furthermore, vanadium oxide composites have the great potential to accelerate water dissociation kinetics and reduce charge-transfer resistance.243,244 Tao et al.245 synthesized MoS2/VO2 hybrids by using a two-step hydrothermal method. The phase transition of VO2 exhibits a significant effect on hydrogen evolution properties of the heterostructures (an onset potential of 99 mV and a Tafel slope of 85 mV dec–1). The enhanced performance is mainly due to the faster electron transport as well as the strain effect on MoS2. Tu et al.246 fabricated Co3O4/VO2 heterogeneous nanosheet structures on carbon cloth (Co3O4/VO2/CC) by the combination of hydrothermal and electrodeposition methods. The Co3O4/VO2/CC composites gave good HER performance with a low overpotential of 108 mV at 10 mA cm–2 and a Tafel slope of 98 mV dec–1, which results from the abundant active sites, effective electron transport, and improved hydrogen binding energy. Najafi et al.247 prepared room temperature-stable metallic rutile VO2 nanosheets by the topochemical transformation of two-dimensional VSe2. By an O2 plasma pretreatment of the VSe2 nanosheets, the obtained VO2 nanosheets show a porous structure, which shows good HER and OER performances in either acidic or alkaline media. The symmetric two-electrode water splitting cell based on the porous VO2 nanosheets as both the anode and cathode delivers a current density of 10 mA cm–2 at cell voltages of 1.710 and 1.660 V in 0.5 M H2SO4 and 1 M KOH, respectively.

7. V2O5

7.1. Structures and Synthesis

V2O5 has the highest oxygen state in vanadium–oxygen systems and is the most stable member of the series of vanadium oxides. V2O5 has multiple distinctive polymorphs, including α-V2O5 (orthorhombic), β-V2O5 (monoclinic or tetragonal), and γ-V2O5 (orthorhombic). Among them, the most common α-V2O5 is the thermodynamically stable phase (the unit cell structure belonging to the Pmnm space group with lattice parameters of a = 11.150 Å, b = 3.563 Å, and c = 4.370 Å), and the other two phases (β-V2O5 and γ-V2O5) can be transformed from the α-V2O5 phase under high pressure and high-temperature conditions.248 The orthorhombic structure of α-V2O5 is shown in Figure 30, in which each single layer of V2O5 consists of edge- and corner-sharing square pyramids, and the adjacent layers are bonded together along the c-axis by weak van der Waals bonds between the vanadium and oxygen of neighboring pyramids. Additionally, three different oxygens atoms, O1, O2, O3, have different coordinations depending on the position in each layer. The terminal/apical and bridging (corner-sharing) coordinated vanadyl oxygen atoms O1 and O2 have V–O bond lengths about of 1.54 and 1.77 Å, respectively; the triply coordinated O3 links three vanadium atoms via edge-sharing VO5 square pyramids, and the three corresponding V–O bond lengths are 1.88, 1.88, and 2.02 Å.249

Figure 30.

Figure 30

Perspective view (a) along the b-axis and (b) along the c-axis of two layers of V2O5, V atoms are gray balls, O atoms are red balls, and weak van der Waals bonds are omitted for clarity. Reproduced with permission from ref (249). Copyright 2018 Elsevier.

The outstanding characteristics of V2O5, such as a layered structure, a direct band gap in the visible-light region, high chemical and thermal stability, electrochemical safety, low cost, and easy preparation, make V2O5 a suitable material for electrochemical energy conversion and storage,250 catalysis,251253 solar cells,254 gas sensors,31 electrochromic devices,255 and optoelectronic devices.256 Compared with bulk V2O5, the nanostructured ones have higher surface to volume ratios, which is beneficial to improve various performances. Over the past few years, a variety of methods, such as sol–gel, hydrothermal, chemical vapor deposition, magnetron sputtering, and atomic layer deposition, have been developed to prepare V2O5 nanostructures.

The sol–gel method has been used to fabricate V2O5 thin films and nanopowders through V2O5 sols, which were prepared by ion exchange, alkoxide hydrolysis, peroxide-assisted hydrolysis, and melt-quenching. The disadvantage of using an ion exchange is the difficulty to control the vanadium concentration as it varies throughout the whole process. In addition, some foreign ions such as Na+ may remain in the gel after ion exchange.257 The alkoxide hydrolysis always involves some expensive raw materials, and the molten V2O5 quenching process will produce toxic gas.

Hence the synthesis of V2O5 sol by peroxide-assisted hydrolysis stands out due to its advantages of being environmental friendly, inexpensive, and requiring simple fabrication. Vanadium metal or commercial V2O5 powders are commonly used as a vanadium source that can be dissolved vigorously in a solution of hydrogen peroxide. According to Alonso et al. the dissolution of V2O5 into H2O2 produces unstable diperoxo [VO(O2)2] and then is dissociated to the aqueous solution of [VO2]+ and [H2V10O28]4–.258 Etman et al. also found [H2V10O28]4– is the main species via real-time nuclear magnetic resonance.259 It is noted that, not related to the preparation method, the V2O5 sols are comprised of a fibrous structure, dissimilar from many inorganic sols that are typified by a random aggregate of particle structure.260 These fibrous structures can self-assemble into V2O5 nanofibers upon long-term aging through a coagulation mechanism.261 The obtained V2O5 sol was applied onto substrates via coating techniques, including spin coating, dip-coating, and spray process, and the subsequent drying and heat treatment are necessary to obtain V2O5 films.262264Figure 31a highlight the evaporation, hydrolysis, and subsequent solidification procedure during the formation of V2O5 thin films by dip-coating.265Figure 31b,c shows TEM diffraction patterns and corresponding FFT of V2O5 thin films formed from low concentration dilution (LCP) and high concentration dilution (HCP) using PEG-400 as an additive (HCP-PEG), respectively. V2O5 thin films formed from an LCP precursor show the formation of grains of orthorhombic V2O5 with defined grain boundaries (Figure 31b), while the HCP-PEG samples have a polycrystalline structure without uniformed grain (Figure 31c). Meanwhile, as shown in AFM images in Figure 31d–g, thin films formed from higher concentration precursors have a larger surface roughness (RS). Thermal treatment can further increase the RS due to the formation of crystallites on the thin film surface. The thin film formed from LCP-PEG precursor has the lowest RS of 0.2 nm. Liu et al. studied the substrate effect on the structure and electrical properties of nanocrystalline V2O5 thin films prepared by the sol–gel method.266 They found the annealed V2O5 film on the Si substrate exhibited more uniform rod-like morphology, and electrical measurements indicated the typical n-type semiconducting behavior. Senapati et al. prepared nanoscale V2O5 films having thicknesses ranging from 92 to 137 nm by spin coating V2O5 sol at different stages of aging.267 They reported the decrease of strain in the films with aging, and the electrical conductivity increased with aging due to the improved crystallinity of the films.

Figure 31.

Figure 31

(a) The formation process of V2O5 thin film through a combination of evaporation, hydrolysis, and solidification is shown optically and schematically; TEM diffraction patterns and corresponding FFT of V2O5 thin films formed from (b) low concentration dilution (LCP) and (c) high concentration dilution using PEG-400 as an additive (HCP-PEG); AFM surface images of as-deposited and post-thermally treated orthorhombic V2O5 films that were dip-coated from (d) LCP, (e) LCP-PEG, (f) HCP, and (g) HCP-PEG solutions. Reproduced with permission from ref (265). Copyright 2015 under CC BY 4.0 license.

Besides thin films, the sol–gel method also provides good control over the size, morphology, doping, and chemical composition of V2O5 powders.268271 Li et al. reported the flower-like V2O5 powders prepared by coagulating V2O5 sol and subsequent annealing crystallization.272 V2O5/graphene hybrid aerogel was prepared by Wu et al. at ambient pressure through a simple one-pot sol–gel method from commercial V2O5 powder.273Figure 32a illustrates the fabrication process and images of VOx nanofibers and graphene oxide sheets. First, graphene oxide (GO) aqueous solution was added to the V2O5 sol under vigorous stirring to induce hydrolysis and in situ recombination of the GO sheets and VOx oligomers. Then, a dark red VOx/GO hybrid gel was obtained after about 5 min because of the rapid formation of intermediate vanadium phases and the growth of nanofibers. After aging for 2 days, the gel gradually changed to deep green, and the V2O5 nanofibers are anchored and in situ grown on the graphene surfaces. The V2O5/graphene hybrid aerogel is so light that it can be lifted by a feather, but it is strong enough to support the weight of 100 g (Figure 32b).

Figure 32.

Figure 32

(a) The fabrication process and structure of the V2O5/graphene hybrid aerogel and the corresponding digital images of different formation stages: (i) hydrolysis of VOx oligomers and self-assembled coordination of VOx nanofibers and graphene oxide sheets; (ii) aging of VOx/graphene oxide gel and growth of VOx nanofibers along graphene oxide sheets; (iii) solvent replacement and drying, and (iv) thermal reduction of graphene oxide, oxidation, and partly crystallization of V2O5. (b) Lightweight V2O5/graphene hybrid aerogel standing on a feather; it can support the weight of 100 g. Reproduced with permission from ref (273). Copyright 2015 Royal Society of Chemistry.

The hydrothermal method has been widely used for the synthesis of a vast range of V2O5 nanostructures with a desired size and morphology, such as nanoparticles,274 nanowires,275 nanotube,276 nanosheets,277 and micro-/nanostructures.278,279 The importance of solubility of precursors, the pH value, the surfactant, as well as the hydrothermal temperature, reaction time, and solution filling factor are highlighted in many references. Li’s research group conducted extensive research on the hydrothermal treatment of V2O5 sol.280 They prepared V2O5 nanoparticles and ultralong nanobelts with the usage of an inorganic V2O5 sol precursor (Figure 33a,b).261,274 The obtained single-crystalline V2O5 nanobelts have a large specific surface area, with width and thickness of 30–200 nm and length in millimeters or even longer (Figure 33b). Strong evidence suggested that the oriented attachment growth mechanism was responsible for the formation of V2O5 nanobelts. Pan et al. reported a one-step solvothermal method to form V2O5 hollow spheres without adding surfactants.281,282 As shown in Figure 33c, the SEM image of the V2O5 hollow spheres has a uniform size of around 1 μm in diameter. They investigated the time-dependent interior structural evolution by TEM and gave the possible growth mechanism of the V2O5 microspheres (Figure 33d): vanadium oxide nanoparticles are first generated by the hydrolysis of VOC2O4 and then aggregation to form solid microspheres in stage I. The solid spheres undergo the initial inside-out Ostwald ripening process and transform to the yolk–shelled structure (stage II). With extended solvothermal reaction, secondary Ostwald-ripening takes place on the preformed solid cores, resulting in the formation of a multishelled structure (stage III). Finally, completely hollow microspheres are obtained as a result of the thorough dissolution and recrystallization of the less stable interior architectures (stage IV). Hence, the interior structure of the VO2 microspheres could be effectively tailored by simply controlling the reaction duration and concentration of the precursor.

Figure 33.

Figure 33

V2O5 with different morphologies: (a) Nanoparticles. Reproduced with permission from ref (274). Copyright 2015 Elsevier. (b) Ultralong nanobelts. Reproduced with permission from ref (261). Copyright 2011 Royal Society of Chemistry. (c) Hollows spheres and (d) growth mechanism. Reproduced with permission from ref (281). Copyright 2013 Wiley. (e) 2D nanosheets. Reproduced with permission from ref (283). Copyright 2015 Elsevier. (f) Urchin-like microflowers. Reproduced with permission from ref (284). Copyright 2012 American Chemical Society. (g) Nanoring. Reproduced with permission from ref (285). Copyright 2010 under CC BY 2.0 license.

The fabrication of two-dimensional (2D) V2O5 nanosheets has been studied by Cao et al.283 As cathode materials for lithium-ion batteries, the resulting 2D V2O5 nanosheets (Figure 33e) exhibit remarkable electrochemical performances, including high reversible capacity, good cyclic stability, and great rate capability. 3D hierarchical vanadium oxide microstructures, including urchin-like microflowers (Figure 33f), have been successfully synthesized by Lou et al. through a solvothermal method.284 The morphologies of the microstructures can be easily tailored by varying the concentration of the vanadium oxalate solution, and the obtained V2O5 microflowers are highly porous with a surface area of 33.64 m2 g–1 giving high lithium storage capacity, and enhanced cycling stability and rate capability. V2O5 nanorings and microloops are rarely reported, but they are very interesting morphologies (Figure 33g). The cation-induced asymmetric strain is the main driving force in making a layered V2O5 coil into a ring structure.285

CVD is widely used for depositing high-quality and high-performance solid materials. As shown in Figure 34a, CVD involves the transfer of precursor molecules, which are either liquid or gaseous to a reaction chamber by a carrier gas (step 1), and then it is followed by the reaction and/or decomposing on the surface of the substrate to produce the desired films (steps 2a, 3, and 4) or powders (step 2b), and the byproducts and unreacted precursor are transported out from the chamber at the end of the process (step 5a, 5b).19 There are several types of CVD systems such as atmospheric pressure CVD, aerosol-assisted chemical vapor deposition (AACVD), atomic layer CVD, plasma enhanced CVD, metal–organic CVD, and so forth.286 The morphology, size, crystal phase, and specific surface area of V2O5 can be affected by various parameters, namely the reaction time, substrate temperature, pressure, precursor properties, and reaction position during the CVD method. SEM images in Figure 34b display the effect of growth temperature on the morphological characteristics of V2O5 coatings. V2O5 grown at 350 and 375 °C has rod-like structures of nonuniform thickness and width, while at 400 °C pellet-like features of V2O5 are observed, and the morphology evolution could be due to the coexistence of both α-V2O5 and β-V2O5.287 Chun et al. prepared V2O5 nanosheets via the reaction of VCl3 vapor with oxygen in the CVD system without a vacuum system.288Figure 34c shows representative optical images and their corresponding SEM images of the obtained V2O5 nanosheets and three distinguished shapes: hexagons, triangles, and truncated triangles. Wang et al. used the CVD method to control the morphologies of V2O5 by changing the reaction distance from the source position using vanadyl acetylacetonate (VO(acac)2) as the vanadium precursor.289 They found that VOx vapor and VO(acac)2 vapor existed simultaneously during the growth process, and the different supersaturation distributions of these two vapors led to three main growth areas. 1. V2O5 thin-films were formed at a high concentration and supersaturation of VOx in the region near the source material (Figure 34d); 2. nanowires with a length of about 10 μm and width about 200 nm were formed at a distance of 18 cm from the source due to the low vapor concentration of VOx (Figure 34e); 3. nanospheres with a diameter of about 200–500 nm were obtained when the source material was far away (about 30 cm) due to the high concentration and supersaturation of VO(acac)2 that was oxidized to V2O5 nanospheres (Figure 34f).289

Figure 34.

Figure 34

(a) Schematic diagram of the CVD process. Reproduced with permission from ref (19). Copyright 2018 Elsevier. (b) SEM images of AACVD grown V2O5 films at 350 °C, 375 °C, 400 °C. Reproduced with permission from ref (287). Copyright 2016 Elsevier. (c) Representative optical and SEM images of the as-synthesized V2O5 nanosheets with well-defined shapes, such as hexagon, triangle, and truncated triangle. All scale bars are 3 μm. Reproduced with permission from ref (288). Copyright 2018 American Chemical Society. (d) The growth schematic diagram of the V2O5 nanomaterials prepared by chemical vapor deposition using VO(acac)2 powder as the precursor, and the formed (e) nanowire and (f) nanospheres at a distance of 18 and 30 cm away from the source, respectively. Reproduced with permission from ref (289). Copyright 2010 Institute of Physics.

ALD is considered a specific type of CVD, which was first introduced in the 1960s and is currently receiving ever-growing attention as a method of choice for the growth of conformal coatings on nanostructures with high aspect ratios.12Figure 35a depicts the typical ALD process, in which the precursor gases sequentially react with a surface to form an ultrathin film through a self-limiting process, and all byproducts and unreacted precursor molecules are purged out of the reactor.290 The primary advantages of ALD lie in subnanometer film thickness and conformality control that profit from the cyclic, self-saturating nature of ALD processes. Moreover, the ALD is unique as it is able to coat complex 3D structures with a high degree of uniformity and smoothness.291,292 Chen et al. successfully fabricated a multiwall carbon nanotube (MWCNT)/V2O5 core/shell sponge by ALD.291Figure 35b shows the experimental flow schematically: the MWCNT sponge structure exhibits a very low density (∼7 mg/cm3) and high porosity (>99%), allowing for a high amount of active material loading; the V2O5 layer of about 16 nm is subsequently deposited on the MWCNT sponge by 1000 cycles of H2O-based ALD; finally, the MWCNT/V2O5 sponge is compressed and assembled in a coin cell battery, which enables de/lithiation in active material within a very short time. Figure 35c,d shows SEM images of the MWCNT sponge before and after ALD V2O5 coating, giving uniform and smooth V2O5 coating. Two different oxidants, O3 and H2O, have been studied during the ALD process, and it was found that as the ALD cycle numbers increased from 100 to 2500, the roughness for the O3-based films kept increasing from 0.7 to 10.4 nm while that for H2O-based films only increased from 0.4 to 1.9 nm (Figure 35e).293 Østreng et al. prepared V2O5 films by ALD using the β-diketonate VO(thd)2 and ozone as precursors.294 They found that the crystallographic orientation, optical properties, band gap, and surface roughness of the V2O5 films were correlated and could be varied by controlling the deposition temperature and film thickness.

Figure 35.

Figure 35

(a) Schematic of a single ALD cycle consisting of half-cycles of Precursor A and Precursor B separated by purge steps to remove excess precursor and byproducts. Reproduced with permission from ref (290). Copyright 2020 Royal Society of Chemistry. (b) Schematic of experimental flow to fabricate MWCNT/V2O5 sponge, and SEM images of MWCNT sponge (c) before and (d) after ALD V2O5 coating. Reproduced with permission from ref (291). Copyright 2012 American Chemical Society; (e) Compares the RMS roughness of the O3-based and the H2O-based films prepared by ALD as a function of cycle number. Reproduced with permission from ref (293). Copyright 2013 Royal Society of Chemistry.

PVD techniques (Figure 36a) involve evaporation and many different modes of physical sputter deposition, in which the primary source of the depositing species is a solid or liquid, as opposed to generally gaseous precursors in CVD. However, chemical reactions can and do occur in PVD systems, such as in the reactive sputtering deposition. The presence of the reactive gas (oxygen or nitrogen) in the chamber can significantly alter the PVD source.295 PVD possesses some unique advantages for the creation of uniform and dense solid thin films that strongly adhere to the substrates. Meanwhile, the thickness, composition, crystallinity, and crystal orientation of the thin film can be well controlled by changing the growth conditions with minimal risk of contamination due to the absence of organic reactants. Another advantage is to sequentially deposit several materials to form well-defined multilayer systems as well as special alloy compositions and structures.296 Magnetron sputtering is one of the most used PVD methods to fabricate a large range of materials, including metal oxides. The application of a negative voltage to the cathode will generate positively charged argon ions that can bombard the target ions to be ejected toward the substrate to form a film. Magnets are used in order to increase ion bombardment. This technique has been developed on an industrial scale to make large surface deposits with a wide variety of materials. V2O5 film consisted of fine long strip particles deposited by radio frequency magnetron sputtering, and the thickness of the V2O5 film was determined to be approximately 150 nm according to the cross-sectional SEM image (Figure 36b). V2O5 films underwent four different thermal transition behaviors to other vanadium oxides that were closely related to the oxygen proportion of the annealing ambient.297 Amorphous V2O5 film can be used as a hole injection layer in quantum dot light-emitting diodes, which exhibited a maximum luminance of 198.5 cd/m2, a turn-on voltage of 1.7 V, and a max external quantum efficiency of about 8.3%.298

Figure 36.

Figure 36

(a) Schematic illustration of different physical vapor deposition techniques: sputter deposition, pulsed laser deposition, electron beam evaporation, and thermal evaporation. In evaporation, atoms are removed from the source by thermal means, while in sputtering they are dislodged from a solid target via bombardment by gaseous ions. Reproduced with permission from ref (296). Copyright 2021 John Wiley and Sons. SEM images of V2O5 prepared by (b) megnetron sputtering. Reproduced with permission from ref (297). Copyright 2021 Elsevier. (c) PLD. Reproduced with permission from ref (299). Copyright 2021 Elsevier. (d) Electron beam evaporation. Reproduced with permission from ref (301). Copyright 2017 Elsevier. (e) Thermal evaporation. Reproduced with permission from ref (306). Copyright 2019 American Chemical Society.

Pulsed laser deposition (PLD) is another PVD method, which has been preferred to grow different structures ranging from high-quality epitaxial thin films to various nanostructured layers. During PLD, a high-power pulsed laser beam is focused on a target of the desired composition. Material vaporized from the target is deposited as a thin film on a substrate that faces the target in an ultrahigh vacuum (UHV) environment. Polycrystalline V2O5 thin film in the desired orientation can be prepared by PLD, which has aligned nanorod morphology on a flexible stainless steel substrate (Figure 36c).299 Huotari et al. found the film surface morphology varied largely according to oxygen partial pressure: lower O2 partial pressures resulted in a denser and thinner film, while higher O2 partial pressures gave a film surfaces formed with randomly agglomerated nanoparticles or agglomerates with pillar-like morphology.300

Electron-beam deposition (EBD) is another form of PVD where a target anode is bombarded with a high-energy electron beam that is given off from a charged tungsten filament under a high vacuum. The electron beam leads to joule heating and converts the target into the gaseous phase, which subsequently precipitates into the solid form on the desired substrate. Han et al.301 reported the growth of nanocolumnar V2O5 molecules that were aggregated with each other and collapsed after annealing treatment (Figure 36d). Most of the nanosized V2O5 columns’ structure could retain its original shape during the annealing process by changing the source from V2O5 to VO2. Highly oriented V2O5 thin films with nanosized grains were grown by EBD, and the film thickness was found to be in the range of 800–1200 nm that varied by adjusting the substrate position.302 Meanwhile, the mobility and carrier concentration of the oriented V2O5 thin films increased with the increase of V2O5 film thickness. Thermal evaporation is the vaporization of a material by heating to a temperature that the vapor pressure becomes appreciable, and the materials are sublimated from the target surface in a vacuum. By this method, heterojunctions,303 nanorods,304 nanoparticles,305 or highly crystalline V2O5 films306 have been studied in several reports. Wang et al. synthesized Ga-doped V2O5 nanorods by thermal evaporation at 850 °C and found interstitial Ga and Ga–O phases influence the photoluminescence properties of V2O5 nanorods.304 Berouaken et al. used thermal evaporation to prepare V2O5 nanoplatelets on the quartz crystal microbalance, followed by rapid thermal annealing.307 The obtained V2O5 nanoplatelets exhibited good sensing performance toward NH3 vapor at room temperature: a fast response time, a short recovery time, good stability, reproducibility, reversibility, and linearity. Velmurugan et al. prepared highly crystalline V2O5 films with a controlled thickness of about 530 nm and an average particle size of around 560 nm using a thermal evaporation process (Figure 36e).306 The films were fabricated in electrochemical microcapacitors and subjected to various electrochemical characterizations, which display improved reliability and excellent capacitance retention.

7.2. Applications

7.2.1. Batteries

V2O5 is a feasible cathode material for metal-ion storage due to its high output voltage and unique crystal structure with large interlayer spacing (4.4 Å). Mai et al.308 designed and synthesized a V2O5 hollow microclew (V2O5–HM) through a facile solvothermal assisted calcination method. The amorphous V2O5–HM precursor can convert into crystalline V2O5 through calcination (Figure 37a). Compared with crystalline V2O5 (V2O5-Ms) and V2O5 nanowires (V2O5–NWs), the V2O5–HMs exhibit the best Li+ storage performance (145.3 and 94.8 mAh g–1 at 0.67 and 65 C, respectively), which is due to the 3D hierarchical microstructure with intertangled nanowires (Figure 37b). This 3D hierarchical microstructure not only inherits fast electrolyte penetration, and short ionic and electronic transport pathways, but also significantly alleviates the strain during the Li+ intercalation/deintercalation. Meanwhile, compared to disordered V2O5 nanowires, a unique V2O5–HM structure effectively helps improve the tap density, which is more suitable for commercial applications (Figure 37c). V2O5 can be used as other metal ions (Na+/K+/Zn2+) electrode material except for LIB storage material. Chung group synthesized a nanosized V2O5/C composite cathode by ball milling the nanosized V2O5 with acetylene black and investigating the reaction mechanism in the NIB system.309 Generally, compared with other vanadium oxides (VO2, V2O3, et al.), V2O5 consists of the square-based pyramid with the highly distorted environment and exhibits the highest pre-edge intensity from the X-ray absorption near edge structure (XANES) result. Thus, the average vanadium oxidation state of the calcinated V2O5/C sample is +4.83 (Figure 37d), which may be due to the calcined carbon reduction (a slight difference with standard V2O5 in the pre-edge peak position). They used ex-situ XRD to demonstrate the major NaV2O5 with a minor Na2V2O5 phase formed at the first discharge state, accompanying a c lattice parameter increase by 9.09% and unit cell volume increase by 9.2%. At the subsequent charge, the NaV2O5 + Na2V2O5 will transform into NaV2O5 + V2O5 along with a V4+ → V5+ change (Figure 37e). The V2O5/C delivers an initial discharge capacity of 195 mAh g–1, and increases to 255 mAh g–1 at the 10th cycle corresponding to 1.7 Na+ inserts into per unit formula (Figure 37f). Besides LIB/NIB cathodes, V2O5 can be investigated as an aqueous zinc-ion battery (ZIB) cathode material, and the intercalation of water into the vanadium oxide can increase the interlayer distance, which is in favor of expanding the gallery for Zn2+ intercalation. Yang et al.310 reported V2O5·H2O/graphene (VOG) synthesized via a freeze-drying method. They investigated the critical role (“lubricating” effect) of structural H2O on the Zn2+ intercalation into bilayer V2O5·nH2O, and H2O-solvated Zn2+ possesses a largely reduced effective charge and improves electrochemical performance. VOG delivers a high capacity of 381 mAh g–1 at 60 mA g–1 and maintains 248 mAh g–1 at a high current density of 30 A g–1, which are much higher than those of most aqueous ZIB cathode materials (Figure 37g,h). The interlayer distances of VOG are 12.6, 13.5, and 10.4 Å at initial, discharge, and charge states by ex-situ XRD and MAS NMR (Figure 37i). These results demonstrate that water in vanadium oxide layers plays an important role in the performance of an aqueous ion battery.

Figure 37.

Figure 37

XRD patterns (a), rate performances (b) of V2O5–HMs, V2O5-Ms, and V2O5–NWs. Schematic illustration of the V2O5–HMs during charge/discharge process (c). Reproduced with permission from ref (308). Copyright 2016 John Wiley & Sons. The XANES spectra of as prepared and after calcination at 400 °C are plotted with solid lines (d), the Na+ de/intercalation channels crystal structure (e), charge/discharge profiles and cycling performance (f) of orthorhombic V2O5/C. Reproduced with permission from ref (309). Copyright 2016 American Chemical Society. The CV curves (g), rate performances (h), and the proposed crystal structures at different states (i) of V2O5·H2O/graphene. Reproduced with permission from ref (310). Copyright 2018 John Wiley & Sons.

To investigate the electrical conductivity and structural mechanism during lithium insertion/deinsertion of V2O5, Yoon et al.311 developed a 3D V2O5/rGO/CNT with short Li+ diffusion, and high continuous 3D conductive network, and investigated its structural mechanism during Li+ intercalation/deintercalation by in situ XRD/XANES analysis (Figure 38a). The 3D V2O5/rGO/CNT delivers a high discharge capacity of 100 mAh g–1 at 20 C, which is much higher than 2D V2O5/rGO (68 mAh g–1). There are numerous metastable phases of LixV2O5 during Li+ intercalation into V2O5. The α-Li0.26V2O5, ε-Li0.93V2O5, δ-Li1.27V2O5, γ-Li1.93V2O5, and ω-Li2.65V2O5 phase form in turn during the first discharge to 3.4, 3.3, 3.19, 2.28, and 2.01 V, respectively. In addition, the most reflection will return to the same position at a pristine state during the subsequent charge process, which can confirm the high structural reversibility of V2O5 in the ternary composite upon Li+ intercalation/deintercalation (Figure 38b). The 3D V2O5/rGO/CNT delivers a high discharge capacity of 100 mAh g–1 at 20 C, which is higher than 2D V2O5/rGO (68 mAh g–1) (Figure 38c). The preintercalation interlayer metal ions can act as pillars to increase the electronic conductivity, ion diffusion rate, and stability of layered vanadium oxides. Mai et al.312 designed and assembled Na0.33V2O5 (NVO) and V2O5 single nanowire devices to investigate the effect on the intrinsic electrical conductivity of Na+ intercalation. The conductivity of NVO is 5.9 × 104 S m–1, while the conductivity of V2O5 is 7.3 S m–1, which indicates that the electronic conductivity of V2O5 is greatly improved by the Na+ intercalation (Figure 38d). Stucky et al.313 fabricated a series of nanostructured Mn-doped V2O5 cathode materials and found that the larger Mn doping in the modified V2O5 structure can increase the cell volume, which facilitates high Li+ diffusion and improves the electronic conductivity (Figure 38e). The Mn0.01V1.99O5 delivers a high discharge capacity (251 mAh g–1 at 1 C) and excellent cycling stability (80% after 50 cycles), which is much higher than V2O5 (215 mAh g–1 vs. 70%) (Figure 38f,g). Fan et al.314 reported lightweight, freestanding V2O5 nanoarray-based positive electrodes (UGF-V2O5/PEDOT), which were prepared by growing a V2O5 nanobelt array directly on 3D ultrathin graphite foam (UGF), followed by coating the V2O5 with a mesoporous thin layer of the conducting polymer poly(3,4-ethylenedioxythiophene) (PEDOT) (Figure 38h). In addition, the PEDOT coating constructs an integrated conductive network for the V2O5, providing decreased electrode polarization, improved charge transfer kinetics, and a prolonged discharge plateau of V2O5 (Figure 38i,j), and therefore it can lead to an increased proportion of high-voltage capacity and energy density than that without PEDOT.

Figure 38.

Figure 38

FESEM image (a), in situ XRD (b), and rate performance (c) of 3D V2O5/rGO/CNT composite. Reproduced with permission from ref (311). Copyright 2016 under CC BY 4.0 license. The IV curves of Na0.33V2O5 (NVO) and V2O5 (d). Reproduced with permission from ref (312). Copyright 2018 John Wiley and Sons. Rietveld refined XRD patterns of the MnxV2–xO5 compounds (e), rate performance (f), and cycle performance (g) at 300 mA g–1 of Mn0.01V1.99O5 and V2O5. Reproduced with permission from ref (313). Copyright 2015 American Chemical Society. The SEM image of UGF-V2O5/PEDOT (h), the charge/discharge profiles (i), Nyquist plots at fully charged stage (j) of UGF-V2O5/PEDOT and UGF-V2O5. Reproduced with permission from ref (314). Copyright 2014 John Wiley and Sons.

7.2.2. Supercapacitors

Among all types of vanadium oxides, V2O5 has attracted attention for the application of SCs due to its broad oxidation states, high specific capacitance, and low acquisition cost. Palani et al. fabricated RuO2 nanoparticle-decorated V2O5 nanoflakes by a solvothermal method.315Figure 39a illustrates the corresponding schematic of the fabricated asymmetric cell that exhibited a high specific capacitance of 421 F g–1 at a current density of 1 A g–1 with excellent cyclic retention of 94.6% over 10000 cycles (Figure 39b). The symmetric device of V2O5||PVA-KOH||V2O5 was fabricated using thin flexible substrate by Velmurugan et al. (Figure 39c), where both annealed (A-500) and as-prepared (RT) V2O5 films were used as electrode material separately.306 A schematic representation of the structural image of the V2O5 is provided in Figure 39d. The CV curves in Figure 39e denote that the symmetric A-500 gives a larger area under the curve than the symmetric-RT, suggesting the improved performance of the annealed sample. The A-500 devices display a maximum energy density of 0.68 μWh cm–2, which is obviously much higher than that of the RT electrodes (0.05 μWh cm–2, Figure 39f). Moreover, the symmetric A-500 shows excellent cycle life up to 30000 cycles with a Coulombic efficiency of 99%. As shown in Figure 39g, the practical feasibility of the as-fabricated devices was demonstrated by lighting blue light-emitting diodes. Several groups reported the hybrid structure of V2O5 with carbon materials fabricated by different strategies for enhancing the SC property. For example, Sahu et al.316 synthesized graphene nanoribbon @V2O5 nanostrip composites to improve the conductive property of V2O5, which displays a high energy density of 42.09 Wh kg–1 and power density of 475 W kg–1. W. Sun et al.317 prepared a 3D monolithic aerogel composed of uniform carbon nanofibers/V2O5 core/shell nanostructures. The composite aerogel exhibits high specific capacitance (595.1 F g–1), excellent energy density (82.65 Wh kg–1), and good cycling behavior (>12000). Zhu et al. proposed a simple “liquid phase impregnation template” strategy to successfully synthesize hierarchically porous V2O5/C nanocomposites that exhibits a specific capacitance of 492.1 F g–1, as well as an energy density of 87.6 Wh kg–1.318 Yao et al. successfully synthesized SCs based on 3D networks hybrids of reduced graphene oxide and V2O5 nanobelts through a simple hydrothermal method.319 The rGO/V2O5 hybrid aerogel electrodes showed a high energy density of 249.7 W kg–1 and excellent long-term cycle stability (remaining 90.2% after 5000 cycles).

