Next Article in Journal
Microstructure and Mechanical Properties of Stainless Steel/6082 Aluminum Alloy Heterogeneous Laser Welded Joint
Next Article in Special Issue
Technological Advances and Market Developments of Solid-State Batteries: A Review
Previous Article in Journal
Strength and Microscopic Mechanism of Cement–Fly Ash–Slag–Desulfurization Gypsum Solidified Mica Schist Weathered Soil
Previous Article in Special Issue
Design and Optimization Strategies for Flexible Quasi-Solid-State Thermo-Electrochemical Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polymorphs of Nb2O5 Compound and Their Electrical Energy Storage Applications

1
School of Material Science and Engineering, University of Jinan, Jinan 250022, China
2
State Key Laboratory of Crystal Materials, Institute of Novel Semiconductors, Shandong University, Jinan 250100, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(21), 6956; https://doi.org/10.3390/ma16216956
Submission received: 30 September 2023 / Revised: 26 October 2023 / Accepted: 27 October 2023 / Published: 30 October 2023
(This article belongs to the Special Issue Advances in Smart Materials for Energy Storage and Conversion)

Abstract

:
Niobium pentoxide (Nb2O5), as an important dielectric and semiconductor material, has numerous crystal polymorphs, higher chemical stability than water and oxygen, and a higher melt point than most metal oxides. Nb2O5 materials have been extensively studied in electrochemistry, lithium batteries, catalysts, ionic liquid gating, and microelectronics. Nb2O5 polymorphs provide a model system for studying structure–property relationships. For example, the T-Nb2O5 polymorph has two-dimensional layers with very low steric hindrance, allowing for rapid Li-ion migration. With the ever-increasing energy crisis, the excellent electrical properties of Nb2O5 polymorphs have made them a research hotspot for potential applications in lithium-ion batteries (LIBs) and supercapacitors (SCs). The basic properties, crystal structures, synthesis methods, and applications of Nb2O5 polymorphs are reviewed in this article. Future research directions related to this material are also briefly discussed.

1. Introduction

Recently, enthusiasm surrounding research on the structure, properties, and applications of Nb2O5 has grown. In particular, its potential use as catalysts, photocatalysts, supercapacitors (SCs), lithium-ion batteries (LIBs), etc., has been investigated. In nature, niobium (Nb) does not occur in a free state but rather is typically found in a mineral form, (Fe, Mn)Nb2O6, known as columbite [1,2,3]. Research on Nb2O5 commenced in the 1940s, with its crystal form the first to be evaluated [4]. The historical research progress on Nb2O5 is shown in Figure 1. Nb2O5 exhibits plentiful polymorphs linked to NbO6 or NbO7 polyhedra by corner- or edge-sharing forms, leading to exceptional properties and various applications [1,2]. As a white powder, Nb2O5 exhibits redox properties and is non-toxic and insoluble in water and acid but soluble in molten potassium bisulfate, alkali metal carbonates, and hydroxides [5].
Besides its structure, Nb2O5 has garnered significant attention due to its morphology, size, and corresponding properties. The nanostructured Nb2O5 often has distinct physical and chemical properties compared to its bulk forms due to its quantum effect and high specific surface area ratio [6]. Nb2O5 nanostructures and thin films have undergone thorough investigation as promising electrode materials for LIBs and SCs. Nb2O5 has charge/discharge plateaus (1–2 V vs. Li0/Li+) and two redox couples (Nb5+/Nb4+ and Nb4+/Nb3+), resulting in higher specific capacities [7]. Orthorhombic Nb2O5 possesses a two-dimensional Li-ion transport pathway without kinetics limitations, facilitating a rapid charge rate [8]. As such, research has focused considerably on controlling the crystal phase, nanostructured Nb2O5 synthesis, and composite formation [9].
This article provides a comprehensive overview of Nb2O5, encompassing its crystal structure, preparation techniques, fundamental properties, and diverse applications. It places a particular emphasis on the in-depth exploration of Nb2O5’s usage in LIBs and SCs. The final section discusses the potential for further research on Nb2O5 to expand its application scope.

2. Polymorphs of Nb2O5 and Synthesis Methods

The subsequent section provides a detailed description of the fundamental properties of Nb2O5, encompassing its synthesis methods, crystal structure, and product morphology. An overview of the physical and chemical properties of Nb2O5 is also provided, covering the fundamental aspects of this material.

2.1. Polymorphs of Nb2O5

Nb2O5 is a white crystal or powder with a wide band gap that boasts a relative density of 4.6 and an exceptionally high melting point of 1520 °C. Upon heating, it becomes yellow and exhibits solubility in hydrofluoric acid and sulfuric acid while remaining insoluble in water. Nb2O5 has diverse physical and chemical properties, making it highly valuable in various applications. Meanwhile, it remains stable in air, insoluble in water, complex in structure, and exhibits extensive polymorphism [1]. In recent decades, researchers have identified more than 15 distinct crystalline phases of Nb2O5 (as shown in Table 1), including TT-, T-, B-, M-, N-, P-, R-, and H-Nb2O5. These diverse configurations have unique properties and can be utilized in a wide range of applications [10].
However, not all crystal phases of Nb2O5 are equally prevalent. The most common include TT-Nb2O5, T-Nb2O5, and H-Nb2O5; their corresponding lattice parameters are shown in Table 2. These crystal structures possess distinct characteristics and can be utilized in various applications [1,19]. Notably, in certain articles, the crystal phases of Nb2O5 have been re-named as γ (i.e., T), β (i.e., M), and α (i.e., H) [1].
Amorphous Nb2O5 is typically obtained through various low-temperature synthesis methods and can be subsequently crystallized into either the TT or T phase at approximately 500 °C (Figure 2). At intermediate temperatures (~800 °C), the material can transform into the M phase (tetragonal), while the H phase forms at ≥1000 °C. It is important to note that, besides temperature, other factors can also influence the formation of Nb2O5 crystals. The phase stability is affected by environmental temperature and pressure, while the existence of polymorphs depends on the heating and preparation methods [22]. However, the nature of the starting material and the presence of impurities are also important factors affecting the formation of Nb2O5 crystals [11,23]. As a result, temperature should be considered an indicative factor rather than the sole determinant of the crystalline phase of Nb2O5.
The different polymorphs of Nb2O5 include distorted octahedra (NbO6); the degree of distortion is influenced by the type of connection between the octahedra, whether through edges, angles or a combination of both [24]. Schäfer et al. [11] identified multiple modes of attachment between the octahedra while maintaining an O/Nb ratio of 2.5, confirming the existence of various arrangements of Nb2O5 and polycrystalline forms of niobium oxide. Nb atoms exhibit 6-fold (NbO6) and 7-fold (NbO7) coordination in the T and TT phases, along with distorted octahedral and pentagonal bipyramidal sites.
Undoubtedly, different crystalline phases of Nb2O5 exhibit different properties. For instance, the dielectric constant of the H phase can reach 100. The T-phase has applications in electrochemistry owing to its excellent electrochemical properties, including its resistance to reactions with other substances and cyclic stability. Extensive research has also been conducted to assess the application of the TT phase as an electrochromic material.

2.1.1. T- Nb2O5

The T-Nb2O5 phase possesses an orthorhombic crystal structure with a Pbam space group. In its conventional cell, it comprises 16.8 Nb and 42 O atoms. Among these, sixteen Nb ions occupy four Wyckoff positions 8i with half occupancy, while eleven O atoms are distributed across one 2b, four 4g, and six 4h positions. The remaining 0.8 Nb atoms exhibit random occupancies of 0.08, 0.08, and 0.04 across three 4g Wyckoff positions. The orthorhombic T-Nb2O5 phase comprises Nb atoms surrounded by six or seven oxygen atoms, forming twisted octahedral or pentagonal bipyramidal shapes [25] (Figure 3a). In the T and TT phases, Nb atoms have six-fold (NbO6) and seven-fold (NbO7) coordination, which can lead to the formation of distorted octahedrons and pentagonal bipyramids, respectively [24] (Figure 3a,d).
Serghiou et al. first discovered that the T phase of Nb2O5 becomes amorphous under certain conditions, including pressure and temperatures up to 19.2 GPa and 300 K, respectively. During this amorphous process, the oxide becomes amorphous and reduced [26].
In the 2010s, T-Nb2O5 gradually became a research hotspot. Li et al. reported a chemical method without catalyst topology, combined with molten salt synthesis (MSS), to achieve the large-scale synthesis of rod-like H-Nb2O5 and sheet-like T-Nb2O5 single crystals. Subsequently, Raman spectroscopy, SEM, X-ray diffraction, and TEM were employed to investigate the structural changes involved in the process [27].
In the mid-2010s, research on applying T-Nb2O5 intercalated pseudocapacitors gained popularity. Augustyn et al. recently demonstrated that mesoporous and nanocrystalline films of T-Nb2O5 (i.e., trapezoidal Nb2O5) exhibit behavior consistent with pseudocapacitance upon inserting lithium ions [28]. Meanwhile, Kong et al. employed a hydrothermal synthesis technique to anchor T-Nb2O5 nanocrystals onto conductive graphene sheets, fabricating asymmetric SCs. The T-Nb2O5/graphene nanocomposites and mesoporous carbon were used as negative/positive electrodes and exhibited high-rate responses, significantly enhancing the performance of SCs [29]. According to Lim et al., The Nb2O5@C NCs core-shell nanocrystals exhibit excellent electrochemical performance. More specifically, the core-shell structure provides a high surface area and efficient electron/ion transport pathways, with the carbon shell acting as a buffer to accommodate volume changes during cycling. The one-pot synthesis method offers a simple and scalable approach for producing high-performance anode materials for next-generation LIBs [30].
In the late 2010s, T-Nb2O5 gained attention as a promising anode material for sodium-ion batteries (NIBs) due to its unique surface frame and large interplanar lattice spacing [31]. In addition, research into applying T-Nb2O5 in potassium-ion batteries (KIBs) began. For the first time, Li et al. studied nanostructured T-Nb2O5 as a cathode material for KIBs [32].
During the late 2010s and early 2020s, most studies focused on modifying the T-Nb2O5 anode and cathode of hybrid supercapacitor (HSC) or optimizing synthesis methods to improve its defects. Additionally, the application of T-Nb2O5 in lithium-sulfur batteries was explored. Wang et al. [33] reported on a novel niobium oxide matrix that incorporates amorphous/crystalline hetero-conjunctions (A/T-Nb2O5), serving as a two-in-one host for a Li-S system. This matrix significantly improves the electrochemical performance of the Li-S system, with A-Nb2O5 providing a high surface area for sulfur adsorption and T-Nb2O5 offering a stable framework for sulfur confinement. The amorphous/crystalline hetero-conjunctions further enhance the conductivity of the matrix and facilitate the transport of lithium ions, presenting a promising strategy for developing high-performance electrode materials in next-generation energy storage systems.

2.1.2. TT-Nb2O5

At low temperatures, the two phases of Nb2O5, TT, and T, share many similarities. For example, the X-ray diffraction patterns of the TT and T phases of Nb2O5 exhibit remarkable resemblance (Figure 4a,b). However, a key distinguishing factor is that the XRD pattern of the TT-Nb2O5 phase includes a peak corresponding to a split reflection, which is not present in the T-Nb2O5 phase [1,23]. The emergence of the TT phase of Nb2O5 is not necessarily associated with using pure components as starting materials. Rather, TT may represent a less crystalline form of T, possibly stabilized by impurities [23]. In certain cases, oxygen atoms within Nb2O5 can be substituted by monovalent species like OH or Cl or even vacancies. This substitution process helps stabilize the TT phase of Nb2O5.
In TT-Nb2O5, the unit cell of Nb2O5 contains half of its formula equivalent, along with a constitutional defect involving the absence of one oxygen atom per unit cell [25] (Figure 3d). The Nb atom is positioned at the center of 4, 5, or 6 oxygen atoms within the ab plane. The crystal structure of Nb2O5 along the a-axis comprises two in-plane niobium layers interconnected by oxygen bonds, forming a relatively stacked structure. Meanwhile, the b-axis exhibits a tunnel-like structure with an approximate diameter of 4 Å, extending throughout the entire structure. This unique feature of Nb2O5 is responsible for its enhanced ion diffusion properties [35,36,37]. Along the c-axis, the crystal structure of Nb2O5 exhibits a disordered hexagon formed by oxygen atoms in six faces and an Nb-O-Nb-O chain structure. The absence of oxygen in the structure leads to the deformation of these polyhedra.
During the 1950s, Frevel and Rinn conducted studies on Nb2O5 and discovered a “low temperature” phase, now known as TT-Nb2O5. Their research utilized X-ray powder diffraction and identified a pseudo-hexagonal phase with lattice constants a = 3.607 Å and c = 3.925 Å. This phase transforms into T-Nb2O5 when heated to 973 K. Since its discovery, TT-Nb2O5 has been the subject of considerable scientific debate. Some believe the TT phase is merely a variant of T-Nb2O5, suggesting it is less crystalline than T-Nb2O5 [38]. Others say it is not even a separate phase [39]. Meanwhile, Terao was the first to identify the orthogonal structure of T-Nb2O5 and suggested that TT-Nb2O5 is a metal oxide with crystallization defects instead of a poorly crystallized T-phase [40]. Tamura suggested that TT-Nb2O5 possesses monoclinic cells and is stabilized by trace impurities in the sample [41]. Weissman et al. proposed that the structures of TT-Nb2O5 and T-Nb2O5 are highly similar, noting the absence of peak splitting and the presence of broad peaks in TT-Nb2O5, indicating that niobium resides in a more symmetrical 4h Wyckoff position rather than the 8i position found in T-Nb2O5 [42]. Their model suggests that TT-Nb2O5 can be indexed within the orthorhombic system of T-Nb2O5 and comprises subcell domains within a superlattice. Despite the distortions leading to unusual crystallization, TT-Nb2O5 displays pseudo-hexagonal subcell symmetry, while the superlattice follows an orthonormal crystal system [42]. Recently, Košutová et al. demonstrated an improved TT structure using hexagonal cells [43]. TT-Nb2O5 exhibits a complex structure with a disordered pseudohexagonal subcell of β = 120°, resulting in an orthogonal superlattice that resembles the T-phase. Niobium atoms occupy a more symmetrical Wyckoff position in this phase. The disordered polyhedron in TT-Nb2O5 forms a tunnel structure along the b-axis. Given its significance in the performance and applications of TT-Nb2O5, further investigation into its structure is warranted [44].

2.1.3. H-Nb2O5

H-Nb2O5 exhibits a high degree of order, with its structure divided into blocks [23]. In this monoclinic lattice, H-Nb2O5 comprises groups of ReO3-type blocks, encompassing 3 × 4 and 3 × 5 blocks. These blocks are arranged in a layered structure, with the 3 × 4 blocks forming the first layer and the 3 × 5 blocks forming the second layer. It contains the NbO6 octahedron [25] (Figure 3c). The ReO3-type blocks in H-Nb2O5 are coupled by edge-sharing and move half-unit cell size along the c-axis. The NbO6 units are connected through shared angles within a block [25]. For every 28 Nb atoms in the cell, one resides at the tetrahedral site, typically at certain block junctions.
This variant of Nb2O5 is readily available. Indeed, H-Nb2O5 can be produced in any other form when heated in air to ~1100 °C. If M-type single crystals are heated at 1100 °C, they retain their single-crystal properties during the transition to H-type, whereas all other Nb2O5 crystals convert to polycrystalline H-Nb2O5 [11].
H-Nb2O5 was initially identified as a high-temperature variant of Nb2O5 by Brauer in 1941. Later, in 1964, Gatehouse and Wadsley determined the crystal structure of H-Nb2O5 using Patterson and differential synthesis methods. Subsequently, Busing et al. used ORFLS to refine the structure by the full matrix least square method. Kato conducted additional refinements of the H-Nb2O5 structure in 1975 [16]. In the 21st century, extensive research has been conducted on the applications of H-Nb2O5.

2.1.4. M-Nb2O5

M-Nb2O5 shares a crystal structure similar to H-Nb2O5 and contains 4 × 4 ReO3-type blocks. The close resemblance between M- and H-type X-ray powder patterns accounts for this similarity.
M-Nb2O5 belongs to the tetragonal system with lattice constants a = 20.44 Å and b = 3.832 Å. The space group is I4/mmm. Presumed metastable variants of Nb2O5 can be obtained through various pathways within the temperature range of ~900 °C. Single crystals are obtained using chemical transport techniques. In the structure of M-Nb2O5 (Figure 3b), Nb-O octahedrons are arranged in a blocky manner and can be considered part of the ReO3 lattice. Formally, the structure of M-Nb2O5 is derived from that of N-Nb2O5. Perpendicular to the C-axis, the dimension of the block measures 4 × 4 octahedral diagonal units. In this plane, adjacent blocks are connected by edges. In the c direction, the octahedrons are connected by angles. In M-Nb2O5, four structures are interconnected to form a set of four edge-connected octahedrons (NbO4/4O2/2).

2.1.5. B-Nb2O5

B-Nb2O5 features a monoclinic crystal structure indexed in the space group C2/c, with four formula units in each cell. The structure comprises eight cations (Nb) located at the 8f Wyckoff position and 20 anionic oxygen ions (O) located at three Wyckoff locations: 4e (O1) and two 8f (O2 and O3). The crystal structure of the B-Nb2O5 phase consists of twisted NbO6 octahedral blocks arranged in a string of pairs of shared edge octahedra connected in a zigzag pattern of shared angle octahedrons (Figure 3e). According to the study of Pinto et al., DFT calculations and experimental evidence confirm that the B phase is the more stable phase at lower temperatures [14].

2.1.6. Other Nb2O5 Phases

In addition to the common crystal phases of Nb2O5, there are several less common or less studied phases, including N-Nb2O5, P-Nb2O5, and R-Nb2O5, among others. The presence of twin domains observed in previous studies likely results from the crystal structure and growth conditions during synthesis. The presence of coherent twins at 90° from each other suggests that the crystal structure of N-Nb2O5 possesses a high degree of symmetry. Regions containing blocks of different sizes may be attributed to defects or impurities within the crystal structure. Moreover, N-Nb2O5 and M-Nb2O5 differ only in structural arrangement. The synthesis of N-Nb2O5 was first investigated by Schafer et al. in 1964 [45], with its crystal structure reported by Andersson et al. in 1967 [17]. Meanwhile, Uyeda et al. (1984) [46] had earlier published HRTEM images of N-Nb2O5 and utilized the compound as a standard specimen; however, they did not describe the various types of twins in N-Nb2O5. Overall, these studies have provided insights into the microstructure of N-Nb2O5, emphasizing the importance of understanding crystal structure and defects in materials science.
RS-Nb2O5 is a nanostructured rock salt Nb2O5 electrode, reported by Barnes et al. [47]. This crystalline phase has excellent electrical properties through the amorphous to crystalline transition during the repeated electrochemical cycle with Li+.

