Next Article in Journal
Investigation on Physico Chemical and X-ray Shielding Performance of Zinc Doped Nano-WO3 Epoxy Composite for Light Weight Lead Free Aprons
Next Article in Special Issue
Tailoring the Microstructure of Porous Carbon Spheres as High Rate Performance Anodes for Lithium-Ion Batteries
Previous Article in Journal
Machining of Triangular Holes in D2 Steel by the Use of Non-Conventional Electrodes in Die-Sinking Electric Discharge Machining
Previous Article in Special Issue
Optimized Pinecone-Squama-Structure MoS2-Coated CNT and Graphene Framework as Binder-Free Anode for Li-Ion Battery with High Capacity and Cycling Stability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanostructured TiO2 Arrays for Energy Storage

by
Pingyun Si
1,
Zhilong Zheng
2,
Yijie Gu
3,
Chao Geng
1,*,
Zhizhong Guo
1,
Jiayi Qin
1 and
Wei Wen
1,*
1
School of Mechanical and Electrical Engineering, Collaborative Innovation Center of Ecological Civilization, Hainan University, Haikou 570228, China
2
Zhanjiang Power Supply Bureau of Guangdong Power Grid Co., Ltd., Zhanjiang 524001, China
3
College of Electronics and Information, Hangzhou Dianzi University, Hangzhou 310018, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(10), 3864; https://doi.org/10.3390/ma16103864
Submission received: 11 April 2023 / Revised: 14 May 2023 / Accepted: 14 May 2023 / Published: 20 May 2023
(This article belongs to the Special Issue Advances in Organic Framework Materials: Syntheses and Applications)

Abstract

:
Because of their extensive specific surface area, excellent charge transfer rate, superior chemical stability, low cost, and Earth abundance, nanostructured titanium dioxide (TiO2) arrays have been thoroughly explored during the past few decades. The synthesis methods for TiO2 nanoarrays, which mainly include hydrothermal/solvothermal processes, vapor-based approaches, templated growth, and top-down fabrication techniques, are summarized, and the mechanisms are also discussed. In order to improve their electrochemical performance, several attempts have been conducted to produce TiO2 nanoarrays with morphologies and sizes that show tremendous promise for energy storage. This paper provides an overview of current developments in the research of TiO2 nanostructured arrays. Initially, the morphological engineering of TiO2 materials is discussed, with an emphasis on the various synthetic techniques and associated chemical and physical characteristics. We then give a brief overview of the most recent uses of TiO2 nanoarrays in the manufacture of batteries and supercapacitors. This paper also highlights the emerging tendencies and difficulties of TiO2 nanoarrays in different applications.

1. Introduction

Due to its numerous benefits, including affordability, earthiness, and superior chemical stability, titanium dioxide is a crucial multifunctional substance with several uses in batteries, photocatalysis, and sensors [1,2,3,4,5]. Due to their increased specific surface area and decreased diffusion length, nanostructured TiO2 materials perform better electrochemically when compared to bulk TiO2 materials [6]. Compared to bulk materials, nanostructured TiO2 obtains a larger specific contact area and a decreased diffusion length and therefore performs better electrochemically. In this regard, nanostructured TiO2 is substantially studied as the electrode for Li-ion batteries, Na-ion batteries, supercapacitors, and emerging aqueous batteries [7]. In contrast to thin-film electrodes, which could boost conductivity at the interface between active materials with current collectors despite having low specific surface areas, traditional powder-based electrodes have high specific surface areas but inevitably require binders and conductive agents [8]. Nanostructured arrays have the ability to be free of binders and conductive materials while possessing a significant area of specific surface. Several previous reviews have been reported on the regulation and preparation of TiO2 hierarchical morphologies [9] and their applications [10,11]. Yet, it remains challenging to controllably create nanostructured TiO2 arrays nowadays. Although there are several evaluations concerning TiO2 materials [12], there are very few assessments of the energy storage capabilities of nanostructured TiO2 arrays.
In this study, we emphasize the significant developments in the creation of nanostructured TiO2 arrays in various sizes [13,14,15,16,17]. The uses of nanostructured TiO2 arrays for energy storage are then discussed, with a focus on methods for enhancing electrochemical performance [6,18,19,20,21,22]. It is possible to summarize and predict the optimization of energy storage capabilities by contrasting the electrochemical and morphological characteristics of various TiO2 nanostructured arrays.

2. Morphology of Nanostructured TiO2 Nanoarrays

The phase structures of titania mainly include rutile, anatase, and brookite. At high temperatures, metastable phases such as anatase and brookite will thermodynamically change into rutile with excellent thermodynamic stability. Well-aligned TiO2 nanostructures can be classified as one-dimensional (1D) [13,14,23], two-dimensional (2D) [24], and three-dimensional (3D) [25] nanostructured arrays (Figure 1).

2.1. 1D Nanostructured Arrays

One-dimensional TiO2 nanoarrays grown horizontally on conductive materials and have been thoroughly investigated and utilized as negative anodes for energy storage devices because of their high ion transfer rate [7]. One-dimensional nanostructured TiO2 arrays mainly include nanotubes [26], nanorods [27,28], and nanowires [29,30].
(1)
Nanowire arrays
Alkaline hydrothermal techniques are frequently used to prepare nanowire arrays [31,32]. The alkaline hydrothermal method uses diluted alkaline solution to heat Ti foil at moderately high temperatures (usually 150–220 °C) in a Teflon-lined autoclave to produce a vertical alignment of protonated sodium titanate nanowire arrays on Ti foil. With further proton exchange and annealing, the phase changes from sodium titanate to anatase or rutile TiO2 [32]. The alkaline hydrothermal preparation usually results in nanotubes, nanowires or nanorods, and nanoribbons [33]. Nanosheets are usually observed at an early stage of the preparation or as a small impurity in the final product. Nanobelts are also usually produced at relatively high temperatures (for example, 180 °C) during hydrothermal treatment. Nanowires are often obtained by a calcination of nanotubes at temperatures above 400 °C or a hydrothermal reaction at high temperatures (typical above 200 °C). The size and shape of the arrays can be controlled by varying the hydrothermal parameters and solution concentration (e.g., reaction time can affect the length of the nanowires) [34,35]. Liu et al. produced single crystalline nanowires orientated in the (100) direction, as shown in Figure 2a,b [31]. Similarly, Boercker et al. prepared polycrystalline TiO2 nanowires grown on titanium foil [29]. Moreover, by adjusting the calcination temperatures, TiO2-B nanowires that ranged in diameter from 20 to 40 nm were created by Armstrong’s group [36]. The solvent ethanol was also used to create TiO2-B nanowires with similar size as the former [37].
A variety of methods for creating TiO2 nanowire arrays under mild circumstances have been disclosed. By employing the electrospinning technique on an FTO substrate, Krishnamoorthy’s group created a simple and efficient to create a vertical alignment of anatase TiO2 nanowires [38]. Aligned TiO2 nanoribbons with a length of 25 μm were manufactured by applying an improved electrospinning method on a TCO substrate. The vertical TiO2 nanowires were approximately 27 μm in length and had an average wire width of 90 nm (Figure 2c,d) [39]. In addition, Wu et al. reported an efficient and inexpensive technique to grow TiO2 nanowire arrays on arbitrary substrates (e.g., stainless steel and carbon cloth) by using H2O2 solutions at a low temperature (Figure 2e–i) [23].
(2)
Nanorod arrays
Rutile TiO2 nanorod arrays can be produced by employing concentrated HCl solution [40,41,42], while SO42−, CH3CO2−, and C2O42− cloud create anatase titania [43]. The quasi-1D anatase TiO2 nanorod array made of orientated nanocrystals on a FTO glass was created by a solvothermal technique with tetrabutyl titanate and H2SO4 [43]. Orientated mono-crystal rutile TiO2 nanorod films were manufactured via a simple hydrothermal technique (concentrated HCl) (Figure 3a,b) [27]. The morphology of the nanorods can be modified by adjusting experimental factors such as reaction circumstances and solution concentration and additives. By adjusting the growth parameters, such as growth time, reaction temperature, initial reactant concentration, acidity, and additives, the diameter, length, and density of the nanorods may be altered [44]. Additionally, the epitaxial relationship (lattice mismatch) between active materials and substrate significantly affects the nucleation and development of nanorods [45]. Titanium metal substrates can also be used to grow vertical rutile TiO2 nanorods grown along the [001]-axis, as illustrated in Figure 3c–f [46]. Different reaction conditions can lead to the formation of different crystal structures of 1D TiO2. In general, the nanorods obtained by the acid hydrothermal methods are rutile phases [47]. Those synthesized by alkaline hydrothermal methods [48] and hydrogen peroxide methods [49] are sodium titanate or titanic acid; subsequently, titanic acid can be transformed to titanium dioxide after heat treatment, and the crystal phases of the obtained TiO2 depends on the heat treatment temperature (TiO2(B) at low temperatures, anatase at moderate temperatures, and rutile at high temperatures) [50]. Mixed phases can be obtained by controlling the heat treatment temperature.
Wu et al. produced ordered titania nanorod arrays on titanium surfaces using a method similar to that described above for creating TiO2 nanowire arrays [14]. A coating of condensed anatase nanoparticles (2 μm thick) was laid on the basis of the aligned nanorod arrays (1 μm thick). The majority of the as-deposited nanorods were a combination of rutile and anatase. Rutile nanorods cultivated alongside the path of the [001]-axis.
(3)
Nanotube arrays
Titanium dioxide nanotube arrays can be made by using numerous methods. The main routes can be divided into anodic self-organization, templating, and electrospinning methods [51,52,53].
Anodization techniques can successfully create TiO2 nanotube arrays. In contrast to the uncontrollable hydrothermal method, the anodic oxidation method can modulate the size of nanotubes. The nanotubular structures are affected by several factors, including the application of potential, the electrolyte’s concentration, and the fluoride content [54,55,56]. For instance, the applied voltage can be used to regulate the tube diameter [55], and the anodization period affects the tube length (larger tubular diameters and layer thickness are observed for larger voltage and longer anodization times) [54,57]. The regulation of electrolytes can also affect the tube width and length [54]. The tube length is limited to 500–600 nm in electrolytes at lower pH because of the faster etching rate in acidic electrolytes [58]. In aqueous electrolytes, tubes with widths between 10 and 100 nm generally develop when a voltage of 1 to 25 V is applied in an electrolyte containing 0.1 to 0.5 wt % F [59,60].
Larger diameters can be obtained in organic solutions (e.g., F with a concentration of <1 M in C6H10O4 [61,62,63,64]); the diameter of the tubes can reach 800 nm in the most ideal organic electrolyte [64]. However, aqueous solutions containing F have severe electrochemical etching rates for electrodes, resulting in the length of nanotubes in aqueous electrolytes being capped at a few micrometers. The electrolyte temperature is another important factor that greatly influences the thickness of tubes [65,66]; in fact, by performing anodizing at a low temperature, it can even be possible to practically seal the interior of the tubes and create rod-like structures [66]. Complexing agents such as EDTA (41 μm h−1) [67] or lactic acid (1200 μm h−1) [68] can also be employed to speed up the formation of nanotubes (Figure 4a,b). Under certain anodizing conditions, this can lead to peculiar morphologies, such as the tube-in-tube structure [69]. Titanium dioxide nanotube arrays can be also achieved by alkaline hydrothermal methods. The formation process of the nanotubes can be divided into several stages: the dissolution of TiO2 raw material in alkaline solution occurs simultaneously with the epitaxial growth of sodium trititanate layered nanosheets, followed by the crystallization of dissolved titanate on the exfoliated nanosheets and the generation of physical tension, which induces bending of the nanosheets and the formation of nanotubes [70].
In template methods for the synthesis of nanotubes, modulating the hydrolysis rate of the titanium-containing compound solution (containing the template agent) enables the polymerization of TiO2 in or on the surface of self-assembled template molecules, followed by the selective removal of the template agent. High-aspect-ratio cellulose, micelles, or hard templates can be used as templates to synthesize tubular structures by a range of deposition methods, such as atomic layer deposition [71,72,73,74,75]. Hoyer et al. prepared titanium dioxide nanotubes by electrodeposition on ordered alumina templates [76]. The sol-gel technique can also be used to manufacture TiO2 [77,78,79,80]. An alumina template, for instance, can be selectively dissolved before TiO2 sol has been absorbed into its pores [80]. By reproducing different ZnO nanomorphologies, complex hollow TiO2 nanostructures are successfully created (Figure 4c–i). For example, NH3 saturated deposition can produce TiO2 nanotube arrays with 70 nm diameters [81].
By creating arrays of randomly oriented nanotube electrodes Figure 5a–d), both sealed (Figure 5e–h) and unsealed (Figure 5i–l), Han’s team systematically studied variables that influence the performance of TiO2 electrodes [82]. Their theoretical study and experimental findings revealed that the highly ordered titania nanotube electrode outperforms conventionally random electrodes and cap-sealed electrodes in terms of rate capability. Additionally, electrospinning is another efficient technique for producing TiO2 nanotube arrays. By adding the precursor solution through a capillary spinneret, Li’s group obtained an array of TiO2 hollow nanotubes [83].

2.2. 2D Nanostructured Arrays

Two-dimensional materials include nanobelts and nanosheets featuring a significant aspect ratio [84,85]. Additionally, 2D materials, such as graphene and MoS2, have a variety of unusual physical, chemical, optical, electrical, and magnetic properties [84]. Larger specified areas, more exposed active sites, and certain specifically exposed crystal surfaces can all be provided by 2D TiO2 nanostructured arrays.
(1)
Nanosheets arrays
TiO2 nanosheets with special exposed surfaces such as (010), (101), (001), and (105) facets are attractive for renewable energy due to their different reaction activities [86,87,88,89]. For instance, by etching with hydrofluoric acid, Yang et al. prepared a highly exposed (001) surface of single-crystalline anatase titanium dioxide, and reactive (001) facets have promising applications in photocatalysis [89]. Liu et al. [90] hydrothermally processed Ti foil in a 5 M NaOH solution to create titanate nanosheet arrays on Ti foil. These titanate nanosheets underwent the proper post-processing to produce anatase TiO2/brookite TiO2 heterostructures. Yang et al. [91] used tetrabutyl titanate and ammonium hexafluorotitanate as Ti precursors, where F functions acted as termination agents, producing anatase TiO2 nanosheet arrays with optimal exposure (001) facets on FTO. Lu et al. [92] reported a two-step method to fabricate anatase TiO2 nanosheet arrays with the preferred exposure of [93] facets, where tetrabutyl titanate in hydrofluoric acid/toluene was hydrothermally treated to yield hexagonal TiOF2 nanosheets at first, and then the TiO2 nanosheets could be obtained by calcinating the TiOF2 nanosheets at 500 °C. Nevertheless, each of the aforementioned techniques used a strongly acidic or basic solution, which is corrosive and dangerous. Therefore, Zhong et al. exploited a novel approach to grow anatase TiO2 nanosheet arrays on FTO substrate via a relative green approach by using Na2EDTA and TEOA as co-coordination agents under weak basic conditions (Figure 6a–c) [94]. Gan et al. also used the intermediate kassite [CaTi2O4(OH)2] to create the nanoporous hexagonal TiO2 nanosheet arrays [95]. The developed process involved a TiO nanorod-derived synthesis of upstanding hexagonal kassite nanosheet arrays and a transformation of the kassite to TiO2. The single-crystalline hexagonal kassite was hydrothermally treated with diluted HNO3 aqueous solution and transformed into nanoporous rutile TiO2 nanosheet arrays with shape preservation (Figure 6d). The Ti-O bond of TiO2 nanorods was broken under the hydrothermal treatment of concentrated NaOH, and the binding of Ca2+ ions released from the substrate into the solution with Ti2O4(OH)2− caused the heterogeneous nucleation and growth of CaTi2O4(OH)2. After the nucleation, CaTi2O4(OH)2 nanosheets grew along the (100) direction to reduce the lattice mismatch with the FTO substrate. Subsequently, the density of TiO2 nanorods decreased, the yield and size of nanosheets increased, and finally, TiO2 nanorods were completely dissolved and converted into nanosheets (Figure 6e).
(2)
Nanobelt arrays
TiO2 nanobelt arrays are typically fabricated through an alkaline hydrothermal treatment, followed by a subsequent proton exchange and heat treatment. For instance, Zhuo et al. investigated the route of Ti foil with the alkaline solution (5 M) [16]. Titanate was then heated to transform titanate into TiO2 nanobelt arrays without changing the morphology. Unlike conventional alkali-hydrothermal approaches to titanates, we demonstrated the synthesis of titanate ultrathin nanobelt arrays using a novel and robust H2O2-assisted wet-chemistry route at ambient conditions (Figure 7a–f) [96]. In the absence of any seed layers, the synthesis technique is effective for synthesizing thin films of one- or three-dimensional arrays on a variety of substrates at low temperatures.
TiO2 nanobelt arrays can also serve as a substrate for supporting other active materials. For instance, Luo’s group synthesized the core-shell structure of TiO2 nanobelts@MnO2 nanosheets by the hydrothermal method (Figure 8a,b) [97]. Benefiting from this unique three-dimensional structure, the ultrathin nanosheets are uniformly dispersed on the nanobelts, resulting in a larger contact area. Meanwhile, the vertically grown nanobelt arrays not only act as a strong supporter but also facilitate the charge transport. Similar structures include graphene-wrapped TiO2@Co3O4 coaxial nanobelt arrays, as illustrated in Figure 8c,d [98].