Figure 39.

Figure 39

(a) Schematic illustration of the 3 wt % RuO2 in V2O5 asymmetric supercapacitor device and (b) cyclic stability 3 A g–1. Reproduced with permission from ref (315). Copyright 2021 Royal Society of Chemistry. (c) As-fabricated V2O5 flexible thin film, and the inset view is the FESEM image of the film. (d) Schematic representation of the symmetric V2O5 capacitor. (e) Comparison of CV curves of the symmetric SCs with V2O5 electrode prepared at room temperature (RT) and 50 °C (A-500) at a scan rate of 50 mV s–1. (f) Ragone plot comparison of both symmetric-RT and symmetric A-500 capacitors. (g) Lighting of LED using symmetric A-500 connected in series (inset: various bent position of the symmetric A-500). Reproduced with permission from ref (306). Copyright 2019 American Chemical Society.

7.2.3. Catalysts

7.2.3.1. OER

V2O5 exhibited good OER performance due to the multivalent states of the V element, which can enrich active intermediates (*OH, *O, and *OOH) by regulating the valence electron structure of the V element.275,320 The OER performance could be enhanced by fabricating the composites with other materials. Lan et al.320 synthesized CoV2O6-V2O5/nitrogen-doped reduced graphene oxide composites (CoV2O6-V2O5/NRGO) by a one-pot hydrothermal method integrating polyoxovanadate, ethylenediamine (EN), and graphene oxide (GO) for the precursor and postcalcined process (Figure 40a). Without V2O5, the CoV2O6/NRGO delivered a relatively acceptable OER performance with an overpotential of 379 mV at a current density of 10 mA cm–2, which is comparable to that of IrO2 (337 mV). By adding V2O5, the OER performance could be enhanced with an overpotential of 239 mV at a current density of 10 mA cm–2 (Figure 40b). Furthermore, the CoV2O6–V2O5/NRGO exhibits a higher current density (47.08 mA cm–2) at an overpotential of 300 mV compared with CoV2O6/NRGO (0.45 mA cm–2) and IrO2 (3.95 mA cm–2) (Figure 40c). Meanwhile, the CoV2O6–V2O5/NRGO shows the fastest reaction kinetics with a Tafel slope of 49.7 mV dec–1, which could be due to the enhanced charge transport (Figure 40d). The CoV2O6–V2O5/NRGO gives good stability from the polarization curves, which are almost overlapping before and after 1000 cycles (Figure 40e). The theoretical calculation found that the existence of the hydrogen bond between V2O5 and intermediate HOO* of OER decreases the adsorption energy, which may be responsible for the low overpotential.

Figure 40.

Figure 40

(a) Schematic illustration of the synthetic process of CoV2O6–V2O5/NRGO composite, (b) polarization curves, (c) comparison of catalysts’ overpotential at a current density of 10 mA cm–2, corresponding current density at an overpotential of 300 mV, and (d) corresponding Tafel plots. (e) Initial polarization curves of CoV2O6–V2O5/NRGO and after 1000 CV cycles. Inset: time-dependent current density curve of CoV2O6–V2O5/NRGO under a potential of 239 mV for 24 h. Reproduced with permission from ref (320). Copyright 2017 American Chemical Society.

7.2.3.2. HER

In general, V2O5 is generally considered as an HER-inactive material due to the weak H* adsorption on V sites, which limits the formation of H* at the active site.321,322 Three main methods can be adopted to promote the HER performance of V2O5: defects creation, interfacial engineering, and forming composites. First of all, creating defects, especially vacancies, is regarded as an efficient way to enhance the HER performance.323325 Oxygen vacancy (VO) in transition metal oxides can accelerate the adsorption of the H* intermediates by activating the delocalized electrons of the metal center, which could lead to a better HER performance.326,327 Li et al.328 synthesized the V2O5 nanosheet arrays with VO on Co foam through a hydrothermal reaction. The highest VO concentration of 34.2% of the V2O5 nanosheet arrays on the Co foam (Co–V2O5–H) is easily obtained by controlling the pH = 1 of the NH4VO3 precursor solution. The Co–V2O5–H exhibits a low overpotential of 51 mV at a current density of 10 mA cm–2, which shows better performance compared to the V2O5 powder and low VO concentration samples (Figure 41a). Meanwhile, the catalyst shows a negligible potential drop at different current densities, indicating long-time stability (Figure 41b). The Co–V2O5–H is employed as both a cathode and anode to establish a two-electrode alkaline electrolyzer for overall water splitting, which can maintain a steady output current at a cell voltage of 1.6 V for 24 h (Figure 41c). Second, the interfacial engineering between V2O5 and transition metal also provides a route to improve the HER performance. Kim et al.329 directly grew the V2O5 particles on Ni foam via a one-step hydrothermal method. The as-prepared V2O5/Ni(OH)2@NF catalyst shows a low overpotential of 39 mV at a current density of 10 mA cm–2, which is comparable to Pt (35 mV @ 10 mA cm–2). The Ni(OH)2@NF samples show an overpotential of 188 mV at 10 mA cm–2, which indicates that the V2O5 plays an important role in the HER performance (Figure 41d). DFT calculation was performed to investigate the active sites of the V2O5/Ni(OH)2@NF catalyst. The exposed facet of V2O5 is the (010), (001), and (310) planes, which is confirmed by TEM (Figure 41e). The H adsorption energy of the V2O5 (001) and (310) surfaces is larger than that of V2O5 (010), which indicates that V2O5 (010) is the active surface. Furthermore, the ΔGH* values of the Ni-sites at the edges of interfaces in Ni(OH)2@Ni and V2O5@Ni and the O-site of V2O5 (corresponding to 2, 4, and 5, respectively, as shown in Figure 41f) are close to zero, which is nearly equal to that of the Pt(111)-surface. The calculation indicated the edges of the interfaces in V2O5/Ni(OH)2@NF play a significant role in the HER process. Moreover, the hierarchical V2O5@Ni3S2 hybrid nanoarray also exhibited a good overpotential of 95 mV at 10 mA cm–2,330 which is also due to the interfaces between V2O5 and Ni3S2. Third, V2O5 is also used as an additive with other materials to form a composite and further enhance the HER performance. By doping phosphorus and adding V2O5 into Pt/graphene, the prepared catalysts exhibit a good HER performance of the initial potential of 32 mV and a Tafel slope of 23 mV dec–1.331

Figure 41.

Figure 41

(a) HER polarization curves of the Co–V2O5–H sample, (b) stability of the Co–V2O5–H at 10–100 mA cm–2, (c) stability of the Co–V2O5–H // Co–V2O5–H couple at 1.6 V for overall water splitting. Reproduced with permission from ref (328). Copyright 2020 Elsevier. (d) Polarization curves of NF, 20% Pt/C@NF, V2O5@NF, Ni(OH)2@NF, and V2O5/Ni(OH)2@NF at a scan rate of 2 mV s–1, (e) HRTEM image of V2O5/Ni(OH)2@NF, (f) calculated free energies at sites 1 to 6 (1: the reconstructed Ni-surface, 2: the Ni-site of Ni(OH)2@Ni, 3: the Ni-site between V2O5@Ni and Ni(OH)2@Ni, 4: the Ni-site of V2O5@Ni, 5: the O-site on the V2O5 surface, 6: the O-site inside the V2O5 channel) and various individual pristine materials. Reproduced with permission from ref (329). Copyright 2019 Royal Society of Chemistry.

7.2.3.3. Photocatalysis

It is well documented that V2O5 has a typical narrow band gap (∼2.3 eV) and wide optical absorption range and is a high electron mobility semiconductor, which exhibits good photoresponsive properties by capturing visible light and is widely used in the electro-photocatalytic field, such as for hydrogen production, environmental pollutant degradation, etc.332334 Garcia et al.335 found that the morphology of V2O5 nano-/microparticles dramatically affected the hydrogen production by photocatalysis. The V2O5 with microbelts, nanoplates, and nanopillars morphology can be obtained by using sunflowers’ petals and the center of the sunflower as biodegradable templates during the synthesis.335 The nanopillar V2O5 delivers a maximum hydrogen generation rate in the presence of Na2SO3 as a sacrificial agent with 65.5 μmol g–1 h–1, which is higher than microbelts (26.3 μmol g–1 h–1) and nanoplates V2O5 (19.4 μmol g–1 h–1) (Figure 42a). The high hydrogen generation rate is attributed to the large surface area, high absorbance in the UV–vis range, high photocurrent, and high content of defects of the V2O5 nanopillars. By fabricating the nanocomposite with V2O5 containing 1 or 2 heterojunctions, the hydrogen generation rate will be further improved, which is due to the delayed electron–hole recombination.335 For example, graphitic carbon nitride nanosheets/V2O5 composites (578 μmol g–1 h–1),336 Na2Ti3O7/V2O5/g-C3N4 composites (11000 μmol g–1 h–1),337 Na2TiO3/V2O5/g-C3N4 composites (567 μmol g–1 h–1),337 and Nb-doped SnO2/V2O5 (1346 μmol g–1 h–1).338 Transition metal oxides (TiO2, ZnO, etc.) have been applied in the removal of environmental pollutants in water, which could completely decompose organic pollutants into CO2 and H2O by photocatalytic oxidation.339,340 Hollow V2O5 microspheres consisting of randomly packed platelets showed an enhanced UV light absorption compared to the commercial V2O5 powder, which led to the highest activity for degrading rhodamine B under UV light.341 Composites, consisting of 1D V2O5 nanorod and 2D carbon-based materials, are promising photocatalysts for environmental pollutant degradation. The V2O5 nanorods/graphene oxide and V2O5 nanorods/graphene nanocomposites showed good degradation performance of Victoria blue dye and methylene blue dye (>95% degradation within 90 min), respectively.342,343 Especially, the nanocomposites exhibited the best degradation performance under direct sunlight irradiation compared to UV and visible light. The g-C3N4 is also used to construct the heterojunctions with V2O5 for a high-performance degradation catalyst.344 The photoexcited electron in the conduction band of g-C3N4 shows a strong reducing ability, while the photoexcited hole on the valence band of V2O5 exhibits a strong oxidizing ability. Thus, the g-C3N4 /V2O5 heterojunctions exhibited efficient degradation performance of rhodamine B, methyl orange (MO), and methylene blue (MB) dyes under visible light (Figure 42b). Besides the graphene or g-C3N4, the V2O5 composites with other inorganic photocatalytic oxides also showed enhanced photocatalytic properties for degradation of organic pollutants, such as V2O5/BiVO4,332,345 V2O5/CeO2,346 V2O5/TiO2,347,348 and so on. The V2O5-based composite can degrade not only the dye molecules in solution, but also some small solvent molecules in the gas phase. The ternary V2O5/BiVO4/TiO2 nanocomposites exhibited a well-aligned band structure and increasing photoinduced charge carriers through the charges separation across their multiple interfaces, which resulted in good light absorption from the UV to the visible region and better photocatalytic activity for the decomposition of gaseous toluene compared to pure TiO2 and V2O5/BiVO4 under visible light irradiation (Figure 42c).349

Figure 42.

Figure 42

(a) The hydrogen generation curves for VO-microbelts, VO-nanoplates, and VO-nanopillars with a sacrificial agent. Reproduced with permission from ref (335). Copyright 2021 Elsevier. (b) The photodegradation performance of g-C3N4/V2O5 photocatalyst for RhB, MO, and MB degradation under visible light irradiation. Reproduced with permission from ref (344). Copyright 2016 Elsevier. (c) Schematic diagram of lectron–hole pairs separation of the V2O5/BiVO4/TiO2 nanocomposites under visible-light irradiation. Reproduced with permission from ref (349). Copyright 2014 American Chemical Society.

7.2.4. Electrochromism

Since Colton et al. did the pioneering work on the electrochromism of V2O5 in 1976–1977,350,351 V2O5 has attracted increasing attention among transition metal oxides because it can exhibit both anodic and cathodic coloration.352 From 1975 to 1999, studies focused on the absorption-transmission spectra modulation in visible and infrared light, while, after 2000, V2O5 with various micro- and nanostructures were emerging, and researchers paid attention to improve some important key figures of merit to evaluate the electrochromic performance, such as the coloration efficiency, cycling life, switching time, and so on.353358 Recently, several reviews were published on the electrochromic application of V2O5 film, which provide more specific and detailed information on this topic.256,352,359,360 For V2O5, the discoloration mechanism is explained as the result of injection/extraction of electrons and electrolyte cations and variation of valence change of vanadium ions, which is widely accepted.361,362

In order to improve the performance of V2O5 as an electrochromic device, many strategies have been designed. Especially, V2O5 nanostructures with small sizes and large specific surface areas are expected to facilitate the ion intercalation/deintercalation process, thereby enhancing the electrochromic properties. Panagopoulou et al.363 successfully prepared Mg-doped V2O5 thin films using RF sputtering, and found the 15 atom % Mg-doped films displayed optimal electrochromic properties with the fastest switching time of tc = 10/4 s (intercalation/deintercalation), the best coloration efficiency of 71.3 cm2 C1– at 560 nm, higher visible transmittance of 85%, and the highest contrast value between the coloration states of ΔT (34.4% @ 560 nm). Qi et al.364 fabricated flexible V2O5 nanosheets/graphene oxide films, which exhibit ultrafast coloring response time (1.6 s) and bleaching time (2 s) attributed to the reduced charge transport distances of the ultrathin nanosheet structure (4–40 nm). They also display an excellent transmittance contrast of 57.5% at 425 nm and reversible yellow/green/blue-gray multicolor changes. Tong et al.365 fabricated a 3D crystalline V2O5 nanorod architecture on ITO substrates by a colloidal crystal-assisted electrodeposition method. Such architecture exhibits a highly reversible Li-ion insertion/extraction process (Columbic efficiency up to 96.9%), five distinct color change, good transmittance modulation ΔT (38.48% @ 460 nm), and acceptable response times (8.8 s for coloration and 9.3 s for bleaching), making it a promising film electrode for electrochromic devices. Kim et al.366 prepared highly crystalline 2D V2O5 nanosheets by using single-layer V2CTx film as a sacrifice template (Figure 43a). The mean thickness and lateral size of V2CTx are 2.38 nm and 0.78 μm, respectively (Figure 43b). Figure 43c shows the SEM image of V2O5 nanosheets film after annealing of V2CTx at 350 °C. The 2D V2O5 nanosheets based electrochromic device has sharp multicolor transformations with a robust optical contrast from yellow to green to blue (Figure 43d). The corresponding optimal electrochromic performance shows a high optical contrast (53.98% @ 700 nm) and a fast response time (6.5 s for coloration and 5.0 s for bleaching).

Figure 43.

Figure 43

Fabrication of a 2D V2O5 nanosheet based electrochromic device. (a) Schematic for fabrication of an electrochromic device based on a V2CTx-derived 2D V2O5 nanosheet, resulting in minimized optical scattering and enhanced ion diffusion. (b) Representative AFM image of single layer V2CTx with a submicron lateral size. (c) SEM image of cross-section of 2D V2O5 nanosheet based electrochromic layer and corresponding EDS analysis in terms of vanadium, carbon, and oxygen. (d) Digital image of electrochromic device operating from +1.9 V to −1.2 V. Reproduced with permission from ref (366). Copyright 2023 Elsevier.

8. Other Vanadium Oxides

8.1. V2O2

V2O2 is generally used as a tool to investigate the nature of metal–oxygen bonds, which is crucial to provide a proper rationalization of the relationship between structure and properties at an atomic scale.367 Stable V2O2 could not be synthesized by a traditional solid-state reaction or solution methods. However, it can be observed by IR spectroscopy on V, Ne, and O2 codeposited matrices.368 The calculation results show two possible vanadium bonding situations: 1) no bonds between the two vanadium atoms; 2) a short distance between the two vanadium atoms, which indicated multiple bonding (Figure 44a).368370 However, Himmel et al.368 found that there is a multiple vanadium–vanadium bond in V2O2 molecules. There are three V–V bonding and antibonding orbitals, which are occupied by 1.70, 1.58, 1.50 electrons and 0.42, 0.40, 0.49 electrons, respectively. Thus, 4.78 electrons are in V–V bonding orbitals, and 1.31 electrons are in V–V antibonding orbitals. Furthermore, five bond critical points (four points between the oxygen and the vanadium atoms and one point in the center between the vanadium atoms) and two ring critical points (between the center of the cluster and the oxygen atoms) can be identified in V2O2 molecules (Figure 44b), which confirms the presence of a V–V bond. The presence of the strong V–V bond may lead to unique optical and magnetic properties.

Figure 44.

Figure 44

(a) Two possible Lewis representations highlighting the unclear bonding situation in V2O2. (b) Plot of the electron density within the molecular plane. Rectangles: bond critical points; triangles: ring critical points. Reproduced with permission from ref (368). Copyright 2017 John Wiley and Sons.

8.2. V4O9

V4O9 has an orthorhombic structure with a space group of Cmcm, which processes three types of VO polyhedra (Figure 45a). The VO5 pyramids and VO6 octahedra make pairs, which are connected by corner oxygen atoms from the VO4 tetrahedra.371 The valence of the vanadium ion in the octahedron and pyramid is +4, while that in the tetrahedron is +5. V4O9 is difficult to synthesize by a solid-state reaction from the mixture of binary V2O5 and V2O3 (or VO2), but it can be synthesized by the reduction of V2O5 using reducing agents of carbon, SO2, and sulfur.371,372 The amount of reducing agents dramatically affects the final products, which may consist of other vanadium oxides, such as V6O13 and VO2. Consequently, reducing V2O5 by the solvothermal method is a facile way to obtain V4O9. Different solvents (tetraethylene glycol, 2-propanol, and tetrahydrofuran) are used to synthesize V4O9 with different morphologies, such as nanoflakes, nanosheets, and so on.373375 Liang et al.375 investigated the aqueous zinc ion batteries of V4O9. It is found that the V4O9 exhibits fast zinc ion and electron diffusion, which is due to the unique tunnel structure and the V5+/V4+ mixed-valences induced metallic behavior. The V4O9 cathode shows a high reversible discharge capacity (420 mA h g–1 at 0.5 C). Even at a high current density of 50 C, it also exhibits an impressive discharge capacity of 234.4 mA h g–1, suggesting a fast Zn2+ storage ability of V4O9 (Figure 45b). Furthermore, V4O9 delivers an energy density of 175.8 W h kg–1 at a high power of 17625 W kg–1, which gives a high power density compared with other vanadium-based cathode materials in aqueous zinc ion batteries, such as V2O5,376 V6O13,377 LiV3O8,378 Na3V2(PO4)3,379 VO2,380 and VS2.381

Figure 45.

Figure 45

(a) Crystal structure of V4O9 with VO6 octahedron (yellow), VO5 pyramid (green), and VO4 tetrahedron (blue). (b) Rate performance at various currents ranging from 0.5 to 50 C of the aqueous Zn//V4O9 second battery. Reproduced with permission from ref (375). Copyright 2021 Royal Society of Chemistry. (c) The cycling stability of the specific capacitance of flower V4O9 at 2 A g–1. The inset shows the charge–discharge curves at different cycle numbers. (d) UV–vis absorbance curve (with polynomial fitting) of 75 mg flower V4O9 (blue) at varying H2O2 concentrations. Reproduced with permission from ref (373). Copyright 2013 Royal Society of Chemistry.

The 2D single layered V4O9 nanosheet assembled 3D microflowers exhibit good supercapacitor performance with a specific capacitance of 392 F g–1 at a current density of 0.5 A g–1 and 75% retained capacitance after 2000 cycles (Figure 45c).373 The flower-like structure assembled from ultrathin and well-separated nanosheets was unchanged during the charge–discharge cycles, which is responsible for the high capacitance and stability. Meanwhile, the V4O9 flower demonstrates a good ability to sense H2O2 and methanol with a detection limit of ∼0.1 μM and ∼60 μM, respectively (Figure 45d).

8.3. V6O13

The mixed-valence V6O13 attracts extensive attention because it can be used in a variety of ion batteries, such as Li+, Na+, Mg2+, Zn2+, and so on. As shown in Figure 46a, V6O13 is composed of alternating single and double vanadium oxide layers. There are two types of VO6 octahedra: V4+ occupied the yellow octahedra, and V5+ occupied the blue octahedra. All the octahedra are connected with corner O atoms, which form a tunnel-like structure.49,382 The solvothermal reaction is widely used to obtain V6O13 nanostructures, particles, and related composites, such as V6O13 nanogrooves,383 V6O13@hollow carbon microspheres,384 nest-like V6O13,385 V6O13 nanosheets,386 V6O13 nanowires,387 V6O13 nanorods,388 and so on (Figure 46b). Cao et al.389 recently developed a new strategy to synthesize V6O13 nanosheets by microwaves, which could reduce the thickness of V6O13 nanosheets greatly compared to that prepared by a hydrothermal method (Figure 46c,d).

Figure 46.

Figure 46

(a) Crystal structure of V6O13 with V two types of VO6 octahedra. (b) Different morphologies of V6O13 synthesized by a solvothermal method. Reproduced with permission from ref (385) (Copyright 2020 Royal Society of Chemistry), ref (384) (Copyright 2021 Elsevier), ref (383) (Copyright 2015 American Chemical Society), and ref (390) (Copyright 2019 Elsevier). (c, d) V6O13 nanosheet obtained by hydrothermal and microwave-assisted synthesis. Reproduced with permission from ref (386) (Copyright 2021 Elsevier) and ref (389) (Copyright 2022 Elsevier).

Due to the tunnel-like structure and mixed-valence, V6O13 exhibits a metallic character at room temperature, which is beneficial for high-rate charge and discharge. When it is used as an LIB electrode material, 8 mol Li+ intercalated per formula unit endowing a high theoretical specific capacity of 417 mAh g–1 and energy density of 900 Wh kg–1. The Yu group383 synthesized a 3D V6O13 nanotextile with interconnected 1D nanogrooves via a facile solution-redox-based self-assembly route at room temperature (Figure 47a). They confirmed that the precursor concentration affected the mesh size in the textile structure. The 3D V6O13 delivers a high capacity of 326 mAh g–1 at 20 mA g–1 and maintains 80% capacity after 100 cycles at 500 mA g–1. The energy density can reach 780 Wh kg–1, which is much higher than those of commercialization cathodes (LiFePO4 and LiCoO2) (Figure 47b). The excellent electrochemical performance is due to the unique structure of 3D textiles, which can be maintained upon cycling and are beneficial for ion transport and cycle stability (Figure 47c). In addition, Mai et al.391 synthesized a novel ultrathin prelithiated V6O13 nanosheet by a secondary hydrothermal prelithiation process (Figure 47d). A single-nanosheet device was employed to in situ probe the intrinsic advantages of prelithiated nanosheets. Compared with nonlithiated V6O13 nanosheets, the ultrathin prelithiated V6O13 nanosheets exhibit a higher electrical conductivity and maintain the same conductance level after the Li+ intercalation (Figure 47e). Meanwhile, the specific capacity of the ultrathin prelithiated V6O13 nanosheets can be maintained at 98% after 150 cycles at a current density of 1000 mA g–1, which is much higher than 46% capacity of nonlithiated V6O13 nanosheets (Figure 47f). These results demonstrate that prelithiation is a strategy to obtain high-energy and long-cycling energy storage cathode materials.

Figure 47.

Figure 47

XRD pattern and SEM image (a), different CV curves (b), and the schematic diagram of the Li+ intercalation process (c) of 3D V6O13 nanotextile electrodes in LIB. Reproduced with permission from ref (383). Copyright 2015 American Chemical Society. The TEM image of ultrathin lithiated V6O13 nanosheets (d), the transport properties (e), and cycling performance at 1000 mA g–1 (f) of nonlithiated ultrathin and ultrathin lithiated V6O13 nanosheets. Reproduced with permission from ref (391). Copyright 2014 Elsevier. Optimized geometry of Zn intercalated V6O13 with water (g) and without water (h), the galvanostatic voltage-capacity profiles for V6O13 cycled in electrolytes with different water contents in 1 M Zn(CF3SO3)2 acetonitrile (i). Reproduced with permission from ref (392). Copyright 2019 John Wiley & Sons. (j) Charge/discharge curves of K–V6O13 at different rates. (k) The migration energy barriers of K–V6O13 and V6O13. Reproduced with permission from ref (394). Copyright 2022 Royal Society of Chemistry. (l) Rate capability under various currents of V6O13 on carbon cloth. Reproduced with permission from ref (385). Copyright 2020 Royal Society of Chemistry.

Doping of various metal ions can improve the electrochemical performance and achieve a good capacity of the battery. The V6O13 with Al/Ga, Al/Fe, and Al/Na doping delivered an initial discharge specific capacity of 411.5 mA h g–1, 426.9 mA h g–1, and 514 mA h g–1 at 0.1 C, respectively.388 However, the electrochemical performances were poor at a high current density. V6O13 was also employed in multivalent ion (Zn2+, Mg2+) batteries.384,385,389,392,393 Choi et al.392 applied V6O13 as the ZIB cathode material and investigated its electrochemical behavior for Zn2+. In particular, they analyzed the effect of water in the electrolyte on the Zn2+ storage to investigate the physicochemical characteristics of V6O13. DFT calculation results demonstrate that the coordination environments of Zn show a big difference with/without water (Figure 47g,h). It will form octahedral coordination with water, but undercoordination without water. The Zn2+ storage of V6O13 increases with increasing water content in the electrolyte, and it can deliver a high capacity of 360 mAh g–1 and be maintained at 92% after 2000 cycles in an aqueous Zn(CF3SO3)2 (∼1 M) electrolyte (Figure 47i). Even at a high current density of 24 A g–1, it maintains a relatively high capacity of 145 mAh g–1.16 This work highlights that cointercalating water molecules play a vital role in enhancing the electrochemical performance of the aqueous ion storage system. Zhao et al.394 found that the K+ intercalated V6O13 exhibited a specific capacity of 367 mAh g–1 at 0.5 A g–1 and 198.8 mAh g–1 at 10 A g–1 (Figure 47j). Meanwhile, the capacity could be maintained at 90% after 2000 cycles at 10 A g–1. The DFT calculation suggested that K+ intercalation could significantly contribute to the reduction of the Zn2+ diffusion energy barrier (from 0.94 to 0.33 eV), which enables Zn ions to migrate away from the intercalation sites more easily (Figure 47k). Thus, K+ intercalated V6O13 showed good battery performance. Furthermore, by growing the V6O13 on carbon cloth and using the ZnSO4 as the electrolyte, the V6O13 cathode showed a capacity of 520 mAh g–1 (at a current density of 0.5 A g–1) and good cycle life (a stable capacity of 335 mAh g–1 over 1000 cycles) (Figure 47l).385

8.4. Other Vanadium Oxides for Energy-Related Application

V5O12·6H2O is another layered monoclinic vanadium oxide separated by water pillars with a large interlayer spacing of 1.179 nm. Wang et al.395 added a small amount of platinum (Pt, 1.5 wt %) into the interlayer of V5O12·6H2O. The obtained V5O12·6H2O–Pt electrode delivers a specific capacity of 440 mAh g–1 at 500 mA g–1, and increases to 489 mAh g–1 for the third cycle, much higher than V5O12·6H2O electrodes (270 mAh g–1 at 500 mA g–1) (Figure 48a,b). They also demonstrated that the Pt additive makes no contribution, and is even counterproductive to conductivity, but facilitates a significant enhancement of pseudocapacitance. Therefore, it is clear that there is a strong relationship between Pt and the new phase Zn4SO4(OH)6·5H2O. The in situ XRD patterns show that obvious characteristic peaks shifted to lower angles during the discharge process and returned to the pristine state during charging, indicating the interlayer spacing gradual enlargement and recovery upon Zn2+ intercalation/deintercalation. Meanwhile, the formation/disappearance of the zinc hydroxyl complex is accompanied by Zn2+ insertion/extraction (Figure 48c).

Figure 48.

Figure 48

XRD pattern (a), charge/discharge profiles (b), and in situ XRD patterns and 2D contour map of peak intensities of V5O12·6H2O–Pt (c). Reproduced with permission from ref (395). Copyright 2021 John Wiley & Sons.

Mai et al.396 artificially constructed a VOx cluster/reduced graphene oxide (rGO) cathode material with interfacially inserted Zn2+ repelling the pristine-bonded C atom into the plane of the rGO and constructing interfacial V–O–C bonds (Figure 49a). Meanwhile, the VOx consists of subnanoclusters (less than 1 nm per dimension) and partial nanoclusters (slightly larger than 1 nm). In combination with the electrons transferred to the rGO during the discharge process (Figure 49b–d), the reduced degree of defect is additional proof of the interfacial Zn2+ storage. As a result, they have discovered a new mechanism in which Zn2+ ions are stored mainly at the interface between VOx and rGO, which leads to anomalous valence changes compared to conventional mechanisms and exploits the storage capacity of the nonenergy storing active but highly conductive rGO. The obtained VOx-G heterostructure delivers a superior rate performance with a capacity of 174.4 mAh g–1 at an ultrahigh current density of 100 A g–1, with capacity retention of 39.4% for a 1000-fold increase in current density (Figure 49e). To the best of our knowledge, the rate performance is one of the best among ZIBs.

Figure 49.

Figure 49

Characterization of the interfacial configuration in the VOx-G heterostructure and its anomalous Zn2+ storage mechanism (a), DFT simulations in pristine (b) and fully discharged (c) states, formation energy evolution during Zn2+ ion insertion into the interface (d), galvanostatic discharge curves varying from 0.1 to 100 A g–1 (e) of VOx-G heterostructure. Reproduced with permission from ref (396). Copyright 2021 John Wiley & Sons.

Besides the battery application, the nanostructured VOx materials/composites have been considered as a good catalyst for the HER due to the creation of the more active site and modification of the adsorption and desorption of H atoms.397 Guan et al.398 prepared 2D Zn-VOx-Co ultrathin nanosheets on carbon fiber paper by an electrodeposition method. The Zn-VOx-Co electrocatalysts contain the amorphous Co metal phase and crystalline Zn–Co alloy phase, which gives the materials a good HER performance (an overpotential of 46 mV at 10 mA cm–2 and a Tafel slope of 75 mV dec–1). Furthermore, Zhao et al.397 adopted the same method to fabricate a Ni(Cu)VOx catalyst by changing Zn and Co to Ni and Cu (Figure 50a). The Ni(Cu)VOx electrode displays a small overpotential of 21 mV at a current density of 10 mA cm–2 and a Tafel slope of 28 mV dec–1 (Figure 50b), which is comparable to the commercial 20% Pt/C catalyst (15 mV @ j = 10 mA cm–2 and 25 mV dec–1). Meanwhile, the Ni(Cu)VOx electrode also remains relatively stable for more than 100 h HER at 100 mA cm–2 (Figure 50c). The good HER performance is due to the existence of Ni–O–VOx sites, which promote the formation of highly disordered metallic Ni structures and further induce electron transfer from Ni to VOx.

Figure 50.

Figure 50

(a) Schematic to prepare the Ni(Cu)VOx electrode for HER electrolysis. (b) LSV curves for Ni(Cu)VOx and control samples without iR compensation. The black dotted curve is the HER activity on the Ni(Cu)VOx electrode with iR correction. (c) Chronopotentiometric curve of Ni(Cu)VOx in comparison with Ni(Cu). Reproduced with permission from ref (397). Copyright 2020 under CC BY license.