2.2. Synthesis Methods

Researchers have taken a keen interest in Nb2O5, a non-toxic substance with ecological friendliness and a robust oxidation capacity. Due to its broad potential applications, extensive investigations have been performed to explore the various preparation methods for Nb2O5. The primary motivation for selecting the appropriate synthesis method lies in its capability to modify the properties of Nb2O5, particularly its crystal phase and morphology.
Significant attention has been devoted to developing niobium oxide films or particles, revealing their unique properties. Various techniques have been employed, including pulsed laser decomposition (PLD), electrodeposition, magnetron sputtering, plasma immersion ion implantation, and the sol-gel process. These methods have been utilized to prepare Nb2O5 nanostructures. Indeed, the characteristics of the surface nanostructures formed by these different synthesis methods are unique (Figure 5). For example, the hydrothermal method can be used to synthesize Nb2O5 nanorods and sea urchin nanostructures (Figure 5a,b). Meanwhile, the precipitation rule formed a nanosphere layer with a porous structure (Figure 5c). The samples prepared by the sol-gel method are coral-like nanostructures (Figure 5d), whereas longer Nb2O5 nanowires are synthesized by electrospinning (Figure 5e). Still further, filamentous nanostructures with chaotic surfaces can be formed via electrospinning (Figure 5f). Additionally, a self-organized microstructure of niobium oxide has been produced through potentiostatic anodization.

2.2.1. Hydrothermal and Solvothermal Methods

The hydrothermal and solvothermal methods are widely used due to their simplicity, low cost, and high yield. Typically, the reaction occurs within an autoclave or a Teflon-lined stainless steel vessel filled with water or organic solvents. The Nb ion arises from the reaction of niobium with acidic or basic solutions or by dissolving niobium salts; the solution is then heated to 100–600 °C for several hours or even days, during which Nb2O5 nanostructures grow [54,55,56]. Nb2O5 polycrystal hollow nanospheres and single-crystal nanotubes have been prepared by adjusting the molar ratio of Nb/Ti and the amount of F ions used [56]. Meanwhile, tree-like Nb2O5 nanotrees have been fabricated by secondary nucleation of Nb2O5 nanowires under hydrothermal conditions [57]. Additionally, monoclinic Nb2O5 nanotube arrays from pseudo-hexagonal Nb2O5 nanorod arrays have been reported. The phase transformation was due to energy differences between the pseudo-hexagonal and monoclinic Nb2O5 nanostructures [58]. Future developments may further optimize preparation methods and properties of Nb2O5 nanoparticles, potentially through doping with other elements or surface modifications with organic molecules to improve their electrochemical properties.

2.2.2. Anodization Method

The anodization method is a prevalent nanofabrication technique renowned for creating highly porous and well-ordered oxide structures. The anodization process is initiated by immersing the working and counter electrodes (usually a platinum sheet) in an electrolyte. Current and voltage are then applied, initiating chemical reactions at the interface between the electrolyte and electrode, forming a thin film structure. The shape and size of the resulting film are related to myriad factors, including the electrolyte composition, electrolyte temperature, amount of current or voltage applied, and the conduction time.
Anodized films are typically amorphous, however, they can be crystallized through annealing. Research on anodized Nb2O5 film dates back to the 1960s. Draper and colleagues studied the formation of oxide films on niobium tablets, investigating how the electrolyte’s composition affects the film material’s growth, particularly with regard to structural irregularities [57]. Subsequently, studies reported on the properties of anode Nb2O5 films (i.e., resistivity, dielectric constant). Other initial investigations into the surface morphology of anodized Nb2O5 and crystal analysis using electron microscopy electron beam crystallization have also been documented [58,59].
The anodization process of Nb has been extensively investigated using various electrolytes, including sulfuric acid, phosphoric acid, NaOH, Na2CO3, HF, glycerol, and phosphate-based solutions. These studies aimed to understand the electrochemical behavior and optimize the formation of oxide layers on the Nb surfaces for different applications. Most reported anodic Nb2O5 films possess a highly nanoporous structure on an Nb foil substrate [60]. One of the most significant studies in this area was conducted by Habazaki and colleagues, who discussed the influence of water content and elevated temperatures on the formation of porous Nb2O5 during the anodization process using a K2HPO4-glycerol electrolyte. They explored how these factors affect the morphology, structure, and properties of the resulting porous Nb2O5 films [60]. They reported that maintaining the water content of the electrolyte at 0.08 mass% at 160 °C significantly increased the film’s thickness. This was attributed to intensified field strength arising from the heightened concentration of phosphorus species in the electrolyte solution. Meanwhile, Ou and colleagues developed an Nb2O5 crisscross nanoporous network by high-temperature anodization [61].
Throughout these processes, ion diffusion during anodization is affected by several factors. Apart from nanoporous structures, microcones can also be produced through Nb anodization in an electrolyte solution with low hydrofluoric acid content in deionized water or a glycerol electrolyte solution containing K2HPO4 [62,63]. Wei and colleagues improved the anodic oxidation of Nb in NH4F-based glycerol electrolytes and successfully fabricated Nb2O5 nanotubes up to 4 mm [64]. Lee et al. first reported the anodic oxidation process of highly ordered Nb2O5 nanochannels grown at 180 °C in a glycerol electrolyte solution containing K2HPO4 [65]. Lee et al. preheated the electrolyte to 200 °C before anodic oxidation to reduce the water content and successfully grew Nb2O5 thick nanochannel films. Similarly, Abdul Rani and colleagues reported the synthesis of Nb2O5 thin films on anodic nanochannels [66].
Other studies have reported on the production of nanoporous and nanochannel Nb2O5 thin films through Nb thin film anodization and deposition via RF sputtering on FTO glass substrates [67,68]. For the first time, Liu and colleagues successfully demonstrated the preparation of Nb2O5 nanotube powders by a simple electrochemical anodization method. In the anodizing process, the anodic oxide was released continuously and spontaneously into the electrolyte, which was collected to produce a white powder. The resulting powder primarily comprised nanotubes, with lengths ranging from ~50 to 100 nm and diameters of approximately 20 to 30 nm [69]. Meanwhile, Khairir et al. prepared a nanoporous thin film structure of Nb2O5 by the anodic oxidation method under different oxidation times; longer anodic oxidation times led to larger pore sizes [70].
Pligovka et al. conducted X-ray diffraction analysis on niobium oxide nanostructures, including defective non-uniform arrays and nanocolumns, fabricated via electrochemical anodization [71]. Gorokh et al. further studied the composition of columnar niobium oxide nanostructures (CNONS), synthesized initial samples by vacuum magnetron sputtering Nb/Al targets on a silicon substrate, and then synthesized samples by anodization and repeated anodization (reanodization) in different electrolytes. Finally, the growth mechanism of CNONS was proposed by analyzing the samples. The infrared spectroscopy analysis of CNONS identified three oxide phases in the NB-O system: NbO, NbO2, and Nb2O5. Moreover, the top of the column was composed primarily of two Nb2O5 modifications: α-Nb2O5 and β-Nb2O5 [72]. In 2021, Pligovka et al. analyzed the morphology of niobium oxide nanocrystals formed by an electrochemical anodizing Al/Nb system with a similar synthesis method. They formed three different embryonic morphological types: skittle-, medusa- and goblet-like (Figure 6). Moreover, the morphological characteristics of the region between the top and bottom of the formed nanostructures were closely related to the electrolyte and the formation voltage of the anodic oxidation process [73].
The electrophysical properties of columnar niobium oxide nanostructures have also been studied. The measurement results showed that the current-voltage I-U curve is nonlinear and asymmetrical. An increase in temperature resulted in an increase in current. This behavior may represent a p-n junction or a metal-semiconductor junction [74]. Recently, its application in nano-optical biosensors has also been preliminarily studied [75].

2.2.3. Sol-Gel Methods

Since the sol-gel method was discovered in the 1970s, it has been rapidly developed and successfully applied in various material engineering fields [76]. Ulrich in 1988 [77] and Hench in 1990 [78] described the broad prospect of this method. To date, its application has been extensive, particularly in wet chemistry, as it allows for the precise control of composition and homogeneity during material synthesis at low temperatures. Furthermore, it offers a simple process with the potential for cost-effective large-scale production.
To generate sols, alcohol solutions are typically hydrolyzed with water, followed by polymerization or aggregation to produce dispersed fine particles. Depending on the desired sol composition and properties, these precursor compounds can be inorganic or metal-organic. The fundamental chemistry of sol-gel processes has been thoroughly covered in other reviews [79]. As per the summary of Bokov et al., during the sol-gel process, molecular precursors such as metallic alkoxides undergo hydrolysis or alcoholysis, resulting in gel formation from solutions in water or alcohol via heating and agitation. Since the gels resulting from these processes usually retain some water and solvent, it is necessary to dry them using appropriate methods that align with their intended properties and applications. For example, if an alcohol solution is utilized, drying can be accomplished by burning off the alcohol. Once dried, the resulting gel is often powdered and subjected to calcination [80]. In brief, the sol particles connect, forming a network of inorganic polymers known as gels, which retain some residual water and solvent. A dry gel is formed or coated in a transition window from sol to gel before removing the remaining water and solvent. The gel is then heat-treated to form the final dense product. Depositing oxide films using the sol-gel method is typically accomplished through dipping or spin coating techniques. However, the sol-gel method does have certain limitations, including the potential for uneven film thickness due to weak bonding. Moreover, it is difficult to control the reaction rate and porosity [81].
Alquier et al. first reported the preparation of Nb2O5 salt and its solvent by the sol-gel method in 1986 [82]. Post-annealing treatment is typically necessary to induce the crystallization of Nb2O5 films prepared via these processes. Schmitt et al. prepared these thin films by dissolving NbCl5 powder in a mixture of butanol and acetic acid [83]. Following the preparation of the Nb2O5 film by dip-coating, the Nb2O5 coating was annealed at different temperatures from 400 to 600 °C, causing its structure to shift from amorphous to crystalline (i.e., a TT structure). Subsequently, others have demonstrated analogous synthetic procedures using alternative precursors, such as niobium ethanol and ammonium niobium oxalate. For instance, Melo et al. prepared Nb2O5 and Nb2O5: Li+ films by the sol-gel method with acoustic catalysis. They employed NbCl5 as the precursor and butanol as the solvent, adding lithium salt LiCF3SO3 into the precursor solution to obtain thin films with different electrochemical properties [84]. In 2021, Xu et al. studied the impact of crystallinity on the optical properties of sol-gel-prepared Nb2O5. TG-DSC analysis revealed amorphous Nb2O5 up to 460 °C. Meanwhile, XRD results indicated the transformation of amorphous Nb2O5 into a pseudo-hexagonal phase, with higher temperatures resulting in improved crystallinity. UV-VIS and Raman spectra results demonstrated a progression in the arrangement of Nb2O5 atoms, starting with short-range order and progressing to medium-range, ultimately reaching a long-range order. That is, as the temperature increased, atoms became long-range ordered structures by connecting structural units [85].

2.2.4. Electrodeposition

Electrodeposition is widely used to manufacture nanostructured materials in various forms, such as powder, composite, and thin films. It is a well-established technique that produces a thin and uniform coating by applying an electric potential between two electrodes immersed in a solution, causing metal ions to deposit onto a substrate. In contrast to the anodization method described in Section 2.2.2, the metal oxide film formed by electrodeposition is at the cathode. Electroreduction is commonly used to produce low-cost (metal) materials, while electrooxidation is typically employed for high-value material production [86].
The electrodeposition process can be tailored by adjusting several parameters, including current density/voltage, temperature, and the addition of agents to the electrolyte solution. Additionally, modifying the substrate’s surface properties can achieve the desired electrodeposition characteristics. The current density/voltage pair is pivotal in controlling electrodeposition characteristics and determining the threshold for different reactions. A significant benefit of electrodeposition is the ability to regulate the reaction using an external circuit, enabling accurate and simple control over material deposition [86]. Niobium ions and hydrogen peroxide are typically required in the aqueous electrolytes for the electrodeposition of Nb2O5 [87].
Cathodic electrodeposition of niobium oxide (NbOx) is hindered by Nb3+’s extremely negative reduction potential (−1.1 V vs. NHE), leading to the coevolution of destructive H2 gas when the reaction occurs in aqueous solutions [88]. Crayston et al. [89] avoided this issue by using electrically generated OH to electrodeposit NbOx, precipitating niobium ions, and immobilizing them in porous films prepared using the sol-gel technique. Meanwhile, in the late 1990s, Zhitomirsky electrodeposited Nb2O5 thin films on platinum-coated silicon or platinum substrates using peroxide as a precursor. They applied a stable 20 mA cm−2 current during the 20-min electrodeposition process, which occurred at 1 °C. The electrolyte mixture comprised NbCl5 and H2O2 [90,91]. The prepared Nb2O5 film had a uniform thickness and strong adhesion to the underlying material surface. Electrodeposition can also be performed using non-aqueous electrolytes, although a supporting electrolyte is typically required to facilitate current conduction in these systems.
In the study by Kamada et al., Nb served as the anode and platinum as the cathode. Electrodeposition was carried out in 0.01 M I2− or Br2− acetone as the solvent under a constant pressure of 50 V at room temperature and direct current. Introducing iodine to acetone led to anodization of the metal anode, however, electrochemical dissolution of the anode and its cathodic deposition did not occur. Conversely, including bromine in the solution promoted anodic dissolution and electrodeposition [92]. Meanwhile, Zhao et al. [93] applied the electrodeposition method to deposit T-Nb2O5 quantum dots onto a Ti nanorod array, creating a Ti@T-Nb2O5 core-shell array electrode with good electrical properties. The size of T-Nb2O5 particles can be manipulated by modifying the deposition current density. For example, with a deposition current of 6 mA cm−2, quantum dots measuring several nanometers in diameter can be synthesized. Jha et al. improved Zhitomirsky’s electrodeposition method to obtain NbOx colloids by rapidly injecting methanol-dissolved niobium salt into a hydrogen peroxide solution cooled to approximately 2 °C. This led to the production of a pure phase T-Nb2O5 film with a uniform and controllable thickness, high porosity, and good cycling efficiency [88].

2.2.5. Vapor Phase Deposition

Vapor deposition is a technique used to produce material layers by condensing vaporized source material under specific conditions. Vapor deposition can be classified into two primary categories: physical vapor deposition (PVD) and chemical vapor deposition (CVD). PVD involves evaporating solid or liquid material sources in a vacuum environment. This results in the formation of gaseous atoms, molecules, or partially ionized ions. These particles are then deposited onto the substrate surface as a gas (in a low-pressure state) or as plasma, producing films with different functions [94,95]. The PVD methods include vacuum evaporation, sputter coating [96], ion beam-assisted deposition [97,98], thermal evaporation [99], pulsed laser deposition (PLD) [100], etc. PVD technology has advanced to the point where it is possible to deposit various films, including those composed of metals, alloys, compounds, ceramics, semiconductors, and polymers, among other materials.
Sputtering is employed to prepare Nb2O5 thin films as it can use a direct current (DC) [95,101] or radio frequency (RF) [102,103,104,105] power supply in an oxygen (O2) environment with metal niobium (Nb) [106] or Nb2O5 [102,103,105] as the target, using argon (Ar) and other carrier gas preparations. By manipulating the deposition parameters, such as pressure, target-substrate distance, substrate temperature, discharge voltage, and RF power, these methods can produce Nb2O5 thin films with precisely controlled size, crystallinity, and grain size [107]. Meanwhile, post-annealing treatment can improve the crystallinity of sputtered films. However, the deposition rate of this technique is low and time-consuming. However, further investigations have revealed that using pulsed magnetron sputtering and PLD technology, Nb2O5 thin films can be deposited at tens of microns per hour while maintaining the original stoichiometric ratio of the bulk target material.
CVD is a widely-used technique involving chemical reactions of gaseous or vapor-state substances to form solid deposits at the interface between gas and solid or between two gas phases. The main difference between CVD and PVD is that chemical reactions occur during CVD deposition. CVD is a gas-phase chemical growth process wherein various raw gas materials are introduced into a reaction chamber. Chemical reactions occur among the gases, generating new materials, subsequently deposited onto the substrate surface. As such, CVD is an effective method for preparing Nb2O5, allowing precise control over the thickness, morphology, shape, and composition of Nb2O5 films by regulating reaction conditions to meet diverse application requirements. In addition, CVD technology is widely applied in fabricating thin films, particularly conformal coatings and nanostructures [108]. Various CVD techniques include atmospheric pressure chemical vapor deposition (APCVD), low-pressure chemical vapor deposition (LPCVD) [109], vapor phase epitaxy (VPE) [110], ultra-high vacuum chemical vapor deposition (UHVCVD) [111], laser-induced chemical vapor deposition (LCVD) [112], among others.
Typically, Nb2O5 films are deposited using CVD by introducing precursor chemicals like NbCl5 and pentaethoxy niobium [Nb(OC2H5)5] into a reaction chamber via a carrier gas. They undergo thermal decomposition on the surface of a heated substrate to form the desired film. O’Neill et al. reported that the growth of Nb2O5 films via CVD is significantly influenced by the substrate’s temperature [113]. Other studies using the CVD method provide an excellent alternative to the production of layered Nb2O5, especially for fiber coating applications [114]. For example, Silveira et al. exposed glass fiber to air after reacting with NbCl5, heat treating at 300 °C for 2 h, and heat treating and hydrolyzing in water to achieve a well-coated glass fiber with Nb2O5 [115].
Spray pyrolysis is another technique commonly employed to produce Nb2O5 films with various thicknesses through an aerosol-assisted CVD process [116,117]. This method features simple experimental equipment and high production efficiency. However, when depositing Nb2O5 films at low temperatures, post-annealing treatment is often required to induce crystallization and improve the structural quality of the films.
It is worth mentioning that HV-CVD is also an effective method for synthesizing Nb2O5 thin films [108]. According to Yury Kuzminykh’s research, HV-CVD technology offers unique advantages in Nb2O5 preparation, achieving deposition rates up to 500 nm/h, markedly higher than that of the MBE process. The film thickness of vegetation by this method is up to several microns. Moreover, HV-CVD requires lower substrate temperatures than MO-CVD, opening the possibility for in-situ high vacuum characterization techniques [108].

2.2.6. Thermal Oxidation

The preparation of Nb2O5 via the thermal oxidation process is typically a straightforward procedure involving using Nb metal/foil or powder as the starting material. The material is subsequently loaded into a furnace and heated in a high-oxygen or pure O2 environment to temperatures up to 1000 °C to facilitate oxidation and the formation of Nb2O5 [118,119,120]. The thermal oxidation process relies on the diffusion of oxidizing agents like O2 onto the substrate at elevated temperatures. These agents then react with the Nb material, leading to the direct creation of nanoscale Nb2O5 structures on the surface. This approach enables the fabrication of Nb2O5 nanowires, which can extend to lengths up to 20 mm [121]. The resulting nanostructures are influenced by factors such as temperature, time, metal catalysts and gas atmosphere.

3. Properties of Nb2O5

In the following section, other basic properties of Nb2O5 including electrical, optical and thermal properties will be discussed.