2.3. 3D Nanostructured Arrays

Three-dimensional nanoarrays have a larger specific surface area than that of 1D and 2D arrays. Additionally, they are usually prepared by one-step or two-step methods, namely one-step growth [99] and the multi-step surface branching decoration of pre-formed 1D morphology [100]. The preparation is simpler and more convenient in the former, but the controllability is not as good as that of the latter.
Decorating trunks with various branches, such as nanowires, nanosheets, nanorods, or nanoparticles, is useful to increase the contact surface area. By modifying the experimental parameters (e.g., precursor solution [101] and reaction rate [102]), the shape of branches can be controlled, for example, in the case of rutile TiO2 nanowire branches on a rutile TiO2 nanowire [88] and a hyperbranched hierarchical anatase TiO2 nanowire on pre-performed anatase TiO2 nanowire trunks [103]. Yang’s group developed rutile TiO2 nanoarrays by the acidic evaporation method, and the rutile (101) twinning structure promoted the form of nanotrees (Figure 9a–f) [104]. The reaction mechanism of this process was the oxidation of rutile TiO2 by HCl vapor and the induction of its nucleation and growth along the c-direction. Then, the acidic vapor was able to induce (101) twinning of the sample, thus inducing the nucleation and growth of branches on the (100) side of the surface with higher surface energy. With the growth of the second branch and chemical erosion on both sides of the branch, the third branches could be generated on the initially grown nanorods at the previously eroded sites. In addition, we produced 3D anatase TiO2 nanostructured arrays with TiO2 nanoparticle branches (Figure 9g,h) [105]. Specifically, TiO2 nanowire grown by H2O2-assisted method was subsequently deposited in a liquid deposition technology to obtain a three-dimensional structure with abundant anatase TiO2 nanoparticles.
Typical cases include rutile TiO2 nanorods on pre-formed rutile TiO2 nanorod trunks [106] and anatase TiO2 nanorod/nanosheet branches on pre-performed anatase TiO2 nanowire [107,108]. Using the multi-step hydrothermal preparation method, Wu et al. constructed directional layered heterogeneous TiO2 hyperbranched array materials. These three-dimensional structures were composed of basic units such as nanowires, nanosheets, and nanorods. (Figure 10a–f) [109], and the 3D monoblocks led to a substantially enlarged contact area.
Liu et al. built nanotube networks, namely hierarchical TiO2 nanorod arrays composed of rutile trunks and anatase nanotube branches, by using ZnO nanorods served as the template (Figure 10g–i) [25]. The backbones of three-dimensional TiO2 arrays mainly include nanowires, nanotubes, and nanorods. A pre-formed nanowire trunk is often created using the alkaline hydrothermal process as the aforementioned method for 1D nanowire arrays [103]. Additionally, using an H2O2-assisted approach [105], we can obtain bramble-like anatase TiO2 nanoarrays [110]. We also obtained three-dimensional TiO2 nanorod arrays with a biphasic mixture (rutile/anatase) of balsam-pear-like shapes through a series of simple and green preparation methods [111].
Using nanotube arrays as structural support allows for an increased surface area (more inner surface exposure) and decreased electrolyte/electrode interfacial resistance. Roh’s team disclosed a simple method for creating hierarchical TiO2 nanotube arrays (Figure 11a–c) [112]. The hierarchical TiO2 nanotube arrays are made up of short nanorod branches and long nanotube trunk with a vertical orientation. Various morphologies for the hierarchical TiO2 nanotube arrays can be obtained by adjusting the preparation parameters. We prepared a 3D electrode that consists of anatase TiO2 mesocrystal branches and single-crystal-like TiN nanowire trunks (Figure 11d,e) [113]. These trunks provide a stable structure during charging and enhance electron migration. Anatase/rutile TiO2 nanoflower arrays were also synthesized by proton exchange of titanate with the support of K2S2O8 (Figure 12a–c) [114].
Single-crystal nanoarrays are highly attractive because they can simultaneously enlarge the contact area and enhance the reaction kinetics of ions. Sheng’s team developed single-crystal rutile TiO2 nanowire arrays [88]. The branches grew through the backbone along the (100) and (010) crystallographic planes in four symmetric directions of epitaxy, with a typical angle of 65° between the c-axes of the backbone and branches. These branched nanowires extending from the backbone not only expanded the contact area between the electrode material and the electrolyte but also accelerated the charge transport. In addition, strain engineering can optimize the physical and chemical properties of TiO2 [116]. We prepared quasi-single-crystal anatase TiO2 branched nanowire arrays and adjusted the lattice constant by the “convertible precursor induced growth” method, extending the lattice constant α by 0.37% (Figure 12d–f) [115]. During the growth of anatase TiO2 mesocrystals on the surface of single-crystal sodium titanate using the liquid-phase deposition method, an extended lattice parameter a (3.804 Å) was introduced because the lattice parameter b of sodium titanate was slightly higher than the lattice parameter a of anatase TiO2 (Figure 12g,h). In addition, TiO2 3D nanoarrays can also be used for the synthesis of TiN nanoarrays with unique structures. We uncovered a novel “surface-induced” Kirkendall effect that causes titanate-decorated TiO2 3D nanoarrays to be nitridated to form distinctive hollow TiN nanotrees (Figure 13a–f) [117]. The stacking of titanate nanosheets can accelerate chemical reactions and transform into a nitride layer. The hollow structures are caused by the Kirkendall effect. Table 1 provides brief summary of various synthesis methods of TiO2 nanoarrays.

3. Energy Storage Applications of Nanostructured TiO2 Arrays

Using aligned TiO2 nanoarray materials as electrodes for energy storage has many benefits [118,119,120,121]. Firstly, nano-array structures reduce the impedance at the interfaces of electrodes and electrolytes. Secondly, aligned nanostructures can provide direct electrical transfer pathways. Finally, the direct connection of electrodes to the current collectors eliminates the usage of binders and conducting additives [122,123,124].

3.1. Lithium-Ion Batteries

Battery performance is highly influenced by the characteristics of the electrodes. There is a need for fundamental advancements in terms of high energy and low costs. Since TiO2 has an appropriate discharge plateau (approximately 1.5–1.75 V vs. lithium), which prevents organic electrolytes from decomposition, it has been regarded as a good alternative to graphite [125,126,127]. However, it has been found that the achievable capacities of bulk TiO2 are quite low [128,129] due to the repulsive force among Li ions. Highly ordered TiO2 nanoarrays facilitate Li-ion transportation [130,131,132,133]. The relevant works are outlined in this section based on the TiO2 polymorphs.
Since lithium ions are more easily transported along the c-direction [134,135], rutile TiO2 can improve the electrochemical performance by adjusting the growth direction of nanoarrays [136,137,138]. The c-direction TiO2 nanorod arrays displayed outstanding cycling performance, with 93% capacity containing after 600 cycles [139]. Similarly, Dong et al. also prepared rutile TiO2 nanorod arrays growing along the direction of the c-axis. The capacity of the active materials for Li-ions insertion/extraction is 10 times higher than that of the compact layer. And it remained to 133 μAh cm−2 (15 μA cm−2) after 50th cycle [46]. Several studies have reported other nanoarray configurations, including nanotubes, with outstanding electrochemical performances. For instance, Guan et al. developed rutile TiO2 nanotube arrays by anodizing Ti foils. These arrays exhibited enhanced performances, maintaining 77% capacity after 100th cycle [140].
When charged and discharged with low currents, an anatase phase can reversibly absorb 0.5 Li per formula unit of TiO2 through a biphasic process [141]. Tang et al. produced TiO2 nanowire arrays with dense nanocavities with 305.8 mAh g−1 (0.2 C) [142].
Compared to the above two phases, TiO2-B is more suitable for storing lithium ions [143,144], and many works showed the excellent electrochemical performances of TiO2-B nanoarrays. Considerable efforts have been devoted to the synthesis of TiO2-B nanomaterials with various morphologies through different methods such as hydrothermal, sol-gel, and solvothermal methods [104,126,145,146,147]. Liu et al. grew vertically oriented single-crystalline TiO2-B nanowire arrays by placing Ti foil in a hydrothermal solution during the hydrothermal process. Similar single-crystal TiO2-B nanowire arrays by Liu’s group had an outstanding cycling calendar with 120 mAh g−1 even at 1.8 C [148]. Other approaches, such as Tang’s team’s, synthesized TiO2(B) nanowire arrays with 124.9 mAh/g at 2 C [149].
Phase mixing further enhances the lithium ions’ storage [150,151,152,153]. We constructed 3D TiO2 nanotrees by depositing ultrathin nanoribbons of the anatase/TiO2-B hybrid phase on single-crystal anatase nanowire arrays. The anatase is responsible for electron reception, while the TiO2(B) performs the intercalation of ions, leading to the effective separation of ions and electrons, thus enhancing the battery’s performance. In addition, the ultra-long branching structure of the nanoribbons results in the excellent electrochemical performance of the electrode (Figure 14a–e) [154].
The morphology of TiO2 nanoarrays also has a huge impact on the performances of batteries [121]. For example, layered structures can ensure high capacities and good structural stability during Li ion insertion/extraction, which is attributed to shorter transport lengths for Li+. In addition, a larger specific surface area could increase the electrode–electrolyte contact area and facilitate the lithium ions’ insertion and extraction [157,158,159,160,161,162,163].
Element doping and compositing TiO2 nanostructured arrays with other materials can also be good strategies to enhance the performance of electrode materials [164,165]. For instance, Kyeremateng’s group developed Sn doping TiO2 nanotube arrays. The phase transformation of TiO2 was affected by Sn (From anatase to rutile) [166,167,168,169]. With a current density of 70 μA cm−2 (1 C), Sn-doped TiO2 nanotubes delivered much higher capacity values compared to TiO2 nanotubes. An electrochemical test showed that at 70 μA cm−2, TiO2 nanotubes with Sn doping possessed a significantly larger capacity than conventional nanotubes. High-capacity Fe2O3 was deposited onto the surface of 1D TiO2 nanoarrays to improve the capacity of the electrodes [170]. To boost the electrode’s capacity, composition with high-capacity Fe2O3 is an additional option. Furthermore, we prepared core-shell nanowire arrays to enhance the electrical conductivity and ion diffusion rate [113]. The single-crystal TiN nanowires acted as conductive collectors, while the branching shells were composed of nanoporous anatase TiO2. This specially structured 3D array exhibited pseudocapacitive-dominated charge storage with excellent electrochemical properties (Figure 15a–g).
Post-annealing TiO2 nanoarrays in a reducing atmosphere can also substantially improved the capacity [172,173]. Lu et al. demonstrated that hydrogenation treatment processed an abundance of oxygen vacancies, which improved the electronic conductivity and thus resulted in excellent rate performances [21]. Table 2 provides an overview of the performance parameters of nanoarrays of TiO2 in lithium-ion batteries.

3.2. Sodium-Ion Batteries

Due to their availability, affordability, and environmental friendliness, sodium-ion batteries have drawn a lot of interest [194,195]. Nonetheless, there is still a critical demand for the discovery of superior anode materials [196,197].
Highly ordered TiO2 nanoarrays feature the intrinsic merits for a larger contact area and the ability to release inner stress [197]. Ruffo et al. synthesized nanostructured TiO2 with different morphologies. Thus, the exposition of different crystalline surfaces brought essentially different performances [198]. Bella et al. produced and compared the amorphous and anatase TiO2 nanotubular arrays obtained by an anodic oxidation [199]. After the first cycling, anatase TiO2 had superior electrochemical properties compared to its amorphous counterpart. In addition, they also demonstrated that anatase had optimal cyclic stability, and the channels along the [001] direction facilitated the transport of Na ions [200,201].
It was reported that TiO2@C nanotube arrays exhibited an excellent capacity with 232 mAh g−1 after 500th [202]. And three-dimensional Ni@TiO2 core/shell nanoarrays showed ∼312 mAh g−1 after 100th at 50 mA g−1 [203]. The co-doping technique (such as Ni and N doping) can further improve the Na-ion storage [19]. It was also reported that compositing with the high capacity and conductivity of 2D MoS2 nanosheets could improve the electronical performance [204]. The capacity can also be enhanced by the treatment of functionalized salts. Ni et al. produced phosphorylated TiO2 nanotube arrays with a large capacitance of 147 mAh g−1 at 3350 mA g−1 and an excellent cycling calendar [205]. Table 3 provides an overview of the performance parameters of nanoarrays of TiO2 in sodium-ion batteries.

3.3. Supercapacitors

TiO2 nanoarrays can also be used for aqueous supercapacitors. Aqueous supercapacitors are safer and can be charged and discharged fast, despite having a low energy density. TiO2 nanoarrays are favorable for their faster charge transport and interfacial ion mobility, apart from a higher surface area for ion adsorption [208,209].
For example, Kim’s group applied TiO2 nanotube array electrode materials to an EDLC-type device, and the experimental results were 2.4 mF cm−2 at 50 mV s−1 [210]. The TiO2 nanoarray structure used directly as an electrode material had an area capacitance (1 mV s−1/911 μF cm−2) several orders magnitudes larger when compared to the capacitance of nanoparticles (1 mV s−1/180 μF cm−2) [211]. Additionally, the performance of the supercapacitor was proportional to the aspect ratio of the nanotubes [212].
Despite the inherent low area capacitance properties of TiO2, relatively high performance can be obtained by a range of methods [208,213,214,215]. Compositing with C materials [216,217,218], H2 treatment [219], plasma treatment [220], electrochemical doping [221,222,223], and synthesizing black or blue TiO2 [224,225], have also been performed to improve the conductivity and capacitances of TiO2 nanoarrays. For example, Kim et al. reported that NiO-TiO2 nanotube arrays were obtained by the anodic oxidation of a Ni–Ti alloy, which showed an excellent performance for supercapacitors [226]. Additionally, Patil et al. fabricated 3D–1D TiO2 hierarchical nanostructures by using straightforward chemical methods, which had the benefits of improved electron transportation properties (contributed by 1D TiO2 nanotubes) and high surface area (as a result of the 3D TiO2 flower) for improved area capacitance [227].
In recent times, we can also enhance the performance of supercapacitors by surface fluorine-modified anatase TiO2 nanoarrays [105]. With an improved capacity of more than 6.4-fold, the calculation results showed that the fluorine modification altered the storing sites and transport routes, which caused a substantial reduction in the diffusion potential (Figure 16a–g). Table 4 provides an overview of the performance parameters of nanoarrays of TiO2 in supercapacitors.

3.4. Other Batteries

Due to their enormous theoretical energy density, multivalent ion batteries, such as those based on magnesium [246], aluminum [247], and zinc [248], have attracted a lot of research. The Al3+ with a small radius (53.5 pm) can perform the intercalation reaction [248]. For instance, anatase TiO2 nanotube arrays possess reversible capacity of 75 mAh g−1 in 1 M AlCl3 aqueous solution [249]. In addition, the experimental results show that ions can reversibly insert/extract into/from TiO2.
Battery voltage is constrained by the water’s electrochemical window [250], thus limiting the energy density. By increasing the salt concentration, the activity of water can be drastically reduced, and the potential window can be extended [251]. Zhou’s group developed TiO2-B nanotube arrays as an anode in the electrolyte of 50 M LiTFSI and 25 M TMBTFSI, and 194.5 mAh/g and 150 Wh/kg could be obtained when the anode mated with LiMn2O4 [252].
Hydrogen ions, which possess the smallest ionic radius, are ideal charge carriers [253]. However, H+ are normally present as a form of H3O+ with high dehydration energy [254]. Recently we have demonstrated that anatase TiO2 (001) enables the decomposition of H3O+ to H+ [110], and the anatase TiO2 nanowire array functions superbly as the anode of the proton battery (Figure 17).

4. Summary and Outlook

This review summarizes and compares the design principles, synthesis, and applications of TiO2 nanoarrays. Emerging opportunities for TiO2 nanostructured arrays in a variety of domains are also discussed in detail. The goal of this review is to figure out the current applications and challenges associated with TiO2 nanostructured arrays, proving an in-depth look at TiO2 nanostructured arrays. TiO2 nanostructured arrays represent an important category of nanomaterials in energy applications, such as lithium-ion batteries, sodium-ion batteries, supercapacitors, and other emerging batteries. However, the preparation scale is still at the laboratory level, and the controllability of the material growth is still insufficient. For industrial applications, simple and controllable scale production methods need to be developed in the future. The current material growth mechanism is also not very clear, which impedes the accurate design of the material. It needs in situ spherical differential TEM, in situ synchrotron radiation, and other advanced characterizations to reveal the relevant mechanism, which is conducive to the accurate design and controllable preparation of titanium dioxide arrays. Moreover, there has not been much research showing that these materials have a solid theoretical design. Scientists will have the capability to forecast the characteristics of TiO2 nanostructured arrays based on their material components, size, and morphology when the theoretical foundations and computational power are improved. Therefore, TiO2 nanostructured arrays’ components and structure may be tailored to meet the desired characteristics for various applications. The possibility for more multidisciplinary research on this subject will increase, resulting in further TiO2 nanostructured array applications in several new fields.

Funding

This research was funded by the Hainan Provincial Natural Science Foundation of China (No. 521RC740 and 2019RC047), National Natural Science Foundation of China (No. 51862005 and 52201211), and Hainan Province Science and Technology Special Fund (No. ZDYF2020175).