The VOx based composites also exhibit good OER performance. Dong et al.399 synthesized 3D nanoflower-like VOx nanosheets (VOx/NiS/NF) by a hydrothermal method, which showed the OER performance with a low overpotential (330 mV at 50 mA cm–2) and a small Tafel slope (121 mV dec–1). Yang et al.400 synthesized a ternary Co-VOx-P catalyst with a nanoflower structure directly onto the Ni foam. The as-prepared materials demonstrated excellent catalytic performance for both the HER (an overpotential of 98 mV at 10 mA cm–2 and a Tafel slope of 59 mV dec–1) and OER (an overpotential of 230 mV at 100 mA cm–2 and a Tafel slope of 64 mV dec–1) under an alkaline environment.

9. Conclusions and Future Outlook

In the past two decades, there has been accelerated development of vanadium oxides due to the fact that they are the most promising candidates in versatile applications, such as batteries, energy saving smart windows, sensing, catalysts, optoelectronic devices, etc. In this review, we have discussed the V–O binary phase diagram, the structure and synthesis methods of five thermodynamically stable vanadium oxides (V2O3, V3O5, VO2, V3O7, V2O5) and some metastable vanadium oxides (V2O2, V4O9, V6O13) with selected applications on hydrogen evolution catalysis, supercapacitors, batteries, smart windows, and some other aspects. The battery, supercapacitor, and HER/OER performances of these vanadium oxides are summarized for comparison to provide an overview of the research in the field (Tables 1, 2, and 3).

Table 1. Electrochemical Properties of the Typical Vanadium Oxides for Metal-Ion Batteries.

materials battery types Ccapacity (mAh g–1)/current density (A g–1) retention (%)/Ccycle numbers/current density (A g–1) rate capacity (mAh g–1)/current density (A g–1) ref
VOx–rGO heterostructure ZIBs 443/0.1 ∼92%/1000/20 174.4/100 (396)
V2O3 hollow spheres LIBs 785/0.1 Over 100%/700/2 361/2 (76)
V2O3 porous nanofibers KIBs 240/0.05 94.5%/500/0.05 134/1 (68)
V2O3 ZIBs 625/0.1 97.7%/1000/5 486/20 (77)
V2O3 with CNTs SIBs 612/0.1 100%/6000/2 207/10 (361)
70%/10000/10
Carbon-confined V2O3 ZIBs 633.1/0.2 90.3%/10000/12 271.4/24 (401)
V3O5 LIBs 412/0.2 78.5%/2000/50 125/50 (98)
Graphene quantum dots coated onto the VO2 surfaces LIBs 421/0.1 94%/1500/18 151/36 (237)
Graphene quantum dots coated onto the VO2 surfaces NIBs 306/0.1 88%/1500/18 93/36 (237)
VO2 (B) nanorods KIBs 209.2/0.05 86%/500/0.5 141.4/2 (238)
VO2 (B) nanorods MIBs 391/0.025 41.9%/60/0.85 341/0.1 (402)
VO2 hollow microspheres LIBs 203/0.1 80%/1000/2 134/2 (240)
VO2 interwoven nanowires coated with Carbon quantum dots LIBs 427/0.1 112%/500/19.2 168/19.2 (403)
VO2 with graphene ribbons LIBs 415/0.4 93%/1000/37.2 204/37.2 (404)
VO2 (B) nanobelt forest LIBs 475/0.1 63%/47/0.1 100/27 (405)
VO2 nanorods ZIBs 325.6/0.05 86%/5000/3 72/5 (239)
VO2 nanofibers ZIBs 357/0.1 N.A. 171/51.2 (406)
V3O7 nanowire templated graphene scrolls LIBs 321/0.1 87.3%/400/2 162/3 (120)
V3O7·H2O ZIBs 370/0.375 80%/200/3 270/3 (121)
H2V3O8 MIBs 231/0.01 77%/100/0.04 97/0.08 (108)
H2V3O8 nanowires ZIBs 423.8/0.1 94.3%/1000/5 113.9/5 (407)
H2V3O8 nanowires with graphene ZIBs 394/0.1 87%/2000/6 270/6 (408)
V4O7 LIBs 291/0.05 ∼100%/100/3 159/3 (409)
V4O9 ZIBs 420/0.235 78.8%/1000/9.4 233.4/23.5 (375)
3D V6O13 nanotextiles LIBs 326/0.02 80%/100/0.5 134/0.5 (383)
V6O13 nanosheet LIBs 331/0.1 98%/150/1 150/2 (391)
V6O13 ZIBs 360/0.2 92%/2000/4 145/24 (392)
V6O13 microflowers SIBs 159.8/0.16 73.5%/30/0.16 N.A. (410)
Oxygen-deficient V6O13 ZIBs 401/0.2 95%/200/0.2 223/5 (411)
V5O12·6H2O ZIBs 440/0.5 ∼95%/400/10 158/15 (395)
V5O12·6H2O nanobelt ZIBs 354.8/0.5 ∼94%/1000/2 228/5 (412)
V10O24·12H2O ZIBs 164.5/0.2 90.1%/3000/10 80/10 (413)
V10O24·12H2O ZIBs 365.3/0.2 83.2%/3000/5 127.2/80 (414)
V7O16 nanotube ZIBs 314.6/0.1 80.5%/950/2.4 87.8/9.6 (415)
V2O5 hollow microclew LIBs 145.3/0.1 94.4%/50/0.1 94.8/10 (308)
V2O5 hollow nanosphere SIBs 159.3/0.04 72.6%/100/0.16 112.4/0.64 (416)
V2O5 nanowires AIBs 305/0.125 89.5%/20/0.125 N.A. (417)
V2O5 ZIBs 470/0.2 91.1%/4000/5 386/10 (418)
V2O5 nanosheet ZIBs 224/0.1 81.3%/30/0.1 N.A. (419)
V2O5 nanofibers ZIBs 319/0.02 81%/500/0.6 104/3 (376)
Porous V2O5 LIBs 142/0.075 ∼90%/100/0.075 86.7/8.2 (420)
V2O5/C SIBs 255/0.015 95%/30/0.015 170/0.294 (309)
V2O5-polyaniline superlattice MIBs 270/0.1 61.5%/500/4 130/4 (421)
Polyaniline intercalated V2O5 NIBs 307/0.5 42%/100/5 69/20 (422)
V2O5·nH2O SIBs 338/0.05 73%/50/0.5 96/1 (423)
V2O5·nH2O ZIBs 372/0.3 71%/900/6 248/30 (310)
V2O5·nH2O CIBs 204/0.14 86%/350/0.7 28/2.8 (424)
3D V2O5/RGO/CNT LIBs 304/0.0294 90%/80/0.294 100/5.88 (311)
Na0.33V2O5 ZIBs 367.1/0.1 93%/1000/1 96.4/2 (312)
Mn0.01V1.99O5 LIBs 251/0.294 80%/50/0.294 171/1.47 (313)
UGF-V2O5/PEDOT LIBs 297/0.294 98%/1000/17.64 85/23.52 (314)

Table 2. Electrochemical Properties of the Typical Vanadium Oxides for Supercapacitors.

materials electrolyte Specific capacitance (F g–1)/Current density (A g–1) Retention (%)/Cycle numbers ref
VOOH nanosheets 1 M LiClO4/PPC 323/0.2 70%/2000 (425)
V2O3 nanoflakes@C core–shell composites 1 M NaNO3 205/0.05 76%/500 (94)
V2O3/C nanocomposites 5 M LiCl 458.6/0.5 86%/1000 (93)
V2O3@C core–shell nanorods 5 M LiCl 228/0.5 81%/1000 (426)
V2O3@C core–shell nanorods 1 M Na2SO4 192/1 66%/1000 (427)
V2O3@C core–shell composites 1 M Na2SO4 223/0.1 39.7%/100 (428)
N-doped carbon coated nest-like V3O7 1 M Na2SO4 660.63/0.5 80.47%/4000 (112)
    187.72/50    
V3O7 nanowires on carbon fiber cloth 1 M Na2SO4 198/1 97%/100000 (125)
V3O7·H2O nanobelts/CNT/rGO composites 5 M LiCl/PVA 685/0.5 99.7%/10000 (126)
V3O7-rGO-polyaniline composites 1 M H2SO4 579/0.2 95%/2500 (429)
VO2(B) nanobelts/rGO composites 0.5 M K2SO4 353/1 78%/10000 (235)
VO2 nanosheet 1 M LiClO4/PPC 405/1 82%/6000 (430)
VO2 nanosheet 6 M KOH 663/10 99.4%/9000 (431)
VO2(B)/C core–shell composites 1 M Na2SO4 203/0.2 10.4%/100 (432)
Graphene foam/VO2 nanoflakes/hydrogen molybdenum bronze composites 1 M K2SO4 485/2 97.5%/5000 (433)
    306/32    
VO2 microarrays 1 M Na2SO4 265/1 100%/3000 (434)
VO2@polyaniline coaxial nanobelts 0.5 M Na2SO4 246/0.5 28.6%/1000 (435)
V4O9 nanosheets 1.5 M KOH 392/0.5 75%/2000 (373)
V6O13@C 1 M Na2SO4 545/0.5 88.3%/2000 (436)
V6O13 sheets 1 M NaNO3 285/0.05 96.7%/300 (437)
V6O13 1 M LiNO3 456/0.6 65%/2000 (438)
Sulfur-doped V6O13–x@C 5 M LiCl 1353/1.9 92.3%/10000 (439)
RuO2 nanoparticle decorated V2O5 nanoflakes 1 M KCl 421/1 94.6%/10000 (315)
Graphene nanoribbons @ V2O5 nanostrips 0.5 M Na2SO4 335.8/1 ∼98.5%/10000 (316)
N-doped carbon nanofibers/V2O5 core/shell 1 M Na2SO4 595.1/0.5 97%/12000 (317)
Interconnected V2O5 Nanoporous Network 0.5 M K2SO4 304/0.1 24%/600 (440)
V2O5/rGO nanocomposites 8 M LiCl 537/1 84%/1000 (441)
V2O5/rGO hybrids 1 M Na2SO4 468.5/1 91.5%/10000 (442)
V2O5/graphene hybrid aerogels 1 M Na2SO4 486/0.5 90%/20000 (273)
V2O5/rGO composite hydrogel 0.5 M Na2SO4 320/1 70%/1000 (443)
Carbon coated V2O5 nanorods 0.5 M K2SO4 417/0.5 76%/1000 (444)
    341/10    
Hollow spherical V2O5 5 M LiNO3 559/3 70%/100 (445)
V2O5 nanorods 1 M LiClO4 347/1 94.3%/10000 (446)
V2O5 nanorods/rGO 0.75 M NaPF6 289/0.01 85%/1000 (447)
V2O5 nanosheets/rGO 1 M KCl 635/1 94%/3000 (448)
V2O5 nanobelts 1 M LiClO4/ PPC 132.5/1 95%/500 (449)
V2O5 nanobelts/rGO 0.5 M K2SO4 310.1/1 90.2%/5000 (319)
V2O5 microtubules 1 M LiNO3 680/1 70%/10000 (271)
V2O5 nanoparticles 1 M LiClO4/ PPC 545/1 70%/500 (450)
V2O5/graphene hybrid 1 M Na2SO4 484/0.6 80%/10000 (451)
V2O5/graphene hybrid aerogel composite 1 M LiClO4/ PPC 384/0.1 82.2%/10000 (452)
    197/2    
V2O5/MWCNT core/shell hybrid aerogels 1 M Na2SO4 625/0.5 120%/20000 (453)
Carbon coated flowery V2O5 1 M K2SO4 417/0.5 100%/2000 (454)
V2O5/mesoporous carbon microspheres 1 M Al2(SO4)3 290/0.5 88%/10000 (455)
Graphene-wrapped V2O5 nanospheres 1 M Na2SO4 612.5/1 89.6%/10000 (456)
V2O5/polypyrrole 5 M LiNO3 448/0.8 81%/1000 (457)
V2O5/polyaniline 0.5 M LiClO4/ PPC 1115/1 90%/4000 (458)
V2O5@Ni3S2 1 M KOH 854/1 60%/1000 (330)
V2O5/Na0.33V2O5 1 M LiClO4 334/1 96%/1000 (459)
V2O5/TiO2 1 M LiNO3 587/0.5 92%/5000 (460)
V2O5 nanowire arrays/N-doped graphene aerogel 8 M LiCl 710/0.5 95%/20000 (461)
V2O5 nanofibers/conductive polymer 1 M Na2SO4 614/0.5 111%/15000 (462)
V2O5/nanoporous carbon network 0.5 M K2SO4 314.6/0.2 89.5%/5000 (463)
V2O5/Ni foam 1 M KOH 399.7/0.01 96.1%/2000 (464)
V2O5/WO3 1 M H2SO4 386/0.1 104%/5000 (465)
V2O5/g-C3N4 1 M KOH 192.3/0.5 85.7%/5000 (466)
V2O5@Ti 1 M LiCl 1520/1.5 99%/12000 (467)
V2O5/vertically aligned CNT 1 M Na2SO4 284/2 76%/5000 (468)
V2O5 nanosheets/carbon fiber felt 5 M LiCl 475.5/1 89.7%/6000 (469)

Table 3. A Brief Survey of Typical Vanadium Oxides Electrocatalysts.

catalyst catalyst type electrolyte η @ j = 10 mA cm–2 (mV) Tafel slope (mV dec–1) stability (cycle number or time @ current density) ref
V2O3/ MoSx /CC HER 0.5 M H2SO4 146 45 1000 cycles (82)
NiFe@V2O3 HER 1 M KOH 84 85 24 h @ 10 mA cm–2 (80)
NiFe@V2O3 OER 1 M KOH 255 51 1000 cycles (80)
24 h @ 10 mA cm–2
Ni0.8/V2O3 HER 1 M KOH 44 38 24 h @ 10 mA cm–2 (470)
V2O3–Ni3N HER 1 M KOH 57 50 24 h @ 10 mA cm–2 (471)
V2O3@Ni HER 1 M KOH 47 74 1000 cycles (472)
10 h @ 10 mA cm–2
V2O3–CoFe2O4 HER 1 M KOH 61 58 80 h @ 500 mA cm–2 (473)
V2O3–CoFe2O4 OER 1 M KOH 226 56 80 h @ 500 mA cm–2 (473)
Ni4Mo–V2O3 HER 1 M PBS 40 66 1000 cycles (474)
          5.5 h @ 50 mA cm–2  
MoS2/VO2 HER 0.5 M H2SO4 99 @ 1 mA cm–2 85 N. A. (245)
Ni3S2/VO2 HER 1 M KOH 100 114 15 h @ 10 mA cm–2 (244)
Ni3S2/VO2 OER 1 M KOH 150 47 15 h @ 10 mA cm–2 (244)
Co3O4/VO2/CC HER 1 M KOH 108 98 10 h @ 10 mA cm–2 (246)
porous VO2 nanosheets HER 0.5 M H2SO4 184 70 120 h @ 90 mA cm–2 (247)
porous VO2 nanosheets OER 1 M KOH 209 92 120 h @ 70 mA cm–2 (247)
CoV2O6 - V2O5/NRGO OER 1 M KOH 239 50 1000 cycles (320)
Co–V2O5 HER 1 M KOH 51 42 24 h @ 10 mA cm–2 (328)
V2O5/Ni(OH)2@NF HER 1 M KOH 39 44 10000 cycles (329)
Pt (P)–V2O5/graphene HER 0.5 M H2SO4 32 23 1000 cycles (331)
V2O5@Ni3S2 HER 1 M KOH 95 108 9000 cycles (330)
Ni–Co–P/V2O5–TiO2/GO HER 1 M KOH 101 @ 100 mA cm–2 36 N. A. (475)
Zn-VOx-Co HER 1 M KOH 46 75 36 h @ 100 mA cm–2 (398)
Ni(Cu)VOx HER 1 M KOH 21 28 125 h @ 100 mA cm–2 (397)
VOx/NiS/NF OER 1 M KOH 330 @ 50 mA cm–2 121 1000 cycles (399)
Co-VOx-P HER 1 M KOH 98 59 1000 cycles (400)
24 h @ 30 mA cm–2
Co-VOx-P OER 1 M KOH 230 @ 100 mA cm–2 64 1000 cycles (400)
          25 h @ 100 mA cm–2  
VOx@NiFe/NiCoP/TM HER 1 M KOH 45 34 1000 cycles (476)
          16 h @ 100 mA cm–2  
VOx@NiFe/NiCoP/TM OER 1 M KOH 215 37 20 h @ 500 mA cm–2 (476)
Co(VOx) HER 1 M KOH 178 @ 100 mA cm–2 40 60 h @ 100 mA cm–2 (477)

While vanadium oxides are important materials for many applications, more detailed, mechanistic and systematical studies are needed to fully explore their potential as the bottleneck is increasingly related to the materials’ quality and device fabrications. We propose a few future works in this field which could be further developed in the following aspects (Figure 51):

  • (1)

    Searching the applications of vanadium oxides in underexplored areas. One example is biological thermal imaging by leveraging its thermal phase transition characteristics as the insulator-to-metal phase transition of some vanadium oxides gives sharply enhanced optical absorption above their critical temperature. It has potential applications in biochemistry, especially in the fields of bioimaging. Another example is that some nanostructured vanadium oxides also possess optimal physicochemical properties (e.g., optical, thermal, magnetic properties), which give the opportunity to be applied in intelligent medicine based on micro-/nanorobots.

  • (2)

    Development of high-level theoretical calculations to guide the rational design of vanadium oxides and their composites and to better understand the fundamentals of high-performance devices. The use of advanced computational tools and theories enables researchers to understand the origin of complex and interacting phenomena at multiple scales, which could accelerate our understanding of the fundamentals of high-performance devices and improve the operation and design of new materials systems. Additionally, machine learning is a useful toolkit for designing and exploring new vanadium oxides with desired properties. It can also aid the understanding of the complex correlations between structures and properties in vanadium oxides.

  • (3)

    Fabrication of ultrathin or two-dimensional vanadium oxides, which could be applied in some new electronic devices. Owing to the atomic-scale thickness of single layers, the 2D materials exhibit tunable electrical properties and bandgaps. Therefore, the 2D vanadium oxides may hold promise for a wide range of applications in low-power electronics, flexible electronics, optoelectronics, catalysis, batteries, and so on. Furthermore, the 2D vanadium oxides may also form novel 2D heterostructures with other ultrathin 2D nanomaterials, which would be of certain interest.

  • (4)

    Exploration of new approaches to stabilize the vanadium oxides related devices, especially thermal, light, moisture, and oxygen environmental stability. Long-term device stability is one of the most important challenges for all devices. Most vanadium oxides are sensitive to oxygen and moisture, which may lead to the degradation of the device performances. Furthermore, most stability measurements of the devices are performed under ambient conditions, which limits their application in high temperature, high humidity, and high light intensity conditions. Meanwhile, it is also important to understand the degradation mechanisms in the different types of devices based on vanadium oxides, which could be the key to improving the device stabilities.

  • (5)

    Exploring new green chemistry synthesis methods to eliminate the toxicity of vanadium oxides. Vanadium oxides cause a variety of toxic effects such as biochemical changes, neurobehavioral injury, and functional lesions in the liver and bones. Especially, vanadium oxides in breathing air can cause pulmonary problems and DNA damage in leukocytes. The toxicity is more related to the phase structure, stoichiometric ratio, concentration, particle size, and crystalline degree, which could be considered in all the processes of application of vanadium oxides. Therefore, exploring new green synthesis methods and avoiding risks to humans and the environment during vanadium oxides’ production, use, and disposal processes deserve more systematic and comprehensive studies. First, it is important to develop new synthesis protocols with minimum steps to prepare vanadium oxides, such as one-pot synthesis, completely enclosed-system synthesis, and so on. Second, the encapsulation of the devices to confine the toxicity needs to be considered when designing a new device. Third, the release of vanadium oxides to the atmosphere should be controlled during normal operation. For example, the marked decrease in toxicity is confirmed via silica coating on vanadium oxides due to the perception that the toxicity of vanadium oxide is closely related to the solubility and the robust silica barrier can isolate air and water to reduce the solubility.

  • (6)

    Realization of the bifunctional, trifunctional, or even multifunctional vanadium oxides to achieve integrated functionality. Different device integrations based on the same materials is expected to greatly reduce the cost and the incompatibility of different materials. Meanwhile, multifunctional materials would reduce the complexity of designing devices and promote the application in designated situations.

Figure 51.

Figure 51

Future development and research directions for vanadium oxides.

Overall, we believe that vanadium oxides are great candidates for future applications in related fields and help solve key challenges in the global warming crisis. Moreover, the integration of vanadium oxides with multidisciplinary fields such as material science, device physics, civil engineering, mechanical design, and bioscience would continue to attract the interest of many scientists from different disciplines for new fascinating fields.

Acknowledgments

Peng Hu acknowledges the financial support from the National Natural Science Foundation of China (No. 51803168), the Key Research and Development Program of Shaanxi Province (2022GY-356), and the Youth Innovation Team of Shaanxi Universities. L.M. acknowledges the financial support from the National Key Research and Development Program of China (Grant No. 2020YFA0715000), the National Natural Science Foundation of China (Grant No. 52127816), and the Foshan Xianhu Laboratory of the Advanced Energy Science and Technology Guangdong Laboratory (XHT2020-003). Ping Hu acknowledges the financial support from the Guangdong Basic and Applied Basic Research Foundation (2021A1515110059). Y.L. would like to acknowledge the funding support from MOE-T2EP50221-0014, Ministry of Education, Singapore. Y.L. would like to acknowledge the funding support from MOE-T2EP50221-0014, Minster of Education, Singapore and Global STEM Professorship Scheme sponsored by the Government of the Hong Kong Special Administrative Region.

Glossary

Abbreviations

1D

one-dimensional

2D

two-dimensional

3D

three-dimensional

4D

four-dimensional

AACVD

aerosol-assisted chemical vapor deposition

AC

activated carbon

AFM

atomic force microscope

AIB

aluminum-ion battery

ALD

atomic layer deposition

AM 1.5

air mass 1.5

BP

black paint

CC

carbon cloth

Cdl

double-layer capacitance

CIB

calcium-ion battery

CNC

carbon nanocoils

CNT

carbon nanotube

CV

cyclic voltammetry

CVD

chemical vapor deposition

DC

direct current

DFT

density functional theory

DNA

DNA

EBD

electron-beam deposition

EN

ethylenediamine

GCD

galvanostatic charge/discharge

GF

graphene foam

GITT

galvanostatic intermittent titration technique

GQD

graphene quantum dot

GVG

graphene foam supported graphene quantum dot anchored VO2 arrays

FeFET

ferroelectric field-effect transistor

FET

field-effect transistor

GO

graphene oxide

HCP

high concentration dilution

HER

hydrogen evolution reaction

High-E

high emissivity

HM

hollow microclew

IL

ionic liquid

IR

infrared

ITO

indium tin oxide

KIB

potassium-ion battery

LCP

low concentration dilution

LED

light emitting diode

LIB

lithium-ion battery

LiTFSI

lithium bis-trifluoromethanesulfonimide

Low-E

low emissivity

LWIR

longwave infrared

LSPR

localized surface plasmonic resonance

MAS NMR

magic-angle-spinning nuclear magnetic resonance

MB

methylene blue

MCE

magnetocaloric effect

MD

molecular dynamics

MEMS

microelectromechanical systems

MIB

magnesium-ion battery

MIT

metal–insulator transition

MO

methyl orange

MWCNT

multiwalled carbon nanotube

N.A.

not applicable

NaTFSI

sodium bis-trifluoromethanesulfonimide

NDR

negative differential resistance

NF

Ni foam

NIB

ammonium-ion battery

NIR

near-infrared

NMR

nuclear magnetic resonance

NP

nanoparticle

NRGO

nitrogen-doped reduced graphene oxide

NW

nanowire

OER

oxygen evolution reaction

PBS

phosphate-buffered saline

PDMS

polydimethylsiloxane

PEDOT

poly(3,4-ethylenedioxythiophene)

PEG

polyethylene glycol

PLD

pulsed laser deposition

PNCNF

porous N-doped carbon nanofiber

PPC

propylene carbonate

PS

polystyrene

PVA

poly(vinyl alcohol)

PVC

polyvinyl chloride

PVD

physical vapor deposition

PVDF-HFP

polyvinylidene fluoride-hexafluoropropylene

PVP

polyvinylpyrrolidone

PZT

Pb(Zr0.52Ti0.48)O3

RC

radiative cooling

RCRT

radiative cooling regulating thermochromic

ReRAM

resistive random-access memory

rGO

reduced graphene oxide

RhB

rhodamine B

RF

radio frequency

RT

room temperature

SA-VO2

surface amorphized VO2

SC

supercapacitor

SCE

saturated calomel electrode

SCVA

single-crystalline VO2 actuator

SEM

scanning electron microscope

SHE

standard hydrogen electrode

SIB

sodium-ion battery

SWCNT

single- or multiwalled carbon nanotube

TARC

temperature-adaptive radiative coating

TEM

transmission electron microscope

TGA

thermogravimetric analysis

TM

titanium mesh

UGF

ultrathin graphite foam

UHV

ultrahigh vacuum

UV–vis–NIR

ultraviolet–visible-near-infrared

VOG

V2O5·H2O/graphene

VGS

V3O7 nanowire templated graphene scroll

V2O3@NC

nitrogen-doped carbon-confined V2O3

XANES

X-ray absorption near edge structure

XPS

X-ray photoelectron spectroscopy

XRD

X-ray diffraction

ZIB

zinc-ion battery

Biographies

Peng Hu received his PhD degree (under the guidance of Prof. Christian Kloc) in materials science at Nanyang Technological University, Singapore, in 2016. Then, he was a research fellow at Nanyang Technological University, Singapore, in Prof. Christian Kloc’s group (2016–2017) and Dr. Yi Long’s group (2017–2018). He is currently a professor in the School of Physics at Northwest University, China. His research interests focus on the single crystal growth of organics, low-dimensional materials and organic–inorganic hybrids, and their electronic, optoelectronic, and energy applications.

Ping Hu received his PhD degree from Wuhan University of Technology in 2020. He is currently a postdoc at Wuhan University of Technology. His research interests focus on advanced electrode materials for electrochemical energy storage devices.

Tuan Duc Vu is a PhD candidate from Nanyang Technological University (NTU) in Dr. Long Yi and Dr. Xianting Zeng’s group, School of Materials Science and Engineering. He previously obtained his B.Eng. (Materials) from the same school. In 2018, he received the prestigious A*STAR Graduate Scholarship AGS(S), which funded his PhD candidacy in NTU in collaboration with Singapore Institute of Manufacturing Technology (SIMTech). He specializes in thin film processing using a physical vapor deposition process such as thermal evaporation and magnetron sputtering. His current research focuses on the fabrication of high performance and high durability vanadium dioxide smart window glazing.

Ming Li is an Associate Professor in the Laboratory of Nanomaterials & Nanostructures at the Institute of Solid State Physics, Chinese Academy of Science (ISSP, CAS). He received his PhD in condensed matter physics from the ISSP, CAS. His research focuses on developing thermochromic materials of vanadium oxides for energy-efficient windows. His research topics cover the synthesis of nanomaterials, the regulation of photoelectric properties, and the design of thermochromic devices.

Shancheng Wang is a research fellow at Nanyang Technological University under the supervision of Dr. Yi Long. He obtained his PhD degree from NTU in 2021. His research interests include the synthesis of stimuli-response materials and design of stimuli-response smart structures for optical and thermal applications. In 2021, Shancheng received the MSE Doctorate Technopreneur Award from the School of Materials Science and Engineering, NTU.

Yujie Ke is a scientist at the Institute of Materials Research and Engineering, Agency for Science, Technology and Research (A*STAR) Singapore. He received his Ph.D. from NTU under the supervision of Dr. Long Yi in 2019 and his master’s degree from University at Buffalo (SUNY-Buffalo) in 2015. He is interested in phase-change materials, mechano-/thermochromics, plasmonics and photonic crystals, mechanical metamaterials, smart window, and building energy.

Xianting Zeng is Director of the Knowledge Transfer Office (KTO) of the Singapore Institute of Manufacturing Technology (SIMTech), of the Agency for Science, Technology and Research (A*STAR). He obtained his PhD degree in Thin Film Physics from the Chinese University of Hong Kong in 1995, and MSc and BSc degrees from the Huazhong University of Science and Technology, China in 1987 and 1982, respectively. He has been an active research scientist in the fields of nanocomposite coating materials and plasma enhanced PVD thin film processes. After joining SIMTech in 1995, he established the KTO in 2009 to transfer knowledge and capability to industry through case-studies and hands-on practical training to meet the technology and skills gap in productivity improvement, skills, and capabilities upgrading and business transformation.

Liqiang Mai is a Chair Professor of Materials Science and Engineering at the Wuhan University of Technology. He received his PhD degree from Wuhan University of Technology in 2004 and carried out his postdoctoral research at the Georgia Institute of Technology in 2006–2007. He worked as an advanced research scholar at Harvard University in 2008–2011 and the University of California, Berkeley in 2017. His current research interests are focused on new nanomaterials for electrochemical energy storage and micro-/nanoenergy devices.

Yi Long is a Fellow of the Royal Society of Chemistry. She obtained her PhD from the University of Cambridge, UK. She is currently an professor in the Department of Electronic Engineering, The Chinese University of Hong Kong. Her research focuses on nanostructured functional materials for different applications. She has successfully implemented technology transfer from lab to industry for a hard-disk company in her early career. Her recent research is to develop various smart materials by manipulation of the structure at the nanoscale to achieve unusual properties. She is the recipient of GreenAwards Top 3 London and Falling Walls 10 Breakthrough in Engineering and Technology Berlin in 2022.

Author Contributions

# Peng Hu and Ping Hu contributed equally to this work. CRediT: Tuan Duc Vu writing-original draft; Ming Li writing-original draft; Shancheng Wang writing-original draft; Yujie Ke writing-original draft; Xianting Zeng supervision; Liqiang Mai funding acquisition, supervision, writing-review & editing; Yi Long conceptualization, funding acquisition, supervision, writing-review & editing.

The authors declare no competing financial interest.