3.1. Electrical Properties

The properties of metal oxides play a critical role in developing electronic devices, with key parameters such as band structure, electrical conductivity, and permittivity holding particular importance. In electronic applications, niobium oxide, especially Nb2O5, the most thermodynamically stable niobium oxide, has garnered significant attention.
The dielectric constants for Nb2O5 have shown significant variation and inconsistency, ranging from 41 to 120. And its band gap spans from 3.4 to 5.3 eV [122], classifying Nb2O5 as an N-type semiconductor with a wide band gap. The conduction band is formed by the vacant 4d orbitals of Nb5+ and is approximately 0.2–0.4 eV greater than that of TiO2. [123]. Nb2O5 is frequently employed in synthesizing alkali metal niobates like MNbO3(M = Li, K, Na), commonly used in optoelectronic devices [124]. Due to its excellent dielectric properties, Ding et al. also applied it to Y2Ti2O7 microwave dielectric ceramics to improve its dielectric properties [125].
Nb2O5 exhibits varying bandgap energies (Eg) ranging from 3.1 eV (semiconducting behavior) to 5.3 eV (insulating behavior). The bandgap energy can be adjusted by adding foreign ions and other methods. The nanostructured Nb2O5 can significantly affect its electrical properties, potentially causing a blue shift in the band gap. According to Clima et al., different structures in sol-gel deposited Nb2O5 correspond to different permittivities, such as ortho, δA hexagonal, and δB hexagonal [126]. Simultaneously, their examination of the permittivity tensor reveals that Nb2O5 exhibits anisotropic dielectric behavior.
When studying the dielectric constant (ε′) of Nb2O5, different structures yield different values [127]. For instance, amorphous Nb2O5 films formed via anodic oxidation can have dielectric constants ranging from 41 to 120. Meanwhile, sintered Nb2O5 pellets exhibit dielectric constants ranging from 38 to 165 depending on the processing routes [128]. Furthermore, Graca and colleagues [124] found that T-Nb2O5 and H-Nb2O5 powder compacts exhibit dielectric constant values of approximately 80 and 17, respectively, and a mixture of these phases reached ~600. Soares and colleagues [128] discovered that the dielectric constant of sintered Nb2O5 pellets differs based on the processing method. The electrical properties of bulk Nb2O5 are closely related to its crystal structure and material density, which are related to sintering conditions [124,128].
The electrical conductivity of niobium oxide is related to the formation of electrons in oxygen vacancies, with oxygen in the formation route determining conductivity [129]. Therefore, the electrical characteristics can significantly vary based on the experimental conditions. Under non-oxygen atmospheres, the conductivity of Nb2O5 increases owing to the creation of additional oxygen vacancies. When heated in air, however, the conductivity declines at ~550 K, attributed to molecular absorption reducing vacancies. Therefore, these alterations in transmission properties also relate to oxygen vacancies. This reliance on vacancies forms the foundation of the material’s technological applications.
In addition, electron mobility is an important parameter in the development of semiconductor devices, with studies indicating that the mobility of Nb2O5 increases with an increase in temperature. Indeed, the conductivity of Nb2O5 crystal exhibits an exponential dependence on temperature [130].

3.2. Optical Properties

Recently, there has been a growing interest and increasing scientific attention toward Nb2O5 due to its versatility as a multifaceted material. It is widely recognized as a transparent oxide semiconductor material [1,131,132], which exhibits transparency in the ultraviolet region owing to its wide band gap, insolubility in water, and stability in air. Due to the interesting characteristics of Nb2O5 films, including their wide band gap, high refractive index, excellent thermochemistry, and stability, they are currently used in several applications. These applications encompass, but are not limited to, photoelectric devices, catalysis, gas sensors, high-performance oxide glasses [133], and EC devices [106]. Nb2O5 has garnered attention due to its potential applications in solar cells, batteries, photodetectors, and other electronic devices [9,134].
Le and colleagues conducted studies to determine the band gap of crystalline Nb2O5 nanofibers, including H-Nb2O5, O-Nb2O5, and M-Nb2O5, which were sintered at three different temperatures (773 K, 1073 K, and 1373 K). The band gaps for H-Nb2O5, O-Nb2O5, and M-Nb2O5 were measured at 3.85 eV, 3.77 eV, and 3.79 eV, respectively [135]. In a separate study, Abe reported the band gap values of orthotropic Nb2O5, synthesized via powder, to be 3.4 eV, while that for monoclinic Nb2O5 was 3.1 eV [136]. Furthermore, Abe studied how incorporating varying concentrations of Ge into monoclinic Nb2O5 films affected their band gap. The optical absorption edge of H-Nb2O5, without Ge doping, was measured at 3.1 eV. However, increased Ge concentration shifted the absorption edge to 3.35 eV [136]. During the early 2010s, Liu and colleagues utilized a two-step solution method to fabricate single-crystal porous Nb2O5 nanotubes. They also prepared homogeneous single-crystal Nb2O5 nanorods. The band gap values for Nb2O5 nanotubes and nanorods were 3.97 eV and 3.72 eV, respectively. The observed variation in band gap values, specifically the 0.25 eV difference between porous nanotubes and solid nanorods, can be attributed to the blue shift of the absorption edges in the former, caused by the quantum confinement effect in the hollow Nb2O5 nanotubes [137]. Brayner and colleagues similarly noticed a blue shift [138]. Agarwal and colleagues also investigated how grain size influences the electronic properties of Nb2O5. They proposed that the energy of the absorption edge and local coordination can be affected by variations in grain size [139].
As previously mentioned, Nb2O5 possesses a band gap ranging from 3.1 to 4.0 eV. Its optical properties are highly dependent on its crystallinity and grain morphology, allowing it to effectively absorb light in the UV and near-UV spectra or serve as a transparent material for ultraviolet light. Niobium oxide films exhibit electrochromism, transitioning from transparency to shades of brown, gray, or blue when Li+ or H+ plasma is introduced. Nb2O5 is capable of modulating its optical transmission state, switching between high transmittance (T ~ 85%) in a quasi-transparent state and low transmittance (<T ~ 10%) in the UV, visible, or near-infrared (IR) ranges. Moreover, the material’s coloration can vary between blue and brown, depending on its crystallinity structure [140].
The study of Nb2O5’s electrochromic properties dates back to 1980, when Reichman and Bard reported its excellent chemical stability and resistance to acid and alkali corrosion. These characteristics have prompted extensive research into its application as an electrochromic material. This electrochromic phenomenon is related to the continuous and reversible optical changes produced by electrochemistry, which macroscopically manifest as color changes. However, electrochromic materials are typically evaluated based on color rendering efficiency, encompassing transmittance contrast, response time, and chemical stability when transitioning from colored to bleached states under an electric field and charge injection. Thus, the material properties of Nb2O5 films play a significant role in determining these characteristics. Nb2O5 thin sheets exhibit a low refractive index (reported value of ~2–2.3) [141], with the degree of crystallinity influencing the refractive index, which can decrease from 2.30 to 2.20 when the material is subjected to heat treatment at 700 °C under ambient conditions [97].

3.3. Mechanical Properties

In addition to its optical and electrical properties, the mechanical strength of Nb2O5 plays a pivotal role in producing electronic devices, particularly in manufacturing actuators and flexible mechanical components. Thin films often experience stresses and strains due to limitations imposed by the deposition method or substrate.
Sputtered Nb2O5 films typically exhibit an average hardness (H) ranging from 5.6 to 6.8 GPa, with a Young’s modulus (Er) reaching 117 to 268 GPa. These mechanical properties are influenced by the crystalline structure of the film [97,142]. The MIM capacitor, incorporating a sputtered Nb2O5 film, underwent the collapse radius test, a commonly employed bending test technique. The findings indicated that the device exhibited remarkable robustness and resilience, enduring repeated testing for up to 2500 times [143].

4. Application of Nb2O5

Over the past few decades, the excessive exploitation and utilization of fossil fuels by humans has led to resource depletion and significant environmental challenges. Therefore, new energy sources and technologies are needed to alleviate the pressure on energy sources and the environment. The replacement of conventional internal combustion engines with mobile energy storage systems is crucial for phasing out fossil fuels. For these applications, it is necessary to have high energy density energy storage materials [144].
Lithium-ion batteries (LIBs) and supercapacitors (SCs), the two main devices for storing electric energy, are vital to daily life. LIBs are widely used as the primary energy supply for handheld devices, such as mobile phones and laptops. Additionally, they have gained widespread adoption in electric vehicles, garnering significant global attention [145,146]. Moreover, due to the unique electrical properties of Nb2O5, extensive research has been conducted on its applications in LIBs and SCs. Researchers have achieved performance optimization in both LIBs and SCs by modifying the properties of Nb2O5 thin films. The various applications of Nb2O5 reported in the literature are described in the following sections.

4.1. Lithium-Ion Battery

Batteries and HSCs are crucial in electric vehicles, smartphones, and more, driving advancements in related industries. Developing sustainable and high-capacity anode materials is a key parameter of high-efficiency battery technology. Graphite (Gr) is the most widely utilized anode material in LIBs [147]. Carbon nanotubes and graphene also hold significance as anode materials [148]. However, these carbonaceous materials are limited by high production costs, complex, large-scale manufacturing processes, and unsustainable routes.
Consequently, it is imperative to explore the development of novel anode materials that are environmentally friendly and cost-effective [149]. Such advancements are essential to ensuring the sustainable growth of battery technology [150]. This push toward developing affordable and sustainable alternatives highlights the need for ongoing research and innovation in this field [151].
Recently, there has been significant interest in utilizing niobium oxide-based materials and their composites in LIBs. This is due to the unique properties that render them particularly suitable for this application, including their quasi-two-dimensional networks, ion insertion/extraction, abundant REDOX chemistry, and excellent chemical stability. Therefore, this discussion delves into various aspects of Nb2O5 application, including their classification, advantages, disadvantages, and research progress.
The anode materials for LIBs are assigned to three primary classifications: alloy, conversion, and insertion types [152]. While the specific capacity of insertion-type electrode material is theoretically lower than the other two, it possesses many advantageous properties that make it an ideal anode material for LIBs. In particular, it exhibits excellent cycling stability and the ability to maintain high capacity under high power conditions, critical factors for determining the performance of a battery.
The insertion lithium storage mechanism exhibited by niobium (Nb)-based oxides positions them as highly promising candidates for replacing high-rate negative electrode materials [153]. The crystal structure of T-Nb2O5 comprises two alternating atomic layers, namely the loosely arranged 4g layer and the densely arranged 4h layer (Figure 7a). Due to the large atomic space, the 4g layer is a preferred storage and transport site for lithium ions. Theoretical calculations reveal that the lithium ions occupying the 4g layer are relatively stable. Based on the neighboring Nb-O bond structure in T-Nb2O5, two distinct Li-ion diffusion pathways exist: Path A and Path B (Figure 7b). These diffusion pathways offer direct transport channels, ample transport space, and low-energy barriers, facilitating rapid Li-ion transportation [36].
First, most Nb-based oxides operate within a potential range from 1.0 to 2.0 V (relative to Li+/Li), with REDOX reactions involving Nb5+/Nb4+ and Nb4+/Nb3+ being the primary mechanisms. This operating potential range is selected to mitigate electrolyte deposition and the formation of lithium dendrites, critical for ensuring battery safety under high current densities and overcharging conditions. Second, compared to LTO, Nb oxides demonstrate a higher specific capacity, typically ranging from 200 to 400 mAh g−1. This enhances the appeal of Nb-based oxides as negative electrode materials for LIBs. Third, Nb oxide has a high Li+ diffusion coefficient, attributable to its unique structure, making it an attractive option as a negative electrode material in LIBs, as it can improve the battery’s overall performance [99,154]. Fourth, the low volume expansion rate of Nb oxide reduces the risk of structural damage caused by the high levels of expansion and contraction during lithium insertion. This property increases the stability and longevity of the battery (Figure 8) [155].
In addition, even at micron sizes, monoclinic Nb oxide possesses a remarkable energy storage capacity [22,157]. The findings from these studies provide strong evidence that Nb-based oxide materials are highly advantageous as alternative anodes in high-performance LIB applications. In particular, they boast suitable operating potentials, high specific capacities, high Li+ diffusion coefficients, and low volume expansion rates. Indeed, Nb-based oxides have the potential to play a major role in improving the performance and safety of LIBs for future energy storage applications.

4.1.1. Lithium-Ion Battery Anode Performance

There has been a recent increased focus on Nb-based oxide materials, particularly Nb2O5, as candidates for various energy storage applications. These materials have been thoroughly investigated for their potential utilization in technologies such as LIBs, SIBs, and SCs [158,159]. This heightened interest can be attributed to the unique set of advantages that these materials offer, including a high specific capacity, high Li+ diffusion coefficient (Figure 9), low volume expansion rate, and stable cycling performance. As such, Nb-based oxides are widely recognized as a promising avenue for developing advanced energy storage systems. In addition, the high operating voltage of Nb oxides generally exceeds 1.0 V vs. Li+/Li, representing a critical advantage that mitigates the decomposition of organic electrolytes and the development of solid–electrolyte–interface (SEI) films in batteries. These complications can arise from the instability of certain materials at lower voltages, leading to unwanted reactions or side effects that limit the performance and lifespan of the battery. Thus, the high operating voltage of Nb-based oxides can enhance the safety and efficiency of energy storage systems. Rapid energy storage and release are widely regarded as fundamental technologies for developing next-generation battery systems. Achieving fast energy storage performance can be realized through the use of advanced electrode materials, novel electrolyte formulations, or innovative cell designs. In particular, the utilization of Nb-based oxides has shown promise in enhancing the energy storage performance of batteries, owing to their high specific capacity, low volume expansion rate, and excellent cycling stability (Figure 8). By incorporating these materials into next-generation batteries, it is possible to achieve faster charging times, longer lifespans, and higher energy densities, ultimately paving the way toward more efficient and sustainable energy storage technologies [160].
Electrode active materials and their properties are crucial in influencing the overall performance of LIBs as they are essential in facilitating the reversible flow of ions during charging and discharging cycles [162]. In the early 1980s, Nb2O5 was first studied as a LIB material [163,164]. A fundamental discovery was made by Reichman et al., who reported that Nb2O5 could intercalate with Li+ ions, suggesting the possibility of reversible charging and discharging [163]. Kumagai’s group further explored the use of Nb2O5 as a cathode material for LIBs, employing advanced techniques, including XPS, XRD, and XAFS. Their investigations focused on characterizing the structural changes in the Nb2O5 cathode during charging and discharging and elucidating the underlying mechanisms governing its electrochemical behavior [165]. Among different crystal structures, tetragonal-Nb2O5 exhibits the most favorable cycling performance. In fact, it can achieve a discharge capacity of up to 190 mAh g−1 and sustain up to 30 cycles. XRD analysis revealed that the orthorhombic and tetragonal structures of Nb2O5 maintain their original lattice despite experiencing slight changes in cell volume following Li+ embedding. Meanwhile, the unique two-dimensional and layered structure of tetragonal Nb2O5 allows it to accommodate very high concentrations of embedded ions [165]. In 2011, researchers, led by Goodenough proposed TiNb2O7 as a notable material within the TiO2-Nb system, demonstrating its potential as a negative electrode material for LIBs [166]. In 2013, Dunn’s group reported that orthorhombic Nb2O5 (T-Nb2O5) exhibits exceptional electrochemical energy storage capabilities at high rates, primarily due to the utilization of Li+ embedded pseudocapacitance [28]. Following this discovery, Nb2O5-related anode materials regained significant attention due to their promising application prospects in fast-charge energy storage devices.
Research on Nb2O5 LIBs has also included exploring various electrolytes and electrode compositions with varying amounts of graphite. Reducing the graphite content within the electrode composition leads to a decrease in discharge capacity. Furthermore, researchers have explored the application of sputtered Nb2O5 as a LIB electrode. For example, Nakazawa and colleagues reported a battery featuring a sputtered Nb2O5 negative electrode with a thickness of 100 nm, achieving highly favorable charging and discharging performance, maintaining a capacity range of 310–380 mA h cm−3 over 500 cycles [167].