Acknowledgments

This work was financially supported by the Hainan Provincial Natural Science Foundation of China (No. 521RC740 and 2019RC047), National Natural Science Foundation of China (No. 51862005 and 52201211), and Hainan Province Science and Technology Special Fund (No. ZDYF2020175).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Banerjee, S.; Dionysiou, D.D.; Pillai, S.C. Self-Cleaning Applications of TiO2 by Photo-Induced Hydrophilicity and Photocatalysis. Appl. Catal. B Environ. 2015, 176–177, 396–428. [Google Scholar] [CrossRef]
  2. Bai, J.; Zhou, B. Titanium Dioxide Nanomaterials for Sensor Applications. Chem. Rev. 2014, 114, 10131–10176. [Google Scholar] [CrossRef] [PubMed]
  3. Shi, H.; Magaye, R.; Castranova, V.; Zhao, J. Titanium Dioxide Nanoparticles: A Review of Current Toxicological Data. Part. Fibre Toxicol. 2013, 10, 15. [Google Scholar] [CrossRef] [PubMed]
  4. Shetti, N.P.; Bukkitgar, S.D.; Reddy, K.R.; Reddy, C.V.; Aminabhavi, T.M. Nanostructured Titanium Oxide Hybrids-Based Electrochemical Biosensors for Healthcare Applications. Colloids Surf. B Biointerfaces 2019, 178, 385–394. [Google Scholar] [CrossRef]
  5. Henderson, M.A. A Surface Science Perspective on TiO2 Photocatalysis. Surf. Sci. Rep. 2011, 66, 185–297. [Google Scholar] [CrossRef]
  6. Su, X.; Wu, Q.; Zhan, X.; Wu, J.; Wei, S.; Guo, Z. Advanced Titania Nanostructures and Composites for Lithium Ion Battery. J. Mater. Sci. 2012, 47, 2519–2534. [Google Scholar] [CrossRef]
  7. Wang, X.; Li, Z.; Shi, J.; Yu, Y. One-Dimensional Titanium Dioxide Nanomaterials: Nanowires, Nanorods, and Nanobelts. Chem. Rev. 2014, 114, 9346–9384. [Google Scholar] [CrossRef]
  8. Yang, Z.; Choi, D.; Kerisit, S.; Rosso, K.M.; Wang, D.; Zhang, J.; Graff, G.; Liu, J. Nanostructures and Lithium Electrochemical Reactivity of Lithium Titanites and Titanium Oxides: A Review. J. Power Sources 2009, 192, 588–598. [Google Scholar] [CrossRef]
  9. Cargnello, M.; Gordon, T.R.; Murray, C.B. Solution-Phase Synthesis of Titanium Dioxide Nanoparticles and Nanocrystals. Chem. Rev. 2014, 114, 9319–9345. [Google Scholar] [CrossRef]
  10. Liu, G.; Yang, H.G.; Pan, J.; Yang, Y.Q.; Lu, G.Q.; Cheng, H.M. Titanium Dioxide Crystals with Tailored Facets. Chem. Rev. 2014, 114, 9559–9612. [Google Scholar] [CrossRef]
  11. Paramasivam, I.; Jha, H.; Liu, N.; Schmuki, P. A Review of Photocatalysis Using Self-Organized TiO2 Nanotubes and Other Ordered Oxide Nanostructures. Small 2012, 8, 3073–3103. [Google Scholar] [CrossRef] [PubMed]
  12. Wu, W.-Q.; Feng, H.-L.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. Recent Advances in Hierarchical Three-Dimensional Titanium Dioxide Nanotree Arrays for High-Performance Solar Cells. J. Mater. Chem. A 2017, 5, 12699–12717. [Google Scholar] [CrossRef]
  13. Cao, X.; Wu, J.-M.; Zhang, Z.; Qin, J.; Fu, Z.; Hai, J.; Wang, Z.; Ye, Z.; Wen, W. Electrochemical Lithium Intercalation in Titania Nanowire Arrays for Boosting Photocatalytic Activity. J. Electron. Mater. 2023, 52, 23–30. [Google Scholar] [CrossRef]
  14. Wu, J.-M.; Zhang, T.-W.; Zeng, Y.-W.; Hayakawa, S.; Tsuru, K.; Osaka, A. Large-Scale Preparation of Ordered Titania Nanorods with Enhanced Photocatalytic Activity. Langmuir 2005, 21, 6995–7002. [Google Scholar] [CrossRef]
  15. Ge, M.; Li, Q.; Cao, C.; Huang, J.; Li, S.; Zhang, S.; Chen, Z.; Zhang, K.; Al-Deyab, S.S.; Lai, Y. One-Dimensional TiO2 Nanotube Photocatalysts for Solar Water Splitting. Adv. Sci. 2017, 4, 1600152. [Google Scholar] [CrossRef]
  16. Zhuo, Y.; Wu, C.; Han, S.; Chi, B.; Pu, J.; Jin, T.; Jian, L. Oriented Nanostructured Titanates Array from Low Concentration Alkaline Solution via Hydrothermal Process. J. Nanosci. Nanotechnol. 2011, 11, 2298–2304. [Google Scholar] [CrossRef]
  17. Yu, Y.; Li, J.; Geng, D.; Wang, J.; Zhang, L.; Andrew, T.L.; Arnold, M.S.; Wang, X. Development of Lead Iodide Perovskite Solar Cells Using Three-Dimensional Titanium Dioxide Nanowire Architectures. ACS Nano 2015, 9, 564–572. [Google Scholar] [CrossRef]
  18. Fang, D.; Huang, K.; Liu, S.; Li, Z. Electrochemical Properties of Ordered TiO2 Nanotube Loaded with Ag Nano-Particles for Lithium Anode Material. J. Alloys Compd. 2008, 464, L5–L9. [Google Scholar] [CrossRef]
  19. Yan, D.; Yu, C.; Zhang, X.; Li, J.; Li, J.; Lu, T.; Pan, L. Enhanced Electrochemical Performances of Anatase TiO2 Nanotubes by Synergetic Doping of Ni and N for Sodium-Ion Batteries. Electrochim. Acta 2017, 254, 130–139. [Google Scholar] [CrossRef]
  20. Zhang, J.; Wang, Y.; Wu, J.; Shu, X.; Yu, C.; Cui, J.; Qin, Y.; Zhang, Y.; Ajayan, P.M.; Wu, Y. Remarkable Supercapacitive Performance of TiO2 Nanotube Arrays by Introduction of Oxygen Vacancies. Chem. Eng. J. 2017, 313, 1071–1081. [Google Scholar] [CrossRef]
  21. Lu, Z.; Yip, C.-T.; Wang, L.; Huang, H.; Zhou, L. Hydrogenated TiO2 Nanotube Arrays as High-Rate Anodes for Lithium-Ion Microbatteries. ChemPlusChem 2012, 77, 991–1000. [Google Scholar] [CrossRef]
  22. Bi, Z.; Paranthaman, M.P.; Menchhofer, P.A.; Dehoff, R.R.; Bridges, C.A.; Chi, M.; Guo, B.; Sun, X.-G.; Dai, S. Self-Organized Amorphous TiO2 Nanotube Arrays on Porous Ti Foam for Rechargeable Lithium and Sodium Ion Batteries. J. Power Sources 2013, 222, 461–466. [Google Scholar] [CrossRef]
  23. Li, B.; Wu, J.-M.; Guo, T.-T.; Tang, M.-Z.; Wen, W. A Facile Solution Route to Deposit TiO2 Nanowire Arrays on Arbitrary Substrates. Nanoscale 2014, 6, 3046. [Google Scholar] [CrossRef]
  24. Lei, B.-X.; Zheng, X.-F.; Qiao, H.; Li, Y.; Wang, S.-N.; Huang, G.-L.; Sun, Z.-F. A Novel Hierarchical Homogeneous Nanoarchitecture of TiO2 Nanosheets Branched TiO2 Nanosheet Arrays for High Efficiency Dye-Sensitized Solar Cells. Electrochim. Acta 2014, 149, 264–270. [Google Scholar] [CrossRef]
  25. Liu, B.; Sun, Y.; Wang, X.; Zhang, L.; Wang, D.; Fu, Z.; Lin, Y.; Xie, T. Branched Hierarchical Photoanode of Anatase TiO2 Nanotubes on Rutile TiO2 Nanorod Arrays for Efficient Quantum Dot-Sensitized Solar Cells. J. Mater. Chem. A 2015, 3, 4445–4452. [Google Scholar] [CrossRef]
  26. Zhou, Q.; Fang, Z.; Li, J.; Wang, M. Applications of TiO2 Nanotube Arrays in Environmental and Energy Fields: A Review. Microporous Mesoporous Mater. 2015, 202, 22–35. [Google Scholar] [CrossRef]
  27. Liu, B.; Aydil, E.S. Growth of Oriented Single-Crystalline Rutile TiO2 Nanorods on Transparent Conducting Substrates for Dye-Sensitized Solar Cells. J. Am. Chem. Soc. 2009, 131, 3985–3990. [Google Scholar] [CrossRef]
  28. Zhao, L.; Huang, X.; Lin, G.; Peng, Y.; Chao, J.; Yi, L.; Huang, X.; Li, C.; Liao, W. Structure Engineering in Hexagonal Tungsten Trioxide/Oriented Titanium Dioxide Nanorods Arrays with High Performances for Multi-Color Electrochromic Energy Storage Device Applications. Chem. Eng. J. 2021, 420, 1385–8947. [Google Scholar] [CrossRef]
  29. Boercker, J.E.; Enache-Pommer, E.; Aydil, E.S. Growth Mechanism of Titanium Dioxide Nanowires for Dye-Sensitized Solar Cells. Nanotechnology 2008, 19, 095604. [Google Scholar] [CrossRef]
  30. Ji, Q.; Hou, Y.; Wei, S.; Liu, Y.; Du, P.; Luo, L.; Li, W.P. Excellent Energy Storage Performance in Bilayer Composites Combining Aligned TiO2 Nanoarray and Random TiO2 Nanowires with Poly(vinylidene fluoride). J. Phys. Chem. 2020, 124, 2864–2871. [Google Scholar] [CrossRef]
  31. Liu, B.; Boercker, J.E.; Aydil, E.S. Oriented Single Crystalline Titanium Dioxide Nanowires. Nanotechnology 2008, 19, 505604. [Google Scholar] [CrossRef]
  32. Liao, J.-Y.; Lei, B.-X.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. Oriented Hierarchical Single Crystalline Anatase TiO2 Nanowire Arrays on Ti-Foil Substrate for Efficient Flexible Dye-Sensitized Solar Cells. Energy Env. Sci 2012, 5, 5750–5757. [Google Scholar] [CrossRef]
  33. Bavykin, D.V.; Friedrich, J.M.; Walsh, F.C. Protonated Titanates and TiO2 Nanostructured Materials: Synthesis, Properties, and Applications. Adv. Mater. 2006, 18, 2807–2824. [Google Scholar] [CrossRef]
  34. Lei, B.-X.; Luo, Q.-P.; Yu, X.-Y.; Wu, W.-Q.; Su, C.-Y.; Kuang, D.-B. Hierarchical TiO2 Flowers Built from TiO2 Nanotubes for Efficient Pt-Free Based Flexible Dye-Sensitized Solar Cells. Phys. Chem. Chem. Phys. 2012, 14, 13175. [Google Scholar] [CrossRef]
  35. Wu, W.-Q.; Xu, Y.-F.; Rao, H.-S.; Su, C.-Y.; Kuang, D.-B. A Double Layered TiO2 Photoanode Consisting of Hierarchical Flowers and Nanoparticles for High-Efficiency Dye-Sensitized Solar Cells. Nanoscale 2013, 5, 4362. [Google Scholar] [CrossRef] [PubMed]
  36. Armstrong, A.R.; Armstrong, G.; Canales, J.; Bruce, P.G. TiO2–B Nanowires as Negative Electrodes for Rechargeable Lithium Batteries. J. Power Sources 2005, 146, 501–506. [Google Scholar] [CrossRef]
  37. Wang, Y.; Wu, M.; Zhang, W.F. Preparation and Electrochemical Characterization of TiO2 Nanowires as an Electrode Material for Lithium-Ion Batteries. Electrochim. Acta 2008, 53, 7863–7868. [Google Scholar] [CrossRef]
  38. Krishnamoorthy, T.; Thavasi, V.; Subodh, G.M.; Ramakrishna, S. A First Report on the Fabrication of Vertically Aligned Anatase TiO2 Nanowires by Electrospinning: Preferred Architecture for Nanostructured Solar Cells. Energy Environ. Sci. 2011, 4, 2807. [Google Scholar] [CrossRef]
  39. Liu, N.; Chen, X.; Zhang, J.; Schwank, J.W. A Review on TiO2-Based Nanotubes Synthesized via Hydrothermal Method: Formation Mechanism, Structure Modification, and Photocatalytic Applications. Catal. Today 2014, 225, 34–51. [Google Scholar] [CrossRef]
  40. Feng, X.; Shankar, K.; Varghese, O.K.; Paulose, M.; Latempa, T.J.; Grimes, C.A. Vertically Aligned Single Crystal TiO2 Nanowire Arrays Grown Directly on Transparent Conducting Oxide Coated Glass: Synthesis Details and Applications. Nano Lett. 2008, 8, 3781–3786. [Google Scholar] [CrossRef]
  41. Liu, J.; Sheng, X.; Guan, F.; Li, K.; Wang, D.; Chen, L.; Feng, X. Length-Independent Charge Transport of Well-Separated Single-Crystal TiO2 Long Nanowire Arrays. Chem. Sci. 2018, 9, 7400–7404. [Google Scholar] [CrossRef] [PubMed]
  42. Lv, M.; Zheng, D.; Ye, M.; Xiao, J.; Guo, W.; Lai, Y.; Sun, L.; Lin, C.; Zuo, J. Optimized Porous Rutile TiO2 Nanorod Arrays for Enhancing the Efficiency of Dye-Sensitized Solar Cells. Energy Environ. Sci. 2013, 6, 1615. [Google Scholar] [CrossRef]
  43. Liao, J.-Y.; Lei, B.-X.; Wang, Y.-F.; Liu, J.-M.; Su, C.-Y.; Kuang, D.-B. Hydrothermal Fabrication of Quasi-One-Dimensional Single-Crystalline Anatase TiO2 Nanostructures on FTO Glass and Their Applications in Dye-Sensitized Solar Cells. Chem. Eur. J. 2011, 17, 1352–1357. [Google Scholar] [CrossRef]
  44. Albu, S.P.; Roy, P.; Virtanen, S.; Schmuki, P. Self-Organized TiO2 Nanotube Arrays: Critical Effects on Morphology and Growth. Isr. J. Chem. 2010, 50, 453–467. [Google Scholar] [CrossRef]
  45. Howard, C.J.; Sabine, T.M.; Dickson, F. Structural and Thermal Parameters for Rutile and Anatase. Acta Crystallogr. B 1991, 47, 462–468. [Google Scholar] [CrossRef]
  46. Dong, S.; Wang, H.; Gu, L.; Zhou, X.; Liu, Z.; Han, P.; Wang, Y.; Chen, X.; Cui, G.; Chen, L. Rutile TiO2 Nanorod Arrays Directly Grown on Ti Foil Substrates towards Lithium–Ion Micro-Batteries. Thin Solid Films 2011, 519, 5978–5982. [Google Scholar] [CrossRef]
  47. Li, J.; Cao, K.; Li, Q.; Xu, D. Tetragonal Faceted-Nanorods of Anatase TiO2 with A Large Percentage of Active {100} Facets and Their Hierarchical Structure. CrystEngComm 2012, 14, 83–85. [Google Scholar] [CrossRef]
  48. Tian, J.; Zhao, Z.; Kumar, A.; Boughton, R.I.; Liu, H. Recent Progress in Design, Synthesis, and Applications of One-Dimensional TiO2 Nanostructured Surface Heterostructures: A Review. Chem. Soc. Rev. 2014, 43, 6920–6937. [Google Scholar] [CrossRef]
  49. Yang, C.; Xiaoli, C. Titanium Dioxide Anode Materials for Lithium-ion Batteries. Prog. Chem. 2021, 33, 1249–1269. [Google Scholar] [CrossRef]
  50. Li, W.; Elzatahry, A.; Aldhayan, D.; Zhao, D. Core-Shell Structured Titanium Dioxide Nanomaterials for Solar Energy Utilization. Chem. Soc. Rev. 2018, 47, 8203–8237. [Google Scholar] [CrossRef]
  51. Arul, R.; Dong, J.; Simpson, M.C.; Gao, W. LIPSS-Sticks: Laser Induced Double Self Organization Enhances the Broadband Light Absorption of TiO2 Nanotube Arrays. Adv. Photonics Res. 2021, 2, 2000133. [Google Scholar] [CrossRef]
  52. Qiu, J.; Yu, W.; Gao, X.; Li, X. Sol–Gel Assisted ZnO Nanorod Array Template to Synthesize TiO2 Nanotube Arrays. Nanotechnology 2006, 17, 4695–4698. [Google Scholar] [CrossRef] [PubMed]
  53. Zheng, Z.; Zhuge, F.; Wang, Y.; Zhang, J.; Gan, L.; Zhou, X.; Li, H.; Zhai, T. Decorating Perovskite Quantum Dots in TiO2 Nanotubes Array for Broadband Response Photodetector. Adv. Funct. Mater. 2017, 27, 1703115. [Google Scholar] [CrossRef]
  54. Macak, J.M.; Tsuchiya, H.; Ghicov, A.; Yasuda, K.; Hahn, R.; Bauer, S.; Schmuki, P. TiO2 Nanotubes: Self-Organized Electrochemical Formation, Properties and Applications. Curr. Opin. Solid State Mater. Sci. 2007, 11, 3–18. [Google Scholar] [CrossRef]
  55. Roy, P.; Berger, S.; Schmuki, P. TiO2 Nanotubes: Synthesis and Applications. Angew. Chem. Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef] [PubMed]
  56. Zhou, X.; Nguyen, N.T.; Özkan, S.; Schmuki, P. Anodic TiO2 Nanotube Layers: Why Does Self-Organized Growth Occur—A Mini Review. Electrochem. Commun. 2014, 46, 157–162. [Google Scholar] [CrossRef]
  57. Macák, J.M.; Tsuchiya, H.; Schmuki, P. High-Aspect-Ratio TiO2 Nanotubes by Anodization of Titanium. Angew. Chem. Int. Ed. 2005, 44, 2100–2102. [Google Scholar] [CrossRef]
  58. Jiang, H.B.; Cuan, Q.; Wen, C.Z.; Xing, J.; Wu, D.; Gong, X.Q.; Li, C.; Yang, H.G. Anatase TiO2 Crystals with Exposed High-Index Facets. Angew. Chem. Int. Ed. Engl. 2011, 50, 3764–3768. [Google Scholar] [CrossRef]