References

  1. Moskalyk R. R.; Alfantazi A. M. Processing of Vanadium: a Review. Miner. Eng. 2003, 16, 793–805. 10.1016/S0892-6875(03)00213-9. [DOI] [Google Scholar]
  2. Sutradhar M.; Da Silva J. A. L.; Pombeiro A. J. L.. Chapter 1 Introduction: Vanadium, Its Compounds and Applications. In Vanadium Catalysis; The Royal Society of Chemistry, 2021; pp 1–11. [Google Scholar]
  3. Langeslay R. R.; Kaphan D. M.; Marshall C. L.; Stair P. C.; Sattelberger A. P.; Delferro M. Catalytic Applications of Vanadium: A Mechanistic Perspective. Chem. Rev. 2019, 119, 2128–2191. 10.1021/acs.chemrev.8b00245. [DOI] [PubMed] [Google Scholar]
  4. Liu M.; Su B.; Tang Y.; Jiang X.; Yu A. Recent Advances in Nanostructured Vanadium Oxides and Composites for Energy Conversion. Adv. Energy Mater. 2017, 7, 1700885. 10.1002/aenm.201700885. [DOI] [Google Scholar]
  5. Wu C.; Xie Y. Promising Vanadium Oxide and Hydroxide Nanostructures: from Energy Storage to Energy Saving. Energy Environ. Sci. 2010, 3, 1191–1206. 10.1039/c0ee00026d. [DOI] [Google Scholar]
  6. Kianfar E. Recent Advances in Synthesis, Properties, and Applications of Vanadium Oxide Nanotube. Microchem. J. 2019, 145, 966–978. 10.1016/j.microc.2018.12.008. [DOI] [Google Scholar]
  7. Wang Y.; Cao G. Synthesis and Enhanced Intercalation Properties of Nanostructured Vanadium Oxides. Chem. Mater. 2006, 18, 2787–2804. 10.1021/cm052765h. [DOI] [Google Scholar]
  8. Wang Y.; Cao G. Developments in Nanostructured Cathode Materials for High-Performance Lithium-Ion Batteries. Adv. Mater. 2008, 20, 2251–2269. 10.1002/adma.200702242. [DOI] [Google Scholar]
  9. Rehder D. The Future of/for Vanadium. Dalton Trans. 2013, 42, 11749–11761. 10.1039/c3dt50457c. [DOI] [PubMed] [Google Scholar]
  10. Lamsal C.; Ravindra N. M.. Vanadium Oxides: Synthesis, Properties, and Applications. In Semiconductors: Synthesis, Properties and Applications; Pech-Canul M. I., Ravindra N. M., Eds.; Springer International Publishing, 2019; pp 127–218. [Google Scholar]
  11. Cotton F. A.; Wilkinson G.; Murillo C. A.; Bochmann M.. Advanced Inorganic Chemistry; John Wiley & Sons: New York, 1999. [Google Scholar]
  12. Prasadam V. P.; Bahlawane N.; Mattelaer F.; Rampelberg G.; Detavernier C.; Fang L.; Jiang Y.; Martens K.; Parkin I. P.; Papakonstantinou I. Atomic Layer Deposition of Vanadium Oxides: Process and Application Review. Mater. Today Chem. 2019, 12, 396–423. 10.1016/j.mtchem.2019.03.004. [DOI] [Google Scholar]
  13. Wei J.; Ji H.; Guo W.; Nevidomskyy A. H.; Natelson D. Hydrogen Stabilization of Metallic Vanadium Dioxide in Single-Crystal Nanobeams. Nat. Nanotechnol. 2012, 7, 357–362. 10.1038/nnano.2012.70. [DOI] [Google Scholar]
  14. Jang H. W.; Felker D. A.; Bark C. W.; Wang Y.; Niranjan M. K.; Nelson C. T.; Zhang Y.; Su D.; Folkman C. M.; Baek S. H.; et al. Metallic and Insulating Oxide Interfaces Controlled by Electronic Correlations. Science 2011, 331, 886–889. 10.1126/science.1198781. [DOI] [PubMed] [Google Scholar]
  15. Lee J. H.; Kim J.-M.; Kim J.-H.; Jang Y.-R.; Kim J. A.; Yeon S.-H.; Lee S.-Y. Energy Storage: Toward Ultrahigh-Capacity V2O5 Lithium-Ion Battery Cathodes via One-Pot Synthetic Route from Precursors to Electrode Sheets. Adv. Mater. Interfaces 2016, 3, 1600173. 10.1002/admi.201600173. [DOI] [Google Scholar]
  16. Cao A.-M.; Hu J.-S.; Liang H.-P.; Wan L.-J. Self-Assembled Vanadium Pentoxide (V2O5) Hollow Microspheres from Nanorods and Their Application in Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2005, 44, 4391–4395. 10.1002/anie.200500946. [DOI] [PubMed] [Google Scholar]
  17. Liu G.; Yu J. C.; Lu G. Q.; Cheng H.-M. Crystal Facet Engineering of Semiconductor Photocatalysts: Motivations, Advances and Unique Properties. Chem. Commun. 2011, 47, 6763–6783. 10.1039/c1cc10665a. [DOI] [PubMed] [Google Scholar]
  18. Foo C. Y.; Sumboja A.; Tan D. J. H.; Wang J.; Lee P. S. Flexible and Highly Scalable V2O5-rGO Electrodes in an Organic Electrolyte for Supercapacitor Devices. Adv. Energy Mater. 2014, 4, 1400236. 10.1002/aenm.201400236. [DOI] [Google Scholar]
  19. Wang S.; Owusu K. A.; Mai L.; Ke Y.; Zhou Y.; Hu P.; Magdassi S.; Long Y. Vanadium Dioxide for Energy Conservation and Energy Storage Applications: Synthesis and Performance Improvement. Appl. Energy 2018, 211, 200–217. 10.1016/j.apenergy.2017.11.039. [DOI] [Google Scholar]
  20. Wu C.; Feng F.; Xie Y. Design of Vanadium Oxide Structures with Controllable Electrical Properties for Energy Applications. Chem. Soc. Rev. 2013, 42, 5157–5183. 10.1039/c3cs35508j. [DOI] [PubMed] [Google Scholar]
  21. Xu X.; Xiong F.; Meng J.; Wang X.; Niu C.; An Q.; Mai L. Vanadium-Based Nanomaterials: A Promising Family for Emerging Metal-Ion Batteries. Adv. Funct. Mater. 2020, 30, 1904398. 10.1002/adfm.201904398. [DOI] [Google Scholar]
  22. McNulty D.; Buckley D. N.; O’Dwyer C. Synthesis and Electrochemical Properties of Vanadium Oxide Materials and Structures as Li-Ion Battery Positive Electrodes. J. Power Sources 2014, 267, 831–873. 10.1016/j.jpowsour.2014.05.115. [DOI] [Google Scholar]
  23. Armer C. F.; Yeoh J. S.; Li X.; Lowe A. Electrospun Vanadium-Based Oxides as Electrode Materials. J. Power Sources 2018, 395, 414–429. 10.1016/j.jpowsour.2018.05.076. [DOI] [Google Scholar]
  24. Cheng Y.; Xia Y.; Chen Y.; Liu Q.; Ge T.; Xu L.; Mai L. Vanadium-Based Nanowires for Sodium-Ion Batteries. Nanotechnology 2019, 30, 192001. 10.1088/1361-6528/aaff82. [DOI] [PubMed] [Google Scholar]
  25. Cheng F.; Chen J. Transition Metal Vanadium Oxides and Vanadate Materials for Lithium Batteries. J. Mater. Chem. 2011, 21, 9841–9848. 10.1039/c0jm04239k. [DOI] [Google Scholar]
  26. Gonçalves J. M.; Ireno da Silva M.; Angnes L.; Araki K. Vanadium-Containing Electro and Photocatalysts for the Oxygen Evolution Reaction: a Review. J. Mater. Chem. A 2020, 8, 2171–2206. 10.1039/C9TA10857B. [DOI] [Google Scholar]
  27. Ke Y.; Wang S.; Liu G.; Li M.; White T. J.; Long Y. Vanadium Dioxide: The Multistimuli Responsive Material and Its Applications. Small 2018, 14, 1802025. 10.1002/smll.201802025. [DOI] [PubMed] [Google Scholar]
  28. Ke Y.; Yin Y.; Zhang Q.; Tan Y.; Hu P.; Wang S.; Tang Y.; Zhou Y.; Wen X.; Wu S.; et al. Adaptive Thermochromic Windows from Active Plasmonic Elastomers. Joule 2019, 3, 858–871. 10.1016/j.joule.2018.12.024. [DOI] [Google Scholar]
  29. Chen D.; Li J.; Wu Q. Review of V2O5-Based Nanomaterials as Electrode for Supercapacitor. J. Nanopart. Res. 2019, 21, 201. 10.1007/s11051-019-4645-8. [DOI] [Google Scholar]
  30. Qin H.; Liang S.; Chen L.; Li Y.; Luo Z.; Chen S. Recent Advances in Vanadium-based Nanomaterials and their Composites for Supercapacitors. Sustain. Energy Fuels 2020, 4, 4902–4933. 10.1039/D0SE00897D. [DOI] [Google Scholar]
  31. Mounasamy V.; Mani G. K.; Madanagurusamy S. Vanadium Oxide Nanostructures for Chemiresistive Gas and Vapour Sensing: a Review on State of the Art. Microchim. Acta 2020, 187, 253. 10.1007/s00604-020-4182-2. [DOI] [PubMed] [Google Scholar]
  32. Morin F. J. Oxides Which Show a Metal-to-Insulator Transition at the Neel Temperature. Phys. Rev. Lett. 1959, 3, 34–36. 10.1103/PhysRevLett.3.34. [DOI] [Google Scholar]
  33. Zylbersztejn A.; Mott N. F. Metal-Insulator Transition in Vanadium Dioxide. Phys. Rev. B 1975, 11, 4383–4395. 10.1103/PhysRevB.11.4383. [DOI] [Google Scholar]
  34. Nadkarni G. S.; Shirodkar V. S. Experiment and Theory for Switching in Al/V2O5/Al Devices. Thin Solid Films 1983, 105, 115–129. 10.1016/0040-6090(83)90200-6. [DOI] [Google Scholar]
  35. Kosuge K.; Takada T.; Kachi S. Phase Diagram and Magnetism of V2O3-V2O5 System. J. Phys. Soc. Jpn. 1963, 18, 318–319. 10.1143/JPSJ.18.318. [DOI] [Google Scholar]
  36. Heidemann A.; Kosuge K.; Ueda Y.; Kachi S. Hyperfine Interaction in V3O7. Phys. Status Solidi A 1977, 39, K37–K40. 10.1002/pssa.2210390152. [DOI] [Google Scholar]
  37. Schwingenschlögl U.; Eyert V. The Vanadium Magnéli Phases VnO2n-1. Ann. Phys. 2004, 516, 475–510. 10.1002/andp.20045160901. [DOI] [Google Scholar]
  38. Selbin J. The Chemistry of Oxovanadium(IV). Chem. Rev. 1965, 65, 153–175. 10.1021/cr60234a001. [DOI] [Google Scholar]
  39. Chirayil T.; Zavalij P. Y.; Whittingham M. S. Hydrothermal Synthesis of Vanadium Oxides. Chem. Mater. 1998, 10, 2629–2640. 10.1021/cm980242m. [DOI] [Google Scholar]
  40. Livage J. Hydrothermal Synthesis of Nanostructured Vanadium Oxides. Materials 2010, 3, 4175–4195. 10.3390/ma3084175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Bahlawane N.; Lenoble D. Vanadium Oxide Compounds: Structure, Properties, and Growth from the Gas Phase. Chem. Vap. Deposition 2014, 20, 299–311. 10.1002/cvde.201400057. [DOI] [Google Scholar]
  42. Hess C. Nanostructured Vanadium Oxide Model Catalysts for Selective Oxidation Reactions. ChemPhysChem 2009, 10, 319–326. 10.1002/cphc.200800585. [DOI] [PubMed] [Google Scholar]
  43. Carrero C. A.; Schloegl R.; Wachs I. E.; Schomaecker R. Critical Literature Review of the Kinetics for the Oxidative Dehydrogenation of Propane over Well-Defined Supported Vanadium Oxide Catalysts. ACS Catal. 2014, 4, 3357–3380. 10.1021/cs5003417. [DOI] [Google Scholar]
  44. Artiglia L.; Agnoli S.; Granozzi G. Vanadium Oxide Nanostructures on Another Oxide: The Viewpoint from Model Catalysts Studies. Coord. Chem. Rev. 2015, 301–302, 106–122. 10.1016/j.ccr.2014.12.015. [DOI] [Google Scholar]
  45. Shvets P.; Dikaya O.; Maksimova K.; Goikhman A. A Review of Raman Spectroscopy of Vanadium Oxides. J. Raman Spectrosc. 2019, 50, 1226–1244. 10.1002/jrs.5616. [DOI] [Google Scholar]
  46. Amiri V.; Roshan H.; Mirzaei A.; Sheikhi M. H. A Review of Nanostructured Resistive-based Vanadium Oxide Gas Sensors. Chemosensors 2020, 8, 105. 10.3390/chemosensors8040105. [DOI] [Google Scholar]
  47. Anjass M.; Lowe G. A.; Streb C. Molecular Vanadium Oxides for Energy Conversion and Energy Storage: Current Trends and Emerging Opportunities. Angew. Chem., Int. Ed. 2021, 60, 7522–7532. 10.1002/anie.202010577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Mai L.; Xu X.; Xu L.; Han C.; Luo Y. Vanadium Oxide Nanowires for Li-Ion Batteries. J. Mater. Res. 2011, 26, 2175–2185. 10.1557/jmr.2011.171. [DOI] [Google Scholar]
  49. Chernova N. A.; Roppolo M.; Dillon A. C.; Whittingham M. S. Layered Vanadium and Molybdenum Oxides: Batteries and Electrochromics. J. Mater. Chem. 2009, 19, 2526–2552. 10.1039/b819629j. [DOI] [Google Scholar]
  50. Rashad M.; Asif M.; Ahmed I.; He Z.; Yin L.; Wei Z. X.; Wang Y. Quest for Carbon and Vanadium Oxide Based Rechargeable Magnesium-Ion Batteries. J. Magnes. Alloy. 2020, 8, 364–373. 10.1016/j.jma.2019.09.010. [DOI] [Google Scholar]
  51. Johnston B.; Henry H.; Kim N.; Lee S. B. Mechanisms of Water-Stimulated Mg2+ Intercalation in Vanadium Oxide: Toward the Development of Hydrated Vanadium Oxide Cathodes for Mg Batteries. Front. Energy Res. 2021, 8, 611391. 10.3389/fenrg.2020.611391. [DOI] [Google Scholar]
  52. Li Y.; Zhang D.; Huang S.; Yang H. Y. Guest-Species-Incorporation in Manganese/Vanadium-based Oxides: Towards High Performance Aqueous Zinc-Ion Batteries. Nano Energy 2021, 85, 105969. 10.1016/j.nanoen.2021.105969. [DOI] [Google Scholar]
  53. Zhang X.; Sun X.; Li X.; Hu X.; Cai S.; Zheng C. Recent Progress in Rate and Cycling Performance Modifications of Vanadium Oxides Cathode for Lithium-Ion Batteries. J. Energy Chem. 2021, 59, 343–363. 10.1016/j.jechem.2020.11.022. [DOI] [Google Scholar]
  54. Mathew V.; Sambandam B.; Kim S.; Kim S.; Park S.; Lee S.; Alfaruqi M. H.; Soundharrajan V.; Islam S.; Putro D. Y.; et al. Manganese and Vanadium Oxide Cathodes for Aqueous Rechargeable Zinc-Ion Batteries: A Focused View on Performance, Mechanism, and Developments. ACS Energy Lett. 2020, 5, 2376–2400. 10.1021/acsenergylett.0c00740. [DOI] [Google Scholar]
  55. Liu Z.; Sun H.; Qin L.; Cao X.; Zhou J.; Pan A.; Fang G.; Liang S. Interlayer Doping in Layered Vanadium Oxides for Low-cost Energy Storage: Sodium-Ion Batteries and Aqueous Zinc-Ion Batteries. ChemNanoMat 2020, 6, 1553–1566. 10.1002/cnma.202000384. [DOI] [Google Scholar]
  56. Jia X.; Liu C.; Neale Z. G.; Yang J.; Cao G. Active Materials for Aqueous Zinc Ion Batteries: Synthesis, Crystal Structure, Morphology, and Electrochemistry. Chem. Rev. 2020, 120, 7795–7866. 10.1021/acs.chemrev.9b00628. [DOI] [PubMed] [Google Scholar]
  57. Banerjee J.; Dutta K. An Overview on the Use of Metal Vanadium Oxides and Vanadates in Supercapacitors and Rechargeable Batteries. Int. J. Energy Res. 2022, 46, 3983–4000. 10.1002/er.7492. [DOI] [Google Scholar]
  58. Ke Y.; Chen J.; Lin G.; Wang S.; Zhou Y.; Yin J.; Lee P. S.; Long Y. Smart Windows: Electro-, Thermo-, Mechano-, Photochromics, and Beyond. Adv. Energy Mater. 2019, 9, 1902066. 10.1002/aenm.201902066. [DOI] [Google Scholar]
  59. Ke Y. J.; Ye S. S.; Hu P.; Jiang H.; Wang S. C.; Yang B.; Zhang J. H.; Long Y. Unpacking the Toolbox of Two-Dimensional Nanostructures Derived from Nanosphere Templates. Mater. Horizons 2019, 6, 1380–1408. 10.1039/C9MH00065H. [DOI] [Google Scholar]
  60. Wriedt H. A. The O-V (Oxygen-Vanadium) System. Bull. Alloy Phase Diagrams 1989, 10, 271–277. 10.1007/BF02877512. [DOI] [Google Scholar]
  61. Cao Z.; Li S.; Xie W.; Du G.; Qiao Z. Critical Evaluation and Thermodynamic Optimization of the V-O System. CALPHAD 2015, 51, 241–251. 10.1016/j.calphad.2015.10.003. [DOI] [Google Scholar]
  62. Yang Y.; Mao H.; Selleby M. Thermodynamic Assessment of the V-O System. CALPHAD 2015, 51, 144–160. 10.1016/j.calphad.2015.08.003. [DOI] [Google Scholar]
  63. Davydov D. A.; Rempel A. A. Refinement of the V-O Phase Diagram in the Range 25–50 at% Oxygen. Inorg. Mater. 2009, 45, 47–54. 10.1134/S0020168509010087. [DOI] [Google Scholar]
  64. Katzke H.; Tolédano P.; Depmeier W. Theory of Morphotropic Transformations in Vanadium Oxides. Phys. Rev. B 2003, 68, 024109. 10.1103/PhysRevB.68.024109. [DOI] [Google Scholar]
  65. Magnéli A. The Crystal Structures of Mo9O26 (Beta’-Molybdenum Oxide) and Mo8O23 (Beta-Molybdenum Oxide). Acta Chem. Scand. 1948, 2, 501–517. 10.3891/acta.chem.scand.02-0501. [DOI] [Google Scholar]
  66. Robinson W. High-Temperature Crystal Chemistry of V2O3 and 1% Chromium-Doped V2O3. Acta. Crystallogr. B 1975, 31, 1153–1160. 10.1107/S0567740875004700. [DOI] [Google Scholar]
  67. Zhu K.; Wei S.; Shou H.; Shen F.; Chen S.; Zhang P.; Wang C.; Cao Y.; Guo X.; Luo M.; Zhang H.; Ye B.; Wu X.; He L.; Song L.; et al. Defect Engineering on V2O3 Cathode for long-Cycling Aqueous Zinc Metal Batteries. Nat. Commun. 2021, 12, 6878. 10.1038/s41467-021-27203-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Jin T.; Li H.; Li Y.; Jiao L.; Chen J. Intercalation Pseudocapacitance in Flexible and Self-Standing V2O3 Porous Nanofibers for High-Rate and Ultra-Stable K Ion Storage. Nano Energy 2018, 50, 462–467. 10.1016/j.nanoen.2018.05.056. [DOI] [Google Scholar]
  69. Kong F. Y.; Li M.; Li D. B.; Xu Y.; Zhang Y. X.; Li G. H. Synthesis and Characterization of V2O3 Nanocrystals by Plasma Hydrogen Reduction. J. Cryst. Growth 2012, 346, 22–26. 10.1016/j.jcrysgro.2012.02.039. [DOI] [Google Scholar]
  70. Corr S. A.; Grossman M.; Furman J. D.; Melot B. C.; Cheetham A. K.; Heier K. R.; Seshadri R. Controlled Reduction of Vanadium Oxide Nanoscrolls: Crystal Structure, Morphology, and Electrical Properties. Chem. Mater. 2008, 20, 6396–6404. 10.1021/cm801539f. [DOI] [Google Scholar]
  71. Qi J.; Ning G.; Zhao Y.; Tian M.; Xu Y.; Hai H. Synthesis and Characterization of V2O3 Microcrystal Particles Controlled by Thermodynamic Parameters. Mater. Sci. Poland 2010, 28, 535–543. [Google Scholar]
  72. Liu X.; Zhang Y.; Yi S.; Huang C.; Liao J.; Li H.; Xiao D.; Tao H. Preparation of V2O3 Nanopowders by Supercritical Fluid Reduction. J. Supercrit. Fluid 2011, 56, 194–200. 10.1016/j.supflu.2010.11.012. [DOI] [Google Scholar]
  73. Mounasamy V.; Mani G. K.; Ponnusamy D.; Tsuchiya K.; Prasad A. K.; Madanagurusamy S. Template-Free Synthesis of Vanadium Sesquioxide (V2O3) Nanosheets and Their Room-Temperature Sensing Performance. J. Mater. Chem. A 2018, 6, 6402–6413. 10.1039/C7TA10159G. [DOI] [Google Scholar]
  74. Pinna N.; Antonietti M.; Niederberger M. A Novel Nonaqueous Route to V2O3 and Nb2O5 Nanocrystals. Colloids Surf., A 2004, 250, 211–213. 10.1016/j.colsurfa.2004.04.078. [DOI] [Google Scholar]
  75. Chen J.; Liu X.; Su Z. Facile Synthesis and Characterisation of Dandelion-Like V2O3 Core-Shell Microspheres. Micro Nano Lett. 2011, 6, 102–105. 10.1049/mnl.2010.0207. [DOI] [Google Scholar]
  76. Han C.; Liu F.; Liu J.; Li Q.; Meng J.; Shao B.; He Q.; Wang X.; Liu Z.; Mai L. Facile Template-Free Synthesis of Uniform Carbon-Confined V2O3 Hollow Spheres for Stable and Fast Lithium Storage. J. Mater. Chem. A 2018, 6, 6220–6224. 10.1039/C8TA01695J. [DOI] [Google Scholar]
  77. Luo H.; Wang B.; Wang F.; Yang J.; Wu F.; Ning Y.; Zhou Y.; Wang D.; Liu H.; Dou S. Anodic Oxidation Strategy toward Structure-Optimized V2O3 Cathode via Electrolyte Regulation for Zn-Ion Storage. ACS Nano 2020, 14, 7328–7337. 10.1021/acsnano.0c02658. [DOI] [PubMed] [Google Scholar]
  78. Ge Y.; He T.; Wang Z.; Han D.; Li J.; Wu J.; Wu J. Chemical Looping Oxidation of CH4 with 99.5% CO Selectivity over V2O3-Based Redox Materials Using CO2 for Regeneration. AIChE J. 2020, 66, e16772 10.1002/aic.16772. [DOI] [Google Scholar]
  79. Zhang Y.; Zhang J.; Nie J.; Zhong Y.; Liu X.; Huang C. Facile Synthesis of V2O3/C Composite and the Effect of V2O3 and V2O3/C on Decomposition of Ammonium Perchlorate. Micro Nano Lett. 2012, 7, 782–785. 10.1049/mnl.2012.0422. [DOI] [Google Scholar]
  80. Xu H.; Liu L.; Gao J.; Du P.; Fang G.; Qiu H.-J. Hierarchical Nanoporous V2O3 Nanosheets Anchored with Alloy Nanoparticles for Efficient Electrocatalysis. ACS Appl. Mater. Interfaces 2019, 11, 38746–38753. 10.1021/acsami.9b13305. [DOI] [PubMed] [Google Scholar]
  81. Zhang J.; Zhou R.-J.; Chang Q.-Y.; Sui Z.-J.; Zhou X.-G.; Chen D.; Zhu Y.-A. Tailoring Catalytic Properties of V2O3 to Propane Dehydrogenation through Single-Atom Doping: A DFT Study. Catal. Today 2021, 368, 46–57. 10.1016/j.cattod.2020.02.023. [DOI] [Google Scholar]
  82. Hu M.; Huang J.; Li Q.; Tu R.; Zhang S.; Yang M.; Li H.; Goto T.; Zhang L. Self-Supported MoSx/V2O3 Heterostructures as Efficient Hybrid Catalysts for Hydrogen Evolution Reaction. J. Alloys Compd. 2020, 827, 154262. 10.1016/j.jallcom.2020.154262. [DOI] [Google Scholar]
  83. Zhai W.; Ma Y.; Chen D.; Ho J. C.; Dai Z.; Qu Y. Recent Progress on the Long-Term Stability of Hydrogen Evolution Reaction Electrocatalysts. InfoMat 2022, 4, e12357 10.1002/inf2.12357. [DOI] [Google Scholar]
  84. Shi F.; Gao W.; Shan H.; Li F.; Xiong Y.; Peng J.; Xiang Q.; Chen W.; Tao P.; Song C.; et al. Strain-Induced Corrosion Kinetics at Nanoscale Are Revealed in Liquid: Enabling Control of Corrosion Dynamics of Electrocatalysis. Chem. 2020, 6, 2257–2271. 10.1016/j.chempr.2020.06.004. [DOI] [Google Scholar]
  85. Masa J.; Andronescu C.; Schuhmann W. Electrocatalysis as the Nexus for Sustainable Renewable Energy: The Gordian Knot of Activity, Stability, and Selectivity. Angew. Chem., Int. Ed. 2020, 59, 15298–15312. 10.1002/anie.202007672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Karmodak N.; Andreussi O. Catalytic Activity and Stability of Two-Dimensional Materials for the Hydrogen Evolution Reaction. ACS Energy Lett. 2020, 5, 885–891. 10.1021/acsenergylett.9b02689. [DOI] [Google Scholar]
  87. Binninger T.; Mohamed R.; Waltar K.; Fabbri E.; Levecque P.; Kötz R.; Schmidt T. J. Thermodynamic Explanation of the Universal Correlation Between Oxygen Evolution Activity and Corrosion of Oxide Catalysts. Sci. Rep. 2015, 5, 12167. 10.1038/srep12167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Mauger A.; Julien C. M. V2O5 Thin Films for Energy Storage and Conversion. AIMS Mater. Sci. 2018, 5, 349–401. 10.3934/matersci.2018.3.349. [DOI] [Google Scholar]
  89. Huang A. L.; El-Kady M. F.; Chang X. Y.; Anderson M.; Lin C. W.; Turner C. L.; Kaner R. B. Facile Fabrication of Multivalent VOx/Graphene Nanocomposite Electrodes for High-Energy-Density Symmetric Supercapacitors. Adv. Energy Mater. 2021, 11, 2100768. 10.1002/aenm.202100768. [DOI] [Google Scholar]
  90. Pan X.; Ren G.; Hoque M. N. F.; Bayne S.; Zhu K.; Fan Z. Fast Supercapacitors Based on Graphene-Bridged V2O3/VOx Core-Shell Nanostructure Electrodes with a Power Density of 1 MW kg–1. Adv. Mater. Interfaces 2014, 1, 1400398. 10.1002/admi.201400398. [DOI] [Google Scholar]
  91. Hassan N.; Riaz J.; Qureshi M. T.; Razaq A.; Rahim M.; Toufiq A. M.; Shakoor A. Vanadium Oxide (V2O3) for Energy Storage Applications through Hydrothermal Route. J. Mater. Sci. Mater. Electron. 2018, 29, 16021–16026. 10.1007/s10854-018-9689-5. [DOI] [Google Scholar]
  92. Yan Y.; Li B.; Guo W.; Pang H.; Xue H. Vanadium Based Materials as Electrode Materials for High Performance Supercapacitors. J. Power Sources 2016, 329, 148–169. 10.1016/j.jpowsour.2016.08.039. [DOI] [Google Scholar]
  93. Zheng J.; Zhang Y.; Meng C.; Wang X.; Liu C.; Bo M.; Pei X.; Wei Y.; Lv T.; Cao G. V2O3/C Nanocomposites with Interface Defects for Enhanced Intercalation Pseudocapacitance. Electrochim. Acta 2019, 318, 635–643. 10.1016/j.electacta.2019.06.125. [DOI] [Google Scholar]
  94. Li H.-Y.; Jiao K.; Wang L.; Wei C.; Li X.; Xie B. Micelle Anchored in Situ Synthesis of V2O3 Nanoflakes@C Composites for Supercapacitors. J. Mater. Chem. A 2014, 2, 18806–18815. 10.1039/C4TA04062G. [DOI] [Google Scholar]
  95. Zhou C.; Wu C.; Liu D.; Yan M. Metal-Organic Framework Derived Hierarchical Co/C@V2O3 Hollow Spheres as a Thin, Lightweight, and High-Efficiency Electromagnetic Wave Absorber. Chem.—Eur. J. 2019, 25, 2234. 10.1002/chem.201805565. [DOI] [PubMed] [Google Scholar]
  96. Andersson G. Studies on Vanadium Oxides. I. Phase Analysis. Acta Chem. Scand. 1954, 8, 1599. 10.3891/acta.chem.scand.08-1599. [DOI] [Google Scholar]
  97. Gray M. L.; Kershaw R.; Croft W.; Dwight K.; Wold A. Crystal Chemistry of V3O5 and Related Structures. J. Solid State Chem. 1986, 62, 57–63. 10.1016/0022-4596(86)90216-1. [DOI] [Google Scholar]
  98. Chen D.; Tan H.; Rui X.; Zhang Q.; Feng Y.; Geng H.; Li C.; Huang S.; Yu Y. Oxyvanite V3O5: A New Intercalation-Type Anode for Lithium-Ion Battery. InfoMat 2019, 1, 251–259. 10.1002/inf2.12011. [DOI] [Google Scholar]
  99. Abdullaev M. A.; Kamilov I. K.; Terukov E. I. Preparation and Properties of Stoichiometric Vanadium Oxides. Inorg. Mater. 2001, 37, 271–273. 10.1023/A:1004121515671. [DOI] [Google Scholar]
  100. Fisher B.; Patlagan L.; Chashka K. B.; Makarov C.; Reisner G. M. V3O5: Insulator-Metal Transition and Electric-Field-Induced Resistive-Switching. Appl. Phys. Lett. 2016, 109, 103501. 10.1063/1.4962334. [DOI] [Google Scholar]
  101. Nagasawa K. Crystal Growth of VnO2n-1 (3 ≤ n ≤ 8) by the Chemical Transport Reaction and Electrical Properties. Mater. Res. Bull. 1971, 6, 853–863. 10.1016/0025-5408(71)90122-X. [DOI] [Google Scholar]
  102. Nagasawa K.; Bando Y.; Takada T. Growth of V3O5 and V6O11 Single Crystals. Jpn. J. Appl. Phys. 1969, 8, 1267–1267. 10.1143/JJAP.8.1267. [DOI] [Google Scholar]
  103. Kumar N.; Rúa A.; Lu J.; Fernández F.; Lysenko S. Ultrafast Excited-State Dynamics of V3O5 as a Signature of a Photoinduced Insulator-Metal Phase Transition. Phys. Rev. Lett. 2017, 119, 057602. 10.1103/PhysRevLett.119.057602. [DOI] [PubMed] [Google Scholar]
  104. Rúa A.; Díaz R. D.; Kumar N.; Lysenko S.; Fernández F. E. Metal-Insulator Transition and Nonlinear Optical Responseof Sputter-Deposited V3O5 Thin Films. J. Appl. Phys. 2017, 121, 235302. 10.1063/1.4986486. [DOI] [Google Scholar]
  105. Perucchi A.; Baldassarre L.; Postorino P.; Lupi S. Optical Properties Across the Insulator to Metal Transitions in Vanadium Oxide Compounds. J. Phys. (Paris) 2009, 21, 323202. 10.1088/0953-8984/21/32/323202. [DOI] [PubMed] [Google Scholar]
  106. Lysenko S.; Fernández F.; Rúa A.; Liu H. Ultrafast Light Scattering Imaging of Multi-Scale Transition Dynamics in Vanadium Dioxide. J. Appl. Phys. 2013, 114, 153514. 10.1063/1.4826074. [DOI] [Google Scholar]
  107. Waltersson K.; Forslund B.; Wilhelmi K.-A.; Andersson S.; Galy J. The Crystal Structure of V3O7. Acta. Crystallogr. B 1974, 30, 2644–2652. 10.1107/S0567740874007722. [DOI] [Google Scholar]