4.1.2. Lithium Storage Mechanism

The electrochemical behavior of Nb2O5 electrode materials, which exhibit distinct crystal structures resulting from variations in temperature and pH value, demonstrates noticeable differences. These distinctions can be attributed to the specific material processing method employed and the underlying principles of the fuzzy reaction. Further understanding of the reaction mechanism can be obtained by exploring the storage of Li+. Bard et al. conducted the primary investigation on the electrochemical behavior of Li+ in Nb2O5 [101]. They found that the Nb2O5 electrode exhibited electrical conductivity in an acetonitrile solution containing 0.8 M LiClO4, and an electrochemical redox process occurred between Nb2O5 and Li+. Hence, Li+ could undergo a reversible reaction with Nb2O5, resulting in the formation of LixNb2O5. The redox reaction mechanism can be represented by the following equation:
N b 2 O 5 + x L i + + x e L i x N b 2 O 5 x = 0 2
When x = 2, it corresponds to a maximum theoretical capacity of 200 mA h g−1.
Typically, the theoretical capacity of Nb2O5 is determined using Equation (2), which is derived from the two-electron redox reaction, as follows:
Q t h e o r e t i c a l = n F M W = 2 × 96485.3   C   m o l 1 265.8   g   m o l 1 = 725.9   C   g 1 = 201.6   m A   h   g 1
where “n” represents the number of electrons involved in the reaction, “F” denotes the Faraday’s constant, “Mw” signifies the molecular weight of the active material, and 3.6 serves as the conversion factor between coulombs (C) and conventional milliampere-hours (mA h).
Among the various configurations of Nb2O5, the pseudo-hexagonal system TT-Nb2O5, orthorhombic system T/O-Nb2O5, and the monoclinic system H-Nb2O5 have undergone extensive investigation as anode materials for energy storage systems, each with distinct mechanisms for storing lithium. In 1999, Kumagai and colleagues [168] evaluated the electrochemical behavior of three different Nb2O5 configurations obtained at different heating temperatures. These configurations were designated the hexagonal, orthogonal, and monoclinic phases. The heating temperature during the synthesis process significantly impacted the crystal structure and properties of Nb2O5. They obtained Nb2O5 samples with different crystal structures by controlling the heating temperature. The Nb2O5 compounds exhibited comparable electrical activity, demonstrating discharge capacities ranging from 160 to 180 mA h g−1 within the potential range of 1.2 to 3.0 V (vs Li+/Li).
Hexagonal and orthorhombic Nb2O5 electrodes do not exhibit distinct potential plateaus during charging and discharging. Instead, the potential varies linearly with the x value of LixNb2O5 (Figure 10a,b). On the contrary, the charge–discharge curve of the monoclinic Nb2O5 electrode displays multiple potential platforms (Figure 10c). Based on the research findings, the hexagonal and orthorhombic Nb2O5 undergo a single ternary phase formation of LixNb2O5 while embedding Li+. Meanwhile, the lithiation process of monoclinic Nb2O5 involves two-phase reactions, transitioning from x = 0.8 to x = 1.8 within LixNb2O5.
The crystal structure of T-Nb2O5, characterized by its NbO6 and NbO7 polyhedra and the irregular arrangement of Nb and O ions, significantly influences its properties [28], as discussed in Section 2.1.1. Chen et al. [36] investigated the rapid lithium storage mechanism of T-Nb2O5 by combining theoretical calculations and experimental measurements. The T-Nb2O5 structure consists of alternating atomic layers that exhibit two distinct arrangements. Approximately 40% of the O2− ions occupy the 4g Wyckoff site, resulting in relatively loosely packed 4g layers. The remaining ions occupy the 4h Wyckoff site, forming more densely packed 4h layers. All niobium ions are situated in the relatively dense 4h layer, while lithium ions are positioned in the loosely packed 4g layer, providing space to accommodate Li+ ions, while the dense 4h layer facilitates the formation of a well-defined bridge coordination between oxygen and Li+ ions. This unique arrangement of bridging sites enables the formation of rapid Li+ diffusion paths with low migration barriers [36].
Kodama and colleagues demonstrated that a continuous transition from Nb5+ to Nb4+ occurs during the insertion of Li+ ions. Additionally, the cell volume of orthorhombic Nb2O5 undergoes only slight changes after the inclusion of Li+ ions [165]. According to Dunn et al., the preferential insertion of Li+ ions occurs through the (180) and (001) planes in T-Nb2O5. Furthermore, the redox reaction between Li+ ions and T-Nb2O5 occurs within a single-phase system [169], where the {001} group planes form a channel conducive to Li+ migration. Therefore, the orthorhombic T-Nb2O5 system boasts a unique crystal structure that features a two-dimensional Li+ channel unaffected by Li+ reincarceration behavior. This facilitates rapid Li removal in T-Nb2O5. Furthermore, theoretical calculations indicate that the (001) surface of T-Nb2O5 facilitates the transport of ions along degenerate paths with low energy barriers due to the large oxygen–oxygen distance of 3.9 Å [168,170]. In addition, the unique “room-and-pillar” NbO6/NbO7 framework of T-Nb2O5 offers a stable structure for embedding Li+ ions, ensuring that no phase transitions occur during the ion insertion process [22]. Nb2O5‘s special layered structure allows bulk T-Nb2O5 to facilitate rapid Li+ diffusion, exhibiting excellent intercalation pseudocapacitance [165].
T-Nb2O5 and TT-Nb2O5 are considered pseudocapacitor materials due to their similar crystal structure [155]. However, as TT-Nb2O5 possesses a more disordered crystal structure, it encounters more structural obstacles during Li+ embedding [171]. T-Nb2O5 exhibits a higher reversible specific capacity than TT-Nb2O5, leading to a noticeable difference in their performance.
The lithium storage mechanism of H-Nb2O5 differs from that of T-Nb2O5 and TT-Nb2O5, showcasing distinct characteristics in their respective behaviors. In contrast to T-Nb2O5 and TT-Nb2O5, monoclinic H-Nb2O5 features a densely packed oxygen array and exhibits a more significant repulsion intercalation effect. As previously mentioned, the extensive intercalation of lithium within the H-Nb2O5 structure can result in phase separation [168]. This phenomenon is generally recognized as the primary factor contributing to the inferior rate performance observed in numerous studies when comparing H-Nb2O5 to T-Nb2O5. However, Ding et al. [160] found that the performance failure of H-Nb2O5 was attributable to the anisotropy of electron and ion conduction within the H-Nb2O5 crystals. This asynchronous phase transition during lithium (de)intercalation is a direct consequence of the phase separation, further impacting the performance of H-Nb2O5 compared to T-Nb2O5 [172]. Researchers have employed a strategy involving the application of uniform amorphous carbon shells as a coating on the surface of micron-scale single-crystal H-Nb2O5 particles. This serves to homogenize electron and Li+ transport. This optimization technique significantly enhanced the fast lithium storage performance of the negative electrode composed of H-Nb2O5, surpassing the performance of many reported negative electrodes based on LTO and T-Nb2O5.
In conclusion, the inherent structural advantages of Nb2O5 have attracted considerable attention in advanced electrochemical energy storage applications. For example, the high operating voltage of Nb2O5 (>1.0 V vs. Li+/Li) can inhibit the decomposition of the electrolyte and the formation of SEI film and lithium dendrites, ensuring the safety of the battery. Continuous and high-demand research efforts will be focused on exploring more effective strategies for enhancing the conductivity of niobium-based oxides. Structural optimization, surface engineering, and carbon modification remain the primary avenues for investigating this field. Meanwhile, the composite utilization of anode electrode materials with high theoretical capacity and synergistic effects is significant.

4.1.3. Effect of Nb2O5 Nanostructures

Zero-dimensional (0D) Nb2O5 nanostructures, specifically large-sized nanoscale particles, have been the subject of recent studies. These investigations have primarily focused on exploring the unique properties and applications of such nanostructures. However, it is worth noting that the repetitive usage of these large-sized 0D Nb2O5 particles may lead to a decrease in performance or capacity due to potential degradation or other cycling-related factors [173,174,175,176]. The uniform distribution of particles within a battery is essential as it provides a large surface area and abundant reaction sites. This prevents particle agglomeration during the charging and discharging process, ultimately improving the electrochemical performance of the battery. The amorphous Nb2O5 precursors prepared in the air by the sol-gel method were calcined at different temperatures to obtain different crystalline phases of Nb2O5 (amorphous, T, TT). The grain size of these phases (hexagonal, orthorhombic, and monoclinic) increases with rising temperature. Extensive testing has demonstrated that Nb2O5 exhibits remarkable cyclic stability. In addition, Liu and colleagues [175] synthesized T-Nb2O5−x particles using a straightforward sol-gel process, followed by post-etching and calcination. However, a small portion of cell recombination occurred due to the removal of Sr and Ca atoms during the acid etching process. The significant stress associated with this process leads to the formation of high-density defects, contributing to the generation of more reaction sites so that the final sample has good magnification performance and cycle stability. Chen et al. conducted a solvothermal method to obtain Nb2O5 nanomaterials with varying graphene contents (1, 2, 3, 4, and 5 wt.%) and superphosphorus as conductors [176].
One-dimensional Nb2O5 nanostructure: Using pure Nb2O5 with one-dimensional nanostructures as an anode material for LIBs is not widespread. This may be attributed to the intricate preparation requirements and challenging control of the associated processes. Most research efforts in one-dimensional nanostructures have predominantly focused on nanofibers [52] and nanorods [48]. Despite the challenges associated with their preparation, the distinctive aspect ratio of pure Nb2O5 in one-dimensional nanostructures offers an effective pathway for transmitting Li+ ions. This unique characteristic continues to garner significant interest within the research community. In some instances, regulating heating rates during the heat treatment is considered an effective approach to fabricating nanofibers with exceptional properties.
Two-dimensional Nb2O5 nanostructure: Generally, two-dimensional Nb2O5 nanomaterials used for LIBs include nano-thin films [177], nano-sheets [178,179] and nanoribbons [180,181]. One advantage of controlling the heating rate during the treatment process is its ability to create a continuous framework that facilitates efficient ion and electron diffusion; more Li+ storage sites are on the open edge. Consequently, by optimizing reaction kinetics, controlling the heating rate can enhance the overall performance of these materials. Furthermore, the anisotropic growth of two-dimensional nanomaterials offers additional opportunities for REDOX reactions in batteries. Nano-thin film electrode materials can be prepared through physical or chemical deposition methods. Due to the large energy density area, two-dimensional flexible nanofilms grown directly on the substrate without additives and conductive agents can be used in micro-energy storage devices.
In the study conducted by Zhou et al. [182], metallic Nb powder served as a precursor for synthesizing Nb2O5 nanoribbons. Subsequent investigation primarily focused on evaluating the electrochemical characteristics of these nanoribbons when employed as anode materials in LIBs. Liu et al. successfully fabricated Nb2O5 nanosheets with dimensions of 50 nm in thickness and 500 nm in length through a hydrothermal reaction [178]. The excellent rapid energy storage capacity of these two-dimensional Nb2O5 nanostructures can be attributed to their large active surface area and short Li+ diffusion distance.
In addition to the zero-dimensional (0D), one-dimensional (1D), and two-dimensional (2D) structures, Nb2O5 exhibits various heterogeneous multi-level complex structures. Notable examples include hollow porous spheres [183], sea urchin-shaped spheres [184,185,186], micro-structures composed of nanoparticles [187], 3D pore frames [188,189], and self-assembled nanosheets [190]. The casting process employed to generate complex structures, such as hollow porous spheres in Nb2O5, offers several advantages that compensate for the specific advantages that individual structures alone cannot achieve. The casting method allows for precise control over the structure’s shape, size, and composition, enabling enhanced properties such as improved structural stability, increased surface area, and optimized electrolyte penetration. The synergy created by incorporating various structures within a single material can lead to superior performance in various applications. Indeed, complex microsphere structures in Nb2O5 can be synthesized using relatively straightforward methods, such as precipitation, hydrothermal, and solvothermal techniques. Sun et al. [183] showed that a simple oil bath process can also generate nanostructures. They designed a technique for preparing hollow mesoporous Nb2O5 (HM-Nb2O5) nanospheres with a diameter of ~300 nm by calcination at 600 °C. The thickness of the nanosphere shell was achieved by extending the reaction time caused by the controlled hydrolysis of urea (Figure 11a). Liu et al. successfully synthesized sea urchin Nb2O5 microspheres with a 20 nm diameter in a glycerol-isopropyl alcohol mixture using a simple solvothermal method through the self-assembly of nanorods (Figure 11b) [187]. Lou et al. [188] achieved successful synthesis of three-dimensionally ordered macroporous (3DOM) T-Nb2O5 nanostructures using self-assembled polystyrene (PS) microsphere colloidal crystals as a hard template. The process involved immersing the template in a precursor solution to form the desired T-Nb2O5 nanostructures. This approach yielded a substantial quantity of tightly packed macropores measuring 200 nm in diameter. The surface of the synthesized material exhibited many additional macropores, ranging from 60 to 130 nm, along with mesoporous structures (as illustrated in Figure 11c,d).
Overall, the 0D nanostructure offers a significant specific surface area that enhances the contribution of pseudocapacitors. However, it is limited by particle agglomeration resulting from repeated high-rate charging and discharging, which negatively impacts electrode lifespan, contradicting its intended purpose. In the case of 1D nanomaterials grown along (001) surfaces, their unique aspect ratio is crucial in facilitating the rapid transmission of Li+ ions, significantly enhancing Li storage dynamics. Nevertheless, the inherent stable chemical properties of Nb2O5 contribute to the complexity and lack of control observed during the preparation process. Generally, 2D nanostructures with expandable anisotropic ion shuttle paths and open edges offer larger Li attachment sites, enabling higher capacity storage in LIBs. Researchers favor complex nanostructures, including heterogeneous multistage mesoporous structures, due to their inherent versatility and ability to address specific requirements that cannot be met by a single structure alone. This versatility extends to promoting solvent diffusion and fulfilling diverse application needs. However, complex nanostructures are also constrained by their structural durability deficiencies, resulting in a low volume packing density. Therefore, an increasingly popular approach to address these limitations involves optimizing structural engineering through the formation of composite materials (Figure 12).
Conclusion: Nb oxide plays a significant role in energy storage thanks to its distinctive crystal structure, similar to other important materials, and its exceptional chemical stability. Among its various forms, Nb2O5 offers clear advantages over commercial graphite anode materials. Nb2O5 boasts high power, enhanced safety, facilitation of abundant REDOX reactions, and minimal volume expansion. These advantages contribute to its potential for commercial applications, particularly in small batteries and the intercalated pseudocapacitor lithium storage mechanism. However, it is essential to acknowledge that while employing a high-voltage window can prevent the formation of lithium dendrites, it can also lead to a trade-off in terms of lower energy density. As a result, Nb2O5 faces many challenges in its electrochemistry application due to its low conductivity as a semiconductor material and harsh synthesis conditions.

4.2. Supercapacitor

Electric vehicles are witnessing a significant shift toward using batteries and HSCs, with applications in consumer electronics, large power stations, and hybrid electric vehicles (HEVs) [191,192]. SCs offer distinct advantages compared to batteries, including significantly higher power density, longer cycle life, and rapid charging capabilities. However, the primary disadvantage of SCs lies in their relatively low energy density due to limitations in electrode size, restricting their storage capacity. One potential strategy to address these challenges is to focus on developing high-energy-density electrodes.
The primary benefits of Nb2O5-based material being used in SCs and LIBs can be summarized as follows: (1) Nb2O5-based materials will not precipitate metal Li due to their high potential platform; (2) Nb2O5-based materials are non-toxic with higher specific capacity than Li4Ti5O12; (3) Nb2O5-based material exhibit high thermodynamic stability, ensuring good safety performance of LIBs; (4) Nb2O5-based materials feature fast electrochemical reaction kinetics and long cycle life.
However, it has a drawback in its inherent poor electronic conductivity. In addition, Nb2O5-based materials lack a flat charge-discharge platform and possess a lower theoretical capacity than other alloy-type or converted compounds [153,193]. Furthermore, a power capacity imbalance exists between fast non-Faraday capacitive cathodes and slow Faraday battery-type anodes [30]. Despite these limitations, Nb2O5-based materials hold promise in EESC devices due to their unique structural advantages, strong rate performance, and cycle stability.

4.2.1. High Rate Electrode

Kim et al. conducted a study to assess the impact of crystallinity on the capacitance response of Nb2O5 for SCs [171]. Meanwhile, Augustyn and colleagues demonstrated the utilization of electrodes with a thickness of up to 40 mm to achieve high-speed charge storage devices through intercalated pseudocapacitors [28]. Lubimtsev et al. conducted an additional study to investigate the underlying factors contributing to the high-rate behavior of intercalated pseudocapacitors in Nb2O5 crystals [194]. An effective approach to address these challenges to developing high-energy-density electrodes is combining conductive carbon materials, doping techniques, and morphology control in the design of composite materials [195].
Lim and colleagues [30] employed a general process to prepare core-shell T-Nb2O5@carbon nanocrystals (T-Nb2O5/C NCs) and TT-Nb2O5/C composite materials. In a study by Lim et al. [196] the anode was fabricated using mesoporous T-Nb2O5@carbon nanocomposites. The resulting HSCs, assembled with activated carbon (AC) as the cathode, exhibited an impressive energy density of 74 Wh kg−1 and a power density of 18.51 kW kg−1 (Figure 13a). After subjecting the HSCs to 1000 cycles within the voltage range of 1–3.5 V at a current density of 1 A g−1, the capacity retention rate remained approximately 90%. This indicates that the electrode material exhibited commendable cycling stability and retained a significant portion of its initial capacity even after prolonged use. In another study, Wang et al. [197] utilized thin carbon-coated T-Nb2O5 nanowires to assemble 3 V T-Nb2O5@C||AC HSCs. The HSCs exhibited a prominent energy density of ~43 Wh kg−1 at 7.5 kW kg−1 and significant cycle stability (Figure 13b). Wang and colleagues developed an SC electrode comprising a closely mixed network of carbon nanotubes (CNTs) and Nb2O5 nanocrystals. This innovative electrode design aimed to achieve SCs with high capacitance, excellent rate performance, and cycling capability [198]. Furthermore, Zhang and colleagues [199] successfully prepared T-Nb2O5@carbon (Nb2O5@C) and T-Nb2O5@mesoporous carbon (Nb2O5@MC). Both samples were coated with porous carbon shells, enhancing structural stability and electrochemical performance (Figure 13c,d). The improved electrochemical performance of Nb2O5@MC can be attributed to the Nb2O5 nanoparticles with a high specific surface area (SSA), enabling more charge storage capacity. The interlinked mesoporous carbon shells also facilitate efficient ion diffusion and enhance charge transfer kinetics within the electrode material.
Recently, researchers have been exploring using metal-organic frameworks (MOFs) as a promising platform for preparing porous carbon materials and porous tetramethyl orthosilicate (TMOs). MOFs are renowned for their high porosity, large specific surface area, and well-defined porous structures. These properties make MOFs an attractive choice for creating materials with desirable porosity and surface area, with applications in diverse fields, including energy storage and catalysis. For example, Liu et al. [200] prepared T-Nb2O5 quantum dots (QDs) encapsulated in N-doped porous carbon derived from ZIF-8, referred to as NQD-NC. The resulting T-Nb2O5 QD was a unique composite material with enhanced properties. Indeed, the design of carbon nanotube (CNT) composites has emerged as an effective strategy to enhance the electrochemical performance of electrode materials. CNTs possess high electron conductivity and a large specific surface area, making them ideal for improving the performance across various electrochemical reactions. The high electron conductivity of CNTs facilitates the rapid movement of electrons within the composite, enabling efficient transfer during electrochemical processes. This promotes faster reaction kinetics and can lead to enhanced electrochemical performance [201]. For instance, Wang [202] and colleagues prepared T-Nb2O5 nanocrystals grown in situ on carbon nanotubes. Graphene, known for its substantial specific surface area, impressive flexibility, superconductivity, and wide voltage window, is being extensively investigated as a potential electrode material for SCs [203]. The doping process can be utilized to improve the electronic conductivity of materials by reducing the band gap, a feature particularly crucial in the context of SCs [204,205].
In summary, the challenge of low electrical conductivity in Nb2O5 can be effectively addressed through two common strategies: fabricating nanostructures with diverse morphologies and integrating Nb2O5 with carbon-based materials. These approaches have demonstrated significant effectiveness in overcoming the conductivity limitations associated with Nb2O5.

4.2.2. Hybrid Supercapacitors

In contrast to rechargeable batteries, SCs exhibit an extended operational lifespan and capacity to facilitate rapid charge and discharge processes. These qualities contribute to their high power density and long cycle life. However, SCs typically possess a lower energy density than rechargeable batteries. Meanwhile, LIBs are known for their high energy density, providing significant stored energy. However, they are often limited in power density and cycle life. A hybrid supercapacitor (HSC) comprising a positive metal oxide electrode with pseudocapacitance behavior and a negative activated carbon electrode with EDLC behavior is an effective method to improve the energy density of EDLC.
The operation principle of this HSC involves reversible non-Faraday reactions, such as the adsorption of Li+ and Na+ ions on the surface of active and porous carbon materials at the positive electrode, facilitating charge storage. Simultaneously, reversible Faraday reactions take place at the metal oxide electrodes. SCs adopting this asymmetric configuration can accumulate charge utilizing the Faraday REDOX electrochemical reaction process. This enhances the specific capacitance and extends the operating voltage range, ultimately improving the energy density of the hybrid capacitor.
Previously, extensive research on RuO2 as an electrode material for HSCs has been primarily driven by its excellent specific capacitance, cyclic stability, and conductivity [206]. Nonetheless, the adoption of RuO2 as an electrode material for HSCs is hindered by its high cost, which restricts its widespread application.
Recent studies have revealed that bulk Nb2O5 has a higher capacity than the conventional Li4Ti5O12 under high magnification conditions. This suggests that Nb2O5 exhibits pseudocapacitive electrochemical characteristics. This phenomenon is explicitly observed on the surface of Nb2O5 rather than throughout the entire bulk crystal.
DFT calculations have indicated that forming a solid solution by incorporating lithium atoms at specific sites enables the selective provision of electrons to adjacent atoms. This results in the reduction of niobium within that region. The observed high specific capacitance in the case of Nb2O5 may be attributed to its nanoparticle structure. The presence of nanoparticles provides a larger surface area for electrochemical reactions, resulting in enhanced pseudocapacitance and overall improved performance at high magnification [194]. These results also provide directions for improving the performance of nanoporous Nb2O5.
Among the various crystal structures of Nb2O5, orthogonal Nb2O5 exhibits the highest relative capacity. However, although the nanoparticle structure of Nb2O5 contributes to its exceptional performance at high magnifications, the formation of this structure at temperatures above 600 °C can potentially lead to nanoparticle aggregation. Nano-scale rhombic system Nb2O5 is considered a favorable material for enhancing anode system dynamics. Despite the challenges associated with synthesizing small-sized rhombic Nb2O5 nanoparticles, alternative approaches have been developed to incorporate Nb2O5 into nanocomposite structures. One example is synthesizing mesoporous Nb2O5/carbon (m-Nb2O5-C) nanocomposites using a one-pot method assisted by block copolymer self-assembly [196]. Nb2O5@C NCs with controllable crystal phases, including the rhombic system (T) and pseudohexagonal system (TT), were synthesized using a one-pot microemulsion method. The pH condition of the water-in-oil microemulsion system significantly influences the control of the crystalline phase of Nb2O5 [30]. Under acidic conditions, T-Nb2O5@C nanocrystals are formed, while under alkaline conditions, TT-Nb2O5@C nanocrystals are obtained.
Therefore, utilizing Nb2O5 as an SC electrode and combining it with conductive carbon materials to improve charge transfer has emerged as a new research direction. For example, materials such as T-Nb2O5/graphene and Nb2O5/CNTs have demonstrated remarkable power density and cycling performance [198,202,207,208].
In addition, selecting suitable electrolyte solutions presents a challenge when employing Nb2O5 for SC applications. Recent research has indicated that using an electrolyte consisting of 1 M lithium perchlorate in a mixture of ethyl carbonate and dimethyl carbonate leads to the attainment of the highest capacitance and Coulomb efficiencies [208]. Moreover, although the Nb2O5 electrode has excellent lithium storage performance, due to the lack of lithium resources and the low cost and abundant sodium resources, the application of sodium ion batteries (SIB) has been widely studied recently. Meanwhile, the application of its energy storage performance in SIB has declined sharply due to its larger Na-ion radius and sluggish Na-ion diffusion. Therefore, preparing new Nb2O5 electrodes using various modification strategies to improve the electrochemical performance of SIB is an important research direction.