  59. Park, J.; Bauer, S.; von der Mark, K.; Schmuki, P. Nanosize and Vitality: TiO2 Nanotube Diameter Directs Cell Fate. Nano Lett. 2007, 7, 1686–1691. [Google Scholar] [CrossRef]
  60. Bauer, S.; Kleber, S.; Schmuki, P. TiO2 Nanotubes: Tailoring the Geometry in H3PO4/HF Electrolytes. Electrochem. Commun. 2006, 8, 1321–1325. [Google Scholar] [CrossRef]
  61. Lan, M.-Y.; Liu, C.-P.; Huang, H.-H.; Chang, J.-K.; Lee, S.-W. Diameter-Sensitive Biocompatibility of Anodic TiO2 Nanotubes Treated with Supercritical CO2 Fluid. Nanoscale Res. Lett. 2013, 8, 150. [Google Scholar] [CrossRef] [PubMed]
  62. Albu, S.P.; Ghicov, A.; Macak, J.M.; Schmuki, P. 250 Μm Long Anodic TiO2 Nanotubes with Hexagonal Self-Ordering. Phys. Status Solidi RRL—Rapid Res. Lett. 2007, 1, R65–R67. [Google Scholar] [CrossRef]
  63. Ni, J.; Noh, K.; Frandsen, C.J.; Kong, S.D.; He, G.; Tang, T.; Jin, S. Preparation of near Micrometer-Sized TiO2 Nanotube Arrays by High Voltage Anodization. Mater. Sci. Eng. C 2013, 33, 259–264. [Google Scholar] [CrossRef] [PubMed]
  64. Albu, S.P.; Schmuki, P. TiO2 Nanotubes Grown in Different Organic Electrolytes: Two-Size Self-Organization, Single vs. Double-Walled Tubes, and Giant Diameters. Phys. Status Solidi RRL—Rapid Res. Lett. 2010, 4, 215–217. [Google Scholar] [CrossRef]
  65. Enachi, M.; Tiginyanu, I.; Sprincean, V.; Ursaki, V. Self-Organized Nucleation Layer for the Formation of Ordered Arrays of Double-Walled TiO2 Nanotubes with Temperature Controlled Inner Diameter. Phys. Status Solidi RRL—Rapid Res. Lett. 2010, 4, 100–102. [Google Scholar] [CrossRef]
  66. Ruff, T.; Hahn, R.; Schmuki, P. From Anodic TiO2 Nanotubes to Hexagonally Ordered TiO2 Nanocolumns. Appl. Surf. Sci. 2011, 257, 8177–8181. [Google Scholar] [CrossRef]
  67. Banerjee, S.; Misra, M.; Mohapatra, S.K.; Howard, C.; Mohapatra, S.K.; Kamilla, S.K. Formation of Chelating Agent Driven Anodized TiO2 Nanotubular Membrane and Its Photovoltaic Application. Nanotechnology 2010, 21, 145201. [Google Scholar] [CrossRef]
  68. So, S.; Lee, K.; Schmuki, P. Ultrafast Growth of Highly Ordered Anodic TiO2 Nanotubes in Lactic Acid Electrolytes. J. Am. Chem. Soc. 2012, 134, 11316–11318. [Google Scholar] [CrossRef]
  69. Ali, G.; Kim, H.J.; Kim, J.J.; Cho, S.O. Controlled Fabrication of Porous Double-Walled TiO2 Nanotubes Via Ultraviolet-Assisted Anodization. Nanoscale 2014, 6, 3632–3637. [Google Scholar] [CrossRef]
  70. Ocampo, R.A.; Echeverría, F.E. TiO2 Nanotubes Produced on Curved Titanium Surfaces Using Aqueous Electrolytes with Carboxymethyl Cellulose. Physica E 2021, 125, 114391. [Google Scholar] [CrossRef]
  71. Sander, M.S.; Côté, M.J.; Gu, W.; Kile, B.M.; Tripp, C.P. Template-Assisted Fabrication of Dense, Aligned Arrays of Titania Nanotubes with Well-Controlled Dimensions on Substrates. Adv. Mater. 2004, 16, 2052–2057. [Google Scholar] [CrossRef]
  72. Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K. Titania Nanotubes Prepared by Chemical Processing. Adv. Mater. 1999, 11, 1307–1311. [Google Scholar] [CrossRef]
  73. Suzuki, Y.; Yoshikawa, S. Synthesis and Thermal Analyses of TiO2-Derived Nanotubes Prepared by the Hydrothermal Method. J. Mater. Res. 2004, 19, 982–985. [Google Scholar] [CrossRef]
  74. Adachi, M.; Murata, Y.; Harada, M.; Yoshikawa, S. Formation of Titania Nanotubes with High Photo-Catalytic Activity. Chem. Lett. 2000, 29, 942–943. [Google Scholar] [CrossRef]
  75. Adachi, M.; Murata, Y.; Okada, I.; Yoshikawa, S. Formation of Titania Nanotubes and Applications for Dye-Sensitized Solar Cells. J. Electrochem. Soc. 2003, 150, G488. [Google Scholar] [CrossRef]
  76. Hoyer, P. Formation of a Titanium Dioxide Nanotube Array. Langmuir 1996, 12, 1411–1413. [Google Scholar] [CrossRef]
  77. Chakarvarti, S.K.; Vetter, J. Template Synthesisda Membrane Based Technology for Generation of Nano-/Micro Materials: A Review. Radiat. Meas. 1998, 29, 149–159. [Google Scholar] [CrossRef]
  78. Heo, S.Y.; Kim, D.J.; Jeon, H.; Jung, B.; Kang, Y.S.; Kim, J.H. Vertically Aligned Anatase TiO2 Nanotubes on Transparent Conducting Substrates Using Polycarbonate Membranes. RSC Adv. 2013, 3, 13681. [Google Scholar] [CrossRef]
  79. Zhang, M.; Bando, Y.; Wada, K. Sol-Gel Template Preparation of TiO2 Nanotubes and Nanorods. J. Mater. Sci. Lett. 2001, 20, 167–170. [Google Scholar] [CrossRef]
  80. Lakshmi, B.B.; Dorhout, P.K.; Martin, C.R. Sol−Gel Template Synthesis of Semiconductor Nanostructures. Chem. Mater. 1997, 9, 857–862. [Google Scholar] [CrossRef]
  81. Yu, Y.; Yin, X.; Kvit, A.; Wang, X. Evolution of Hollow TiO2 Nanostructures via the Kirkendall Effect Driven by Cation Exchange with Enhanced Photoelectrochemical Performance. Nano Lett. 2014, 14, 2528–2535. [Google Scholar] [CrossRef] [PubMed]
  82. Han, H.; Song, T.; Lee, E.-K.; Devadoss, A.; Jeon, Y.; Ha, J.; Chung, Y.-C.; Choi, Y.-M.; Jung, Y.-G.; Paik, U. Dominant Factors Governing the Rate Capability of a TiO2 Nanotube Anode for High Power Lithium Ion Batteries. ACS Nano 2012, 6, 8308–8315. [Google Scholar] [CrossRef] [PubMed]
  83. Li, D.; Xia, Y. Direct Fabrication of Composite and Ceramic Hollow Nanofibers by Electrospinning. Nano Lett. 2004, 4, 933–938. [Google Scholar] [CrossRef]
  84. Chang, C.; Chen, W.; Chen, Y.; Chen, Y.; Chen, Y.; Ding, F.; Fan, C.; Jin Fan, H.; Fan, Z.; Gong, C.; et al. Recent Progress on Two-Dimensional Materials. Acta Phys. Chim. Sin. 2021, 37, 2108017. [Google Scholar] [CrossRef]
  85. ten Elshof, J.E.; Yuan, H.; Gonzalez Rodriguez, P. Two-Dimensional Metal Oxide and Metal Hydroxide Nanosheets: Synthesis, Controlled Assembly and Applications in Energy Conversion and Storage. Adv. Energy Mater. 2016, 6, 1600355. [Google Scholar] [CrossRef]
  86. Xu, H.; Reunchan, P.; Ouyang, S.; Tong, H.; Umezawa, N.; Kako, T.; Ye, J. Anatase TiO2 Single Crystals Exposed with High-Reactive {111} Facets Toward Efficient H2 Evolution. Chem. Mater. 2013, 25, 405–411. [Google Scholar] [CrossRef]
  87. Liu, M.; Piao, L.; Zhao, L.; Ju, S.; Yan, Z.; He, T.; Zhou, C.; Wang, W. Anatase TiO2 Single Crystals with Exposed 001 and 110 Facets: Facile Synthesis and Enhanced Photocatalysis. Chem. Commun. 2010, 46, 1664–1666. [Google Scholar] [CrossRef] [PubMed]
  88. Sheng, X.; He, D.; Yang, J.; Zhu, K.; Feng, X. Oriented Assembled TiO2 Hierarchical Nanowire Arrays with Fast Electron Transport Properties. Nano Lett. 2014, 14, 1848–1852. [Google Scholar] [CrossRef] [PubMed]
  89. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 Single Crystals with A Large Percentage of Reactive Facets. Nature 2008, 453, 638–641. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, C.; Zhang, X.; Zhang, Y.; Jia, Y.; Yang, J.; Sun, P.; Liu, Y. Hydrothermal Growth of Layered Titanate Nanosheet Arrays on Titanium Foil and Their Topotactic Transformation to Heterostructured TiO2 Photocatalysts. J. Phys. Chem. C 2011, 115, 22276–22285. [Google Scholar] [CrossRef]
  91. Feng, S.; Yang, J.; Zhu, H.; Liu, M.; Zhang, J.; Wu, J.; Wan, J. Synthesis of Single Crystalline Anatase TiO2(001) Tetragonal Nanosheet-Array Films on Fluorine-Doped Tin Oxide Substrate: Rapid Communications of the American Ceramic Society. J. Am. Ceram. Soc. 2011, 94, 310–315. [Google Scholar] [CrossRef]
  92. Baikie, T.; Fang, Y.; Kadro, J.M.; Schreyer, M.; Wei, F.; Mhaisalkar, S.G.; Graetzel, M.; White, T.J. Synthesis and Crystal Chemistry of the Hybrid Perovskite (CH3NH3)PbI3 for Solid-State Sensitised Solar Cell Applications. J. Mater. Chem. A 2013, 1, 5628. [Google Scholar] [CrossRef]
  93. Jiang, C.; Wei, M.; Qi, Z.; Kudo, T.; Honma, I.; Zhou, H. Particle Size Dependence of the Lithium Storage Capability and High Rate Performance of Nanocrystalline Anatase TiO2 Electrode. J. Power Sources 2007, 166, 239–243. [Google Scholar] [CrossRef]
  94. Zhong, D.; Jiang, Q.; Huang, B.; Zhang, W.-H.; Li, C. Synthesis and Characterization of Anatase TiO2 Nanosheet Arrays on FTO Substrate. J. Energy Chem. 2015, 24, 626–631. [Google Scholar] [CrossRef]
  95. Gan, X.; Gao, X.; Qiu, J.; He, P.; Li, X.; Xiao, X. TiO2 Nanorod-Derived Synthesis of Upstanding Hexagonal Kassite Nanosheet Arrays: An Intermediate Route to Novel Nanoporous TiO2 Nanosheet Arrays. Cryst. Growth Des. 2012, 12, 289–296. [Google Scholar] [CrossRef]
  96. Wen, W.; Wu, J.; Jiang, Y.; Yu, S.; Bai, J.; Cao, M.; Cui, J. Anatase TiO2 Ultrathin Nanobelts Derived from Room-Temperature-Synthesized Titanates for Fast and Safe Lithium Storage. Sci. Rep. 2015, 5, 11804. [Google Scholar] [CrossRef]
  97. Luo, Y.; Kong, D.; Luo, J.; Chen, S.; Zhang, D.; Qiu, K.; Qi, X.; Zhang, H.; Li, C.M.; Yu, T. Hierarchical TiO2 Nanobelts@MnO2 Ultrathin Nanoflakes Core–Shell Array Electrode Materials for Supercapacitors. RSC Adv. 2013, 3, 14413. [Google Scholar] [CrossRef]
  98. Luo, Y.; Luo, J.; Zhou, W.; Qi, X.; Zhang, H.; Yu, D.Y.W.; Li, C.M.; Fan, H.J.; Yu, T. Controlled Synthesis of Hierarchical Graphene-Wrapped TiO2 @Co3O4 Coaxial Nanobelt Arrays for High-Performance Lithium Storage. J. Mater. Chem. A 2013, 1, 273–281. [Google Scholar] [CrossRef]
  99. Yang, Q.; Liu, Y.; Deng, C.; Yan, M.; Shi, W. MOF-Derived 3D Hierarchical Nanoarrays Consisting of NiCoZn-S Nanosheets Coupled with Granular NiCo2S4 Nanowires for High-Performance Hybrid Supercapacitors. J. Mater. Chem. A 2019, 7, 26131–26138. [Google Scholar] [CrossRef]
  100. Luo, W.; Zeng, W.; Quan, H.; Pan, M.; Wang, Y.; Chen, D. Carbon Dots Decorated NiCo Hydroxycarbonate Hierarchical Nanoarrays on Carbon Cloth with High Areal Capacitance as Pseudocapacitor Electrode. J. Alloys Compd. 2021, 868, 159048. [Google Scholar] [CrossRef]
  101. Wang, J.; Li, L.; Meng, L.; Wang, L.; Liu, Y.; Li, W.; Sun, W.; Li, G. Morphology Engineering of Nickel Molybdate Hydrate Nanoarray for Electrocatalytic Overall Water Splitting: From Nanorod to Nanosheet. RSC Adv. 2018, 8, 35131–35138. [Google Scholar] [CrossRef] [PubMed]
  102. Wei, G.; Yan, L.; Huang, H.; Yan, F.; Liang, X.; Xu, S.; Lan, Z.; Zhou, W.; Guo, J. The Hetero-Structured Nanoarray Construction of Co3O4 Nanowires Anchored on Nanoflakes as a High-Performance Electrode for Supercapacitors. Appl. Surf. Sci. 2021, 538, 147932. [Google Scholar] [CrossRef]
  103. Pan, Z.; Qiu, Y.; Yang, J.; Liu, M.; Zhou, L.; Xu, Y.; Sheng, L.; Zhao, X.; Zhang, Y. Synthesis of Three-Dimensional Hyperbranched TiO2 Nanowire Arrays with Significantly Enhanced Photoelectrochemical Hydrogen Production. J. Mater. Chem. A 2015, 3, 4004–4009. [Google Scholar] [CrossRef]
  104. Yang, X.; Zhuang, J.; Li, X.; Chen, D.; Ouyang, G.; Mao, Z.; Han, Y.; He, Z.; Liang, C.; Wu, M.; et al. Hierarchically Nanostructured Rutile Arrays: Acid Vapor Oxidation Growth and Tunable Morphologies. ACS Nano 2009, 3, 1212–1218. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, Z.; Hai, J.; Wu, J.-M.; Kobayashi, H.; Sun, T.; Zhang, Z.; Geng, C.; Zhang, Z.; Wen, W. Surface Diffusion Enhancement of Titania by Surface Fluorine Modification to Boost Energy Storage Capacities. Appl. Surf. Sci. 2023, 612, 155843. [Google Scholar] [CrossRef]
  106. Cho, I.S.; Chen, Z.; Forman, A.J.; Kim, D.R.; Rao, P.M.; Jaramillo, T.F.; Zheng, X. Branched TiO2 Nanorods for Photoelectrochemical Hydrogen Production. Nano Lett. 2011, 11, 4978–4984. [Google Scholar] [CrossRef] [PubMed]
  107. Chu, L.; Li, L.; Su, J.; Tu, F.; Liu, N.; Gao, Y. A General Method for Preparing Anatase TiO2 Treelike-Nanoarrays on Various Metal Wires for Fiber Dye-Sensitized Solar Cells. Sci. Rep. 2014, 4, 4420. [Google Scholar] [CrossRef] [PubMed]
  108. Wu, W.-Q.; Rao, H.-S.; Feng, H.-L.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. A Family of Vertically Aligned Nanowires with Smooth, Hierarchical and Hyperbranched Architectures for Efficient Energy Conversion. Nano Energy 2014, 9, 15–24. [Google Scholar] [CrossRef]
  109. Wu, W.-Q.; Feng, H.-L.; Rao, H.-S.; Xu, Y.-F.; Kuang, D.-B.; Su, C.-Y. Maximizing Omnidirectional Light Harvesting in Metal Oxide Hyperbranched Array Architectures. Nat. Commun. 2014, 5, 3968. [Google Scholar] [CrossRef]
  110. Geng, C.; Sun, T.; Wang, Z.; Wu, J.-M.; Gu, Y.-J.; Kobayashi, H.; Yang, P.; Hai, J.; Wen, W. Surface-Induced Desolvation of Hydronium Ion Enables Anatase TiO2 as an Efficient Anode for Proton Batteries. Nano Lett. 2021, 21, 7021–7029. [Google Scholar] [CrossRef]
  111. Wen, W.; Yao, J.-C.; Gu, Y.-J.; Sun, T.-L.; Tian, H.; Zhou, Q.-L.; Wu, J.-M. Balsam-Pear-like Rutile/Anatase Core/Shell Titania Nanorod Arrays for Photoelectrochemical Water Splitting. Nanotechnology 2017, 28, 465602. [Google Scholar] [CrossRef] [PubMed]
  112. Roh, D.K.; Chi, W.S.; Jeon, H.; Kim, S.J.; Kim, J.H. High Efficiency Solid-State Dye-Sensitized Solar Cells Assembled with Hierarchical Anatase Pine Tree-like TiO2 Nanotubes. Adv. Funct. Mater. 2014, 24, 379–386. [Google Scholar] [CrossRef]
  113. Wen, W.; Wu, J.-M.; Jiang, Y.-Z.; Lai, L.-L.; Song, J. Pseudocapacitance-Enhanced Li-Ion Microbatteries Derived by a TiN@TiO2 Nanowire Anode. Chem 2017, 2, 404–416. [Google Scholar] [CrossRef]
  114. Qin, J.; Cao, X.; Huang, H.; Fu, Z.; Wu, J.-M.; Zhang, P.; Ye, Z.; Wen, W. Modulation of Titania Nanoflower Arrays Transformed from Titanate Nanowire Arrays to Boost Photocatalytic Cr(ⅤⅠ) Detoxification. New J. Chem. 2022, 46, 21599–21604. [Google Scholar] [CrossRef]
  115. Wen, W.; Hai, J.; Yao, J.; Gu, Y.-J.; Kobayashi, H.; Tian, H.; Sun, T.; Chen, Q.; Yang, P.; Geng, C.; et al. Univariate Lattice Parameter Modulation of Single-Crystal-like Anatase TiO2 Hierarchical Nanowire Arrays to Improve Photoactivity. Chem. Mater. 2021, 33, 1489–1497. [Google Scholar] [CrossRef]