  108. Rastgoo-Deylami M.; Chae M. S.; Hong S.-T. H2V3O8 as a High Energy Cathode Material for Nonaqueous Magnesium-Ion Batteries. Chem. Mater. 2018, 30, 7464–7472. 10.1021/acs.chemmater.8b01381. [DOI] [Google Scholar]
  109. An Q.; Sheng J.; Xu X.; Wei Q.; Zhu Y.; Han C.; Niu C.; Mai L. Ultralong H2V3O8 Nanowire Bundles as a Promising Cathode for Lithium Batteries. New J. Chem. 2014, 38, 2075–2080. 10.1039/C3NJ01134H. [DOI] [Google Scholar]
  110. Li C.; Isobe M.; Ueda H.; Matsushita Y.; Ueda Y. Crystal Growth and Anisotropic Magnetic Properties of V3O7. J. Solid State Chem. 2009, 182, 3222–3225. 10.1016/j.jssc.2009.09.011. [DOI] [Google Scholar]
  111. Wen P.; Liu T.; Wei F.; Ai L.; Yao F. Soft Chemical Topotactic Synthesis and Crystal Structure Evolution from Two-Dimensional KV3O8 Plates to One-Dimensional V3O7 Nanobelts. CrystEngComm 2016, 18, 8880–8886. 10.1039/C6CE01696K. [DOI] [Google Scholar]
  112. Zhao D.; Zhu Q.; Chen D.; Li X.; Yu Y.; Huang X. Nest-Like V3O7 Self-Assembled by Porous Nanowires as an Anode Supercapacitor Material and its Performance Optimization through Bonding with N-Doped Carbon. J. Mater. Chem. A 2018, 6, 16475–16484. 10.1039/C8TA06820H. [DOI] [Google Scholar]
  113. Berenguer R.; Guerrero-Pérez M. O.; Guzmán I.; Rodríguez-Mirasol J.; Cordero T. Synthesis of Vanadium Oxide Nanofibers with Variable Crystallinity and V5+/V4+ Ratios. ACS Omega 2017, 2, 7739–7745. 10.1021/acsomega.7b01061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Chine M.; Sediri F.; Gharbi N. Hydrothermal Synthesis of V3O7·H2O Nanobelts and Study of Their Electrochemical Properties. Mater. Sci. Appl. 2011, 2, 964–970. 10.4236/msa.2011.28129. [DOI] [Google Scholar]
  115. Zhang Y.; Liu X.; Xie G.; Yu L.; Yi S.; Hu M.; Huang C. Hydrothermal Synthesis, Characterization, Formation Mechanism and Electrochemical Property of V3O7·H2O Single-Crystal Nanobelts. Mater. Sci. Eng., B 2010, 175, 164–171. 10.1016/j.mseb.2010.07.023. [DOI] [Google Scholar]
  116. Zhang Y.; Fan M.; Zhou M.; Huang C.; Chen C.; Cao Y.; Xie G.; Li H.; Liu X. Controlled Synthesis and Electrochemical Properties of Vanadium Oxides with Different Nanostructures. Bull. Mater. Sci. 2012, 35, 369–376. 10.1007/s12034-012-0311-9. [DOI] [Google Scholar]
  117. Li G. C.; Pang S. P.; Wang Z. B.; Peng H. R.; Zhang Z. K. Synthesis of H2V3O8 Single-Crystal Nanobelts. Eur. J. Inorg. Chem. 2005, 2005, 2060–2063. 10.1002/ejic.200400967. [DOI] [Google Scholar]
  118. Zhang X.; Yu M.; Zhao S.; Li F.; Hu X.; Guo S.; Lu X.; Tong Y. 3D V3O7·H2O/Partially Exfoliated Carbon Nanotube Composites with Significantly Improved Lithium Storage Ability. Part. Part. Syst. Charact. 2016, 33, 531–537. 10.1002/ppsc.201500130. [DOI] [Google Scholar]
  119. Cao Z.; Chu H.; Zhang H.; Ge Y.; Clemente R.; Dong P.; Wang L.; Shen J.; Ye M.; Ajayan P. M. An in Situ Electrochemical Oxidation Strategy for Formation of Nanogrid-Shaped V3O7·H2O with Enhanced Zinc Storage Properties. J. Mater. Chem. A 2019, 7, 25262–25267. 10.1039/C9TA09116E. [DOI] [Google Scholar]
  120. Yan M.; Wang F.; Han C.; Ma X.; Xu X.; An Q.; Xu L.; Niu C.; Zhao Y.; Tian X.; et al. Nanowire Templated Semihollow Bicontinuous Graphene Scrolls: Designed Construction, Mechanism, and Enhanced Energy Storage Performance. J. Am. Chem. Soc. 2013, 135, 18176–18182. 10.1021/ja409027s. [DOI] [PubMed] [Google Scholar]
  121. Kundu D.; Hosseini Vajargah S.; Wan L.; Adams B.; Prendergast D.; Nazar L. F. Aqueous vs. Nonaqueous Zn-Ion Batteries: Consequences of the Desolvation Penalty at the Interface. Energy Environ. Sci. 2018, 11, 881–892. 10.1039/C8EE00378E. [DOI] [Google Scholar]
  122. Boldyrev V. V. Thermal Decomposition of Ammonium Perchlorate. Thermochim. Acta 2006, 443, 1–36. 10.1016/j.tca.2005.11.038. [DOI] [Google Scholar]
  123. Jacobs P. W. M.; Whitehead H. M. Decomposition and Combustion of Ammonium Perchlorate. Chem. Rev. 1969, 69, 551–590. 10.1021/cr60260a005. [DOI] [Google Scholar]
  124. Zhang Y.; Liu X.; Chen D.; Yu L.; Nie J.; Yi S.; Li H.; Huang C. Fabrication of V3O7·H2O@C Core-Shell Nanostructured Composites and the Effect of V3O7·H2O and V3O7·H2O@C on Decomposition of Ammonium Perchlorate. J. Alloys Compd. 2011, 509, L69–L73. 10.1016/j.jallcom.2010.10.154. [DOI] [Google Scholar]
  125. Manikandan R.; Raj C. J.; Rajesh M.; Kim B. C.; Nagaraju G.; Lee W.-g.; Yu K. H. Rationally Designed Spider Web-Like Trivanadium Heptaoxide Nanowires on Carbon Cloth as a New Class of Pseudocapacitive Electrode for Symmetric Supercapacitors with High Energy Density and Ultra-Long Cyclic Stability. J. Mater. Chem. A 2018, 6, 11390–11404. 10.1039/C8TA03011A. [DOI] [Google Scholar]
  126. Hu T.; Liu Y.; Zhang Y.; Chen M.; Zheng J.; Tang J.; Meng C. 3D Hierarchical Porous V3O7·H2O Nanobelts/CNT/Reduced Graphene Oxide Integrated Composite with Synergistic Effect for Supercapacitors with High Capacitance and Long Cycling Life. J. Colloid Interface Sci. 2018, 531, 382–393. 10.1016/j.jcis.2018.07.060. [DOI] [PubMed] [Google Scholar]
  127. Zhu J.; Cao L.; Wu Y.; Gong Y.; Liu Z.; Hoster H. E.; Zhang Y.; Zhang S.; Yang S.; Yan Q.; et al. Building 3D Structures of Vanadium Pentoxide Nanosheets and Application as Electrodes in Supercapacitors. Nano Lett. 2013, 13, 5408–5413. 10.1021/nl402969r. [DOI] [PubMed] [Google Scholar]
  128. Yin Z.; Xu J.; Ge Y.; Jiang Q.; Zhang Y.; Yang Y.; Sun Y.; Hou S.; Shang Y.; Zhang Y. Synthesis of V2O5 Microspheres by Spray Pyrolysis as Cathode Material for Supercapacitors. Mater. Res. Express. 2018, 5, 036306. 10.1088/2053-1591/aab424. [DOI] [Google Scholar]
  129. Li H.-Y.; Wei C.; Wang L.; Zuo Q.-S.; Li X.; Xie B. Hierarchical Vanadium Oxide Microspheres Forming from Hyperbranched Nanoribbons as Remarkably High Performance Electrode Materials for Supercapacitors. J. Mater. Chem. A 2015, 3, 22892–22901. 10.1039/C5TA06088E. [DOI] [Google Scholar]
  130. Zhang Y.; Wang X.; Jing X.; Meng C. In-situ Synthesis of V2O5 Hollow Spheres Coated Ni-Foam as Binder-Free Electrode for High-Performance Symmetrical Supercapacitor. Mater. Lett. 2019, 248, 101–104. 10.1016/j.matlet.2019.04.014. [DOI] [Google Scholar]
  131. Liu P.; Zhu K.; Bian K.; Xu Y.; Zhang F.; Zhang W.; Zhang J.; Huang W. 3D Hierarchical Porous Sponge-Like V2O5 Micro/Nano-Structures for High-Performance Li-Ion Batteries. J. Alloys Compd. 2018, 765, 901–906. 10.1016/j.jallcom.2018.06.314. [DOI] [Google Scholar]
  132. Song Y.; Liu T.-Y.; Yao B.; Kou T.-Y.; Feng D.-Y.; Liu X.-X.; Li Y. Amorphous Mixed-Valence Vanadium Oxide/Exfoliated Carbon Cloth Structure Shows a Record High Cycling Stability. Small 2017, 13, 1700067. 10.1002/smll.201700067. [DOI] [PubMed] [Google Scholar]
  133. Cristopher M.; Karthick P.; Sivakumar R.; Gopalakrishnan C.; Sanjeeviraja C.; Jeyadheepan K. On the Preparation of Tri-Vanadium Hepta-Oxide Thin Films for Electrochromic Applications. Vacuum 2019, 160, 238–245. 10.1016/j.vacuum.2018.11.042. [DOI] [Google Scholar]
  134. Lu Z.; Zhong X.; Liu X.; Wang J.; Diao X. Energy Storage Electrochromic Devices in the Era of Intelligent Automation. Phys. Chem. Chem. Phys. 2021, 23, 14126–14145. 10.1039/D1CP01398J. [DOI] [PubMed] [Google Scholar]
  135. Deb S. K. Optical and Photoelectric Properties and Colour Centres in Thin Films of Tungsten Oxide. Philos. Mag.: J. Theor. Exp. Appl. Phys. 1973, 27, 801–822. 10.1080/14786437308227562. [DOI] [Google Scholar]
  136. Deb S. K. Opportunities and Challenges of Electrochromic Phenomena in Transition Metal Oxides. Sol. Energy Mater. Sol. Cells 1992, 25, 327–338. 10.1016/0927-0248(92)90077-3. [DOI] [Google Scholar]
  137. Jourdani R.; Jadoual L.; Ait El Fqih M.; El Boujlaidi A.; Aouchiche H.; Kaddouri A. Effects of Lithium Insertion and Deinsertion into V2O5 Thin Films: Optical, Structural, and Absorption Properties. Surf. Interface Anal. 2018, 50, 52–58. 10.1002/sia.6331. [DOI] [Google Scholar]
  138. Mjejri I.; Rougier A. Color Switching in V3O7·H2O Films Cycled in Li and Na Based Electrolytes: Novel Vanadium Oxide Based Electrochromic Materials. J. Mater. Chem. C 2020, 8, 3631–3638. 10.1039/C9TC06753A. [DOI] [Google Scholar]
  139. Li M.; Magdassi S.; Gao Y.; Long Y. Hydrothermal Synthesis of VO2 Polymorphs: Advantages, Challenges and Prospects for the Application of Energy Efficient Smart Windows. Small 2017, 13, 1701147. 10.1002/smll.201701147. [DOI] [PubMed] [Google Scholar]
  140. Wentzcovitch R. M.; Schulz W. W.; Allen P. B. VO2: Peierls or Mott-Hubbard? A View from Band Theory. Phys. Rev. Lett. 1994, 72, 3389–3392. 10.1103/PhysRevLett.72.3389. [DOI] [PubMed] [Google Scholar]
  141. Baum P.; Yang D.-S.; Zewail A. H. 4D Visualization of Transitional Structures in Phase Transformations by Electron Diffraction. Science 2007, 318, 788–792. 10.1126/science.1147724. [DOI] [PubMed] [Google Scholar]
  142. Mott N. Metal-Insulator Transition. Rev. Mod. Phys. 1968, 40, 677. 10.1103/RevModPhys.40.677. [DOI] [Google Scholar]
  143. Wei J.; Wang Z.; Chen W.; Cobden D. H. New Aspects of the Metal-Insulator Transition in Single-Domain Vanadium Dioxide Nanobeams. Nat. Nanotechnol. 2009, 4, 420–424. 10.1038/nnano.2009.141. [DOI] [PubMed] [Google Scholar]
  144. Whittaker L.; Patridge C. J.; Banerjee S. Microscopic and Nanoscale Perspective of the Metal-Insulator Phase Transitions of VO2: Some New Twists to an Old Tale. J. Phys. Chem. Lett. 2011, 2, 745–758. 10.1021/jz101640n. [DOI] [Google Scholar]
  145. Yao T.; Zhang X.; Sun Z.; Liu S.; Huang Y.; Xie Y.; Wu C.; Yuan X.; Zhang W.; Wu Z.; Pan G.; Hu F.; Wu L.; Liu Q.; Wei S. Understanding the Nature of the Kinetic Process in a VO2 Metal-Insulator Transition. Phys. Rev. Lett. 2010, 105, 226405. 10.1103/PhysRevLett.105.226405. [DOI] [PubMed] [Google Scholar]
  146. Shao Z.; Cao X.; Luo H.; Jin P. Recent Progress in the Phase-Transition Mechanism and Modulation of Vanadium Dioxide Materials. NPG Asia Mater. 2018, 10, 581–605. 10.1038/s41427-018-0061-2. [DOI] [Google Scholar]
  147. Lee S.; Ivanov I. N.; Keum J. K.; Lee H. N. Epitaxial Stabilization and Phase Instability of VO2 Polymorphs. Sci. Rep. 2016, 6, 19621. 10.1038/srep19621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Ma H.; Hou J.; Wang X.; Zhang J.; Yuan Z.; Xiao L.; Wei Y.; Fan S.; Jiang K.; Liu K. Flexible, All-Inorganic Actuators Based on Vanadium Dioxide and Carbon Nanotube Bimorphs. Nano Lett. 2017, 17, 421–428. 10.1021/acs.nanolett.6b04393. [DOI] [PubMed] [Google Scholar]
  149. Driscoll T.; Kim H.-T.; Chae B.-G.; Kim B.-J.; Lee Y.-W.; Jokerst N. M.; Palit S.; Smith D. R.; Di Ventra M.; Basov D. N. Memory Metamaterials. Science 2009, 325, 1518–1521. 10.1126/science.1176580. [DOI] [PubMed] [Google Scholar]
  150. Bae S. H.; Lee S.; Koo H.; Lin L.; Jo B. H.; Park C.; Wang Z. L. The Memristive Properties of a Single VO2 Nanowire with Switching Controlled by Self-Heating. Adv. Mater. 2013, 25, 5098–5103. 10.1002/adma.201302511. [DOI] [PubMed] [Google Scholar]
  151. Nakano M.; Shibuya K.; Okuyama D.; Hatano T.; Ono S.; Kawasaki M.; Iwasa Y.; Tokura Y. Collective Bulk Carrier Delocalization Driven by Electrostatic Surface Charge Accumulation. Nature 2012, 487, 459–462. 10.1038/nature11296. [DOI] [PubMed] [Google Scholar]
  152. Beaumont A.; Leroy J.; Orlianges J.-C.; Crunteanu A. Current-Induced Electrical Self-Oscillations Across out-of-plane Threshold Switches Based on VO2 Layers Integrated in Crossbars Geometry. J. Appl. Phys. 2014, 115, 154502. 10.1063/1.4871543. [DOI] [Google Scholar]
  153. Liu M.; Hwang H. Y.; Tao H.; Strikwerda A. C.; Fan K.; Keiser G. R.; Sternbach A. J.; West K. G.; Kittiwatanakul S.; Lu J.; Wolf S. A.; Omenetto F. G.; Zhang X.; Nelson K. A.; Averitt R. D. Terahertz-Field-Induced Insulator-to-Metal Transition in Vanadium Dioxide Metamaterial. Nature 2012, 487, 345–348. 10.1038/nature11231. [DOI] [PubMed] [Google Scholar]
  154. Gao Y.; Luo H.; Zhang Z.; Kang L.; Chen Z.; Du J.; Kanehira M.; Cao C. Nanoceramic VO2 Thermochromic Smart glass: A Review on Progress in Solution Processing. Nano Energy 2012, 1, 221–246. 10.1016/j.nanoen.2011.12.002. [DOI] [Google Scholar]
  155. Seyfouri M. M.; Binions R. Sol-Gel Approaches to Thermochromic Vanadium Dioxide Coating for Smart Glazing Application. Sol. Energy Mater. Sol. Cells 2017, 159, 52–65. 10.1016/j.solmat.2016.08.035. [DOI] [Google Scholar]
  156. Vu T. D.; Chen Z.; Zeng X.; Jiang M.; Liu S.; Gao Y.; Long Y. Physical Vapour Deposition of Vanadium Dioxide for Thermochromic Smart Window Applications. J. Mater. Chem. C 2019, 7, 2121–2145. 10.1039/C8TC05014G. [DOI] [Google Scholar]
  157. Liu K.; Lee S.; Yang S.; Delaire O.; Wu J. Recent Progresses on Physics and Applications of Vanadium Dioxide. Mater. Today 2018, 21, 875–896. 10.1016/j.mattod.2018.03.029. [DOI] [Google Scholar]
  158. Shi R.; Shen N.; Wang J.; Wang W.; Amini A.; Wang N.; Cheng C. Recent Advances in Fabrication Strategies, Phase Transition Modulation, and Advanced Applications of Vanadium Dioxide. Appl. Phys. Rev. 2019, 6, 011312. 10.1063/1.5087864. [DOI] [Google Scholar]
  159. Chang T.-C.; Cao X.; Bao S.-H.; Ji S.-D.; Luo H.-J.; Jin P. Review on Thermochromic Vanadium Dioxide Based Smart Coatings: from Lab to Commercial Application. Adv. Manuf. 2018, 6, 1–19. 10.1007/s40436-017-0209-2. [DOI] [Google Scholar]
  160. Wan C.; Zhang Z.; Woolf D.; Hessel C. M.; Rensberg J.; Hensley J. M.; Xiao Y.; Shahsafi A.; Salman J.; Richter S.; et al. On the Optical Properties of Thin-Film Vanadium Dioxide from the Visible to the Far Infrared. Ann. Phys. 2019, 531, 1900188. 10.1002/andp.201900188. [DOI] [Google Scholar]
  161. Granqvist C. Spectrally Selective Coatings for Energy Efficiency and Solar Applications. Phys. Scr. 1985, 32, 401. 10.1088/0031-8949/32/4/026. [DOI] [Google Scholar]
  162. Huovila P.; Ala-Juusela M.; Melchert L.; Pouffary S.; Cheng C.-C.; Ürge-Vorsatz D.; Koeppel S.; Svenningsen N.; Graham P.. Buildings and Climate Change: Summary for Decision-Makers; United Nations Environment Programme: Nairobi, Kenya, 2009. [Google Scholar]
  163. Cui Y.; Ke Y.; Liu C.; Chen Z.; Wang N.; Zhang L.; Zhou Y.; Wang S.; Gao Y.; Long Y. Thermochromic VO2 for Energy-Efficient Smart Windows. Joule 2018, 2, 1707–1746. 10.1016/j.joule.2018.06.018. [DOI] [Google Scholar]
  164. Chen Z.; Gao Y.; Kang L.; Du J.; Zhang Z.; Luo H.; Miao H.; Tan G. VO2-Based Double-Layered Films for Smart Windows: Optical Design, All-Solution Preparation and Improved Properties. Sol. Energy Mater. Sol. Cells 2011, 95, 2677–2684. 10.1016/j.solmat.2011.05.041. [DOI] [Google Scholar]
  165. Qian X.; Wang N.; Li Y.; Zhang J.; Xu Z.; Long Y. Bioinspired Multifunctional Vanadium Dioxide: Improved Thermochromism and Hydrophobicity. Langmuir 2014, 30, 10766–10771. 10.1021/la502787q. [DOI] [PubMed] [Google Scholar]
  166. Taylor A.; Parkin I.; Noor N.; Tummeltshammer C.; Brown M. S.; Papakonstantinou I. A Bioinspired Solution for Spectrally Selective Thermochromic VO2 Coated Intelligent Glazing. Opt. Express 2013, 21, A750–A764. 10.1364/OE.21.00A750. [DOI] [PubMed] [Google Scholar]
  167. Ke Y.; Balin I.; Wang N.; Lu Q.; Tok A. I. Y.; White T. J.; Magdassi S.; Abdulhalim I.; Long Y. Two-Dimensional SiO2/VO2 Photonic Crystals with Statically Visible and Dynamically Infrared Modulated for Smart Window Deployment. ACS Appl. Mater. Interfaces 2016, 8, 33112–33120. 10.1021/acsami.6b12175. [DOI] [PubMed] [Google Scholar]
  168. Ke Y.; Zhang Q.; Wang T.; Wang S.; Li N.; Lin G.; Liu X.; Dai Z.; Yan J.; Yin J.; et al. Cephalopod-Inspired Versatile Design Based on Plasmonic VO2 Nanoparticle for Energy-Efficient Mechano-Thermochromic Windows. Nano Energy 2020, 73, 104785. 10.1016/j.nanoen.2020.104785. [DOI] [Google Scholar]
  169. Li S.-Y.; Niklasson G. A.; Granqvist C.-G. Nanothermochromics: Calculations for VO2 Nanoparticles in Dielectric Hosts Show Much Improved Luminous Transmittance and Solar Energy Transmittance Modulation. J. Appl. Phys. 2010, 108, 063525. 10.1063/1.3487980. [DOI] [Google Scholar]
  170. Kang L.; Gao Y.; Luo H.; Chen Z.; Du J.; Zhang Z. Nanoporous Thermochromic VO2 Films with Low Optical Constants, Enhanced Luminous Transmittance and Thermochromic Properties. ACS Appl. Mater. Interfaces 2011, 3, 135–138. 10.1021/am1011172. [DOI] [PubMed] [Google Scholar]
  171. Liu C.; Balin I.; Magdassi S.; Abdulhalim I.; Long Y. Vanadium Dioxide Nanogrid Films for High Transparency Smart Architectural Window Applications. Opt. Express 2015, 23, A124–A132. 10.1364/OE.23.00A124. [DOI] [PubMed] [Google Scholar]
  172. Lu Q.; Liu C.; Wang N.; Magdassi S.; Mandler D.; Long Y. Periodic Micro-Patterned VO2 Thermochromic Films by Mesh Printing. J. Mater. Chem. C 2016, 4, 8385–8391. 10.1039/C6TC02694J. [DOI] [Google Scholar]
  173. Zhou C.; Li D.; Tan Y.; Ke Y.; Wang S.; Zhou Y.; Liu G.; Wu S.; Peng J.; Li A.; et al. 3D Printed Smart Windows for Adaptive Solar Modulations. Adv. Opt. Mater. 2020, 8, 2000013. 10.1002/adom.202000013. [DOI] [Google Scholar]
  174. Wang S.; Jiang T.; Meng Y.; Yang R.; Tan G.; Long Y. Scalable Thermochromic Smart Windows with Passive Radiative Cooling Regulation. Science 2021, 374, 1501–1504. 10.1126/science.abg0291. [DOI] [PubMed] [Google Scholar]
  175. Tang K.; Dong K.; Li J.; Gordon M. P.; Reichertz F. G.; Kim H.; Rho Y.; Wang Q.; Lin C.-Y.; Grigoropoulos C. P.; et al. Temperature-Adaptive Radiative Coating for All-Season Household Thermal Regulation. Science 2021, 374, 1504–1509. 10.1126/science.abf7136. [DOI] [PubMed] [Google Scholar]
  176. Ke Y.; Li Y.; Wu L.; Wang S.; Yang R.; Yin J.; Tan G.; Long Y. On-Demand Solar and Thermal Radiation Management Based on Switchable Interwoven Surfaces. ACS Energy Lett. 2022, 7, 1758–1763. 10.1021/acsenergylett.2c00419. [DOI] [Google Scholar]
  177. Wang S.; Zhou Y.; Jiang T.; Yang R.; Tan G.; Long Y. Thermochromic Smart Windows with Highly Regulated Radiative Cooling and Solar Transmission. Nano Energy 2021, 89, 106440. 10.1016/j.nanoen.2021.106440. [DOI] [Google Scholar]
  178. Vu T. D.; Xie H.; Wang S.; Hu J.; Zeng X.; Long Y. Durable Vanadium Dioxide with 33-Year Service Life for Smart Windows Applications. Mater. Today Energy 2022, 26, 100978. 10.1016/j.mtener.2022.100978. [DOI] [Google Scholar]
  179. Zhou X.; Meng Y.; Vu T. D.; Gu D.; Jiang Y.; Mu Q.; Li Y.; Yao B.; Dong Z.; Liu Q.; et al. A New Strategy of Nanocompositing Vanadium Dioxide with Excellent Durability. J. Mater. Chem. A 2021, 9, 15618–15628. 10.1039/D1TA02525B. [DOI] [Google Scholar]
  180. Xiao L.; Ma H.; Liu J.; Zhao W.; Jia Y.; Zhao Q.; Liu K.; Wu Y.; Wei Y.; Fan S.; Jiang K. Fast Adaptive Thermal Camouflage Based on Flexible VO2/Graphene/CNT Thin Films. Nano Lett. 2015, 15, 8365–8370. 10.1021/acs.nanolett.5b04090. [DOI] [PubMed] [Google Scholar]
  181. Wang S.; Liu G.; Hu P.; Zhou Y.; Ke Y.; Li C.; Chen J.; Cao T.; Long Y. Largely Lowered Transition Temperature of a VO2/Carbon Hybrid Phase Change Material with High Thermal Emissivity Switching Ability and Near Infrared Regulations. Adv. Mater. Interfaces 2018, 5, 1801063. 10.1002/admi.201801063. [DOI] [Google Scholar]
  182. Ji H.; Liu D.; Zhang C.; Cheng H. VO2/ZnS Core-Shell Nanoparticle for the Adaptive Infrared Camouflage Application with Modified Color and Enhanced Oxidation Resistance. Sol. Energy Mater. Sol. Cells 2018, 176, 1–8. 10.1016/j.solmat.2017.11.037. [DOI] [Google Scholar]
  183. Taylor S.; Yang Y.; Wang L. Vanadium Dioxide Based Fabry-Perot Emitter for Dynamic Radiative Cooling Applications. J. Quant. Spectrosc. Radiat. Transfer 2017, 197, 76–83. 10.1016/j.jqsrt.2017.01.014. [DOI] [Google Scholar]
  184. Kim H.; Cheung K.; Auyeung R. C. Y.; Wilson D. E.; Charipar K. M.; Pique A.; Charipar N. A. VO2-based Switchable Radiator for Spacecraft Thermal Control. Sci. Rep. 2019, 9, 11329. 10.1038/s41598-019-47572-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Gomez-Heredia C. L.; Ramirez-Rincon J. A.; Ordonez-Miranda J.; Ares O.; Alvarado-Gil J. J.; Champeaux C.; Dumas-Bouchiat F.; Ezzahri Y.; Joulain K. Thermal Hysteresis Measurement of the VO2 Emissivity and its Application in Thermal Rectification. Sci. Rep. 2018, 8, 8479. 10.1038/s41598-018-26687-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Katase T.; Endo K.; Ohta H. Infrared-Transmittance Tunable Metal-Insulator Conversion Device with Thin-Film-Transistor-Type Structure on a Glass Substrate. APL Mater. 2017, 5, 056105. 10.1063/1.4983276. [DOI] [Google Scholar]
  187. Zhang P.; Zhang W.; Wang J.; Jiang K.; Zhang J.; Li W.; Wu J.; Hu Z.; Chu J. The Electro-Optic Mechanism and Infrared Switching Dynamic of the Hybrid Multilayer VO2/Al: ZnO Heterojunctions. Sci. Rep. 2017, 7, 1–14. 10.1038/s41598-017-04660-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Zhou Y.; Layani M.; Boey F. Y. C.; Sokolov I.; Magdassi S.; Long Y. Electro-Thermochromic Devices Composed of Self-Assembled Transparent Electrodes and Hydrogels. Adv. Mater. Technol. 2016, 1, 1600069. 10.1002/admt.201600069. [DOI] [Google Scholar]
  189. Fan L.; Chen Y.; Liu Q.; Chen S.; Zhu L.; Meng Q.; Wang B.; Zhang Q.; Ren H.; Zou C. Infrared Response and Optoelectronic Memory Device Fabrication Based on Epitaxial VO2 Film. ACS Appl. Mater. Interfaces 2016, 8, 32971–32977. 10.1021/acsami.6b12831. [DOI] [PubMed] [Google Scholar]
  190. Briggs R. M.; Pryce I. M.; Atwater H. A. Compact Silicon Photonic Waveguide Modulator Based on the Vanadium Dioxide Metal-Insulator Phase Transition. Opt. Express 2010, 18, 11192–11201. 10.1364/OE.18.011192. [DOI] [PubMed] [Google Scholar]
  191. Lee J.; Lee D.; Cho S. J.; Seo J.-H.; Liu D.; Eom C.-B.; Ma Z. Epitaxial VO2 Thin Film-based Radio-Frequency Switches with Thermal Activation. Appl. Phys. Lett. 2017, 111, 063110. 10.1063/1.4998452. [DOI] [Google Scholar]
  192. Sun M.; Taha M.; Walia S.; Bhaskaran M.; Sriram S.; Shieh W.; Unnithan R. R. A Photonic Switch Based on a Hybrid Combination of Metallic Nanoholes and Phase-Change Vanadium Dioxide. Sci. Rep. 2018, 8, 11106. 10.1038/s41598-018-29476-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Ding F.; Zhong S.; Bozhevolnyi S. I. Vanadium Dioxide Integrated Metasurfaces with Switchable Functionalities at Terahertz Frequencies. Adv. Opt. Mater. 2018, 6, 1701204. 10.1002/adom.201701204. [DOI] [Google Scholar]
  194. Liu M.; Xu Q.; Chen X.; Plum E.; Li H.; Zhang X.; Zhang C.; Zou C.; Han J.; Zhang W. Temperature-Controlled Asymmetric Transmission of Electromagnetic Waves. Sci. Rep. 2019, 9, 4097. 10.1038/s41598-019-40791-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Wegkamp D.; Stähler J. Ultrafast Dynamics During the Photoinduced Phase Transition in VO2. Prog. Surf. Sci. 2015, 90, 464–502. 10.1016/j.progsurf.2015.10.001. [DOI] [Google Scholar]
  196. Guo P.; Weimer M. S.; Emery J. D.; Diroll B. T.; Chen X.; Hock A. S.; Chang R. P.; Martinson A. B.; Schaller R. D. Conformal Coating of a Phase Change Material on Ordered Plasmonic Nanorod Arrays for Broadband All-Optical Switching. ACS Nano 2017, 11, 693–701. 10.1021/acsnano.6b07042. [DOI] [PubMed] [Google Scholar]
  197. Lei D. Y.; Appavoo K.; Ligmajer F.; Sonnefraud Y.; Haglund Jr R. F.; Maier S. A. Optically-Triggered Nanoscale Memory Effect in a Hybrid Plasmonic-Phase Changing Nanostructure. ACS Photonics 2015, 2, 1306–1313. 10.1021/acsphotonics.5b00249. [DOI] [Google Scholar]
  198. Ke Y.; Wen X.; Zhao D.; Che R.; Xiong Q.; Long Y. Controllable Fabrication of Two-Dimensional Patterned VO2 Nanoparticle, Nanodome, and Nanonet Arrays with Tunable Temperature-Dependent Localized Surface Plasmon Resonance. ACS Nano 2017, 11, 7542–7551. 10.1021/acsnano.7b02232. [DOI] [PubMed] [Google Scholar]
  199. Ke Y.; Zhang B.; Wang T.; Zhong Y.; Vu T. D.; Wang S.; Liu Y.; Magdassi S.; Ye X.; Zhao D.; et al. Manipulating Atomic Defects in Plasmonic Vanadium Dioxide for Superior Solar and Thermal Management. Mater. Horizons 2021, 8, 1700–1710. 10.1039/D1MH00413A. [DOI] [PubMed] [Google Scholar]
  200. Kim B.-J.; Lee Y. W.; Chae B.-G.; Yun S. J.; Oh S.-Y.; Kim H.-T.; Lim Y.-S. Temperature Dependence of the First-Order Metal-Insulator Transition in VO2 and Programmable Critical Temperature Sensor. Appl. Phys. Lett. 2007, 90, 023515. 10.1063/1.2431456. [DOI] [Google Scholar]
  201. Yang Z.; Ko C.; Balakrishnan V.; Gopalakrishnan G.; Ramanathan S. Dielectric and Carrier Transport Properties of Vanadium Dioxide Thin Films Across the Phase Transition Utilizing Gated Capacitor Devices. Phys. Rev. B 2010, 82, 205101. 10.1103/PhysRevB.82.205101. [DOI] [Google Scholar]