5. Conclusions and Perspectives

This review provided an overview of the fundamental crystal structure and physical and chemical properties of Nb2O5. Various synthesis methods for fabricating Nb2O5 polycrystals and the general applications of Nb2O5 in lithium batteries and supercapacitors (SCs) were summarized.
Nb2O5 exhibits polymorphs, including pseudohexagonal (TT-Nb2O5), orthorhombic (T-Nb2O5), and monoclinic (H-Nb2O5) etc. The crystal structures of these different phases were presented, and the research progress pertaining to their crystal structures and applications was discussed. The TT phase Nb2O5 was given particular emphasis. Various nanostructured Nb2O5, including nanopores, nanosheets, nanorods, nanochannels, and nanowires, have been successfully synthesized using different methods. However, it’s worth noting that there are still unexplored and uncharacterized nanoforms.
Similar to many other transition metal materials, pure Nb2O5 primarily exhibits poor conductivity. To address this limitation, incorporating heteroatoms, carbon, metals, and conductive polymers has proven effective in enhancing the electrical conductivity of Nb2O5. Research on forming oxygen vacancies by doping non-metallic elements in Nb2O5 remains relatively limited. Further investigations in this area are necessary to improve the electrochemical performance of such materials.
Nb2O5 plays an essential role in energy storage due to its unique crystal structure for fast charging, rich REDOX reaction capability, intercalation-based storage mechanism, high chemical stability, ultra-small volume expansion, and high potential window, which often served as electrode material for Li-ion batteries or supercapacitors. However, the electrochemical performance of Nb2O5-based electrodes still faces two key challenges. The first challenge is that Nb2O5 exhibits low conductivity and a high synthesis temperature. The second challenge is that Nb2O5 has a relatively low theoretical capacity and energy density. The T-Nb2O5 has fast rate capability with limited space for cations intercalation, while H-Nb2O5 offers more cation intercalation sites with slow rate capability. Active particles are frequently reduced to nanoscale dimensions to overcome the relatively slow solid-state ion diffusion and achieve rapid charging and high power. Optimizing the crystal phase or nanostructure of Nb2O5 is a promising approach to enhancing both rate and capacity performances. Another effective strategy involves doping other materials with Nb2O5.
In addition, Nb2O5 has been utilized as a sensing material for sensors. The synergistic integration of Nb2O5 with biomedicine has been explored to develop innovative solutions for diagnostics, therapeutics, and medical devices. Recently, Nb2O5 has been investigated for application in transistors, memristors, and superconductors. A better understanding of the Nb2O5 structure–property relationship is needed for further research efforts. This will facilitate the full harnessing of its capabilities and optimize its performance.