  116. Potapenko, D.V.; Li, Z.; Kysar, J.W.; Osgood, R.M. Nanoscale Strain Engineering on The Surface of A Bulk TiO2 Crystal. Nano Lett 2014, 14, 6185–6189. [Google Scholar] [CrossRef]
  117. Wen, W.; Yao, J.; Tan, H.; Wu, J.-M. Hollow TiN Nanotrees Derived from a Surface-Induced Kirkendall Effect and Their Application in High-Power Supercapacitors. J. Mater. Chem. A 2019, 7, 21378–21385. [Google Scholar] [CrossRef]
  118. Che, G.; Lakshmi, B.B.; Fisher, E.R.; Martin, C.R. Carbon Nanotubule Membranes for Electrochemical Energy Storage and Production. Nature 1998, 393, 346–349. [Google Scholar] [CrossRef]
  119. Nishizawa, M.; Mukai, K.; Kuwabata, S.; Martin, C.R.; Yoneyama, H. Template Synthesis of Polypyrrole-Coated Spinel LiMn2O4 Nanotubules and Their Properties as Cathode Active Materials for Lithium Batteries. J. Electrochem. Soc. 1997, 144, 1923–1927. [Google Scholar] [CrossRef]
  120. Sides, C.R.; Martin, C.R. Nanostructured Electrodes and the Low-Temperature Performance of Li-Ion Batteries. Adv. Mater. 2005, 17, 125–128. [Google Scholar] [CrossRef]
  121. Wei, T.-T.; Wang, F.-F.; Li, X.-Z.; Zhang, J.-H.; Zhu, Y.-R.; Yi, T.-F. Towards High-Performance Battery Systems By Regulating Morphology of TiO2 Materials. Sustain. Mater. Technol. 2021, 30, 2214–2232. [Google Scholar] [CrossRef]
  122. Li, Y.; Tan, B.; Wu, Y. Mesoporous Co3O4 Nanowire Arrays for Lithium Ion Batteries with High Capacity and Rate Capability. Nano Lett. 2008, 8, 265–270. [Google Scholar] [CrossRef] [PubMed]
  123. Chan, C.K.; Zhang, X.F.; Cui, Y. High Capacity Li Ion Battery Anodes Using Ge Nanowires. Nano Lett. 2008, 8, 307–309. [Google Scholar] [CrossRef] [PubMed]
  124. Chan, C.K.; Peng, H.; Liu, G.; McIlwrath, K.; Zhang, X.F.; Huggins, R.A.; Cui, Y. High-Performance Lithium Battery Anodes Using Silicon Nanowires. Nat. Nanotechnol. 2008, 3, 31–35. [Google Scholar] [CrossRef] [PubMed]
  125. Lu, J.; Chen, Z.; Pan, F.; Cui, Y.; Amine, K. High-Performance Anode Materials for Rechargeable Lithium-Ion Batteries. Electrochem. Energy Rev. 2018, 1, 35–53. [Google Scholar] [CrossRef]
  126. Zhu, G.-N.; Wang, Y.-G.; Xia, Y.-Y. Ti-Based Compounds as Anode Materials for Li-Ion Batteries. Energy Environ. Sci. 2012, 5, 6652. [Google Scholar] [CrossRef]
  127. Sun, C.H.; Yang, X.H.; Chen, J.S.; Li, Z.; Lou, X.W.; Li, C.; Smith, S.C.; Lu, G.Q.; Yang, H.G. Higher Charge/Discharge Rates of Lithium-Ions across Engineered TiO2 Surfaces Leads to Enhanced Battery Performance. Chem. Commun. 2010, 46, 6129. [Google Scholar] [CrossRef]
  128. Nuspl, G.; Yoshizawa, K.; Yamabe, T. Lithium Intercalation in TiO2 Modifications. J. Mater. Chem. 1997, 7, 2529–2536. [Google Scholar] [CrossRef]
  129. Kavan, L.; Rathouský, J.; Grätzel, M.; Shklover, V.; Zukal, A. Surfactant-Templated TiO2(Anatase): Characteristic Features of Lithium Insertion Electrochemistry in Organized Nanostructures. J. Phys. Chem. B 2000, 104, 12012–12020. [Google Scholar] [CrossRef]
  130. Tang, Y.; Yang, L.; Qiu, Z.; Huang, J. Template-Free Synthesis of Mesoporous Spinel Lithium Titanate Microspheres and Their Application in High-Rate Lithium Ion Batteries. J. Mater. Chem. 2009, 19, 5980. [Google Scholar] [CrossRef]
  131. Sudant, G.; Baudrin, E.; Larcher, D.; Tarascon, J.-M. Electrochemical Lithium Reactivity with Nanotextured Anatase-Type TiO2. J. Mater. Chem. 2005, 15, 1263–1269. [Google Scholar] [CrossRef]
  132. Kubiak, P.; Geserick, J.; Hüsing, N.; Wohlfahrt-Mehrens, M. Electrochemical Performance of Mesoporous TiO2 Anatase. J. Power Sources 2008, 175, 510–516. [Google Scholar] [CrossRef]
  133. Fattakhova-Rohlfing, D.; Wark, M.; Brezesinski, T.; Smarsly, B.M.; Rathouský, J. Highly Organized Mesoporous TiO2 Films with Controlled Crystallinity: A Li-Insertion Study. Adv. Funct. Mater. 2007, 17, 123–132. [Google Scholar] [CrossRef]
  134. Gligor, F.; Deleeuw, S. Lithium Diffusion in Rutile Structured Titania. Solid State Ion. 2006, 177, 2741–2746. [Google Scholar] [CrossRef]
  135. Johnson, O.W. One-Dimensional Diffusion of Li in Rutile. Phys. Rev. 1964, 136, A284–A290. [Google Scholar] [CrossRef]
  136. Baudrin, E.; Cassaignon, S.; Koelsch, M.; Jolivet, J.; Dupont, L.; Tarascon, J. Structural Evolution during the Reaction of Li with Nano-Sized Rutile Type TiO2 at Room Temperature. Electrochem. Commun. 2007, 9, 337–342. [Google Scholar] [CrossRef]
  137. Jiang, C.; Honma, I.; Kudo, T.; Zhou, H. Nanocrystalline Rutile TiO2 Electrode for High-Capacity and High-Rate Lithium Storage. Electrochem. Solid-State Lett. 2007, 10, A127. [Google Scholar] [CrossRef]
  138. Reddy, M.A.; Kishore, M.S.; Pralong, V.; Caignaert, V.; Varadaraju, U.V.; Raveau, B. Room Temperature Synthesis and Li Insertion into Nanocrystalline Rutile TiO2. Electrochem. Commun. 2006, 8, 1299–1303. [Google Scholar] [CrossRef]
  139. Yu, Y.; Sun, D.; Wang, H.; Wang, H. Electrochemical Properties of Rutile TiO2 Nanorod Array in Lithium Hydroxide Solution. Nanoscale Res. Lett. 2016, 11, 448. [Google Scholar] [CrossRef]
  140. Guan, D.; Cai, C.; Wang, Y. Amorphous and Crystalline TiO2 Nanotube Arrays for Enhanced Li-Ion Intercalation Properties. J. Nanosci. Nanotechnol. 2011, 11, 3641–3650. [Google Scholar] [CrossRef]
  141. Cava, B.R.J.; Murphy, D.W.; Zahurak, S. The Crystal Structures of the Lithium-Inserted Metal Oxides Li0.5TiO2 Anatase, LiTi2O4 Spinel, and Li2Ti2O4. J. Solid State Chem. 1984, 53, 64–75. [Google Scholar] [CrossRef]
  142. Yiping, T.; Xiaoxu, T.; Guangya, H.; Huazhen, C.; Guoqu, Z. Synthesis of Dense Nanocavities inside TiO2 Nanowire Array and Its Electrochemical Properties as a Three-Dimensional Anode Material for Li-Ion Batteries. Electrochim. Acta 2012, 78, 154–159. [Google Scholar] [CrossRef]
  143. Liu, S.; Jia, H.; Han, L.; Wang, J.; Gao, P.; Xu, D.; Yang, J.; Che, S. Nanosheet-Constructed Porous TiO2-B for Advanced Lithium Ion Batteries. Adv. Mater. 2012, 24, 3201–3204. [Google Scholar] [CrossRef] [PubMed]
  144. Pham, T.N.; Bui, V.K.H.; Lee, Y.C. Recent Advances in Hierarchical Anode Designs of TiO2-Bnanostructures for Lithium-ion Batteries. Int. J. Energy Res. 2021, 45, 17532–17562. [Google Scholar] [CrossRef]
  145. Liu, G.; Wu, H.-H.; Meng, Q.; Zhang, T.; Sun, D.; Jin, X.; Guo, D.; Wu, N.; Liu, X.; Kim, J.-K. Role of the anatase/TiO2(B) heterointerface for ultrastable high-rate lithium and sodium energy storage performance. Nanoscale Horiz. 2020, 5, 150–162. [Google Scholar] [CrossRef]
  146. Shen, W.; Huo, X.; Zhang, M.; Guo, M. Synthesis of Oriented Core/Shell Hexagonal Tungsten Oxide/Amorphous Titanium Dioxide Nanorod Arrays and its Electrochromic-Pseudocapacitive Properties. Appl. Surf. Sci. 2020, 515, 146034. [Google Scholar] [CrossRef]
  147. Huang, M.; Xi, B.; Shi, N.; Wei, R.; Li, H.; Feng, J.; Xiong, S. Systematic Study of Alkali Cations Intercalated Titanium Dioxide Effect on Sodium and Lithium Storage. Small 2020, 16, e2001391. [Google Scholar] [CrossRef]
  148. Liu, B.; Deng, D.; Lee, J.Y.; Aydil, E.S. Oriented Single-Crystalline TiO2 Nanowires on Titanium Foil for Lithium Ion Batteries. J. Mater. Res. 2010, 25, 1588–1594. [Google Scholar] [CrossRef]
  149. Tang, Y.; Hong, L.; Wu, Q.; Li, J.; Hou, G.; Cao, H.; Wu, L.; Zheng, G. TiO2(B) Nanowire Arrays on Ti Foil Substrate as Three-Dimensional Anode for Lithium-Ion Batteries. Electrochim. Acta 2016, 195, 27–33. [Google Scholar] [CrossRef]
  150. Fu, S.; Chen, J.; Wang, X.; He, Q.; Tong, S.; Wu, M. Free-Standing Crystalline@Amorphous Core-Shell Nanoarrays for Efficient Energy Storage. Small 2020, 16, e2000040. [Google Scholar] [CrossRef]
  151. Wu, Q.; Xu, J.; Yang, X.; Lu, F.; He, S.; Yang, J.; Fan, H.J.; Wu, M. Ultrathin Anatase TiO2 Nanosheets Embedded with TiO2-B Nanodomains for Lithium-Ion Storage: Capacity Enhancement by Phase Boundaries. Adv. Energy Mater. 2015, 5, 1401756. [Google Scholar] [CrossRef]
  152. Wang, R.; Xue, X.; Lu, W.; Liu, H.; Lai, C.; Xi, K.; Che, Y.; Liu, J.; Guo, S.; Yang, D. Tuning and Understanding the Phase Interface of TiO2 Nanoparticles for More Efficient Lithium Ion Storage. Nanoscale 2015, 7, 12833–12838. [Google Scholar] [CrossRef] [PubMed]
  153. Zhukovskii, Y.F.; Balaya, P.; Kotomin, E.A.; Maier, J. Evidence for Interfacial-Storage Anomaly in Nanocomposites for Lithium Batteries from First-Principles Simulations. Phys. Rev. Lett. 2006, 96, 058302. [Google Scholar] [CrossRef] [PubMed]
  154. Wen, W.; Wu, J.-M.; Jiang, Y.-Z.; Bai, J.-Q.; Lai, L.-L. Titanium Dioxide Nanotrees for High-Capacity Lithium-Ion Microbatteries. J. Mater. Chem. A 2016, 4, 10593–10600. [Google Scholar] [CrossRef]
  155. Cheah, S.K.; Perre, E.; Rooth, M.; Fondell, M.; Hårsta, A.; Nyholm, L.; Boman, M.; Gustafsson, T.r.; Lu, J.; Simon, P. Self-Supported Three-Dimensional Nanoelectrodes for Microbattery Applications. Nano Lett. 2009, 9, 3230–3233. [Google Scholar] [CrossRef]
  156. Wang, W.; Tian, M.; Abdulagatov, A.; George, S.M.; Lee, Y.C.; Yang, R. Three-Dimensional Ni/TiO2 Nanowire Network for High Areal Capacity Lithium Ion Microbattery Applications. Nano Lett 2012, 12, 655–660. [Google Scholar] [CrossRef]
  157. Tang, Y.; Zhang, Y.; Xie, G.; Zheng, Y.; Yu, J.; Gao, L.; Liu, B. Construction of Rutile-TiO2 Nanoarray Homojuction for Non-Contact Sensing of TATP Under Natural Light. Coat. 2020, 10, 409. [Google Scholar] [CrossRef]
  158. Zhang, Q.; Lee, I.; Joo, J.B.; Zaera, F.; Yin, Y. Core–Shell Nanostructured Catalysts. Acc. Chem. Res. 2013, 46, 1816–1824. [Google Scholar] [CrossRef]
  159. Liu, R.; Shen, Z.; Wan, Z.; Zhu, L.; Chen, J.; Dong, C.; Chen, W.; Cao, W.; Chen, B.; Yuan, X.; et al. Nanoarray Heterojunction and Its Efficient Solar Cells without Negative Impact of Photogenerated Electric Field. Commun. Phys. 2021, 4, 177. [Google Scholar] [CrossRef]
  160. Ying, J.Y.; Sun, T. Research Needs Assessment on Nanostructured Catalysts. J. Electroceram. 1997, 1, 219–238. [Google Scholar] [CrossRef]
  161. Cai, H.; Yan, S.; Dong, X.; Cao, F.; Wang, G. Significantly Enhanced Energy Storage Performance By Constructing TiO2 Nanowire Arrays In Pbzro3-Based Antiferroelectric Films. Ceram. Int. 2020, 46, 6436–6442. [Google Scholar] [CrossRef]
  162. Lu, X.; Li, M.; Hoang, S.; Suib, S.L.; Gao, P.-X. Solvent Effects on the Heterogeneous Growth of TiO2 Nanostructure Arrays By Solvothermal Synthesis. Catal. Today 2021, 360, 275–283. [Google Scholar] [CrossRef]
  163. Zhang, L.; Yang, X.; Jiang, B.; Sun, Y.; Gong, Z.; Zhang, N.; Hou, S.; Li, J.; Yang, N. Superhydrophilic and Underwater Superoleophobic Ti Foam with Robust Nanoarray Structures of TiO2 For Effective Oil-In-Water Emulsion Separation. Sep. Purif. Technol. 2020, 252, 117437. [Google Scholar] [CrossRef]
  164. Hu, C.; Chen, L.; Hu, Y.; Chen, A.; Chen, L.; Jiang, H.; Li, C. Light-Motivated SnO2/TiO2 Heterojunctions Enabling the Breakthrough in Energy Density for Lithium-ion Batteries. Adv Mater 2021, 33, e2103558. [Google Scholar] [CrossRef]
  165. Jamil, M.; Wei, S.; Taylor, M.P.; Chen, J.J.J.; Kennedy, J.V. Hybrid Anode Materials for Rechargeable Batteries-A Review of Sn/TiO2 Based Nanocomposites. Energy Rep. 2021, 7, 2836–2848. [Google Scholar] [CrossRef]
  166. Trang, T.N.Q.; Doanh, T.T.; Vinh, L.Q.; Thu, V.T.H. A Hybrid Ag/TiO2 Nanoarray-Based in Situ Charge Transfer Toward Multi-Functional Active-Platform. Ceram. Int. 2021, 47, 27524–27534. [Google Scholar] [CrossRef]
  167. Yu, J.; Zeng, M.; Zhou, J.; Chen, H.; Cong, G.; Liu, H.; Ji, M.; Zhu, C.; Xu, J. A One-Pot Synthesis of Nitrogen Doped Porous Mxene/TiO2 Heterogeneous Film for High-Performance Flexible Energy Storage. Chem. Eng. J. 2021, 426, 130765. [Google Scholar] [CrossRef]
  168. Qiu, S.Y.; Wang, C.; Jiang, Z.X.; Zhang, L.S.; Gu, L.L.; Wang, K.X.; Gao, J.; Zhu, X.D.; Wu, G. Rational Design of Mxene@TiO2 Nanoarray Enabling Dual Lithium Polysulfide Chemisorption Towards High-Performance Lithium-Sulfur Batteries. Nanoscale 2020, 12, 16678–16684. [Google Scholar] [CrossRef]
  169. Kyeremateng, N.A.; Vacandio, F.; Sougrati, M.T.; Martinez, H.; Jumas, J.C.; Knauth, P.; Djenizian, T. Effect of Sn-Doping on the Electrochemical Behaviour of TiO2 Nanotubes As Potential Negative Electrode Materials for 3D Li-ion Micro Batteries. J. Power Sources 2013, 224, 269–277. [Google Scholar] [CrossRef]
  170. Luo, J.; Xia, X.; Luo, Y.; Guan, C.; Liu, J.; Qi, X.; Ng, C.F.; Yu, T.; Zhang, H.; Fan, H.J. Rationally Designed Hierarchical TiO2 @Fe2O3 Hollow Nanostructures for Improved Lithium Ion Storage. Adv. Energy Mater. 2013, 3, 737–743. [Google Scholar] [CrossRef]
  171. Eustache, E.; Tilmant, P.; Morgenroth, L.; Roussel, P.; Patriarche, G.; Troadec, D.; Rolland, N.; Brousse, T.; Lethien, C. Silicon-Microtube Scaffold Decorated with Anatase TiO2 As A Negative Electrode for A 3D Litium-Ion Microbattery. Adv. Energy Mater. 2014, 4, 1301612. [Google Scholar] [CrossRef]
  172. Liu, D.; Xiao, P.; Zhang, Y.; Garcia, B.B.; Zhang, Q.; Guo, Q.; Champion, R.; Cao, G. TiO2 Nanotube Arrays Annealed in N2 for Efficient Lithium-Ion Intercalation. J. Phys. Chem. C 2008, 112, 11175–11180. [Google Scholar] [CrossRef]
  173. Liu, D.; Zhang, Y.; Xiao, P.; Garcia, B.B.; Zhang, Q.; Zhou, X.; Jeong, Y.-H.; Cao, G. TiO2 Nanotube Arrays Annealed in CO Exhibiting High Performance for Lithium Ion Intercalation. Electrochim. Acta 2009, 54, 6816–6820. [Google Scholar] [CrossRef]
  174. Li, Y.; Lv, X.; Li, J. High Performance Binderless TiO2 Nanowire Arrays Electrode for Lithium-Ion Battery. Appl. Phys. Lett. 2009, 95, 113102. [Google Scholar] [CrossRef]
  175. Ortiz, G.F.; Hanzu, I.; Knauth, P.; Lavela, P.; Tirado, J.L.; Djenizian, T. TiO2 Nanotubes Manufactured By Anodization of Ti Thin Films for On-Chip Li-ion 2D Microbatteries. Electrochim. Acta 2009, 54, 4262–4268. [Google Scholar] [CrossRef]
  176. Ortiz, G.F.; Hanzu, I.; Djenizian, T.; Lavela, P.; Tirado, J.L.; Knauth, P. Alternative Li-Ion Battery Electrode Based on Self-Organized Titania Nanotubes. Chem. Mater. 2009, 21, 63–67. [Google Scholar] [CrossRef]
  177. Xiong, H.; Yildirim, H.; Shevchenko, E.V.; Prakapenka, V.B.; Koo, B.; Slater, M.D.; Balasubramanian, M.; Sankaranarayanan, S.K.R.S.; Greeley, J.P.; Tepavcevic, S.; et al. Self-Improving Anode for Lithium-Ion Batteries Based on Amorphous to Cubic Phase Transition in TiO2 Nanotubes. J. Phys. Chem. C 2012, 116, 3181–3187. [Google Scholar] [CrossRef]
  178. Liu, J.; Chen, J.S.; Wei, X.; Lou, X.W.; Liu, X.-W. Sandwich-Like, Stacked Ultrathin Titanate Nanosheets for Ultrafast Lithium Storage. Adv. Mater. 2011, 23, 998–1002. [Google Scholar] [CrossRef]
  179. Liao, J.-Y.; Chabot, V.; Gu, M.; Wang, C.; Xiao, X.; Chen, Z. Dual Phase Li4Ti5O12–TiO2 Nanowire Arrays as Integrated Anodes for High-Rate Lithium-Ion Batteries. Nano Energy 2014, 9, 383–391. [Google Scholar] [CrossRef]
  180. Luo, Y.; Luo, J.; Jiang, J.; Zhou, W.; Yang, H.; Qi, X.; Zhang, H.; Fan, H.J.; Yu, D.Y.W.; Li, C.M.; et al. Seed-Assisted Synthesis of Highly Ordered TiO2@α-Fe2O3 Core/Shell Arrays on Carbon Textiles for Lithium-Ion Battery Applications. Energy Environ. Sci. 2012, 5, 6559. [Google Scholar] [CrossRef]
  181. Ortiz, G.F.; Hanzu, I.; Lavela, P.; Knauth, P.; Tirado, J.L.; Djenizian, T. Nanoarchitectured TiO2/SnO: A Future Negative Electrode for High Power Density Li-Ion Microbatteries? Chem. Mater. 2010, 22, 1926–1932. [Google Scholar] [CrossRef]
  182. Wang, C.; Wu, L.; Wang, H.; Zuo, W.; Li, Y.; Liu, J. Fabrication and Shell Optimization of Synergistic TiO2-MoO3 Core-Shell Nanowire Array Anode for High Energy and Power Density Lithium-Ion Batteries. Adv. Funct. Mater. 2015, 25, 3524–3533. [Google Scholar] [CrossRef]
  183. Du, G.; Guo, Z.; Zhang, P.; Li, Y.; Chen, M.; Wexler, D.; Liu, H. SnO2 Nanocrystals on Self-Organized TiO2 Nanotube Array as Three-Dimensional Electrode for Lithium Ion Microbatteries. J. Mater. Chem. 2010, 20, 5689. [Google Scholar] [CrossRef]
  184. Wu, X.; Zhang, S.; Wang, L.; Du, Z.; Fang, H.; Ling, Y.; Huang, Z. Coaxial SnO2@TiO2 Nanotube Hybrids: From Robust Assembly Strategies to Potential Application in Li+ Storage. J. Mater. Chem. 2012, 22, 11151. [Google Scholar] [CrossRef]
  185. Kyeremateng, N.A.; Lebouin, C.; Knauth, P.; Djenizian, T. The Electrochemical Behaviour of TiO2 Nanotubes with Co3O4 or NiO Submicron Particles: Composite Anode Materials for Li-Ion Micro Batteries. Electrochim. Acta 2013, 88, 814–820. [Google Scholar] [CrossRef]
  186. Han, H.; Song, T.; Bae, J.-Y.; Nazar, L.F.; Kim, H.; Paik, U. Nitridated TiO2 Hollow Nanofibers as an Anode Material for High Power Lithium Ion Batteries. Energy Environ. Sci. 2011, 4, 4532. [Google Scholar] [CrossRef]
  187. Zeng, L.; Zheng, C.; Xia, L.; Wang, Y.; Wei, M. Ordered Mesoporous TiO2–C Nanocomposite as an Anode Material for Long-Term Performance Lithium-Ion Batteries. J. Mater. Chem. A 2013, 1, 4293. [Google Scholar] [CrossRef]
  188. Wang, B.; Xin, H.; Li, X.; Cheng, J.; Yang, G.; Nie, F. Mesoporous CNT@ TiO2-C Nanocable with Extremely Durable High Rate Capability for Lithium-Ion Battery Anodes. Sci. Rep. 2014, 4, 3729. [Google Scholar] [CrossRef]
  189. Zhu, C.; Xia, X.; Liu, J.; Fan, Z.; Chao, D.; Zhang, H.; Fan, H.J. TiO2 Nanotube @ SnO2 Nanoflake Core–Branch Arrays for Lithium-Ion Battery Anode. Nano Energy 2014, 4, 105–112. [Google Scholar] [CrossRef]
  190. Hou, X.; Wang, X.; Liu, B.; Wang, Q.; Wang, Z.; Chen, D.; Shen, G. SnO2@TiO2 Heterojunction Nanostructures for Lithium-Ion Batteries and Self-Powered UV Photodetectors with Improved Performances. ChemElectroChem 2014, 1, 108–115. [Google Scholar] [CrossRef]
  191. Jeun, J.-H.; Park, K.-Y.; Kim, D.-H.; Kim, W.-S.; Kim, H.-C.; Lee, B.-S.; Kim, H.; Yu, W.-R.; Kang, K.; Hong, S.-H. SnO2@TiO2 Double-Shell Nanotubes for a Lithium Ion Battery Anode with Excellent High Rate Cyclability. Nanoscale 2013, 5, 8480. [Google Scholar] [CrossRef] [PubMed]
  192. Wang, L.; Zhang, S.; Wu, X. Synthesis and Lithium Storage Properties of NiO@TiO2 Nanotube Heterojunction Arrays. Chem. Lett. 2011, 40, 1428–1430. [Google Scholar] [CrossRef]
  193. Zhu, K.; Wang, Q.; Kim, J.-H.; Pesaran, A.A.; Frank, A.J. Pseudocapacitive Lithium-Ion Storage in Oriented Anatase TiO2 Nanotube Arrays. J. Phys. Chem. C 2012, 116, 11895–11899. [Google Scholar] [CrossRef]
  194. Dunn, B.; Kamath, H.; Tarascon, J.M. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef]
  195. Kim, S.-W.; Seo, D.-H.; Ma, X.; Ceder, G.; Kang, K. Electrode Materials for Rechargeable Sodium-Ion Batteries: Potential Alternatives to Current Lithium-Ion Batteries. Adv. Energy Mater. 2012, 2, 710–721. [Google Scholar] [CrossRef]
  196. Slater, M.D.; Kim, D.; Lee, E.; Johnson, C.S. Sodium-Ion Batteries. Adv. Funct. Mater. 2013, 23, 947–958. [Google Scholar] [CrossRef]
  197. Cao, Y.; Xiao, L.; Wang, W.; Choi, D.; Nie, Z.; Yu, J.; Saraf, L.V.; Yang, Z.; Liu, J. Reversible Sodium Ion Insertion in Single Crystalline Manganese Oxide Nanowires with Long Cycle Life. Adv. Mater. 2011, 23, 3155–3160. [Google Scholar] [CrossRef]
  198. Longoni, G.; Pena Cabrera, R.L.; Polizzi, S.; D’Arienzo, M.; Mari, C.M.; Cui, Y.; Ruffo, R. Shape-Controlled TiO2 Nanocrystals for Na-Ion Battery Electrodes: The Role of Different Exposed Crystal Facets on the Electrochemical Properties. Nano Lett. 2017, 17, 992–1000. [Google Scholar] [CrossRef]
  199. Bella, F.; Muñoz-García, A.B.; Meligrana, G.; Lamberti, A.; Destro, M.; Pavone, M.; Gerbaldi, C. Unveiling the Controversial Mechanism of Reversible Na Storage in TiO2 Nanotube Arrays: Amorphous versus Anatase TiO2. Nano Res. 2017, 10, 2891–2903. [Google Scholar] [CrossRef]
  200. Bella, F.; Muñoz-García, A.B.; Colò, F.; Meligrana, G.; Lamberti, A.; Destro, M.; Pavone, M.; Gerbaldi, C. Combined Structural, Chemometric, and Electrochemical Investigation of Vertically Aligned TiO2 Nanotubes for Na-Ion Batteries. ACS Omega 2018, 3, 8440–8450. [Google Scholar] [CrossRef]
  201. Wei, W.; Valvo, M.; Edström, K.; Nyholm, L. A Facile Vapor-Phase Hydrothermal Method for Direct Growth of Nanotube Anodes for Sodium-Ion Batteries. ChemElectroChem 2018, 5, 674–684. [Google Scholar] [CrossRef]
  202. Yang, J.; Chen, Z.; Wang, H.; Liang, F.; Chen, R.; Wu, R. Highly Ordered Three-Dimensional TiO2@C Nanotube Arrays as Freestanding Electrode for Sodium-Ion Battery. Mater. Lett. 2017, 207, 149–152. [Google Scholar] [CrossRef]
  203. Xu, Y.; Zhou, M.; Wen, L.; Wang, C.; Zhao, H.; Mi, Y.; Liang, L.; Fu, Q.; Wu, M.; Lei, Y. Highly Ordered Three-Dimensional Ni-TiO2 Nanoarrays as Sodium Ion Battery Anodes. Chem. Mater. 2015, 27, 4274–4280. [Google Scholar] [CrossRef]
  204. Liao, J.-Y.; Dew Luna, B.; Manthiram, A. TiO2-B Nanowire Arrays Coated with Layered MoS2 Nanosheets for Lithium and Sodium Storage. J. Mater. Chem. A 2016, 4, 801–806. [Google Scholar] [CrossRef]
  205. Ni, J.; Fu, S.; Yuan, Y.; Ma, L.; Jiang, Y.; Li, L.; Lu, J. Boosting Sodium Storage in TiO2 Nanotube Arrays through Surface Phosphorylation. Adv. Mater. 2018, 30, 1704337. [Google Scholar] [CrossRef] [PubMed]
  206. Ni, J.; Fu, S.; Wu, C.; Maier, J.; Yu, Y.; Li, L. Self-Supported Nanotube Arrays of Sulfur-Doped TiO2 Enabling Ultrastable and Robust Sodium Storage. Adv. Mater. 2016, 28, 2259–2265. [Google Scholar] [CrossRef]
  207. Xiong, H.; Slater, M.D.; Balasubramanian, M.; Johnson, C.S.; Rajh, T. Amorphous TiO2 Nanotube Anode for Rechargeable Sodium Ion Batteries. J. Phys. Chem. Lett. 2011, 2, 2560–2565. [Google Scholar] [CrossRef]
  208. Raj, C.C.; Prasanth, R. Review—Advent of TiO2 Nanotubes as Supercapacitor Electrode. J. Electrochem. Soc. 2018, 165, E345–E358. [Google Scholar] [CrossRef]
  209. He, Z.-K.; Lu, Y.; Zhao, C.; Zhao, J.; Gao, Z.; Song, Y.-Y. Surface-Charge Regulated TiO2 Nanotube Arrays As Scaffold for Constructing Binder-Free High-Performance Supercapacitor. Appl. Surf. Sci. 2021, 567, 150832. [Google Scholar] [CrossRef]
  210. Kim, M.S.; Lee, T.-W.; Park, J.H. Controlled TiO2 Nanotube Arrays as an Active Material for High Power Energy-Storage Devices. J. Electrochem. Soc. 2011, 2, 2560–2565. [Google Scholar] [CrossRef]
  211. Salari, M.; Aboutalebi, S.H.; Konstantinov, K.; Liu, H.K. A Highly Ordered Titania Nanotube Array as a Supercapacitor Electrode. Phys. Chem. Chem. Phys. 2011, 13, 5038. [Google Scholar] [CrossRef] [PubMed]
  212. Raj, C.C.; Sundheep, R.; Prasanth, R. Enhancement of Electrochemical Capacitance by Tailoring the Geometry of TiO2 Nanotube Electrodes. Electrochim. Acta 2015, 176, 1214–1220. [Google Scholar] [CrossRef]
  213. Heng, I.; Low, F.W.; Lai, C.W.; Juan, J.C.; Tiong, S.K. Hybrid Graphene Titanium Nanocomposites and Their Applications in Energy Storage Devices: A Review. J. Electron. Mater. 2020, 49, 1777–1786. [Google Scholar] [CrossRef]
  214. Chowdhury, R.; Saini, S.K.; Roy, J. Bio-Fabrication of TiO2 Nanomaterials and Their Applications in Electronics Devices. J. Electron. Mater. 2021, 50, 6087–6101. [Google Scholar] [CrossRef]
  215. Ates, M.; Kuzgun, O.; Candan, I. Supercapacitor Performances of Titanium–Polymeric Nanocomposites: A Review Study. Iran. Polym. J. 2022, 31, 31–57. [Google Scholar] [CrossRef]
  216. Ramadoss, A.; Kim, G.-S.; Kim, S.J. Fabrication of Reduced Graphene Oxide/ TiO2 Nanorod/Reduced Graphene Oxide Hybrid Nanostructures as Electrode Materials for Supercapacitor Applications. CrystEngComm 2013, 15, 10222. [Google Scholar] [CrossRef]
  217. Ke, Q.; Liao, Y.; Yao, S.; Song, L.; Xiong, X. A Three-Dimensional TiO2/Graphene Porous Composite with Nano-Carbon Deposition for Supercapacitor. J. Mater. Sci. 2016, 51, 2008–2016. [Google Scholar] [CrossRef]
  218. Guo, Q.; Li, J.; Zhang, B.; Nie, G.; Wang, D. High-Performance Asymmetric Electrochromic-Supercapacitor Device Based on Poly(Indole-6-Carboxylicacid)/TiO2 Nanocomposites. ACS Appl. Mater. Interfaces 2019, 11, 6491–6501. [Google Scholar] [CrossRef]
  219. Lu, X.; Wang, G.; Zhai, T.; Yu, M.; Gan, J.; Tong, Y.; Li, Y. Hydrogenated TiO2 Nanotube Arrays for Supercapacitors. Nano Lett. 2012, 12, 1690–1696. [Google Scholar] [CrossRef]
  220. Wu, H.; Xu, C.; Xu, J.; Lu, L.; Fan, Z.; Chen, X.; Song, Y.; Li, D. Enhanced Supercapacitance in Anodic TiO2 Nanotube Films by Hydrogen Plasma Treatment. Nanotechnology 2013, 24, 455401. [Google Scholar] [CrossRef]
  221. Zhou, H.; Zhang, Y. Electrochemically Self-Doped TiO2 Nanotube Arrays for Supercapacitors. J. Phys. Chem. C 2014, 118, 5626–5636. [Google Scholar] [CrossRef]
  222. Zhou, H.; Zhang, Y. Enhancing the Capacitance of TiO2 Nanotube Arrays by a Facile Cathodic Reduction Process. J. Power Sources 2013, 239, 128–131. [Google Scholar] [CrossRef]
  223. Wu, H.; Li, D.; Zhu, X.; Yang, C.; Liu, D.; Chen, X.; Song, Y.; Lu, L. High-Performance and Renewable Supercapacitors Based on TiO2 Nanotube Array Electrodes Treated by an Electrochemical Doping Approach. Electrochim. Acta 2014, 116, 129–136. [Google Scholar] [CrossRef]
  224. Kim, C.; Kim, S.; Lee, J.; Kim, J.; Yoon, J. Capacitive and Oxidant Generating Properties of Black-Colored TiO2 Nanotube Array Fabricated by Electrochemical Self-Doping. ACS Appl. Mater. Interfaces 2015, 7, 7486–7491. [Google Scholar] [CrossRef]
  225. Pazhamalai, P.; Krishnamoorthy, K.; Mariappan, V.K.; Kim, S.-J. Blue TiO2 Nanosheets as a High-Performance Electrode Material for Supercapacitors. J. Colloid Interface Sci. 2019, 536, 62–70. [Google Scholar] [CrossRef]
  226. Kim, J.-H.; Zhu, K.; Yan, Y.; Perkins, C.L.; Frank, A.J. Microstructure and Pseudocapacitive Properties of Electrodes Constructed of Oriented NiO-TiO2 Nanotube Arrays. Nano Lett. 2010, 10, 4099–4104. [Google Scholar] [CrossRef]
  227. Patil, J.V.; Mali, S.S.; Shaikh, J.S.; Bhat, T.S.; Hong, C.K.; Kim, J.H.; Patil, P.S. Hydrothermally Grown 3D Hierarchical TiO2 Based on Electrochemically Anodized 1D TiO2 Nanostructure for Supercapacitor. Appl. Phys. A 2018, 124, 592. [Google Scholar] [CrossRef]
  228. Qorbani, M.; Khajehdehi, O.; Sabbah, A.; Naseri, N. Ti-Rich TiO2 Tubular Nanolettuces by Electrochemical Anodization for All-Solid-State High-Rate Supercapacitor Devices. ChemSusChem 2019, 12, 4064–4073. [Google Scholar] [CrossRef]
  229. Li, Z.; Ding, Y.; Kang, W.; Li, C.; Lin, D.; Wang, X.; Chen, Z.; Wu, M.; Pan, D. Reduction Mechanism and Capacitive Properties of Highly Electrochemically Reduced TiO2 Nanotube Arrays. Electrochim. Acta 2015, 161, 40–47. [Google Scholar] [CrossRef]
  230. Zhou, H.; Zhang, Y. Enhanced Electrochemical Performance of Manganese Dioxide Spheres Deposited on a Titanium Dioxide Nanotube Arrays Substrate. J. Power Sources 2014, 272, 866–879. [Google Scholar] [CrossRef]
  231. Song, Y.-Y.; Li, Y.-H.; Guo, J.; Gao, Z.-D.; Li, Y. Facile Method to Synthesize a Carbon Layer Embedded into Titanium Dioxide Nanotubes with Metal Oxide Decoration for Electrochemical Applications. J. Mater. Chem. A 2015, 3, 23754–23759. [Google Scholar] [CrossRef]
  232. Xie, Y.; Huang, C.; Zhou, L.; Liu, Y.; Huang, H. Supercapacitor Application of Nickel Oxide–Titania Nanocomposites. Compos. Sci. Technol. 2009, 69, 2108–2114. [Google Scholar] [CrossRef]
  233. Ray, R.S.; Sarma, B.; Misra, M. Random Shaped ZnO Supported on a Porous Substrate as Supercapacitor. Mater. Lett. 2015, 155, 102–105. [Google Scholar] [CrossRef]
  234. Sarkar, A.; Singh, A.K.; Sarkar, D.; Khan, G.G.; Mandal, K. Three-Dimensional Nanoarchitecture of BiFeO3 Anchored TiO2 Nanotube Arrays for Electrochemical Energy Storage and Solar Energy Conversion. ACS Sustain. Chem. Eng. 2015, 3, 2254–2263. [Google Scholar] [CrossRef]
  235. Yang, Y.; Kim, D.; Yang, M.; Schmuki, P. Vertically Aligned Mixed V2O5–TiO2 Nanotube Arrays for Supercapacitor Applications. Chem. Commun. 2011, 47, 7746. [Google Scholar] [CrossRef] [PubMed]
  236. Gao, Z.; Cui, Z.; Zhu, S.; Liang, Y.; Li, Z.; Yang, X. Design and Synthesis of MWNTs-TiO2 Nanotube Hybrid Electrode and Its Supercapacitance Performance. J. Power Sources 2015, 283, 397–407. [Google Scholar] [CrossRef]
  237. Siuzdak, K.; Bogdanowicz, R.; Sawczak, M.; Sobaszek, M. Enhanced Capacitance of Composite TiO2 Nanotube/Boron-Doped Diamond Electrodes Studied by Impedance Spectroscopy. Nanoscale 2015, 7, 551–558. [Google Scholar] [CrossRef] [PubMed]
  238. Tang, H.; Xiong, M.; Qu, D.; Liu, D.; Zhang, Z.; Xie, Z.; Wei, X.; Tu, W.; Qu, D. Enhanced Supercapacitive Performance on TiO2@C Coaxial Nano-Rod Array through a Bio-Inspired Approach. Nano Energy 2015, 15, 75–82. [Google Scholar] [CrossRef]
  239. Chen, J.; Xia, Z.; Li, H.; Li, Q.; Zhang, Y. Preparation of Highly Capacitive Polyaniline/Black TiO2 Nanotubes as Supercapacitor Electrode by Hydrogenation and Electrochemical Deposition. Electrochim. Acta 2015, 166, 174–182. [Google Scholar] [CrossRef]
  240. Ambade, R.B.; Ambade, S.B.; Shrestha, N.K.; Nah, Y.-C.; Han, S.-H.; Lee, W.; Lee, S.-H. Polythiophene Infiltrated TiO2 Nanotubes as High-Performance Supercapacitor Electrodes. Chem. Commun. 2013, 49, 2308. [Google Scholar] [CrossRef]
  241. Gao, B.; Li, X.; Ma, Y.; Cao, Y.; Hu, Z.; Zhang, X.; Fu, J.; Huo, K.; Chu, P.K. MnO2–TiO2/C Nanocomposite Arrays for High-Performance Supercapacitor Electrodes. Thin Solid Films 2015, 584, 61–65. [Google Scholar] [CrossRef]
  242. Yang, F.; Yao, J.; Liu, F.; He, H.; Zhou, M.; Xiao, P.; Zhang, Y. Ni–Co Oxides Nanowire Arrays Grown on Ordered TiO2 Nanotubes with High Performance in Supercapacitors. J. Mater. Chem. A 2013, 1, 594–601. [Google Scholar] [CrossRef]
  243. Gobal, F.; Faraji, M. Electrodeposited Polyaniline on Pd-Loaded TiO2 Nanotubes as Active Material for Electrochemical Supercapacitor. J. Electroanal. Chem. 2013, 691, 51–56. [Google Scholar] [CrossRef]
  244. Xie, S.; Gan, M.; Ma, L.; Li, Z.; Yan, J.; Yin, H.; Shen, X.; Xu, F.; Zheng, J.; Zhang, J.; et al. Synthesis of Polyaniline-Titania Nanotube Arrays Hybrid Composite via Self-Assembling and Graft Polymerization for Supercapacitor Application. Electrochim. Acta 2014, 120, 408–415. [Google Scholar] [CrossRef]
  245. Zhi, J.; Zhao, W.; Lin, T.; Huang, F. Boosting Supercapacitor Performance of TiO2 Nanobelts by Efficient Nitrogen Doping. ChemElectroChem 2017, 4, 2328–2335. [Google Scholar] [CrossRef]
  246. Pan, Z.; Liu, X.; Yang, J.; Li, X.; Liu, Z.; Loh, X.J.; Wang, J. Aqueous Rechargeable Multivalent Metal-Ion Batteries: Advances and Challenges. Adv. Energy Mater. 2021, 11, 2100608. [Google Scholar] [CrossRef]
  247. Novák, P.; Scheifele, W.; Haas, O. Magnesium Insertion Batteries-an Alternative to Lithium? J. Power Sources 1995, 54, 479–482. [Google Scholar] [CrossRef]
  248. Li, C.; Hou, C.-C.; Chen, L.; Kaskel, S.; Xu, Q. Rechargeable Al-Ion Batteries. EnergyChem 2021, 3, 100049. [Google Scholar] [CrossRef]
  249. Liu, S.; Hu, J.J.; Yan, N.F.; Pan, G.L.; Li, G.R.; Gao, X.P. Aluminum Storage Behavior of Anatase TiO2 Nanotube Arrays in Aqueous Solution for Aluminum Ion Batteries. Energy Environ. Sci. 2012, 5, 9743. [Google Scholar] [CrossRef]
  250. Xiao, D.; Zhang, L.; Li, Z.; Dou, H.; Zhang, X. Design Strategies and Research Progress for Water-in-Salt Electrolytes. Energy Storage Mater. 2022, 44, 10–28. [Google Scholar] [CrossRef]
  251. Chen, M.; Feng, G.; Qiao, R. Water-in-Salt Electrolytes: An Interfacial Perspective. Curr. Opin. Colloid Interface Sci. 2020, 47, 99–110. [Google Scholar] [CrossRef]
  252. Zhou, A.; Liu, Y.; Zhu, X.; Li, X.; Yue, J.; Ma, X.; Gu, L.; Hu, Y.-S.; Li, H.; Huang, X.; et al. TiO2(B) Anode for High-Voltage Aqueous Li-Ion Batteries. Energy Storage Mater. 2021, 42, 438–444. [Google Scholar] [CrossRef]
  253. Yan, L.; Huang, J.; Guo, Z.; Dong, X.; Wang, Z.; Wang, Y. Solid-State Proton Battery Operated at Ultralow Temperature. ACS Energy Lett. 2020, 5, 685–691. [Google Scholar] [CrossRef]
  254. Donald, W.A.; Leib, R.D.; Demireva, M.; O’Brien, J.T.; Prell, J.S.; Williams, E.R. Directly Relating Reduction Energies of Gaseous Eu(H2O)n3+ ,n = 55–140, to Aqueous Solution: The Absolute SHE Potential and Real Proton Solvation Energy. J. Am. Chem. Soc. 2009, 131, 13328–13337. [Google Scholar] [CrossRef] [PubMed]
  255. Guo, Z.; Huang, J.; Dong, X.; Xia, Y.; Yan, L.; Wang, Z.; Wang, Y. An Organic/Inorganic Electrode-Based Hydronium-Ion Battery. Nat. Commun. 2020, 11, 959. [Google Scholar] [CrossRef]
  256. Jiang, H.; Shin, W.; Ma, L.; Hong, J.J.; Wei, Z.; Liu, Y.; Zhang, S.; Wu, X.; Xu, Y.; Guo, Q. A High-Rate Aqueous Proton Battery Delivering Power Below− 78 °C via an Unfrozen Phosphoric Acid. Adv. Energy Mater. 2020, 10, 2000968. [Google Scholar] [CrossRef]
  257. Wang, S.; Zhao, X.; Yan, X.; Xiao, Z.; Liu, C.; Zhang, Y.; Yang, X. Regulating Fast Anionic Redox for High-Voltage Aqueous Hydrogen-Ion-Based Energy Storage. Angew. Chem. 2019, 131, 211–216. [Google Scholar] [CrossRef]
  258. Liang, G.; Mo, F.; Yang, Q.; Huang, Z.; Li, X.; Wang, D.; Liu, Z.; Li, H.; Zhang, Q.; Zhi, C. Commencing an Acidic Battery Based a Copper Anode with Ultrafast Proton-Regulated Kinetics and Superior Dendrite-Free Property. Adv. Mater. 2019, 31, 1905873. [Google Scholar] [CrossRef]
Figure 1. The different dimensional morphology of nanostructured TiO2 arrays.
Figure 1. The different dimensional morphology of nanostructured TiO2 arrays.
Materials 16 03864 g001
Figure 2. (a,b) The cross-sectional views of TiO2 nanowires at different magnifications. Reprinted with permission from [31], Copyright 2008 IOP Publishing. (c,d) The cross-sectional views of electrospun TiO2 nanowires. Reprinted with permission from [39], Copyright 2013 Elsevier. (eh) FESEM images of anatase TiO2 nanowires on different substrates (The insets correspond low-magnification images) and (i) rutile TiO2 nanorod arrays. Reprinted with permission from [23], Copyright 2014 The Royal Society of Chemistry.
Figure 2. (a,b) The cross-sectional views of TiO2 nanowires at different magnifications. Reprinted with permission from [31], Copyright 2008 IOP Publishing. (c,d) The cross-sectional views of electrospun TiO2 nanowires. Reprinted with permission from [39], Copyright 2013 Elsevier. (eh) FESEM images of anatase TiO2 nanowires on different substrates (The insets correspond low-magnification images) and (i) rutile TiO2 nanorod arrays. Reprinted with permission from [23], Copyright 2014 The Royal Society of Chemistry.
Materials 16 03864 g002
Figure 3. (a,b) The top and cross-sectional FESEM images of rutile TiO2 nanorod arrays film grown on FTO substrate. Reprinted with permission from [27], Copyright 2009 American Chemical Society. (ce) FESEM image and (f) TEM image of rutile TiO2 nanorod arrays film grown on Ti substrate. Reproduced with permission from [46], Copyright 2011 Elsevier.
Figure 3. (a,b) The top and cross-sectional FESEM images of rutile TiO2 nanorod arrays film grown on FTO substrate. Reprinted with permission from [27], Copyright 2009 American Chemical Society. (ce) FESEM image and (f) TEM image of rutile TiO2 nanorod arrays film grown on Ti substrate. Reproduced with permission from [46], Copyright 2011 Elsevier.
Materials 16 03864 g003
Figure 4. (a,b) The SEM of TiO2 nanotubes in different electrolyte. Reprinted with permission from [68], Copyright 2012 American Chemical Society. (c) SEM image of TiO2 3D nanoforest. (d) SEM image, (e) TEM images of hollow nanostructures in the cap and nanowire regions are shown in (iii) and (iv) respectively, (f) SEM images after 600 depositions, (g) SEM image and (h,i) Low- and high-(ii) magnification TEM images of titanium dioxide nanowires using ZnO nanowires as a template. Reprinted with permission from [81], Copyright 2014 American Chemical Society.
Figure 4. (a,b) The SEM of TiO2 nanotubes in different electrolyte. Reprinted with permission from [68], Copyright 2012 American Chemical Society. (c) SEM image of TiO2 3D nanoforest. (d) SEM image, (e) TEM images of hollow nanostructures in the cap and nanowire regions are shown in (iii) and (iv) respectively, (f) SEM images after 600 depositions, (g) SEM image and (h,i) Low- and high-(ii) magnification TEM images of titanium dioxide nanowires using ZnO nanowires as a template. Reprinted with permission from [81], Copyright 2014 American Chemical Society.
Materials 16 03864 g004
Figure 5. Schematic illustration, SEM images, and TEM images of (ad) random, (eh) sealed, and (il) unsealed TiO2 nanotubes. Reproduced with permission from [82], Copyright 2012 American Chemical Society.
Figure 5. Schematic illustration, SEM images, and TEM images of (ad) random, (eh) sealed, and (il) unsealed TiO2 nanotubes. Reproduced with permission from [82], Copyright 2012 American Chemical Society.
Materials 16 03864 g005
Figure 6. Top view (ac) cross-sectional view SEM images for layered TiO2 nanosheet arrays. Reprinted with permission from [94], Copyright 2015 Elsevier. (d) Overall strategy and SEM images toward 2D TiO2 nanosheet arrays. (e) Schematic illustration for the formation of nanosheet arrays. Reprinted with permission from [95], Copyright 2011 American Chemical Society.
Figure 6. Top view (ac) cross-sectional view SEM images for layered TiO2 nanosheet arrays. Reprinted with permission from [94], Copyright 2015 Elsevier. (d) Overall strategy and SEM images toward 2D TiO2 nanosheet arrays. (e) Schematic illustration for the formation of nanosheet arrays. Reprinted with permission from [95], Copyright 2011 American Chemical Society.
Materials 16 03864 g006
Figure 7. SEM images of (a,b) TiO2 nanoarrays film, (c,d) the core-shell branched nanowire arrays and (e,f) the core-shell branched nanobelt arrays. Reprinted with permission from [96], Copyright 2015 Springer Nature.
Figure 7. SEM images of (a,b) TiO2 nanoarrays film, (c,d) the core-shell branched nanowire arrays and (e,f) the core-shell branched nanobelt arrays. Reprinted with permission from [96], Copyright 2015 Springer Nature.
Materials 16 03864 g007
Figure 8. (a,b) SEM images of TiO2 nanobelts. Reprinted with permission from [97], Copyright 2013 The Royal Society of Chemistry. (c,d) SEM images of the TiO2@MnO2 nanobelt arrays. Reprinted with permission from [98], Copyright 2013 The Royal Society of Chemistry.
Figure 8. (a,b) SEM images of TiO2 nanobelts. Reprinted with permission from [97], Copyright 2013 The Royal Society of Chemistry. (c,d) SEM images of the TiO2@MnO2 nanobelt arrays. Reprinted with permission from [98], Copyright 2013 The Royal Society of Chemistry.
Materials 16 03864 g008
Figure 9. (af) SEM images of nanotrees. Reprinted with permission from [104], Copyright 2009 American Chemical Society. (g) FESEM image and (h) TEM image of the F decorated TiO2 nanowires. Reprinted with permission from [103], Copyright 2023 Elsevier.
Figure 9. (af) SEM images of nanotrees. Reprinted with permission from [104], Copyright 2009 American Chemical Society. (g) FESEM image and (h) TEM image of the F decorated TiO2 nanowires. Reprinted with permission from [103], Copyright 2023 Elsevier.
Materials 16 03864 g009
Figure 10. Schematic illustrations and SEM/TEM images of (ac) nanosheet branches and (df) nanosheet and nanorod branches. Reprinted with permission from [109], Copyright 2014 Springer Nature. Schematic illustration of the H-TiO2 NRAs photoanode: (g) TiO2 NRAs, (h) branched ZnO/TiO2 NRAs, and (i) H-TiO2 NRAs. Reprinted with permission from [25], Copyright 2015 The Royal Society of Chemistry.
Figure 10. Schematic illustrations and SEM/TEM images of (ac) nanosheet branches and (df) nanosheet and nanorod branches. Reprinted with permission from [109], Copyright 2014 Springer Nature. Schematic illustration of the H-TiO2 NRAs photoanode: (g) TiO2 NRAs, (h) branched ZnO/TiO2 NRAs, and (i) H-TiO2 NRAs. Reprinted with permission from [25], Copyright 2015 The Royal Society of Chemistry.
Materials 16 03864 g010
Figure 11. (a,b) FESEM images and (c) TEM images of hierarchical TiO2 nanotube arrays. Reprinted with permission from [112], Copyright 2014 Wiley. (d) FESEM images and (e) TEM images of TiN and TiN/TiO2 nanowires. Reprinted with permission from [113], Copyright 2017 Elsevier.
Figure 11. (a,b) FESEM images and (c) TEM images of hierarchical TiO2 nanotube arrays. Reprinted with permission from [112], Copyright 2014 Wiley. (d) FESEM images and (e) TEM images of TiN and TiN/TiO2 nanowires. Reprinted with permission from [113], Copyright 2017 Elsevier.
Materials 16 03864 g011
Figure 12. (a) Schematic diagram and (b,c) FESEM images of TiO2 nanoflower arrays. Reprinted with permission from [114], Copyright 2022 The Royal Society of Chemistry. (d) FESEM image and (e,f) TEM image of 3D TiO2 nanoarrays. (g,h) Characterizations of the change in the lattice parameters and its generation mechanism. Reprinted with permission from [115], Copyright 2021 American Chemical Society.
Figure 12. (a) Schematic diagram and (b,c) FESEM images of TiO2 nanoflower arrays. Reprinted with permission from [114], Copyright 2022 The Royal Society of Chemistry. (d) FESEM image and (e,f) TEM image of 3D TiO2 nanoarrays. (g,h) Characterizations of the change in the lattice parameters and its generation mechanism. Reprinted with permission from [115], Copyright 2021 American Chemical Society.
Materials 16 03864 g012
Figure 13. (a) Diagram, (b) TEM image, (c,d) EDS mapping images, (e) HAADF-STEM image, and (f) FESEM image (the inset shows the cross-section) of TiN nanotrees. Reprinted with permission from [117], Copyright 2019 The Royal Society of Chemistry.
Figure 13. (a) Diagram, (b) TEM image, (c,d) EDS mapping images, (e) HAADF-STEM image, and (f) FESEM image (the inset shows the cross-section) of TiN nanotrees. Reprinted with permission from [117], Copyright 2019 The Royal Society of Chemistry.
Materials 16 03864 g013
Figure 14. Electrochemical performances of TiO2 nanotrees. (a) The CV curves. (b) The cycling performance. (c) Cycling stability at 1.0 mA cm−2. (d) The rate capability. (e) Comparison of the rate capability of TiO2 nanotrees and other materials [46,154,153,154,155,156]. Reprinted with permission from [154], Copyright 2016 The Royal Society of Chemistry.
Figure 14. Electrochemical performances of TiO2 nanotrees. (a) The CV curves. (b) The cycling performance. (c) Cycling stability at 1.0 mA cm−2. (d) The rate capability. (e) Comparison of the rate capability of TiO2 nanotrees and other materials [46,154,153,154,155,156]. Reprinted with permission from [154], Copyright 2016 The Royal Society of Chemistry.
Materials 16 03864 g014
Figure 15. Electronical performance of the TiN/TiO2 Nanowire Arrays. (a) The CV curves. (b) The GCD profiles. (c) Determination of b value using the relationship between peak current and scan rate. (d) Voltage offset (ΔEp) of TiN/TiO2 nanowire arrays. (e) Contribution ratio of the capacitive and diffusion-controlled capacities. (f) The rate capability of TiN/TiO2 nanowire arrays [113,155,156,171]. (g) Cycling performance. Reprinted with permission from [113], Copyright 2017 Elsevier.
Figure 15. Electronical performance of the TiN/TiO2 Nanowire Arrays. (a) The CV curves. (b) The GCD profiles. (c) Determination of b value using the relationship between peak current and scan rate. (d) Voltage offset (ΔEp) of TiN/TiO2 nanowire arrays. (e) Contribution ratio of the capacitive and diffusion-controlled capacities. (f) The rate capability of TiN/TiO2 nanowire arrays [113,155,156,171]. (g) Cycling performance. Reprinted with permission from [113], Copyright 2017 Elsevier.
Materials 16 03864 g015
Figure 16. The electrochemical test of F-decorated TiO2 nanoarrays. (a) The CV curves. (b) Determination of the b value using the relationship between peak current and scan rate. (c) The GCD profiles and (d) areal capacitances of different samples. (e) The GCD profiles. (f) Comparing the areal capacitances with other TiO2 materials [105,218,219,220,223,225,226,228]. (g) The cycling stability. Reprinted with permission from [105], Copyright 2023 Elsevier.
Figure 16. The electrochemical test of F-decorated TiO2 nanoarrays. (a) The CV curves. (b) Determination of the b value using the relationship between peak current and scan rate. (c) The GCD profiles and (d) areal capacitances of different samples. (e) The GCD profiles. (f) Comparing the areal capacitances with other TiO2 materials [105,218,219,220,223,225,226,228]. (g) The cycling stability. Reprinted with permission from [105], Copyright 2023 Elsevier.
Materials 16 03864 g016
Figure 17. The electrochemical performances of anatase TiO2 nanoarrays. (a) The CV curves. (b) The GCD profiles. (c) The specific capacities of the half-cell. (d) Schematic diagram of the full cell. (e) The GCD profiles of the full cell. (f) Ragone plots of batteries [110,253,255,256,257,258]. Reprinted with permission from [110], Copyright 2021 American Chemical Society.
Figure 17. The electrochemical performances of anatase TiO2 nanoarrays. (a) The CV curves. (b) The GCD profiles. (c) The specific capacities of the half-cell. (d) Schematic diagram of the full cell. (e) The GCD profiles of the full cell. (f) Ragone plots of batteries [110,253,255,256,257,258]. Reprinted with permission from [110], Copyright 2021 American Chemical Society.
Materials 16 03864 g017
Table 1. Brief summary of various synthesis methods of TiO2 nanoarrays.
Table 1. Brief summary of various synthesis methods of TiO2 nanoarrays.
SampleSynthesis MethodCrystal StructureMorphologyRef.
Oriented single crystalline TiO2 nanowire arraysAlkali hydrothermal and ion-exchange reactionAnataseDiameter: 105 nm and length: 12–16 μm[31]
Mesoporous TiO2 nanowire arraysAlkali hydrothermal and ion-exchange reactionAnataseDiameter: 20 nm and length: 7 μm[29]
Oriented single-crystalline TiO2 nanorod arraysHydrothermal methodRutileDiameter: 90 nm, length: 1.9 μm[27]
Ordered TiO2 nanorod arraysChemical oxidation with
30 mass % H2O2 solution
Anatase and rutileDiameter: 20–30 nm and length: 150 nm[14]
TiO2 nanotube arraysElectrochemical depositionAnataseTube length: 8 μm, inner diameter: 70 nm, and wall thickness: 25 nm[76]
Sealed TiO2 nanotube arraysTemplate assisted methodsAnataseTube length: 2 μm, inner diameter: 80 nm, and wall thickness: 20 nm[82]
Long and hollow
TiO2 nanofiber arrays with uniform, circular cross-sections
ElectrospinningAnataseTube length: 4 μm, inner diameter: 200 nm, and wall thickness: 50 nm[83]
Layered TiO2 nanosheet arraysHydrothermal treatmentTiO2(B)/AnataseLength: 6 μm and thickness: 10 nm[90]
Nanoporous TiO2
nanosheet arrays
Hydrothermal treatment with diluted HNO3 aqueous solutionRutileDiameter: 12 μm, thickness: 200–300 nm and a smooth surface[95]
TiO2 ultrathin nanobelt arraysH2O2-asisted dissolution/precipitation processAnataseThickness: 1–2 nm and a high specific surface area: 193 cm3 g−1[96]
Oriented assembled TiO2 hierarchical nanowire arraysSolvothermal method for trunks and hydrothermal treatment for branchesAnataseA length of branches: 70 nm and an average thickness: 3 μm[88]
Hierarchically tunable TiO2 nanoarraysAcid vapor oxidationRutileBranches diameter: 100 nm, length: 300 nm and trees reach 3 μm in height and have a diameter of 280 nm[104]
TiO2 nanowire arrays with F-decorated TiO2 nanoparticlesH2O2-asisted dissolution/precipitation processAnataseDiameter: 100 nm and the thickness: 2.3 μm[105]
Branched hierarchical TiO2 nanotubes on TiO2 nanorod arraysHydrothermal method and sol–gel methodAnatase for trunks and rutile for branchesBranches diameter: 100 nm, length: 500 nm and nanorods reach 1 μm in height and have a diameter of 50–100 nm[25]
Single-crystal-like
TiO2 hierarchical nanowire arrays
Hydrothermal synthesis and solvothermal methodAnataseDiameter: 130–180 nm and the size of nanoparticles is 10–20 nm[115]
Table 2. Capacitive performances of TiO2 nanoarrays for Li-ion batteries.
Table 2. Capacitive performances of TiO2 nanoarrays for Li-ion batteries.
Types of TiO2 Nanostructured ArraysCurrent DensitySpecific CapacitanceCycle Number and RetentionRef.
TiO2 nanowire arrays70 mA g−1320 mAh g−1230 mAh g−1 (20th)[174]
Amorphous TiO2 nanotube on Si5 μA cm−2196 μAh cm−256 μAh cm−2 (50th)[175]
Crystalline TiO2 nanotube on Si5 μA cm−2165 μAh cm−240 μAh cm−2 (50th)[175]
Amorphous TiO2 nanotube on Ti foil5 μA cm−2129 μAh cm−237 μAh cm−2 (50th)[175]
Crystalline TiO2 nanotube on Ti foil5 μA cm−283 μAh cm−229 μAh cm−2 (50th)[175]
TiN@TiO2 nanowire arrays0.11 mA cm−20.156 mAh cm−20.151 mAh cm−2 (300th)[113]
Rutile TiO2 nanorod arrays15 μA cm−2133 μAh cm−2130 μAh cm−2 (50th)[46]
Hydrogenated TiO2 nanotube arrays200 μA cm−20.2 mAh cm−20.18 mAh cm−2 (100th)[21]
TiO2 nanotube arrays annealed in N2320 mA g−1240 mAh g−1170 mAh g−1 (50th)[172]
Self-organized TiO2 nanotubes5 µA cm−20.14 mAh cm−20.07 mAh cm−2 (50th)[176]
TiO2 nanotube arrays annealed in CO320 mA g−1223 mAh g−1179 mAh g−1 (50th)[173]
Self-organized amorphous TiO2 nanotube arrays10 μA cm−2103 μAh cm−2101μAh cm−2 (100th)[22]
TiO2 nanotubes (from Amorphous to Cubic Phase)7 A g−1230 mAh g−1220 mAh g−1 (600th)[177]
TiO2 nanotrees1.0 mA cm−2159 mAh cm−2152 mAh cm−2 (400th)[154]
Sandwich-like, stacked TiO2 nanosheets10 C175 mAh g−1160 mAh g−1 (150th)[178]
Dual-phase Li4Ti5O12-TiO2 nanowire arrays10 C135.5 mAh g−1129.3 mAh g−1 (100th)[179]
TiO2@α-Fe2O3 core/shell arrays120 mA g−1475 mAh g−1480 mAh g−1 (150th)[180]
Sn/SnO@TiO2 nanowire arrays50 μA cm−2140 μAh cm−2120 μAh cm−2 (50th)[169]
Sn-doping TiO2 nanotube arrays70 μA cm−270 μAh cm−262 μAh cm−2 (50th)[181]
TiO2-MoO3 Core-Shell nanowire array250 mA g−1600 mAh g−1500 mAh g−1 (100th)[182]
SnO2 nanocrystals@TiO2 nanotubes20 μA cm −255 μAh cm−235 μAh cm−2 (100th)[183]
Coaxial SnO2@TiO2 nanotube array100 μA cm−2225 μAh cm−2150 μAh cm−2 (50th)[184]
TiO2 nanotubes with Co3O4/NiO particles70 μAh cm−2110 μAh cm−2103 μAh cm−2 (25th)[185]
Nitridated TiO2 hollow nanofibers0.2 C180 mAh g−1170 mAh g−1 (100th)[186]
Sealed TiO2 nanotubes array0.2 C190 mAh g−1185 mAh g−1 (100th)[82]
unsealed TiO2 nanotubes array0.2 C195 mAh g−1190 mAh g−1 (100th)[82]
randomly oriented TiO2 nanotubes0.2 C140 mAh g−1139 mAh g−1 (100th)[82]
Ordered mesoporous TiO2-C nanocomposite1 C175 mAh g−1166 mAh g−1 (900th)[187]
Mesoporous CNT@TiO2-C nanocable50 C150 mAh g−1127 mAh g−1 (2000th)[188]
TiO2@SnO2 nanoflake nanotube arrays1.6 A g−1620 mAh g−1530 mAh g−1 (50th)[189]
SnO2@TiO2 heterojunction nanotubes20 μA cm−250 μAh cm−235 μAh cm−2 (30th)[190]
SnO2@TiO2 hollow microtubes (array)200 mA g−1900 mAh g−1800 mAh g−1 (100th)[36]
SnO2@TiO2 double-shell nanotubes (array)1500 mA g−1250 mAh g−1232 mAh g−1 (30th)[191]
NiO@TiO2 nanotube heterojunction arrays0.02 mA cm−2325 μAh cm−2275 μAh cm−2 (10th)[192]
Anatase TiO2 ultrathin nanobelts1 C204 mAh g−1198 mAh g−1 (60th)[96]
Oriented anatase TiO2 nanotube arrays0.25 C250 mAh g−1190 mAh g−1(10th)[193]
Table 3. Capacitive performances of TiO2 nanoarrays for Na-ion batteries.
Table 3. Capacitive performances of TiO2 nanoarrays for Na-ion batteries.
Types of TiO2 Nanostructured ArraysCurrent DensitySpecific CapacitanceCycle Number and RetentionRef.
Sulfur-doped TiO2 nanotube arrays10 C140 mAh g−1130 mAh g−1 (4400th)[206]
Ni/N-doped anatase TiO2 nanotube50 mA g−1310 mAh g−1303 mAh g−1 (500th)[19]
Amorphous TiO2 nanotube0.05 A g−1100 mAh g−1140 mAh g−1 (50th)[207]
Crystalline (anatase) TiO2 nanotubular arrays0.1 mA cm–2125 μAh cm–2115 μAh cm–2 (50th)[199]
Monolithic anatase TiO2 nanotube arraysC/5161 mAh g−1156 mAh g−1 (350th)[201]
Ni-TiO2 core-shell nanoarrays50 mA g−1250 mAh g−1200 mAh g−1 (100th)[203]
TiO2-B/MoS2 nanoarraysC/10350 mAh g−1191 mAh g−1 (100th)[204]
Surface phosphorylated TiO2 nanotube arrays67 mA g−1334 mAh g−1270 mAh g−1 (100th)[205]
Table 4. Capacitive performances of TiO2 nanoarrays electrodes for supercapacitors.
Table 4. Capacitive performances of TiO2 nanoarrays electrodes for supercapacitors.
Types of TiO2 Nanostructured ArraysCurrent DensitySpecific CapacitanceCycle Number and RetentionRef.
3D-1D TiO2 microflowers5 mV s−166.50 F g−154.09 F g−1 (2000th)[227]
Oriented NiO-TiO2 nanotube arrays0.4 mA cm−22.6 F cm−23.0 F cm−2 (500th)[226]
Highly ordered TiO2 nanotube array1 mV s−1911 μF cm−2600 μF cm−2 (500th)[211]
The reduced MnCo2O4 TiO2 nanotube arrays by introduction of oxygen vacancies1 mA cm−220 mF cm−218 mF cm−2 (5000th)[20]
Pristine TiO2 nanotube arrays50 mV s−12.4 mF cm−22.0 mF cm−2 (1000th)[210]
Plasma treatment TiO2 nanotube arrays2 mA cm−27.22 mF cm−27.0 mF cm−2 (10000th)[220]
Electrochemical reduction TiO2 nanotube arrays0.01 mA cm−24 mF cm−23.8 mF cm−2 (5000th)[223]
Hydrogenation TiO2 nanotube arrays10 mVs−124 mF cm−28 mF cm−2 (1000th)[229]
Black TiO2 nanotube arrays10 mV s−120 mF cm−218 mF cm−2 (100th)[224]
MnO2/TiO2 nanotube arrays100 mV s−11.8 mF cm−21.7 mF cm−2 (100th)[221]
MnO2/TiO2 nanotube arrays100 mV s−1101 mF cm−295 mF cm−2 (100th)[230]
RuO2/TiO2 nanotube arrays5 mV s−131.82 F g−128 F g−1 (100th)[231]
NiO/TiO2 nanotube arrays0.5 mA cm−272.7 mF cm−260.2 mF cm−2 (100th)[232]
ZnO/TiO2 nanotube arrays20 mV s−1302 F g−1278 F g−1 (100th)[233]
MoO3/TiO2 nanotube arrays5 mV s−1209.6 mF cm−2201.5 mF cm−2 (100th)[233]
BiFeO3/TiO2 nanotube arrays1.1 A g−1440 F g−1423 F g−1 (100th)[234]
V2O5/TiO2 nanotube arrays0.2 mA cm−2220 F g−1210 F g−1 (100th)[235]
MWCNT/TiO2 nanotube arrays0.1 mA cm−24.4 mF cm−23.9 mF cm−2 (100th)[236]
BDD/TiO2 nanotube arrays10 mV s−17.46 mF cm−27.01 mF cm−2 (100th)[237]
C Nanorod/TiO2 nanotube arrays0.2 mA cm−240.75 mF cm−235.7 mF cm−2 (100th)[238]
PANI/TiO2 nanotube arrays0.6 A g−1993 F g−1863 F g−1 (100th)[239]
PTh/TiO2 nanotube arrays2 A g−1640 F g−1580 F g−1 (1000th)[240]
MnO2/TiO2/CNT nanotube arrays2.6 A g−1580 F g−1550 F g−1 (100th)[241]
Ni-Co/TiO2 nanotube arrays2.5 A g−12353 F g−12153 F g−1 (3000th)[242]
Pd/PANI/TiO2 nanotube arrays2.0 A g−11060 F g−1980 F g−1 (100th)[243]
PANI/APTES/TiO2 nanotube arrays0.5 A g−1380 F g−1340 F g−1 (1000th)[244]
Nitrogen doping TiO2 nanobelts1 A g−1216 F g−1198 F g−1 (10000th)[245]
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

Si, P.; Zheng, Z.; Gu, Y.; Geng, C.; Guo, Z.; Qin, J.; Wen, W. Nanostructured TiO2 Arrays for Energy Storage. Materials 2023, 16, 3864. https://doi.org/10.3390/ma16103864

AMA Style

Si P, Zheng Z, Gu Y, Geng C, Guo Z, Qin J, Wen W. Nanostructured TiO2 Arrays for Energy Storage. Materials. 2023; 16(10):3864. https://doi.org/10.3390/ma16103864

Chicago/Turabian Style

Si, Pingyun, Zhilong Zheng, Yijie Gu, Chao Geng, Zhizhong Guo, Jiayi Qin, and Wei Wen. 2023. "Nanostructured TiO2 Arrays for Energy Storage" Materials 16, no. 10: 3864. https://doi.org/10.3390/ma16103864

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