  202. Hou J.; Wang Z.; Ding Z.; Zhang Z.; Zhang J. Facile Synthesize VO2 (M1) Nanorods for a Low-Cost Infrared Photodetector Application. Sol. Energy Mater. Sol. Cells 2018, 176, 142–149. 10.1016/j.solmat.2017.11.030. [DOI] [Google Scholar]
  203. Takeya H.; Frame J.; Tanaka T.; Urade Y.; Fang X.; Kubo W. Bolometric Photodetection Using Plasmon-Assisted Resistivity Change in Vanadium Dioxide. Sci. Rep. 2018, 8, 12764. 10.1038/s41598-018-30944-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Hu B.; Ding Y.; Chen W.; Kulkarni D.; Shen Y.; Tsukruk V. V.; Wang Z. L. External-Strain Induced Insulating Phase Transition in VO2 Nanobeam and Its Application as Flexible Strain Sensor. Adv. Mater. 2010, 22, 5134–5139. 10.1002/adma.201002868. [DOI] [PubMed] [Google Scholar]
  205. Liang J.; Zhao Y.; Zhu K.; Guo J.; Zhou L. Synthesis and Room Temperature NO2 Gas Sensitivity of Vanadium Dioxide Nanowire Structures by Chemical Vapor Deposition. Thin Solid Films 2019, 669, 537–543. 10.1016/j.tsf.2018.11.046. [DOI] [Google Scholar]
  206. Evans G. P.; Powell M. J.; Johnson I. D.; Howard D. P.; Bauer D.; Darr J. A.; Parkin I. P. Room Temperature Vanadium Dioxide-Carbon Nanotube Gas Sensors Made via Continuous Hydrothermal Flow Synthesis. Sens. Actuator B 2018, 255, 1119–1129. 10.1016/j.snb.2017.07.152. [DOI] [Google Scholar]
  207. Zhou Y.; Chen X.; Ko C.; Yang Z.; Mouli C.; Ramanathan S. Voltage-Triggered Ultrafast Phase Transition in Vanadium Dioxide Switches. IEEE Electron Device Lett. 2013, 34, 220–222. 10.1109/LED.2012.2229457. [DOI] [Google Scholar]
  208. Li D.; Sharma A. A.; Gala D. K.; Shukla N.; Paik H.; Datta S.; Schlom D. G.; Bain J. A.; Skowronski M. Joule Heating-Induced Metal-Insulator Transition in Epitaxial VO2/TiO2 Devices. ACS Appl. Mater. Interfaces 2016, 8, 12908–12914. 10.1021/acsami.6b03501. [DOI] [PubMed] [Google Scholar]
  209. Yajima T.; Nishimura T.; Toriumi A. Positive-Bias Gate-Controlled Metal-Insulator Transition in Ultrathin VO2 Channels with TiO2 Gate Dielectrics. Nat. Commun. 2015, 6, 1–9. 10.1038/ncomms10104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Shukla N.; Thathachary A. V.; Agrawal A.; Paik H.; Aziz A.; Schlom D. G.; Gupta S. K.; Engel-Herbert R.; Datta S. A Steep-Slope Transistor Based on Abrupt Electronic Phase Transition. Nat. Commun. 2015, 6, 1–6. 10.1038/ncomms8812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  211. Zhang Y.; Xiong W.; Chen W.; Luo X.; Zhang X.; Zheng Y. Nonvolatile Ferroelectric Field Effect Transistor Based on a Vanadium Dioxide Nanowire with Large On- and Off-Field Resistance Switching. Phys. Chem. Chem. Phys. 2020, 22, 4685–4691. 10.1039/C9CP06428A. [DOI] [PubMed] [Google Scholar]
  212. Goldflam M. D.; Liu M. K.; Chapler B. C.; Stinson H. T.; Sternbach A. J.; McLeod A. S.; Zhang J. D.; Geng K.; Royal M.; Kim B.-J.; Averitt R. D.; Jokerst N. M.; Smith D. R.; Kim H-T.; Basov D. N. Voltage Switching of a VO2 Memory Metasurface Using Ionic Gel. Appl. Phys. Lett. 2014, 105, 041117. 10.1063/1.4891765. [DOI] [Google Scholar]
  213. Jeong J.; Aetukuri N.; Graf T.; Schladt T. D.; Samant M. G.; Parkin S. S. Suppression of Metal-Insulator Transition in VO2 by Electric Field-Induced Oxygen Vacancy Formation. Science 2013, 339, 1402–1405. 10.1126/science.1230512. [DOI] [PubMed] [Google Scholar]
  214. Ji H.; Wei J.; Natelson D. Modulation of the Electrical Properties of VO2 Nanobeams Using an Ionic Liquid as a Gating Medium. Nano Lett. 2012, 12, 2988–2992. 10.1021/nl300741h. [DOI] [PubMed] [Google Scholar]
  215. Shibuya K.; Sawa A. Modulation of Metal-Insulator Transition in VO2 by Electrolyte Gating-Induced Protonation. Adv. Electron. Mater. 2016, 2, 1500131. 10.1002/aelm.201500131. [DOI] [Google Scholar]
  216. Leroy J.; Crunteanu A.; Givernaud J.; Orlianges J.-C.; Champeaux C.; Blondy P. Generation of Electrical Self-Oscillations in Two-Terminal Switching Devices Based on the Insulator-to-Metal Phase Transition of VO2 Thin Films. Int. J. Microw. Wirel. Technol. 2012, 4, 101–107. 10.1017/S175907871100095X. [DOI] [Google Scholar]
  217. Zhou Y.; Ramanathan S. Mott Memory and Neuromorphic Devices. Proc. IEEE 2015, 103, 1289–1310. 10.1109/JPROC.2015.2431914. [DOI] [Google Scholar]
  218. Pellegrino L.; Manca N.; Kanki T.; Tanaka H.; Biasotti M.; Bellingeri E.; Siri A. S.; Marré D. Multistate Memory Devices Based on Free-Standing VO2/TiO2 Microstructures Driven by Joule Self-Heating. Adv. Mater. 2012, 24, 2929–2934. 10.1002/adma.201104669. [DOI] [PubMed] [Google Scholar]
  219. Janod E.; Tranchant J.; Corraze B.; Querré M.; Stoliar P.; Rozenberg M.; Cren T.; Roditchev D.; Phuoc V. T.; Besland M.-P.; et al. Resistive Switching in Mott Insulators and Correlated Systems. Adv. Funct. Mater. 2015, 25, 6287–6305. 10.1002/adfm.201500823. [DOI] [Google Scholar]
  220. Yin H.; Yu K.; Song C.; Wang Z.; Zhu Z. Low-Temperature CVD Synthesis of Patterned Core-Shell VO2@ ZnO Nanotetrapods and Enhanced Temperature-Dependent Field-Emission Properties. Nanoscale 2014, 6, 11820–11827. 10.1039/C4NR02661F. [DOI] [PubMed] [Google Scholar]
  221. Li Z.; Guo Y.; Hu Z.; Su J.; Zhao J.; Wu J.; Wu J.; Zhao Y.; Wu C.; Xie Y. Hydrogen Treatment for Superparamagnetic VO2 Nanowires with Large Room-Temperature Magnetoresistance. Angew. Chem., Int. Ed. 2016, 55, 8018–8022. 10.1002/anie.201603406. [DOI] [PubMed] [Google Scholar]
  222. Choi J.; Kim B.-J.; Seo G.; Kim H.-T.; Cho S.; Lee Y. W. Magnetic Field-Dependent Ordinary Hall Effect and Thermopower of VO2 Thin Films. Curr. Appl. Phys. 2016, 16, 335–339. 10.1016/j.cap.2015.11.023. [DOI] [Google Scholar]
  223. Singh D.; Yadav C.; Viswanath B. Magnetoresistance Across Metal-Insulator Transition in VO2 Micro Crystals. Mater. Lett. 2017, 196, 248–251. 10.1016/j.matlet.2017.03.066. [DOI] [Google Scholar]
  224. Mirfakhrai T.; Madden J. D.; Baughman R. H. Polymer Artificial Muscles. Mater. Today 2007, 10, 30–38. 10.1016/S1369-7021(07)70048-2. [DOI] [Google Scholar]
  225. Liu K.; Cheng C.; Cheng Z.; Wang K.; Ramesh R.; Wu J. Giant-Amplitude, High-Work Density Microactuators with Phase Transition Activated Nanolayer Bimorphs. Nano Lett. 2012, 12, 6302–6308. 10.1021/nl303405g. [DOI] [PubMed] [Google Scholar]
  226. Ma H.; Zhang X.; Cui R.; Liu F.; Wang M.; Huang C.; Hou J.; Wang G.; Wei Y.; Jiang K.; et al. Photo-Driven Nanoactuators Based on Carbon Nanocoils and Vanadium Dioxide Bimorphs. Nanoscale 2018, 10, 11158–11164. 10.1039/C8NR03622E. [DOI] [PubMed] [Google Scholar]
  227. Shi R.; Cai X.; Wang W.; Wang J.; Kong D.; Cai N.; Chen P.; He P.; Wu Z.; Amini A.; et al. Single-Crystalline Vanadium Dioxide Actuators. Adv. Funct. Mater. 2019, 29, 1900527. 10.1002/adfm.201900527. [DOI] [Google Scholar]
  228. Manca N.; Pellegrino L.; Kanki T.; Venstra W. J.; Mattoni G.; Higuchi Y.; Tanaka H.; Caviglia A. D.; Marré D. Selective High-Frequency Mechanical Actuation Driven by the VO2 Electronic Instability. Adv. Mater. 2017, 29, 1701618. 10.1002/adma.201701618. [DOI] [PubMed] [Google Scholar]
  229. Torres D.; Wang T.; Zhang J.; Zhang X.; Dooley S.; Tan X.; Xie H.; Sepúlveda N. VO2-Based MEMS Mirrors. J. Microelectromech. Syst. 2016, 25, 780–787. 10.1109/JMEMS.2016.2562609. [DOI] [Google Scholar]
  230. Holsteen A.; Kim I. S.; Lauhon L. J. Extraordinary Dynamic Mechanical Response of Vanadium Dioxide Nanowires Around the Insulator to Metal Phase Transition. Nano Lett. 2014, 14, 1898–1902. 10.1021/nl404678k. [DOI] [PubMed] [Google Scholar]
  231. Rúa A.; Cabrera R.; Coy H.; Merced E.; Sepúlveda N.; Fernández F. E. Phase Transition Behavior in Microcantilevers Coated with M1-Phase VO2 and M2-phase VO2:Cr Thin Films. J. Appl. Phys. 2012, 111, 104502. 10.1063/1.4716191. [DOI] [Google Scholar]
  232. Manca N.; Pellegrino L.; Kanki T.; Yamasaki S.; Tanaka H.; Siri A. S.; Marré D. Programmable Mechanical Resonances in MEMS by Localized Joule Heating of Phase Change Materials. Adv. Mater. 2013, 25, 6430–6435. 10.1002/adma.201302087. [DOI] [PubMed] [Google Scholar]
  233. Schmidt C. N.; Cao G. Properties of Mesoporous Carbon Modified Carbon Felt for Anode of All-Vanadium Redox Flow Battery. Sci. China Mater. 2016, 59, 1037–1050. 10.1007/s40843-016-5114-8. [DOI] [Google Scholar]
  234. Zhang J.; Chen L.; Wang Y.; Cai S.; Yang H.; Yu H.; Ding F.; Huang C.; Liu X. VO(2)(B)/Graphene Composite-Based Symmetrical Supercapacitor Electrode via Screen Printing for Intelligent Packaging. Nanomaterials 2018, 8, 1020. 10.3390/nano8121020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Lv W.; Yang C.; Meng G.; Zhao R.; Han A.; Wang R.; Liu J. VO2(B) Nanobelts/Reduced Graphene Oxide Composites for High-Performance Flexible All-Solid-State Supercapacitors. Sci. Rep. 2019, 9, 10831. 10.1038/s41598-019-47266-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Wu C.; Zhang X.; Dai J.; Yang J.; Wu Z.; Wei S.; Xie Y. Direct Hydrothermal Synthesis of Monoclinic VO2 (M) Single-Domain Nanorods on Large Scale Displaying Magnetocaloric Effect. J. Mater. Chem. 2011, 21, 4509–4517. 10.1039/c0jm03078c. [DOI] [Google Scholar]
  237. Chao D.; Zhu C.; Xia X.; Liu J.; Zhang X.; Wang J.; Liang P.; Lin J.; Zhang H.; Shen Z. X.; et al. Graphene Quantum Dots Coated VO2 Arrays for Highly Durable Electrodes for Li and Na Ion Batteries. Nano Lett. 2015, 15, 565–573. 10.1021/nl504038s. [DOI] [PubMed] [Google Scholar]
  238. Li Y.; Zhang Q.; Yuan Y.; Liu H.; Yang C.; Lin Z.; Lu J. Surface Amorphization of Vanadium Dioxide (B) for K-Ion Battery. Adv. Energy Mater. 2020, 10, 2000717. 10.1002/aenm.202000717. [DOI] [Google Scholar]
  239. Chen L.; Ruan Y.; Zhang G.; Wei Q.; Jiang Y.; Xiong T.; He P.; Yang W.; Yan M.; An Q.; et al. Ultrastable and High-Performance Zn/VO2 Battery Based on a Reversible Single-Phase Reaction. Chem. Mater. 2019, 31, 699–706. 10.1021/acs.chemmater.8b03409. [DOI] [Google Scholar]
  240. Niu C.; Meng J.; Han C.; Zhao K.; Yan M.; Mai L. VO2 Nanowires Assembled into Hollow Microspheres for High-Rate and Long-Life Lithium Batteries. Nano Lett. 2014, 14, 2873–2878. 10.1021/nl500915b. [DOI] [PubMed] [Google Scholar]
  241. Thomas G. A.; Rapkine D. H.; Carter S. A.; Millis A. J.; Rosenbaum T. F.; Metcalf P.; Honig J. M. Observation of the Gap and Kinetic Energy in a Correlated Insulator. Phys. Rev. Lett. 1994, 73, 1529–1532. 10.1103/PhysRevLett.73.1529. [DOI] [PubMed] [Google Scholar]
  242. Wang Y.; Zhang Z.; Zhu Y.; Li Z.; Vajtai R.; Ci L.; Ajayan P. M. Nanostructured VO2 Photocatalysts for Hydrogen Production. ACS Nano 2008, 2, 1492–1496. 10.1021/nn800223s. [DOI] [PubMed] [Google Scholar]
  243. Zhang Q.; Liu B.; Li L.; Ji Y.; Wang C.; Zhang L.; Su Z. Maximized Schottky Effect: The Ultrafine V2O3/Ni Heterojunctions Repeatedly Arranging on Monolayer Nanosheets for Efficient and Stable Water-to-Hydrogen Conversion. Small 2021, 17, 2005769. 10.1002/smll.202005769. [DOI] [PubMed] [Google Scholar]
  244. Lv Q.; Yang L.; Wang W.; Lu S.; Wang T.; Cao L.; Dong B. One-Step Construction of Core/Shell Nanoarrays with a Holey Shell and Exposed Interfaces for Overall Water Splitting. J. Mater. Chem. A 2019, 7, 1196–1205. 10.1039/C8TA10686J. [DOI] [Google Scholar]
  245. Chen G.; Zhang X.; Guan L.; Zhang H.; Xie X.; Chen S.; Tao J. Phase Transition-Promoted Hydrogen Evolution Performance of MoS2/VO2 Hybrids. J. Phys. Chem. C 2018, 122, 2618–2623. 10.1021/acs.jpcc.7b12040. [DOI] [Google Scholar]
  246. Hu M.; Hu J.; Zheng Y.; Zhang S.; Li Q.; Yang M.; Goto T.; Tu R. Heterostructured Co3O4/VO2 Nanosheet Array Catalysts on Carbon Cloth for Hydrogen Evolution Reaction. Int. J. Hydrog. Energy 2022, 47, 18983–18991. 10.1016/j.ijhydene.2022.04.062. [DOI] [Google Scholar]
  247. Najafi L.; Oropesa-Nuñez R.; Bellani S.; Martín-García B.; Pasquale L.; Serri M.; Drago F.; Luxa J.; Sofer Z.; Sedmidubský D.; et al. Topochemical Transformation of Two-Dimensional VSe2 into Metallic Nonlayered VO2 for Water Splitting Reactions in Acidic and Alkaline Media. ACS Nano 2022, 16, 351–367. 10.1021/acsnano.1c06662. [DOI] [PubMed] [Google Scholar]
  248. Zou C. W.; Fan L. L.; Chen R. Q.; Yan X. D.; Yan W. S.; Pan G. Q.; Wu Z. Y.; Gao W. Thermally Driven V2O5 Nanocrystal Formation and the Temperature-Dependent Electronic Structure Study. CrystEngComm 2012, 14, 626–631. 10.1039/C1CE06170D. [DOI] [Google Scholar]
  249. Yao J. H.; Li Y. W.; Masse R. C.; Uchaker E.; Cao G. Z. Revitalized Interest in Vanadium Pentoxide as Cathode Material for Lithium-Ion Batteries and Beyond. Energy Stor. Mater. 2018, 11, 205–259. 10.1016/j.ensm.2017.10.014. [DOI] [Google Scholar]
  250. Zhou Y.; Pan Q.; Zhang J.; Han C.; Wang L.; Xu H. Insights into Synergistic Effect of Acid on orphological Control of Vanadium Oxide: Toward High Lithium Storage. Adv. Sci. 2021, 8, 2002579. 10.1002/advs.202002579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  251. Nguyen T. P. T.; Kim M. H.; Yang K. H. Formation and Depression of N2O in Selective Reduction of NO by NH3 over Fe2O3-Promoted V2O5-WO3/TiO2 Catalysts: Roles of each Constituent and Strongly-Adsorbed NH3 Species. Catal. Today 2021, 375, 565–575. 10.1016/j.cattod.2020.05.008. [DOI] [Google Scholar]
  252. Yang J. H.; Lee H. J.; Lee H. S.; Jeon S. C.; Han Y. S. Precise Control of Heat-Treatment Conditions to Improve the Catalytic Performance of V2O5/TiO2 for H2S Removal. J. Hazard. Mater. 2021, 416, 125974. 10.1016/j.jhazmat.2021.125974. [DOI] [PubMed] [Google Scholar]
  253. Kang T. H.; Youn S.; Kim D. H. Improved Catalytic Performance and Resistance to SO2 over V2O5-WO3/TiO2 Catalyst Physically Mixed with Fe2O3 for Low-Temperature NH3-SCR. Catal. Today 2021, 376, 95–103. 10.1016/j.cattod.2020.07.042. [DOI] [Google Scholar]
  254. Meyer J.; Zilberberg K.; Riedl T.; Kahn A.. Electronic Structure of Vanadium Pentoxide: An Efficient Hole Injector for Organic Electronic Materials. J. Appl. Phys. 2011, 110.033710. 10.1063/1.3611392 [DOI] [Google Scholar]
  255. Zhang W.; Li H.; Yu W. W.; Elezzabi A. Y. Transparent Inorganic Multicolour Displays Enabled by Zinc-Based Electrochromic Devices. Light Sci. Appl. 2020, 9, 121. 10.1038/s41377-020-00366-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. Beke S. A Review of the Growth of V2O5 Films from 1885 to 2010. Thin Solid Films 2011, 519, 1761–1771. 10.1016/j.tsf.2010.11.001. [DOI] [Google Scholar]
  257. Steunou N.; Livage J. Rational Design of One-Dimensional Vanadium(V) Oxide Nanocrystals: An Insight into the Physico-Chemical Parameters Controlling the Crystal Structure, Morphology and Size of Particles. CrystEngComm 2015, 17, 6780–6795. 10.1039/C5CE00554J. [DOI] [Google Scholar]
  258. Alonso B.; Livage J. Synthesis of Vanadium Oxide Gels from Peroxovanadic Acid Solutions: A 51V NMR Study. J. Solid State Chem. 1999, 148, 16–19. 10.1006/jssc.1999.8283. [DOI] [Google Scholar]
  259. Etman A. S.; Pell A. J.; Svedlindh P.; Hedin N.; Zou X.; Sun J.; Bernin D. Insights into the Exfoliation Process of V2O5·nH2O Nanosheet Formation Using Real-Time 51V NMR. ACS Omega 2019, 4, 10899–10905. 10.1021/acsomega.9b00727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Livage J. Vanadium Pentoxide Gels. Chem. Mater. 1991, 3, 578–593. 10.1021/cm00016a006. [DOI] [Google Scholar]
  261. Li M.; Kong F. Y.; Wang H. Q.; Li G. H. Synthesis of Vanadium Pentoxide (V2O5) Ultralong Nanobelts via an Oriented Attachment Growth Mechanism. CrystEngComm 2011, 13, 5317–5320. 10.1039/c1ce05477e. [DOI] [Google Scholar]
  262. Grayli S. V.; Leach G. W.; Bahreyni B. Sol-Gel Deposition and Characterization of Vanadium Pentoxide Thin Films with High TCR. Sens. Actuator A 2018, 279, 630–637. 10.1016/j.sna.2018.07.002. [DOI] [Google Scholar]
  263. Surca A. K.; Drazic G.; Mihelcic M. Low-Temperature V-Oxide Film for a Flexible Electrochromic Device: Comparison of Its Electrochromic, IR and Raman Properties to Those of a Crystalline V2O5 Film. Sol. Energy Mater. Sol. Cells 2019, 196, 185–199. 10.1016/j.solmat.2019.03.017. [DOI] [Google Scholar]
  264. Wan Z. N.; Mohammad H.; Zhao Y. Q.; Yu C.; Darling R. B.; Anantram M. P. Engineering of the Resistive Switching Properties in V2O5 Thin Film by Atomic Structural Transition: Experiment and Theory. J. Appl. Phys. 2018, 124, 105301. 10.1063/1.5045826. [DOI] [Google Scholar]
  265. Glynn C.; Creedon D.; Geaney H.; Armstrong E.; Collins T.; Morris M. A.; O’Dwyer C. Linking Precursor Alterations to Nanoscale Structure and Optical Transparency in Polymer Assisted Fast-Rate Dip-Coating of Vanadium Oxide Thin Films. Sci. Rep. 2015, 5, 11574. 10.1038/srep11574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  266. Liu Y. Q.; Chen Q. Q.; Du X. L.; Liu X. Q.; Li P. Effects of Substrate on the Structure and Properties of V2O5 Thin Films Prepared by the Sol-Gel Method. AIP Adv. 2019, 9, 045028. 10.1063/1.5095718. [DOI] [Google Scholar]
  267. Senapati S.; Panda S. Effect of Aging of V2O5 Sol on Properties of Nanoscale Films. Thin Solid Films 2016, 599, 42–48. 10.1016/j.tsf.2015.12.045. [DOI] [Google Scholar]
  268. Pradeeswari K.; Venkatesan A.; Pandi P.; Prasad K. G.; Karthik K.; Maiyalagan T.; Kumar R. M. Effect of Cerium on Electrochemical Properties of V2O5 Nanoparticles Synthesized via Non-Aqueous Sol-Gel Technique. Ionics 2020, 26, 905–912. 10.1007/s11581-019-03259-z. [DOI] [Google Scholar]
  269. Yu Z.; Zheng J.; Jing X.; Hang G.; Liu Q.; Cai W. Study on the Optical and Electrochemical Performance of V2O5 with Various Morphologies. J. Dispers. Sci. Technol. 2020, 41, 2203–2210. 10.1080/01932691.2019.1656085. [DOI] [Google Scholar]
  270. Bi W.; Wang J.; Jahrman E. P.; Seidler G. T.; Gao G.; Wu G.; Cao G. Interface Engineering V2O5 Nanofibers for High-Energy and Durable Supercapacitors. Small 2019, 15, 1901747. 10.1002/smll.201901747. [DOI] [PubMed] [Google Scholar]
  271. Zhang H.; Han X. R.; Gan R.; Guo Z. X.; Ni Y. H.; Zhang L. A Facile Biotemplate-Assisted Synthesis of Mesoporous V2O5 Microtubules for High Performance Asymmetric Supercapacitors. Appl. Surf. Sci. 2020, 511, 145527. 10.1016/j.apsusc.2020.145527. [DOI] [Google Scholar]
  272. Li Y.; Kuang J. L.; Lu Y.; Cao W. B. Facile Synthesis, Characterization of Flower-Like Vanadium Pentoxide Powders and Their Photocatalytic Behavior. Acta Metall. Sin. 2017, 30, 1017–1026. 10.1007/s40195-017-0611-6. [DOI] [Google Scholar]
  273. Wu Y. J.; Gao G. H.; Wu G. M. Self-Assembled Three-Dimensional Hierarchical Porous V2O5/Graphene Hybrid Aerogels for Supercapacitors with High Energy Density and Long Cycle Life. J. Mater. Chem. A 2015, 3, 1828–1832. 10.1039/C4TA05537C. [DOI] [Google Scholar]
  274. Pan J.; Li M.; Luo Y. Y.; Wu H.; Zhong L.; Wang Q.; Li G. H. Synthesis and SERS Activity of V2O5 Nanoparticles. Appl. Surf. Sci. 2015, 333, 34–38. 10.1016/j.apsusc.2015.01.242. [DOI] [Google Scholar]
  275. De Jesus L. R.; Horrocks G. A.; Liang Y.; Parija A.; Jaye C.; Wangoh L.; Wang J.; Fischer D. A.; Piper L. F. J.; Prendergast D.; Banerjee S. Mapping Polaronic States and Lithiation Gradients in Individual V2O5 Nanowires. Nat. Commun. 2016, 7, 12022. 10.1038/ncomms12022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Liang X.; Gao G. H.; Feng S. Z.; Du Y. C.; Wu G. M. Synthesis and Characterization of Carbon Supported V2O5 Nanotubes and Their Electrochemical Properties. J. Alloys Compd. 2019, 772, 429–437. 10.1016/j.jallcom.2018.09.077. [DOI] [Google Scholar]
  277. Li H.; Tian H. L.; Chang T. H.; Zhang J. Y.; Koh S. N.; Wang X. N.; Wang C. H.; Chen P. Y. High-Purity V2O5 Nanosheets Synthesized from Gasification Waste: Flexible Energy Storage Devices and Environmental Assessment. ACS Sustain. Chem. Eng. 2019, 7, 12474–12484. 10.1021/acssuschemeng.9b02066. [DOI] [Google Scholar]
  278. Yue Y.; Liang H. Micro- and Nano-Structured Vanadium Pentoxide (V2O5) for Electrodes of Lithium-Ion Batteries. Adv. Energy Mater. 2017, 7, 1602545. 10.1002/aenm.201602545. [DOI] [Google Scholar]
  279. Yan B.; Li X. F.; Bai Z. M.; Zhao Y.; Dong L.; Song X. S.; Li D. J.; Langford C.; Sun X. L. Crumpled Reduced Graphene Oxide Conformally Encapsulated Hollow V2O5 Nano/Microsphere Achieving Brilliant Lithium Storage Performance. Nano Energy 2016, 24, 32–44. 10.1016/j.nanoen.2016.04.002. [DOI] [Google Scholar]
  280. Li M.; Li D. B.; Pan J.; Lin J. C.; Li G. H. Selective Synthesis of Vanadium Oxides and Investigation of the Thermochromic Properties of VO2 by Infrared Spectroscopy. Eur. J. Inorg. Chem. 2013, 2013, 1207–1212. 10.1002/ejic.201201118. [DOI] [Google Scholar]
  281. Pan A.; Wu H. B.; Yu L.; Lou X. W. Template-Free Synthesis of VO2 Hollow Microspheres with Various Interiors and Their Conversion into V2O5 for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2013, 52, 2226–2230. 10.1002/anie.201209535. [DOI] [PubMed] [Google Scholar]
  282. Pan J.; Zhong L.; Li M.; Luo Y.; Li G. Microwave-Assisted Solvothermal Synthesis of VO2 Hollow Spheres and Their Conversion into V2O5 Hollow Spheres with Improved Lithium Storage Capability. Chem.—Eur. J. 2016, 22, 1461–1466. 10.1002/chem.201504259. [DOI] [PubMed] [Google Scholar]
  283. Liang S.; Hu Y.; Nie Z.; Huang H.; Chen T.; Pan A.; Cao G. Template-Free Synthesis of Ultra-Large V2O5 Nanosheets with Exceptional Small Thickness for High-Performance Lithium-Ion Batteries. Nano Energy 2015, 13, 58–66. 10.1016/j.nanoen.2015.01.049. [DOI] [Google Scholar]
  284. Pan A.; Wu H. B.; Yu L.; Zhu T.; Lou X. W. Synthesis of Hierarchical Three-Dimensional Vanadium Oxide Microstructures as High-Capacity Cathode Materials for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2012, 4, 3874–3879. 10.1021/am3012593. [DOI] [PubMed] [Google Scholar]
  285. Liu J.; Xue D. Cation-Induced Coiling of Vanadium Pentoxide Nanobelts. Nanoscale Res. Lett. 2010, 5, 1619. 10.1007/s11671-010-9685-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Le T. K.; Kang M.; Kim S. W. A Review on the Optical Characterization of V2O5 Micro-Nanostructures. Ceram. Int. 2019, 45, 15781–15798. 10.1016/j.ceramint.2019.05.339. [DOI] [Google Scholar]
  287. Vernardou D.; Marathianou I.; Katsarakis N.; Koudoumas E.; Kazadojev I. I.; O’Brien S.; Pemble M. E.; Povey I. M. Capacitive Behavior of Ag Doped V2O5 Grown by Aerosol Assisted Chemical Vapour Deposition. Electrochim. Acta 2016, 196, 294–299. 10.1016/j.electacta.2016.02.186. [DOI] [Google Scholar]
  288. Lee S.; Kim J.; Jeon J. H.; Song M.; Kim S.; You Y. G.; Jhang S. H.; Seo S. A.; Chun S. H. Chemical Vapor-Deposited Vanadium Pentoxide Nanosheets with Highly Stable and Low Switching Voltages for Effective Selector Devices. ACS Appl. Mater. Interfaces 2018, 10, 42875–42881. 10.1021/acsami.8b15686. [DOI] [PubMed] [Google Scholar]
  289. Wang Y; Su Q; Chen C H; Yu M L; Han G J; Wang G Q; Xin K; Lan W; Liu X Q Low Temperature Growth of Vanadium Pentoxide Nanomaterials by Chemical Vapour Deposition Using VO(acac)2 as Precursor. J. Phys. D 2010, 43, 185102. 10.1088/0022-3727/43/18/185102. [DOI] [Google Scholar]
  290. Raiford J. A.; Oyakhire S. T.; Bent S. F. Applications of Atomic Layer Deposition and Chemical Vapor Deposition for Perovskite Solar Cells. Energy Environ. Sci. 2020, 13, 1997–2023. 10.1039/D0EE00385A. [DOI] [Google Scholar]
  291. Chen X.; Zhu H.; Chen Y. C.; Shang Y.; Cao A.; Hu L.; Rubloff G. W. MWCNT/V2O5 Core/Shell Sponge for High Areal Capacity and Power Density Li-Ion Cathodes. ACS Nano 2012, 6, 7948–7955. 10.1021/nn302417x. [DOI] [PubMed] [Google Scholar]
  292. Mattelaer F.; Geryl K.; Rampelberg G.; Dendooven J.; Detavernier C. Amorphous and Crystalline Vanadium Oxides as High-Energy and High-Power Cathodes for Three-Dimensional Thin-Film Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 13121–13131. 10.1021/acsami.6b16473. [DOI] [PubMed] [Google Scholar]
  293. Chen X. Y.; Pomerantseva E.; Gregorczyk K.; Ghodssi R.; Rubloff G. Cathodic ALD V2O5 Thin Films for High-Rate Electrochemical Energy Storage. RSC Adv. 2013, 3, 4294. 10.1039/c3ra23031g. [DOI] [Google Scholar]
  294. Østreng E.; Nilsen O.; Fjellvåg H. Optical Properties of Vanadium Pentoxide Deposited by ALD. J. Phys. Chem. C 2012, 116, 19444–19450. 10.1021/jp304521k. [DOI] [Google Scholar]
  295. Rossnagel S. M. Thin Film Deposition with Physical Vapor Deposition and Related Technologies. J. Vac. Sci. Technol. A 2003, 21, S74–S87. 10.1116/1.1600450. [DOI] [Google Scholar]