Author Contributions

Writing—original draft preparation, R.P.; writing—review and editing, Z.W. and K.C.; visualization, R.P. and Z.W.; supervision, J.L. and K.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the Natural Science Foundation of Shandong Province (Grant Nos., ZR2020ME045, ZR2020MEO46 and ZR2020ZD35), “New Universities20” Foundation of Jinan (Grant Nos. 2021GXRCO99, T202204).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Nowak, I.; Ziolek, M. Niobium Compounds:  Preparation, Characterization, and Application in Heterogeneous Catalysis. Chem. Rev. 1999, 99, 3603–3624. [Google Scholar]
  2. Wu, J.; Xue, D. Localized crystallization: A chemical transformation of Nb2O5 rod-like arrays into ordered niobate arrays. CrystEngComm 2011, 13, 1966–1975. [Google Scholar] [CrossRef]
  3. Zhang, B.; Deng, L.; Liebau, M.; Ren, Y.; Luo, C.; Liu, B.; Zhang, S.; Gläser, R. Promotion effect of niobium on ceria catalyst for selective catalytic reduction of NO with NH3. J. Rare Earths 2022, 40, 1535–1545. [Google Scholar]
  4. Brauer, G. Die Oxyde des Niobs. Anorg. Allg. Chem. 1941, 248, 1–31. [Google Scholar] [CrossRef]
  5. Liu, M.; Xue, D. Large-scale fabrication of H2(H2O)Nb2O6 and Nb2O5 hollow microspheres. Mater. Res. Bull. 2010, 45, 333. [Google Scholar] [CrossRef]
  6. Su, K.; Liu, H.; Gao, Z.; Fornasiero, P.; Wang, F. Nb2O5-Based Photocatalysts. Adv. Sci. 2021, 8, 2003156. [Google Scholar]
  7. Yi, T.F.; Sari, H.M.K.; Li, X.; Wang, F.; Zhu, Y.R.; Hu, J.; Zhang, J.; Li, X. A review of niobium oxides based nanocomposites for lithium-ion batteries, sodium-ion batteries and supercapacitors. Nano Energy 2021, 85, 105955. [Google Scholar]
  8. Ohzuku, T.; Sawai, K.; Hirai, T. Electrochemistry of L-niobium pentoxide a lithium/non-aqueous cell. J. Power Sources 1987, 19, 287–299. [Google Scholar] [CrossRef]
  9. Ding, J.; Hu, W.; Paek, E.; Mitlin, D. Review of Hybrid Ion Capacitors: From Aqueous to Lithium to Sodium. Chem. Rev. 2018, 118, 6457–6498. [Google Scholar]
  10. Dhawan, S.; Dhawan, T.; Vedeshwar, A.G. Residual stress induced crystalline to amorphous phase transformation in Nb2O5 quantum dots. J. Appl. Phys. 2014, 116, 043503. [Google Scholar]
  11. Schäfer, H.; Gruehn, R.; Schulte, F. The Modifications of Niobium Pentoxide. Angew. Chem. Int. Ed. 1966, 5, 40–52. [Google Scholar] [CrossRef]
  12. Ikeya, T.; Senna, M. Change in the structure of niobium pentoxide due to mechanical and thermal treatments. J. Non-Cryst. Solids 1988, 105, 243–250. [Google Scholar] [CrossRef]
  13. Huang, H.; Kuai, H.; Ding, X.; Hu, B.; Chen, Y.; Zhou, Q.; Xiong, X. Single crystal H-Nb2O5 growing along the [001] crystal direction for ultrafast lithium storage. J. Mater. Chem. A 2023, 11, 642–648. [Google Scholar] [CrossRef]
  14. Pinto, M.B.; Soares, A.L., Jr.; Quintão, M.C.; Duarte, H.A.; De Abreu, H.A. Unveiling the Structural and Electronic Properties of the B-Nb2O5 Surfaces and Their Interaction with H2O and H2O2. J. Phys. Chem. C 2018, 122, 6618–6628. [Google Scholar] [CrossRef]
  15. Mertin, W.; Andersson, S.; Gruehn, R. Über die Kristallstruktur von M-Nb2O5. J. Solid State Chem. 1970, 1, 419–424. [Google Scholar] [CrossRef]
  16. Kato, K. Structure refinement of H-Nb2O5. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1976, 32, 764–767. [Google Scholar] [CrossRef]
  17. Andersson, S. The Crystal Structure of N-Nb2O5, prepared in the presence of small amounts of LiF. Z. Anorg. Allg. Chem. 1967, 351, 106–112. [Google Scholar] [CrossRef]
  18. Israelsson, M.; Kihlborg, L. The crystal structure of monoclinic wolfram vanadium oxide, W3V5O20, an OD structure related to R-Nb2O5. J. Solid State Chem. 1970, 1, 469–477. [Google Scholar] [CrossRef]
  19. Reznichenko, L.A.; Akhnazarova, V.V.; Shilkina, L.A.; Razumovskaya, O.N.; Dudkina, S.I. Invar effect in N-Nb2O5, αht-Nb2O5, and L-Nb2O5. Crystallogr. Rep. 2009, 54, 483–491. [Google Scholar] [CrossRef]
  20. Bowman, A.; Wallace, T.; Yarnell, J.; Wenzel, R. The crystal structure of niobium monoxide. Acta Crystallogr. 1966, 21, 843. [Google Scholar] [CrossRef]
  21. Park, H.; Lee, D.; Song, T. High capacity monoclinic Nb2O5 and semiconducting NbO2 composite as high-power anode material for Li-Ion batteries. J. Power Sources 2019, 414, 377–382. [Google Scholar] [CrossRef]
  22. Griffith, K.J.; Forse, A.C.; Griffin, J.M.; Grey, C.P. High-rate intercalation without nanostructuring in metastable Nb2O5 bronze phases. J. Am. Chem. Soc. 2016, 138, 8888–8899. [Google Scholar] [CrossRef] [PubMed]
  23. Ko, E.I.; Weissman, J.G. Structures of niobium pentoxide and their implications on chemical behavior. Catal. Today 1990, 8, 27–36. [Google Scholar] [CrossRef]
  24. Pinto, M.B.; Soares, A.L.; Orellana, A.M.; Duarte, H.A.; De Abreu, H.A. Structural, Electronic, and Thermodynamic Properties of the T and B Phases of Niobia: First-Principle Calculations. J. Phys. Chem. A 2017, 121, 2399–2409. [Google Scholar] [CrossRef] [PubMed]
  25. Zhao, Y.; Zhou, X.; Ye, L.; Chi Edman Tsang, S. Nanostructured Nb2O5 catalysts. Nano Rev. 2012, 3, 17631. [Google Scholar] [CrossRef]
  26. Serghiou, G.C.; Winters, R.R.; Hammack, W.S. Pressure-induced amorphization and reduction of T-Nb2O5. Phys. Rev. Lett. 1992, 68, 3311–3314. [Google Scholar] [CrossRef]
  27. Li, L.; Deng, J.; Yu, R.; Chen, J.; Wang, X.; Xing, X. Phase evolution in low-dimensional niobium oxide synthesized by a topochemical method. Inorg. Chem. 2010, 49, 1397–1403. [Google Scholar] [CrossRef]
  28. Augustyn, V.; Come, J.; Lowe, M.A.; Kim, J.W.; Taberna, P.L.; Tolbert, S.H.; Abruna, H.D.; Simon, P.; Dunn, B. High-rate electrochemical energy storage through Li+ intercalation pseudocapacitance. Nat. Mater. 2013, 12, 518–522. [Google Scholar] [CrossRef]
  29. Kong, L.P.; Zhang, C.F.; Zhang, S.M.; Wang, J.T.; Cai, R.; Lv, C.X.; Qiao, W.M.; Ling, L.C.; Long, D.H. High-power and high-energy asymmetric supercapacitors based on Li+-intercalation into a T-Nb2O5/graphene pseudocapacitive electrode. J. Mater. Chem. A 2014, 2, 17962–17970. [Google Scholar] [CrossRef]
  30. Lim, E.; Jo, C.; Kim, H.; Kim, M.-H.; Mun, Y.; Chun, J.; Ye, Y.; Hwang, J.; Ha, K.-S.; Roh, K.C.; et al. Facile Synthesis of Nb2O5@Carbon Core–Shell Nanocrystals with Controlled Crystalline Structure for High-Power Anodes in Hybrid Supercapacitors. ACS Nano 2015, 9, 7497–7505. [Google Scholar] [CrossRef]
  31. Liu, F.F.; Cheng, X.L.; Xu, R.; Wu, Y.; Jiang, Y.; Yu, Y. Binding Sulfur-Doped Nb2O5 Hollow Nanospheres on Sulfur-Doped Graphene Networks for Highly Reversible Sodium Storage. Adv. Funct. Mater. 2018, 28, 1800394. [Google Scholar] [CrossRef]
  32. Li, N.; Zhang, F.; Tang, Y.B. Hierarchical T-Nb2O5 nanostructure with hybrid mechanisms of intercalation and pseudocapacitance for potassium storage and high-performance potassium dual-ion batteries. J. Mater. Chem. A 2018, 6, 17889–17895. [Google Scholar] [CrossRef]
  33. Wang, J.Y.; Li, G.R.; Luo, D.; Zhao, Y.; Zhang, Y.G.; Zhou, G.F.; Shui, L.L.; Wang, X. Amorphous-crystalline-heterostructured niobium oxide as two-in-one host matrix for high-performance lithium-sulfur batteries. J. Mater. Chem. A 2021, 9, 11160–11167. [Google Scholar] [CrossRef]
  34. Hu, Z.; He, Q.; Liu, Z.; Liu, X.; Qin, M.; Wen, B.; Shi, W.; Zhao, Y.; Li, Q.; Mai, L. Facile formation of tetragonal-Nb2O5 microspheres for high-rate and stable lithium storage with high areal capacity. Sci. Bull. 2020, 65, 1154–1162. [Google Scholar] [CrossRef]
  35. Liao, J.; Tan, R.; Kuang, Z.; Cui, C.; Wei, Z.; Deng, X.; Yan, Z.; Feng, Y.; Li, F.; Wang, C. Controlling the morphology, size and phase of Nb2O5 crystals for high electrochemical performance. Chin. Chem. Lett. 2018, 29, 1785–1790. [Google Scholar] [CrossRef]
  36. Chen, D.; Wang, J.-H.; Chou, T.-F.; Zhao, B.; El-Sayed, M.A.; Liu, M. Unraveling the nature of anomalously fast energy storage in T-Nb2O5. J. Am. Chem. Soc. 2017, 139, 7071–7081. [Google Scholar] [CrossRef]
  37. Zhang, S.; Liu, G.; Qiao, W.; Wang, J.; Ling, L. Oxygen vacancies enhance the lithium ion intercalation pseudocapacitive properties of orthorhombic niobium pentoxide. J. Colloid Interface Sci. 2020, 562, 193–203. [Google Scholar] [CrossRef]
  38. Holtzberg, F.; Reisman, A.; Berry, M.; Berkenblit, M. Chemistry of the Group VB Pentoxides. VI. The Polymorphism of Nb2O5. J. Am. Chem. Soc. 1957, 79, 2039–2043. [Google Scholar] [CrossRef]
  39. Shafer, M.W.; Roy, R. The Polymorphism of Nb2O5. Z. Kristallogr Cryst. Mater. 1958, 110, 241–248. [Google Scholar] [CrossRef]
  40. Terao, N. Structures des Oxydes de Niobium. Jpn. J. Appl. Phys. 1963, 2, 156. [Google Scholar] [CrossRef]
  41. Tamura, S. High-pressure phase research on Nb2O5. J. Mater. Sci. 1972, 7, 298–302. [Google Scholar] [CrossRef]
  42. Weissman, J.G.; Ko, E.I.; Wynblatt, P.; Howe, J.M. High-resolution electron microscopy and image simulation of TT-,T-, and H-niobia and model silica-supported niobium surface oxides. Chem. Mater. 1989, 1, 187–193. [Google Scholar] [CrossRef]
  43. Košutová, T.; Horák, L.; Pleskunov, P.; Hanuš, J.; Nikitin, D.; Kúš, P.; Cieslar, M.; Gordeev, I.; Burazer, S.; Choukourov, A.; et al. Thermally-driven morphogenesis of niobium nanoparticles as witnessed by in-situ X-ray scattering. Mater. Chem. Phys. 2022, 277, 125466. [Google Scholar] [CrossRef]
  44. Taques Tractz, G.; Staciaki da Luz, F.; Regina Masetto Antunes, S.; do Prado Banczek, E.; Taras da Cunha, M.; Rogério Pinto Rodrigues, P. Nb2O5 synthesis and characterization by Pechini method to the application as electron transport material in a solar device. Sol. Energy 2021, 216, 1–6. [Google Scholar] [CrossRef]
  45. Schäfer, H.; Schulte, F.; Gruehn, R. Neue Nb2O5-Modifikationen. Angew. Chem. 1964, 76, 536–537. [Google Scholar] [CrossRef]
  46. Uyeda, N.; Fujiyoshi, Y.; Ishizuka, K. High resolution imaging and interpretation of regular and irregular structures in N-niobium pentoxide crystal. Ultramicroscopy 1984, 15, 139–149. [Google Scholar] [CrossRef]
  47. Barnes, P.; Zuo, Y.; Dixon, K.; Hou, D.; Lee, S.; Ma, Z.; Connell, J.G.; Zhou, H.; Deng, C.; Smith, K. Electrochemically induced amorphous-to-rock-salt phase transformation in niobium oxide electrode for Li-ion batteries. Nat. Mater. 2022, 21, 795–803. [Google Scholar] [CrossRef]
  48. Shi, C.; Xiang, K.; Zhu, Y.; Zhou, W.; Chen, X.; Chen, H. Box-implanted Nb2O5 nanorods as superior anode materials in lithium ion batteries. Ceram. Int. 2017, 43, 12388–12395. [Google Scholar] [CrossRef]
  49. Wang, Y.; Xin, F.; Yin, X.; Song, Y.; Xiang, T.; Wang, J. Arginine-Assisted Hydrothermal Synthesis of Urchin-like Nb2O5 Nanostructures Composed of Nanowires and Their Application in Cyclohexanone Ammoximation. J. Phys. Chem. C 2018, 122, 2155–2164. [Google Scholar] [CrossRef]
  50. Zhou, Y.; Qiu, Z.; Lü, M.; Zhang, A.; Ma, Q. Preparation and characterization of porous Nb2O5 nanoparticles. Mater. Res. Bull. 2008, 43, 1363–1368. [Google Scholar] [CrossRef]
  51. Ristić, M.; Popović, S.; Musić, S. Sol–gel synthesis and characterization of Nb2O5 powders. Mater. Lett. 2004, 58, 2658–2663. [Google Scholar] [CrossRef]
  52. Cheong, J.Y.; Jung, J.-W.; Youn, D.-Y.; Kim, C.; Yu, S.; Cho, S.-H.; Yoon, K.R.; Kim, I.-D. Mesoporous orthorhombic Nb2O5 nanofibers as pseudocapacitive electrodes with ultra-stable Li storage characteristics. J. Power Sources 2017, 360, 434–442. [Google Scholar] [CrossRef]
  53. Yan, C.; Nikolova, L.; Dadvand, A.; Harnagea, C.; Sarkissian, A.; Perepichka, D.F.; Xue, D.; Rosei, F. Multiple NaNbO3/Nb2O5 heterostructure nanotubes: A new class of ferroelectric/semiconductor nanomaterials. Adv. Mater. 2010, 22, 1741–1745. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, F.; Xue, D. Controlled fabrication of Nb2O5 hollow nanospheres and nanotubes. Modern Phys. Lett. B 2009, 23, 3769–3775. [Google Scholar] [CrossRef]
  55. Liu, F.; Xue, D. Fabrication of Nb2O5 nanotrees with controlled branching degrees. Phys. Scr. 2010, T139, 014074. [Google Scholar] [CrossRef]
  56. Yan, C.; Xue, D. Formation of Nb2O5 Nanotube Arrays through Phase Transformation. Adv. Mater. 2008, 20, 1055–1058. [Google Scholar] [CrossRef]
  57. Draper, P.H.G.; Harvey, J. The structure of anodic films—I. An electron diffraction examination of the products of anodic oxidation on tantalum, niobium and zirconium. Acta Metall. 1963, 11, 873–879. [Google Scholar] [CrossRef]
  58. Arora, M.R.; Kelly, R. The structure and stoichiometry of anodic films on V, Nb, Ta, Mo and W. J. Mater. Sci. 1977, 12, 1673–1684. [Google Scholar] [CrossRef]
  59. Aladjem, A.; Brandon, D.G.; Yahalom, J.; Zahavi, J. Electron-beam crystallization of anodic oxide films. Electrochim. Acta 1970, 15, 663–671. [Google Scholar] [CrossRef]
  60. Habazaki, H.; Oikawa, Y.; Fushimi, K.; Aoki, Y.; Shimizu, K.; Skeldon, P.; Thompson, G.E. Importance of water content in formation of porous anodic niobium oxide films in hot phosphate–glycerol electrolyte. Electrochim. Acta 2009, 54, 946–951. [Google Scholar] [CrossRef]
  61. Ou, J.Z.; Rani, R.A.; Ham, M.-H.; Field, M.R.; Zhang, Y.; Zheng, H.; Reece, P.; Zhuiykov, S.; Sriram, S.; Bhaskaran, M.; et al. Elevated Temperature Anodized Nb2O5: A Photoanode Material with Exceptionally Large Photoconversion Efficiencies. ACS Nano 2012, 6, 4045–4053. [Google Scholar] [CrossRef] [PubMed]
  62. Karlinsey, R.L. Preparation of self-organized niobium oxide microstructures via potentiostatic anodization. Electrochem. Commun. 2005, 7, 1190–1194. [Google Scholar] [CrossRef]
  63. Karlinsey, R.L. Self-assembled Nb2O5 microcones with tailored crystallinity. J. Mater. Sci. 2006, 41, 5017–5020. [Google Scholar] [CrossRef]
  64. Wei, W.; Lee, K.; Shaw, S.; Schmuki, P. Anodic formation of high aspect ratio, self-ordered Nb2O5 nanotubes. Chem. Commun. 2012, 48, 4244–4246. [Google Scholar] [CrossRef] [PubMed]
  65. Lee, K.; Yang, Y.; Yang, M.; Schmuki, P. Formation of Highly Ordered Nanochannel Nb Oxide by Self-Organizing Anodization. Chem.–A Eur. J. 2012, 18, 9521–9524. [Google Scholar] [CrossRef] [PubMed]
  66. Abdul Rani, R.; Zoolfakar, A.S.; Subbiah, J.; Ou, J.Z.; Kalantar-zadeh, K. Highly ordered anodized Nb2O5 nanochannels for dye-sensitized solar cells. Electrochem. Commun. 2014, 40, 20–23. [Google Scholar] [CrossRef]
  67. Rani, R.A.; Zoolfakar, A.S.; Ou, J.Z.; Kadir, R.A.; Nili, H.; Latham, K.; Sriram, S.; Bhaskaran, M.; Zhuiykov, S.; Kaner, R.B.; et al. Reduced impurity-driven defect states in anodized nanoporous Nb2O5: The possibility of improving performance of photoanodes. Chem. Commun. 2013, 49, 6349–6351. [Google Scholar] [CrossRef]
  68. Yao, D.D.; Rani, R.A.; O’Mullane, A.P.; Kalantar-zadeh, K.; Ou, J.Z. High Performance Electrochromic Devices Based on Anodized Nanoporous Nb2O5. J. Phys. Chem. C 2014, 118, 476–481. [Google Scholar] [CrossRef]
  69. Liu, X.L.; Yuan, R.L.; Liu, Y.S.; Zhu, S.; Lin, J.; Chen, X.F. Niobium pentoxide nanotube powder for efficient dye-sensitized solar cells. New J. Chem. 2016, 40, 6276–6280. [Google Scholar] [CrossRef]
  70. Khairir, N.S.; Rani, R.A.; Ab Kadir, R.; Soin, N.; Abdullah, W.F.H.; Mamat, M.H.; Rusop, M.; Zoolfakar, A.S. Electrical Behavior of a Nanoporous Nb2O5/Pt Schottky Contact at Elevated Temperatures. J. Electron. Mater. 2019, 48, 611–620. [Google Scholar] [CrossRef]
  71. Pligovka, A.; Yunin, P.; Hoha, A.; Korolyov, S.; Gorokh, G.; Skorokhodov, E. Morphology and Structure of Defected Niobium Oxide Nonuniform Arrays Formed by Anodizing Bilayer Al/Nb Systems. Tech. Phys. 2020, 65, 1771–1776. [Google Scholar] [CrossRef]
  72. Gorokh, G.G.; Pligovka, A.N.; Lozovenko, A. Columnar niobium oxide nanostructures: Mechanism of formation, microstructure, and electrophysical properties. Tech. Phys. 2019, 64, 1657–1665. [Google Scholar] [CrossRef]
  73. Pligovka, A.; Hoha, A.; Turavets, U.; Poznyak, A.; Zakharau, Y. Formation features, morphology and optical properties of nanostructures via anodizing Al/Nb on Si and glass. Mater. Today Proc. 2021, 37, A8–A15. [Google Scholar] [CrossRef]
  74. Pligovka, A.; Lazavenka, A.; Zakhlebayeva, A. Electro-physical properties of niobia columnlike nanostructures via the anodizing of Al/Nb layers. In Proceedings of the 2018 IEEE 18th International Conference on Nanotechnology (IEEE-NANO), Cork, Ireland, 23–26 July 2018; IEEE: Piscataway, NJ, USA, 2018; pp. 1–5. [Google Scholar]
  75. Pligovka, A.; Lazavenka, A.; Turavets, U.; Hoha, A.; Salerno, M. Two-Level 3D Column-like Nanofilms with Hexagonally–Packed Tantalum Fabricated via Anodizing of Al/Nb and Al/Ta Layers—A Potential Nano-Optical Biosensor. Materials 2023, 16, 993. [Google Scholar] [CrossRef]
  76. Matsuda, A.; Kawamura, G. Sol-Gel Nano-/Micropatterning Process. In Handbook of Sol-Gel Science and Technology; Springer: Cham, Switzerland, 2016. [Google Scholar]
  77. Ulrich, D.R. Prospects of sol-gel processes. J. Non-Cryst. Solids 1988, 100, 174–193. [Google Scholar] [CrossRef]
  78. Hench, L.L.; West, J.K. The sol-gel process. Chem. Rev. 1990, 90, 33–72. [Google Scholar] [CrossRef]
  79. Nistico, R.; Scalarone, D.; Magnacca, G. Sol-gel chemistry, templating and spin-coating deposition: A combined approach to control in a simple way the porosity of inorganic thin films/coatings. Microporous Mesoporous Mater. 2017, 248, 18–29. [Google Scholar] [CrossRef]
  80. Bokov, D.; Jalil, A.T.; Chupradit, S.; Suksatan, W.; Ansari, M.J.; Shewael, I.H.; Valiev, G.H.; Kianfar, E. Nanomaterial by Sol-Gel Method: Synthesis and Application. Adv. Mater. Sci. Eng. 2021, 2021, 5102014. [Google Scholar] [CrossRef]
  81. Niederberger, M.; Pinna, N. Metal Oxide Nanoparticles in Organic Solvents: Synthesis, Formation, Assembly and Application; Springer Science & Business Media: Cham, Switzerland, 2009. [Google Scholar]
  82. Alquier, C.; Vandenborre, M.T.; Henry, M. Synthesis of niobium pentoxide gels. J. Non-Cryst. Solids 1986, 79, 383–395. [Google Scholar] [CrossRef]
  83. Schmitt, M.; Heusing, S.; Aegerter, M.A.; Pawlicka, A.; Avellaneda, C. Electrochromic properties of Nb2O5 sol-gel coatings. Sol. Energy Mater. Sol. Cells 1998, 54, 9–17. [Google Scholar] [CrossRef]
  84. Melo, L.; Avellaneda, C.O.; Caram, R.; Sichieri, E.; Pawlicka, A. Electrochromic Properties of Sol-gel Coating of Nb2O5 and Nb2O5: Li+. Mater. Res. 2002, 5, 43–46. [Google Scholar] [CrossRef]
  85. Xu, Z.Y.; Shen, X.M.; Wu, X.R.; Cao, F.B.; Li, L.S. Effect of crystallization degree on optical characteristics of Nb2O5 by sol-gel process. Mod. Phys. Lett. B 2021, 35, 2150171. [Google Scholar] [CrossRef]
  86. Pu, J.; Shen, Z.H.; Zhong, C.L.; Zhou, Q.W.; Liu, J.Y.; Zhu, J.; Zhang, H.G. Electrodeposition Technologies for Li-Based Batteries: New Frontiers of Energy Storage. Adv. Mater. 2020, 32, 1903808. [Google Scholar] [CrossRef] [PubMed]