  296. Lobe S.; Bauer A.; Uhlenbruck S.; Fattakhova-Rohlfing D. Physical Vapor Deposition in Solid-State Battery Development: From Materials to Devices. Adv. Sci. 2021, 8, 2002044. 10.1002/advs.202002044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  297. Xia F.; Yang L.; Dai B.; Yang Z. H.; Xu L. G.; Gao G.; Sun C. Q.; Song Z. C.; Ralchenko V.; Zhu J. Q. Thermal Transition Behaviors of Vanadium Pentoxide Film During Post-Deposition Annealing. Surf. Coat. Technol. 2021, 405, 126637. 10.1016/j.surfcoat.2020.126637. [DOI] [Google Scholar]
  298. Choi S. G.; Seok H. J.; Rhee S.; Hahm D.; Bae W. K.; Kim H. K. Magnetron-Sputtered Amorphous V2O5 Hole Injection Layer for High Performance Quantum Dot Light-Emitting Diode. J. Alloys Compd. 2021, 878, 160303. 10.1016/j.jallcom.2021.160303. [DOI] [Google Scholar]
  299. Vijay V. S.; Varghese R.; Sakunthala A.; Rajesh S.; Vidhya B. Highly Crystalline V2O5 and V6O13 Thin Films by PLD and a Study on Morphology Transition of V2O5 by Post Annealing. Vacuum 2021, 187, 110097. 10.1016/j.vacuum.2021.110097. [DOI] [Google Scholar]
  300. Huotari J.; Lappalainen J. Nanostructured Vanadium Pentoxide Gas Sensors for SCR Process Control. J. Mater. Sci. 2017, 52, 2241–2253. 10.1007/s10853-016-0517-0. [DOI] [Google Scholar]
  301. Han S. D.; Moon H. G.; Noh M. S.; Pyeon J. J.; Shim Y. S.; Nahm S.; Kim J. S.; Yoo K. S.; Kang C. Y. Self-Doped Nanocolumnar Vanadium Oxides Thin Films for Highly Selective NO2 Gas Sensing at Low Temperature. Sens. Actuator B 2017, 241, 40–47. 10.1016/j.snb.2016.10.029. [DOI] [Google Scholar]
  302. Thiagarajan S.; Thaiyan M.; Ganesan R. Physical Property Exploration of Highly Oriented V2O5 Thin Films Prepared by Electron Beam Evaporation. New J. Chem. 2015, 39, 9471–9479. 10.1039/C5NJ01582K. [DOI] [Google Scholar]
  303. Deniz A. R. The Temperature Dependence of Current-Voltage Characteristics of V2O5/p-Si Heterojunction Diode. J. Mater. Sci. Mater. Electron. 2021, 32, 18886–18899. 10.1007/s10854-021-06406-3. [DOI] [Google Scholar]
  304. Wang C. C.; Lu C. L.; Shieu F. S.; Shih H. C. Enhanced Photoluminescence Properties of Ga-Doped V2O5 Nanorods via Defect Structures. Chem. Phys. Lett. 2020, 738, 136864. 10.1016/j.cplett.2019.136864. [DOI] [Google Scholar]
  305. Uddin M. N.; Rahman M. S.; Shumi W.; Hossain M. K.; Ullah A. K. M. A. Characterization, Microbial and Catalytic Application of V2O5 Nanoparticles Prepared from Schiff Base Complexes as Precursor. J. Chem. Sci. 2020, 132, 131. 10.1007/s12039-020-01840-y. [DOI] [Google Scholar]
  306. Velmurugan R.; Premkumar J.; Pitchai R.; Ulaganathan M.; Subramanian B. Robust, Flexible, and Binder Free Highly Crystalline V2O5 Thin Film Electrodes and Their Superior Supercapacitor Performances. ACS Sustain. Chem. Eng. 2019, 7, 13115–13126. 10.1021/acssuschemeng.9b02302. [DOI] [Google Scholar]
  307. Berouaken M.; Talbi L.; Yaddadene C.; Maoudj M.; Menari H.; Alkama R.; Gabouze N. Room Temperature Ammonia Gas Sensor Based on V2O5 Nanoplatelets/Quartz Crystal Microbalance. Appl. Phys. A: Mater. Sci. Process. 2020, 126, 949. 10.1007/s00339-020-04129-6. [DOI] [Google Scholar]
  308. Zhang P.; Zhao L.; An Q.; Wei Q.; Zhou L.; Wei X.; Sheng J.; Mai L. A High-Rate V2O5 Hollow Microclew Cathode for an All-Vanadium-based Lithium-Ion Full Cell. Small 2016, 12, 1082–1090. 10.1002/smll.201503214. [DOI] [PubMed] [Google Scholar]
  309. Ali G.; Lee J. H.; Oh S. H.; Cho B. W.; Nam K.-W.; Chung K. Y. Investigation of the Na Intercalation Mechanism into Nanosized V2O5/C Composite Cathode Material for Na-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 6032–6039. 10.1021/acsami.5b11954. [DOI] [PubMed] [Google Scholar]
  310. Yan M.; He P.; Chen Y.; Wang S.; Wei Q.; Zhao K.; Xu X.; An Q.; Shuang Y.; Shao Y.; et al. Water-Lubricated Intercalation in V2O5·nH2O for High-Capacity and High-Rate Aqueous Rechargeable Zinc Batteries. Adv. Mater. 2018, 30, 1703725. 10.1002/adma.201703725. [DOI] [PubMed] [Google Scholar]
  311. Palanisamy K.; Um J. H.; Jeong M.; Yoon W.-S. Porous V2O5/RGO/CNT Hierarchical Architecture as a Cathode Material: Emphasis on the Contribution of Surface Lithium Storage. Sci. Rep. 2016, 6, 31275. 10.1038/srep31275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  312. He P.; Zhang G.; Liao X.; Yan M.; Xu X.; An Q.; Liu J.; Mai L. Sodium Ion Stabilized Vanadium Oxide Nanowire Cathode for High-Performance Zinc-Ion Batteries. Adv. Energy Mater. 2018, 8, 1702463. 10.1002/aenm.201702463. [DOI] [Google Scholar]
  313. Zeng H.; Liu D.; Zhang Y.; See K. A.; Jun Y.-S.; Wu G.; Gerbec J. A.; Ji X.; Stucky G. D. Nanostructured Mn-Doped V2O5 Cathode Material Fabricated from Layered Vanadium Jarosite. Chem. Mater. 2015, 27, 7331–7336. 10.1021/acs.chemmater.5b02840. [DOI] [Google Scholar]
  314. Chao D.; Xia X.; Liu J.; Fan Z.; Ng C. F.; Lin J.; Zhang H.; Shen Z. X.; Fan H. J. A V2O5/Conductive-Polymer Core/Shell Nanobelt Array on Three-Dimensional Graphite Foam: A High-Rate, Ultrastable, and Freestanding Cathode for Lithium-Ion Batteries. Adv. Mater. 2014, 26, 5794–5800. 10.1002/adma.201400719. [DOI] [PubMed] [Google Scholar]
  315. Palani N. S.; Kavitha N. S.; Venkatesh K. S.; Kumar K. A.; Senthilkumar M.; Pandurangan A.; Ilangovan R. The Synergistic Effect of the RuO2 Nanoparticle-Decorated V2O5 Heterostructure for High-Performance Asymmetric Supercapacitors. New J. Chem. 2021, 45, 14598–14607. 10.1039/D1NJ00011J. [DOI] [Google Scholar]
  316. Sahu V.; Goel S.; Tomar A. K.; Singh G.; Sharma R. K. Graphene Nanoribbons @ Vanadium Oxide Nanostrips for Supercapacitive Energy Storage. Electrochim. Acta 2017, 230, 255–264. 10.1016/j.electacta.2017.01.188. [DOI] [Google Scholar]
  317. Sun W.; Gao G. H.; Zhang K.; Liu Y. D.; Wu G. M. Self-Assembled 3D N-CNFs/V2O5 Aerogels with Core/Shell Nanostructures through Vacancies Control and Seeds Growth as an Outstanding Supercapacitor Electrode Material. Carbon 2018, 132, 667–677. 10.1016/j.carbon.2018.03.004. [DOI] [Google Scholar]
  318. Zhu C. X.; Hu D.; Liu Z. Interconnected Three-Dimensionally Hierarchical Heterostructures with Homogeneously-Dispersed V2O5 Nanocrystals and Carbon for High Performance Supercapacitor Electrodes. Electrochim. Acta 2017, 229, 155–165. 10.1016/j.electacta.2017.01.144. [DOI] [Google Scholar]
  319. Yao L.; Zhang C.; Hu N.; Zhang L.; Zhou Z.; Zhang Y. Three-Dimensional Skeleton Networks of Reduced Graphene Oxide Nanosheets/Vanadium Pentoxide Nanobelts Hybrid for High-Performance Supercapacitors. Electrochim. Acta 2019, 295, 14–21. 10.1016/j.electacta.2018.10.134. [DOI] [Google Scholar]
  320. Shen F.-C.; Wang Y.; Tang Y.-J.; Li S.-L.; Wang Y.-R.; Dong L.-Z.; Li Y.-F.; Xu Y.; Lan Y.-Q. CoV2O6-V2O5 Coupled with Porous N-Doped Reduced Graphene Oxide Composite as a Highly Efficient Electrocatalyst for Oxygen Evolution. ACS Energy Lett. 2017, 2, 1327–1333. 10.1021/acsenergylett.7b00229. [DOI] [Google Scholar]
  321. Ling T.; Yan D.-Y.; Wang H.; Jiao Y.; Hu Z.; Zheng Y.; Zheng L.; Mao J.; Liu H.; Du X.-W.; Jaroniec M.; Qiao S.-Z. Activating Cobalt(II) Oxide Nanorods for Efficient Electrocatalysis by Strain Engineering. Nat. Commun. 2017, 8, 1509. 10.1038/s41467-017-01872-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  322. Gong M.; Wang D.-Y.; Chen C.-C.; Hwang B.-J.; Dai H. A Mini Review on Nickel-Based Electrocatalysts for Alkaline Hydrogen Evolution Reaction. Nano Res. 2016, 9, 28–46. 10.1007/s12274-015-0965-x. [DOI] [Google Scholar]
  323. Hu P.; Long G.; Chaturvedi A.; Wang S.; Tan K.; He Y.; Zheng L.; Liu G.; Ke Y.; Zhou Y.; et al. Agent-Assisted VSSe Ternary Alloy Single Crystals as an Efficient Stable Electrocatalyst for the Hydrogen Evolution Reaction. J. Mater. Chem. A 2019, 7, 15714–15721. 10.1039/C9TA04287C. [DOI] [Google Scholar]
  324. Gong Q.; Cheng L.; Liu C.; Zhang M.; Feng Q.; Ye H.; Zeng M.; Xie L.; Liu Z.; Li Y. Ultrathin MoS2(1-x)Se2x Alloy Nanoflakes For Electrocatalytic Hydrogen Evolution Reaction. ACS Catal. 2015, 5, 2213–2219. 10.1021/cs501970w. [DOI] [Google Scholar]
  325. Tan C.; Luo Z.; Chaturvedi A.; Cai Y.; Du Y.; Gong Y.; Huang Y.; Lai Z.; Zhang X.; Zheng L.; et al. Preparation of High-Percentage 1T-Phase Transition Metal Dichalcogenide Nanodots for Electrochemical Hydrogen Evolution. Adv. Mater. 2018, 30, 1705509. 10.1002/adma.201705509. [DOI] [PubMed] [Google Scholar]
  326. Li Y. H.; Liu P. F.; Pan L. F.; Wang H. F.; Yang Z. Z.; Zheng L. R.; Hu P.; Zhao H. J.; Gu L.; Yang H. G. Local Atomic Structure Modulations Activate Metal Oxide as Electrocatalyst for Hydrogen Evolution in Acidic Water. Nat. Commun. 2015, 6, 8064. 10.1038/ncomms9064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  327. Bi W.; Jahrman E.; Seidler G.; Wang J.; Gao G.; Wu G.; Atif M.; AlSalhi M.; Cao G. Tailoring Energy and Power Density through Controlling the Concentration of Oxygen Vacancies in V2O5/PEDOT Nanocable-Based Supercapacitors. ACS Appl. Mater. Interfaces 2019, 11, 16647–16655. 10.1021/acsami.9b03830. [DOI] [PubMed] [Google Scholar]
  328. Zhang J.; Zhang H.; Liu M.; Xu Q.; Jiang H.; Li C. Cobalt-Stabilized Oxygen Vacancy of V2O5 Nanosheet Arrays with Delocalized Valence Electron for Alkaline Water Splitting. Chem. Eng. Sci. 2020, 227, 115915. 10.1016/j.ces.2020.115915. [DOI] [Google Scholar]
  329. Meena A.; Ha M.; Chandrasekaran S. S.; Sultan S.; Thangavel P.; Harzandi A. M.; Singh B.; Tiwari J. N.; Kim K. S. Pt-Like Hydrogen Evolution on a V2O5/Ni(OH)2 Electrocatalyst. J. Mater. Chem. A 2019, 7, 15794–15800. 10.1039/C9TA03627J. [DOI] [Google Scholar]
  330. Zhong X.; Zhang L.; Tang J.; Chai J.; Xu J.; Cao L.; Yang M.; Yang M.; Kong W.; Wang S.; et al. Efficient Coupling of a Hierarchical V2O5@Ni3S2 Hybrid Nanoarray for Pseudocapacitors and Hydrogen Production. J. Mater. Chem. A 2017, 5, 17954–17962. 10.1039/C7TA04755J. [DOI] [Google Scholar]
  331. Zou X.; Li Z.; Xie Y.; Wu H.; Lin S. Phosphorus-Doping and Addition of V2O5 into Pt/Graphene Resulting in Highly-Enhanced Electro-Photo Synergistic Catalysis for Oxygen Reduction and Hydrogen Evolution Reactions. Int. J. Hydrog. Energy 2020, 45, 30647–30658. 10.1016/j.ijhydene.2020.09.165. [DOI] [Google Scholar]
  332. Su J.; Zou X.-X.; Li G.-D.; Wei X.; Yan C.; Wang Y.-N.; Zhao J.; Zhou L.-J.; Chen J.-S. Macroporous V2O5-BiVO4 Composites: Effect of Heterojunction on the Behavior of Photogenerated Charges. J. Phys. Chem. C 2011, 115, 8064–8071. 10.1021/jp200274k. [DOI] [Google Scholar]
  333. Akbarzadeh R.; Umbarkar S. B.; Sonawane R. S.; Takle S.; Dongare M. K. Vanadia-Titania Thin Films for Photocatalytic Degradation of Formaldehyde in Sunlight. Appl. Catal., A 2010, 374, 103–109. 10.1016/j.apcata.2009.11.035. [DOI] [Google Scholar]
  334. Arunachalam M.; Ahn K.-S.; Kang S. H. Oxygen Evolution NiOOH Catalyst Assisted V2O5@BiVO4 Inverse Opal Hetero-Structure for Solar Water Oxidation. Int. J. Hydrog. Energy 2019, 44, 4656–4663. 10.1016/j.ijhydene.2019.01.024. [DOI] [Google Scholar]
  335. Gomez-Solis C.; Oliva J.; Puentes-Prado E.; Badillo F.; Garcia C. R. Effect of Morphology on the Hydrogen Production of V2O5 Nano/Micro-Particles Synthesized by a Biodegradable Template. J. Phys. Chem. Solids 2021, 152, 109977. 10.1016/j.jpcs.2021.109977. [DOI] [Google Scholar]
  336. Vattikuti S. V. P.; Reddy P. A. K.; Shim J.; Byon C. Synthesis of Vanadium-Pentoxide-Supported Graphitic Carbon Nitride Heterostructure and Studied Their Hydrogen Evolution Activity Under Solar Light. J. Mater. Sci. Mater. Electron. 2018, 29, 18760–18770. 10.1007/s10854-018-0001-5. [DOI] [Google Scholar]
  337. Vattikuti S. V. P.; Nam N. D.; Shim J. Graphitic Carbon Nitride/Na2Ti3O7/V2O5 Nanocomposite as a Visible Light Active Photocatalyst. Ceram. Int. 2020, 46, 18287–18296. 10.1016/j.ceramint.2020.05.045. [DOI] [Google Scholar]
  338. Sadeghzadeh-Attar A. Enhanced Photocatalytic Hydrogen Evolution by Novel Nb-Doped SnO2/V2O5 Heteronanostructures Under Visible Light with Simultaneous Basic Red 46 Dye Degradation. J. Asian Ceram. Soc. 2020, 8, 662–676. 10.1080/21870764.2020.1773621. [DOI] [Google Scholar]
  339. Akimov A. V.; Neukirch A. J.; Prezhdo O. V. Theoretical Insights into Photoinduced Charge Transfer and Catalysis at Oxide Interfaces. Chem. Rev. 2013, 113, 4496–4565. 10.1021/cr3004899. [DOI] [PubMed] [Google Scholar]
  340. Chen X.; Mao S. S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. 10.1021/cr0500535. [DOI] [PubMed] [Google Scholar]
  341. Fei H.-L.; Zhou H.-J.; Wang J.-G.; Sun P.-C.; Ding D.-T.; Chen T.-H. Synthesis of Hollow V2O5 Microspheres and Application to Photocatalysis. Solid State Sci. 2008, 10, 1276–1284. 10.1016/j.solidstatesciences.2007.12.026. [DOI] [Google Scholar]
  342. Shanmugam M.; Alsalme A.; Alghamdi A.; Jayavel R. Enhanced Photocatalytic Performance of the Graphene-V2O5 Nanocomposite in the Degradation of Methylene Blue Dye under Direct Sunlight. ACS Appl. Mater. Interfaces 2015, 7, 14905–14911. 10.1021/acsami.5b02715. [DOI] [PubMed] [Google Scholar]
  343. Beaula Ruby Kamalam M.; Inbanathan S. S. R.; Sethuraman K.; Umar A.; Algadi H.; Ibrahim A. A.; Rahman Q. I.; Garoufalis C. S.; Baskoutas S. Direct Sunlight-Driven Enhanced Photocatalytic Performance of V2O5 Nanorods/ Graphene Oxide Nanocomposites for the Degradation of Victoria Blue Dye. Environ. Res. 2021, 199, 111369. 10.1016/j.envres.2021.111369. [DOI] [PubMed] [Google Scholar]
  344. Hong Y.; Jiang Y.; Li C.; Fan W.; Yan X.; Yan M.; Shi W. In-Situ Synthesis of Direct Solid-State Z-Scheme V2O5/g-C3N4 Heterojunctions with Enhanced Visible Light Efficiency in Photocatalytic Degradation of Pollutants. Appl. Catal., B 2016, 180, 663–673. 10.1016/j.apcatb.2015.06.057. [DOI] [Google Scholar]
  345. Jiang H.; Nagai M.; Kobayashi K. Enhanced Photocatalytic Activity for Degradation of Methylene Blue over V2O5/BiVO4 Composite. J. Alloys Compd. 2009, 479, 821–827. 10.1016/j.jallcom.2009.01.051. [DOI] [Google Scholar]
  346. Saravanan R.; Joicy S.; Gupta V. K.; Narayanan V.; Stephen A. Visible Light Induced Degradation of Methylene Blue Using CeO2/V2O5 and CeO2/CuO Catalysts. Mater. Sci. Eng., C 2013, 33, 4725–4731. 10.1016/j.msec.2013.07.034. [DOI] [PubMed] [Google Scholar]
  347. Liu J.; Yang R.; Li S. Preparation and Characterization of the TiO2-V2O5 Photocatalyst with Visible-Light Activity. Rare Metals 2006, 25, 636–642. 10.1016/S1001-0521(07)60005-9. [DOI] [Google Scholar]
  348. Wang Y.; Su Y. R.; Qiao L.; Liu L. X.; Su Q.; Zhu C. Q.; Liu X. Q. Synthesis of One-Dimensional TiO2/V2O5 Branched Heterostructures and Their Visible Light Photocatalytic Activity towards Rhodamine B. Nanotechnology 2011, 22, 225702. 10.1088/0957-4484/22/22/225702. [DOI] [PubMed] [Google Scholar]
  349. Sun J.; Li X.; Zhao Q.; Ke J.; Zhang D. Novel V2O5/BiVO4/TiO2 Nanocomposites with High Visible-Light-Induced Photocatalytic Activity for the Degradation of Toluene. J. Phys. Chem. C 2014, 118, 10113–10121. 10.1021/jp5013076. [DOI] [Google Scholar]
  350. Colton R. J.; Guzman A. M.; Rabalais J. W. Photochromism and Electrochromism in Amorphous Transition Metal Oxide Films. Acc. Chem. Res. 1978, 11, 170–176. 10.1021/ar50124a008. [DOI] [Google Scholar]
  351. Colton R. J.; Guzman A. M.; Rabalais J. W. Electrochromism in Some Thin-Film Transition-Metal Oxides Characterized by X-Ray Electron Spectroscopy. J. Appl. Phys. 1978, 49, 409–416. 10.1063/1.324349. [DOI] [Google Scholar]
  352. Wang Z.; Wang X.; Cong S.; Geng F.; Zhao Z. Fusing Electrochromic Technology with Other Advanced Technologies: A New Roadmap for Future Development. Mater. Sci. Eng. R 2020, 140, 100524. 10.1016/j.mser.2019.100524. [DOI] [Google Scholar]
  353. Mjejri I.; Gaudon M.; Song G.; Labrugère C.; Rougier A. Crystallized V2O5 as Oxidized Phase for Unexpected Multicolor Electrochromism in V2O3 Thick Film. ACS Appl. Energy Mater. 2018, 1, 2721–2729. 10.1021/acsaem.8b00386. [DOI] [Google Scholar]
  354. Mjejri I.; Rougier A.; Gaudon M. Low-Cost and Facile Synthesis of the Vanadium Oxides V2O3, VO2, and V2O5 and Their Magnetic, Thermochromic and Electrochromic Properties. Inorg. Chem. 2017, 56, 1734–1741. 10.1021/acs.inorgchem.6b02880. [DOI] [PubMed] [Google Scholar]
  355. Mjejri I.; Manceriu L. M.; Gaudon M.; Rougier A.; Sediri F. Nano-vanadium pentoxide films for electrochromic displays. Solid State Ion. 2016, 292, 8–14. 10.1016/j.ssi.2016.04.023. [DOI] [Google Scholar]
  356. Mjejri I.; Gaudon M.; Rougier A. Mo addition for improved electrochromic properties of V2O5 thick films. Sol. Energy Mater. Sol. Cells 2019, 198, 19–25. 10.1016/j.solmat.2019.04.010. [DOI] [Google Scholar]
  357. Salek G.; Bellanger B.; Mjejri I.; Gaudon M.; Rougier A. Polyol Synthesis of Ti-V2O5 Nanoparticles and Their Use as Electrochromic Films. Inorg. Chem. 2016, 55, 9838–9847. 10.1021/acs.inorgchem.6b01662. [DOI] [PubMed] [Google Scholar]
  358. Mjejri I.; Duttine M.; Buffière S.; Labrugère-Sarroste C.; Rougier A. From the Irreversible Transformation of VO2 to V2O5 Electrochromic Films. Inorg. Chem. 2022, 61, 18496–18503. 10.1021/acs.inorgchem.2c02722. [DOI] [PubMed] [Google Scholar]
  359. Le T. K.; Pham P. V.; Dong C.-L.; Bahlawane N.; Vernardou D.; Mjejri I.; Rougier A.; Kim S. W. Recent Advances in Vanadium Pentoxide (V2O5) Towards Related Applications in Chromogenics and Beyond: Fundamentals, Progress, and Perspectives. J. Mater. Chem. C 2022, 10, 4019–4071. 10.1039/D1TC04872D. [DOI] [Google Scholar]
  360. Surca A. K.; Dražić G.; Mihelčič M. Spectroelectrochemistry in the Investigation of Sol-Gel Electrochromic V2O5 Films. J. Sol-Gel Sci. Technol. 2020, 95, 587–598. 10.1007/s10971-020-05337-5. [DOI] [Google Scholar]
  361. Talledo A.; Granqvist C. G. Electrochromic Vanadium-Pentoxide-Based Films: Structural, Electrochemical, and Optical Properties. J. Appl. Phys. 1995, 77, 4655–4666. 10.1063/1.359433. [DOI] [Google Scholar]
  362. Prasad A. K.; Park J.-Y.; Kang S.-H.; Ahn K.-S. Electrochemically Co-Deposited WO3-V2O5 Composites for Electrochromic Energy Storage Applications. Electrochim. Acta 2022, 422, 140340. 10.1016/j.electacta.2022.140340. [DOI] [Google Scholar]
  363. Panagopoulou M.; Vernardou D.; Koudoumas E.; Katsarakis N.; Tsoukalas D.; Raptis Y. S. Tunable Properties of Mg-Doped V2O5 Thin Films for Energy Applications: Li-Ion Batteries and Electrochromics. J. Phys. Chem. C 2017, 121, 70–79. 10.1021/acs.jpcc.6b09018. [DOI] [Google Scholar]
  364. Qi Y.; Qin K.; Zou Y.; Lin L.; Jian Z.; Chen W. Flexible Electrochromic Thin Films with Ultrafast Responsion Based on Exfoliated V2O5 Nanosheets/Graphene Oxide via Layer-by-Layer Assembly. Appl. Surf. Sci. 2020, 514, 145950. 10.1016/j.apsusc.2020.145950. [DOI] [Google Scholar]
  365. Tong Z.; Zhang X.; Lv H.; Li N.; Qu H.; Zhao J.; Li Y.; Liu X.-Y. From Amorphous Macroporous Film to 3D Crystalline Nanorod Architecture: A New Approach to Obtain High-Performance V2O5 Electrochromism. Adv. Mater. Interfaces 2015, 2, 1500230. 10.1002/admi.201500230. [DOI] [Google Scholar]
  366. Kim J.; Lee K. H.; Lee S.; Park S.; Chen H.; Kim S. K.; Yim S.; Song W.; Lee S. S.; Yoon D. H.; et al. Minimized Optical Scattering of MXene-Derived 2D V2O5 Nanosheet-Based Electrochromic Device with High Multicolor Contrast and Accuracy. Chem. Eng. J. 2023, 453, 139973. 10.1016/j.cej.2022.139973. [DOI] [Google Scholar]
  367. Calatayud M.; Andrés J.; Beltrán A. A Systematic Density Functional Theory Study of VxOy+ and VxOY (X = 2–4, Y = 2–10) Systems. J. Phys. Chem. A 2001, 105, 9760–9775. 10.1021/jp011535x. [DOI] [Google Scholar]
  368. Hübner O.; Himmel H.-J. Multiple Metal-Metal Bond or No Bond? The Electronic Structure of V2O2. Angew. Chem., Int. Ed. 2017, 56, 12340–12343. 10.1002/anie.201706266. [DOI] [PubMed] [Google Scholar]
  369. Jakubikova E.; Rappé A. K.; Bernstein E. R. Density Functional Theory Study of Small Vanadium Oxide Clusters. J. Phys. Chem. A 2007, 111, 12938–12943. 10.1021/jp0745844. [DOI] [PubMed] [Google Scholar]
  370. Wang Y.; Gong X.; Wang J. Comparative DFT Study of Structure and Magnetism of TMnOm (TM = Sc-Mn, n = 1–2, m = 1–6) Clusters. Phys. Chem. Chem. Phys. 2010, 12, 2471–2477. 10.1039/b920033a. [DOI] [PubMed] [Google Scholar]
  371. Yamazaki S.; Li C.; Ohoyama K.; Nishi M.; Ichihara M.; Ueda H.; Ueda Y. Synthesis, Structure and Magnetic Properties of V4O9—A Missing Link in Binary Vanadium Oxides. J. Solid State Chem. 2010, 183, 1496–1503. 10.1016/j.jssc.2010.04.007. [DOI] [Google Scholar]
  372. Kawashima K.; Kosuge K.; Kachi S. Isothermal Reduction of V2O5 by SO2 Gas. Chem. Lett. 1975, 4, 1131–1136. 10.1246/cl.1975.1131. [DOI] [Google Scholar]
  373. Pang H.; Dong Y.; Ting S. L.; Lu J.; Li C. M.; Kim D.-H.; Chen P. 2D Single- or Double-Layered Vanadium Oxide Nanosheet Assembled 3D Microflowers: Controlled Synthesis, Growth Mechanism, and Applications. Nanoscale 2013, 5, 7790–7794. 10.1039/c3nr02651e. [DOI] [PubMed] [Google Scholar]
  374. Chine M. K.; Sediri F.; Gharbi N. Solvothermal Synthesis of V4O9 Flake-Like Morphology and Its Photocatalytic Application in the Degradation of Methylene Blue. Mater. Res. Bull. 2012, 47, 3422–3426. 10.1016/j.materresbull.2012.07.016. [DOI] [Google Scholar]
  375. Wang Q.; Sun T.; Zheng S.; Li L.; Ma T.; Liang J. A new Tunnel-Type V4O9 Cathode for High Power Density Aqueous Zinc Ion Batteries. Inorg. Chem. Front. 2021, 8, 4497–4506. 10.1039/D1QI00747E. [DOI] [Google Scholar]
  376. Chen X.; Wang L.; Li H.; Cheng F.; Chen J. Porous V2O5 Nanofibers as Cathode Materials for Rechargeable Aqueous Zinc-Ion Batteries. J. Energy Chem. 2019, 38, 20–25. 10.1016/j.jechem.2018.12.023. [DOI] [Google Scholar]
  377. Shan L.; Zhou J.; Zhang W.; Xia C.; Guo S.; Ma X.; Fang G.; Wu X.; Liang S. Highly Reversible Phase Transition Endows V6O13 with Enhanced Performance as Aqueous Zinc-Ion Battery Cathode. Energy Technol. 2019, 7, 1900022. 10.1002/ente.201900022. [DOI] [Google Scholar]
  378. Alfaruqi M. H.; Mathew V.; Song J.; Kim S.; Islam S.; Pham D. T.; Jo J.; Kim S.; Baboo J. P.; Xiu Z.; et al. Electrochemical Zinc Intercalation in Lithium Vanadium Oxide: A High-Capacity Zinc-Ion Battery Cathode. Chem. Mater. 2017, 29, 1684–1694. 10.1021/acs.chemmater.6b05092. [DOI] [Google Scholar]
  379. Li G.; Yang Z.; Jiang Y.; Jin C.; Huang W.; Ding X.; Huang Y. Towards Polyvalent Ion Batteries: A Zinc-Ion Battery Based on NASICON Structured Na3V2(PO4)3. Nano Energy 2016, 25, 211–217. 10.1016/j.nanoen.2016.04.051. [DOI] [Google Scholar]
  380. Wei T.; Li Q.; Yang G.; Wang C. An Electrochemically Induced Bilayered Structure Facilitates Long-Life Zinc Storage of Vanadium Dioxide. J. Mater. Chem. A 2018, 6, 8006–8012. 10.1039/C8TA02090F. [DOI] [Google Scholar]
  381. He P.; Yan M.; Zhang G.; Sun R.; Chen L.; An Q.; Mai L. Layered VS2 Nanosheet-Based Aqueous Zn Ion Battery Cathode. Adv. Energy Mater. 2017, 7, 1601920. 10.1002/aenm.201601920. [DOI] [Google Scholar]
  382. Xue K. H.; Yang H.; Zhou Y. M.; Li G.; Skotheim T. A.; Lee H. S.; Yang X. Q.; McBreen J. A Study of the Zn/V6 O13 Secondary Battery. J. Electrochem. Soc. 1993, 140, 3413–3417. 10.1149/1.2221104. [DOI] [Google Scholar]
  383. Ding Y.-L.; Wen Y.; Wu C.; van Aken P. A.; Maier J.; Yu Y. 3D V6O13 Nanotextiles Assembled From Interconnected Nanogrooves as Cathode Materials for High-Energy Lithium Ion Batteries. Nano Lett. 2015, 15, 1388–1394. 10.1021/nl504705z. [DOI] [PubMed] [Google Scholar]
  384. Hu J.; Chen H.; Xiang K.; Xiao L.; Chen W.; Liao H.; Chen H. Preparation for V6O13@hollow Carbon Microspheres and Their Remarkable Electrochemical Performance for Aqueous Zinc-Ion Batteries. J. Alloys Compd. 2021, 856, 157085. 10.1016/j.jallcom.2020.157085. [DOI] [Google Scholar]
  385. He P.; Liu J.; Zhao X.; Ding Z.; Gao P.; Fan L.-Z. A Three-Dimensional Interconnected V6O13 Nest with a V5+-Rich State for Ultrahigh Zn Ion Storage. J. Mater. Chem. A 2020, 8, 10370–10376. 10.1039/D0TA03165H. [DOI] [Google Scholar]
  386. Wu M.; Zhu K.; Liang P.; Yao Z.; Shi F.; Zhang J.; Yan K.; Liu J.; Wang J. Uniform Rotate Hydrothermal Synthesis of V6O13 Nanosheets as Cathode Material for Lithium-Ion Battery. J. Alloys Compd. 2021, 877, 160174. 10.1016/j.jallcom.2021.160174. [DOI] [Google Scholar]
  387. Wu X.; Zou Z.; Li S.; Zhang Y. Solvothermal Preparation of Ga-Doped V6O13 Nanowires as Cathode Materials for Lithium-Ion Batteries. Ionics 2019, 25, 4557–4565. 10.1007/s11581-019-03039-9. [DOI] [Google Scholar]
  388. Zhang S.; Zou Z.; Zhang H.; Liu J.; Zhong S. Al/Ga Co-Doped V6O13 Nanorods with High Discharge Specific Capacity as Cathode Materials for Lithium-Ion Batteries. J. Electroanal. Chem. 2021, 890, 115220. 10.1016/j.jelechem.2021.115220. [DOI] [Google Scholar]
  389. Wang Z.; Zhang Y.; Xiao S.; Zhai H.; Zhu Y.; Cao C. Microwave-Assisted Synthesis of Metallic V6O13 Nanosheet as High-Capacity Cathode for Magnesium Storage. Mater. Lett. 2022, 308, 131279. 10.1016/j.matlet.2021.131279. [DOI] [Google Scholar]
  390. Li S.; Zou Z.; Wu X.; Zhang Y. Solvothermal Preparation of Carbon Coated V6O13 Nanocomposite as Cathode Material for Lithium-Ion Battery. J. Electroanal. Chem. 2019, 846, 113173. 10.1016/j.jelechem.2019.05.055. [DOI] [Google Scholar]
  391. Tian X.; Xu X.; He L.; Wei Q.; Yan M.; Xu L.; Zhao Y.; Yang C.; Mai L. Ultrathin Pre-lithiated V6O13 Nanosheet Cathodes with Enhanced Electrical Transport and Cyclability. J. Power Sources 2014, 255, 235–241. 10.1016/j.jpowsour.2014.01.017. [DOI] [Google Scholar]
  392. Shin J.; Choi D. S.; Lee H. J.; Jung Y.; Choi J. W. Hydrated Intercalation for High-Performance Aqueous Zinc Ion Batteries. Adv. Energy Mater. 2019, 9, 1900083. 10.1002/aenm.201900083. [DOI] [Google Scholar]
  393. Lai J.; Zhu H.; Zhu X.; Koritala H.; Wang Y. Interlayer-Expanded V6O13·nH2O Architecture Constructed for an Advanced Rechargeable Aqueous Zinc-Ion Battery. ACS Appl. Energy Mater. 2019, 2, 1988–1996. 10.1021/acsaem.8b02054. [DOI] [Google Scholar]
  394. Chen Q.; Luo Z.; Zhao X. K-Ion Intercalated V6O13 with Advanced High-Rate Long-Cycle Performance as Cathode for Zn-Ion Batteries. J. Mater. Chem. C 2022, 10, 590–597. 10.1039/D1TC04822H. [DOI] [Google Scholar]
  395. Yang G.; Wang C. Platinum-Induced Pseudo-Zn-Air Reaction Massively Increases the Electrochemical Capacity of Aqueous Zn/V5O12·6H2O Batteries. Energy Environ. Mater. 2021, 4, 596–602. 10.1002/eem2.12141. [DOI] [Google Scholar]
  396. Dai Y.; Liao X.; Yu R.; Li J.; Li J.; Tan S.; He P.; An Q.; Wei Q.; Chen L.; et al. Quicker and More Zn2+ Storage Predominantly from the Interface. Adv. Mater. 2021, 33, 2100359. 10.1002/adma.202100359. [DOI] [PubMed] [Google Scholar]
  397. Li Y.; Tan X.; Hocking R. K.; Bo X.; Ren H.; Johannessen B.; Smith S. C.; Zhao C. Implanting Ni-O-VOx Sites into Cu-Doped Ni for Low-Overpotential Alkaline Hydrogen Evolution. Nat. Commun. 2020, 11, 2720. 10.1038/s41467-020-16554-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  398. Chen M.; Liu J.; Kitiphatpiboon N.; Li X.; Wang J.; Hao X.; Abudula A.; Ma Y.; Guan G. Zn-VOx-Co Nanosheets with Amorphous/Crystalline Heterostructure for Highly Efficient Hydrogen Evolution Reaction. Chem. Eng. J. 2022, 432, 134329. 10.1016/j.cej.2021.134329. [DOI] [Google Scholar]
  399. Chai Y.-M.; Zhang X.-Y.; Lin J.-H.; Qin J.-F.; Liu Z.-Z.; Xie J.-Y.; Guo B.-Y.; Yang Z.; Dong B. Three-Dimensional VOx/NiS/NF Nanosheets as Efficient Electrocatalyst for Oxygen Evolution Reaction. Int. J. Hydrog. Energy 2019, 44, 10156–10162. 10.1016/j.ijhydene.2019.02.242. [DOI] [Google Scholar]
  400. Zhu Z.; Xu K.; Guo W.; Zhang H.; Xiao X.; He M.; Yu T.; Zhao H.; Zhang D.; Yang T. Vanadium-Phosphorus Incorporation Induced Interfacial Modification on Cobalt Catalyst and Its Super Electrocatalysis for Water Splitting in Alkaline Media. Appl. Catal., B 2022, 304, 120985. 10.1016/j.apcatb.2021.120985. [DOI] [Google Scholar]
  401. Wang X.; Zhang Z.; Huang M.; Feng J.; Xiong S.; Xi B. In Situ Electrochemically Activated Vanadium Oxide Cathode for Advanced Aqueous Zn-Ion Batteries. Nano Lett. 2022, 22, 119–127. 10.1021/acs.nanolett.1c03409. [DOI] [PubMed] [Google Scholar]
  402. Luo T.; Liu Y.; Su H.; Xiao R.; Huang L.; Xiang Q.; Zhou Y.; Chen C. Nanostructured-VO2(B): A High-Capacity Magnesium-Ion Cathode and Its Electrochemical Reaction Mechanism. Electrochim. Acta 2018, 260, 805–813. 10.1016/j.electacta.2017.12.042. [DOI] [Google Scholar]
  403. Balogun M.-S.; Luo Y.; Lyu F.; Wang F.; Yang H.; Li H.; Liang C.; Huang M.; Huang Y.; Tong Y. Carbon Quantum Dot Surface-Engineered VO2 Interwoven Nanowires: A Flexible Cathode Material for Lithium and Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 9733–9744. 10.1021/acsami.6b01305. [DOI] [PubMed] [Google Scholar]
  404. Yang S.; Gong Y.; Liu Z.; Zhan L.; Hashim D. P.; Ma L.; Vajtai R.; Ajayan P. M. Bottom-up Approach toward Single-Crystalline VO2-Graphene Ribbons as Cathodes for Ultrafast Lithium Storage. Nano Lett. 2013, 13, 1596–1601. 10.1021/nl400001u. [DOI] [PubMed] [Google Scholar]
  405. Ren G.; Hoque M. N. F.; Pan X.; Warzywoda J.; Fan Z. Vertically Aligned VO2(B) Nanobelt Forest and Its Three-Dimensional Structure on Oriented Graphene for Energy Storage. J. Mater. Chem. A 2015, 3, 10787–10794. 10.1039/C5TA01900A. [DOI] [Google Scholar]
  406. Ding J.; Du Z.; Gu L.; Li B.; Wang L.; Wang S.; Gong Y.; Yang S. Ultrafast Zn2+ Intercalation and Deintercalation in Vanadium Dioxide. Adv. Mater. 2018, 30, 1800762. 10.1002/adma.201800762. [DOI] [PubMed] [Google Scholar]
  407. He P.; Quan Y.; Xu X.; Yan M.; Yang W.; An Q.; He L.; Mai L. High-Performance Aqueous Zinc-Ion Battery Based on Layered H2V3O8 Nanowire Cathode. Small 2017, 13, 1702551. 10.1002/smll.201702551. [DOI] [PubMed] [Google Scholar]
  408. Pang Q.; Sun C.; Yu Y.; Zhao K.; Zhang Z.; Voyles P. M.; Chen G.; Wei Y.; Wang X. H2V3O8 Nanowire/Graphene Electrodes for Aqueous Rechargeable Zinc Ion Batteries with High Rate Capability and Large Capacity. Adv. Energy Mater. 2018, 8, 1800144. 10.1002/aenm.201800144. [DOI] [Google Scholar]
  409. Wang X.; Zheng S.; Mu X.; Zhang Y.; Du H. Additive-Free Synthesis of V4O7 Hierarchical Structures as High Performance Cathodes for Lithium Ion Batteries. Chem. Commun. 2014, 50, 6775–6778. 10.1039/C4CC01050G. [DOI] [PubMed] [Google Scholar]
  410. Fei H.; Lin Y.; Wei M. Facile Synthesis of V6O13 Micro-Flowers for Li-Ion and Na-Ion Battery Cathodes with Good Cycling Performance. J. Colloid Interface Sci. 2014, 425, 1–4. 10.1016/j.jcis.2014.03.028. [DOI] [PubMed] [Google Scholar]
  411. Liao M.; Wang J.; Ye L.; Sun H.; Wen Y.; Wang C.; Sun X.; Wang B.; Peng H. A Deep-Cycle Aqueous Zinc-Ion Battery Containing an Oxygen-Deficient Vanadium Oxide Cathode. Angew. Chem., Int. Ed. 2020, 59, 2273–2278. 10.1002/anie.201912203. [DOI] [PubMed] [Google Scholar]
  412. Zhang N.; Jia M.; Dong Y.; Wang Y.; Xu J.; Liu Y.; Jiao L.; Cheng F. Hydrated Layered Vanadium Oxide as a Highly Reversible Cathode for Rechargeable Aqueous Zinc Batteries. Adv. Funct. Mater. 2019, 29, 1807331. 10.1002/adfm.201807331. [DOI] [Google Scholar]
  413. Wei T.; Li Q.; Yang G.; Wang C. High-Rate and Durable Aqueous Zinc Ion Battery Using Dendritic V10O24·12H2O Cathode Material with Large Interlamellar Spacing. Electrochim. Acta 2018, 287, 60–67. 10.1016/j.electacta.2018.08.040. [DOI] [Google Scholar]
  414. Li S.; Wei X.; Chen H.; Lai G.; Wang X.; Zhang S.; Wu S.; Tang W.; Lin Z. A Mixed-Valent Vanadium Oxide Cathode with Ultrahigh Rate Capability for Aqueous Zinc-Ion Batteries. J. Mater. Chem. A 2021, 9, 22392–22398. 10.1039/D1TA04420F. [DOI] [Google Scholar]
  415. Yang F.; Zhu Y.; Xia Y.; Xiang S.; Han S.; Cai C.; Wang Q.; Wang Y.; Gu M. Fast Zn2+ Kinetics of Vanadium Oxide Nanotubes in High-Performance Rechargeable Zinc-Ion Batteries. J. Power Sources 2020, 451, 227767. 10.1016/j.jpowsour.2020.227767. [DOI] [Google Scholar]
  416. Su D. W.; Dou S. X.; Wang G. X. Hierarchical Orthorhombic V2O5 Hollow Nanospheres as High Performance Cathode Materials for Sodium-Ion Batteries. J. Mater. Chem. A 2014, 2, 11185–11194. 10.1039/c4ta01751j. [DOI] [Google Scholar]
  417. Jayaprakash N.; Das S. K.; Archer L. A. The Rechargeable Aluminum-Ion Battery. Chem. Commun. 2011, 47, 12610–12612. 10.1039/c1cc15779e. [DOI] [PubMed] [Google Scholar]
  418. Zhang N.; Dong Y.; Jia M.; Bian X.; Wang Y.; Qiu M.; Xu J.; Liu Y.; Jiao L.; Cheng F. Rechargeable Aqueous Zn-V2O5 Battery with High Energy Density and Long Cycle Life. ACS Energy Lett. 2018, 3, 1366–1372. 10.1021/acsenergylett.8b00565. [DOI] [Google Scholar]
  419. Zhou J.; Shan L.; Wu Z.; Guo X.; Fang G.; Liang S. Investigation of V2O5 as a Low-Cost Rechargeable Aqueous Zinc Ion Battery Cathode. Chem. Commun. 2018, 54, 4457–4460. 10.1039/C8CC02250J. [DOI] [PubMed] [Google Scholar]
  420. Wang S.; Li S.; Sun Y.; Feng X.; Chen C. Three-Dimensional Porous V2O5 Cathode with Ultra High Rate Capability. Energy Environ. Sci. 2011, 4, 2854–2857. 10.1039/c1ee01172c. [DOI] [Google Scholar]
  421. Zuo C.; Xiao Y.; Pan X.; Xiong F.; Zhang W.; Long J.; Dong S.; An Q.; Luo P. Organic-Inorganic Superlattices of Vanadium Oxide@Polyaniline for High-Performance Magnesium-Ion Batteries. ChemSusChem 2021, 14, 2093–2099. 10.1002/cssc.202100263. [DOI] [PubMed] [Google Scholar]
  422. Han C.; Zhu J.; Fu K.; Deng D.; Luo W.; Mai L. A High-Capacity Polyaniline-Intercalated Layered Vanadium Oxide for Aqueous Ammonium-Ion Batteries. Chem. Commun. 2022, 58, 791–794. 10.1039/D1CC05677H. [DOI] [PubMed] [Google Scholar]
  423. Wei Q.; Liu J.; Feng W.; Sheng J.; Tian X.; He L.; An Q.; Mai L. Hydrated Vanadium Pentoxide with Superior Sodium Storage Capacity. J. Mater. Chem. A 2015, 3, 8070–8075. 10.1039/C5TA00502G. [DOI] [Google Scholar]
  424. Chae M. S.; Heo J. W.; Hyoung J.; Hong S.-T. Double-Sheet Vanadium Oxide as a Cathode Material for Calcium-Ion Batteries. ChemNanoMat 2020, 6, 1049–1053. 10.1002/cnma.202000011. [DOI] [Google Scholar]
  425. Xu C.; Li M.; Li K.; Du Z.; Chen J.; Zou F.; Xu S.; Li N.; Li G. VOOH Nanosheets with Enhanced Capacitance as Supercapacitor Electrode. J. Alloys Compd. 2021, 869, 159367. 10.1016/j.jallcom.2021.159367. [DOI] [Google Scholar]
  426. Hu T.; Liu Y.; Zhang Y.; Nie Y.; Zheng J.; Wang Q.; Jiang H.; Meng C. Encapsulating V2O3 Nanorods into Carbon Core-Shell Composites with Porous Structures and Large Specific Surface Area for High Performance Solid-State Supercapacitors. Microporous Mesoporous Mater. 2018, 262, 199–206. 10.1016/j.micromeso.2017.11.044. [DOI] [Google Scholar]
  427. Zhang Y. Designed Synthesis and Supercapacitor Electrode of V2O3@C Core-Shell Structured Nanorods with Excellent Pseudo-Capacitance in Na2SO4 Neutral Electrolyte. ChemistrySelect 2018, 3, 1577–1584. 10.1002/slct.201702705. [DOI] [Google Scholar]
  428. Zheng J.; Zhang Y.; Jing X.; Liu X.; Hu T.; Lv T.; Zhang S.; Meng C. Synthesis of Amorphous Carbon Coated on V2O3 Core-Shell Composites for Enhancing the Electrochemical Properties of V2O3 as Supercapacitor Electrode. Colloids Surf., A 2017, 518, 188–196. 10.1016/j.colsurfa.2017.01.035. [DOI] [Google Scholar]
  429. Yasoda K. Y.; Mikhaylov A. A.; Medvedev A. G.; Kumar M. S.; Lev O.; Prikhodchenko P. V.; Batabyal S. K. Brush Like Polyaniline on Vanadium Oxide Decorated Reduced Graphene Oxide: Efficient Electrode Materials for Supercapacitor. J. Energy Storage 2019, 22, 188–193. 10.1016/j.est.2019.02.010. [DOI] [Google Scholar]
  430. Rakhi R. B.; Nagaraju D. H.; Beaujuge P.; Alshareef H. N. Supercapacitors Based on Two Dimensional VO2 Nanosheet Electrodes in Organic Gel Electrolyte. Electrochim. Acta 2016, 220, 601–608. 10.1016/j.electacta.2016.10.109. [DOI] [Google Scholar]
  431. Ndiaye N. M.; Madito M. J.; Ngom B. D.; Masikhwa T. M.; Mirghni A. A.; Manyala N. High-Performance Asymmetric Supercapacitor Based on Vanadium Dioxide and Carbonized Iron-Polyaniline Electrodes. AIP Adv. 2019, 9, 055309. 10.1063/1.5091799. [DOI] [Google Scholar]
  432. Zhang Y.; Zheng J.; Hu T.; Tian F.; Meng C. Synthesis and Supercapacitor Electrode of VO2(B)/C Core-Shell Composites with a Pseudocapacitance in Aqueous Solution. Appl. Surf. Sci. 2016, 371, 189–195. 10.1016/j.apsusc.2016.02.199. [DOI] [Google Scholar]
  433. Xia X.; Chao D.; Ng C. F.; Lin J.; Fan Z.; Zhang H.; Shen Z. X.; Fan H. J. VO2 Nanoflake Arrays for Supercapacitor and Li-Ion Battery Electrodes: Performance Enhancement by Hydrogen Molybdenum Bronze as an Efficient Shell Material. Mater. Horizons 2015, 2, 237–244. 10.1039/C4MH00212A. [DOI] [Google Scholar]
  434. Fan Y.; Ouyang D.; Li B.-W.; Dang F.; Ren Z. Two-Dimensional VO2 Mesoporous Microarrays for High-Performance Supercapacitor. Nanoscale Res. Lett. 2018, 13, 142. 10.1186/s11671-018-2557-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  435. Nie G.; Lu X.; Zhu Y.; Chi M.; Gao M.; Chen S.; Wang C. Reactive Template Synthesis of Inorganic/Organic VO2@Polyaniline Coaxial Nanobelts for High-Performance Supercapacitors. ChemElectroChem. 2017, 4, 1095–1100. 10.1002/celc.201600830. [DOI] [Google Scholar]
  436. Yang W.; Zeng J.; Xue Z.; Ma T.; Chen J.; Li N.; Zou H.; Chen S. Synthesis of Vanadium Oxide Nanorods Coated with Carbon Nanoshell for a High-Performance Supercapacitor. Ionics 2020, 26, 961–970. 10.1007/s11581-019-03203-1. [DOI] [Google Scholar]
  437. Zeng H. M.; Zhao Y.; Hao Y. J.; Lai Q. Y.; Huang J. H.; Ji X. Y. Preparation and Capacitive Properties of Sheet V6O13 for Electrochemical Supercapacitor. J. Alloys Compd. 2009, 477, 800–804. 10.1016/j.jallcom.2008.10.100. [DOI] [Google Scholar]
  438. Ling W.; Zhang S.; Peng X.; Zhong S.; Liang F.; Geng J.; Zou Z. The Prospected Application of V6O13 in Lithium-Ion Supercapacitors Based on Its Researches in Lithium-Ion Batteries and Supercapacitors. Ionics 2021, 27, 4961–4981. 10.1007/s11581-021-04271-y. [DOI] [Google Scholar]
  439. Zhai T.; Lu X.; Ling Y.; Yu M.; Wang G.; Liu T.; Liang C.; Tong Y.; Li Y. A New Benchmark Capacitance for Supercapacitor Anodes by Mixed-Valence Sulfur-Doped V6O13-x. Adv. Mater. 2014, 26, 5869–5875. 10.1002/adma.201402041. [DOI] [PubMed] [Google Scholar]
  440. Saravanakumar B.; Purushothaman K. K.; Muralidharan G. Interconnected V2O5 Nanoporous Network for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2012, 4, 4484–4490. 10.1021/am301162p. [DOI] [PubMed] [Google Scholar]
  441. Li M.; Sun G.; Yin P.; Ruan C.; Ai K. Controlling the Formation of Rodlike V2O5 Nanocrystals on Reduced Graphene Oxide for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2013, 5, 11462–11470. 10.1021/am403739g. [DOI] [PubMed] [Google Scholar]
  442. Chen Y.; Lian P.; Feng J.; Liu Y.; Wang L.; Liu J.; Shi X. Tailoring Defective Vanadium Pentoxide/Reduced Graphene Oxide Electrodes for All-Vanadium-Oxide Asymmetric Supercapacitors. Chem. Eng. J. 2022, 429, 132274. 10.1016/j.cej.2021.132274. [DOI] [Google Scholar]
  443. Zhang H.; Xie A.; Wang C.; Wang H.; Shen Y.; Tian X. Bifunctional Reduced Graphene Oxide/V2O5 Composite Hydrogel: Fabrication, High Performance as Electromagnetic Wave Absorbent and Supercapacitor. ChemPhysChem 2014, 15, 366–373. 10.1002/cphc.201300822. [DOI] [PubMed] [Google Scholar]
  444. Saravanakumar B.; Purushothaman K. K.; Muralidharan G. High Performance Supercapacitor Based on Carbon Coated V2O5 Nanorods. J. Electroanal. Chem. 2015, 758, 111–116. 10.1016/j.jelechem.2015.10.031. [DOI] [Google Scholar]
  445. Yang J.; Lan T.; Liu J.; Song Y.; Wei M. Supercapacitor Electrode of Hollow Spherical V2O5 with a High Pseudocapacitance in Aqueous Solution. Electrochim. Acta 2013, 105, 489–495. 10.1016/j.electacta.2013.05.023. [DOI] [Google Scholar]
  446. Karade S. S.; Lalwani S.; Eum J.-H.; Kim H. Coin Cell Fabricated Symmetric Supercapacitor Device of Two-Steps Synthesized V2O5 Nanorods. J. Electroanal. Chem. 2020, 864, 114080. 10.1016/j.jelechem.2020.114080. [DOI] [Google Scholar]
  447. Kiruthiga R.; Nithya C.; Karvembu R.; Venkata Rami Reddy B. Reduced Graphene Oxide Embedded V2O5 Nanorods and Porous Honey Carbon as High Performance Electrodes for Hybrid Sodium-Ion Supercapacitors. Electrochim. Acta 2017, 256, 221–231. 10.1016/j.electacta.2017.10.049. [DOI] [Google Scholar]
  448. Nagaraju D. H.; Wang Q.; Beaujuge P.; Alshareef H. N. Two-Dimensional Heterostructures of V2O5 and Reduced Graphene Oxide as Electrodes for High Energy Density Asymmetric Supercapacitors. J. Mater. Chem. A 2014, 2, 17146–17152. 10.1039/C4TA03731F. [DOI] [Google Scholar]
  449. Zhang S.; Chen S.; Luo Y.; Yan B.; Gu Y.; Yang F.; Cao Y. Large-Scale Preparation of Solution-Processable One-Dimensional V2O5 Nanobelts with Ultrahigh Aspect Ratio for Bifunctional Multicolor Electrochromic and Supercapacitor Applications. J. Alloys Compd. 2020, 842, 155882. 10.1016/j.jallcom.2020.155882. [DOI] [Google Scholar]
  450. Zhang Y.; Zheng J.; Wang Q.; Zhang S.; Hu T.; Meng C. One-Step Hydrothermal Preparation of (NH4)2V3O8/Carbon Composites and Conversion to Porous V2O5 Nanoparticles as Supercapacitor Electrode with Excellent Pseudocapacitive Capability. Appl. Surf. Sci. 2017, 423, 728–742. 10.1016/j.apsusc.2017.06.249. [DOI] [Google Scholar]
  451. Yilmaz G.; Lu X.; Ho G. W. Cross-Linker Mediated Formation of Sulfur-Functionalized V2O5/Graphene Aerogels and their Enhanced Pseudocapacitive Performance. Nanoscale 2017, 9, 802–811. 10.1039/C6NR08233E. [DOI] [PubMed] [Google Scholar]
  452. Liu Z.; Zhang H.; Yang Q.; Chen Y. Graphene/V2O5 Hybrid Electrode for an Asymmetric Supercapacitor with High Energy Density in an Organic Electrolyte. Electrochim. Acta 2018, 287, 149–157. 10.1016/j.electacta.2018.04.212. [DOI] [Google Scholar]
  453. Wu Y.; Gao G.; Yang H.; Bi W.; Liang X.; Zhang Y.; Zhang G.; Wu G. Controlled Synthesis of V2O5/MWCNT Core/Shell Hybrid Aerogels through a Mixed Growth and Self-Assembly Methodology for Supercapacitors with High Capacitance and Ultralong Cycle Life. J. Mater. Chem. A 2015, 3, 15692–15699. 10.1039/C5TA02708J. [DOI] [Google Scholar]
  454. Balasubramanian S.; Purushothaman K. K. Carbon Coated Flowery V2O5 Nanostructure as Novel Electrode Material for High Performance Supercapacitors. Electrochim. Acta 2015, 186, 285–291. 10.1016/j.electacta.2015.10.160. [DOI] [Google Scholar]
  455. Tian M.; Li R.; Liu C.; Long D.; Cao G. Aqueous Al-Ion Supercapacitor with V2O5 Mesoporous Carbon Electrodes. ACS Appl. Mater. Interfaces 2019, 11, 15573–15580. 10.1021/acsami.9b02030. [DOI] [PubMed] [Google Scholar]
  456. Ghaly H. A.; El-Deen A. G.; Souaya E. R.; Allam N. K. Asymmetric Supercapacitors Based on 3D Graphene-Wrapped V2O5 Nanospheres and Fe3O4@3D Graphene Electrodes with High Power and Energy Densities. Electrochim. Acta 2019, 310, 58–69. 10.1016/j.electacta.2019.04.071. [DOI] [Google Scholar]
  457. Qian T.; Xu N.; Zhou J.; Yang T.; Liu X.; Shen X.; Liang J.; Yan C. Interconnected Three-Dimensional V2O5/Polypyrrole Network Nanostructures for High Performance Solid-State Supercapacitors. J. Mater. Chem. A 2015, 3, 488–493. 10.1039/C4TA05769D. [DOI] [Google Scholar]
  458. Aamir A.; Ahmad A.; Shah S. K.; Ain N.; Mehmood M.; Khan Y.; Rehman Z. Electro-Codeposition of V2O5-Polyaniline Composite on Ni Foam as an Electrode for Supercapacitor. J. Mater. Sci. Mater. Electron. 2020, 31, 21035–21045. 10.1007/s10854-020-04616-9. [DOI] [Google Scholar]
  459. Manikandan R.; Justin Raj C.; Rajesh M.; Kim B. C.; Park S. Y.; Cho B.-B.; Yu K. H. Polycrystalline V2O5/Na0.33V2O5 Electrode Material for Li+ Ion Redox Supercapacitor. Electrochim. Acta 2017, 230, 492–500. 10.1016/j.electacta.2017.02.031. [DOI] [Google Scholar]
  460. Xu J.; Zheng F.; Xi C.; Yu Y.; Chen L.; Yang W.; Hu P.; Zhen Q.; Bashir S. Facile Preparation of Hierarchical Vanadium Pentoxide (V2O5)/Titanium Dioxide (TiO2) Heterojunction Composite Nano-Arrays for High Performance Supercapacitor. J. Power Sources 2018, 404, 47–55. 10.1016/j.jpowsour.2018.10.005. [DOI] [Google Scholar]
  461. Sun W.; Gao G.; Du Y.; Zhang K.; Wu G. A Facile Strategy for Fabricating Hierarchical Nanocomposites of V2O5 Nanowire Arrays on a Three-Dimensional N-Doped Graphene Aerogel with a Synergistic Effect for Supercapacitors. J. Mater. Chem. A 2018, 6, 9938–9947. 10.1039/C8TA01448E. [DOI] [Google Scholar]
  462. Bi W.; Huang J.; Wang M.; Jahrman E. P.; Seidler G. T.; Wang J.; Wu Y.; Gao G.; Wu G.; Cao G. V2O5-Conductive Polymer Nanocables with Built-in Local Electric Field Derived from Interfacial Oxygen Vacancies for High Energy Density Supercapacitors. J. Mater. Chem. A 2019, 7, 17966–17973. 10.1039/C9TA04264D. [DOI] [Google Scholar]
  463. Jiao Y.; Wan C.; Wu Y.; Han J.; Bao W.; Gao H.; Wang Y.; Wang C.; Li J. Ultra-High Rate Capability of Nanoporous Carbon Network@V2O5 Sub-Micron Brick Composite as a Novel Cathode Material for Asymmetric Supercapacitors. Nanoscale 2020, 12, 23213–23224. 10.1039/D0NR04000B. [DOI] [PubMed] [Google Scholar]
  464. Patil M. D.; Dhas S. D.; Mane A. A.; Moholkar A. V. Clinker-Like V2O5 Nanostructures Anchored on 3D Ni-Foam for Supercapacitor Application. Mater. Sci. Semicond. Process. 2021, 133, 105978. 10.1016/j.mssp.2021.105978. [DOI] [Google Scholar]
  465. Latha K.; Anbuselvi S.; Periasamy P.; Sudha R.; Velmurugan D. Microwave-Assisted hybridised WO3/V2O5 Rod Shape Nanocomposites for Electrochemical Supercapacitor Applications. Inorg. Chem. Commun. 2021, 133, 108927. 10.1016/j.inoche.2021.108927. [DOI] [Google Scholar]
  466. Devarayapalli K. C.; Lee K.; Do H. B.; Dang N. N.; Yoo K.; Shim J.; Prabhakar Vattikuti S. V. Mesostructured g-C3N4 Nanosheets Interconnected with V2O5 Nanobelts as Electrode for Coin-Cell-Type-Asymmetric Supercapacitor Device. Mater. Today Energy 2021, 21, 100699. 10.1016/j.mtener.2021.100699. [DOI] [Google Scholar]
  467. Javed M. S.; Najim T.; Hussain I.; Batool S.; Idrees M.; Mehmood A.; Imran M.; Assiri M. A.; Ahmad A.; Ahmad Shah S. S. 2D V2O5 Nanoflakes as a Binder-Free Electrode Material for High-Performance Pseudocapacitor. Ceram. Int. 2021, 47, 25152–25157. 10.1016/j.ceramint.2021.05.181. [DOI] [Google Scholar]
  468. Sun G.; Ren H.; Shi Z.; Zhang L.; Wang Z.; Zhan K.; Yan Y.; Yang J.; Zhao B. V2O5/Vertically-Aligned Carbon Nanotubes as Negative Electrode for Asymmetric Supercapacitor in Neutral Aqueous Electrolyte. J. Colloid Interface Sci. 2021, 588, 847–856. 10.1016/j.jcis.2020.11.126. [DOI] [PubMed] [Google Scholar]
  469. You M.; Zhang W.; Yan X.; Jiang H.; Miao J.; Li Y.; Zhou W.; Zhu Y.; Cheng X. V2O5 Nanosheets Assembled on 3D Carbon Fiber Felt as a Free-Standing Electrode for Flexible Asymmetric Supercapacitor with Remarkable Energy Density. Ceram. Int. 2021, 47, 3337–3345. 10.1016/j.ceramint.2020.09.175. [DOI] [Google Scholar]
  470. Zhou P.; Lv X.; Gao Y.; Cui Z.; Liu Y.; Wang Z.; Wang P.; Zheng Z.; Dai Y.; Huang B. Enhanced Electrocatalytic HER Performance of Non-Noble Metal Nickel by Introduction of Divanadium Trioxide. Electrochim. Acta 2019, 320, 134535. 10.1016/j.electacta.2019.07.046. [DOI] [Google Scholar]
  471. Zhou P.; Zhai G.; Lv X.; Liu Y.; Wang Z.; Wang P.; Zheng Z.; Cheng H.; Dai Y.; Huang B. Boosting the Electrocatalytic HER Performance of Ni3N-V2O3 via the Interface Coupling Effect. Appl. Catal., B 2021, 283, 119590. 10.1016/j.apcatb.2020.119590. [DOI] [Google Scholar]
  472. Hu M.; Huang J.; Zhang S.; Liu Z.; Li Q.; Yang M.; Li H.; Goto T.; Tu R. In situ Synthesis of V2O3@Ni as an Efficient Hybrid Catalyst for the Hydrogen Evolution Reaction in Alkaline and Neutral Media. Int. J. Hydrog. Energy 2021, 46, 9101–9109. 10.1016/j.ijhydene.2020.12.205. [DOI] [Google Scholar]
  473. Zhang H.; Qian G.; Chen X.; Jiang W.; Yu T.; Wang Y.; Luo L.; Yin S. V2O3-Decorated Spinel CoFe2O4 with Carbon-Encapsulated Mesoporous Nanosheets for Efficient Water Splitting. ACS Sustain. Chem. Eng. 2021, 9, 980–986. 10.1021/acssuschemeng.0c08477. [DOI] [Google Scholar]
  474. Xie Z.; Wang W.; Ding D.; Zou Y.; Cui Y.; Xu L.; Jiang J. Accelerating Hydrogen Evolution at Neutral pH by Destabilization of Water with a Conducting Oxophilic Metal Oxide. J. Mater. Chem. A 2020, 8, 12169–12176. 10.1039/D0TA04241B. [DOI] [Google Scholar]
  475. Sarika S.; Abhilash S.; Sumi V. S.; Rijith S. Synthesis and Characterization of Transition Metal Mixed Oxide Doped Graphene Embedded Durable Electrocatalyst for Hydrogen Evolution Reaction. Int. J. Hydrog. Energy 2021, 46, 16387–16403. 10.1016/j.ijhydene.2021.01.137. [DOI] [Google Scholar]
  476. Yang X.; Wang X.; Zhao T.; Ma Y.; Wang Z.; Zhao C. Electrodeposited of Ultrathin VOx-Doped NiFe Layer on Porous NiCo Phosphide for Efficient Overall Water Splitting. Appl. Phys. Lett. 2021, 119, 103902. 10.1063/5.0061856. [DOI] [Google Scholar]
  477. Li Y.; Tan X.; Yang W.; Bo X.; Su Z.; Zhao T.; Smith S. C.; Zhao C. Vanadium Oxide Clusters Decorated Metallic Cobalt Catalyst for Active Alkaline Hydrogen Evolution. Cell Rep. Phys. Sci. 2020, 1, 100275. 10.1016/j.xcrp.2020.100275. [DOI] [Google Scholar]

Articles from Chemical Reviews are provided here courtesy of American Chemical Society

RESOURCES