  87. Jang, J.-H.; Kim, T.-Y.; Kim, N.-J.; Lee, C.-H.; Park, E.-M.; Park, C.; Suh, S.-J. Preparation and characterization of Nb2O5–Al2O3 composite oxide formed by cathodic electroplating and anodizing. Mater. Sci. Eng. B 2011, 176, 1505–1508. [Google Scholar] [CrossRef]
  88. Jha, G.; Tran, T.; Qiao, S.P.; Ziegler, J.M.; Ogata, A.F.; Dai, S.; Xu, M.J.; Le Thai, M.; Chandran, G.T.; Pan, X.Q.; et al. Electrophoretic Deposition of Mesoporous Niobium(V)Oxide Nanoscopic Films. Chem. Mater. 2018, 30, 6549–6558. [Google Scholar] [CrossRef]
  89. RobertáLee, G. Studies on the electrochemical deposition of niobium oxide. J. Mater. Chem. 1996, 6, 187–192. [Google Scholar]
  90. Zhitomirsky, I. Electrolytic deposition of oxide films in the presence of hydrogen peroxide. J. Eur. Ceram. Soc. 1999, 19, 2581–2587. [Google Scholar] [CrossRef]
  91. Zhitomirsky, I. Electrolytic deposition of niobium oxide films. Mater. Lett. 1998, 35, 188–193. [Google Scholar] [CrossRef]
  92. Kamada, K.; Mukai, M.; Matsumoto, Y. Anodic dissolution of tantalum and niobium in acetone solvent with halogen additives for electrochemical synthesis of Ta2O5 and Nb2O5 thin films. Electrochim. Acta 2004, 49, 321–327. [Google Scholar] [CrossRef]
  93. Zhao, G.Y.; Ye, C.; Zhang, L.; Li, C.L.; Sun, K.N. T-Nb2O5 quantum dots prepared by electrodeposition for fast Li ion intercalation/deintercalation. Nanotechnology 2017, 28, 215401. [Google Scholar] [CrossRef]
  94. Yang, D. Nanocomposite Films for Gas Sensing; INTECH Open Access Publisher: Rijeka, Croatia, 2011. [Google Scholar]
  95. Yap, Y. Physical Vapor Deposition, Encyclopedia of Nanotechnology: Springer Reference; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  96. Moreto, J.; Gelamo, R.; Nascimento, J.; Taryba, M.; Fernandes, J. Improving the corrosion protection of 2524-T3-Al alloy through reactive sputtering Nb2O5 coatings. Appl. Surf. Sci. 2021, 556, 149750. [Google Scholar] [CrossRef]
  97. Çetinörgü-Goldenberg, E.; Klemberg-Sapieha, J.-E.; Martinu, L. Effect of postdeposition annealing on the structure, composition, and the mechanical and optical characteristics of niobium and tantalum oxide films. Appl. Opt. 2012, 51, 6498–6507. [Google Scholar] [CrossRef] [PubMed]
  98. Tien, C.-L. Effect of sputtering anisotropic ejection on the optical properties and residual stress of Nb2O5 thin films. Appl. Surf. Sci. 2010, 257, 481–486. [Google Scholar] [CrossRef]
  99. Reichman, B.; Bard, A.J. Electrochromism at niobium pentoxide electrodes in aqueous and acetonitrile solutions. J. Electrochem. Soc. 1980, 127, 241–242. [Google Scholar] [CrossRef]
  100. Vinnichenko, M.; Rogozin, A.; Grambole, D.; Munnik, F.; Kolitsch, A.; Möller, W.; Stenzel, O.; Wilbrandt, S.; Chuvilin, A.; Kaiser, U. Highly dense amorphous Nb2O5 films with closed nanosized pores. Appl. Phys. Lett. 2009, 95, 081904. [Google Scholar] [CrossRef]
  101. Pillis, M.F.; Geribola, G.A.; Scheidt, G.; de Araújo, E.G.; de Oliveira, M.C.L.; Antunes, R.A. Corrosion of thin, magnetron sputtered Nb2O5 films. Corros. Sci. 2016, 102, 317–325. [Google Scholar] [CrossRef]
  102. Coşkun, Ö.D.; Demirela, S. The optical and structural properties of amorphous Nb2O5 thin films prepared by RF magnetron sputtering. Appl. Surf. Sci. 2013, 277, 35–39. [Google Scholar] [CrossRef]
  103. Chen, K.-N.; Hsu, C.-M.; Liu, J.; Liou, Y.-C.; Yang, C.-F. Investigation of antireflection Nb2O5 thin films by the sputtering method under different deposition parameters. Micromachines 2016, 7, 151. [Google Scholar] [CrossRef]
  104. Al-Baradi, A.M.; El-Nahass, M.; Hassanien, A.; Atta, A.; Alqahtani, M.S.; Aldawsari, A.O. Influence of RF sputtering power on structural and optical properties of Nb2O5 thin films. Optik 2018, 168, 853–863. [Google Scholar] [CrossRef]
  105. Dhar, A.; Alford, T. Optimization of Nb2O5/Ag/Nb2O5 multilayers as transparent composite electrode on flexible substrate with high figure of merit. J. Appl. Phys. 2012, 112, 103113. [Google Scholar] [CrossRef]
  106. Venkataraj, S.; Drese, R.; Kappertz, O.; Jayavel, R.; Wuttig, M. Characterization of niobium oxide films prepared by reactive DC magnetron sputtering. Phys. Status Solidi A 2001, 188, 1047–1058. [Google Scholar] [CrossRef]
  107. Bussamara, R.; Eberhardt, D.; Feil, A.; Migowski, P.; Wender, H.; de Moraes, D.P.; Machado, G.; Papaléo, R.M.; Teixeira, S.R.; Dupont, J. Sputtering deposition of magnetic Ni nanoparticles directly onto an enzyme surface: A novel method to obtain a magnetic biocatalyst. Chem. Commun. 2013, 49, 1273–1275. [Google Scholar] [CrossRef] [PubMed]
  108. Kuzminykh, Y.; Dabirian, A.; Reinke, M.; Hoffmann, P. High vacuum chemical vapour deposition of oxides: A review of technique development and precursor selection. Surf. Coat. Technol. 2013, 230, 13–21. [Google Scholar] [CrossRef]
  109. Jung, S.-C.; Imaishi, N.; Park, H.-C. Reaction engineering modeling of low-pressure metalorganic chemical vapor deposition of Nb2O5 thin film. Jpn. J. Appl. Phys. 1995, 34, L775. [Google Scholar] [CrossRef]
  110. Breiland, W.G.; Coltrin, M.E.; Creighton, J.R.; Hou, H.Q.; Moffat, H.K.; Tsao, J.Y. Organometallic vapor phase epitaxy (OMVPE). Mater. Sci. Eng. R Rep. 1999, 24, 241–274. [Google Scholar] [CrossRef]
  111. Mueller, N.S.; Morfa, A.J.; Abou-Ras, D.; Oddone, V.; Ciuk, T.; Giersig, M. Growing graphene on polycrystalline copper foils by ultra-high vacuum chemical vapor deposition. Carbon 2014, 78, 347–355. [Google Scholar] [CrossRef]
  112. Baum, T.H.; Larson, C.E.; Jackson, R.L. Laser-induced chemical vapor deposition of aluminum. Appl. Phys. Lett. 1989, 55, 1264–1266. [Google Scholar] [CrossRef]
  113. O’Neill, S.A.; Parkin, I.P.; Clark, R.J.; Mills, A.; Elliott, N. Atmospheric pressure chemical vapour deposition of thin films of Nb2O5 on glass. J. Mater. Chem. 2003, 13, 2952–2956. [Google Scholar] [CrossRef]
  114. de Oliveira, A.F.; da Silveira, C.B.; de Campos, S.D.; de Campos, E.A.; Carasek, E. Niobium (V) oxide coated on thin glass–ceramic rod as a solid phase microextraction fiber. Talanta 2005, 66, 74–79. [Google Scholar] [CrossRef]
  115. da Silveira, C.; de Oliveira, A.; de Campos, S.; de Campos, E.; Fraporti, A. Nb2O5 coating of glass fibres applied by chemical vapour deposition. Surf. Eng. 2012, 28, 68–72. [Google Scholar] [CrossRef]
  116. Kovendhan, M.; Joseph, D.P.; Manimuthu, P.; Ganesan, S.; Sambasivam, S.; Maruthamuthu, P.; Suthanthiraraj, S.A.; Venkateswaran, C.; Mohan, R. Spray deposited Nb2O5 thin film electrodes for fabrication of dye sensitized solar cells. Trans. Indian Inst. Met. 2011, 64, 185–188. [Google Scholar] [CrossRef]
  117. Xia, J.; Masaki, N.; Jiang, K.; Yanagida, S. Fabrication and characterization of thin Nb2O5 blocking layers for ionic liquid-based dye-sensitized solar cells. J. Photochem. Photobiol. A 2007, 188, 120–127. [Google Scholar] [CrossRef]
  118. Wang, Z.; Hu, Y.; Wang, W.; Zhang, X.; Wang, B.; Tian, H.; Wang, Y.; Guan, J.; Gu, H. Fast and highly-sensitive hydrogen sensing of Nb2O5 nanowires at room temperature. Int. J. Hydrogen Energy 2012, 37, 4526–4532. [Google Scholar] [CrossRef]
  119. Tamang, R.; Varghese, B.; Mhaisalkar, S.G.; Tok, E.S.; Sow, C.H. Probing the photoresponse of individual Nb2O5 nanowires with global and localized laser beam irradiation. Nanotechnology 2011, 22, 115202. [Google Scholar] [CrossRef] [PubMed]
  120. Nico, C.; Rino, L.; Matos, M.; Monteiro, R.; Costa, F.M.; Monteiro, T.; Graça, M.P. NbO/Nb2O5 core–shells by thermal oxidation. J. Eur. Ceram. Soc. 2013, 33, 3077–3083. [Google Scholar] [CrossRef]
  121. Lim, J.H.; Choi, J. Formation of niobium oxide nanowires by thermal oxidation. J. Ind. Eng. Chem. 2009, 15, 860–864. [Google Scholar] [CrossRef]
  122. De Sá, A.; Rangel, C.; Skeldon, P.; Thompson, G. Semiconductive properties of anodic niobium oxides. Port. Electrochim. Acta 2006, 24, 305. [Google Scholar] [CrossRef]
  123. Luo, H.; Song, W.; Hoertz, P.G.; Hanson, K.; Ghosh, R.; Rangan, S.; Brennaman, M.K.; Concepcion, J.J.; Binstead, R.A.; Bartynski, R.A. A sensitized Nb2O5 photoanode for hydrogen production in a dye-sensitized photoelectrosynthesis cell. Chem. Mater. 2013, 25, 122–131. [Google Scholar] [CrossRef]
  124. Graça, M.P.F.; Meireles, A.; Nico, C.; Valente, M.A. Nb2O5 nanosize powders prepared by sol–gel–Structure, morphology and dielectric properties. J. Alloys Compd. 2013, 553, 177–182. [Google Scholar] [CrossRef]
  125. Jiayu, D.; Zhefei, W.; Zhijie, G.; Yuan, X.; Lixi, W.; ZHANG, Q. Optimization on dielectric properties of Y2Ti2O7 ceramics with Bi2O3-Nd2O3-Nb2O5 co-doping. J. Rare Earths 2011, 29, 683–687. [Google Scholar]
  126. Clima, S.; Pourtois, G.; Van Elshocht, S.; De Gendt, S.; Heyns, M.M.; Wouters, D.J.; Kittl, J.A. Dielectric response of Ta2O5, NbTaO5 and Nb2O5 from first-principles investigations. ECS Trans. 2009, 19, 729. [Google Scholar] [CrossRef]
  127. García, H.; Castán, H.; Perez, E.; Dueñas, S.; Bailón, L.; Blanquart, T.; Niinistö, J.; Kukli, K.; Ritala, M.; Leskelä, M. Influence of growth and annealing temperatures on the electrical properties of Nb2O5-based MIM capacitors. Semicond. Sci. Technol. 2013, 28, 055005. [Google Scholar] [CrossRef]
  128. Soares, M.; Leite, S.; Nico, C.; Peres, M.; Fernandes, A.; Graça, M.; Matos, M.; Monteiro, R.; Monteiro, T.; Costa, F. Effect of processing method on physical properties of Nb2O5. J. Eur. Ceram. Soc. 2011, 31, 501–506. [Google Scholar] [CrossRef]
  129. Greener, E.; Hirthe, W. Electrical Conductivity of Nonstoichiometric α-Nb2O5. J. Electrochem. Soc. 1962, 109, 600. [Google Scholar] [CrossRef]
  130. Nanda, G.; Awin, E.W.; Gasyak, T.; Koroleva, E.; Filimonov, A.; Vakhrushev, S.; Sujith, R.; Kumar, R. Temperature dependent conductivity and broadband dielectric response of precursor-derived Nb2O5. Ceram. Int. 2020, 46, 9512–9518. [Google Scholar] [CrossRef]
  131. Lazarova, K.; Vasileva, M.; Marinov, G.; Babeva, T. Optical characterization of sol–gel derived Nb2O5 thin films. Opt. Laser Technol. 2014, 58, 114–118. [Google Scholar] [CrossRef]
  132. Abood, M.; Wahid, M.; Saimon, J.; Salim, E. Physical properties of Nb2O5 thin films prepared at 12M ammonium concentration. Int. J. Nanoelectron. Mater. 2018, 11, 237–244. [Google Scholar]
  133. Sun, S.; Mottram, L.; Hyatt, N.C. On the existence of the compound “Ce3NbO7+δ” prepared under air atmosphere. J. Rare Earths 2021, 39, 596–599. [Google Scholar] [CrossRef]
  134. Fakhri, M.A.; Al-Douri, Y.; Hashim, U.; Salim, E.T. Optical investigations of photonics lithium niobate. Sol. Energy 2015, 120, 381–388. [Google Scholar] [CrossRef]
  135. Le Viet, A.; Jose, R.; Reddy, M.; Chowdari, B.; Ramakrishna, S. Nb2O5 photoelectrodes for dye-sensitized solar cells: Choice of the polymorph. J. Phys. Chem. C 2010, 114, 21795–21800. [Google Scholar] [CrossRef]
  136. Abe, S. Formation of Nb2O5 matrix and Vis-NIR absorption in Nb-Ge-O thin film. Nanoscale Res. Lett. 2012, 7, 1–6. [Google Scholar]
  137. Liu, J.; Xue, D.; Li, K. Single-crystalline nanoporous Nb2O5 nanotubes. Nanoscale Res. Lett. 2011, 6, 1–8. [Google Scholar]
  138. Brayner, R.; Bozon-Verduraz, F. Niobium pentoxide prepared by soft chemical routes: Morphology, structure, defects and quantum size effect. Phys. Chem. Chem. Phys. 2003, 5, 1457–1466. [Google Scholar] [CrossRef]
  139. Agarwal, G.; Reddy, G. Study of surface morphology and optical properties of Nb2O5 thin films with annealing. J. Mater. Sci. Mater. Electron. 2005, 16, 21–24. [Google Scholar] [CrossRef]
  140. Aegerter, M.A.; Schmitt, M.; Guo, Y. Sol-gel niobium pentoxide coatings: Applications to photovoltaic energy conversion and electrochromism. Int. J. Photoenergy 2002, 4, 108237. [Google Scholar] [CrossRef]
  141. Xiao, X.; Dong, G.; Xu, C.; He, H.; Qi, H.; Fan, Z.; Shao, J. Structure and optical properties of Nb2O5 sculptured thin films by glancing angle deposition. Appl. Surf. Sci. 2008, 255, 2192–2195. [Google Scholar] [CrossRef]
  142. Shcherbina, O.; Palatnikov, M.; Efremov, V. Mechanical properties of Nb2O5 and Ta2O5 prepared by different procedures. Inorg. Mater. 2012, 48, 433–438. [Google Scholar] [CrossRef]
  143. Rani, R.A.; Zoolfakar, A.S.; O’Mullane, A.P.; Austin, M.W.; Kalantar-Zadeh, K. Thin films and nanostructures of niobium pentoxide: Fundamental properties, synthesis methods and applications. J. Mater. Chem. A 2014, 2, 15683–15703. [Google Scholar]
  144. Cavers, H.; Molaiyan, P.; Abdollahifar, M.; Lassi, U.; Kwade, A. Perspectives on improving the safety and sustainability of high voltage lithium-ion batteries through the electrolyte and separator region. Adv. Energy Mater. 2022, 12, 2200147. [Google Scholar]
  145. Armand, M.; Tarascon, J.-M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef]
  146. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [PubMed]
  147. Zhang, H.; Yang, Y.; Ren, D.; Wang, L.; He, X. Graphite as anode materials: Fundamental mechanism, recent progress and advances. Energy Storage Mater. 2021, 36, 147–170. [Google Scholar]
  148. Ren, J.; Ren, R.P.; Lv, Y.K. A New Anode for Lithium-Ion Batteries Based on Single-Walled Carbon Nanotubes and Graphene: Improved Performance through a Binary Network Design. Chem.–A Asian J. 2018, 13, 1223–1227. [Google Scholar]
  149. Abdollahifar, M.; Molaiyan, P.; Perovic, M.; Kwade, A. Insights into Enhancing Electrochemical Performance of Li-Ion Battery Anodes via Polymer Coating. Energies 2022, 15, 8791. [Google Scholar]
  150. Molaiyan, P.; Dos Reis, G.S.; Karuppiah, D.; Subramaniyam, C.M.; Garcia-Alvarado, F.; Lassi, U. Recent progress in biomass-derived carbon materials for Li-ion and Na-ion batteries—A review. Batteries 2023, 9, 116. [Google Scholar]
  151. Välikangas, J.; Laine, P.; Hu, T.; Tynjälä, P.; Selent, M.; Molaiyan, P.; Jürgen, K.; Lassi, U. Effect of Secondary Heat Treatment after a Washing on the Electrochemical Performance of Co-Free LiNi0.975Al0.025O2 Cathodes for Li-Ion Batteries. Small 2023, 2305349. [Google Scholar] [CrossRef]
  152. Yuan, C.; Wu, H.B.; Xie, Y.; Lou, X.W. Mixed transition-metal oxides: Design, synthesis, and energy-related applications. Angew. Chem. Int. Ed. 2014, 53, 1488–1504. [Google Scholar]
  153. Deng, Q.; Fu, Y.; Zhu, C.; Yu, Y. Niobium-based oxides toward advanced electrochemical energy storage: Recent advances and challenges. Small 2019, 15, 1804884. [Google Scholar] [CrossRef]
  154. Zhou, Z.; Lou, S.; Cheng, X.; Cui, B.; Si, W.; Ding, F.; Ma, Y.; Zuo, P.; Du, C.; Wang, J. Superior Electrochemical Performance of WNb2O8 Nanorods Triggered by Ultra-Efficient Li+ Diffusion. ChemistrySelect 2020, 5, 1209–1213. [Google Scholar] [CrossRef]
  155. Lin, J.; Yuan, Y.; Su, Q.; Pan, A.; Dinesh, S.; Peng, C.; Cao, G.; Liang, S. Facile synthesis of Nb2O5/carbon nanocomposites as advanced anode materials for lithium-ion batteries. Electrochim. Acta 2018, 292, 63–71. [Google Scholar]
  156. Shen, F.; Sun, Z.; He, Q.; Sun, J.; Kaner, R.B.; Shao, Y. Niobium pentoxide based materials for high rate rechargeable electrochemical energy storage. Mater. Horiz. 2021, 8, 1130–1152. [Google Scholar]
  157. Griffith, K.J.; Wiaderek, K.M.; Cibin, G.; Marbella, L.E.; Grey, C.P. Niobium tungsten oxides for high-rate lithium-ion energy storage. Nature 2018, 559, 556–563. [Google Scholar] [PubMed]
  158. Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the development of advanced Li-ion batteries: A review. Energy Environ. Sci. 2011, 4, 3243–3262. [Google Scholar]
  159. Reddy, M.V.; Subba Rao, G.V.; Chowdari, B.V.R. Metal Oxides and Oxysalts as Anode Materials for Li Ion Batteries. Chem. Rev. 2013, 113, 5364–5457. [Google Scholar] [PubMed]
  160. Ding, H.; Song, Z.; Zhang, H.; Zhang, H.; Li, X. Niobium-based oxide anodes toward fast and safe energy storage: A review. Mater. Today Nano 2020, 11, 100082. [Google Scholar]
  161. Yang, M.; Li, S.; Huang, J. Three-Dimensional Cross-Linked Nb2O5 Polymorphs Derived from Cellulose Substances: Insights into the Mechanisms of Lithium Storage. ACS Appl. Mater. Interfaces 2021, 13, 39501–39512. [Google Scholar]
  162. Li, M.; Lu, J.; Chen, Z.; Amine, K. 30 years of lithium-ion batteries. Adv. Mater. 2018, 30, 1800561. [Google Scholar]
  163. Reichman, B.; Bard, A.J. The Application of Nb2O5 as a Cathode in Nonaqueous Lithium Cells. J. Electrochem. Soc. 1981, 128, 344. [Google Scholar] [CrossRef]
  164. Kumagai, N.; Tanno, K.; Nakajima, T.; Watanabe, N. Structural changes of Nb2O5 and V2O5 as rechargeable cathodes for lithium battery. Electrochim. Acta 1983, 28, 17–22. [Google Scholar]
  165. Kodama, R.; Terada, Y.; Nakai, I.; Komaba, S.; Kumagai, N. Electrochemical and in situ XAFS-XRD investigation of Nb2O5 for rechargeable lithium batteries. J. Electrochem. Soc. 2006, 153, A583. [Google Scholar] [CrossRef]
  166. Han, J.-T.; Huang, Y.-H.; Goodenough, J.B. New anode framework for rechargeable lithium batteries. Chem. Mater. 2011, 23, 2027–2029. [Google Scholar] [CrossRef]
  167. Nakazawa, H.; Sano, K.; Abe, T.; Baba, M.; Kumagai, N. Charge–discharge characteristics of all-solid-state thin-filmed lithium-ion batteries using amorphous Nb2O5 negative electrodes. J. Power Sources 2007, 174, 838–842. [Google Scholar] [CrossRef]
  168. Kumagai, N.; Koishikawa, Y.; Komaba, S.; Koshiba, N. Thermodynamics and Kinetics of Lithium Intercalation into Nb2O5 Electrodes for a 2 V Rechargeable Lithium Battery. J. Electrochem. Soc. 1999, 146, 3203. [Google Scholar] [CrossRef]
  169. Come, J.; Augustyn, V.; Kim, J.W.; Rozier, P.; Taberna, P.-L.; Gogotsi, P.; Long, J.W.; Dunn, B.; Simon, P. Electrochemical kinetics of nanostructured Nb2O5 electrodes. J. Electrochem. Soc. 2014, 161, A718. [Google Scholar] [CrossRef]
  170. Liu, C.-P.; Zhou, F.; Ozolins, V. First principles study for lithium intercalation and diffusion behavior in orthorhombic Nb2O5 electrochemical supercapacitor. In APS March Meeting Abstracts; American Physical Society: College Park, MD, USA, 2012; p. B26-003. [Google Scholar]
  171. Kim, J.W.; Augustyn, V.; Dunn, B. The effect of crystallinity on the rapid pseudocapacitive response of Nb2O5. Adv. Energy Mater. 2012, 2, 141–148. [Google Scholar] [CrossRef]
  172. Song, Z.; Li, H.; Liu, W.; Zhang, H.; Yan, J.; Tang, Y.; Huang, J.; Zhang, H.; Li, X. Ultrafast and stable Li-(de) intercalation in a large single crystal H-Nb2O5 anode via optimizing the homogeneity of electron and ion transport. Adv. Mater. 2020, 32, 2001001. [Google Scholar] [CrossRef]
  173. Lübke, M.; Sumboja, A.; Johnson, I.D.; Brett, D.J.L.; Shearing, P.R.; Liu, Z.; Darr, J.A. High power nano-Nb2O5 negative electrodes for lithium-ion batteries. Electrochim. Acta 2016, 192, 363–369. [Google Scholar] [CrossRef]
  174. Li, S.; Xu, Q.; Uchaker, E.; Cao, X.; Cao, G. Comparison of amorphous, pseudohexagonal and orthorhombic Nb2O5 for high-rate lithium ion insertion. CrystEngComm 2016, 18, 2532–2540. [Google Scholar] [CrossRef]
  175. Liu, Z.; Dong, W.; Wang, J.; Dong, C.; Lin, Y.; Chen, I.-W.; Huang, F. Orthorhombic Nb2O5−x for durable high-rate anode of Li-ion batteries. iScience 2020, 23, 100767. [Google Scholar] [CrossRef]
  176. Chen, H.; Zhang, H.; Wu, Y.; Zhang, T.; Guo, Y.; Zhang, Q.; Zeng, Y.; Lu, J. Nanostructured Nb2O5 cathode for high-performance lithium-ion battery with Super-P and graphene compound conductive agents. J. Electroanal. Chem. 2018, 827, 112–119. [Google Scholar] [CrossRef]
  177. Ye, R.; Ohta, K.; Baba, M. Electrochemical properties of amorphous Nb2O5 thin film and its application to rechargeable thin film lithium ion batteries. ECS Trans. 2016, 73, 49. [Google Scholar] [CrossRef]
  178. Liu, M.; Yan, C.; Zhang, Y. Fabrication of Nb2O5 nanosheets for high-rate lithium ion storage applications. Sci. Rep. 2015, 5, 8326. [Google Scholar] [CrossRef] [PubMed]
  179. Uchida, S.; Zettsu, N.; Hirata, K.; Kami, K.; Teshima, K. High-voltage capabilities of ultra-thin Nb2O5 nanosheet coated LiNi1/3Co1/3Mn1/3O2 cathodes. RSC Adv. 2016, 6, 67514–67519. [Google Scholar] [CrossRef]
  180. Liu, X.; Liu, G.; Chen, H.; Ma, J.; Zhang, R. Facile synthesis of Nb2O5 nanobelts assembled from nanorods and their applications in lithium ion batteries. J. Phys. Chem. Solids 2017, 111, 8–11. [Google Scholar] [CrossRef]
  181. Huang, C.; Fu, J.; Song, H.; Li, X.; Peng, X.; Gao, B.; Zhang, X.; Chu, P.K. General fabrication of mesoporous Nb2O5 nanobelts for lithium ion battery anodes. RSC Adv. 2016, 6, 90489–90493. [Google Scholar] [CrossRef]
  182. Wei, M.; Wei, K.; Ichihara, M.; Zhou, H. Nb2O5 nanobelts: A lithium intercalation host with large capacity and high rate capability. Electrochem. Commun. 2008, 10, 980–983. [Google Scholar] [CrossRef]
  183. Sun, Y.-G.; Piao, J.-Y.; Hu, L.-L.; Bin, D.-S.; Lin, X.-J.; Duan, S.-Y.; Cao, A.-M.; Wan, L.-J. Controlling the reaction of nanoparticles for hollow metal oxide nanostructures. J. Am. Chem. Soc. 2018, 140, 9070–9073. [Google Scholar] [CrossRef]
  184. Lu, H.; Xiang, K.; Bai, N.; Zhou, W.; Wang, S.; Chen, H. Urchin-shaped Nb2O5 microspheres synthesized by the facile hydrothermal method and their lithium storage performance. Mater. Lett. 2016, 167, 106–108. [Google Scholar] [CrossRef]
  185. Liu, S.; Zhou, J.; Cai, Z.; Fang, G.; Pan, A.; Liang, S. Nb2O5 microstructures: A high-performance anode for lithium ion batteries. Nanotechnology 2016, 27, 46LT01. [Google Scholar] [CrossRef]
  186. Liu, X.; Liu, G.; Liu, Y.; Sun, R.; Ma, J.; Guo, J.; Hu, M. Urchin-like hierarchical H-Nb2O5 microspheres: Synthesis, formation mechanism and their applications in lithium ion batteries. Dalton Trans. 2017, 46, 10935–10940. [Google Scholar] [CrossRef]
  187. Liu, G.; Jin, B.; Bao, K.; Xie, H.; Guo, J.; Ji, X.; Zhang, R.; Jiang, Q. Facile synthesis of porous Nb2O5 microspheres as anodes for lithium-ion batteries. Int. J. Hydrogen Energy 2017, 42, 6065–6071. [Google Scholar] [CrossRef]
  188. Lou, S.; Cheng, X.; Wang, L.; Gao, J.; Li, Q.; Ma, Y.; Gao, Y.; Zuo, P.; Du, C.; Yin, G. High-rate capability of three-dimensionally ordered macroporous T-Nb2O5 through Li+ intercalation pseudocapacitance. J. Power Sources 2017, 361, 80–86. [Google Scholar] [CrossRef]
  189. Li, S.; Schmidt, C.N.; Xu, Q.; Cao, X.; Cao, G. Macroporous nanostructured Nb2O5 with surface Nb4+ for enhanced lithium ion storage properties. ChemNanoMat 2016, 2, 675–680. [Google Scholar] [CrossRef]
  190. Chen, J.; Wang, H.; Zhang, X.; Liu, B.; Xu, L.; Zhang, Z.; Zhang, Y. 2D ultrathin nanosheet-assembled Nb2O5 microflowers for lithium ion batteries. Mater. Lett. 2018, 227, 112–115. [Google Scholar] [CrossRef]
  191. Li, X.; Sun, X.; Hu, X.; Fan, F.; Cai, S.; Zheng, C.; Stucky, G.D. Review on comprehending and enhancing the initial Coulombic efficiency of anode materials in lithium-ion/sodium-ion batteries. Nano Energy 2020, 77, 105143. [Google Scholar]
  192. Han, X.; Meng, Q.; Wan, X.; Sun, B.Y.; Zhang, Y.; Shen, B.C.; Gao, J.L.; Ma, Y.L.; Zuo, P.J.; Lou, S.F.; et al. Intercalation pseudocapacitive electrochemistry of Nb-based oxides for fast charging of lithium-ion batteries. Nano Energy 2021, 81, 105635. [Google Scholar] [CrossRef]
  193. Yan, L.; Rui, X.; Chen, G.; Xu, W.; Zou, G.; Luo, H. Recent advances in nanostructured Nb-based oxides for electrochemical energy storage. Nanoscale 2016, 8, 8443–8465. [Google Scholar]
  194. Lubimtsev, A.A.; Kent, P.R.; Sumpter, B.G.; Ganesh, P. Understanding the origin of high-rate intercalation pseudocapacitance in Nb2O5 crystals. J. Mater. Chem. A 2013, 1, 14951–14956. [Google Scholar] [CrossRef]
  195. Chen, P.-C.; Shen, G.; Shi, Y.; Chen, H.; Zhou, C. Preparation and characterization of flexible asymmetric supercapacitors based on transition-metal-oxide nanowire/single-walled carbon nanotube hybrid thin-film electrodes. ACS Nano 2010, 4, 4403–4411. [Google Scholar] [CrossRef]
  196. Lim, E.; Kim, H.; Jo, C.; Chun, J.; Ku, K.; Kim, S.; Lee, H.I.; Nam, I.-S.; Yoon, S.; Kang, K. Advanced hybrid supercapacitor based on a mesoporous niobium pentoxide/carbon as high-performance anode. ACS Nano 2014, 8, 8968–8978. [Google Scholar] [CrossRef]
  197. Wang, X.; Yan, C.; Yan, J.; Sumboja, A.; Lee, P.S. Orthorhombic niobium oxide nanowires for next generation hybrid supercapacitor device. Nano Energy 2015, 11, 765–772. [Google Scholar] [CrossRef]
  198. Wang, X.; Li, G.; Chen, Z.; Augustyn, V.; Ma, X.; Wang, G.; Dunn, B.; Lu, Y. High-performance supercapacitors based on nanocomposites of Nb2O5 nanocrystals and carbon nanotubes. Adv. Energy Mater. 2011, 1, 1089–1093. [Google Scholar] [CrossRef]
  199. Zhang, S.; Wu, J.; Wang, J.; Qiao, W.; Long, D.; Ling, L. Constructing T-Nb2O5@Carbon hollow core-shell nanostructures for high-rate hybrid supercapacitor. J. Power Sources 2018, 396, 88–94. [Google Scholar] [CrossRef]
  200. Liu, S.; Zhou, J.; Cai, Z.; Fang, G.; Cai, Y.; Pan, A.; Liang, S. Nb2O5 quantum dots embedded in MOF derived nitrogen-doped porous carbon for advanced hybrid supercapacitor applications. J. Mater. Chem. A 2016, 4, 17838–17847. [Google Scholar]
  201. Ding, Y.; Zeng, M.; Fu, L. Low-temperature synthesis of sp2 carbon nanomaterials. Sci. Bull. 2019, 64, 1817–1829. [Google Scholar]
  202. Wang, X.; Li, G.; Tjandra, R.; Fan, X.; Xiao, X.; Yu, A. Fast lithium-ion storage of Nb2O5 nanocrystals in situ grown on carbon nanotubes for high-performance asymmetric supercapacitors. RSC Adv. 2015, 5, 41179–41185. [Google Scholar]
  203. Ma, Y.; Han, J.; Wang, M.; Chen, X.; Jia, S. Electrophoretic deposition of graphene-based materials: A review of materials and their applications. J. Mater. 2018, 4, 108–120. [Google Scholar]
  204. Chen, Z.; Li, H.; Lu, X.; Wu, L.; Jiang, J.; Jiang, S.; Wang, J.; Dou, H.; Zhang, X. Nitrogenated Urchin-like Nb2O5 Microspheres with Extraordinary Pseudocapacitive Properties for Lithium-Ion Capacitors. ChemElectroChem 2018, 5, 1516–1524. [Google Scholar]
  205. Wang, J.; Li, H.; Shen, L.; Dong, S.; Zhang, X. Nb2O5 nanoparticles encapsulated in ordered mesoporous carbon matrix as advanced anode materials for Li ion capacitors. RSC Adv. 2016, 6, 71338–71344. [Google Scholar] [CrossRef]
  206. Zheng, J.; Cygan, P.; Jow, T. Hydrous ruthenium oxide as an electrode material for electrochemical capacitors. J. Electrochem. Soc. 1995, 142, 2699. [Google Scholar] [CrossRef]
  207. Arunkumar, P.; Ashish, A.G.; Babu, B.; Sarang, S.; Suresh, A.; Sharma, C.H.; Thalakulam, M.; Shaijumon, M.M. Nb2O5/graphene nanocomposites for electrochemical energy storage. RSC Adv. 2015, 5, 59997–60004. [Google Scholar] [CrossRef]
  208. Zhang, C.J.; Maloney, R.; Lukatskaya, M.R.; Beidaghi, M.; Dyatkin, B.; Perre, E.; Long, D.; Qiao, W.; Dunn, B.; Gogotsi, Y. Synthesis and electrochemical properties of niobium pentoxide deposited on layered carbide-derived carbon. J. Power Sources 2015, 274, 121–129. [Google Scholar] [CrossRef]
Figure 1. Milestone events of Nb2O5 research.
Figure 1. Milestone events of Nb2O5 research.
Materials 16 06956 g001
Figure 2. Nb2O5 crystal phase with temperature.
Figure 2. Nb2O5 crystal phase with temperature.
Materials 16 06956 g002
Figure 3. Crystal structures of Nb2O5 materials. (a) Structure of T-Nb2O5 observed along the b-axis and c-axis. (b) Structure of M-Nb2O5 observed along the b-axis and c-axis. (c) Structure of H-Nb2O5 observed along the b an c axes. (d) Crystal structure of TT-Nb2O5 in different directions. (e) Unit cell of B-Nb2O5 phase, On the right side of the picture it is shown how the polyhedra are connected in the crystal structure.
Figure 3. Crystal structures of Nb2O5 materials. (a) Structure of T-Nb2O5 observed along the b-axis and c-axis. (b) Structure of M-Nb2O5 observed along the b-axis and c-axis. (c) Structure of H-Nb2O5 observed along the b an c axes. (d) Crystal structure of TT-Nb2O5 in different directions. (e) Unit cell of B-Nb2O5 phase, On the right side of the picture it is shown how the polyhedra are connected in the crystal structure.
Materials 16 06956 g003
Figure 4. XRD patterns of various structures of Nb2O5 materials. Different niobium oxide phases (a) TT-Nb2O5 and (b) T-Nb2O5 were obtained by oxidizing niobium powder at 500 and 600 °C, respectively. Reprinted with permission from Ref. [34], Copyright 2020 Elsevier.
Figure 4. XRD patterns of various structures of Nb2O5 materials. Different niobium oxide phases (a) TT-Nb2O5 and (b) T-Nb2O5 were obtained by oxidizing niobium powder at 500 and 600 °C, respectively. Reprinted with permission from Ref. [34], Copyright 2020 Elsevier.
Materials 16 06956 g004
Figure 5. (a) TEM images of Nb2O5 nanorods synthesized by the hydrothermal method. Reprinted with permission from Ref. [48], Copyright 2017 Elsevier. Hydrothermal method (b) Nb2O5 nanostructure. Reprinted with permission from Ref. [49], Copyright 2018 American Chemical Society. Precipitation method (c) porous Nb2O5 nanoparticles. Reprinted with permission from Ref. [50], Copyright 2007 Elsevier. (d) N2-650 samples prepared by the sol-gel method. Reprinted with permission from Ref. [51], Copyright 2004 Elsevier. (e) High power SEM image of m-T-Nb2O5 NFs prepared by electrospinning. Reprinted with permission from Ref. [52], Copyright 2017 Elsevier. (f) Electrostatic spun multiple NaNbO3/Nb2O5 heterostructure nanotubes. Reprinted with permission from Ref. [53], Copyright 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 5. (a) TEM images of Nb2O5 nanorods synthesized by the hydrothermal method. Reprinted with permission from Ref. [48], Copyright 2017 Elsevier. Hydrothermal method (b) Nb2O5 nanostructure. Reprinted with permission from Ref. [49], Copyright 2018 American Chemical Society. Precipitation method (c) porous Nb2O5 nanoparticles. Reprinted with permission from Ref. [50], Copyright 2007 Elsevier. (d) N2-650 samples prepared by the sol-gel method. Reprinted with permission from Ref. [51], Copyright 2004 Elsevier. (e) High power SEM image of m-T-Nb2O5 NFs prepared by electrospinning. Reprinted with permission from Ref. [52], Copyright 2017 Elsevier. (f) Electrostatic spun multiple NaNbO3/Nb2O5 heterostructure nanotubes. Reprinted with permission from Ref. [53], Copyright 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Materials 16 06956 g005
Figure 6. SEM views of surface and cross-sectional, schematic 3D views (ac) skittle- and (df) medusa-like embryos formed by anodizing bilayer Al/Nb systems in 0.4 mol dm−3 oxalic acid aqueous solution at 37 and 53 V. SEM views of surface and cross-sectional, schematic 3D views (gi) medusa—and (jl) goblet-like embryos formed by anodizing bilayer Al/Nb systems in 0.4 mol dm−3 phosphoric acid aqueous solution at 100 and 150 V. The images were obtained after the alumina layer had been dissolved away (“alumina-free” samples). Reprinted with permission from Ref. [73], Copyright 2021 Elsevier.
Figure 6. SEM views of surface and cross-sectional, schematic 3D views (ac) skittle- and (df) medusa-like embryos formed by anodizing bilayer Al/Nb systems in 0.4 mol dm−3 oxalic acid aqueous solution at 37 and 53 V. SEM views of surface and cross-sectional, schematic 3D views (gi) medusa—and (jl) goblet-like embryos formed by anodizing bilayer Al/Nb systems in 0.4 mol dm−3 phosphoric acid aqueous solution at 100 and 150 V. The images were obtained after the alumina layer had been dissolved away (“alumina-free” samples). Reprinted with permission from Ref. [73], Copyright 2021 Elsevier.
Materials 16 06956 g006
Figure 7. (a) Structure of lithiated T-Nb2O5 after geometry optimization. (b) Two kinds of Li-ion transport path topologies (Path A and Path B). Reprinted with permission from Ref. [36], Copyright 2017 American Chemical Society.
Figure 7. (a) Structure of lithiated T-Nb2O5 after geometry optimization. (b) Two kinds of Li-ion transport path topologies (Path A and Path B). Reprinted with permission from Ref. [36], Copyright 2017 American Chemical Society.
Materials 16 06956 g007
Figure 8. Electrochemical responses of the Nb2O5 polymorphs (H, B, M, T and TT). (a) Cycling performance tests at 1C (T, TT) or C/10 (B, H). (b) Differential capacity plots derived from the discharge/charge profiles. Reprinted with permission from Ref. [156], Copyright 2021 Royal Society of Chemistry. (c) Rate performances of Nb2O5 (H, M, T and TT) tested at various current densities ranging from 0.1, 0.2, 0.5, 1, 2, 5 and back to 0.1 A g−1. (d) Long cycling performances of Nb2O5 (H, M, T and TT) at a current density of 200 mA g−1. Reprinted with permission from Ref. [34], Copyright 2020 Elsevier.
Figure 8. Electrochemical responses of the Nb2O5 polymorphs (H, B, M, T and TT). (a) Cycling performance tests at 1C (T, TT) or C/10 (B, H). (b) Differential capacity plots derived from the discharge/charge profiles. Reprinted with permission from Ref. [156], Copyright 2021 Royal Society of Chemistry. (c) Rate performances of Nb2O5 (H, M, T and TT) tested at various current densities ranging from 0.1, 0.2, 0.5, 1, 2, 5 and back to 0.1 A g−1. (d) Long cycling performances of Nb2O5 (H, M, T and TT) at a current density of 200 mA g−1. Reprinted with permission from Ref. [34], Copyright 2020 Elsevier.
Materials 16 06956 g008
Figure 9. (ad) GITT plots of TT-, T-, M- and H-Nb2O5, respectively. Reprinted with permission from Ref. [161], Copyright 2021 American Chemical Society.
Figure 9. (ad) GITT plots of TT-, T-, M- and H-Nb2O5, respectively. Reprinted with permission from Ref. [161], Copyright 2021 American Chemical Society.
Materials 16 06956 g009
Figure 10. Initial discharge curves of (a) hexagonal, (b) orthorhombic, and (c) monoclinic Nb2O5 powder pressed electrodes as a function of x in LixNb2O5 [168].
Figure 10. Initial discharge curves of (a) hexagonal, (b) orthorhombic, and (c) monoclinic Nb2O5 powder pressed electrodes as a function of x in LixNb2O5 [168].
Materials 16 06956 g010
Figure 11. (a)TEM image of HM-Nb2O5 collected 20 min after reaction. Reprinted with permission from Ref. [183], Copyright 2018 American Chemical Society (b) Field emission SEM image of Nb2O5 microspheres. Reprinted with permission from Ref. [186], Copyright 2007 Royal Society of Chemistry. (c) SEM image and (d) TEM image of 3DOM T-Nb2O5. Reprinted with permission from Ref. [188], Copyright 2017 Elsevier.
Figure 11. (a)TEM image of HM-Nb2O5 collected 20 min after reaction. Reprinted with permission from Ref. [183], Copyright 2018 American Chemical Society (b) Field emission SEM image of Nb2O5 microspheres. Reprinted with permission from Ref. [186], Copyright 2007 Royal Society of Chemistry. (c) SEM image and (d) TEM image of 3DOM T-Nb2O5. Reprinted with permission from Ref. [188], Copyright 2017 Elsevier.
Materials 16 06956 g011
Figure 12. Classification of Nb2O5 nanostructure surface morphologies. The pictures are from the literature.
Figure 12. Classification of Nb2O5 nanostructure surface morphologies. The pictures are from the literature.
Materials 16 06956 g012
Figure 13. (a) CV curves of T-Nb2O5@C NCs at various sweep rates from 0.1 to 1.0 mV s−1. Reprinted with permission from Ref. [30], Copyright 2015 American Chemical Society. (b) Cyclic voltammetry curves of sample C-T-Nb2O5 for first three cycles at a scan rate of 0.1 mV s−1. Reprinted with permission from Ref. [197], Copyright 2014 Elsevier Cycle voltammetry curves: (c) Nb2O5@C and (d) Nb2O5@MC. Reprinted with permission from Ref. [199], Copyright 2018 Elsevier.
Figure 13. (a) CV curves of T-Nb2O5@C NCs at various sweep rates from 0.1 to 1.0 mV s−1. Reprinted with permission from Ref. [30], Copyright 2015 American Chemical Society. (b) Cyclic voltammetry curves of sample C-T-Nb2O5 for first three cycles at a scan rate of 0.1 mV s−1. Reprinted with permission from Ref. [197], Copyright 2014 Elsevier Cycle voltammetry curves: (c) Nb2O5@C and (d) Nb2O5@MC. Reprinted with permission from Ref. [199], Copyright 2018 Elsevier.
Materials 16 06956 g013
Table 1. Classification of Nb2O5 polymorphs.
Table 1. Classification of Nb2O5 polymorphs.
PolymorphismDesignation of the Nb2O5 FormCrystallization Temperature (K)Ref.
TT-Nb2O5Tief-tief773–873[11]
T-Nb2O5Tief (γ)873–1073[12,13]
B-Nb2O5Blätter (ζ)1023–1123[14]
M-Nb2O5Medium (β)1173–1223[15]
H-Nb2O5High (α)1273[16]
N-Nb2O5Needles1103[17]
P-Nb2O5Prisms (η)1023[11]
R-Nb2O5neutral-[18]
ε-Nb2O5-1708[11]
Nb2O5-I-high-1558[11]
Nb2O5-II-1153–1223[11]
The oxI to oxVI Forms of—Nb2O5oxidation1573[11]
Table 2. Structure parameter of various Nb2O5 structures.
Table 2. Structure parameter of various Nb2O5 structures.
MaterialsCrystal StructureCell ParameterSpace GroupRef.
TT-Nb2O5Pseudohexagonala = b = 3.607 Å/a = b = 3.600 Å
c = 3.925 Å/c = 3.919 Å
P6/mmm (No. 191)[12,20]
T-Nb2O5Orthorhombica = 6.75 Å/a = 6.144 Å
b = 29.175 Å/b = 29.194 Å
c = 3.930 Å/c = 3.940 Å
Pbam (No. 55)[12,13]
M-Nb2O5Tetragonala = b = 20.44 Å
c = 3.832 Å
I4/mmm (No. 139)[21]
H-Nb2O5Monoclinica = 21.153 Å/a = 21.163 Å
b = 3.8233 Å/b = 3.824 Å
c = 19.356 Å/c = 19.355 Å
P2/m (No. 10)[13,21]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pang, R.; Wang, Z.; Li, J.; Chen, K. Polymorphs of Nb2O5 Compound and Their Electrical Energy Storage Applications. Materials 2023, 16, 6956. https://doi.org/10.3390/ma16216956

AMA Style

Pang R, Wang Z, Li J, Chen K. Polymorphs of Nb2O5 Compound and Their Electrical Energy Storage Applications. Materials. 2023; 16(21):6956. https://doi.org/10.3390/ma16216956

Chicago/Turabian Style

Pang, Rui, Zhiqiang Wang, Jinkai Li, and Kunfeng Chen. 2023. "Polymorphs of Nb2O5 Compound and Their Electrical Energy Storage Applications" Materials 16, no. 21: 6956. https://doi.org/10.3390/ma16216956

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop