Next Article in Journal
Co3O4 Nanostructured Sensor for Electrochemical Detection of H2O2 as a Stress Biomarker in Barley: Fe3O4 Nanoparticles-Mediated Enhancement of Salt Stress Tolerance
Next Article in Special Issue
Underground Ink: Printed Electronics Enabling Electrochemical Sensing in Soil
Previous Article in Journal
Effect of Metal Oxide Semiconductor Field-Effect Transistor Output Parasitic Capacitance on Efficiency in Full-Bridge LLC DC/DC Converters
Previous Article in Special Issue
Characterization of a Heterojunction Silicon Solar Cell by Means of Impedance Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fabrication of Li4Ti5O12 (LTO) as Anode Material for Li-Ion Batteries

Institut de Minéralogie, de Physique des Matériaux et de Cosmochimie (IMPMC), Sorbonne Université, CNRS-UMR 7590, 4 place Jussieu, 75252 Paris, France
*
Author to whom correspondence should be addressed.
Micromachines 2024, 15(3), 310; https://doi.org/10.3390/mi15030310
Submission received: 13 November 2023 / Revised: 19 January 2024 / Accepted: 17 February 2024 / Published: 23 February 2024

Abstract

:
The most popular anode material in commercial Li-ion batteries is still graphite. However, its low intercalation potential is close to that of lithium, which results in the dendritic growth of lithium at its surface, and the formation of a passivation film that limits the rate capability and may result in safety hazards. High-performance anodes are thus needed. In this context, lithium titanite oxide (LTO) has attracted attention as this anode material has important advantages. Due to its higher lithium intercalation potential (1.55 V vs. Li+/Li), the dendritic deposition of lithium is avoided, and the safety is increased. In addition, LTO is a zero-strain material, as the volume change upon lithiation-delithiation is negligible, which increases the cycle life of the battery. Finally, the diffusion coefficient of Li+ in LTO (2 × 10−8 cm2 s−1) is larger than in graphite, which, added to the fact that the dendritic effect is avoided, increases importantly the rate capability. The LTO anode has two drawbacks. The energy density of the cells equipped with LTO anode is lower compared with the same cells with graphite anode, because the capacity of LTO is limited to 175 mAh g−1, and because of the higher redox potential. The main drawback, however, is the low electrical conductivity (10−13 S cm−1) and ionic conductivity (10−13–10−9 cm2 s−1). Different strategies have been used to address this drawback: nano-structuration of LTO to reduce the path of Li+ ions and electrons inside LTO, ion doping, and incorporation of conductive nanomaterials. The synthesis of LTO with the appropriate structure and the optimized doping and the synthesis of composites incorporating conductive materials is thus the key to achieving high-rate capability. That is why a variety of synthesis recipes have been published on the LTO-based anodes. The progress in the synthesis of LTO-based anodes in recent years is such that LTO is now considered a substitute for graphite in lithium-ion batteries for many applications, including electric cars and energy storage to solve intermittence problems of wind mills and photovoltaic plants. In this review, we examine the different techniques performed to fabricate LTO nanostructures. Details of the synthesis recipes and their relation to electrochemical performance are reported, allowing the extraction of the most powerful synthesis processes in relation to the recent experimental results.

1. Introduction

Lithium-ion batteries (LIBs) are rechargeable power sources built on the idea of the flow of Li+ ions back and forth between two intercalation electrodes with different redox potentials [1]. In most commercial LIBs, the negative electrode (anode) is a carbon-based material (e.g., graphite) and the positive electrode (cathode) is a lithiated transition-metal oxide. Graphite has the advantages of exceptional kinetics, low cost, and high abundance, but its low Li-intercalation potential (approaching 0 V vs. Li+/Li towards the end of charging) poses serious safety issues related to Li dendrite growth on the anode surface during the overcharge process. Therefore, extensive research has been conducted on the alternative anode [2]. Titanium-based materials (i.e., TiO2, LiTi2O4, Li2TiO3, Li4Ti5O12) belong to the class of alternative materials to graphite as anode materials of LIBs that operate at a potential above 0.8 V vs. Li+/Li, where a stable solid-electrolyte interphase (SEI) layer is not required. Since the 1950s, Jonker reported the first study of spinel Li4Ti5O12, which is a stable phase of the Li2O–TiO2 system also denoted as Li1.33Ti1.67O4, Li4/3Ti5/3O4, or [Li]8a[Li1/3Ti5/3]16d[O4]32e. Li4Ti5O12 (with the popular acronym LTO) appeared in the form of a well-crystallized black powder, in which the titanium is in the oxidation state +4 [3]. Subsequently, Deschanvres et al. [4] first reported a comprehensive structural study of single-crystal LTO. Materials with a spinel structure have a general chemical formula of AB2O4, in which the oxygen ions (the Wyckoff 32e sites in the space group Fd3m (No. 227)) form a cubic-close-packed (ccp) array with tetrahedrally (8a, 8b, 48c) and octahedrally (16c, 16d) coordinated interstices partially occupied by the A and B cations (Figure 1a). In 1973, Li3+yTi6−yO12 spinels (0 ≤ y ≤ 1) were under investigation because of their superconductivity at a relatively high transition temperature [5]. Interestingly, by tuning the oxygen concentration in the sample preparation, the spinel phase changes from LiTi2O4−δ to Li4Ti5O12 along with the superconductor-insulator phase transition. In 1983, Murphy et al. first reported the capability of Li4Ti5O12 (a = 8.357 Å) to react with n-butyllithium (n-BuLi) giving Li7Ti5O12 (a = 8.368 Å), in which all the lithium ions are likely on octahedral sites [6]. Figure 1b shows the ternary phase diagram between Li2TiO3, TiO2, and LiTiO2, on which the composition Li[Li1/3Ti5/3]O4 is converted to Li2[Li1/3Ti5/3]O4 upon lithiation corresponding to the reaction:
[Li]8a[Li1/3Ti5/3]16d[O4]32e + Li+ + e ⇆ [□]8a[Li2]16c[Li1/3Ti5/3]16d[O4]32e.
Since the 1980s, LTO has been widely investigated as a Li insertion compound. Colbow et al. first measured the galvanostatic charge–discharge properties of a Li//Li4Ti5O12 half-cell and found that LTO reacted with one Li atom [7]. In 1995, Ferg et al. [8] evaluated the LTO anode associated with spinel-type cathodes (Li1.03Mn1 97O4 and LiZn0.025Mn1 95O4) or LiCoO2 in room-temperature lithium cells. In 2001, Amatucci et al. [9] introduced the aqueous asymmetric Li4Ti5O12//AC cell as a safer hybrid pseudocapacitor showing 10–15% capacity loss after 5000 cycles at 10C charge/discharge rates. A typical plot of the Rietveld refinement of the X-ray diffraction (XRD) pattern of pristine LTO powders is displayed in Figure 1c.
The Li4Ti5O12 spinel-framework structure (white in color, Fd3m, a = 8.36 Å) can be electrochemically reduced to Li2[Li1/3Ti5/3]O4 (dark blue, Fd3m, a = 8.37 Å) at a voltage of 1.55 V and the reaction is highly reversible. Ohzuku and co-workers [10] reported that the lattice dimension, determined by X-ray diffraction measurements, did not change during the reaction in Equation (1). Lithium insertion causes a first-order displacement of the tetrahedrally-coordinated Li ions in the Li[Li1/3Ti5/3]O4 framework into octahedral sites to generate the ordered rock-salt phase Li2[Li1/3Ti5/3]O4 (Li7Ti5O12). During lithium insertion (discharging), three Li atoms move from 8a sites to 16c sites, and the inserted three Li ions move to the 16c sites via 8a sites with the simultaneous redox reaction of Ti4+/Ti3+ [11]. From experiment investigations of the structural properties of the lithiated Li4Ti5O12 electrode using neutron diffraction, Liu et al. [12] demonstrated that Li4Ti5O12 transforms to [Li0.16]8a[Li1Ti5]16d[Li5.84]16c[O12]32e (Fd3m S.G.) after the 1.55 V plateau in the discharge. Continued discharging of LTO to 0.01 V not only takes one extra Li into the LTO bulk but also promotes electrolyte reduction at the LTO surface to form a thick SEI layer. The fully discharged LTO is noted as [Li0.62]8a[Li1Ti5]16d[Li6]16c[Li0.38]48f[O12]32e (Fd3m S.G.), with 8a and 48f sites being partially occupied (Figure 1d) [12,13,14]. The reaction (Equation (1) is called a zero-strain insertion reaction. Thus, negative electrodes made of Li4Ti5O12 material can undergo many hundreds of cycles without structural disintegration. Lithium dendrite growth is completely avoided at this moderate voltage, enhancing the safety of the battery [15]. Furthermore, the voltage of a Li//Li4+xTi5O12 cell changes abruptly at the end of discharge and charge. Thus, a Li4+xTi5O12 spinel electrode provides very sharp end-of-charge and end-of-discharge indicators, which are useful for controlling cell operation and preventing overcharge abuse and over-discharge.
Figure 1. (a) The crystal structure of spinel Li4Ti5O12. The tetrahedral A site is occupied by Li+, while the octahedral B site is occupied by 1/6 Li+ and 5/6 Ti4+. Reproduced from Ref. [13]. Copyright 2017 under the terms of the Creative Commons Attribution 4.0 Int. License. (b) Ternary phase diagram for the Li–Ti–O system in the region bounded by Li2TiO3, TiO2, and LiTiO2. The line corresponding to the lithiation process between Li[Li1/3Ti5/3]O4 and Li2[Li1/3Ti5/3]O4 occurs at 1.55 V vs. Li+/Li. (c) Rietveld refinement of XRD pattern of spinel LTO powders with space group Fd3m. Reproduced with permission from [14]. (d) Schematic of the energy storage mechanism of the spinel LTO host lattice Reproduced with permission from [12]. Copyright 2019 under the ACS Free to read License.
Figure 1. (a) The crystal structure of spinel Li4Ti5O12. The tetrahedral A site is occupied by Li+, while the octahedral B site is occupied by 1/6 Li+ and 5/6 Ti4+. Reproduced from Ref. [13]. Copyright 2017 under the terms of the Creative Commons Attribution 4.0 Int. License. (b) Ternary phase diagram for the Li–Ti–O system in the region bounded by Li2TiO3, TiO2, and LiTiO2. The line corresponding to the lithiation process between Li[Li1/3Ti5/3]O4 and Li2[Li1/3Ti5/3]O4 occurs at 1.55 V vs. Li+/Li. (c) Rietveld refinement of XRD pattern of spinel LTO powders with space group Fd3m. Reproduced with permission from [14]. (d) Schematic of the energy storage mechanism of the spinel LTO host lattice Reproduced with permission from [12]. Copyright 2019 under the ACS Free to read License.
Micromachines 15 00310 g001
Due to its outstanding rate capability, the LTO battery is most promising to equip devices that require rapid charge and discharge. LTO has the ability for a recharge efficiency of up to 98%, which is considerably more than conventional energy storage mechanisms, and can be extensively used in electric vehicle (EV) charging stations, renewable energy storage power, etc. Li-ion batteries based on the Li4Ti5O12//LiFePO4 full cell, with nanostructured carbon-coated electrodes operating at a flat voltage of 1.85 V are a promising option for potential stationary storage and electric vehicle applications [16,17,18]. A Li4Ti5O12//LiFePO4 18650-type battery, which delivered a capacity of 800 mAh, showed outstanding cycling stability, with a definitively stable capacity after 20,000 cycles at a 10C (6 min) charge rate and a 5C (12 min) discharge rate, and even 95% capacity retention was observed after 30,000 cycles at a 15 C (4 min) charge rate and 5C discharge-rate. This battery was used in a real car with the charge time reduced to 5 min using a three-level charger in parallel (500 V, 125 A) [18]. Another aspect to be considered in the fast charge/discharge process is the natural increase of temperature of the batteries, which must be preserved to a precise limit to avoid thermal runaway. In practice, it is critical to keep the Li-ion cells at an operating temperature of 25–40 °C and a temperature uniformity of less than 5 °C for batteries used in EVs. Fast recharge usually requires a liquid cooling system used in different EVs, for instance in Tesla, BYD, and BMW i8 cars. In this context, the LTO battery, which consists of a Li(Ni1−y−zCoyMnz)O2 (named NCM) cathode and LTO anode, is used as a model to study the efficiency of different cooling systems [19,20]. Recently, Behi et al. studied experimentally a hybrid thermal management system (TMS) using phase change material (PCM) heat buffer plate and liquid cooling on such a LTO battery consisting of 30 23 Ah prismatic cells connected in series with no space between the cells [21]. The tests, conducted at a constant initial temperature of 25 °C and 4C charging/discharging rate between 10% and 90% of SOC, show that the temperature of each cell never exceeds 32 °C for the charging and discharging process. Moreover, this hybrid TMS was effective in providing temperature uniformity inside the module. It thus allows commercial use of the LTO battery at a 4C rate both for the charge and the discharge. LTO is also a competitive choice for grid-scale energy storage systems (ESS). Yang et al. developed a 10 Ah lithium–titanate battery with lithium cobalt oxide–lithium nickel cobalt manganese oxide dual-phase cathode and investigated its application in 100 kWh-level ESS. This battery demonstrated a specific capacity of 79 Wh kg−1 with a high-capacity retention rate of 91.8% after 1000 cycles at 55 °C and >80% capacity retention at 15 Ccell. The 125 kWh ESS shows an energy efficiency of 97.82% and 89.97% at 0.2 and 1.5 CSystem, respectively. Self-discharge and constant power tests confirmed that this battery is suitable for storing intermittent power from renewable sources such as solar and wind, and its ability to smooth the fluctuating power is confirmed by coordinating with a wind power system [22]. The use of NCM as the cathode is the first solution to increase the energy density of the LTO-based lithium-ion batteries since NCM belongs to the family of high-voltage cathode elements, compensating for the operating voltage of 1.55 V of LTO. Another high-voltage cathode element that has been considered is the spinel LiNi0.5Mn1.5O4 (LNMO) cathode having a high voltage of 4.7 V vs. Li+/Li. The use of LNMO has been postponed due to the oxidation of electrolytes present on the surface of LNMO under a high voltage and Mn cation dissolution. However, Piao et al. solved this problem by designing a nonflammable all-fluorinated electrolyte denoted as AFE with a composition of 1 mol L−1 lithium hexafluorophosphate (LiPF6) in a mixture of fluoroethylene carbonate (FEC), 3,3,3-fluoroethylmethyl carbonate (FEMC), 1,1,2,2-tetrafluoroethyl-2,2,3,3-tetrafluoropropylether (TTE) and 0.02 mol L−1 lithium difluoro(oxalato)borate (LiDFOB) [23]. At 5C, The LNMO//LTO cell is a high-energy cell with an operating voltage of 3.3 V. With this electrolyte, it delivered a capacity of 114 mAh g−1 with a capacity retention of 90.3% and an average Coulombic efficiency of 99.9% after 1500 cycles.
The performance of the LTO anode was achieved by overcoming the drawback of Li4Ti5O12 material, namely its low intrinsic electronic conductivity (ranging from 1.0 × 10−8 to 1.0 × 10−7 S cm−1, arising from the empty Ti 3d state), which prevents its electrochemical rate performances [24,25,26,27]. One solution is the nano-structuration of LTO to reduce the path of Li+ ions and electrons inside LTO. The reduction of the crystallite size L to nanometer scale can decrease the duration τ of the insertion (discharge) and deinsertion (charge) reaction (Equation (2)):
τ = L 2 4 π D ~ ,
where D ~ is the diffusion coefficient of ions. It is believed that the kinetically induced effective two-phase reaction in Li4+xTi5O12, resulting from the mixed 8a/16c occupation, is most likely responsible for the high-rate capabilities [28]. Nanosizing can also increase the interfacial surface contact between the electrode and electrolyte, leading to a high charge-discharge rate [29]. Kavan et al. [30] analyzed systematically the dependence of the particle size on the electrochemical properties of the LTO using particles ranging from 1 µm to 9 nm in thin electrode films. They found an optimum battery performance for particles with a size around 20 nm (∼100 m2 g−1) in a potential interval of 2.5–0.9 V. Improved rate performance of LTO anode can be also achieved by other strategies such as: (i) modification of the material morphology (i.e., particle shape, porosity, aggregates), (ii) introduction of an electronically conductive phase such as carbon, metal or metal oxide, and (iii) doping with foreign cations or anions, which can change the energy level distribution. Several promising routes have been recently suggested in a series of works aiming to rational design of LTO anodes [31], in relation to the electrochemical properties [32,33,34,35,36,37].
In this review, we examine the different techniques performed to fabricate LTO nanostructures. Details of the synthesis recipes and their relation to electrochemical performance are reported, allowing the extraction of the most powerful synthesis processes in relation to the recent experimental results (See Scheme 1). The electrochemical properties of LTO are dependent on synthesizing methods to a great extent. The synthesis of LTO with the appropriate structure and the optimized doping and the synthesis of composites incorporating conductive materials is thus the key to achieving high-rate capability. That is why a variety of synthesis recipes have been published on the LTO-based anodes. For each synthesis route, various parameters should be considered: starting reagents (e.g., titanium precursor or lithium salt), morphology of the precursor, presence of additive, reaction temperatures and reaction time, duration, temperature, and atmosphere of calcination, etc. The present article aims to review the techniques of fabrication of LTO anode materials. The influence of reaction conditions on the products is investigated in detail. It is shown that uniform, monodisperse, and stable mesoporous nanoparticles can be prepared in mild synthesis conditions and the particle size can be controlled from 25 to 200 nm by varying synthesis parameters and adding suitable additive agents including alcohols, amine, inorganic bases, and inorganic salts. Novel synthesis methods are also used to control the particle size (see Refs. [38,39,40,41,42]). In this work, we examine the different techniques performed to fabricate LTO nanostructures. Details of the synthesis recipes and their relation to the electrochemical performance are reported, allowing the extraction of the most powerful synthesis processes in relation to the recent experimental results.

2. Synthesis Methods

Various strategies have been reported for the synthesis of LTO with the aim of finding facile and simple processes for large-scale production. The main objective is the development of a method, which would provide a product with good homogeneity, regular morphology, and narrow particle size distribution. For instance, Wang et al. [34] reported that the synthesis of monodisperse LTO nanospheres is the key to reducing the irreversible capacity of LTO materials, which can be solved by the pH regulation of the growth process. This process provides LTO anode materials with high electrochemical performance (109 mAh g−1 at 60C rate; 92% retention after 500 cycles). Additionally, the high-temperature post-treatment is not desirable in terms of fabrication expenses. Conventional synthesis techniques include solid-state reaction (SSR) [25,43,44,45,46,47,48,49], hydrothermal [50,51,52,53,54,55,56,57,58], solvothermal method [59,60], sol–gel method [61,62,63,64,65,66,67,68,69,70], biphasic interfacial reaction [71], spray pyrolysis [72,73,74,75,76,77,78], molten-salt method [79,80,81,82], microwave heating method [83,84], template method [85,86,87,88,89], emulsion-gel process [90], reflux method [91], flux method [92], mechanochemical synthesis [93,94,95], electrospinning method [95,96], solution combustion synthesis [97,98,99,100,101,102,103], sonochemical method [104,105,106,107,108], single step metal organic precursor (SSMO) method [109], template-free hydrothermal process (TFHP) [110], spray drying process [111,112], rheological phase method [113], supercritical fluids [114], and electrospray pyrolysis method [115].
Various types of nanosized LTO materials have been synthesized. They can be classified according to their dimensionality from 0D to 3D architectures (see Table 1): 0D-LTO nanomaterials include nanoparticles [97,103,116,117,118,119], nanocrystals [120,121], nanospheres [39], and hollow spheres [85]. 1D LTO nanomaterials are composed of nanorods [122,123,124,125,126,127], nanowires [128,129,130,131,132], nanofibers [95,133,134,135,136,137,138,139,140], nanotubes [54,141], nanoflakes [142], and nanobelts [143,144,145,146]. 2D LTO nanomaterials encompass nanosheets [136,147,148,149,150,151], nanoplates [152], and wave-like structures [153]. 3D LTO nanomaterials include nanoflowers [153], nanoporous structures [55,154,155,156,157], mesoporous structures [67], hierarchical structures [158], and nanoarrays [159,160,161,162,163,164,165,166,167,168,169].
Currently, the growth of LTO nano-powders is achieved by using successive techniques, which can include the synthesis of a precursor at the start and at least one calcination step at the end. For example, Lu et al. [121] prepared nanosized LTO using a three-step process. First, α-Li2TiO3 was hydrothermally prepared (at 160 °C for 48 h) from 0.5 g anatase TiO2 (4.5 nm particles) and added to a 2 mol L−1, 40 mL LiOH aqueous solution. Second, cubic NaCl-type (Li0.4H0.6)2TiO3 was formed by mixing the white-colored α-Li2TiO3 with a 1 mol L−1 HCl aqueous solution. Finally, the product was calcined at 400 °C for 2 h to obtain nanosized LTO (40 nm in size). Another approach consists of the growth of a precursor with the requested final morphology. For instance, a novel Li4Ti5O12@carbon nanotubes (CNT) composite, composed of Li4Ti5O12 nanowires intertwined with CNT (4 wt%), was synthesized by a simple method. The Li4Ti5O12@CNT shows impressive electrochemical performances, such as high initial coulombic efficiency (ICE, ~94%), acceptable capacity (147.6 mAh g−1) and ultra-long life (3000 cycles) [128]. Similarly, Kim et al. [132] prepared LTO nanowires via a two-step ionic exchange process using Na2Ti3O7 nanowires formed by hydrothermal treatment of TiO2 powders dispersed in a 10 mol L−1 NaOH aqueous solution and subjected to ion exchange in an HCl aqueous solution to produce H2Ti3O7 nanowires. Next, the H2Ti3O7 powders were again hydrothermally treated in a LiOH aqueous solution at 100 °C for 24 h to form lithiated titanates via hydrothermal ionic exchange.
Sorensen et al. [167] fabricated 3D ordered macroporous Li4Ti5O12 samples (3DOM-LTO) synthesized using poly(methyl methacrylate) colloidal crystal templates and metal-organic aqueous precursors via the mixture (5:4 v/v) of titanyl oxalate solution and lithium acetate. The 3DOM structure, with an interconnected network of nanometer-thick walls, is an attractive architecture for battery materials. The 3DOM architecture has two types of porosity: large macropores, on the order of 100 nm, formed by the spheres, and smaller mesopores, on the order of 15 nm, formed by contact points between the spheres. The surrounding walls of the 3DOM structure are not solid monoliths but consist of crystallites that form an irregular wall surface. Surface roughness can work to the advantage of a battery electrode because it creates more defects for lithium to enter the electrode material.
Raman spectroscopy is a powerful tool for the study of the local bonding configurations in Li4Ti5O12 [170]. Moreover, a low amount (<1%) of other titanates (anatase and rutile TiO2) can be easily detected in Raman spectra, which have also been identified as a unique tool to probe structural defects (Figure 2).
Following the theoretical calculation of A[B2]O4 spinel-type compound, spinel phase Li4Ti5O12 consists of a symmetric group Oh7 with the five expected 3F2g + Eg + A1g Raman active vibrations [171,172]. The Raman spectrum of nanocrystalline LTO (Figure 2) displays characteristic bands at 234, 265, 347, 427, 517, 674, and 754 cm−1. For a complete understanding, see the analysis of the structural heterogeneity of compacted lithium titanate in Ref. [173]. The Raman band attribution is given in Table 2, together with frequency values reported in the literature [174,175,176,177].

2.1. Starting Materials

LTO synthesis techniques have been developed using various titanium and lithium raw materials with the main objective of a low-cost process and good control of the final product morphology. Thus, the choice of precursor for the LTO synthesis is crucial for its application as anode materials for lithium-ion batteries because the particle dispersibility and particle size have peculiar features yielding high Li+ diffusion coefficient and better kinetics of Li+ ions during charge transfer reactions [178,179]. In the pseudo-binary phase diagram of the Li2O-TiO2 system, the Li4Ti5O12 region is extremely narrow, making the growth of phase-pure LTO difficult. LTO is usually produced via the formation (or use) of an intermediate Ti-oxide phase converted to the final product through thermal treatment.

2.1.1. Titanium Alkoxides

An alkoxide is the conjugate base of an alcohol and consists of an organic group bonded to a negatively charged oxygen atom (noted RO). Alkoxides of titanium (IV), which are widely employed in organic synthesis of LTO, include titanium ethoxide Ti4(OC2H5)16 (noted Ti(OEt)4 or TET), titanium butoxide Ti(OC4H9)4 (noted Ti(OBu)4 or TBT), titanium isopropoxide Ti(OC3H7)4 (noted Ti(O-i-Pr)4 or TIP) and tetraethyl ortho-titanate (Ti(OC2H5)4 (noted TOT). All titanium alkoxides react with water to deposit TiO2 in the form of powders or thin films, which can be expressed with the balanced equation:
Ti(OC3H7)4 + 2H2O → TiO2 + 4OC3H8.
Generally, the crystallinity and morphology of the TiO2 product can be modified by the addition of a weak acid such as acetic acid and by the change in hydrolysis ratio. To provide considerable amounts of unreacted TiO2 phase (or other products such as Li1+xTi2−xO4, Li2TiO3), inventors developed methods comprising an organo-titanium compound selected from titanic acid esters in an organic solvent. Particularly preferred alkoxides as starting reagents are titanium(IV) isopropoxide and titanium(IV) butoxide (or tetrabutyl titanate). For example, ultrafine LTO nanoparticles can be easily synthesized by hydrothermal and solvothermal methods by reacting TBT or TIP with lithium hydroxide or lithium acetate. However, these processes involve high energy consumption and cost, which prevent them from being used in large-scale productions and practical applications [180,181]. For the fabrication of the LTO/CNT composite, the LTO precursor was prepared by the mixture of 47.5 mmol TBT with 66.5 mmol lithium nitrate dissolved in a mixture of 54 mL ethanol and 4.2 mL (36% in H2O) hydrochloric acid, which produces a slightly yellow clear solution [182]. In a typical approach to fabricating LTO-coated TiO2 nanotube arrays, the LTO precursor was prepared by mixing three solutions. Stoichiometric amounts of tetrabutyl titanate (Ti(OC4H9)4) (solution A) and citric acid (solution B) were dissolved in alcohol, respectively. In solution B, the citric acid to total metal ions of solution A ratio was 0.8. An alcoholic solution of CH3COOLi·2H2O was added to a mixed solution of A and B, the Ti/Li = 0.8, and a transparent sol was obtained after stirring for 4 h [183]. Recently, Wang et al. [184] experimented successfully with an energy-saving solid-phase synthesis route for LTO powders using metatitanic acid (H2TiO3) as a titanium source. Then, 0.7832 g H2TiO3 powder was mixed with 0.2560 g Li2CO3 (excessive 4%) in an agate mortar directly. Then, the precursor powder was heated at 500 °C for 2 h and heated to 700 °C with a heating rate of 5 °C min−1 for 6 h in Ar atmosphere.

2.1.2. TiO2 as Ti Source

With most titanium salts, the starting precursors are easily hydrolyzed to form TiO2. Various processes have adopted the use of a TiO2 precursor to develop low-cost LTO [39,93,102,115,179,184,185,186,187,188,189,190,191,192,193,194]. Commercial or homemade TiO2 (anatase or rutile phase) and tetrabutyl titanate (Ti(C4H9O)4) are the most popular titanium sources used for the preparation of LTO. The use of micro-sized TiO2 as a starting material can efficiently reduce the synthesis cost of LTO. Nanosized TiO2 (size 20 nm) commercialized as P25 is composed of 85% rutile and 15% anatase phase (price of €212 per kg). Amorphous TiO2 has been also employed [193]. The precursor TiO2 can be prepared from H2TiO3 synthesized through the following procedures: (1) industrial titanyl sulfate solution was diluted with de-ionized water to obtain 20 g L−1 (TiO2) solution; (2) the diluted solution was boiled in a 1000 mL round-bottomed flask attaching to a refluxing condenser with a constant temperature of 105 °C; (3) the solution was then hydrolyzed for 2 h under vigorous stirring; (4) a white precipitate formed gradually and the H2TiO3 precipitate was washed 2 times with a sulfuric acid aqueous solution of pH 1–2, then washed with de-ionized water several times until no sulfate ion was present (determined by 0.5 mol L−1 barium chloride solution), and finally dried in an oven at 80 °C. H2TiO3 was calcined at 850 °C for 5 h in a tubular furnace to synthesize the precursor TiO2. To reduce the synthesis cost and increase the efficiency of LTO, Hong and co-workers [193] prepared TiO2 microspheres (1–2 µm in size) assembled with nanoparticles through the urea-based hydrolysis of titanium (IV) oxysulfate (TiOSO4), which is generally produced during the sulfate process in titanium mining of ilmenite (FeTiO3). Table 3 provides some examples of Ti and Li raw materials used in LTO synthesis techniques.
The conversion of nanocrystalline TiO2 (anatase) into LTO was explored by a reaction of colloidal TiO2 with LiOH. However, this strategy was not successful, nor its variants employing Li2CO3, CH3COOLi, and LiNO3 in combination with the stoichiometric amount of colloidal anatase in acidic or alkaline media at temperatures up to 250 °C (in autoclave). In all cases, the product contained Li1+xTi2−xO4 with considerable amounts of unreacted anatase [195]. Li et al. [179] used an axiolitic TiO2 to prepare LTO nanoparticles. The axiolitic TiO2 phase, which was assembled by 10–20 µm particles, was synthesized by solvothermal reaction using the mixture of acetic acid (50 mL) and tetrabutyl titanate (TBT, 2 mL) forming a white suspension. After stirring for 15 min, the suspension was hydrothermally treated at 180 °C for 24 h in a 100 mL Teflon-lined stainless-steel autoclave and collected by centrifugation, washed with ethanol several times, and dried at 60 °C for 24 h. Finally, the precursor of anatase TiO2 was obtained after calcination at 500 °C for 4 h. It is rarely reported that nano-sized materials can be synthesized when ordinary TiO2 is used as the source of titanium. For example, Han et al. prepared spinel Li4Ti5O12 with the size of ~1 μm using ordinary TiO2 and Li2CO3 through solid-state reaction, which showed poor high-rate capacity [196]. On the contrary, the use of the precursor with special morphology or special treatment makes it possible to obtain as-prepared LTO particles of the same morphology and size as those of the precursors, displaying excellent high-rate capacity [197]. Industrial titanyl sulfate solution is an intermediate product in the commercial preparation of TiO2 by sulfate route, which is obtained by acidulating ilmenite (FeTiO3) with sulfuric acid that can solubilize titanium to form its sulfate [198].
Numerous groups of researchers have controlled the morphology of the precursor using as-prepared TiO2 instead of commercial titanium dioxide [57,107,181,199,200]. Tsai and co-workers [199] fabricated monodispersed sphere-like TiO2 precursors (~15 nm) as Ti source for the growth of LTO via a sol–gel method using titanium isopropoxide (TIP) with carboxylic acid addition. In a typical synthesis procedure, acetic acid (3.75 mL, 65 mmol) was injected into the 25 mL anhydride ethanol, followed by the addition of 0.3 mL (1 mmol) TIP. The solution was then heated at 85 °C for 6 h under reflux in air. After that, hydrolysis and condensation were initiated by adding a 7 mL solution consisting of 2 mL DI water and 5 mL ethanol causing the particles to gradually precipitate over 4 h, turning the solution turbid. The residual ethanol and DI water were then pumped out in a vacuum system. TiO2 submicron spheres were also prepared with different volumes of butyric (0.3~1.25 mL), valeric (0.5~2 mL), and octanoic (0.5~2 mL) acids, which resulted in larger particle sizes. Jin et al. [107] prepared small and uniformly sized LTO particles (50 to 100 nm) from TiO2 powders prepared through the urea-forced hydrolysis/precipitation route below 100 °C, which have extraordinarily large surface areas of more than 250 m2 g−1. Wang and co-workers [57] used a microemulsion method to prepare ultrafine TiO2 precursor for the synthesis of very fine LTO nanoparticles through moderate calcination. TiO2 particles were fabricated from the mixture of two solutions: (i) a milky solution obtained by ultrasonic mixing of a certain amount of cetyltrimethyl ammonium bromide (CTAB) in 150 mL deionized water, and (ii) titanium isopropoxide (TIP) and oleic acid with a mole ratio of 1:0.06 dissolved into 60 mL n-hexane. The concentration of Ti4+ is 0.13 mol L−1 and the mole ratio of TIP and CTAB is 1:1.17. The yellowish precipitate was collected by filtrating, washing several times, and drying at 80 °C to produce ultrafine TiO2 monodispersed nanoparticles with a particle size of 3~5 nm.
TiO2 nanospheres were prepared via the mixture of Ti(OC4H9)4 with (CH2OH)2 poured into a solution of ethanol and DI water (63:1 in volume) and stirred for 2 h. The obtained white precipitate harvested by centrifugation was washed with ethanol and refluxed for 30 min to form the final product [200].
The microemulsion method was also used successfully by Zhang et al. [201] who synthesized LTO/graphene (G) composites by interfacial electrostatic self-assembly in a water-in-oil (W/O) microemulsion system. In the W/O microemulsion system, the aqueous phase was uniformly dispersed into discontinuous nanoscale “water pools”. The growth of Ti(OH)4 was effectively suppressed by the space confinement of the water pools, and the uniform distribution of the water pools made the titanium source hydrolyze uniformly. So, Ti(OH)4 colloid was electrostatically self-assembled with the graphene uniformly at the water-oil interface. after adding to this precursor an appropriate amount of LiOH solution, and followed by drying and calcining, LTO particles with a size of less than 50 nm were uniformly anchored on the graphene sheets. This anode demonstrated a capacity of 174 mAh g−1 (304.5 mAh cm−3) and 152 mAh g−1 (266 mAh cm−3) at 1C and 10C rate, respectively, and a 97% capacity retention after 600 cycles at 10C rate.
Liao et al. [202] synthesized highly mesoporous anatase phase hierarchical TiO2 spheres via a solvothermal method. In a typical synthesis, 0.65 mL TBT was added to 40 mL acetic acid (HAc) with stirring. After a few minutes, the mixture of TBT and HAc was transferred into a 50 mL Teflon-lined stainless-steel autoclave placed in an electronic oven and maintained at 160 °C for 12 h
In many cases, researchers used a few weight-percent excess Li sources to compensate for lithia volatilization during high-temperature heating. Few research works have addressed the issue of nanoporous or nanocrystalline LTO synthesized by soft chemical methods by dipping producing in situ lithium ethoxide that reacts with titanium alkoxide to form a double Li-Ti alkoxide without particle precipitation [67]. Lithium ethoxide (CH3CH2OLi) was obtained by the gradual exothermic reaction between metallic Li beads in a bath of ethanol cooled at −16 °C with the production of hydrogen according to the relation:
CH3CH2OHexc +Li → CH3CH2OLi + ½H2.
A similar procedure was carried out by Yu et al. [203], who dissolved 50 mg (7.2 mmol) of metallic lithium in 20 mL benzyl alcohol at 50 °C forming a yellowish solution. Several studies have shown the influence of the synthesis parameters on the structural and electrochemical properties of LTO anode materials [204,205,206,207]. Mahmoud et al. [207] investigated the effect of the gel-drying temperature, annealing time at 900 °C, and calcination temperature on the purity phase, particle size, agglomeration of the particles, and porosity of LTO grown by the sol–gel method. Xu et al. [206] explored the properties of LTO prepared by solid-state reaction with subsequent heat treatment in air and N2 atmosphere at 800 °C for 3 h. The calcination carried out in N2 ambient results in a larger lattice parameter, which can be attributed to the reduction of some Ti4+ ions into Ti3+ when the process occurred under low-oxygen partial pressure (darker color of LTO powders). Thus, the presence of Ti3+ ions is ascribed to the increase in electron concentration, which results from non-stoichiometry Li4Ti5O12−δ (Equation (5)):
O o x   1 2 O 2 g + V o ° ° + 2 e .
For example, defective mesoporous Li4Ti5O12−y anode material with improved high-rate performance was prepared by solvothermal method with oxygen vacancies and Ti3+–O2−–Ti4+ pairs formed by annealing the product in N2 at 500 °C for 2 h [208]. Defects of the samples were identified by EPR which is sensitive to the electrons and holes trapped at defect sites. The EPR spectrum for N2-treated LTO displays one intense signal at g = 2.00, which confirms the presence of oxygen vacancies and Ti3+. The high-rate performance is represented by a discharge capacity of 139 mAh g−1 at a high rate of 20C with a capacity retention of 91.4% over 300 cycles. Similar results have been reported for LTO heat treated under H2/Ar atmosphere [16]. The nanostructure of LTO material was found to depend on the preceding intermediate phase formation conditions and the applied annealing protocol.

2.1.3. Intermediate Phase

The process for producing LTO includes the steps of synthesizing a lithium titanate hydrate intermediate via aqueous chemical processing and thermally treating it to produce LTO. Several workers reported the synthesis of LTO nanoparticles via the thermal transformation (annealing) of the intermediate orthorhombic phase, lithium titanate hydrate Li1.81H0.19Ti2O5·xH2O (LTH), obtained by hydrothermal process [209,210,211,212,213]. One advantage of converting the LTH intermediate phase into LTO is its theoretical Li/Ti ratio, 0.905, higher than 0.8 LTO. The over-stoichiometric ratio with uniform distribution in atomic scale can compensate for Li loss during sintering, and provide high homogeneity and purity in the final LTO product. In the patent filed by Demopoulos et al. [211], the LTH intermediate was synthesized as follows: a volume of 20 mL of a 2 mol L−1 TiCl4 aqueous solution was added dropwise to 180 mL of an ice-cold 1.33 mol L−1 LiOH stirred aqueous solution (Li/Ti molar ratio equals 6). The temperature was maintained below 10 °C during addition. The terminal pH of the reaction was within the range of 11.5–12, and the mixture was stirred for an additional 2 h. The as-neutralized precipitate was collected by centrifugation and washed 3 times with deionized (DI) water. The recovered product was transferred to a closed vessel for aging at 80 °C without stirring for 36 h. The LTH intermediate was then further dried in an oven at 80 °C. Also, Chiu and Demopoulos [210] reported the production of nanoflower-structured LTO with “petals” 17–50 nm thick by annealing the LTH (x = 2) intermediate phase at 400 °C. Liu et al. [214] reported a hydrothermal treatment of a solution containing titanium isopropoxide as a precursor, LiOH, plus H2O2, carried out at 130–170 °C to prepare an intermediate that is subsequently transformed to LTO by calcination at 550 °C. Xu et al. [215] analyzed the intermediate LTH phase (white powder) obtained by hydrothermal reaction at 180 °C for 36 h of 1.7 mL (5 mmol) of tetrabutyl titanate, and 0.189 g of LiOH·H2O thoroughly mixed in 20 mL of ethanol and 25 mL DI water at room temperature (XRD pattern shown in Figure 3). Calcined at 700 °C for 6 h in air, the white hydrothermal LTH product is transformed into cubic LTO nanosheets (9 nm thick).
Another strategy is to disperse TiO2 powders in a concentrated NaOH solution to obtain sodium titanate as an intermediate phase via a hydrothermal reaction. Subsequent treatment in HCl solution for 5 h yields the ion exchange H+ for Na+ and the formation of H2Ti3O7 (HTO), which is like hydrate titanate possessing a monoclinic phase (C2/m space group). Further, this intermediate phase is ultrasonically mixed with LiOH and hydrothermally treated at 150 °C for 12 h providing LTO nanowires with a diameter of 50–100 nm and a length of about 10–20 µm [130].

2.1.4. Lithium Salts

For solid-state synthesis of LTO, researchers have employed varying lithium salts, i.e., Li2CO3, LiOH·H2O. Lithium acetate dihydrate (CH3COOLi·2H2O) noted as LiOAc is used as a precursor solution in many synthesis methods. Lithium nitrate, LiNO3, associated with titanyl nitrate (TiO(NO3)2) as the oxidant precursors and glycine as the fuel are generally used in the solution-combustion method, which provides porous morphology, or nanometer-sized particles [97]. Lithium ethoxide (LiC2H5O) is used in the sol–gel process with Ti(IV) alkoxides to achieve nanoparticles with high surface areas (>50 m2 g−1) and small particle sizes (<20 nm). Lithium tert-butoxide (LiOC(CH3)3) is employed as a lithium source in the solvo-thermal method. Lithium acetate starting solution in 2-propanol is associated with titanium isopropoxide (Ti(OC3H7)4) for the growth of LTO through electrospray pyrolysis [115]. Lithium acetylacetonate (LiC5H7O2) is a Li source for the fabrication of electrospun LTO that is soluble in organic solvents as an organometallic compound, in which acetylacetonate anion complexes by bonding each oxygen atom to the metallic cation to form a chelate ring.
LTO prepared by solid-state reaction with two different Li-containing precursors, i.e., Li2CO3 and LiOAc or different heat treatments, exhibit the same crystal structure as characterized by X-ray diffraction, but show different morphologies, as revealed by scanning electron microscopy (SEM) [216]. SEM images showed that primary LTO/LiOAc particles aggregate to loosely packed small clusters, while LTO/Li2CO3 particles of submicron size are connected to form a percolated network. The effects of bridged grain boundaries on maintaining the occupancy balance between 8a and 16c sites in LTO structures have an impact on the rate performance and over-potentials of electrodes.

2.2. Solid State Reaction

In large-scale industrial synthesis of active material for lithium-ion batteries, solid-state reaction (SSR) is considered one of the most simple and practicable techniques. However, this method has a distinct drawback in that the final particle size of the prepared electrode material increases greatly after high-temperature calcination. It is very difficult to obtain a product with a narrow size distribution. Two types of preparation are utilized: (i) one-step SSR [44,179,217,218] and (ii) two-step synthesis with SSR as the final process [219]. In this process, various methods were used to assist the SSR synthesis, not only for reducing the particle size of reactants but also the increasing the homogeneity of the starting mixture. The formation mechanism of LTO was investigated by [220]. Combined in situ high-temperature powder X-ray diffraction (HT-PXRD) and thermal gravimetry-differential thermal analysis (TG-DTA) showed the origin of the impurity phases and revealed that the formation of LTO from anatase TiO2 and Li2CO3 is a two-stage process. Initially, TiO2 and Li2CO3 react to form monoclinic Li2TiO3, followed at higher temperatures by a reaction with the remaining TiO2 to yield Li4Ti5O12. This is different from the mechanism suggested by Shen et al. [220], arguing that crystalline Li2CO3 changes to an amorphous Li-precursor upon heating and then Li diffuses into TiO2 to form Li2TiO3. It is also demonstrated that Li4Ti5O12 decomposes to some Ti-rich phases and probably Li2O when heated above 1000 °C. A schematic representation of the formation process of Li4Ti5O12 with different anatase crystallite sizes is presented in Figure 4.

2.2.1. One-Step SSR

From a commercial viewpoint, the synthesis of LTO powders via SSR exhibits a potential application due to the simple synthesis route and low synthesis cost. The LTO crystalline phase can be prepared by the one-step SSR method using stoichiometric quantities of starting materials ground together to a fine uniform powder and pressed into pellets fired at high temperatures (>800 °C). Various starting materials are employed [4,7,221]. In method A, the mixture of lithium carbonate and titanium metal is fired at 900 °C, in air, for 20 h but the rutile form of TiO2 is present as an impurity of a few percent:
2Li2CO3 + 5Ti +4O2 → Li4Ti5O12 + 2CO↑.
In method B, lithium hydroxide and titanium dioxide powders are the starting materials. The reaction is carried out at 800 °C in an inert atmosphere:
4LiOH + 5TiO2 → Li4Ti5O12 + 2H2O.
In method C, LTO is synthesized from a mixture of TiO2 and Li2TiO3 intimately ground in an agate mortar and then heated in a platinum crucible in air at temperatures between 800 and 1000 °C:
2Li2TiO3 + 3TiO2 → Li4Ti5O12.
The solid-phase reaction involves cation and anion diffusion. Commonly, the precursors are anatase TiO2 and Li2CO3 (method D). TiO2 and Li2CO3 have melting points of 1640 and 723 °C, respectively. The overall reaction, providing impurity-free LTO, can be expressed by Equation (9):
2Li2CO3 + 5TiO2 → Li4Ti5O12 + 2CO2,
where 16.1% of the weight loss is expected due to CO2 evolution. The diffusivity of Li+ ions being much higher than that of Ti4+ suggests that Li2CO3 is much more active than TiO2 at intermediate temperature ranges. As a result, the formation of LTO is attributed to Li2CO3 reacting with stable TiO2 (see Figure 5) [217]. Duh et al. [187,222] prepared LTO powders by a one-step solid-state reaction using LiCl, H2C2O4, and TiCl4, as the raw materials. First, an appropriate amount of LiCl and 70 wt.% of oxalic acid was mixed together, and then, TiCl4 was added. Mixing of TiCl4 and oxalic acid resulted in the metathesis reaction and an acid mist of HCl was emitted during the process. The precursor was heated at 150 °C for 0.5 h. Next, the precursor was sintered at 400 °C for 3 h in air providing micron-sized particles.
Sorensen et al. [167] prepared titanyl oxalate H2TiO(C2O4)2 in solution as Ti precursor, combining Ti(OH)4 precipitate (formed by 49.6 mL of 2.017 mol L−1 TiOCl2 with 27 mL of aqueous NH4OH) with 0.2 mol of oxalic acid. A mixture (5:4 v/v) of titanyl oxalate solution and lithium acetate was used as a precursor solution for the fabrication of a three-dimensionally ordered microporous LTO material. Wang et al. [223] synthesized a graphitic carbon-coated lithium titanium (LTO/GC) sample by a simple one-step solid-state reaction process with the assistance of sucrose without elevating the sintering temperature. The average grain size of the as-prepared LTO/GC was about 100–200 nm, one order smaller than that of pure LTO prepared similarly. After 300 cycles, the capacity retention was more than 90% at a high rate of 15C.
Several workers stressed the importance of the process of mixing raw materials. Generally, the starting Ti and Li source materials are ground with enough liquid (methanol, ethyl alcohol, acetone) to form a slurry, which is further dried and heat treated to obtain a pure LTO ceramic [43,198]. Growth of LTO powders via SSR using Li2CO3 and anatase TiO2 as starting reagents was reported by Shenouda and Murali [42], who first mixed and grounded the precursor materials in a mortar with some drops of acetone. The powder mixture of samples was dried in an oven at 150 °C for 1/2 h and finally fired at 900 °C for 10 h in air. The resulting particles were crystallized with a spherical shape but SEM images showed a wide particle size distribution (average grain size of ~1.5 µm). Yi et al. [224] used the same procedure but the calcination at 850 °C for 24 h in a flowing air atmosphere yields particle size in the range 400–600 nm. Yang and Gao [225] synthesized LTO powders via a simple solid-state reaction using TiO2-anatase and Li2CO3 (mole ratio of Li:Ti = 4.2:5) as reaction precursors, mixed and heated at 850 °C for 16 h to obtain well-crystallized LTO grains with average size of 480 nm. Kim and Cho [226] fabricated a large quantity of spinel Li4Ti5O12 nanowires (diameter of 150 nm) by firing a mixture of TiO2·1.25H2O nanowires and Li acetates at 800 °C for 3 h. The Brunauer–Emmett–Teller (BET) specific surface area (SSA) of the spinel wire was 38 m2 g−1, which was seven times higher than that of fine particles, which had a SSA of only 4.6 m2 g−1. Wu and co-workers [198] synthesized two kinds of LTO samples via SSR using H2TiO3 and TiO2 as the Ti precursors and Li2CO3. The molar ratio of Li and Ti was 0.84. The process steps were: (1) being initially mixed by grinding in the ethanol for 2 h at room temperature, (2) being dried in the oven at 80 °C for 6 h, and (3) thermal treatment at 850 °C for 16 h in a tubular furnace. Spherical particles with average sizes of about 1.0 and 0.5 µm were grown using the precursor H2TiO3 and TiO2, respectively. The size of TiO2 particles is smaller than that of H2TiO3 particles, which is attributed to the dehydration in the calcination of H2TiO3. Han et al. [227] fabricated 15 kg of LTO to test a mass production line. Here, 4.88 kg coarse Li2CO3 (Dm = 4.50 µm, purity > 99.5%) and anatase-phase TiO2 (Dm = 230 nm, purity > 99.0%) were used as the starting materials, corresponding to a Li/Ti ratio of 4.05/5.00. The mixture was added to 29.91 kg of de-ionized water containing 2 wt.% of the ammonium salt of polycarboxylic acid dispersant. The slurries underwent 6 h of high-energy milling at 1500 rpm using 0.30 mm ZrO2 beads. Then, heat-treated at 800 °C for 3 h in air at a heating rate of 3 °C min−1. The mean particle size was 242 nm. Yao and co-workers [228] synthesized LTO whiskers (~300 nm in diameter) via a solid-state process using TiO2-B and lithium acetate dihydrate as Ti and Li precursors, respectively. TiO2-B was obtained by calcination in a muffle oven at 500 °C in the air for 2 h of an intermediate product, which is the result of the ion-exchange process of hydrated K2Ti2O5 in 0.1 mol L−1 HCl solution. Several researchers investigated the influence of the TiO2 particle size on the properties of Li4Ti5O12 anode [229,230]. Chen et al. [229] proposed a simple approach to synthesize nanostructured LTO with different morphologies (nanorods, hollow spheres, and nanoparticles), in which the TiO2 precursor was first coated with a conductive carbon layer by the chemical vapor deposition (CVD) method, followed by a solid-state reaction with lithium salt. The results indicate that, by employing the carbon pre-coating process, the carbon-coated nanostructured LTO can maintain the initial morphologies of the TiO2 precursors. Table 4 lists the different synthesis conditions for the preparation of LTO anode materials via solid-state reaction.

2.2.2. Two-Step SSR

The mechano-chemical activation process, i.e., ball milling, has been shown to assist SSR synthesis. It allows the intimate mixing and induces the phase reaction of the reactants at room temperature, improving the structural characteristics of LTO particles [46,48,92,238,239,240,241,242,243,244]. Thus, a high compositional homogeneity can be obtained at lower calcination temperatures. Huang et al. [242] showed that, despite first planet ball milling commercial LiCO3 and TiO2 starting materials for 4 h with alcohol as a solvent, the LTO grown via solid-state reaction has a particle size of 0.9 µm. LTO samples were prepared via an SSR process using TiO2 and Li2CO3, successively, together mixed by ball milling using ethanol as a dispersant agent for 5 h, dried at 80 °C, calcined at 800 °C for 12 h, and then ground into fine particles by hand. The displayed particle size was less than 1 µm. Most LTO particles have a spindle shape, like two particles interconnecting together [243]. Becker and co-workers [244] homogenized the mixture of Li2CO3 and anatase TiO2 using a wet ball-milling route with n-pentane at 400 rpm for 1 h. Then, the final LTO white powders were obtained by heat-treatment of the dried mixture in a platinum crucible at 900 °C for 8 h in air. In another experiment, 200 µL of 2-propanol was added as a dispersing agent. Yao et al. [228] synthesized LTO whiskers (300 nm in diameter) from TiO2-B whiskers via a solid-state reaction at 650 °C. The crystal structure of the TiO2-B is easy for lithium diffusion and the process is performed in two separate steps (i.e., diffusion and reaction), which makes it possible to decrease the solid-state reaction temperature down to 650 °C and then maintain the morphologies of whiskers (Figure 6).
A comparison of the electrochemical performance of nanostructured LTO materials synthesized by solid-state reaction with different morphologies is given in Table 5. Lai et al. employed the initial ball milling step to obtain titanate. In a typical experiment, 4.02 g TiOSO4·xH2O and 2.765 g LiOH·H2O were added to the agate tank and then well mixed by ball milling in a planetary ball mill under air atmosphere at a speed of 300 rpm for 4 h [46]. The effects of Li2CO3 particle size (d50 in the range 0.4–4.5 µm) on the properties of LTO were investigated using a ball-milling-assisted solid-state reaction [240]. Then, 31.08 g of Li2CO3 and 80 g of TiO2 (Li:Ti = 0.84) were mechanically dispersed in 200 mL DI water and ground by a conventional ball milling for 1 h at a rotor speed of 500 rpm. The powders were sintered at 800 °C in air for 12 h. Electrochemical tests showed that LTO samples synthesized by fine Li2CO3 particles exhibit better rate capacity and cycle performance. Huang et al. [245] fabricated LTO using a ball-milling-modified solid-state reaction. The precursor of TiO2·xH2O (the hydrolysis product of TBT) coated Li2CoO3 was formed from a TBT, Li2CoO3, HNO3 solution dispersed in an agate jar and well mixed by ball milling performed in air at 400 rpm rotational speed for 10 h. Finally, porous LTO was obtained after calcination at 800 °C for 7 h in air. Natalia et al. [246] prepared an LTO precursor by mixing anatase TiO2 and Li2CO3 at a molar ratio of 5:2 with zirconia balls as crushers. The mixing process was performed in a bill mill at 300 rpm for 2 h and then continued at 480 rpm for 2 h. LTO@CN composite was synthesized via a solid-state process, and the detailed process was described as follows. First, 2.2281 g of Li2CO3, 5.9985 g of TiO2, and 1 g of folic acid were mixed by ball milling in ethyl alcohol for 6 h. Second, the solvent was evaporated by drying the readily prepared mixture at 80 °C for 2 h. Third, the synthesized precursor was heated under an argon atmosphere at 750 °C for 8 h [247]. An alternative way to improve the crystallinity of LTO is the addition of a small amount of salt during synthesis, which is referred to as a salt-assisted solid-state reaction [233]. In this work, TiO2, LiOH, NaCl (2–6 wt.%), and methanol (acting as a dispersant) were mixed in a ball mill for 4 h forming a slurry sintered at 800 °C for 12 h in an O2 atmosphere.
Li et al. [179] studied the effect of precursor size on the morphology of nanosized LTO synthesized by ball-milling/SSR process using axiolitic TiO2 (a-TiO2) as Ti source instead of commercial TiO2. The detailed process using urea as a complexing agent is as follows: a-TiO2, Li2CO3, CO(NH2)2 (molar ratio Ti:Li:urea = 5:4:1), and 10 mL titanium isopropoxide were added into a grinding bowl for 5 h forming a white slurry. After drying at 80 °C in air for 24 h, the obtained precursor was ground fully and calcined at 850 °C in N2 for 12 h. These LTO powders (average particle size of 220 nm) showed excellent capacity of 120 mAh g−1 at 10C even after 20 cycles. Using small-size TiO2 as the precursor significantly decreases the particle size of LTO, which is beneficial for the fast kinetics of Li ions. The solid-state calcination route was also widely used to prepare LTO/C composite. For instance, Hu and co-workers [219] prepared LTO/C composites through a one-step solid-state reaction method using four commonly used organic compounds or organic polymers as carbon sources, i.e., polyacrylate acid (PAA), citric acid (CA), maleic acid (MA), and polyvinyl alcohol (PVA). LiOH·H2O was dissolved into an aqueous solution of PAA with different molar ratios, and powdered TiO2 was added into the PAA-Li solution and stirred for 6 h to yield homogeneous PAA-Li+TiO2 slurries. LTO/C composite with a small particle size of ~300 nm displayed an initial discharge capacity of 161 mAh g−1 and a capacity retention of 95.9% after 50 cycles. Krajewsli et al. [248] performed a three-step solid-state synthesis to produce LTO powders. In the first stage, stoichiometric amounts of Li2CO3 and TiO2 starting reagents were together mixed, ground in an agate mortar, and heated in an alumina crucible to 950 °C and for 10 h in air. In the second step, the obtained powder was ground and heated at 500 °C for 6 h and then at 800 °C for the additional 20 h in air. Impurities such as Li2TiO3 and rutile TiO2 were detected. In the third stage, the sample was mixed with ethanol medium and mechanically ground for 12 h in a planetary ball mill at a rotation speed of 200 rpm. The mixed reactant was evaporated and subsequently dried at 150 °C for a few hours in air. Finally, the powder was ground and heated at 500 °C for 6 h and then at 800 °C for an additional 20 h under air. After the first and second steps, the powders consisted of large crystallites with well-developed surface faceting ~2−3 µm in size, while particles were much smaller (~200−500 nm) after the third ball-milling process.

2.2.3. SSR-Assisted Method

The rapid microwave-assisted solid-state process is a fast and economical synthetic method that yields uniform Li4Ti5O12 nanocrystallites prepared in large quantities. The microwave system can significantly shorten the synthesis time to minutes [251,252]. Compared to the existing solid-state synthetic methods using conventional heating, microwave irradiation prompts a much more rapid solid-state reaction due to the interaction of the microwave field with reactant molecules. In particular, the reaction time and temperature of consecutive steps can be finely programmed in modern microwave systems, facilitating easy and efficient optimization of reaction conditions and scale-up material preparation. In a typical experiment, TiO2 as raw material has an appropriate dielectric loss tangent (tan δ ≈ 1 × 10−4) in the microwave field, which is preferable for energy absorption during microwave-induced synthesis. The starting materials of TiO2 and Li2CO3 were mixed and ball milled in acetone for 10 h and, then, treated in a microwave system (frequency of 2.45 GHz and power between 200 and 1400 W) at 600−900 °C for 30 min. LTO produced at 800 °C has a distinctively narrower particle size distribution without agglomeration, and the particle size ranges from 100 to 350 nm with an average size of 180 nm.

2.2.4. Sintering Process

Several researchers have investigated the effect of sintering, which is a mandatory step to obtain a well-crystallized final product [205,253]. Ohtake investigated the crystallinity of LTO nanoparticles prepared by a two-step sintering through solid-phase synthesis from lithium acetate and anatase TiO2 starting chemicals Mixing of reagents via ball milling at a rotating speed of 320 rpm for 1 h was followed by a pre-sintering at 400, 450, and 500 °C for 1 h in air, and a precursor was formed [253]. Furthermore, the precursor was mixed by the ball milling for 1 h similarly and was sintered at 700 and 750 °C at 10 °C min−1 rate for 1 h in air. The results showed that the pure phase LTO (fine particles of 110 nm) was obtained under sintering at 750 °C subsequently to a pre-sintering temperature of 500 °C via Li2TiO3 (Figure 7). Yuan et al. [204] evaluated the optimal calcination atmospheres for the synthesis of LTO (air, N2, H2/Ar). The oxygen vacancy created due to the reduction of TiO2 in the reducing atmosphere of hydrogen helps to enhance the oxygen ion diffusivity during calcination, whereas carbon coated over TiO2 in the starting material acts as a reducing agent during calcination in the nitrogen atmosphere. Thus, the LTO samples prepared by heating under diluted hydrogen (H2/Ar) and nitrogen atmosphere showed a remarkably higher rate capability and better cycle stability compared to samples prepared by heating in air.

2.3. Sol–Gel Process

Among the wet-chemical techniques, the sol–gel route has gradually attracted the attention of the electrochemistry community as a versatile way to prepare complex intercalation electrodes. LTO has been synthesized through a variety of sol–gel techniques, including various experimental conditions producing a homogeneous distribution of uniform, submicron-size particles with good stoichiometric control [62,63,188,254,255,256,257,258,259,260,261,262,263,264]. This process yields small particle size, homogeneous and narrow size distribution, low temperature, and short periods of calcination. It is based on the hydrolysis and condensation of molecular precursors [265]. In the sol–gel process, a solid phase is formed through the gelation of a colloidal (gel) suspension. Drying of the gel can then give a “dry gel” (xerogel) state and subsequent heat treatment can be used to remove unreacted organic residues, stabilize the gel, densify it, or introduce crystallinity [266]. This method involves the mixing of metal acetates with a complexing agent (chelate) in an aqueous medium. The complexing agent acts as a fuel during the formation process of transition-metal oxide powders, decomposes the homogeneous precipitate of metal complexes at low temperatures, and yields an impurity-free compound. Thus, the control of the growth of LTO crystal is ensured by the carboxylic acid groups present in the complexing agent, which could form chemical bonds with the metal ions to produce a viscous resin (precursor) upon evaporation of the solvent. Chelating agents such as acetic acid (AA) [62,267], citric acid (CA) [263,268,269,270,271,272,273], oxalic acid (OA) [63,187,274], ethylene diamine tetra-acetic acid (EDTA) [263,275], and triethanolamine (TEA) [64,65,276] are often used to reduce the excessive reactivity of metal alkoxides towards water. Zhou et al. [277] demonstrated that EDTA–CA–metal complexing reaction is a non-ion selective process.
Several workers choose the homopolymer polyvinylpyrrolidone (PVP, (C6H9NO)n) as a chelating agent [278,279,280]. When PVP is added to the metal colloidal dispersion, its non-ionic surfactant property makes PVP readily dissolve in water and the imide group (N and O atoms) establishes a strong affinity to a single unit of metal colloid. This imide and metal bonding stabilizes the particle growth and increases the probability of nucleus formation [281]. Further, the sintering of this complex melts PVP and decomposes. Figure 8 shows the typical 3-solution sol–gel preparation of LTO using tetrabutyl titanate and lithium acetate as Ti and Li raw materials and citric acid as a complexing agent.
In 1998, Bach et al. [254] synthesized spinel LTO via a sol–gel process in non-aqueous media (ethanol). They developed the hydrolysis and condensation of metal alkoxides, Ti(OR)4, which are the result of a direct combination of a metal M with an alcohol ROH. As an example, titanium isopropoxide, Ti(OCH(CH3)2)4 is hydrolyzed according to Equation (10):
Ti(OCH(CH3)2)4 + xH2O → Ti(OH)x[OCH(CH3)2]4−x + (CH3)2CHOH,
and a polymerization-condensation reaction occurs by dehydration. The mixture of 0.082 mol titanium isopropoxide with lithium acetate in 75 mL of ethanol solution results in a yellow solution, which becomes more and more viscous, yielding the formation of a white monolithic gel after 1 h. Mani and co-workers [69] explained the nanostructure formation of LTO by a dissolution–crystallization–self-assembly mechanism by controlling the annealing temperature of the gel obtained from titanium isopropoxide sol. Supersaturated anatase (Ti, Li)O2 is formed at 400 °C, whereas anatase and/or rutile interspersed with Li2TiO3 and Li4Ti5O12 are observed between 450 °C and 800 °C. Most sol–gel syntheses are conducted using lithium acetate (CH3COOLi) or lithium hydroxide (LiOH∙H2O) and tetrabutyl titanate (Ti(OC4H9)4) as starting materials, using the molar ratio Li:Ti = 4:5 to form the sol. After the sol is dried at 120 °C for 10 h, a gel is formed and then calcined in N2 atmosphere at 800 °C for 12 h to obtain the final LTO powders [232,261]. Instead of a pure aqueous solution, Luo et al. [261] employed a solution of 17 mL ethanol and 1 mL DI water, in which 3.46 g of glacial acetic acid was added to slow down the hydrolysis reaction of tetrabutyl titanate. Ma and Cheng [282] investigated by single factor method the effects of chelating agent (triethanolamine), raw material ratio, calcination temperature and calcination time on LTO products with pure phase, uniform grain size, 100–200 nm particle size and no obvious agglomeration. Xia and co-workers [283] used a solution of alcohol and propionic acid (alcohol: propionic acid = 2:1), the propionic acid is as a hydrolysis inhibitor in the solution to suppress the tetrabutyl titanate decomposition. Qiu and co-workers [284] synthesized pure LTO particles using a modified sol–gel method with ethyl acetoacetate as a chelating agent and extended this technique for the preparation of La-doped LTO, which can restrain the agglomeration of particles during heat treatment. Mahmoud et al. [191] prepared LTO nanoparticles (360 nm in size) by using the sol–gel procedure with the stoichiometric amounts of TiCl4 diluted in ethanol, LiOH·H2O as reagents, and TEA as a complexing agent with a Ti/TEA molar ratio of 0.8. The best electrochemical performances were obtained for the sample calcined at 900 °C for 1 h using the heating/cooling rate of 2 °C min−1. Impurity-free nanosized LTO powders were synthesized by the sol–gel process using TIP and LiOAc (LI/Ti = 4.5) mixed in 75 mL of ethanol forming a yellow-colored solution, which turned to white color after stirring for 10 min [255]. The LTO nanoparticles obtained by grounding to fine powder and calcination at 800 °C for 12 h exhibited a specific discharge capacity of 242 mAh g−1 at 0.1C rate in the potential range of 0.01–1.75 V vs. Li+/Li. Examples of starting materials used in wet-chemical methods for the synthesis of LTO anodes are given in Table 6.
To mitigate the formation of agglomerates, surfactants may be used during sol–gel synthesis. A surfactant acts as a template for nanoparticle formation. The adsorption of surfactant-like molecules to nucleated nano-crystals lowers the free energy of the surface and, therefore, the reactivity of the particles. The ratio of surfactant to metal precursor can control the size distribution of the nanoparticles [289,290]. Several researchers employed ionic surfactants to control the crystal growth of LTO such as CTAB, sodium dodecyl benzene sulfonate (SDBS), or triblock copolymer PEO)20(PPO)70(PEO)20 (Pluronic, P123) [288,291,292,293,294]. In a typical process, CTAB is dissolved in 100 mL ethanol under magnetic stirring. Four grams of lithium acetate dihydrate was dissolved in the above solution with continued magnetic stirring. Titanium (IV) isopropoxide is added to the above solution dropwise by keeping the mole ratio Li:Ti = 1:1.25. The temperature of the solution is raised to 90 °C and stirred continuously to form the gel. The gel is aged at 100 °C for 24 h in air and the precursor is decomposed at 400 °C for 4 h followed by calcination at 800 °C for 12 h in air. The average discharge capacity of the prepared LTO taken over 20 cycles is ~60 mAh g−1 at a constant current density of 21.37 mA g−1 [292]. Jiang et al. [70] have prepared highly dispersed LTO nanoparticles by the triblock copolymer P123 surfactant-assisted sol–gel process, exhibiting very good rate capability. By using the P123 surfactant, the LTO sample is very loose, with the individual grains being well dispersed and almost without any aggregation. Zhang et al. [288] synthesized mesoporous LTO nanoparticles (~100 nm) via the sol–gel process by employing a nonionic P123 surfactant as the structure-directing agent. In a typical procedure, the Ti-based sol contains 3.0 g P123 and 6 mL HNO3 (65 wt.%) dissolved in 30 mL anhydrous ethanol, subsequently added by 8.8 mL titanate isopropoxide. Wang and co-workers [287] synthesized LTO nanocrystals with an ultra-fine particle size distribution between 120 and 250 nm through the sol–gel method using lauric acid as a surfactant. A white sticky gel was obtained by mixing three alcoholic solutions containing CH3COOLi∙2H2O, Ti(OC4H9)4, and lauric acid, respectively. Li et al. [293] synthesized hierarchical Li4Ti5O12-TiO2 microspheres using an SDBS/CTAB mixed surfactant. When the critical micelle concentration is reached, the self-assembly of anionic and cationic surfactants is observed and intermediates [Ti–O–Ti]n transforms gradually into spherical mixed micelles.
In 2002, Kavan and Grätzel [285] reported the preparation of nanocrystalline LTO particles in a slurry using lithium ethoxide and Ti(IV) alkoxides. The incorporation of polyethylene glycol (PEG) produced a viscous liquid that was deposited using a doctor-blading technique. The thickness is about 2–4 μm and the electrochemical performance is excellent, at least at the beginning of the charging/discharging cycling. The synthesis of double metal alkoxides containing lithium can be produced through a condensation reaction with the elimination of a volatile side product. The simplest case is the reaction of metallic lithium with an alkoxide in a solution of either pure alcohol or alcohol and some organic solvent. This reaction results in the formation of a double metal alkoxide and the evolution of hydrogen gas [295]. Shen et al. [62] reported the first synthesis of Li4Ti5O12 nanocrystalline via the sol–gel method to obtain nano-scale particles (100 nm) using tetrabutyl titanate (Ti(OC4H9)4) and isopropyl alcohol mixed with a mole ratio of 1:60, and lithium acetate dissolved into the mixture solutions of isopropyl alcohol/deionize water/acetic acid. Kuo and Lin [270] fabricated LTO/C composites via a modified one-pot sol–gel synthesis using citric acid as a chelating agent and carbon source. The gel precursor was obtained by mixing two solutions A and B prepared by dissolving CH3COOLi·H2O and Ti(OC4H9)4 in ethanol solution (99 wt.%) with Li:Ti molar ratio of 4:5 with a third ethanol solution containing citric acid. The resultant mixture (ratio of 4:5:1.25) was heated at 80 °C for 6 h and then calcinated by a two-step treatment at 350 °C for 4 h followed by heating at 800 °C for 12 h in an N2 atmosphere. The final product shows uniform particle shape and size distribution (size of ca. 100–300 nm). Rho and Kanamura used two kinds of sols for the synthesis of LTO films prepared by spin coating: (i) 7.65 mL of titanium isopropoxide into 200 mL ethanol/2-propanol with poly(vinylpyrrolidone) (PVP) as solvent and (ii) 0.6 g LiOH dissolved in DI water. The sols mixed for 1 h convert to gel during the spin coating process at 3000 rpm [278]. Hao et al. [65] studied the influence of various complex agents on the electrochemical properties of Li4Ti5O12 anode material. This method involves the mixing of two solutions, which form a transparent gel after slow hydrolysis for several hours: (i) lithium acetate dissolved in the mixture solutions of ethanol and DI water, and (ii) tetrabutyl titanate dissolved in alcohol and mixed with the complexing agent. The compound synthesized by TEA (see Figure 9 showing the thermogravimetry TG-DTA analysis) had the smallest average particle size of 80 nm with homogeneous distribution and, also yielded the best cycling behavior. At a constant current density of 0.5 mA cm−2, it delivered a discharge capacity of 170 mAh g−1 in the first cycle and 150 mAh g−1 after 30 cycles.
Taking advantage of the sol–gel process accompanied by spinodal decomposition, which yields interconnected porous structure with uniform through-pores, Hasegawa et al. [296] prepared a hierarchically porous LTO anode material via the alkoxy-derived titania sol–gel technique. Porous TiO2 gels were treated in LiOH aqueous under relatively mild conditions, giving rise to platy-layered hydrous lithium titanate crystallites on the macroporous surface. The dissolution-reprecipitation leads to the flower-like structures composed of platy crystallites with a thickness of ≈30 nm when calcinated up to 700 °C. By combining the versatile sol–gel process and a hydrothermal reaction, Wang et al. [297] synthesized hierarchical hollow Li4Ti5O12 urchin-like microspheres with an ultra-high specific surface area of over 140 m2·g−1 and a diameter of more than 500 nm. LTO microspheres delivered a high specific capacity of 120 mAh·g−1 at 20C and long cycling stability of <2% decay after 100 cycles.
Erdas et al. [292] produced LTO powders via a sol–gel method, using citric acid as a chelating agent. Then, 5 mmol of titanium isopropoxide was dissolved in 150 mL of 2-propanol to obtain a saturated solution. Then, 4 mmol of lithium nitrate was added with mild stirring. A saturated aqueous citric acid was then added at a molar ratio of 9 mmol. pH of the final solution was kept at 7.0 by adding ammonium hydroxide. Then, excess ammonia and water were removed by heating the solution to 90 °C with vigorous stirring, until a transparent gel was obtained. After the drying process (12 h in an air oven at 120 °C forming a metal citrate precipitation), the precursors were decomposed at 450 °C for 4 h and 850 °C for 7 h in air to eliminate organic contents. The scanning electron microscopy (SEM) image indicates that the sample has a polyhedron morphology composed of subgrains (particle size ranging from 60 to 90 nm), and a uniform particle size distribution.
Figure 9. TG-DTA curves of the LTO precursors prepared by triethanolamine as complexing agent. Three distinct steps of weight loss are observed. The exothermic peak at 317.7 °C is assigned to the combustion of the xerogel precursor and release of CO2 and H2O gases. The second (444.0 °C) and third (549.3 °C) exothermic peaks are due to the continued combustion of the residue carbonate components and the decomposition of the acetate. LTO crystallization is initiated at ~600 °C (particle size of ~80 nm). Reproduced with permission from [65]. Copyright 2007 Elsevier.
Figure 9. TG-DTA curves of the LTO precursors prepared by triethanolamine as complexing agent. Three distinct steps of weight loss are observed. The exothermic peak at 317.7 °C is assigned to the combustion of the xerogel precursor and release of CO2 and H2O gases. The second (444.0 °C) and third (549.3 °C) exothermic peaks are due to the continued combustion of the residue carbonate components and the decomposition of the acetate. LTO crystallization is initiated at ~600 °C (particle size of ~80 nm). Reproduced with permission from [65]. Copyright 2007 Elsevier.
Micromachines 15 00310 g009
Erdas et al. [298] synthesized a Cu–LTO electrode by combining sol–gel process and electroless copper deposition technique. A one-pot co-precipitation method was carried out by Liu et al. [299] for the fabrication of Li4Ti5O12/V2O5 nanocomposites without surfactant. TBT and LiOH∙H2O were dissolved in ethanol and deionized water (at the molar ratio of Li:Ti = 4.1:5), respectively. V2O5 (2 wt.%) was added to LiOH∙H2O solution and stirred for 4 h, then the mixture was poured drop-wise into the TBT ethanol solution and was slowly stirred for 24 h. It is assumed that VO3 ion formation with the dissolution of the added V2O5 restrained aggregation and dramatically reduced the particle size from 500 nm for pristine to 100 nm for V2O5-coated Li4Ti5O12. LTO was also synthesized using a modified and facile sol–gel method with a combined EDTA-citrate as a chelating agent [263,300,301]. The bi-components EDTA and CA were dropped in a mixed metal ion solution (i.e., stoichiometric amounts of TBT and Li2CO3 dissolved in an ethanol-HNO3 preblended solution in the volume ratio of TBT: ethanol:HNO3 of 1:12:0.5) with the mole ratio of total metal ions to EDTA and to CA is 1:1:2. EDTA usually forms four or six five-membered rings with most metal ions while the chelating reaction is fast, complete, and in one single step. As a result, the product is stable and dispersed and the presence of hydroxyl groups in the citrate structure makes the polymerization reaction between CA and EDTA feasible. This method provided particles of size 200–500 nm after a post-treatment at 750 °C for 5 h in air. The LTO nanoparticles showed a network morphology with high dispersion, giving a capacity of 164 and 108 mAh g−1 at 1C and 10C discharge rates, respectively. The result of the cycling performance showed a high-capacity retention of 97% at 1C and 25 °C after 1000 cycles.
The emulsion-gel process is a derivative of the sol–gel method. The gelation proceeds in a water–oil emulsion, so that both hydrolysis and condensation of metal alkoxides take place in a droplet, and spherical particles of metal oxides can be prepared at relatively low temperatures. The preparation of spherical LTO particles (with very narrow particle size (450 nm) distribution) was conducted as follows. The solution of TIP mixed with 180 mL of n-octanol and 0.3 g of hydroxypropyl cellulose (HPC, Mw = 62,500) was stirred with an emulsifier at 40 °C for 15 min. Subsequent dissolution of TIP (0.02–0.12 mol dm−3) in acetonitrile droplets generates emulsion. Then, 7.2 mL of LIOH aqueous solution was poured into the mixed solution to form a gel after stirring for 1 h [89].
Researchers from Sichuan University reported several works related to the sol–gel synthesis of LTO using various chelating agents. In the first report, various initial conditions were studied in order to find the optimal parameters for the synthesis of LTO using an oxalic acid-assisted sol–gel method with Li2CO3 and Ti(OC4H9)4 as starting materials [63]. LTO powders synthesized with an oxalic acid to titanium ratio of R = 1.0 heat-treated at 800 °C for 20 h have a uniform cubic morphology with an average particle size of 200 nm. In a second work [64], they reported a novel sol–gel method with triethanolamine (TEA) as a chelating agent, which exhibited an initial discharge capacity of 168 mAh g−1 and a subsequent charge capacity of 151 mAh g−1. TEA was used as a chelating agent in the sol– gel process. It could complex with Ti4+ to form a stable complex for preventing the excessively rapid hydrolysis of Ti4+. TEA can be used as a surfactant in many reactions to prepare powders with good dispersivity. The combustion heat generated due to the decomposition of the carbonaceous residue produced from TEA makes the nucleation process complete at the early stage of the sol–gel process, thus leading to a narrow size distribution of the LTO powders. They also reported [267] the synthesis of LTO (average particle size of 500 nm) by a sol–gel method assisted with citric acid in an alcoholic solution as a chelating agent and Li2CO3 and tetrabutyl titanate as starting materials. When the molar ratio R of citric acid to total metal ions was 0.5, LTO exhibited an initial discharge capacity of 167 mAh g−1 at 23.5 mA g−1 and a subsequent charge capacity of 151 mAh g−1. Besides these, they further investigated [63] the preparation of LTO via the oxalic acid-assisted sol–gel method, which exhibits 171 mAh g−1 in the first cycle and 150 mAh g−1 after 35 cycles under an optimal synthesis condition at 800 °C for 20 h, while oxalic acid to titanium ratio R = 1.0. Alias et al. [302] used lithium tert-butoxide (0.05 mol) and titanium isopropoxide (0.03 mol) as starting materials, which were together mixed in ethanol and DI water and kept for one week. The residue was harvested and finally sintered at different temperatures (700–1000 °C) and times (1–5 h). Recently, Pershina et al. [286] developed an optimized sol–gel method for the synthesis of single-phase LTO using tetraethoxytitanium (Ti(C2H5O)4, TET) as the Ti precursor and citric acid as a complexing agent. First, the hydrolysis of TET was carried out by stirring for 3 h followed by the dissolution of a white precipitate of metatitanic acid (TiH4O4) with the addition of diluted HNO3 (1:1), which resulted in a transparent titanyl solution. Then, the titanyl solution was added to a solution of Li2CO3 with citric acid and heated at 80 °C for 12 h to form a gel. Then the gel was heated in air to a temperature of ~200 °C and held for 5 h. Upon subsequent heating to 500 °C and, holding for 1 h, all organic compounds are completely decomposed and volatilized. It was found that the phase formation of lithium titanate is accompanied by the formation of an intermediate phase Li2TiO3 enriched in lithium. The final LTO product, consisting of particles 300–650 nm in size, was sintered at 750 °C for 1 h and 800 °C for 5 h in air. Recently, Llaín-Jiménez et al. [303] synthesized LTO powders (~150 nm average crystal size) via a modified sol–gel method using TBT, LiOAc, and CA as a complexing agent. Different samples were obtained, which differed in the addition rate of three sols: sol-1 formed by TBT added to ethyl alcohol (EtOH), sol-2 consisted of LiAs dissolved in EtOH added to sol-1, and sol-3 consisted of citric acid in EtOH. When sol-2 was gradually added to sol-3 within 45 min, pure LTO was obtained, while a quick mixture within 3 min produced an LTO/rutile (1.7%) hybrid. Recently, Kang et al. [178] investigated the impact of titanium precursors on the formation and electrochemical properties of LTO synthesized via sol–gel method. Lithium acetate dihydrate, anatase TiO2 or tetrabutyl titanate (TBT), and citric acid (CA) were used as Li sources, Ti sources, and chelating agents, respectively. In the synthesis, the proper amount of reagents with the Li:Ti:CA molar ratio of 4:5:4 was separately dissolved in 50 mL of 95% ethanol. The gel precursor was formed after heating the precursor solution at 120 °C under continuous stirring for 8 h and dried at 60 °C for 2 days to remove the residual solvent. Finally, well-crystallized LTO powders were obtained by a thermal pre-treatment at 350 °C for 4 h under air atmosphere and calcination at 800 °C for 12 h under air atmosphere. The sample synthetized with a TBT source revealed better particle dispersibility, and its particles were slightly smaller in size. These specific features resulted in better kinetic of Li+ ions during charge transfer reactions. The specific capacity values for both electrodes equal 150 and 63 mAh g−1 for TBT-LTO and 120 and 58 mAh g−1 for TiO2-LTO at C-rates of 1C and 10C, respectively. Table 7 lists the electrochemical performance of nanostructured LTO materials synthesized by the sol–gel route with different morphologies.

2.4. Pechini Process

The Pechini method is based on a gelation reaction between alcohol and acid used as solvents, unlike the sol–gel process in which the metal alkoxide participates in the gel formation reactions. The precursor is a polymeric resin, in which the cations are well distributed, yielding a nanostructured oxide during calcination. Currently, highly cross-linked resins containing a more uniform distribution of the reacting cations are obtained using a polyacrylic acid with higher functionality. Stenina et al. [275] prepared LTO samples with good ionic conductivity using the Pechini process. In this process, titanium tetrabutylate, citric acid, and Li2CO3 were sequentially added to an ethylene glycol + nitric acid mixture (5:2 volume ratio) with constant stirring. The titanium to citric acid molar ratio was 1:4. The resultant mixture was calcined at 70 and 130 °C for 24 h at each temperature. As a result, the ethylene glycol and citric acid molecules underwent polycondensation to give polymer gel, which was then pyrolyzed at 325 °C for 5 h. In the EDTA–citrate process, titanium tetrabutylate and lithium carbonate were dissolved in an ethanol + nitric acid mixture (5:1 volume ratio). An ammoniacal solution of EDTA and citric acid was added to the resultant solution. The molar ratio of metal:EDTA:citric acid was 1:1:2. The resultant mixture was calcined at 70 °C until a gelatinous substance was formed, which was then heat-treated at 225 °C for 5 h. Mo4+ partially substituted spinel Li4Ti5−xMoxO12 was synthesized by the Pechini method using LiOOCCH3·2H2O, Ti[OCH(CH3)2]4, (NH4)6Mo7O24·4H2O, citric acid and ethylene glycol [309].

2.5. Hydrolysis (Wet-Chemical) Method

Hydrolysis is a chemical reaction of the interaction of chemicals with water, leading to the decomposition of both the substance and water. Generally, to obtain the final LTO phase, an intermediate titanium dioxide phase (either rutile or anatase) is prepared by reactions of hydrolysis using titanium tetrachloride (TiCl4), titanium(IV) butoxide (TBOT, Ti(OBu)4), as Ti source. In a typical process, LTO submicron-spheres are obtained using a three-step synthesis: (i) hydrolysis of TiCl4 to form rutile TiO2, (ii) hydrothermal treatment of rutile TiO2 with LiOH to prepare an intermediate phase of LiTi2O4+δ, and (iii) the calcination of LiTi2O4+δ to obtain spinel Li4Ti5O12 [310]. Three major factors, including LiOH concentration, LiOH dosage, and hydrolysis temperature are optimized to obtain a phase-pure LTO [311]. Li et al. [312] have grown a perfect LTO crystal structure by the reaction of LiOH with Ti(OH)4, which is the hydrolysate of Ti(OC4H9)4 (Equation (11)) in the presence of ammonia to form the precipitation of LiTi5O6(OH)9∙6H2O (Equation (12)):
Ti ( OC 4 H 9 ) 4 + 4 H 2 O   N H 3 · H 2 O Ti ( OH ) 4 + 4 CH 3 ( CH 2 ) 3 OH ,
LiOH + 5Ti(OH)4 ⮀ LiTi5O6(OH)9∙6H2O.
Equation (12) approximates a reversible reaction, which indicates an ordered and even process. However, the amount of ammonia must be controlled to completely consume Ti(OH)4, which could be converted into TiO2 with excessive NH3. Recently, Wang et al. [313] prepared submicron-sized LTO particles through a simple hydrolysis method by using lithium acetate dihydrate (CH3COOLi·2H2O) and titanium (IV) butoxide (C16H36O4Ti) as raw materials and showed that sintering conditions significantly affect purity and dispersion of particle sizes. The optimized LTO product calcined at 650 °C for 20 h demonstrates small particle sizes (570 nm) and excellent electrochemical performances. In the voltage range of 1.0 to 3.0 V, it delivers an initial discharge capacity of 242 mAh g−1 and remains at 117 mAh g−1 over 500 cycles at the current density of 60 mA g−1; its discharge capacity remains at 64 mAh g−1 after 2500 cycles at 1.2 A g−1. Liu et al. [152] synthesized novel wave-like spinel LTO nanosheets using a facile “co-hydrolysis” method, which exhibits a capacity of ~150 mAh g−1 at high charge/discharge rates of 8.5 A g−1 (50C). LTO nanosheets with a scale of about 10 nm in thickness and 400–1000 nm in width/length were obtained by dissolving Li metal in 2-dimethylaminoethanol (DMEA) and 1-butanol reagents. Wang et al. [39] prepared monodisperse LTO nanospheres (diameter of 120 nm) via hydrolysis of peroxo-titanium prepared from TiN powders immersed in DI water, hydrogen peroxide and ammonia solution. After stirring for 30 min, a transparent yellow solution of peroxo-titanium complex was obtained. The peroxo-titanium complex was gradually hydrolyzed as described (Equation (13)):
2[Ti(OH)3O2] → 2TiO2 + 2H2O + O2↑ + 2OH,
that is, the addition of OH anions into the precursor solution can retard the hydrolysis of the peroxo-titanium complex, and thus reduce the aggregation of TiO2 and lead to the formation of monodisperse TiO2 nanospheres. The different preparations of the LTO spinel phases by hydrolysis route are summarized in Table 8.

2.6. Mechanochemical Synthesis

Mechanochemical synthesis, i.e., ball milling, has been widely used to synthesize LTO [92,93,170,205,238,316,317,318,319,320,321]. The ball milling technique meets the requirements of simplicity, ease of scale-up, low manufacturing and material cost, and lack of sophisticated apparatus. It is believed that mechanochemical actions such as ball milling generate local high-pressure and high-temperature environments. Therefore, the ball-mill process is always employed to help prevent aggregation and obtain reduced particle size. Different grinding systems are currently used such as attrition, planetary, or vibratory mills. According to the milling speed, we distinguish conventional ball milling (speeds <1000 rpm) and high-energy ball milling (HEBM), which is an efficient method due to its high rotor turning at speeds up to several thousand rpm using small grinding media. The effects of HEBM on the synthesis of LTO were reported by Zaghib et al. [238] and by Wang et al. [92], who studied the influence of the milling time on particle size. TiO2 anatase (~20 nm), LiOH∙H2O, activated carbon (~1993 m2 g−1), and glycerol were mixed at the weight ratio of 1.2:1.0:6.8:28.5 by ball milling and then calcined in air at 800 °C for 12 h. The product with a particle size of ~150 nm delivered a discharge capacity of 110 mAh g−1 at a 10 C rate. Han and co-workers [93] synthesized 100 g of LTO powder using 32.5 g of Li2CO3 (Dm = 4.5 µm) and 86.8 g of anatase TiO2 (Dm = 230 nm), which corresponds to a Li/Ti = 4.05/5.00, mixed with 200 g of DI water after adding 2 wt.% of the ammonium salt of polycarboxylic acid as dispersant (yellowish-brown liquid, viscosity of 85 cps). They controlled the particle sizes under four different milling conditions (i.e., ZrO2 bead sizes: 0.05, 0.10, 0.45, and 5 mm) at 3000 rpm for 3 h. The mean particle size was 146, 175, 215, and 285 nm, respectively. Pure LTO powders (rutile TiO2 and Li2TiO3 free) were obtained with ball sizes of 0.05 and 0.10 mm after calcination at 800 °C for 3 h. Hong et al. [317] investigated the influence of starting materials using either anatase or rutile TiO2, annealing temperatures (e.g., 700, 800, 900 °C), and mechanochemical activations (e.g., ball milling or high-energy milling). The hydrolysis process was performed using 200 mL DI water added to 2 wt.% of the ammonium salt of polycarboxylic acid, as a dispersant. These studies showed that mechanochemical activation by HEBM of the starting materials was more effective in decreasing the reaction temperature and particle size as well as increasing the LTO content of the final powder than those prepared by conventional ball milling. Thus, rutile TiO2 is more desirable in acquiring pure LTO than anatase due to the anatase-to-rutile phase transition. The changes in structure and morphology of starting materials and the effects, on the properties of the final products, have been reported [322]. In a typical experiment, TiO2 precursors are prepared with a HEBM machine at the speed of 2500 rpm using 0.4 mm ZrO2 beads as grinding media. After 1 h of grinding, the TiO2 particle size (D50) decreased from 1.0 to 0.5 µm. Li and co-workers [312] proposed a new LiCO3-ammonia-ball milling system to achieve the hydrolysis and condensation of Ti and Li sources step by step using ammonia as the adjustment and effectively controlling the number of formed crystal grains. The formed hydrate is in a metastable state, which is transformed into a perfect crystal of LTO with a high stability, e.g., spinel phase. The capacity of obtained LTO reaches 175 mAh g−1 at a current rate of 0.5C and keeps nearly 100% capacities after 300 cycles. Jia and co-workers [323] used a high-temperature (700–800 °C) ball milling process to synthesize LTO with different Li and Ti sources. The best product in terms of electrochemical performance is the sample prepared at 700 °C for 3.5 h using Li2CO3 and amorphous TiO2 as raw materials with the molar ratio Li:Ti = 0.82. In a second word, Jia et al. [324] reported that LTO prepared via a high-temperature ball-milling route without pretreatment contains Li2TiO3 impurities even at T = 800 °C. They suggested optimizing the experimental conditions with the assistance of ultrasonic dispersion, mechanical agitation, and glucose monohydrate (C6H12O6·H2O) as a solid phase dispersant was added to raw materials to refine grain, disperse evenly, and avoid agglomerating. The influence of the high-energy ball-milling time on the structure, morphology, and electrochemical performance of synthesized powders was investigated by Michalska et al. [325]. Remarkably, the electrochemical tests performed at high rates showed significantly improved specific capacity of the powders obtained at a constant speed of 200 rpm and extended processing time of 30 h (Figure 10).
Ball milling was also adopted to assist in the preparation of pure phase LTO. The production cost can be reduced by lowering the firing period [46,205,286,319,320,326,327,328,329,330,331,332]. The ball-milling assistance is capable to increase the reactivity of powders by reducing particle size and increasing particle surface areas, which highly facilitates the completeness of the synthesis. Mesoporous LTO was prepared in a planetary ball-milling-assisted solid-state reaction with titanyl sulfate (TiOSO4∙xH2O) and LiOH as precursors [46]. The sample annealed at 500 °C for 2.5 h shows agglomerates of nanoparticles with a size of ~8 nm, predominant mesopores centered at 1.9 and 3.8 nm, and a large specific surface area of 154 m2 g−1. Guerfi et al. [320] demonstrated that a pre-treatment by dry mechanochemical activation of precursors (anatase TiO2 and Li2CO3 ground during 30 and 300 min) is a successful process that reduces the firing time to 1/6 of that with standard mixed precursors (calcination at 800 °C for 1 h). The homogeneous coloration of powders compared to conventional milling indicates an intimate mixing of powders.
In the synthesis of LTO via a ball-milling-assisted sol–gel route, Yan and co-workers [326] evidenced that ball-milling and calcination temperature had a significant effect on the formation of the phase-pure Li4Ti5O12. The sol prepared from mixing 8.2 g of lithium acetate dissolved in 100 mL ethanol with a TiCl4 solution (Li:Ti ratio of 4:5) and citric acid as a chelating agent dissolved in solution B with stirring to form a transparent gelation was calcined in a muffle furnace at 800 °C for 10 h in air to obtain LTO precursors. Further, these products were ball milled and then calcined in the range 600–900 °C for 10 h in air. The precursor without ball milling exhibits two XRD peaks at 2θ = 27.4° and 54.3°, which were attributed to a secondary phase of rutile TiO2, while, the ball-milled sample is a pure LTO phase (lattice parameter a = 8.367 Å). Investigation of the influence of calcination temperature on the formation of LTO showed that the samples obtained by calcination up to 800 °C led to the single-phase spinel type, while impurities of rutile and anatase TiO2 existed in samples treated at temperatures lower than 700 °C. When the calcination temperature is beyond 900 °C, the impurity phase of TiO2 appears again, which is due to the loss of lithium at high temperatures.
Yuan and co-workers [205] synthesized LTO nanoparticles (grain size of 300–400 nm) from stoichiometric TiO2 anatase (~2 μm secondary particles composed of ~20 nm primary particles) and Li2CO3 mixed in liquid ethanol by HEBM. The milling process was conducted in the air using a planetary mill with grinding bowl sizes of 80 mL at a rotational speed of 400 rpm for 1 h. The wet-ball-milling-assisted solid-state reaction was associated with an ultra-high-speed nano-pulverization pretreatment process to grow LTO nanoparticles [245]. In this process, the precursor of TiO2∙xH2O (the hydrolysis product of tetrabutyl titanate) coated Li2CO3 was obtained by ball milling in air at 400 rpm rotational speed for 10 h. Kim et al. [327] reported the synthesis of LTO by a novel sol–gel method with high-energy ball milling of precursor. LTO anode material exhibited the first discharge capacity of 173 mAh g−1 and excellent cycleability. Suzuki et al. [213] demonstrated the mechanochemical-hydrothermal synthesis of layered lithium titanate hydrate, Li1.81H0.19Ti2O5∙xH2O (LHTO) nanotubes with lengths over 300 nm, using planetary ball milling of LiOH and TiO2 with water at room temperature. The LHTO precursor transformed into LTO nanotubes via a thermal treatment at 500 °C for 2 h. Pristine and La-doped LTO powders were fabricated via a simple ball-milling-assisted modified solid-state method [328]. The three-step procedure was conducted as follows: (i) mixture of a solution obtained by dissolving Li2CO3 in alcohol/deionized water (ratio of 4:1) with an alcoholic solution containing tetrabutyl titanate, (ii) drying the slurry mixture and calcination at 800 °C for 7 h, and (iii) dispersion of powders in a solution of alcohol and DI water and ball milling in air at 400 rpm for 10 h using agate-ball-to-powders weight ratio of 3:1. The white slurry was dried to obtain the final powders.
Recently, Pershina et al. [286] optimized the synthesis of LTO powders using a ball-milling-assisted solid-state reaction with Li2CO3 and rutile TiO2 starting materials. Before sintering, the blend was homogenized in an agate mortar or mechanically activated in a wet ball mill with agate balls 10 mm in diameter at a weight ratio of balls:charge of 3:1 at a speed of 750 rpm for 30 min with the addition of ethyl alcohol. Chauque et al. [331] mentioned that high-energy milling increases the specific capacity if crystalline degree remains high. Zhang et al. [332] mixed and active the LTO precursors by mechanical ball milling using ethanol as dispersant agent for 12 h (rate 330 rpm, ball feed ratio 16:1).
Jang et al. obtained LTO/CNT nanocomposites by complexing bulk LTO particles and carbon nanotubes (CNTs) via mechanofusion. In addition, the surface of the LTO/CNT material was easily fluorinated by physically mixing LTO/CNT with NH4F followed by heating at 300 °C for 2 h in a nitrogen (N2) atmosphere. The optimized surface fluorinated MF-LTO/CNT was obtained by mixing MF-LTO/CNT and NH4F in a mass ratio of 95:5, in which case the capacity reached ~170 mAh g−1 at 0.2 C, and ~140 mAh g−1 at a high rate of 20 C. At 10C, this composite showed a capacity of 128.5 mAh g−1 after 500 cycles. [329]. For comparison of different syntheses of LTO/CNT, we can mention the work of Ye et al. who synthesized composite anode materials with LTO nanoparticles uniformly chained by CNTs by a facile sol–gel process followed by high-temperature calcination step [330]. The best rate capability was obtained for 11% CNTs/LTO. At 0.2C, this composite delivered a capacity of 147.2 mAh g−1, 89.7% of which was retained at 30C. It was coupled with a superactive carbon (AC) cathode to assemble a Li-ion capacitor (LIC), with the mass ratio of AC cathode to 11% CNTs/LTO set at about 2.5. An energy density of 58.6 Wh kg−1 was achieved at a power density of 214.2 W kg−1, and an energy density of 35 Wh kg−1 was maintained even at a high-power density of 7434 W kg−1. This outperforms the properties of many other LTO-anode-based LICs.
Table 9 lists the electrochemical performances of nanostructured LTO materials synthesized by mechano-chemical method with different morphologies. Recently, Akintola and co-workers [288] synthesized LTO nanoparticles in a planetary ball mill with agate balls at a speed of 250 rpm using TBT, LiOAc as reagents, and a triblock copolymer P123 as the surfactant. This method provided well-defined-edge nanoparticles (average size of 350 nm) with a specific surface area of 55.7 m2 g−1 and good dispersion and uniformity.

2.7. Molten-Salt Synthesis

Molten-salt synthesis is a flux growth method, which has been reported to be one of the simplest means to prepare multi-component oxides [80,333,334,335,336,337]. The molten salt method was preferred since it does not need any complicated apparatus for material synthesis. It is also cost effective and no organic additives are needed. The method is characterized by an accelerated reaction rate and controllable particle morphology. The most important advantage is the prevention of particle agglomeration due to the diffusion of molten salts. Molten salts can provide a liquid reaction environment when serving as reaction media, thus accelerating the crystal formation reaction. Cheng et al. [333] synthesized nanosized LTO using LiCl only as a high-temperature flux and adopted this material for a coin-type cell and hybrid capacitor. However, the as-synthesized spinel LTO can deliver a capacity over 150 mAh g−1 (at a current rate of 40 mA g−1 between 1.0 V and 3.0 V) provided that the molar ratio LICl:TiO2 = 16. Zr- and Si-doped LTO microcrystals were synthesized via the molten-salt method using LiOH∙H2O and TiO2 as the starting precursor and LiCl and KCl, which act as fluxes. The stoichiometric quantities of LiOH·H2O and TiO2 were ground well for ½ h. Then the flux LiCl and KCl in the molar ratio of 50:50 was added to the above mixture and mixed for another half an hour to obtain a homogeneous powder. The resultant powder was then heated at 800 °C for 10 h with a heating rate of 10 °C min−1 to obtain the desired product. The particle size histogram infers that the maximum number of particles is lying in the size range of 1.75 μm [334]. Bai and co-workers [80] adopted a low content of LiCl–KCl (the molar ratio of LiCl–KCl to TiO2 is fixed to just 2 and the molar ratio of LiCl:KCl varies from 0.25 to 4.0) to prepare LTO particles. For LiCl:KCl = 1.5, the LTO particles are uniformly distributed (D50 = 3.4 µm) and have a well-developed octahedral shape. This material achieves high initial discharge capacity (169 mAh g−1), charge–discharge efficiency (94%) at 0.2C rate, and good rate performances from 0.2 to 5C.
Pure LTO particles with an average size of 1.3 µm were prepared by using the LiCl-KCl molten-salt method using anatase TiO2 (2 g, 100 nm in average particle diameter), LiOH·H2O (0.85 g), LiCl (10.2 g) and KCl (12.46 g) [336]. Rahman et al. [337] synthesized LTO–TiO2 composites via the molten salt method with minimization of the water content in the starting material mixture for the molten salt (LiNO3 -LiOH∙H2O). The products yield a high capacity (166 mAh g−1 at 0.5C) and good rate capability (110 mAh g−1 at 10C). The excellent performance was partly related to the presence of high grain boundary density among the particles. Nithya et al. [334,335] prepared several LTO anode materials via a single-step molten-salt synthesis using LiCl-KCl as a flux. Specific discharge capacities of 240, 265, 305, and 200 mAh g−1 were delivered by pristine LTO, Li4Ti4.95Mn0.05O12, Li4Ti4.9Mn0.1O4, and 5%Si-95%LTO composite at a current density of 0.1 mA cm−2. Sharmila et al. [338] reported the manufacturing of LTO via a single-step molten salt using LiOH∙H2O and TiO2 as the starting compounds and LiCl–KCl as a flux. The mixture heated at 800 °C for 10 h in air yields a uniform morphology displaying 3D polyhedral shapes with smooth surfaces and particle size of ~1.5–2 µm. A nanocrystalline Li4Ti5O12–TiO2 duplex phase was synthesized by a simple basic molten salt process (BMSP) using a eutectic mixture of LiNO3–LiOH–Li2O2 at 400–500 °C. The sample prepared by heat-treating at 300 °C for 3 h revealed dense agglomerates of ultra-fine nanocrystalline LTO; with heat treatment at 400 °C for 3 h, there was a duplex crystallite size (fine < 10 nm, and coarse > 20 nm) of Li4Ti5O12–TiO2; at 500 °C for 3 h, a much coarser and less-dense distribution of lithium titanate (crystallite size ∼15–30 nm) [339].

2.8. Solution-Combustion Method

The solution-combustion method is a facile and economical technique for the preparation of nanomaterials for lithium-ion batteries [340]. Combustion synthesis implies heating a precursor to a relatively low process temperature, after which the system generates the necessary energy for complete conversion and crystallization to the desired oxide. It has the advantage of relatively simple equipment, formation of high-purity products with size and shape control, and stabilization of metastable phases. The initial reaction medium is an aqueous or nonaqueous solution, the reactants are oxidizers (nitrates) and organic compounds as fuels (glycine, urea, α-amino-acid L-alanine, lactic acid, etc.), which are used as the source of C and H to liberate heat through combustion. Complexes formed with the metal ions could facilitate the homogeneous mixing of cations in solution. The combustion energy that is released upon heating the sample depends on the amounts of fuel and oxidizer in the precursor. Prakash et al. [97] reported a solution-combustion synthesis of LTO nanopowders using titanyl nitrate [TiO(NO3)2] and LiNO3 as precursors. In a typical reaction, which is completed in a few seconds, an aqueous redox mixture containing a stoichiometric amount of titanyl nitrate (0.0362 mol), 2 g of LiNO3 (0.0289 mol), and 4.22 g of glycine (0.0562 mol) were taken in a 120 mL alumina crucible and placed into a muffle furnace preheated to 800 °C. Raja et al. [103] reported an aqueous combustion process using titanium(IV) methoxide (Ti(OCH3)4) and Li2CO3 as reagents, and a common α-amino-acid L-alanine as fuel. Ti(OCH3)4 was dissolved in nitric acid and an aqueous solution of LiNO3, which yielded LTO nanoparticles of uniform morphology with an average particle size in the range of 40–80 nm. Yuan et al. [101,217] studied the mechanism of LTO synthesis using a cellulose-assisted glycine-nitrate combustion process. High-purity and well-crystallized LTO spinels were obtained at a calcination temperature of 750 °C via the sequence, for which the cellulose was first adsorbed by the mixed solution of LiNO3 and glycine followed by the impregnation of TiO2 suspension. Temperature-programmed oxidation experiments demonstrate that cellulose thermal pyrolysis creates a reducing atmosphere, which may facilitate oxygen-ion diffusion. In the combustion process performed by De Sloovere and co-workers [100], the Ti precursor was prepared using lactic acid added to the hydrated titanium hydroxide in a 3:1 molar ratio in aqueous solution. Then, the mixture was refluxed at 80 °C for complete dissolution and adjusted at a pH of ~6.8 by the addition of ammonia (35%). Yuan et al. [99] reported a glycine-nitrate combustion process. LTO calcined at 700 °C showed the best electrochemical performance, which reached a specific capacity of 125 mAh g−1 at a 10C rate with fairly stable cycling performance even at 40 °C. A similar method was used to prepare the La-incorporated Li4Ti5O12 with different La to Ti molar ratios via a one-pot cellulose-assisted combustion method [341]. Intermediate steps include the fabrication of a titanyl nitrate solution (transparent bright yellow) from the mixture of the precipitate TiO(OH)2 with nitric acid and the absorption of the mixed solution into dried activated cotton fibers. LTO nanopowders were prepared by solution−combustion synthesis using a titanyl nitrate [TiO(NO3)2] aqueous solution and LiNO3 as the oxidant precursors and glycine as the fuel, with stoichiometric amounts of titanyl nitrate (0.0362 mol), 2.1 g of LiNO3 (0.0303 mol), and 3.00 g of glycine (0.400 mol) [342]. Electrochemical performances of nanostructured LTO materials synthesized by the combustion method are provided in Table 10.
The samples prepared via the glycine-assisted combustion method show micrometer-sized secondary particles composed of nanometer-sized primary particles, with uniform sizes in the range of 50–100 nm. Similarly, N-doped LTO nanopowders were prepared using urea (2.64–5.35 g) as fuel. HRTEM image showed that a thin shell with a thickness of approximately 5 nm was formed on the surface of the LTO nanocrystals from the thermal decomposition of urea. The nitrogen content of the shell was 8.5 wt.%, which is considerably greater than the average mass ratio (1.1 wt.%) of nitrogen in the LTO material. Chang et al. [280] synthesized LTO anode materials by a gel-combustion method with PVP as the polymer chelating agent and fuel followed by a post-annealing at 800 °C. Excellent reversible capacities of 167, and 144 mAh g−1 were achieved at the current densities of 0.5C and 5C. Yuan et al. [101] used a cellulose-assisted glycine-nitrate combustion process for a production of high-purity and well-crystallized LTO oxides. Particles of 0.2–0.8 µm were obtained after calcination at 750 °C via the sequence, for which the cellulose was first adsorbed by the mixed solution of LiNO3 and glycine followed by the impregnation of TiO2 suspension.

2.9. Hydrothermal Method

Hydrothermal reaction (HTR) is a wet-chemical process using water as a solvent [342,343]. It is an efficient and facile strategy for controlling the chemical composition, and morphology of nanostructured materials (particle shape, and crystallite size) because it can decrease the activation energy of precursors in a simple and inexpensive way, so LTO can be obtained at 450–550 °C approximately. Nevertheless, the conventional hydrothermal process involves a very complicated reaction route and a long hydrothermal process (5–24 h) to synthesize LTO [344,345,346,347,348,349,350,351,352,353,354,355,356]. For example, LTO nanotubes were obtained as white precipitates from a mixture of titanium n-butoxide (TNBT), cetyltrimethyl ammonium chloride 25 wt.% aqueous solution (CTAC), and lithium hydroxide hydrothermally treated at 120 °C for 24 h [54]. However, the choice of precursors is crucial. It is difficult to avoid the formation of impurity phases like TiO2 and Li2TiO3 during the synthesis of LTO. The main reason may be the gradual deviation of the Ti:Li ratio from the desired proportion under a prolonged heating time at high temperatures, even when using different types of nanosized TiO2. This is due to the loss of the lithium salts during the heat treatment process and the less stability of most titanium salts in air. An accurate Ti:Li ratio can be controlled with the assistance of a solvent, which has the excellent ability to form hydrogen bonds with other compounds [342]. A soft synthetic route to obtain nano-LTO is realized by the introduction of organic additives such as surfactants (cetyltrimethylammonium bromide (CTAB), polyethylene glycol (PEG)), biomolecule (urea, vitamin C), ionic liquids, organic acids (citric, oxalic, EDTA), and alcoholic solvents (glycol, ethanol, glycerol). During the synthesis process, organic additives mainly act as the shape controller, reducing agent, and hydrolyzation inhibitor. Li et al. [344] reported that, in the hydrothermal process, the introduction of CTAB as a surfactant significantly improves the rate performance of LTO anode. The LTO/C composite (carbon layer with a thickness of ~1 nm) prepared in the presence of CTAB shows a larger diffusion coefficient of lithium ions (6.82 × 10−12 cm2 s−1) and smaller charge-transfer resistance (Rct) (19.2 Ω) than those of the composite (1.22 × 10−13 cm2 s−1 and 50.2 Ω) free from CTAB in the preparation. LTO nanowires were obtained from H2Ti3O7 prepared powders (0.5 g) through hydrothermal treatment in a LiOH aqueous solution (50 mL) at 100 °C for 24 h, in which lithiated titanate was formed via hydrothermal ionic exchange [132]. After washing the precipitates with ethanol until a pH of 7 was reached, the precipitates were dried at 60 °C and then calcined and heated at 800 °C for 6 h in air to obtain a pure LTO phase. Wu and co-workers studied the effects of the reactant concentration and heat treatment temperature on the phase structure of LTO nanosheets synthesized via hydrothermal method using Ti(OC4H9)4 and LiOH as the raw materials [345]. The best LTO nanostructure was obtained using 2 mol L−1 LiOH and heat treatment at 550 °C for 6 h. This material can deliver a specific capacity of 150 mAh g−1 at a 20C rate when the cut-off voltage is from 0.8 to 2.5 V. The best LTO nanostructure was obtained using 2 mol L−1 LiOH and heat treatment at 550 °C for 6 h. This material can deliver a specific capacity of 150 mAh g−1 at a 20C rate when the cut-off voltage is from 0.8 to 2.5 V.
Wang et al. embedded LTO nanoparticles in a reduced graphene oxide (rGO) conductive network by an in situ electrostatic self-assembly effect using a simple hydrothermal reduction method. The LTO nanoparticles were combined with rGO nanosheets by a Ti–O–C covalent bond, and rGO not only increased the conductivity but also prevented the agglomeration of LTO. In a half cell, this anode showed an outstanding rate capability, with a capacity maintained at ~272 mAh g−1 after the 1000th cycle at 10C when tested in a half battery [357]. For comparison, LTO nano-particles having a size in the range of 20–100 nm, wrapped and distributed uniformly inside the rGO sheets delivered a capacity of 141 mAh g−1 at 10C with a capacity retention rate as high as 97.2% after 1000 cycles [358].
Hierarchical LTO nanoparticles were also prepared using a multi-step hydrothermal synthesis [150,352,359]. Ding et al. [352] fabricated a hierarchical yolk-shell Li4Ti5O12-SnO2 structure by controlling the reaction temperature of the hydrothermal process at 130 °C for 12 h. This composite maintained a specific capacity of 253 mAh g−1 after 200 cycles at a 1C rate. Wang et al. synthesized a nanoflower-like composite by coating a 3D conductive rGO and SnO2 nanoflower on LTO via an in situ electrostatic self-assembly and hydrothermal reduction process. [360]. Here, both the 3D rGO network and the doping of Sn ions on the LTO lattice improved the electron transport capability of LTO. In addition, SnO2 enhanced the mobility of Li in the interlayer of LTO/rGO/SnO2. Consequently, the LTO/rGO/SnO2 electrode showed remarkable rate capability, delivering a capacity of 445 mAh g−1 at a rate of 20C after 1000 cycles. DFT calculations showed excellent structural stability through the formation of Ti-O-C bond, Sn-O-C, and Sn-O-Ti bonds. The full cell with this anode and LiFePO4 anode has a high energy density of 130.56 Wh kg−1 with a high-power density of 2868.5 W kg−1. Zhu et al. used hydrothermal, freeze-drying, and calcination methods to embed SnO2 in carbon-coated LTO nanoparticles. At 0.2 Ag−1, this porous layered SnO2-Li4Ti5O12@C composite showed a capacity of 958 mAhg−1 after 300 cycles, and 766 mAhg−1 after 1000 cycles [361]. Sha et al. [151] investigated the phase development of LTO nanoplates prepared through a two-step hydrothermal method using either benzyl alcohol-NH3·H2O (BN) or ethanol-NH3·H2O (EN) as the Ti-source hydrolysis solvent. Ammonia in the solution increased the pH value of the solvent to facilitate Li+ diffusion and dissolution into TiO2·nH2O. In the benzyl alcohol–ammonia solution Li1.81H0.19Ti2O5·nH2O was successfully formed after a short hydrothermal treatment time of 30 min. The LTO-BN sample hydrothermally treated at 180 °C for 6 h showed superior electrochemical performance. The discharge capacity reached as high as 175 and 157 mAh g−1 at 1C and 40C, respectively; after charge–discharge cycling at 40C rate for 1000 times, capacity retention of 92.5% was still maintained with a decay rate of 0.0078% per cycle. Ultrathin Li4Ti5O12 nanosheet (6.6 ± 0.25 nm thick, specific surface area 178 m2 g−1) based hierarchical microspheres were synthesized through a three-step hydrothermal procedure [353]. First, TiO2 powder (2.0 g, P25) was dispersed in NaOH solution (80 mL, 10 mol L−1) by ultrasonic treatment. The obtained dispersion, i.e., sodium titanate (NaTO), was transferred into a 100 mL Teflon-lined stainless-steel autoclave and maintained at 120 °C for 24 h, and then the NaTO NWs were separated by centrifugation. Second, NaTO NWs (0.2 g) were dispersed in NaOH solution (38.5 mL, 2 mol L−1), and then, H2O2 (1.5 mL, 30%) was added into the solution. The solution was transferred into a 50 mL Teflon-lined stainless-steel autoclave and maintained at 150 °C for 12 h to prepare Na2Ti3O7. Third, the Na2Ti3O7 was added into HNO3 (0.05 mol L−1) solution, stirring for 12 h to exchange the hydrogen ions for Na+ ions, repeated two times, and the H2Ti3O7 intermediate phase (HTO) was achieved. Finally, the HTO was dispersed in LiOH solution (40 mL, 0.3 mol L−1) by ultrasonic treatment and the solution was transferred into a 50 mL Teflon-lined stainless-steel autoclave and maintained at 120 °C for 24 h. The products were separated by centrifugation, washed three times with water, and dried in an oven. After calcination at 400 °C for 4 h, the LiHTO-NS samples were achieved. Hu and co-workers [128] used the same hydrothermal/ion exchange process for the fabrication of surface-fluorinated LTO nanowires/reduced graphene oxide (F-LTO/rGO) composite. The fluorination was performed by an additional hydrothermal treatment of LTO/rGO nanowires in 3 mol L−1 NH4F solution at 160 °C for 6 h followed by a calcination at 500 °C for 4 h under an Ar/H2 atmosphere. The material exhibits nanowire morphology with a diameter of 50–100 nm and a length of ~10–20 µm. Hierarchical LTO porous microspheres with open structures of 4–6 µm diameters were prepared via a two-step hydrothermal process. The products consisted of a bunch of willow leaf-like nanosheets formed of LTO nanoparticles of 20 nm in size [359]. First, TiOx·H2O microspheres were prepared by HTR (at 150 °C for 24 h) of 1 mL tetrabutyl titanate dropped into 50 mL acetic acid. A second HTR (at 150 °C for 12 h) was used to prepare the final LTO product from 0.27 g TiOx·H2O microspheres mixed with 1.75 g of LiOH·H2O dissolved in 50 mL de-ionized water. Further sintering at 750 °C for 5 h in an H2/Ar atmosphere (10:90 in volume) was performed to obtain LTO microspheres; Similarly, LTO@C composites were prepared with the addition of glucose in the second HTR step (see Figure 11).
Nano-sized LTO pore microspheres (primary particles with a size between 30 and 40 nm) were prepared by hydrothermal treatment of Ti-(OC4H9)4 and LiAc·2H2O as raw materials and subsequent calcination at a high temperature of 1050 °C for 1 h, which provided big secondary particles (agglomerates) of 1.5–2.0 µm in diameter [362]. The nano-LTO exhibits a specific capacity of 133 mAh g−1 at a 10C rate and even 80 mAh g−1 at 60C. Porous LTO microspheres were also hydrothermally fabricated by Shen and co-workers [55]. Nano/microspherical superstructures, with sizes of ca. 4 µm in diameter, exhibited rich hierarchical pores and a specific surface area of 57.5 m2 g−1. At the high rate of 20C, the specific charge capacity was still ~92 mAh g−1, and less than 4.8% discharge capacity loss was observed after 200 cycles. Kim et al. [363] reported on Li4Ti5O12 microspheres assembled from nanosheets and synthesized by mixing Ti(IV)isopropoxide, LiOH, and H2O2 at 130 °C for 20 h under hydrothermal conditions; these nanosheets evidence a stable cycling performance, with high-capacity retention (140 mAh g−1) at a current density of 10 A g−1 even after 4000 cycles. Wu et al. [364] synthesized petal-like Li4Ti5O12–TiO2 nanosheets by boiling a lithium titanium peroxide–ammonium solution in an oil bath followed by a low-temperature, short-duration solid-state calcination. The resulting nanosheets give rise to excellent performance with a remarkably high discharge capacity of ~112 mAh g−1 at 20C rate and retain 97.1% of their initial capacity after 100 cycles. “Flower-like” motifs of Li4Ti5O12 consisting of thin petal-like component nanosheets were synthesized by using a facile and large-scale hydrothermal process involving unique Ti foil precursors followed by a short, relatively low-temperature calcination (500 °C for 3 h) in air [192]. The as-prepared LTO anode materials delivered capacities of 141 and 60 mAh g−1 at 10C and 100C, respectively, and capacity retention of 87 mAh g−1 at 20C rate after 300 cycles. Chen et al. [365] reported that a small amount of rGO (only 1.2 wt.%) greatly improves the whole morphology and electrochemical performance of the composite. LTO was prepared by hydrothermal method using 5 mmol TBT dissolved into 25 mL ethylene glycol mixed with 1 mL acetic acid, 0.5 g CTAB, and 2 mg rGO. After complete dissolution, this mixture was added to 0.336 g LiOH·H2O dispersed into 20 mL DI water and hydrothermally treated at 180 °C for 24 h. The nanoparticles uniformly grow on the rGO nanosheets effectively suppressing the agglomeration and enhancing the specific surface area. This anode material delivers high rates discharge capacity of 128 mAh g−1 at 80 C (discharge-charge time only 33s) and retains 50% of its initial capacity after 2000 cycles at 80C.
Several workers showed that the particle growth can be controlled through the hydrothermal route with the addition of glycerol, ethanol, or ethyl alcohol because of their excellent ability to form hydrogen bonds with starting materials [366,367,368]. In a typical ethyl-alcohol-assisted hydrothermal synthesis, 0.05 mol L−1 tetrabutyl titanate was dispersed in 17 mL ethyl alcohol forming a pale-yellow solution. Then, the 20.5 mL of 2 mol L−1 LiOH·H2O solution was added dropwise into this solution to form a white suspension with strong stirring for 15 min and then transferred to a 100 mL stainless-steel autoclave, which was maintained at 180 °C for 34 h to form a white precipitate [368]. Zhang et al. [354] added 2 mol L−1 LiOH in an aqueous solution to another solution formed by the mixture of tetrabutyl titanate and ethyl alcohol (volume ratio of 1:1) to form a white suspension further hydrothermally treated at 180 °C for 24 h. Glycerol has been used to assist the formation of various inorganic nano-materials in hydrothermal synthesis, because of its unique solvent properties. During the synthesis process, glycerol mainly assists in the dispersion of raw materials and acts as the shape controller, reducing agent, and hydrolyzation inhibitor. In a typical procedure of a glycerol-assisted hydrothermal process, Ti(SO4)2 (4 mmol) is dissolved in the solution formed by deionized water (15 mL) and glycerol (5 mL) and mixed with a fixed amount of LiOH∙H2O dissolved in deionized water (20 mL). Ti(SO4)2 is selected as the titanium source due to its good solubility and stability in solution [367]. During the hydrothermal process, glycerol is added as a modifier to assist the dispersion of starting materials, which can help to control the morphology of LTO. The glycerol-assisted samples have a cubic shape and the particle size is relatively uniform at 15–20 nm with a specific surface area of 86 m2 g−1 when calcined at 500 °C for 5 h in air.
The 3D nano/micro hierarchical structure, which combines the merits of nanosized building blocks with the benefits of thermodynamically stable microsized assemblies, represents an ideal host for Li storage. For instance, Tang et al. [146] prepared flower-like LTO nanosheets using 20 mL hot ethylene glycol (90 °C) with 10 mmol titanium tetraisopropoxide and 2 mL ammonia (25–28 wt.%) mixed in 8 mmol LiOH solution (30 mL H2O). After hydrothermal reaction at 170 °C for 36 h, the final product was obtained by calcination at 500 °C for 2 h. These structures exhibit a high reversible capacity and an excellent rate capability of ~166 mAh g−1 at 8C. Tang et al. [51] prepared hollow structured LTO with a shell consisting of nanosheets prepared from amorphous hydrous titanium oxide (AHTO). In a typical process to fabricate an LTO hollow structure, 4 mmol spherical AHTO colloids were dispersed in 20 mL, 0.2 mol L−1 LiOH solution. After stirring for 10 min, the suspension was transferred into a 30 mL Teflon-lined stainless-steel autoclave and then placed in an oven at 180 °C. Figure 12 presents the TEM images of samples obtained for different hydrothermal reaction times (0–10 h). TEM micrographs clearly demonstrate the evolution of structure from solid microspheres to hollow microspheres and finally to separated nanosheets. XRD patterns (Figure 12g) evidence the composition of products from different hydrothermal reaction stages and after calcination. TEM micrographs clearly demonstrate the evolution of structure from solid microspheres to hollow microspheres and finally to separated nanosheets. The AHTO precursor can transform to pure anatase TiO2 after calcination at 500 °C. With the reaction proceeding, the diffraction peaks of anatase TiO2 become weak, while the diffraction peaks of spinel LTO emerge after a 3 h reaction.
Hierarchical mesoporous LTO, with a nest-like structure (thickness of the wall of 50–80 nm), the tiny crystallite size of 8.2 nm, the very large specific surface area of 219 m2 g−1 and pore volume of 0.509 cm3 g−1, were prepared by a hydrothermal reaction, using CTAB as the surfactant, and [NH4+]4[H+]2[Ti4(C2H2O3)4(C2H3O3)2(O2)4O2]6− water-soluble titanium complex as the titanium precursor [292]. Typically, the water-soluble titanium complex was prepared as follows: 2 mmol titanium metal powder was added to an ice-cooled aqueous solution consisting of 8 mL 30% H2O2 and 2 mL 28% NH3. After 3 h, 3 mmol glycolic acid was added, and this solution was heated at 80 ◦C with stirring for 6 h to remove excess H2O2 and NH3 and form a gel-like orange precursor. The 20 mL aqueous solution containing the Ti precursor, 0.3 mol L−1 CTAB, and 0.5 mol L−1 LiOH was stirred for 1 h, and hydrothermally treated into a Teflon-lined stainless autoclave maintained at 180 °C for 12 h. The products show remarkable electrochemical performance (i.e., low polarization and high specific capacities of 135 and 113 mAh g−1 at 2.5 and 10 A g−1, respectively, after 200 cycles). Zhang et al. [369] elaborated a versatile hydrothermal method for the synthesis of hierarchically structured Li4Ti5O12–TiO2 (LTO-TiO2) composites, by which the LTO and TiO2 phases were adjustable in the final products. With control over the time of the hydrothermal reaction at 18 h, an appropriate amount of anatase TiO2 (46.8%) can be obtained, and it possesses a uniform carambola-like framework assembled by numerous ultrathin nanosheets, which enables a relatively large specific surface area along with abundant interlayer channels to favor electrolyte penetration. Typically, 10 mL H2O2 aqueous solution (30%) was added to 150 mL of 0.5 mol L−1 LiOH solution. Then, 4.14 g tetrabutyl titanate was added to the solution. The H2O2 worked as a chelating agent, and could also slow down the hydrolysis of TBT. The hydrothermal reaction was carried out at 130 °C for 18 h (providing a light-yellow precipitate), then washed, dried at 80 °C and; finally, sintered at 500 °C for 3 h in air. Specifically, this anode material delivered a discharge capacity of 91.2 mAh g−1 at a very high current rate of 40C. Using the same synthesis process but prolonging the hydrothermal reaction time to 24 h, an urchin-like morphology was obtained (Figure 13).
Many researchers have optimized the synthesis of LTO using hydrothermal processes combined with other techniques such as microwave [351,370,371], sol–gel [372,373], dispersion [367], and ball milling [374]. Ultrathin LTO nanosheets with ordered microstructures were prepared via a polyether-assisted hydrothermal process [212]. The polyether (Pluronic P123) can impede the growth of Li2TiO3 in the precursor and also act as a structure-directing agent to facilitate the (Li1.81H0.19)Ti2O5·2H2O precursor to form the LTO nanosheets with the ordered microstructure. The synthesis was carried out as follows. First, a suspension was prepared: (i) a yellow solution was formed by TBT added into 15 mL of dehydrated ethanol, (ii) a solution was obtained by P123 dissolved in another 5 mL of ethanol, (iii) the dropwise mixing of solutions under continuous stirring for 4 h, (iv) finally, another 5 mL of LiOH (3.075 mmol) in a water solution was slowly added, where the stoichiometric ratio of Li/Ti was set as 4.1/5 with 2.5% excess Li. Then, the hydrothermal reaction was conducted at 180 °C for 18 h, and the solid product was calcined at 600 °C for 3 h in air. The ultrathin LTO nanosheets showed a rate capability much higher than that of the LTO sample without P123 in a Li battery with over 130 mAh g−1 of capacity remaining at the 64C rate. Hui et al. [371] showed that LTO prepared via microwave-assisted hydrothermal (MWHT) reaction using TiO2 and LiOH as raw materials exhibit interconnected nanoparticles (~40 nm in diameter) with mesoporous morphology (8 nm in pore size). MWHT-prepared LTO (at 170 °C for 2 h) delivered a specific capacity of 90 mAh g−1 at 60C (~10 A g−1), corresponding to a time of 60 s for full charge, which is superior to that of LTO synthesized via HTR at 160 °C for 48 h. Shi et al. proposed an advanced microwave-hydrothermal (MW-HT) method for the preparation of LTO microspheres composed of nanoflakes wrapped in graphene nanosheets [351]. First, 8 mmol LiOH was dissolved in 20 mL deionized water, 1 mL 30% H2O2 was then added to the solution, and finally, 2 mmol titanium isopropoxide was added to the solution and stirred for 20 min. The as-prepared solution was transferred to a polytetrafluoroethylene (PTFE)-lined autoclave, which was sealed for microwave irradiation. The LTO precursor, obtained after treatment at 150 °C for 15 min, was dissolved in 100 mL DI water with 0.25 mL 70% hydrazine hydrate and 0.1 g GO after 2 h of ultrasonic dispersion time to form the LTO/graphene composite. A sol–gel precursor and hydrothermal methods are used for the synthesis of LTO nanoparticles (100–200 nm) [279]. First, 2.38 g C2H3O2Li∙2H2O was dissolved in 10 mL of an anhydrous alcohol (Solution A). Then, 8 mL tetrabutyl titanate was dissolved in 30 mL of isopropyl alcohol (99.7%) (Solution B). Each solution was vigorously stirred at room temperature for 30 min, and solution A was then dumped into solution B under constant stirring. The suspension was placed inside a stainless-steel Teflon-lined reaction vessel (60 mL) heated at 160 °C for 24 h. Finally, the prepared sample was dried and calcined at 800 °C for 10 h under an Ar atmosphere. Recently, a novel self-supported LTO nanobelt array was fabricated on the titanium foil by a two-step hydrothermal process and subsequent ion exchange [145,165].
Ti foil can act as both the Ti precursor for LTO growth and a current collector. First, titanium foils in a mixed NaF/NaOH solution were hydrothermally treated at 180 °C for 24 h resulting in Na0.98H1.02Ti4O9·3.3H2O, which was transformed to H2Ti5O11·3H2O after protonation in 0.1 mol L−1 HCl solution, indicating that Na+ in sodium titanate can be completely replaced by H+. The second hydrothermal process with 0.2 mol L−1 LiOH produced the precursor (Li1.81,H0.19)Ti2O5·2H2O. Finally, the precursor calcined at 450 °C for 1 h in air to form LTO nanobelts. Wu et al. [148] explored the formation of ultrathin dual-phase nanosheets consisting of alternating spinel LTO and rutile-TiO2 lamellas through a facile and scalable hydrothermal method. The white suspension appeared immediately once the LiOH·H2O aqueous solution was slowly dropped into the Ti(OC4H9)4 ethanol solution and further hydrothermally treated at 180 °C for 24 h and calcined at 600 °C for 6 h in a horizontal tube furnace in air. The thickness of constituent lamellas can be controlled exactly by adjusting the mole ratio of Li:Ti in the original reactants. The best electrochemical performance, i.e., the stable discharge capacity of 123 mAh g−1 at a current density of 50C after 500 cycles, was obtained with Li:Ti = 4.1:5. LTO nanorods with sizes of 100–200 nm in diameter and 1–2 µm in length were prepared via a hydrothermal process of TiO2-B nanorods with LiOH in aqueous solution [124]. The final product heat treated at 500 °C for 4 h exhibits a high specific discharge capacity of 101 mAh g−1 after 1000 cycles at a 20C rate.
A double-hydrothermal method was performed to prepare LTO nanobelt arrays. First, stamped titanium foils were hydrothermally treated in NaOH and NaF solution at 180 °C for 24 h giving the product Na0·98H1·02Ti4O9·3.3H2O, which was subsequently transformed into H2Ti5O11·3H2O by protonation in acid. Next, lithiation was obtained by a second hydrothermal reaction resulting in (Li1.81,H0.19)Ti2O5·2H2O, which was finally transformed to pure spinel Li4Ti5O12 after calcining [159]. Nanoparticle-stacked Li4Ti5O12-TiO2 nanowire arrays were constructed via a hydrothermal/ion exchange process [375]. Films obtained by hydrothermal treatment of Ti foils in NaOH+NaCl solution were immersed in a diluted HCl solution for several hours to thoroughly exchange the Na+ with H+. The resultant H2Ti2O5·H2O arrays were placed in a 2 mol L−1 LiOH·H2O solution at 60 °C for 10 h and calcined in N2 at 600 °C for 3 h. This electrode material achieved a specific capacity of ~200 mAh g−1 at a current rate of 0.75C (potential range of 1.0–3.0 V). Huang et al. [14] fabricated pure LTO and LTO-rutile TiO2 (LTO-RTO) composite via HTS by adjusting the amount of TBT as a Ti source. LTO-RTO was obtained using 16.25 g of TBT dissolved in 34 mL alcohol and mixed with 1.68 g of LiOH·H2O dissolved in 20 mL of distilled water. The gained white suspension was transferred into a 100 mL stainless-steel autoclave and then reacted at 180 °C for 36 h. After filtrating and drying at 80 °C for 7 h, the precipitate was calcined at 600 °C for 10h and transferred into final product. By changing the amount of TBT to 15.10 g, the pure LTO was obtained using a similar procedure. The Li4Ti5O12/C composite with lump morphology and excellent rate performance was synthesized using a facile hydrothermal process, in which CTAB was introduced as a surfactant significantly improving the rate performance. The specific capacities of the obtained composite at charge and discharge rates of 0.1, 1, 5, 10, and 20C are 176, 163, 156, 151, and 136 mAh g−1, respectively [344]. Electrochemical performances of nanostructured LTO materials synthesized by hydrothermal reaction are provided in Table 11.

2.10. Supercritical Synthesis

Supercritical water (SCW) has low viscosity, high diffusivity, and zero surface tension. The enhancement of the reaction rate and high supersaturation based on the nucleation theory due to lowering the solubility is another motivation for the use of SCW. These synergetic effects have been used to prepare LTO powders with high—rate capability using supercritical hydrothermal synthesis in batch process. In practice, the reaction media to prepare LTO free of TiO2 and Li2TiO3 impurities is SCW, Tc = 374 °C, Pc = 221 bar [377,378,379]. The effects of various synthesis conditions—feed concentration, reaction time, and calcination—on the particle properties, including particle size, surface area, particle morphology, phase purity, and crystallinity, were carefully analyzed [379]. In this work, each of the Ti-sol and LiOH solutions was fed forward to a reactor by high-pressure pumps X-500 at a flow rate of 5 g min−1, and these two streams were mixed. On the other hand, distilled water was fed by another high-pressure pump at a flow rate of 40 g min−1 and heated to an appropriate temperature by an electrical furnace. Then, the reactant mixture of Ti sol and LiOH was mixed with the SCW. Phase-pure LTO particles are obtained with a long reaction time of 6 h in supercritical water at 400 °C and 300 bar without subsequent calcination, while the anatase TiO2 impurity phase is detected at shorter reaction times of 5 min to 2 h. Nugroho et al. [377] synthesized LTO by reacting lithium hydroxide monohydrate and titanium(IV) isopropoxide (TIP) in supercritical water using a tube reactor of 11 mL inner volume. Free-impurity quasi-spherical LTO nanoparticles were obtained with an SSA of 38 m2 g−1 and a size of 50–100 nm. The SCW LTO products show a high initial discharge capacity of 212 mAh g−1 at a 1C rate in the potential range of 0.1–3.0 V vs. Li+/Li and a capacity retention of 91% after 50 cycles. Further, the same group of researchers fabricated, in a very short time, hierarchical mesoporous LTO microspheres in supercritical methanol [380]. The suspension of lithium hydroxide and TIP in methanol was transferred in a tube reactor maintained at 400 °C, in which the pressure increased to 30 MPa in 1 min. Hayashi et al. [379] reported the synthesis of LTO nanoparticles via supercritical water flow hydrothermal reaction method using TiO2-sols and Ti(SO4)2 as the Ti source. The syntheses were carried out at temperatures from 350 to 410 °C, at pressure of 30 MPa, and for residence times from 10 to 26 s. Particles of BET-specific surface area of 82 m2 g−1 are in the range of several 10 nm to 50 nm and particle size did not change during the annealing process up to 400 °C. Li4Ti5O12 was crystallized directly via a continuous flow supercritical synthesis using lithium ethoxide and TIP as reactants. Crystalline nanoparticles are obtained in a single step and in less than one minute, by mixing the reactants with superheated water in a continuous flow reactor at near- and supercritical conditions. The LTO nanoparticles have an average crystallite size of 4.5 nm with a specific surface area of ≥230 m2 g−1 [381].

2.11. Solvothermal Method

Unlike the hydrothermal technique, non-aqueous solutions are used in the solvothermal synthesis, which allows for easy control of the morphology, particle size, and homogeneity of the products. Since a variety of organic solvents, e.g., ethanol, benzyl alcohol, polyol, etc., with high boiling points can be selected, the temperature can be elevated much higher than that in the hydrothermal method [59,60,382,383,384,385,386,387]. For example, LTO nanosheets [209], nanoflakes [384], sawtooth-like nanosheets [385], and nanoflowers [146] fabricated by solvothermal method exhibit good electrochemical performance. The solvothermal reaction can be performed by 0.1 mol tetrabutyl titanate dissolved in 100 mL ethanol and then followed by the addition of 0.1 or 0.08 mol LiOH. The undissolved LiOH particles were suspended in the orange solution. The alcoholic suspension was solvothermally treated at 140 °C for 24 h in a 150 mL autoclave under pressure of ~747 kPa. The direct solvothermal synthesis of the LTO nanocrystals was achieved through the self-control of the basicity using a two-phase system of water–ethanol and toluene–oleic acid. The Li–Ti–O precursor was produced under relatively high pH conditions (pH > 11) in the initial stage. The spinel phase was then formed through delithiation and dehydration from the precursor by lowering the pH of the system (pH < 10) [386]. Feckl and co-workers [387] created ultrasmall building blocks for a porous spinel framework in the absence of water, using a solvothermal reaction in tert-butanol, which is an excellent reaction medium for the synthesis of ultrasmall crystals at a low temperature of 170 °C. Yu et al. [203] introduced a one-pot template-free solvothermal synthesis of crystalline Li4Ti5O12 nanostructures based on the “benzyl alcohol route”. Spherical particles (1–2 µm size) are constituted of nanocrystals in the size range of a few nm. Microwave wet chemistry not only reduces the chemical reaction times by several orders of magnitude but also offers high reaction yields and reproducibility of synthesis protocols due to the suppression of uncontrolled side reactions [189,388]. Nowack et al. [189] presented a novel, rapid, and low-temperature synthesis protocol to achieve battery-grade LTO microspheres using a solvothermal microwave-assisted method. The solution of metallic lithium dissolved in benzyl alcohol and titanium isopropoxide (Li:Ti = 0.8) was treated in a microwave reactor at 260 °C for a period of less than 2 min. Products are spherical micron-sized particles with an average size of 4.4 µm, which were electrochemically tested at 60 °C, delivering a specific capacity of 136 mAh g−1 at 10C rate. Several researchers combined the solvothermal method with other techniques such as sol–gel and ball milling [372]. Recently, Gangaja et al. [389] synthesized Li4Ti5O12 nanoplates by an off-stoichiometric solvothermal process, which leads to the coexistence of phase-separated crystalline nanoparticles of Li4Ti5O12 and TiO2 exhibiting reasonable high-rate performances. An ultrahigh-charging-rate capability of up to 1200C (60 mAh g−1; discharge limited to 100C) was demonstrated.
Santhoshkumar et al. [390] report a homogeneous composite of integrated LTO-TiO2 (LTO-TO) nano-hybrid synthesis by a solvothermal technique followed by a calcination process. CTAB was used not only as a precursor but also as a surfactant to optimize the composition. The optimized hybrid showed a capacity of 468 and 83 mAh g−1 at current densities100 to 8 A g−1, respectively. After 300 cycles, at a current density of 5 A g−1, the capacity was still 198 mAh g−1. In this work, the calcination temperature was 600 °C. A lower temperature is recommended to avoid the transition from the phase TiO2(B) to the anatase phase [391]. The best results with LTO-TO were obtained by Lu et al. who fabricated TiO2(B) nanosheets on the surface of the LTO microsphere by using a solvothermal method to obtain LTO-TiO2(B) (LTO-TOB) composite after calcination at 300 °C for 2 h. This anode demonstrated superior rate capability (145, 139, and 124 mAh g−1 at 20C, 30C, and 50C, respectively) and superior cycling stability (98.9% capability retention after 500 cycles at 5C) [392]. Meso-structured, micron-sized LTO/C dyad spheres with high surface area were successfully synthesized via solvothermal treatment of tetrabutyl titanate and lithium acetate in the presence of furfural (heterocyclic aldehyde, CHO) allowing for the formation and co-assembly of carbon and LTO crystallites into the interpenetrating nanoarchitecture [60]. In the first step, TiO2 crystallites are formed by a controlled slow nucleation process via a solvothermal reaction of 1.2 mL TBT, 1.0 mL furfural, and 2.5 g LiOAc dissolved in 60 mL ethanol. Afterward, the chemical lithiation of TiO2 and carbonization of furfural completed simultaneously under solvothermal conditions result in the formation and co-assembly of micrometer-sized L-T-O-C intermediate phase. Finally, the L-T-O-C network is converted to mesoporous LTO-C microspheres by a short post-annealing at 550 °C for 2 h in the N2 atmosphere (Figure 14).
Li4Ti5O12/TiO2 (LTO/TO) nanoparticles were synthesized via a one-step solvothermal method followed by further heat treating using LiOH·H2O and TBT as raw materials [384]. Typically, 0.2 g LiOH·H2O was completely dissolved in 50 mL of ethanol after ultrasonication for 30 min as solution A. Then, 4 mL of TBT was dissolved in 10 mL of ethanol as solution B. Solution B was then slowly dropped into solution A with magnetic stirring. The mixed solution was transferred into a 100 mL Teflon-lined stainless-steel autoclave and reacted at 200 °C for 24 h. After natural cooling to ambient temperature, the precursor product was collected by centrifugation, washed with ethanol and ultrapure water several times, and dried overnight at 80 °C. Finally, the LTO/TO sample was obtained via calcining at 600 °C for 3 h in air. Though the same procedure, the pure LTO phase can be synthesized by dissolving 2 mL of TBT in 10 mL of ethanol to form solution B. Electrochemical performances of nanostructured LTO materials synthesized by the solvothermal method are provided in Table 12.
Figure 14. (a) Schematic representation of the solvothermal synthesis of mesoporous LTO-C microspheres. (be) Characterization of mesoporous LTO microspheres. (b) XRD pattern, (c,d) FE-SEM images, and (e) N2 adsorption/desorption isotherm (inset: BJH distribution). Reproduced with permission from [60]. Copyright 2011 The Royal Society of Chemistry.
Figure 14. (a) Schematic representation of the solvothermal synthesis of mesoporous LTO-C microspheres. (be) Characterization of mesoporous LTO microspheres. (b) XRD pattern, (c,d) FE-SEM images, and (e) N2 adsorption/desorption isotherm (inset: BJH distribution). Reproduced with permission from [60]. Copyright 2011 The Royal Society of Chemistry.
Micromachines 15 00310 g014

2.12. Reflux Method

The solution-based reflux method is a green synthesis method used to grow nanostructured inorganic materials. Reflux involves heating the chemical reaction for a specific amount of time, while continually cooling the vapor produced back into liquid form, using a condenser. The vapors produced above the reaction continually undergo condensation, returning to the flask as a condensate [90,393,394]. In 2004, Singhai and Skandan [393] patented a new method to prepare nanostructured LTO powders using a reflux-type synthesis. The process utilized: (i) nanoparticles of TiO2 (~20–25 nm), (ii) a lithium salt, and (iii) an organic solvent such as hexanol (CH3(CH2)5OH) with a boiling point of 158 °C. The solution was refluxed for 15 h (heated to a temperature and at a pressure to facilitate the diffusion of Li ions into the TiO2 nanoparticles). Annealing at 500–800 °C in an O2 atmosphere provides ultrafine powders of LTO with particles of average size of ~500 nm, composed of nanocrystallites with an average size of approximately 30 nm. The effect of the solvent on the phase purity was investigated. Unreacted TiO2 material and other unidentified impurities were identified when LiOH·H2O and nano-TiO2 (or Li2CO3 and nano-TiO2) were refluxed in water. In another attempt, pure LTO particles were obtained by refluxing the TiO2/LiNO3/methanol mixture for approximately 15 h. Kim et al. [90] used a reflux method to obtain LTO nanoparticles (6–8 nm in size), which were prepared from ethylene glycol solution of titanium tetra-isopropoxide (Ti(O-iPr)4) and excess Li2O2 by refluxing at 197 °C for 12 h. The obtained particles were filtered and dried at 100 °C for 12 h, as a post-reaction step, and the dried powder samples further heated. Post-treated at 500 °C for 3 h, the LTO anode exhibited the initial capacity of 320 mAh g−1 at a current density of 0.05 mA·cm−2. Yagi and co-workers [141] reported the growth of LTO nanotubes. First, a colloidal solution of anatase TiO2 in 10 mol L−1 NaOH hydrothermally treated at 150 °C for 24 h resulted in Na+-intercalated hydrogen titanate nanotubes (TNTs). Then, Na+ ions in the interlayer of the synthesized nanotubes were exchanged with Li+ ions by a reflux treatment using 100 mL of a 2.5 mol L−1 LiOH aqueous solution, leading LTO nanotubes (diameter and length of ~ 5 nm and several submicrons, respectively) when post-annealed at 500 °C for 2 h, while cubic powders 10 − 50 nm were obtained at 700 °C. Lin et al. [180] prepared monodispersed mesoporous LTO submicrospheres by the solvothermal process using a water–ethanol (60 vol%) solvent followed by a calcination process at 600 °C, revealing a large sphere size of 660 ± 30 nm with a small primary particle size of 20–100 nm, a large specific surface area of 15.5 m2 g−1, an appropriate pore size of 4.5 nm and an ultra-high tap density of 1.62 g cm−3. They displayed a charge capacity of 179 mAh g−1 at a C/2 rate between 1.0 and 2.5 V vs. Li+/Li and an initial Coulombic efficiency of 93.5%.
Yin et al. obtained in situ self-assembly of LTO submicron spheres by simple one-step liquid phase deposition using a low-cost inexpensive titanium source (H2TiO3). The sample calcined with 4 mL of 25% NH3‧H2O at 700 °C showed the best performance. Its capacity remained at 176.4 mAh g−1 over 100 cycles at 87.5 mA g−1, 136.7 mAh g−1 over 500 cycles at 875, and 105.6 mAh g−1 over 1000 cycles at a current density of 1750 mA g−1 with a retention rate of 96.7% [395].

2.13. Templating Method

Utilization of a template may help the formation of porous structures in materials. Various templates have been used to synthesize LTO composites that resolve the predicament of agglomerates [110,396,397,398]. The synthesis of LTO nanocrystals from solution-phase precursors using a carbon-templated growth process achieves complete crystal growth (all of the nutrients are consumed), and the surfaces are annealed at high temperature at nearly the same rate at which the template burns off [85,86,87,88,397]. Monodisperse Li4Ti5O12 hollow spheres were prepared by using carbon spheres as templates [85]. Scanning electron microscopy images show hollow spheres that have an average outer diameter of 1.0 μm and an average wall thickness of 60 nm.
Lv and co-workers [154] implemented a method involving the in-situ oxidation and gasification of the phenol-formaldehyde (PF) resin as a “template” to prepare porous LTO with a mesopore size of 2–20 nm. TiO2-rutile, Li2CO3, and water-soluble phenol–formaldehyde (10 wt.%) resin (solid content 60%) were used as starting materials. Firstly, PF resin was dissolved in the solution of de-ionized water and alcohol, and then TiO2 and Li2CO3 (at the molar ratio of Li:Ti = 4.2:5) were mixed, respectively, in the solution under magnetic stirring. After 0.5 h stirring, the dark red solution was treated at 65 °C to remove solvents gradually. The mixed slurry was dried at 120 °C in the vacuum drying oven to solidify the PF resin. Then, after ball milling, the precursors were heat-treated at 800 °C for 8 h under an air atmosphere. Alginic acid aquagel was used as a template and carbon source in the one-pot solid-state synthesis of Li4Ti5O12/C nanocomposites [249]. The final product was obtained by the pyrolysis at 800 °C for 3 h under argon flow (50 mL min−1) of the mixture of 470 mg of TiO2, 505 mg of LiOAc, and the resulting aquagel formed by 200 mg of alginic acid gelled in water (10 mL) by heating at 90 °C for 2.5 h.
Polystyrene (PS) beads of 400 nm diameter were used to prepare the template for the LTO inverse hemispheric structure, which was fabricated on a Pt/Ti/SiO2/Si substrate by the sol–gel and dip coating method. A coating solution prepared using precursor sources was dropped on the template-deposited substrates, which were then calcinated at 400 °C [398]. Mesoporous LTO thin-film electrodes with open pores of ~18 nm in diameter were prepared by a soft-templating method derived from sol–gel [188]. Thin films were produced via dip-coating on polar substrates, using a mixture of 333 mg of Ti(OBut)4 in 0.5 mL of dry EtOH and 0.15 mL of glacial acetic acid combined with both 52 mg of Li(CH3COO) and 40 mg of a structure-directing agent (poly(ethylene-co-butylene)-block-poly(ethylene oxide), KLE)) dissolved in 0.5 mL of dry ethanol and 1.0 mL of 2-methoxyethanol. Hermawan et al. [68] synthesized porous LTO samples by sol–gel method using an eggshell membrane as a soft template. Lithium nitrate and titanium tetrachloride as precursors were dissolved separately in DI water and concentrated HCl (37%), respectively, and then added by ammonia solution to form colloidal particles, which were attached to the as-prepared eggshell membrane immersed into the solution for 12 h. Hao et al. [397] applied a solution-based method starting from alkoxide precursors of lithium and titanium to obtain atomic-level mixing from which to synthesize LTO. Nanocrystals (with a high surface area of 27 m2 g−1 and concomitant small particle size of 58 nm) were formed from a templated synthesis using this precursor solution and carbon (acetylene) black (composed of ~70 nm particles) (Figure 15). LTO templated electrodes demonstrated reversible cycling storage ~160 mAh g−1 at a 10C rate and maintaining a capacity >150 mAh g−1 even at 100C.
Polyisoprene-block-poly(ethylene oxide) (PI-b-PEO) with an sp2-hybridized carbon-containing hydrophobic block was employed as a structure-directing agent as well as a carbon source for the synthesis of a mesostructured LTO/C composite [155]. The inorganic precursor solution was prepared by mixing 1.287 mL of titanium tetra-isopropoxide and 0.181g of lithium ethoxide (3.48 mL of 1 mol L−1 LiOC2H5 in tetrahydrofuran stock solution) and added to 0.39 g of oxalic acid. The LTO/C composite was formed after heat-treatment of the mixture at 700 °C under N2 and held for 2 h. Xia and co-workers [283] used butterfly wings as biotemplates, which resulted in a 3D porous architecture with a periodic network structure formed by the interconnected “ridge” and “strut”. Liu et al. [159] adopted a template-hydrolyzation plus glucose decomposition method to prepare LTO/C nanotube arrays grown directly from TiO2 nanotubes on stainless-steel foil. The precursor of TiO2 nanotube arrays was fabricated by a one-step deposition and in situ etching solution route. In a typical synthesis, the as-synthesized ZnO nanorod arrays on stainless steel foil were immersed in an aqueous solution consisting of 0.075 mol L−1 (NH4)2TiF6 and 0.2 mol L−1 H3BO3 at room temperature for 40 min. In this solution, (NH4)2TiF6 hydrolyzed to TiO2 on the surface of individual ZnO nanorods template, while ZnO dissolved simultaneously in the solution with acids produced by (NH4)2TiF6 hydrolysis. The final LTO product obtained by a chemical lithiation route possesses much shorter ion diffusion paths because the electrolyte both inside and outside the nanotubes could simultaneously diffuse leading to accelerated reaction kinetics. The ZnO nanorod array was also used as the sacrificial template for the fabrication of LTO nanotube arrays (TiCl4 and H2O as the Ti and O precursors) via an ADL-lithiation process using 100 mL of 3 mol L−1 LiOH aqueous solution kept at 80 °C for 1 h [399]. The ZnO template was simultaneously etched by the LiOH during the lithiation process. A high specific capacity of 130 mAh g−1 was delivered at 40C after 1600 cycles. In a typical synthesis of nanotube arrays, the ZnO nanorods template was fabricated via a chemical bath deposition consisting of a reaction solution Zn(NO3)2 and ammonia in DI water kept at 95 °C for 5 h. Singh et al. [400] proposed a facile and low-cost templating method using carbonate salts creating 3D interconnected ionic pathways that improve the ionic charge transport without compromising the electrode density significantly. The method was demonstrated for C/Li4Ti5O12 electrode material resulting in excellent capacity retention reaching ~90% at 5C and ~50% at 200C rate combined with high active material electrode densities around 1.45 g cm3. Electrochemical performances of nanostructured LTO materials synthesized by the templating method are listed in Table 13.

2.14. Spray Drying

The spray drying (SD) method is particularly attractive because it has good reproducibility and is easily up-scalable to the industrial level [401]. To demonstrate the scalability of the template-based spray drying approach, Nowack et al. [402] fabricated LTO particles in a pilot-scale spray dryer capable of a production rate of 4 kg h−1. SDM is used to produce powders from aqueous and/or organic-aqueous solutions or suspensions. The most common solvent for solutions is water. Alcohols are also used, either pure or mixed with water. SD involves the atomization of a liquid in a chamber, in which a hot gas recirculates. As the nebulized droplets are exposed to a very large surface with the drying fluid, the drying reaction is almost instantaneous. Two of the important parameters to control the process are the output temperature and the relative humidity after the cyclone collector (Figure 16) [403]. Numerous works report the SD synthesis of LTO [111,112,404,405,406,407,408,409,410,411,412,413,414,415,416,417,418,419,420,421,422,423,424,425,426,427,428,429]. In an early work, Hsiao et al. [409] synthesized porous and dense LTO powders by spray drying followed by solid-state calcination. The spray drying precursor was slurries of LiOH, anatase TiO2 (21 nm particle size), and 5 wt.% (relative to TiO2) BYK-190 dispersant (solution of a high molecular weight block copolymer with pigment affinic groups) in DI water. After ball milling at a speed of 300 rpm for 8 h, the homogeneous slurries were atomized at 250 °C using a two-fluid nozzle with an atomizing pressure of 3 kg cm−2. The final crystalline product was obtained by solid-state calcination in air at 850 °C for 8 h.
In a typical synthesis, LTO powders are prepared by heating at 850 °C for 2 h under air a precursor obtained by spray-drying an aqueous suspension of TiO2 and LiOH∙H2O taking a stoichiometric ratio Li:Ti = 4:5. The spray-drying experiments are conducted using a rotating injector with inlet temperature 190 °C, outlet temperature 110 °C, feed rate 25 mL min−1 and air pressure of 3 bars [430]. Yuan et al. [431] constructed three-dimensional (3D) crumpled graphene sheets wrapped in nano-Li4Ti5O12 (LTO@GS) composites using a one-pot spray-drying assisted solid-phase reaction method with anatase TiO2 as Ti source and LiNO3 (3% mole excess) as Li source. The GO suspension (1 mg L−1) with a mass ratio of LTO to GO of 80:20 was sonicated for 1 h and spray-dried at 200 °C. The specific capacitance of this LTO@GS composite as an anode in a hybrid battery-capacitor BatCap system still retains 90% of the initial value after 20,000 cycles. The LTO@GS//AC BatCap system showed a specific energy of 29.2 Wh kg−1 at a power density of 58.4 W kg−1. Wu et al. [432] developed a low-cost procedure for the synthesis of Cr-doped LTO microspheres (1–5 µm in diameter, BET SSA of 9.39 m2 g−1), which involves a versatile solution-based method, spray drying, which enables ion-level mixing of liquid components. An ammonia solution was used to maintain a pH value of ~10. Zhang et al. [420] synthesized a Li4Ti5O12/rGO composite by spray-drying and annealing processes using aqueous suspensions of LTO and GO vigorously stirred together to form a slurry with LTO:GO = 50:1 by weight. De-ionized water was added to the mixture to adjust the solid content to 2 wt.%. Subsequently, the slurry was stirred, and ultrasonically exposed for 30 min, which was then spray dried at 200 °C to form a solid LTO/GO composite. The composite was heated to 400 °C at a rate of 10 °C·min−1 and sintered at that temperature for 5 h under an Ar atmosphere to form the active LTO/rGO anode material. The LTO primary particles aggregated in micro-sized spherical secondary particles, which were wrapped homogeneously and loosely with a rGO network. This anode material showed excellent rate performance and high cyclic stability; after 300 cycles at 20C, the specific capacity was still 100 mAh g−1. Jia et al. [433] synthesized nanostructured Li4Ti5O12/carbon nanotube (LTO/CNT) composite particles using an aerosol spray drying process followed by thermal annealing. The electrodes delivered a high reversible capacity of 108 mAh g−1 at an extremely high rate of 100C. Moreover, ultralong cycling stability was attained through 8000 rapid charge-discharge cycles with 89% capacity retention. Mesoporous spherical LTO/TiO2 nanocomposite was prepared by aerosol spraying hydrolysis using TBT (1.0 mL), LiNO3 (4/5 by Li/Ti molar ratio), triblock copolymer P123 (0.25 g), and ice-acetic acid (HAc, 0.5 mL) dissolved in ethanol (20 mL). The solution was transferred into a medical-use ultrasonic humidifier (1.7 MHz, 35 W) for the aerosol-spraying process. The final product was obtained by calcining its precursor at 600 °C for 8 h in air. The mesoporous composite had a specific surface area of 43.18 m2 g−1, an average pore diameter at ca. 2.4 nm, and delivered a capacity of 111 mAh g−1 at a 20C rate [434]. Hole-rich Li4Ti5O12/CNTs composites were synthesized by spray drying of titanium peroxide solution and carbon nanotubes as additives in the precursor solution, subsequently followed by calcination at high temperature in air [423]. The SD solution was formed by Li2CO3 and titanium(IV) isopropoxide (TIP) in stoichiometric amounts (molar ratio of Ti/Li = 0.86) dissolved in ethanol and further added to a second solution formed by deionized water containing oxalic acid (OA) with mole ratios TIPP:OA of 3:1. After mixing with CNTs and ultrasonic sonication for 5 h, the obtained sol was subsequently spray-dried by hot air at 230 °C, followed by sintering at 800 °C for 12 h in air. The as-synthesized hole-rich LTO hybrid exhibited an initial discharge capacity of 145 mAh g−1 at a 2C rate and decreased gradually with continuous cycling, retaining 138 mAh g−1 after 200 cycles.
The preparation of the spray-drying aqueous solution proposed by Park et al. [435] involved a mixture of ternary precursor materials: Li2CO3 and rutile-structured TiO2 (~21 nm) with Li:Ti molar ratio 4.1:5 were dissolved into distilled water with glucose (C6H12O6). A polymeric dispersant, i.e., gum Arabic, was mixed into the solution via an ultrasonic bath for 30 min to form a relatively homogenous mixture. The dispersed solution (~15 mL) was sprayed onto a titanium plate, which maintained a uniform surface temperature of 150 °C and evaporated the solvent instantaneously via a controllable hot plate (PC-420D-230, Corning) and a compressor (Monster comp001, China). The Li4Ti5O12/C composite powders were obtained by a spray-drying process, collected, and calcined at 800 °C for 2 h in a furnace purged with Ar gas.
To overcome the drawbacks of the spray drying method, i.e., poor dispersion, insufficient mixing, and too large particle size of the starting precursors, unreacted TiO2 or secondary phase of Li2TiO3, Chang-Jian et al. [42] succeeded in synthesizing pure LTO by controlling the Li and Ti ratios in the range of 0.800–0.900 or changing the parameters of thermal annealing through the spray-drying method. The pilot scale solution was composed of 411.34 g of TiO2 powder, 152.11 g of Li2CO3, and 60 mL of polyvinyl alcohol (PVA) solution (10%) added to 1360 mL of DI water. Lee et al. [436] fabricated Li4Ti5O12/pristine multiwalled carbon nanotube (LTO/P-MWCNT) composites with high-rate capability and compared, two additional LTO composites prepared by using oxidized MWCNTs and surfactant-treated MWCNTs through a similar spray-drying process. The hybrid supercapacitor composed of an LTO/P-MWCNT anode and an activated carbon cathode delivers an energy density of 70.9 Wh kg−1 at a power density of 0.03 kW kg−1 and a maximum power density of 21.8 kW kg−1 is achieved at an energy density of 24.3 Wh kg−1. Chien et al. [437] used two Ti sources, i.e., 100% anatase TiO2 and P25 TiO2 (80% anatase + 20% rutile), and LiOH as Li source to synthesize LTO via spray-drying method with the air inlet and outlet temperature of the spray-dryer were 175 and 113 °C, respectively. The electrochemical performance of the LTO material tested at 0.1C/0.1C prepared from P25 TiO2 (specific capacity of 169 mAh g−1) was superior to that of the LTO prepared from 100% anatase TiO2 (165 mAh g−1). Hsieh et al. [438] evaluated the influence of Li addition on the charge/discharge behavior of LTO. Other LTO products were prepared by spray technique using a mixture of TIP, LiNO3, and isopropanol [439,440].
Nanoscale and highly porous LTO powders (with 200 nm particle size) were synthesized by SD process from Li2CO3 and nanocrystalline anatase type TiO2 (72.75 wt.%, 30−40 nm) as precursors, in aqueous solution containing 3 wt.% of dispersant ammonium polycarboxylate. The homogeneous slurry obtained was spray-dried with the inlet and outlet temperatures of 200 and 100 °C, respectively, and heat treated at 850 °C for 5 h [404]. The influences of Li/Ti atomic ratios (0.784, 0.800, 0.816, and 0.832) on the performance of LTO synthesized through SDM were investigated by He et al. [408]. The results indicate that when the spray-drying precursors at the Li/Ti molar ratio of 0.816 are calcined at 700 °C for 16 h in air, a pure LTO phase (200 nm primary particle size) with a lithium-excess composition is obtained, which shows the best electrochemical properties, i.e., a specific discharge capacity of 135 mAh g−1 after 100 cycles at 5C rate. Ren et al. [416] prepared uniform LTO microspheres with high tap density by a newly designed industrial spray-drying approach at 220 °C. LTO sample obtained with a Li/Ti molar ratio of 0.75 and sintered at 850 °C exhibits the highest specific capacity of 173 mAh g−1 after 400 cycles at 1C rate. Several workers provided proof that Li addition influences the charge-discharge behavior of spray-dried LTO [338,339,340]. Table 14 summarizes the electrochemical performances of nanostructured LTO materials synthesized by the spray-drying method.
The overall reversible capacity in the voltage range of 2.5–0.01 V vs. Li+/Li is limited by the vacant tetrahedral (8a) and octahedral (16c) sites. The decrease in the capacity ratio from the second plateau (1.0–0.01 V) to the first plateau (2.5–1.0 V) with the R = I(311)/I(400) value proves that the ratio of site vacancy greatly affects the capacity contribution. At high C rates, the tetrahedral (8a) sites exhibit a better accessibility to accommodate Li-ions during the intercalation process, compared with the accessibility of octahedral (16c) sites.

2.15. Spray Pyrolysis Method

In conventional spray pyrolysis, the solution is atomized into a hot wall reactor where the aerosol droplets undergo evaporation and solute concentration within the droplet, drying, and thermolysis of the precipitate particle at higher temperature to form a microporous particle, and eventually a dense one by sintering [439,440]. Ceramic nanoparticles with high purity and relatively narrow size distribution can be prepared (see Zhu et al. [441] for experimental details). Spray pyrolysis reactors are routinely used in industry to make a variety of oxides (SiO2, TiO2, Al2O3, etc.). Combustion of liquid precursor droplets is currently used in the synthesis of mixed oxide powders by the so-called flame spray pyrolysis (FSP) process (see Figure 1a). The FSP is capable of producing mixed metal oxide powders in the size range of 1–200 nm from low-cost precursors with production rates up to 250 g h−1 [442]. This method has been successfully applied for the synthesis of LTO [115,117,443,444]. In 2005, Doi and co-authors [115] first reported the preparation of uniform nanosized LTO particles using an electrospray deposition method. The spinel-LTO particles have a fairly-uniform particle size of ~12 nm with a distinct crystal structure.
Bresser et al. [117] fabricated LTO nanoparticles using the FSP method with an average diameter of approximately 20–30 nm and a specific surface area of 91.7 m2 g−1. However, phase impurities such as rutile, anatase TiO2, and Li3Ti3O7 were detected by means of XRD analysis. The advantage of nanoparticles was electrochemically illustrated at a high rate of 100C providing a discharge capacity of 72 mAh g−1. In the FSP process used by Karhunen et al. [444], LiOAc and TIP raw materials were dissolved with a stoichiometric ratio of (4:5) in an equal volume mixture of toluene and 2-ethyl hexanoic acid. The resulting solution with a total metal concentration of 0.5 mol L−1 and the premixed methane-oxygen flamelet ignited the aerosolized precursor solution at temperatures in excess of 1700 °C. LTO particles with a mean size of 80 nm were formed (Figure 17). Ernst and co-authors [75] used the FSP method to synthesize spinel-LTO with primary crystallite sizes of 7–30 nm and a high-temperature stability from lithium tert-butoxide dissolved in tetrahydrofuran (THF) and titanium (IV) isopropoxide dissolved in xylene with Li/Ti ratios in the range of 0.5–1.0. It has been found that the FSP process optimization could be used to further remove impurities. Moreover, FSP-prepared LTO nanoparticles showed good sintering stability at elevated temperatures and the scalability of the process. Waser et al. [445] reported the design of a scalable process for size-controlled flame spray synthesis of LTO particles. The role of air entrainment (AE) during aerosol synthesis was quantitively analyzed by computational fluid dynamics and temperatures were verified by FTIR spectroscopy. The LTO precursor solution (1.0 mol L−1 total metal concentration) consisted of Li-acetylacetonate and Ti(IV)-isopropoxide in a stoichiometric molar ratio of Li:Ti = 4:5, diluted with a 1:3 (by volume) mixture of 2-ethylhexanoic acid and xylene solvents. For AE > 50 L min−1, more than 95% pure LTO was obtained with only minor impurities of anatase and rutile TiO2. This process appears superior to the as-prepared LTO obtained by open FSP [75] with 87 wt.% LTO content and dBET of 20 nm and FSP-made nano LTO by Bresser et al. [117], but similar to that by Karhunen et al. [444].
Gockeln et al. [77] fabricated ultrafine LTO/C composite using combined double flame spray pyrolysis and pressure-based lamination technique. Within the flame spray, the nanostructured LTO particles are formed at temperatures of ≥2000 °C upon nucleation, surface growth, coagulation, and coalescence. To synthesize carbon from xylene, the N2 gas was chosen as the dispersant (1.5 L min−1, nozzle pressure drop 1.5 bar) and the support flame was maintained at under-ventilated fuel-to-air ratio conditions (CH4 of 2.5 mL min−1 and O2 of 0.5 mL min−1) resulting in a strong sooting flame. Ju et al. [72,73,74] controlled the mean size of the LTO powders by changing the concentrations of the spray solution. The precursor solution was prepared by dissolving LiNO3 and Ti[OCH(CH3)2]4 in distilled water using a small amount of nitric acid. The precursor powders had fine size, narrow size distribution, dense inner structure, and homogeneous composition when the flow rate of the carrier gas and the preparation temperature were 10 L min−1 and 800 °C, respectively. high discharge capacities were delivered by particles fabricated with high concentrations above 0.5 mol L−1. In another work, the same group of researchers studied the effects of drying control chemical additives on the properties of LTO using a solution containing dimethylacetamide (DMA) as a drying control chemical additive and citric acid (CA) and ethylene glycol (EG) as organic additives [446]. It was concluded that the addition of DMA to the polymeric spray solutions containing CA and EG helps in the effective control of the morphology of as-sprayed LTO powders. Meierhofer et al. [76] studied the LTO phase purity anode using FSP via precursors dissolved in five different organic solvents, i.e., ethanol (EtOH), benzyl alcohol (BnOH), tetrahydrofuran (THF), xylene, and ethylhexanoic acid (EHA). The effect of precursor and solvent parameters such as chemical reactivity, boiling point, and combustion enthalpy on the particle formation either via gas-to-particle (evaporation/nucleation/growth) or via droplet-to-particle (precipitation/incomplete evaporation) was discussed. The presence of carboxylic acid in the precursor solution resulted in pure (>95 mass %) and homogeneous LTO nanoparticles of size 4−9 nm, for two reasons: (1) stabilization of water-sensitive metal alkoxides precursor and (2) formation of volatile carboxylates from lithium nitrate. In contrast, the absence of carboxylic acids resulted in larger inhomogeneous crystalline TiO2 particles with a significant reduction of LTO content as low as ~34 mass %. LTO thin films (400 nm thick) were prepared on quartz and gold substrates by the spray pyrolysis method by using lithium-acetyl acetonate (Li-acac) and titanil-acetil acetonate (TiO(acac)4) with di-methyl formamide (DMF) solvent as starting materials [447]. Yamada et al. [448] prepared spray-pyrolyzed C/Li4Ti5O12 powders using LiNO3 and titanium isoproxide dissolved in an aqueous solution with a different organic acid such as lactic, malic, and citric acid in 0.1–0.4 mol L−1 concentration. The use of lactic acid as a carbon source was most effective for their improvement. Kim and Kang [449] prepared amorphous LTO particles using inexpensive lithium nitrate (LNT) and TTIP dissolved in a mixture of ethyl alcohol and distilled water (volumetric ratio, 3/7) with a total metal concentration of 0.5 mol L−1. Du et al. [450] synthesized crystalline LTO particles (10–30 nm) using LiOAc and TBT dissolved in ethanol to obtain a 0.5 mol L−1 solution. The molecular-level uniformity of the precursor allows for the synthesis of LTO with a significantly shorter heat treatment compared to conventional solid-state reaction, which in turn saves energy during large-scale production. Spray-pyrolyzed LTO delivered a discharge capacity of 146 mAh g−1 at a 10C rate for up to 500 cycles. Electrochemical performances of nanostructured LTO materials synthesized by the spray-pyrolysis technique are listed in Table 15.
Figure 17. (A) Scheme of a flame spray pyrolysis apparatus. The main elements are the atomizer, preheater, non-premixed slot burner, and collection system. Ultrasonic transducers operating at 1.7 MHz provide the driving force for atomization of the precursor inside the vessel, generating a fine aerosol with a number-based mean diameter of 3.5 μm. A flow of carrier gas comprising air enriched with oxygen transported the solution droplets at a rate of approximately 1 g droplets min−1. The flame synthesis temperature (~1700 °C) is adjusted by the gas mixture CH4/N2 used as fuel. (B) morphology of LTO products: (a) TEM image of single aggregate, (b) High-resolution TEM image. Reproduced with permission from [444]. Karhunen et al. under the Creative Commons Attribution License.
Figure 17. (A) Scheme of a flame spray pyrolysis apparatus. The main elements are the atomizer, preheater, non-premixed slot burner, and collection system. Ultrasonic transducers operating at 1.7 MHz provide the driving force for atomization of the precursor inside the vessel, generating a fine aerosol with a number-based mean diameter of 3.5 μm. A flow of carrier gas comprising air enriched with oxygen transported the solution droplets at a rate of approximately 1 g droplets min−1. The flame synthesis temperature (~1700 °C) is adjusted by the gas mixture CH4/N2 used as fuel. (B) morphology of LTO products: (a) TEM image of single aggregate, (b) High-resolution TEM image. Reproduced with permission from [444]. Karhunen et al. under the Creative Commons Attribution License.
Micromachines 15 00310 g017

2.16. Sonochemical Process

The sonochemical process is an energy-miser route, which yielded nanoscale homogeneous spinel LTO product. Generally, the estimated temperature and pressure in the liquid zone around the collapsing bubble generated in a sonochemical reactor at a driving frequency of 20 kHz with an input power of 179 W are approximately 1000 °C and 500 atm, respectively. Lee et al. [104] synthesized homogeneous LTO nanoparticles using TiO2 nanoparticles coated with a LiOH thin coating (2–5 nm) via a sonochemical method, operated at 20 kHz and 220 W for 20 min. The resulting LTO nanoparticles thermally treated at 500 °C for 1 h, have an average grain size of about 30–40 nm with excellent phase purity. Kim et al. [454] reported the growth of LTO ceramic from hydrogen titanate nanowires as precursors implemented by using TiO2 having a size of 60 nm and NaOH, and performing synthesis at 70 °C for 6 h with a sonochemical method. Ghosh et al. [105] reported the sonochemical synthesis carried out using commercial precursors LiOH·H2O and TiO2 (spherical particles of diameter 135 nm) mixed in ethyl alcohol dispersive media. The ultrasonic process was conducted in the following three steps: (1) first, TiO2 was dispersed in ethyl alcohol using an ultrasonicator for 20 min, (2) LiOH·H2O was dissolved into this mixture with continuous stirring for 10 min, (3) this precursor mixture was kept under the probe sonicator inside an ice bath to perform sonication experiment at room temperature. The sonication operation was carried out by applying sonic wave for 10–40 min at a pulse rate of 2 s with a pause of 1 s using a Ti probe (Φ = 13 mm) operating at 500 W (with amplitude of 35%, 20 kHz). The LTO nanomaterial investigated by Ni et al. [106] was prepared via an ultrasonic-assisted liquid deposition method. Typically, 3.4 g titanium tetrabutoxide (C16H36O4Ti, TBOT) was dispersed in 80 mL ethanol with sonication for 1 h. A mix solution of distilled water (2 mL), ethanol (18 mL) and 0.56 g CH3COOLi was added drop-wise into the above suspension with further sonication for 1 h. The precursor was calcined at 400 °C for 4 h, and subsequently sintered at 750 °C for another 10 h under N2 to obtain well-dispersed nanoparticles (~100 nm in calculated average size). Jin and co-workers [107] used a sonication process for 2 h to disperse TiO2 nanopowders in DI water before mixing them with LiOH, which yields white-colored suspensions. Mao et al. [109] reported a one-step continuous synthesis of spherical LTO/graphene composites through direct aerosolization of a GO suspension mixed with LiOH and titanium(IV) bis(ammonium lactato) dihydroxide as Li and Ti precursors. The mixed solution was nebulized by an ultrasonic nebulizer (2.4 MHz) to form aerosol particles, which were carried by argon gas through a horizontal tube furnace preheated at a desired temperature (750 °C). The TEM image also reveals the LTO nanocrystals are closely packed on the graphene surface, with a diameter of a few nanometers (5–10 nm). The LTO/CG composite, as an anode in LIBs, cycled at 1.25 A g−1 shows a capacity of ca. 100 mAh g−1 and retaining 88 mAh g−1 after 5000 cycles.

2.17. Microwave Synthesis

Microwave (MW) synthesis has been known to provide a fast route for reactions in the solid state, and it often occurs at lower temperatures than the conventional synthesis process. The higher reaction and cooling rates during microwave irradiation make the formation of metastable phases possible. The pioneering research on microwave synthesis of titanates was reported by Bhat and coworkers, and they have obtained lithium lanthanum titanate from metal oxides [455]. Due to the high penetration depths of the microwave, it can be uniformly and rapidly absorbed by LTO, and thus the high temperature needed in a solid-state reaction can be obtained within several minutes [216,351,396,456,457]. Yang et al. [458] implemented a hybrid microwave solid-state synthesis to prepare LTO grains with an average size of ~1 µm starting using a mixture of Li2CO3 and TiO2, as reactants, finely grounded and pressed into pellet with a pressure of 9 MPa and placed in a domestic microwave oven working at the frequency of 2.45 GHz and maximum power of 800 W. Li et al. [456] fabricated regular sphere-like LTO nanoparticles (average size of 40–50 nm) using microwave irradiation at 700 W for 15 min. However, the morphology of the product is also not easy to control in solid-state reactions. The present combination of microwave-assisted hydrothermal and microwave post-annealing methods is a great time-saving method to produce nanomaterials with controlled morphologies. Chou et al. [370] implemented a rapid method for the preparation of microspheres composed of nanoflakes within 1 h by a combination of a microwave-assisted hydrothermal method and a microwave post-annealing process. A polytetrafluoroethylene (Teflon)-lined autoclave was put under microwave-assisted hydrothermal treatment at 150 °C for 15 min to yield the as-prepared LTO precursor, followed by washing three times in deionized water and three times in acetone and finally annealed by microwave heating in air for 20 min.

2.18. Rheological Phase Reaction

The rheological phase reaction is an efficient soft chemistry method to prepare electrode materials. Contrary to the conventional SSR, this method provides homogeneous [459]. A typical rheological phase reaction involves five steps: dissolution, diffusion, reaction, nucleation, and growth. The solid reactants are fully mixed by adding a proper amount of deionized water or other solvents to produce a solid–liquid rheological mixture, in which the solid particles are uniformly distributed in the liquid substance. Yin et al. [113] have reported the synthesis of individual particle Li4Ti5O12 (~140 nm) with a high-rate performance by a modified rheological phase method, which occurs as follows. Stoichiometric amounts of lithium acetate dihydrate (CH3COOLi∙2H2O) and tetra-n-butyl titanate (Ti(OC4H9)4) were sufficiently ground in the mortar at room temperature for half an hour. The hydrolysis of Ti(OC4H9)4 started readily during the grinding process, accompanied by evaporation of part of acetic acid and butanol. As the reaction proceeded, the mixture became mushy and underwent gradual changes in color from colorless to white. Then a trace amount of deionized water was added to accelerate the hydrolysis of Ti(OC4H9)4 and reduce the conglutination resistance during the grinding. After sufficiently grinding for a few minutes, the white rheological phase body was transferred into a container and the container was sealed in a stainless autoclave at 110 °C for 12 h. Subsequently, the ultrafine white powder lithium titanate was obtained by calcination at 800 °C for 12 h in air. Liu and co-workers [460] synthesized LTO particles (average size of 2.1 μm with a narrow size distribution) by a simple rheological phase method using polyvinylbutyral (PVB) as both template and carbon source. Stoichiometric amounts of Li2CO3 and TiO2 were blended homogeneously and then added into the PVB/ethanol (9.2 g L−1) solution to obtain a solid–liquid rheological mixture. After continuous magnetic stirring of the mixture for 4 h, drying at 80 °C for 6 h in a vacuum oven to eliminate ethanol adequately, the sample was sintered at 800 °C for 15 h under argon flow. Wang et al. [461] synthesized an LTO/C composite (particle size of ~500 nm) by starch-sol-assisted rheological phase method using anatase TiO2 and LiOH·H2O as raw materials. The electrochemical tests show that LTO/C presents a capacity retention of 87% (500 cycles at 1C) and 73.0% (2000 cycles at 20C).

2.19. Electrospinning Method

Electrospinning is a synthesis method used to generate micro- or nanofibers from a polymer solution using a high-voltage power supply. In this technique, a precursor solution is loaded into a syringe and is subjected to a high voltage (1–20 kV), which lets the solution be extruded from a nozzle forming a jet. Fibers are formed and deposited on the collector by the jet during drying. Electrospinning creates a high surface area to volume ratios needed for most catalyst systems depending on operation parameters (voltage, flow rate, distance between tip and collector, concentration, and viscosity of the solution). The scheme of the electrospinning synthesis route is shown in Figure 1a. Electrospinning has been successfully employed to prepare 3D net architectures of spinel LTO nanofibers [95,462,463]. In a typical experiment, the electrospun solution consists of a mixture of TIP, lithium acetylacetonate (LiAAc), and polyvinylpyrrolidone (PVP) in isopropyl alcohol. The electrospinning process is conducted in air. A dc voltage of 1.2 kV is applied between a gauge needle and a stainless-steel collector electrode. Finally, PVP is removed from nanofibers by treating them in air at 500 °C for 1 h. Guo and co-workers [464] fabricated two types of LTO/C composites through the electrospinning method (high voltage of 10 kV) of two different solutions: (i) the first composite consists of LTO nanoparticles and aggregates coated by carbon and connected by carbon nanofibers using a solution prepared by adding 0.8 g polyacrylonitrile (PAN, Mw =150,000) and 2 g nano-LTO (primary particle size of 20–60 nm) into 17.2 g dimethylformamide (DMF) solvent, and the second composite is constructed solely by LTO/C fibers using a mixture of 1.52 g tetrabutyl titanate, 0.23 g lithium acetate and 1 g PVP into 18 g isopropyl alcohol solvent. Electrochemical tests show specific capacities of 115 and 120 mAh g−1 at a 2C rate for LTO/C particles/fibers and LTO/C fibers, respectively. LTO@C nanofibers containing tiny nanoparticles were obtained by electrospinning based on an adjustable solution strategy [132]. Typically, 4 mmol of Li(CH3COO)2·H2O, 5 mmol of titanium(IV) oxyacetylacetonate, and 1.4 g poly(vinylpyrrolidone) (PVP, Mw = 1,300,000) were dissolved into 10 mL ethanol. After vigorous stirring for 12 h, the homogenous precursor solution was poured into a syringe connected to a plastic needle, while a copper wire attached to a high-voltage generator was placed in the solution. A tension of 20 kV was applied between the needle and the Al foil target used for collection. Acetate and acetylacetonate were completely decomposed during the high-temperature process, while PVP gel was carbonized into conductive carbon nanofibers. The whole synthetic route is a simple, low-cost, and high-yield process. Finally, the as-collected electrospun fibers were calcined at 400 °C for 2 h and 800 °C for 5 h under an Ar atmosphere to obtain LTO@C hierarchical nanofibers. Wang et al. [465] investigated the growth of LTO nanofibers (average diameter of 230 nm) via thermal treating electrospun precursor fibers at 700 °C. Ji et al. [466] reported the electrospinning preparation of one-dimensional Ce3+-doped Li4Ti5O12 sub-micron belts with a width of approximately 500 nm and a thickness of about 200 nm. Li et al. [467] used a precursor solution of lithium acetate, tetraisopropyl titanate, acetic acid, and different doses of PVP. The solution was loaded into a plastic syringe and subjected to electrospinning. The flow rate was 1.5 mL h−1 and the distance between the tip of the needle and collector was 10 cm, the collector drum speed was 800~1000 rpm, and the experiment voltage was 12 kV. It was demonstrated that the viscosity of LTO precursor solution (from 11 to 117 cP) dramatically affects the morphology of LTO submicron fiber (from 0.17 to 0.58 µm diameter). Jo et al. [137] reported the synthesis of LTO nanofibers with uniform diameter <300 nm, a specific surface area of 9.7 m2 g−1, and an average pore size of ~2 nm through electrospinning. Xu et al. [140] prepared electrospun conformal Li4Ti5O12/C fibers using an as-spun fibrous precursor containing PVP, TIP, and LiAAc stabilized at 350 °C in air and then thermally treated at 800 °C in an N2 atmosphere. Long and continuous fibers with smooth surfaces and an average diameter of approximately 500 nm were observed (Figure 18).
The LTO electrospun nanofibers, with an average diameter of ~250 nm, were prepared by Park et al. [468] from the solution precursor of 1.42 g titanium butoxide mixed with a solution containing 2 mL ethanol and 1.25 mL acetic acid, 0.354 g lithium acetylacetonate and a polymeric solution consisting of 3 mL ethanol and 0.5 g PVP. The feeding rate was set at 0.2 mL h−1 and a high voltage of ~20 kV was exerted to collect nanofibers. Chen and co-workers [469] prepared electrospun LTO fibers using acetic acid, ethanol, butyl titanate, lithium acetate, and PVP (K90) as the raw materials according to the ratio 4.28:14.19:4.16:1:2.45. The electrospinning process was carried out ay 19–21 °C under 20 kV dc high voltage. Zou et al. [470] reported the preparation of electrospun Li3.9Cr0.3Ti4.8O12 nanofibers as anode material for high-rate and low-temperature LIBs. The precursor solution was prepared by dissolving 6.808 g tetrabutyl titanate, 0.507 g chromium nitrate, 1.742 g lithium acetate, and 2.500 g polyvinyl pyrrolidone (PVP, Mw = 1,300,000) in a mixed solution with 30 mL ethanol and 10 mL acetic acid. The dc voltage of 20 kV was loaded on the needles, and the distance between the needle and a collector was 15 cm. The spinning speed was controlled at 2.0 mL h−1 using an auto-syringe pump. The as-spun nanofibers (heat-treated at 700 °C for 3 h in air) exhibited excellent lithium storage performance: 140 mAh g−1 at 10C and 91 mAh g−1 at 50C rates at room temperature, 100 mAh g−1 at 1 C rate at −20 °C. Electrospun bare LTO and Ag-LTO nanofibers were synthesized using 1.48 mL of titanium(Ⅳ) isopropoxide and 0.413 g of lithium acetate dissolved over 1 h in 10 mL of ethanol and 7 mL of acetic acid and added with 0.8 g PVP [471]. The randomly aligned nanofibers have average diameters of 50–200 nm and lengths extending to several tens of micrometers. At the highest rate of 30 C, the specific capacity of the Ag–LTO nanofibers is 107 mAh·g−1, which is significantly higher than that of the bare LTO nanofibers (82.2 mAh·g−1). LTO/Ag composite nanobelts with an average width of ca. 1.12 µm were synthesized by Li et al. [144]. These composites showed good rate capability (specific capacity of 132 mAh g−1 at 15C rate) and cycling stability (172 mAh g−1 after 100 cycles at 0.2C rate). Castano et al. [96] synthesized LTO fibers via a two-step method. First, a precursor solution containing a spinnable polymer, and titanium and lithium salts, which were dissolved in a mixture of solvents, was subjected to a process of electrospinning. Isopropyl alcohol was used as a solvent for PVP as well as for lithium acetate and titanium butoxide precursor salts. Then, the obtained fibers were calcined at temperatures between 650 and 850 °C for 7–10 h in air or argon. Zhang et al. [472] synthesized a self-standing nonwoven flexible Li4Ti5O12/carbon nanofiber composite (denoted LTO/CNF) by using a facile method involving the electrospinning fabrication of CNFs and chemical deposition of LTO over the CNF surface. LTO/CNF film was composed of 50 ± 20 nm diameter LTO polycrystalline mesoporous particles distributed over 300 ± 50 nm diameter CNF nanofibers. As an anode, it delivered a specific capacity of 109 mAh g−1 at a rate of 50C.
Li4Ti5O12/rutile TiO2 (LTO/RT) heterostructured nanorods with tunable oxide phases in situ fabricated by annealing the electrospun nanofiber precursor delivered a high capacity of 160.3 mAh·g−1 at 1C after 200 cycles, 125.5 mAh·g−1 at 10C after 500 cycles and superior capacity retention of 90.3% after 1000 cycles at 30C [473].
Wang et al. [474] utilized an amphiphilic triblock copolymer (Pluronic F127) as a surfactant owing to the hydrophilic group for the synthesis of electrospun mesoporous LTO/C nanofibers. The initial concentration of F127 in the inorganic precursor solution was lower than its critical micelle concentration (cmc). During electrospinning, the progressive evaporation of ethanol concentrated the surfactant to a concentration that is higher than its cmc. Meanwhile, TIP and LiOAc underwent hydrolysis and condensation, leading to the formation of hydroxyl Li–Ti–O precursors. The self-assembly of hydroxyl Li–Ti–O precursor to form mesopores was driven at a higher concentration of surfactant. For the synthesis of LTO nanowires (500 nm diameter, 10 µm length), Liu et al. [475] used a solution of ethanol (5.0 g), acetylacetone (0.5 g), and acetic acid (1.6 g), where 1.4218 g of tetrabutyl titanate and 0.2298 g of lithium acetate were added. After stirring for 2 h, 0.55 g of PVP (Mw = 1,300,000) was dissolved into the solution. A transparent yellow precursor was formed after stirring for 24 h at room temperature. The electrospinning process was performed using a 10 mL plastic needle tube tipped with a 24-gauge stainless-steel needle at a rate of 0.8 mL h−1. The applied high voltage between the injector nozzle and the collector was 15 kV, and the distance was about 7 cm. The nanowires were annealed in air atmosphere at 700 °C for 6 h with a heating rate of 5 °C min−1. Characterizations and electrochemical properties are displayed in Figure 19. In the potential range 1–2 V, the LTO nanowires delivered a specific capacity of 155 and 80 mAh g−1 at 1C and 50C rates, respectively. Table 16 lists the details of the electrospinning synthesis and electrochemical performance of LTO nanofibers.

2.20. Ion Exchange

Ion exchange is a chemical process in which ions of a certain charge contained in a solution (for example cations) are removed from this solution by adsorption on a solid material (the ion exchanger), to be replaced by an equivalent quantity of other ions of the same charge emitted by the solid. Oppositely charged ions are not affected. An ion exchanger is a salt, an acid, or a base, solid and insoluble in water, but hydrated, that is to say, swollen with water like a sponge. The water content of an apparently dry material can be more than 50% of its total mass and the exchange reactions take place in this water, called swelling or hydration water, inside the exchanger. Ion exchange is another method used to prepare nanoscale LTO particles. This method is feasible for mass production compared with hydro- or solvothermal treatments.
A novel process for the fabrication of LTO was proposed by Lu et al. [121]. LTO was prepared by a two-step ion-exchange technique. They first prepared α-Li2TiO3 via the hydrothermal treatment of anatase TiO2 with a uniform particle size of 4.5 nm in a 2 mol L−1 LiOH aqueous solution, which was then reacted with a proper amount of HCl to allow an ion exchange between the α-Li2TiO3 and the solution. This led to the formation of (Li0.4H0.6)2TiO3 nanocrystals. Well-dispersed uniform LTO single crystals (size of ~40 nm) were obtained after the further calcination of the precursor at 400 °C for 2 h. Ma and Cheng [376] used metatitanic acid and lithium hydroxide as raw materials, for the preparation of LTO via one-step hydrothermal ion exchange synthesis method. Yagi et al. [141] prepared Na+-intercalated hydrogen titanate nanotubes as LTO precursors from a colloidal solution by dispersing 0.5 g of anatase TiO2 powder in 10 mL of a 10 mol L−1 NaOH aqueous solution. Li4Ti5O12 with nanotube/nanowire morphology and high surface area was prepared using low-temperature hydrothermal ion exchange processing from hydrogen titanate (H-TiO2) precursor [130]. H-TiO2 nanotubes were prepared from industrial TiO2 powders via a sonochemical-hydrothermal reaction in concentrated 10–15 mol L−1 NaOH at 130–170 °C for 24–72 h. Further hydrothermal treatment of the H-TiO2 nanotube/nanowires in the aqueous LiOH solution allowed an ion exchange between H-TiO2 and the solution to form an LTO precursor. Na+ ions in the interlayer of the synthesized H-TiO2 were exchanged with Li+ ions by a reflux treatment using 100 mL of a 2.5 mol L−1 LiOH aqueous solution, in which 1 g of the Na+-intercalated H-TiO2 powder was dispersed.
LTO nanowires were synthesized via a two-step ionic exchange process from Na2Ti3O7 nanowires obtained by hydrothermal reaction of P-25 TiO2 dispersed in 10 mol L−1 NaOH solution [132]. Na2Ti3O7 nanowires were subjected to ionic exchange in an HCl aqueous solution to produce H2Ti3O7 nanowires. Next, H2Ti3O7 nanowires were treated in a LiOH aqueous solution via hydrothermal ion exchange at 100 °C for 24 h. Kim et al. [477] elaborated the H2Ti2O5·H2O precursor by hydrothermal ion exchange treatment of NaOH on TiO2 with HCl solution. LTO nanofibers were subsequently prepared by a second hydrothermal ion exchange process of H2Ti2O5·H2O and LiOH·H2O and a final calcination at 350–400 °C. Kataoka et al. [478] prepared the lithiated H2Ti12O25 sample by the H+/Li+ ion exchange synthetic technique in the molten LiNO3 at 270 °C using H2Ti12O25 as a starting compound. The obtained sample, H1.05Li0.35Ti12O25−δ having δ = 0.3, exhibits a high discharge capacity of 192 mAh g−1 at 1 A g−1 current density. The completely Li+/H+ ion exchanged sample using LiOH or LiI in high-temperature solutions was unsuccessful, resulting in the formation of Li4Ti5O12 and Li2TiO3 as main products. Nanoparticle-stacked Li4Ti5O12-TiO2 nanowire arrays (SLTO) were synthesized via a two-step ion exchange [375]. First, Ti foils were hydrothermally treated in 1 mol L−1 NaOH and 3 mol L−1 NaCl at 200 °C for 12 h and subsequently immersed in a diluted HCl solution for several hours to thoroughly exchange the Na+ with H+. Second, the obtained H2Ti2O5·H2O nanowire arrays were placed in the solution of 2 mol L−1 LiOH·H2O at 60 °C for 10 h to thoroughly exchange the H+ with Li+. Electrochemical performances of nanostructured LTO materials synthesized by using the two-step ion exchange technique are listed in Table 17.

2.21. Thin-Film Techniques

Li4Ti5O12 thin films were fabricated using various deposition techniques such as sol–gel coating [257], solid-liquid phase epitaxy [481], radio-frequency (RF) magnetron sputtering method [482,483], direct-current (DC) magnetron sputtering epitaxy [484], and laser-pulse deposition (PLD) technique [13,485,486,487,488,489,490,491]. The sputtering deposition method is the most popular technique to grow metal-oxide films since it allows faster deposition rates. Its main advantage comes from the production of good surface uniformity of as-deposited films. Advantages of the PLD technique include easy control of the film composition by tuning the deposition parameters and a good reproducible stoichiometry of the target material in the films. In the PLD technique, a pulsed laser beam (10 ns duration) is focused by a lens to ablate the Li-rich LTO target. The energy of the beam is in the range of 100–500 mJ per pulse (laser fluence of 10 J cm−2) with a laser pulse repetition of 10 Hz. Target and substrates are placed inside the deposition chamber and evacuated to ~1 mPa. To avoid depletion of the deposit at any given spot, the target rotates at 10 rpm. For reactive synthesis, pure oxygen gas is introduced into the chamber with a typical partial pressure (pO2) in the range of 0.1–10 Pa to obtain single-phased and stoichiometric LTO films. The typical deposition rate is in the range of 2–5 Å s−1. Epitaxial thin-films fabricated by PLD could provide an ideal surface with a roughness of less than 1 nm. However, a target ceramic containing excess Li species is required to fabricate stoichiometric Li4Ti5O12 thin films [492].
Rho et al. [256,257,493,494] fabricated LTO films using a sol–gel coating with Ti and Li isopropoxide added by propanol (i-C3H7OH), acetic acid (CH3COOH)and poly(vinylpyrrolidone) (PVP, 55,000 m.w.) as an additive for sol, which is very effective for preparing crack-free thin films. The molar composition of the starting solution was 5:4:100:100:5. The sols were coated on the Au substrate or the quartz glass substrate with a spin coater at 3000 rpm. The prepared sol films were converted to gel films in the course of the spin-coating process. Then, the gel films were heated at 800 °C for 1 h. LTO thin film 0.4 µm thick showed a high electrochemical response until 50 mV s−1. Mosa et al. [174] reported on the sol–gel process and coating preparation of spinel Li4Ti5O12 thin films by dipping using lithium ethoxide in situ that reacts with titanium alkoxide as the starting reagents to produce precursor solution without precipitation. The alkoxide solution method shows promising potential for control over particle size, high purity, good chemical homogeneity, low thermal treatment temperatures and times, and micro-structural uniformity. In situ production of a lithium alkoxide and its reaction with a titanium alkoxide reduces the thermal treatment to obtain a pure phase at only 700 °C and 15 min. Wang et al. [482] reported the fabrication and characterization of LTO films grown by RF magnetron sputtering onto Au(100 nm)/Ti(10 nm)/SiO2/Si substrate maintained at various temperatures in the range of 500–700 °C. The spinel phase of LTO appears at deposition temperatures above 500 °C. The films were deposited at the pressure of 30 mTorr with the mixed Ar/O2 (3:2) gas, and the power density was estimated to be ~4 W cm−2. Wude et al. [495] reported LTO thin films (250 nm thick) deposited by dc-ion beam sputtering at different O2 partial pressures and different substrate temperatures. Thin films prepared at a substrate temperature of 600 °C and an oxygen partial pressure of 3 × 10−4 mbar show the most characteristic LTO features. Kumatani et al. [477] fabricated Li4Ti5O12 (111) epitaxial thin films on α-Al2O3(0001) substrates by RF magnetron sputtering. Thin films of amorphous Li4Ti5O12 were deposited at room temperature, and then the films were annealed at high temperatures for the formation of epitaxial thin films. The RF power applied to the target was maintained at 100 W, and the films were deposited in a mixed Ar/O2 atmosphere at different oxygen partial pressures (pO2), controlled by the flow ratio of Ar to O2. The total pressure was kept at 1.0 Pa, and thin film thicknesses were approximately 100 nm (deposition rate was approximately 1.5 nm/min). After the deposition, the films were annealed at different temperatures (800–1100 °C) in a vacuum (~1.0 × 10−3 Pa).
The first PLD growth of LTO thin films deposited onto Pt/Ti/SiO2/Si substrates using a KrF excimer laser beam (248 nm, 250 mJ) was reported by Deng et al. [485]. Films annealed at 800 °C (410 nm thick) exhibited a cubic structure with a lattice constant 8.375 Å larger than that of the LTO crystal (8.359 Å). The SEM cross-section image (Figure 20a) revealed the porous morphology induced by the high-temperature treatment. The discharge-specific capacity was the largest for films annealed at 700 °C due to the optimized adhesion strength between film and substrate (Figure 20b). The anode films discharged at a current density of 10 µA cm−2 (0.58C rate) showed excellent cycleability; the discharge capacity remained 149 mAh g−1 after 50 cycles. Yu et al. [488] reported the growth of LTO films (545 nm thick) deposited at 600 °C for 90 min on conducting fluorine-doped tin oxide (FTO) by PLD using a KrF excimer laser at an energy density of 5 J cm−2 with a repetition rate of 9 Hz. Using high-purity oxygen kept at 30 P, the films are grown with a grain size of ~250 nm and roughness of 50 nm in the perpendicular direction toward the thin film.
Hirayama et al. [496,497] fabricated epitaxial Li4Ti5O12 thin-films deposited on SrTiO3 single-crystal substrates with (111), (110), and (100) lattice plane orientations using a PLD apparatus equipped with a KrF excimer laser with a wavelength of 248 nm under O2 atmosphere. LTO films have the same orientation as the SrTiO3 substrates: Li4Ti5O12 (111) on SrTiO3 (111), Li4Ti5O12 (110) on SrTiO3 (110), and Li4Ti5O12 (100) on SrTiO3 (100). These epitaxial films contained island structures, and the morphologies of the (111), (110), and (100) films exhibit angular, needle-like, and circular shapes, respectively. Impurity-free epitaxial LTO thin-film was obtained from a Li-rich target Li5.2Ti5O12 when deposited at Ts = 700 °C. The electrochemical features of LTO film anodes (20 nm thick) exhibited discharge capacities of ~200 and ~250 mAh g−1 for (100)- and (111)-orientation, respectively. Epitaxial LTO(110) films were also deposited on Nb:SrTiO3(110) substrate. These films (~28 nm thick) tested by cyclic voltammetry at a scan rate of 1 mV s−1 exhibited redox peaks at 1.53 and 1.60 V corresponding to the insertion and extraction of Li+ ions. As-deposited films at a substrate temperature of 700 °C in 6.6 Pa oxygen partial pressure exhibited high initial capacity (~200 mAh g−1) but poor stability [497]. Kumatani et al. [486] investigated the PLD growth process of epitaxial LTO films deposited on the MgAl2O4 (111) substrate. With Ts = 800 °C and PO2 = 1 × 10−3 Torr LTO films had excellent crystallinity and low resistivity of 3.3 × 10−4 Ω cm. at 25 °C. At lower PO2, the PLD LiTi2O4 film was formed, while at higher PO2 Ti was segregated as TiO2 rutile and Li0.74Ti3O6.
Figure 20. (a) SEM cross-section image of LTO film (410 nm thick) heat treated at 800 °C. (b) Charge-discharge profiles recorded at 20 µA cm−2 (i.e., ~1.15C) current density in the voltage range 1–2 V vs. Li+/Li of PLD films heated at various temperatures. Reproduced from [485] with permission. Copyright 2009 Elsevier.
Figure 20. (a) SEM cross-section image of LTO film (410 nm thick) heat treated at 800 °C. (b) Charge-discharge profiles recorded at 20 µA cm−2 (i.e., ~1.15C) current density in the voltage range 1–2 V vs. Li+/Li of PLD films heated at various temperatures. Reproduced from [485] with permission. Copyright 2009 Elsevier.
Micromachines 15 00310 g020
Studies of the electrochemical performance and kinetic behavior of PLD LTO films deposited on Pt/Ti/SiO2/Si substrates were reported by Deng et al. [489]. Using a Li-rich target (i.e., an excess of 5 wt.% Li2O), the films annealed at 700 °C for 2 h in air were well-crystallized items with densely packed grains. The galvanic charge-discharge plateau was observed around 1.56 V and an initial specific capacity of 159 mAh g−1 was delivered with a retention of 93.7% after 20 cycles. The diffusion coefficient of Li+ ions in such an LTO framework was in the range of 10−15–10−12 cm2 s−1. The energy barrier of the diffusion of lithium ions was estimated to be Ea = 0.11 eV in LTO (111)-oriented PLD thin films (190 nm thick) grown on spinel MgAl2O4 (111) substrate [13]. Zhao et al. reported the optical properties of epitaxially grown LTO films on (001)-oriented MgAl2O4 substrate. The optical bandgap of 3.14 eV was measured for 86 nm thick films (surface roughness of 4.61 nm) [490]. Schichtel et al. fabricated an all-solid-state microbattery with LTO as a positive electrode. PLD films were obtained on various substrates at Ts = 650 °C under a 0.3 Pa pure oxygen atmosphere using a commercially available LTO powder. As-prepared films (650 nm thick) revealed columnar growth that allowed a Coulombic efficiency >97% after the second cycle and a discharge capacity of 33 µAh cm−2 at 3.5 µA cm−2 current density [491]. Pfenninger et al. demonstrated that LTO thin films deposited by PLD on MgO substrate kept at 500 °C using dense Li7.1Ti5O12 target sintered at 1000 °C for 12 h are compatible with Li6.25Al0.25La3Zr2O12 electrolyte pellet. Such films display stable structure and cycleability almost close to 175 mAh g−1. The typical voltage plateau at 1.57 V (oxidation) and 1.53 V (reduction) was observed at the current density of 2.5 mA g−1 [498]. Among the Li1+xTi1−xO4 ternary system, LiTi2O4 thin films were grown by the PLD route in the temperature range of 400–800 °C using a target with a higher Li/Ti ratio of 0.8 [499]. Chopdekar et al. grew epitaxial PLD LiTi2O4 thin films on various crystalline-oriented substrates such as single crystalline substrates of MgAl2O4, MgO, and SrTiO3 [500]. The authors stated the PLD conditions with Ts held at 450–600 °C in a vacuum better than 5 × 10−6 Torr without mention of oxygen partial pressure, while Kumatani determined that stoichiometric LiTi2O4 thin films were obtained at a PO2 of 5 × 10−6 Torr with Ts = 800 °C [486]. Pagani et al. [501] reported the growth of epitaxial, single-crystal, strain-free LTO thin films (150 nm thick) deposited by PLD on MgO (111) single-crystal substrate at 500 °C in an oxygen atmosphere at a pressure of 1.7 Pa. A KrF excimer laser (λ = 248 nm) was employed at a fluence of 2.6 J cm−2 and a frequency set at 10 Hz. For comparison, polycrystalline LTO film, representing a nonideal system including grain boundaries, was deposited on a sputtered polycrystalline MgO film on Si(100). In a second publication, the target for PLD was prepared by mixing and grinding Li2CO3 and TiO2 rutile to obtain an overlithiated target with a composition of Li5.35Ti5O12. Recently, the same group [484] used Li4Ti5O12 and Li2O targets simultaneously in confocal configuration to sputter LTO epitaxial films at a power density of 4.93 and 1.48 W cm−2, respectively, and at a substrate temperature of 700 °C in an argon atmosphere at a pressure of 1.5 mTorr and an argon flow of 30 sccm resulting in a deposition rate of 0.14 Å s−1. Lithium losses were mitigated by Li2O co-sputtering at a rate of 0.02 Å s−1. Yu et al. [487] prepared textured Li4Ti5O12 thin films (thickness from 35 to 210 nm) via PLD. Films were deposited on various substrates maintained at 450 °C under a flowing oxygen atmosphere at 5 Pa. An XeCl excimer laser with a wavelength of 308 nm and laser pulse frequency of 8 Hz was used for thin-film growth. Both the target and substrate holder were rotated at a constant speed to ensure uniform film thickness in each sample. The distance between the target and the substrate was kept at 4.7 cm. Prior to film deposition, the vacuum chamber was evacuated to a background pressure of 1 × 10−4 Pa. The thin films were further annealed at 600 °C in an atmospheric environment for 30 min to increase their crystallinity. The growth rate is estimated to be 2.33 nm min−1. It is shown that the pseudocapacitive charging is remarkably activated by the nanocrystalline microstructure full of defect-rich surface, which can simultaneously promote Na-ion and electron accessibility to the surface/subsurface. A highly reversible charge capacity of 225 mAh g−1 at 1C is achievable for the Li//LTO supercapacitor with 1 mol L−1 LiClO4 in a nonaqueous solution of ethylene carbonate and polycarbonate (50:50, wt.%). Cunha et al. [502] fabricated PLD LTO films deposited on Nd-doped single crystalline SrTiO3 (100), (110) and (111) substrates at Ts = 700 °C under PO2 = 20 Pa from a sintered Li4.8Ti5O12 (20 wt.% excess Li2O) target. It was shown that increasing the relative amount of ⟨111⟩ facets would significantly increase the storage capacity, although the reversibility of this capacity could be limited by irreversible surface reactions at these high compositions.

2.22. 3D Ink Printing

The fabrication of wearable batteries using 3D printing approaches is highly desired because of their capability of printing arbitrary shapes and sizes and configuring multiple materials at different positions as needed. These techniques include lithography-based 3D printing, inkjet printing, direct ink writing, and fused deposition modeling. The composition and rheology of each ink must be optimized to ensure reliable high aspect ratio electrode architectures [503]. Li4Ti5O12 is the most commonly used anode material in 3D-printed batteries [504,505,506,507,508,509,510,511,512,513,514,515]. The first 3D-interdigitated microbattery architectures (3D-IMA) with LTO/LFP materials were developed in 2013 by the Lewis group [504]. LTO ink was well designed by adding deionized water, 20–30 wt.% ethylene glycol, 27 wt.% glycerol, and 9 wt.% cellulose-based viscosifiers. Prior to printing, interdigitated Au current collector patterns were prepared by a combination of lithographic patterning and e-beam deposition. The LTO ink was then deposited onto a pattern to form multilayer electrodes. After the printing was finished and the electrodes were dried, the LTO interdigital structure was heated to 600 °C in an inert gas to remove the organic additives to advance the nanoparticle sintering. The packaged micro-battery showed a capacity of 1.2 mAh cm−2 at a rate of 0.5C and exhibited a high areal energy density of 9.7 J cm−2 at a power density of 2.7 mW cm−2.
Zhao et al. fabricated LTO thin films by using the ink-jet printing technique [505]. The average thickness of 10-layer LTO film was about 1.7~1.8 μm and the active material in the thin film was nano-sized about 50–300 nm. It was also found that the as-prepared thin film exhibited a high discharge capacity of about 174 mAh g−1 and the discharge capacity in the 300th cycle retained 88% of the largest discharge capacity at a current density of 10.4 μA cm−2 in the potential range of 1.0–2.0 V [505]. Chen et al. [506] fabricated a 3D microbattery using an ultraviolet-curable poly(ethylene glycol) (PEG)-based gel polymer as a resin for micro-stereolithography. Active materials, LiFePO4 (cathode) and Li4Ti5O12 (anode) were mixed with carbon black and the GPE resin, which was then flown into the 3D structure. This LFP/PEG-gel polymer/LTO microbattery delivered a specific capacity of 1.4 µAh cm−2 at a 2C rate. Inkjet printing of ionogel was elaborated to construct a lithium microbattery delivering 300 µAh cm−2 over 100 cycles (i.e., 60 mAh g−1 at 0.1C rate) [507]. Homogenous porous composite electrodes, obtained by tape casting an aqueous slurry, contained an active material powder (LTO and LFP), a polymeric binder (carboxyl methyl cellulose (CMC), Mw = 250,000 g mol−1), and an electronic conductor powder (carbon Super P, SSA = 60 m2 g−1). Zhou et al. [508] used the 3D printing technology to optimize the electrode geometry for the sake of enhanced electrochemical performance. 3D direct writing ink composed of LTO, carboxymethylcelluclose sodium (CMC), carbon nanotube, and water achieved LTO electrical conductivity of 2.08 S cm−1 and an ultrahigh areal capacity of 5.05 mAh cm−2. The full cells assembled with the 3D-printed LFP and LTO microlattices showed a stable capacity and still delivered a capacity of 102 mAh g−1 after 35 cycles. In a recent work, Wei et al. prepared LTO functional ink using 30 vol.% LTO with 1.35 vol.% Ketjenblack in 1 mol L−1 LiTFSI/anhydrous propylene carbonate with 1 wt.% PVP (Mw = 40,000 g mol−1). The packaged LIBs composed of thick LFP/LTO electrodes, customized separator, and glassy carbon current collectors delivered an areal capacity of 4.45 mAh cm−2 (second cycle) at a current density of 0.14 mA cm−2 [509]. Wang et al. [510] incorporated LTO nanofibers into a PVDF (Mw = 600,000 g mol−1) dissolved in n-methyl-2-pyrrolidone (NMP) solution with CNT conductive additive to make the anode ink. The all-fiber quasi-solid-state LIB device assembled by twisting LFP and LTO printed electrodes together with gel polymer electrolyte, exhibited a high discharge specific capacity of ≈110 mAh g−1 at 50 mA g−1. To achieve a balance between the electrochemical, mechanical, and rheological properties of ink-printed LTO anode, Kohlmeyer et al. [511] developed a system in which the ink components (i.e., active material, carbon nanofibers (CNFs), and polymer) could be independently varied to tune the ink rheology and the final electrode properties. Each batch consisted of 100 mg of active material and 5 mL solvent (i.e., 1-methyl-2-pyrrolidone (NMP), diethyl carbonate, and isopropanol). Direct ink write printing was performed by extrusion at a pressure ranging from 0.5 to 5 psi. The Li4Ti5O12/CNF/PVDF electrode with composition 40/40/20 displayed a conductivity of ~10 S cm−1 and delivered a specific capacity of 89 mAh g−1 at 5C rate. Ragones et al. [512] utilized a fused-filament fabrication (FFF) method for 3D printable microbatteries (3DPMs). The fabrication of electrochemically active LTO was realized using LTO/carbon/polyester polylactic acid mixture. The reversible capacity of the Li/0.3 mol L−1 LiTFSI–PYR14TFSI/LTO half-cells was 80 mAh g−1 at 30 µA cm−2 after 30 cycles. Recently, Viviani et al. [515] proposed a simple ink formulation that is aqueous-based, non-toxic, and safe to handle. In a typical procedure, 1.5 g of LTO powders, 0.21 g of polyvynilpirrolidone (PVP), and 0.187 g of a carbon-based conductive agent, either carbon black or MWCNTs, were ball milled at 420 rpm for 8 h with 2.5 mL of ethylene glycol and 2.5 mL 2-propanol (IPA). A second ball-milling step was carried out for 1 h at 150 rpm after adding 40 mL aqueous solution of 1 mmol L−1 lithium dodecyl sulfate (LDS) and 0.4 g of lithium polyacrylate (Li-PAA). Inkjet-printed LTO thin-film electrodes manufactured with carbon black and MWCNTs deliver a specific capacity of 150 mAh g−1 at 0.2 C, showing negligible capacity loss for over 100 cycles.

2.23. Miscellaneous Treatments

LTO confined in activated carbon nanopores (LTO/AC) prepared by a vacuum impregnation technique yielded nanocomposite electrodes for asymmetric supercapacitors [516]. Typically, 340 mg of titanium (IV) butoxide and 57 mg of lithium acetate were dissolved in 2 mL of methanol by ultrasonication for 5 min to obtain a transparent solution with a bright yellow color as the precursor. After adding the solution dropwise to 200 mg spherical nanoporous activated carbon, a vacuum was applied during the interval of every second addition. The precursor AC composites were then preheated in air at 300 °C for 1 h, followed by calcination at 800 °C in an argon atmosphere for 10 h. LTO-AC with a large content of 30−50 nm mesopores retains up to 50% of its capacity at the 200C rate. For boosting the Li-ion transport in the spinel framework of LTO, Zhu et al. incorporated oxygen defects by eco-friendly and cost-effective plasma treatment in an H2/N2 atmosphere at a pressure of 5 Pa at 150 °C for 2 h [517]. Oxygen vacancies were evidenced by XRD, XPS, UV-Vis, and ESR measurements (Figure 21). A clear shift of the (111) and (400) X-ray reflections is observed towards lower 2θ angles due to the lattice expansion of ~1% due to the partial reduction of Ti4+ to larger Ti3+ ions to maintain charge neutrality. The oxygen-deficient LTO delivered a capacity of 133 mAh g−1 at a 20C rate after 500 cycles with a Coulombic efficiency of 100% (Figure 21).
Sodium titanium oxide with a spinel-type structure (Na3LiTi5O12, NTO) phase was demonstrated to be a stable structure as an anode material for sodium-ion batteries (SIBs) [518]. NTO was fabricated by the reaction of commercially available LTO powder with a sodium–organic reduction reagent (SOR). SOR was prepared using an excess of sodium-metal flakes dissolved into the biphenyl 1,2-dimethoxyethane solution. The color of the LTO powder soaked into SOR solution for 1 day in Ar atmosphere turned to black from white, confirming that enough sodium insertion reaction has occurred.
Figure 21. Characterization of pristine LTO and oxygen defective LTO. HRTEM images of (a) as-prepared LTO and (b) plasma-treated LTO. (c) Cycling performance at 1C rate of pristine LTO and plasma-treated LTO. Insets show the electron paramagnetic resonance (EPR) spectra and the linear relationship between the peak current of cyclic voltammograms and the square root of scan rate. The high EPR signal at g = 2.003 in plasma-treated LTO originates from the unpaired electrons trapped in O vacancies. Reproduced with permission from [517]. Copyright 2019 The American Chemical Society.
Figure 21. Characterization of pristine LTO and oxygen defective LTO. HRTEM images of (a) as-prepared LTO and (b) plasma-treated LTO. (c) Cycling performance at 1C rate of pristine LTO and plasma-treated LTO. Insets show the electron paramagnetic resonance (EPR) spectra and the linear relationship between the peak current of cyclic voltammograms and the square root of scan rate. The high EPR signal at g = 2.003 in plasma-treated LTO originates from the unpaired electrons trapped in O vacancies. Reproduced with permission from [517]. Copyright 2019 The American Chemical Society.
Micromachines 15 00310 g021
To overcome the drawbacks of pure LTO materials, i.e., low electronic conductivity at room temperature and moderate Li+ diffusion coefficient, surface modification to change the oxygen stoichiometry has been utilized. Thermal nitridation of metal oxides has been known to change the oxygen stoichiometry, to reduce the metal oxidation state, and to introduce nitride thin films [519,520,521]. In 1994, Richard et al. demonstrated the modification of the surface structure of spinel oxides, LiMn2O4 and Li4Ti5O12 via annealing in ammonia [520]. Nitridation should change Li insertion/extraction behaviors because Li4Ti5O12−δ has an average Mn oxidation state of less than 4 (i.e., formation of a mixed-valent intermediate phase) and substantial Li can be extracted from it (i.e., resulting in an increase discharge capacity). From Rietveld refinements, it was found that LTO treated at 150 and 200 °C under a flow of anhydrous NH3 for 1.5 h exhibits oxygen occupation in 32e Wyckoff sites of 3.958 and 3.873, respectively. In addition, NH3 treatment should also make bonding between surface Ti and N simultaneously. As a functional material easily prepared, TiN has high thermal stability, high chemical stability, and, also, high electronic conductivity, and it introduces a significant enhancement in the battery performance. For example, a simple synthesis of nitrogen- and TiN-modified LTO by solid-state reaction of Li2CO3 and TiO2 anatase in an ammonia-containing atmosphere was reported [522]. A uniformly coated conducting TiN/TiOxNy layer was synthesized using electrospinning and subsequent nitridation process on the surface of LTO nanofiber [468]. Wang et al. [193] used a different procedure for the LTO nitridation by heating the powders from room temperature to 700 °C at the rate of 5 C min−1 under Ar atmosphere; then, the gas was changed into NH3 and the temperature was kept at 700 °C for ~2 h, followed by natural cooling to room temperature under flowing Ar gas preventing oxidation. The TiN layer formed on the surface of Li4Ti5O11.7Br0.3 powder after heating in NH3 provided excellent electrochemical performance, i.e., a reversible capacity of 138 and 104 mAh g−1 after 100 cycles at 10C and 20C, respectively. Li et al. have successfully prepared TiN-coated LTO submicrospheres by solvothermal method and subsequent nitridation process in the presence of ammonia [186]. With a TiN coating layer 5 nm thick, the anode material delivered a discharge capacity of 101 mAh g−1 at a 20C rate.
Doping is one of the best and most cost-effective ways to improve the performance of LTO as an anode material. An extensive review of the doping strategy of LTO has been published by Ezhyeh et al., concluding that among the metal dopants Al, Ca, Co, Fe, Nb, Ru, V, W, and Zn present the highest discharge capacity in the range of 190–340 mAh g−1, while in non-metal candidates N with discharge capacity of 226 mAh g−1 shows the best result and finally Al/Mn as a good performance of co-doping displays discharge capacity about 230 mAh g−1 [37]. We can also mention Cu2+ and F co-doping. The best results were obtained with Li4Ti4.75Cu0.25O11.7F0.3, which delivered a capacity of 134.7 mAh g−1 at 1C with a capacity retention of 97.7 % after 100 cycles [523]. Nevertheless, even better results were obtained recently by Ali et al. who succeeded in doping LTO with Cd2+ at the Li(8a) site to obtain Li4–xCdxTi5O12 prepared by a solid-state route up to the limit of x = 0.20. The Cd dopant is at the Li(8a) site in the LTO lattice. At rate 2C, the half-cell obtained with x = 0.20 delivered an initial capacity of 106.8 mAh g−1, with 98.62% retention after 300 cycles [524]. Noreochim et al. also used the solid-state route to prepare lithium iodide-doped LTO by adding 0.2 mol of LiI to the precursors of LTO based on 1 mol LTO. This I-doped LTO delivered a capacity of 123 mAh g−1 at 15C. At 1C, the cell delivered a discharge capacity of 171 mAh g−1 with a capacity retention of 99.15% after 100 cycles [525]. Synergetic effects can be obtained by the synthesis of composites made of doped LTO and a conductive material.
Qasim and Mousa used a solid-state method to produce pure LTO and Mg, Mn, and V-doped LTO nanoparticles, as well as a chemical polymerization method to synthesize polyaniline (PANI) and composites of pure and doped LTO@PANI nanoparticles. The V-LTO@PANI demonstrated the highest performance with pseudo-capacitive behavior. It demonstrated a specific capacitance of 202 F g−1, an energy density of 72.8 Wh kg−1, a power density of 2430 W kg−1, and 82.6% capacitance retained over 3000 cycles at 1 A g−1 [526]. Self-doping is also a promising strategy to improve the LTO properties. In. particular, Yeo et al. synthesized porous zeolitic LTO (Z-LTO) microspheres using TiO2 and LiOH∙H2O via a hydrothermal process combined with Ar/H2 thermal treatment. The increased concentration of Ti3+ self-doping-derived oxygen vacancies improved significantly the electronic conductivity and structural stability, as the anode delivered a capacity of 181 mAh g−1 at a 5C rate after 2000 cycles, corresponding to 90% capacity retention [527].
Tian et al. prepared in-situ growing Ti3C2Tx MXene/LTO (M-LTO) nanocomposite via controllable natural oxidation and hydrothermal lithiation method. The flake-like Ti3C2Tx MXene was used as both the titanium source and conductive skeletons to support LTO. In the process, the monodispersed LTO nanoparticles grew evenly on the surface of Ti3C2Tx MXene. The in situ growing M-LTO composites resulted in an oriented growth between LTO and Ti3C2Tx MXene, i.e., LTO [110]//Ti3C2Tx [001], which improved the electrical conductivity. As a result, the MXene/M-LTO) exhibited a much superior rate performance than the mixture of Ti3C2Tx MXene and LTO or pristine LTO. At 10C, a reversible capacity of 137 mAh g−1 was obtained, with a capacity retention of 87.5% after 1000 cycles [528]. Li et al. reported a different approach to utilizing Ti3C2 in batteries, in which Ti3C2 only serves as an intermediate for the synthesis of LTO. The Ti3C2 with extracted Ti-ion served as the raw material for a one-step hydrothermal process. The resultant LTO product (M-LTO) inherited the original morphology of Ti3C2 and contained the in-situ formed carbon quantum dots increasing the conductivity. At 5C, this anode delivered an initial capacity of 175.4 mAh, maintained at 160.7 mAh g−1 after 5000 cycles, corresponding to a capacity retention of 91.3%. It still delivered capacities of 168.8 and 168.4 mAh g−1 at 30 and 40C, respectively. At the low rate of 0.5C, the capacity is barely affected when the temperature is lowered down to −30 °C [529].
Wang et al. employed the industrial metatitanic acid (H2TiO3, HTO) as the Ti source to be wrapped with the carbon source first, and then mixed with the lithium source, followed by only one sintering process to prepare LTO. The low-cost glucose was introduced to pre-coat HTO, which could not only avoid the generation of Li2SO4 impurity but also suppress the particle growth of LTO. The amorphous crystal structure of HTO is more conducive for the Li+ insertion, making it possible to decrease the calcination temperature to 700 °C, which is hard to achieve for TiO2 as the Ti source in the traditional route. As a result, the LTO particles were uniformly coated with carbon. As an anode, the composite delivered the specific capacity of 132.9 mAh g−1 is retained even under the current rate of 20C. At 2C, the capacity was 157 mAh g−1 with a capacity retention of 84.8% after 1000 cycles [184]. This synthesis process reduces importantly the fabrication cost of the LTO anode.
Meng et al. used the flame melting method to prepare oxygen-vacancy-rich LTO hollow macrospheres of 40–80 μm in diameter, providing a tap density of 1.26 g cm−3. The synthesis proceeded in three steps: i) Li2Ti3O7: Li4Ti5O12 and Li2Ti3O7 were prepared by a conventional solid-state method. ii) Li2Ti3O7 hollow microspheres were prepared by the flame melting method. iii) The Li2Ti3O7 hollow macrospheres were heated under an Ar atmosphere at 600–900 °C for 5 h. This material delivered excellent low-temperature performance even at a high mass loading (5 g cm−2). At −20 °C, the specific capacities achieved at 1C and 5C are 136.5 and 107.6 mAh g−1, respectively, stable over the 500 cycles that have been tested [530]. This synthesis provides insight into the design of an LTO anode for high mass loading at low temperatures, which outperforms the conventional LTO prepared by solid-state reaction.

3. Concluding Remarks

In this article, we have presented a comprehensive overview of the procedures and different strategies used for the synthesis of Li4Ti5O12 anode materials. Spinel Li4Ti5O12 can be synthesized by many different techniques: solid-state reaction, sol–gel methods, microwave-assisted synthesis, spray pyrolysis, and in hydrothermal batch reactors. The sol–gel method can be employed to prepare particles with uniform composition distributions at low temperatures. Spray pyrolysis enables the direct synthesis of powder from precursor solutions, using low processing temperatures and resulting in high purity and homogeneity of the as-synthesized material. However, it is difficult to control the morphology and chemical composition of the products. Hydrothermal synthesis is known for allowing the preparation of nanoparticles at low temperatures, but the products usually have low crystallinity, which might cause a fast decrease in the capacity of nanoparticulate LTO. In addition, organic residues that appear when using organic solvents result in low Coulombic efficiency. For practical applications, the solid-state method is the most widely used method since it is easy to scale up, and the precursors are cheap and abundant.
We show that optimized LTO is a very good electrode material exhibiting excellent reversible lithium-ion intercalation and de-intercalation processes during charge–discharge cycling without structural change (zero-strain insertion material). Additionally, unlike graphite, LTO exhibits an attractive property for the LIB anode since it operates within the stable electrolyte voltage window with reduced solid electrolyte interphase formation. LTO anode has favorable characteristics for automotive applications being explored intensively. Moreover, recent prospects have shown that LTO is a promising candidate as anode material for SIBs [531]. At present, the strategies to improve the sodium storage performance of LTO include: (i) surface coating and ion doping to increase the ion diffusion rate and electron conductivity and to alleviate the lattice distortion during sodium encapsulation; (ii) designing nano-sized LTO materials to improve the properties of materials by shortening the ion diffusion distance and increasing the contact area with electrolyte.
It has been long recognized that the complete electrode microstructure is decisive for the charge transport and for this reason, research effort is aimed at controlling the particle morphology through the choice of an adequate strategy for the material synthesis. Additionally, the best technique must be relatively cost-effective for commercial applications. There is a general trend showing that nanosized LTO has a higher capacity and better rate performance; however, its production is expensive. Therefore, a trade-off must be found between the performance and cost of the final product. Various computer models have been proven to be invaluable in studying the effects of particle size on the electrochemical performance of LTO anode materials and have been used to define appropriate design parameters needed for high-power applications [532]. Newman and co-workers utilized a pseudo-2D (P2D) model to characterize the performance of the LTO electrode [533]. A full-cell mathematical model is used to compare the performance of graphite (LixC6) and LTO, building Ragone plots as a function of the electrode porosity. Kashkooli et al. [534] showed that it was possible to optimize a particle size according to the desired operating rate. A model design with adjustable parameters based on monodispersed active electrode particles reflects real battery performance and provides the optimum nano-particle size, which eliminates the performance loss due to the limited mass transportation. Table 18 lists the advantages and drawbacks of LTO synthesis methods.
For industrial production, the solid-state method is the most widely used technique since it is easy to scale up and the precursors are cheap and abundant. As mentioned by several workers, this method often leads to particle agglomeration or sintering [220,230,535]. For most titanium salts, the starting precursors are easily hydrolyzed to form TiO2, which is necessary to heat with lithium salts at high temperatures for extended periods of time to obtain well-crystallized LTO powders. This process is energy-consuming. Furthermore, at high temperatures, the LTO particle sizes grow to several hundred nanometers or even several microns resulting in an increase of the lithium diffusion path within isolated particles and thereby decreases electrode performance.
Solvo/hydrothermal methods using cost-effective precursors are known for allowing the preparation of nanoparticles at low temperatures, but the products usually have low crystallinity. However, they provide nanosized LTO particles exhibiting highly improved rate performances at C-rates as high as 30C and 60C, when compared to LTO prepared by the solid-state reaction method. The sol–gel synthesis or Pechini process is currently employed to synthesize particles with uniform composition distributions at low temperatures. In such processes, Li and Ti cations are normally trapped homogeneously on the atomic scale throughout the polymer matrix and thus the reaction temperature and time are reduced. Using low processing temperatures, spray pyrolysis has the advantage of the direct synthesis of powder from precursor solutions, resulting in high purity and homogeneity of the as-prepared LTO material. However, controlling the morphology and chemical composition of the powder is difficult. The best electrochemical performances are influenced by the small particle size, good particle dispersion, and high specific surface area of the LTO anode.
Note that in this review, we did not discuss the thermodynamic stability region issues because, even though the solid-state and sol–gel reaction processes are suitable for large-scale fabrication, they cannot be used to synthesize LTO with particle sizes <300 nm. Smaller particles are needed to obtain good electrochemical properties. That is why modern high-energy mechanical milling is employed to facilitate pulverization to obtain an LTO powder with nanoscale particle size by mechanical force that produces materials with metastable and nano-crystalline phases. In such a case, the effort paid to nucleation and growth direction related to surface chemistry before ball milling has little effect after the milling. The referee mentions the phase stability of the precursors, but this has been analyzed for instance in Ref. [76], as reported in the paper, with the stabilization of water-sensitive metal alkoxides precursor, and the importance of the presence of carboxylic acids.
In the spray pyrolysis process, precursor salts are dissolved in a solution that allows uniform mixing of each component at the molecular level, leading to high chemical purity in the products. This technique is cost-effective and industrially scalable, providing nanostructured ceramics and composites with high electrochemical performances. Using appropriately chosen precursor salts, spray pyrolysis is expected to decrease heat treatment times, and thus save energy and cost in large-scale production.
Substantial efforts have been devoted to developing nanostructured Li4Ti5O12 and Li4Ti5O12/carbon nanocomposite to improve the rate performance for high-power Li-ion batteries, but there are still several challenges. First, it is often difficult to match the nano-dimension with the optimal crystallinity of the produced material through a low-temperature synthesis method. Second, the nanoparticles will result in low powder tap density, which consequently causes low volumetric energy density of the cell. Third, the interparticle contact resistance remains a main rate limiting factor, which is attributed to the aggregation of nanoparticles. Making Li4Ti5O12/carbon nanocomposites using carbon coating could improve the rate capability, but it also hinders Li-ion diffusion. Thus, the optimal carbon layer should be as thin as possible. Facile methods suitable for large-scale commercial production of this microscale/nanoscale hybrid materials with uniform carbon coating are still requested.
At this stage, it is possible to prepare LTO-based lithium-ion batteries competitively for the next generation of lithium batteries. For instance, Zaghib et al. [18] studied an 18650-type battery (capacity of 800 mAh) using carbon-coated LiFePO4 as a cathode and carbon-free Li4Ti5O12 derived from a solid-state reaction as an anode. This battery with a permanently stable capacity after 20,000 cycles at a charge rate of 10C and discharge rate of 5C was used in a real car with the charge time reduced to 5 min using a three-level charger in parallel (500 V, 125 A). More recently, batteries with LTO anodes and high-voltage cathodes have been proven to be safe with remarkable rate capability and cycle ability [19,20,23]. Already, some companies, such as Chongqi, Zhuhai, and Shenzhen of China, have used batteries comprised of Li4Ti5O12 as the anode in the EV power system which has driven more than 50,000 km and charged more than 2000 times. It gives evidence that it is now possible to prepare nanostructured LTO with high purity and small size dispersion, in synthesis processes that avoid aggregation.
Other issues, however, need to be improved. Most efforts have been focused on the performance of LTO. Doping is a low-cost strategy for this purpose, and so is the synthesis of composites of LTO with a conductive element. However, both strategies should be used simultaneously to search for a synergetic effect by the synthesis of a composite of doped LTO with a conductive material. More research on such synergetic effects is needed. For practical use, some other aspects should now be considered. The cost is an important parameter for the industrialization. In this respect, the liquid-state method and the sol–gel method needing many organic reagents in the process are more expensive than the solid-state process but allow for better control of the size of the nanoparticles. Other parameters are the tap density and the mass loading, considered only in a few works [416,530]. Further investigations should then focus on these parameters to further extend the industrialization application of Li4Ti5O12 in advanced energy storage devices.

Author Contributions

Conception, writing—original draft preparation, C.M.J.; writing—reviewing and editing, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

AAacetic acid
ACactivated carbon
AHTOamorphous hydrous titanium oxide
BETBrunauer-Emmett-Teller
CAcitric acid
CNTcarbon nanotube
CTABcetyl trimethyl ammonium bromide
CTACcetyltrimethyl ammonium chloride
DI waterdeionized water
DMEA 2-dimethylaminoethanol
DTA differential thermal analysis
EDTA ethylene diamine tetraacetic acid
EtOHethanol
FSPflame spray pyrolysis
HEBM high-energy ball milling
HOP hydroxypropyl
HTShydrothermal synthesis
LiAAc lithium acetylacetonate
LIBlithium-ion battery
LiOAc lithium acetate dihydrate
LTHlithium titanate oxide hydrate
MAmalic acid
MeOHmethanol
MWCNT multiwalled carbon nanotube
MWHT microwave-assisted hydrothermal
PAApolyacrylate acid
PANpolyacrylonitrile
PBI-b-PEOpoly(isobutylene)-b-polyethylene oxide
PEGpolyethylene glycol
PFphenol-formaldehyde
PVApolyvinyl alcohol
PVPpolyvinylpyrrolidone
rGOreduced graphene oxide
SDSpray drying
SDBS sodium dodecyl benzene sulfonate
SEIsolid electrolyte interphase
SEM scanning electron microscopy
SGspace group
SIBsodium-ion battery
SORsodium–organic reduction reagent
SSRsolid-state reaction
TEA triethanolamine
TBTtetrabutyl titanate
TBOTtitanium tetrabutoxide
TEA triethanolamine
TETtetraethoxytitanium
TG thermogravimetry analysis
TIP titanium(IV) isopropoxide
TNBT titanium n-butoxide
XRDX-ray diffraction

References

  1. Mauger, A.; Julien, C.M.; Goodenough, J.B.; Zaghib, K. Tribute to Michel Armand: From rocking chair—Li-ion to solid-state lithium batteries. J. Electrochem. Soc. 2020, 167, 070507. [Google Scholar] [CrossRef]
  2. Mauger, A.; Xie, H.; Julien, C.M. Composite anodes for lithium-ion batteries: Status and trends. AIMS Mater. Sci. 2016, 3, 1054–1106. [Google Scholar] [CrossRef]
  3. Jonker, G.H. Compounds in the system Li2O–TiO2 and their stability. In Proceedings of the Third International Symposium on the Reactivity of Solids, Madrid, Spain, 2–6 April 1956; pp. 413–421. [Google Scholar]
  4. Deschanvres, A.; Raveau, B.; Sekkal, Z. Mise en évidence et étude cristallographique d’une nouvelle solution solide de type spinelle Li1+xTi2−xO4, 0 ≤ x ≤ 0.33. Mater. Res. Bull. 1971, 6, 699–704. [Google Scholar] [CrossRef]
  5. Johnston, D.C.; Prakash, H.; Zachariasen, W.H.; Viswanathan, R. High temperature superconductivity in the Li-Ti-O ternary system. Mater. Res. Bull. 1973, 8, 777–784. [Google Scholar] [CrossRef]
  6. Murphy, D.W.; Cava, R.J.; Zahurak, S.M.; Santoro, A. Ternary LixTiO2 phases from insertion reaction. Solid State Ion. 1983, 9–10, 413–417. [Google Scholar] [CrossRef]
  7. Colbow, K.M.; Dahn, J.R.; Haering, R.R. Structure and electrochemistry of the spinel oxides LiTi2O4 and Li4/3Ti5/3O4. J. Power Sources 1989, 26, 397–402. [Google Scholar] [CrossRef]
  8. Ferg, E.; Gummow, R.J.; de Kock, A. Spinel anodes for lithium-ion batteries. J. Electrochem. Soc. 1995, 141, L147–L150. [Google Scholar] [CrossRef]
  9. Amatucci, G.G.; Badway, F.; Du Pasquier, A.; Zheng, T. An asymmetric hybrid nonaqueous energy storage cell. J. Electrochem. Soc. 2001, 148, A930–A939. [Google Scholar] [CrossRef]
  10. Ohzuku, T.; Ueda, A.; Yamamoto, N. Zero-strain material of Li[Li1/3Ti5/3]O4 for rechargeable lithium cells. J. Electrochem. Soc. 1995, 142, 1431–1435. [Google Scholar] [CrossRef]
  11. Sun, X.; Radovanovic, P.V.; Cui, B. Advances in spinel Li4Ti5O12 anode materials for lithium-ion batteries. New J. Chem. 2015, 39, 38–63. [Google Scholar] [CrossRef]
  12. Liu, H.; Zhu, Z.; Huang, J.; He, X.; Chen, Y.; Zhang, R.; Lin, R.; Li, Y.; Yu, S.; Xing, X.; et al. Elucidating the limit of Li insertion into the spinel Li4Ti5O12. ACS Mater. Lett. 2019, 1, 96–102. [Google Scholar] [CrossRef]
  13. Sugiyama, J.; Umegaki, I.; Uyama, T.; McFadden, R.M.L.; Shiraki, S.; Hitosugi, T.; Salman, Z.; Saadanoui, H.; Morris, G.D.; MacFarlane, W.A.; et al. Lithium diffusion in spinel Li4Ti5O12 and LiTi2O4 films detected with 8Li β-NMR. Phys. Rev. B 2017, 96, 094402. [Google Scholar] [CrossRef]
  14. Huang, Q.; Yang, Z.; Mao, J. Mechanisms of the decrease in low-temperature electrochemical performance of Li4Ti5O12-based anode materials. Sci. Rep. 2017, 7, 15292. [Google Scholar] [CrossRef] [PubMed]
  15. Jung, H.G.; Jang, M.W.; Hassoun, J.; Sun, Y.K.; Scrosati, B. A high-rate long-life Li4Ti5O12/Li[Ni0.45Co0.1Mn1.45]O4 lithium-ion battery. Nat. Commun. 2011, 2, 516. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, W.; Choi, D.; Yang, Z. Li-ion battery with LiFePO4 cathode and Li4Ti5O12 anode for stationary energy storage. Metall. Mater. Trans. A 2013, 44, 21–25. [Google Scholar] [CrossRef]
  17. Yang, C.; Hu, H.; Lin, S.J.; Chien, W. Electrochemical performance of V-doped spinel Li4Ti5O12/C composite anode in Li-half and Li4Ti5O12/LiFePO4–full cell. J. Power Sources 2014, 258, 424–433. [Google Scholar] [CrossRef]
  18. Zaghib, K.; Dontigny, M.; Guerfi, A.; Charest, P.; Rodrigues, I.; Mauger, A.; Julien, C.M. Safe and fast-charging Li-ion battery with long shelf life for power applications. J. Power Sources 2011, 196, 3949–3954. [Google Scholar] [CrossRef]
  19. Behi, H.; Karimi, D.; Behi, M.; Jaguemont, J.; Ghanbarpour, M.; Behnia, M.; Berecibar, M.; Van Mierlo, J. Thermal management analysis using heat pipe in the high current discharging of lithium-ion battery in electric vehicles. J. Energy Storage 2020, 32, 101893. [Google Scholar] [CrossRef]
  20. Behi, H.; Karimi, D.; Gandoman, F.H.; Akbarzadeh, M.; Khaleghi, S.; Kalogiannis, T.; Hosen, M.S.; Jaguemont, J.; Van Mierlo, J.; Berecibar, M. PCM assisted heat pipe cooling system for the thermal management of an LTO cell for high-current profiles. Case Stud. Therm. Eng. 2021, 25, 100920. [Google Scholar] [CrossRef]
  21. Behi, H.; Karimi, D.; Kalogiannis, T.; He, J.; Patil, M.S.; Muller, J.-D.; Haider, A.; Mierlo, J.V.; Berecibar, M. Advanced hybrid thermal management system for LTO battery module under fast charging. Case Stud. Therm. Eng. 2022, 33, 101938. [Google Scholar] [CrossRef]
  22. Yang, W.; Zhang, M.; Ma, S.; Luo, Y.; Zhang, B.; Wang, S.; Yan, L.; Tong, Z.; Lu, T.; Zhou, Y.-N.; et al. Li4Ti5O12-based battery energy storage system with dual-phase cathode. Energy Technol. 2023, 11, 2200899. [Google Scholar] [CrossRef]
  23. Piao, N.; Wang, P.-F.; Chen, L.; Deng, T.; Fan, X.; Wang, L.; He, X. Nonflammable all-fluorinated electrolytes enabling high-power and long-life LiNi0.5Mn1.5O4/Li4Ti5O12 lithium-ion batteries. Nano Energy 2023, 105, 108040. [Google Scholar] [CrossRef]
  24. Wolfenstine, J.; Lee, U.; Allen, J.L. Electrical conductivity and rate-capability of Li4Ti5O12 as a function of heat-treatment atmosphere. J. Power Sources 2006, 154, 287–289. [Google Scholar] [CrossRef]
  25. Wolfenstine, J.; Allen, J.L. Electrical conductivity and charge compensation in Ta doped Li4Ti5O12. J. Power Sources 2008, 180, 582–585. [Google Scholar] [CrossRef]
  26. Thackeray, M.M.; Amine, K. Li4Ti5O12 spinel anodes. Nat. Energy 2021, 6, 683. [Google Scholar] [CrossRef]
  27. Vikram Babu, B.; Vijaya Babu, K.; Tewodros Aregai, G.; Seeta Devi, L.; Madhavi Latha, B.; Sushma Reddi, M.; Samatha, K.; Veeraiah, V. Structural and electrical properties of Li4Ti5O12 anode material for lithium-ion batteries. Results Phys. 2018, 9, 284–289. [Google Scholar] [CrossRef]
  28. Wagemaker, M.; Simon, D.R.; Kelder, E.M.; Schoonman, J.; Ringpfeil, C.; Haake, U.; Lützenkirchen-Hecht, D.; Frahm, R.; Mulder, F.M. A kinetic two-phase and equilibrium solid solution in spinel Li4+xTi5O12. Adv. Mater. 2006, 18, 3169–3173. [Google Scholar] [CrossRef]
  29. Borghols, W.J.H.; Wagemaker, M.; Lafont, U.; Kelder, E.M.; Mulder, F.M. Size effects in the Li4+xTi5O12 spinel. J. Am. Chem. Soc. 2009, 131, 17786–17792. [Google Scholar] [CrossRef]
  30. Kavan, L.; Prochazka, J.; Spitler, T.M.; Kalbac, M.; Zukalova, M.T.; Drezen, T.; Gratzel, M. Li insertion into Li4Ti5O12 (spinel): Charge capability vs. particle size in thin-film electrodes. J. Electrochem. Soc. 2003, 150, A1000–A1007. [Google Scholar] [CrossRef]
  31. Zhang, H.; Yang, Y.; Xu, H.; Wang, L.; Lu, X.; He, X. Li4Ti5O12 spinel anode: Fundamentals and advances in rechargeable batteries. InfoMat 2022, 4, e12228. [Google Scholar] [CrossRef]
  32. Yi, T.-F.; Yang, S.-Y.; Xie, Y. Recent advances of Li4Ti5O12 as a promising next generation anode material for high power lithium-ion batteries. J. Mater. Chem. A 2015, 3, 5750–5777. [Google Scholar] [CrossRef]
  33. Zhao, B.; Ran, R.; Liu, M.; Shao, Z. A comprehensive review of Li4Ti5O12-based electrodes for lithium-ion batteries: The latest advancements and future perspectives. Mater. Sci. Eng. R 2015, 98, 1–71. [Google Scholar] [CrossRef]
  34. Zhu, G.N.; Wang, Y.G.; Xia, Y.Y. Ti-Based compounds as anode materials for Li-ion Batteries. Energy Environ. Sci. 2012, 5, 6652–6667. [Google Scholar] [CrossRef]
  35. Yuan, T.; Tan, Z.; Ma, C.; Yang, J.; Ma, Z.-F.; Zheng, S. Challenges of spinel Li4Ti5O12 for lithium-ion battery industrial applications. Adv. Energy Mater. 2017, 7, 1601625. [Google Scholar] [CrossRef]
  36. Yan, H.; Zhang, D.; Lu, Q.; Duo, X.; Sheng, X. A review of spinel lithium titanate (Li4Ti5O12) as electrode material for advanced energy storage devices. Ceram. Int. 2021, 47, 5870–5895. [Google Scholar] [CrossRef]
  37. Ezhyeh, Z.; Khodaei, M.; Torabi, F. Review on doping strategy in Li4Ti5O12 as an anode material for lithium-ion batteries. Ceram. Int. 2023, 49, 7105–7141. [Google Scholar] [CrossRef]
  38. Shen, Y.; Eltzholtz, J.; Iversen, B. Controlling size, crystallinity, and electrochemical performance of Li4Ti5O12 nanocrystals. Chem. Mater. 2013, 25, 5023–5030. [Google Scholar] [CrossRef]
  39. Wang, C.; Wang, S.; Tang, L.; He, Y.-B.; Gan, L.; Li, J.; Du, H.; Li, B.; Lin, Z.; Kang, F. A robust strategy for crafting monodisperse Li4Ti5O12 nanospheres as superior rate anode for lithium ion batteries. Nano Energy 2016, 21, 133–144. [Google Scholar] [CrossRef]
  40. Lin, C. Improvements of Li4Ti5O12 anode material for lithium-ion batteries. Mater. Res. Found. 2017, 12, 77–93. [Google Scholar]
  41. Meng, W.; Xu, Y.; Yan, B. In situ nano-sized spinel Li4Ti5O12 powder fabricated by a one-step roasting process in molten salts. J. Alloys Compd. 2018, 732, 784–791. [Google Scholar] [CrossRef]
  42. Chang-Jian, C.W.; Cho, E.C.; Huang, J.H.; Huang, J.H.; Chou, J.A.; Ho, B.C.; Lee, K.C.; Hsiao, Y.S. Spray-drying synthesis of Li4Ti5O12 microspheres in pilot scale using TiO2 nanosheets as starting materials and their application in high-rate lithium ion battery. J. Alloys Compd. 2019, 773, 376–386. [Google Scholar] [CrossRef]
  43. Shenouda, A.Y.; Murali, K.R. Electrochemical properties of doped lithium titanate compounds and their performance in lithium rechargeable batteries. J. Power Sources 2008, 176, 332–339. [Google Scholar] [CrossRef]
  44. Panero, S.; Reale, P.; Ronci, F.; Rossi Albertini, V.; Scrosati, B. Structural and electrochemical study on Li(Li1/3Ti5/3)O4 anode material for lithium ion batteries. Ionics 2000, 6, 461–465. [Google Scholar] [CrossRef]
  45. Ge, H.; Li, N.; Li, D.; Dai, C.; Wang, D. Electrochemical characteristics of spinel Li4Ti5O12 discharged to 0.01 V. Electrochem. Commun. 2008, 10, 719–722. [Google Scholar] [CrossRef]
  46. Lai, C.; Wu, Z.; Zhu, Y.; Wu, Q.; Liang Li, L.; Wang, C. Ball-milling assisted solid-state reaction synthesis of mesoporous Li4Ti5O12 for lithium-ion batteries anode. J. Power Sources 2013, 226, 71–74. [Google Scholar] [CrossRef]
  47. Chandrasekhar, J.; Dhananjaya, M.; Hussain, O.M.; Mauger, A.; Julien, C.M. Enhanced electrochemical performance of Li4Ti5O12 by niobium doping for pseudocapacitive applications. Micro 2021, 1, 28–42. [Google Scholar] [CrossRef]
  48. Wang, Y.; Zhou, A.; Dai, X.; Feng, L.; Li, J.; Li, J. Solid-state synthesis of submicron-sized Li4Ti5O12/Li2TiO3 composites with rich grain boundaries for lithium ion batteries. J. Power Sources 2014, 266, 114–120. [Google Scholar] [CrossRef]
  49. Sarantuya, L.; Sevjidsuren, G.; Zltantsog, P.; Tsogbadrakh, N. Synthesis, structure and electronic properties of Li4Ti5O12 anode material for lithium ion batteries. Solid State Phenom. 2018, 271, 9–17. [Google Scholar] [CrossRef]
  50. Zhang, E.; Zhang, H. Hydrothermal synthesis of Li4Ti5O12-TiO2 composites and Li4Ti5O12 and their applications in lithium-ion batteries. Ceram. Int. 2019, 45, 7419–7426. [Google Scholar] [CrossRef]
  51. Ge, H.; Chen, L.; Yuan, W.; Zhang, Y.; Fan, Q.; Osgood, H.; Matera, D.; Song, X.-M.; Wu, G. Unique mesoporous spinel Li4Ti5O12 nanosheets as anode materials for lithium-ion batteries. J. Power Sources 2015, 297, 436–441. [Google Scholar] [CrossRef]
  52. Tang, Y.; Yang, L.; Fang, S.; Qiu, Z. Li4Ti5O12 hollow microspheres assembled by nanosheets as an anode material for high-rate lithium ion batteries. Electrochim. Acta 2009, 54, 6244–6249. [Google Scholar] [CrossRef]
  53. Yan, H.; Zhu, Z.; Zhang, D.; Li, W.; Lu, Q. A new hydrothermal synthesis of spherical Li4Ti5O12 anode material for lithium-ion secondary batteries. J. Power Sources 2012, 219, 45–51. [Google Scholar] [CrossRef]
  54. Lee, S.C.; Lee, S.M.; Lee, J.W.; Lee, J.B.; Han, S.S.; Lee, H.C.; Kim, H.J. Spinel Li4Ti5O12 nanotubes for energy storage materials. J. Phys. Chem. C 2009, 113, 18420–18423. [Google Scholar] [CrossRef]
  55. Shen, L.; Yuan, C.; Luo, H.; Zhang, X.; Xu, K.; Xia, Y. Facile synthesis of hierarchically porous microspheres for high rate lithium ion batteries. J. Mater. Chem. 2010, 20, 6998–7004. [Google Scholar] [CrossRef]
  56. Shen, L.; Yuan, C.; Luo, H.; Zhang, X.; Yang, S.; Lu, X. In situ synthesis of high-loading Li4Ti5O12–graphene hybrid nanostructures for high rate lithium ion batteries. Nanoscale 2011, 3, 572–574. [Google Scholar] [CrossRef] [PubMed]
  57. Wang, Y.; Rong, H.; Li, B.; Xing, L.; Li, X.; Li, W. Microemulsion-assisted synthesis of ultrafine Li4Ti5O12/C nanocomposite with oleic acid as carbon precursor and particle size controller. J. Power Sources 2014, 246, 213–218. [Google Scholar] [CrossRef]
  58. Qian, K.; Tang, L.; Wagemaker, M.; He, Y.-B.; Liu, D.; Li, H.; Shi, R.; Li, B.; Kang, F. A facile surface reconstruction mechanism toward better electrochemical performance of Li4Ti5O12 in lithium-ion battery. Adv. Sci. 2017, 4, 1700205. [Google Scholar] [CrossRef] [PubMed]
  59. Li, Y.; Zhao, H.; Tian, Z.; Qiu, W.; Li, X. Solvothermal synthesis and electrochemical characterization of amorphous lithium titanate materials. J. Alloys Compd. 2008, 455, 471–474. [Google Scholar] [CrossRef]
  60. Shen, L.; Yuan, C.; Luo, H.; Zhang, X.; Chen, L.; Li, H. Novel template-free solvothermal synthesis of mesoporous Li4Ti5O12-C microspheres for high power lithium ion batteries. J. Mater. Chem. 2011, 21, 14414–14416. [Google Scholar] [CrossRef]
  61. Bach, S.; Pereira-Ramos, J.P.; Baffier, N. Electrochemical properties of sol–gel Li4/3Ti5/3O4. J. Power Sources 1999, 81–82, 273–276. [Google Scholar] [CrossRef]
  62. Shen, C.M.; Zhang, X.G.; Zhou, Y.K.; Li, H.L. Preparation and characterization of nanocrystalline Li4Ti5O12 by sol–gel method. Mater. Chem. Phys. 2002, 78, 437–441. [Google Scholar] [CrossRef]
  63. Hao, Y.-J.; Lai, Q.-Y.; Lu, J.-Z.; Wang, H.-L.; Chen, Y.-D.; Ji, X.-Y. Synthesis and characterization of spinel Li4Ti5O12 anode material by oxalic acid-assisted sol gel method. J. Power Sources 2005, 158, 1358–1364. [Google Scholar] [CrossRef]
  64. Hao, Y.J.; Lai, Q.Y.; Xu, Z.; Liu, X.; Ji, X. Synthesis of TEA sol-gel method and electrochemical properties of Li4Ti5O12 anode material for lithium-ion battery. Solid State Ion. 2005, 176, 1201–1206. [Google Scholar] [CrossRef]
  65. Hao, Y.J.; Lai, Q.Y.; Lu, J.Z.; Liu, D.Q.; Ji, X.Y. Influence of various complex agents on electrochemical property of Li4Ti5O12 anode material. J. Alloys Compd. 2007, 439, 330–336. [Google Scholar] [CrossRef]
  66. Miao, X.; He, H.; Shi, L.; Zhao, X.; Fang, J. Sol-gel synthesis of nanocomposite/carbon nanotubes as anode materials for high-rate performance lithium-ion batteries. Adv. Mater. Res. 2014, 833, 45–49. [Google Scholar] [CrossRef]
  67. Mosa, J.; Aparicio, M. Sol-gel synthesis of nanocrystalline mesoporous Li4Ti5O12 thin-films as anodes for Li-ion microbatteries. Nanomaterials 2020, 10, 1369. [Google Scholar] [CrossRef]
  68. Priyono, S.; Sofyan, N.; Subhan, A.; Prihandoko, B.; Yuwono, A.H. Preparation of Al-doped Li4Ti5O12 anode material via sol-gel process with acidic catalyst for lithium-ion batteries. AIP Conf. Proc. 2023, 2538, 070003. [Google Scholar]
  69. Mani, J.; Katzke, H.; Habouti, S.; Moonoosawmy, K.R.; Dietze, M.; Es-Souni, M. A template-free synthesis and structural characterization of hierarchically nano-structured lithium-titanium-oxide films. J. Mater. Chem. 2012, 22, 6632–6638. [Google Scholar] [CrossRef]
  70. Jiang, C.H.; Ichihara, M.; Honma, I.; Zhou, H.S. Effect of particle dispersion on high rate performance of nano-sized Li4Ti5O12 anode. Electrochim. Acta 2007, 52, 6470–6475. [Google Scholar] [CrossRef]
  71. Liu, G.; Zhang, R.; Bao, K.; Xie, H.; Zheng, S.; Guo, J.; Liu, G. Synthesis of nano- anode material for lithium ion batteries by a biphasic interfacial reaction route. Ceram. Int. 2016, 42, 11468–11472. [Google Scholar] [CrossRef]
  72. Ju, S.H.; Kang, Y.C. Effects of preparation conditions on the electrochemical and morphological characteristics of Li4Ti5O12 powders prepared by spray pyrolysis. J. Power Sources 2009, 189, 185–190. [Google Scholar] [CrossRef]
  73. Ju, S.H.; Kang, Y.C. Characteristics of spherical-shaped Li4Ti5O12 anode powders prepared by spray pyrolysis. J. Phys. Chem. Solids 2009, 70, 40–44. [Google Scholar] [CrossRef]
  74. Ju, S.H.; Kang, Y.C. Effects of types of drying control chemical additives on the morphologies and electrochemical properties of Li4Ti5O12 anode powders prepared by spray pyrolysis. J. Alloys Compd. 2010, 506, 913–916. [Google Scholar] [CrossRef]
  75. Ernst, F.O.; Kammler, H.K.; Roessler, A.; Pratsinis, S.E.; Stark, W.J.; Ufheil, J.; Novak, P. Electrochemically active flame-made nanosized spinels: LiMn2O4, Li4Ti5O12 and LiFe5O8. Mater. Chem. Phys. 2007, 101, 372–378. [Google Scholar] [CrossRef]
  76. Meierhofer, F.; Li, H.; Gockeln, M.; Kun, R.; Grieb, T.; Rosenauer, A.; Fritsching, U.; Kiefer, J.; Birkenstock, J.; Madler, L.; et al. Screening precursor−solvent combinations for Li4Ti5O12 energy storage material using flame spray pyrolysis. ACS Appl. Mater. Interfaces 2017, 9, 37760–37777. [Google Scholar] [CrossRef]
  77. Gockeln, M.; Pokhrel, S.; Meierhofer, F.; Glenneberg, J.; Schowalter, M.; Rosenauer, A.; Fritsching, U.; Busse, M.; Mädler, L.; Kun, R. Fabrication and performance of Li4Ti5O12/C Li-ion battery electrodes using combined double flame spray pyrolysis and pressure–based lamination technique. J. Power Sources 2018, 374, 97–106. [Google Scholar] [CrossRef]
  78. Terechshenko, A.; Sanbayeva, A.; Babaa, M.R.; Nurpeissova, A.; Bakenov, Z. Spray-pyrolysis preparation of Li4Ti5O12/Si composites for lithium-ion batteries. Eurasian Chem. Technol. J. 2019, 21, 69–73. [Google Scholar] [CrossRef]
  79. Xie, Z.; Song, Q.; Xie, H.; Yin, H.; Ning, Z. Chemically driven synthesis of Ti3+ self-doped Li4Ti5O12 spinel in molten salt. J. Am. Ceram. Soc. 2021, 104, 753–765. [Google Scholar] [CrossRef]
  80. Bai, Y.; Wang, F.; Wu, F.; Wu, C.; Bao, L. Influence of composite LiCl–KCl molten salt on microstructure and electrochemical performance of spinel Li4Ti5O12. Electrochim. Acta 2008, 54, 322–327. [Google Scholar] [CrossRef]
  81. Sharmila, S.; Senthilkumar, B.; Kalai Selvan, R. Molten-salt synthesis and characterization of Li4Ti5O12. In AIP Conference Proceedings, Proceedings of the 55th DAE Solid State Physics; American Institute of Physics: College Park, MD, USA, 2011; Volume 1349, pp. 1325–1326. [Google Scholar]
  82. Guo, Q.; Li, S.; Wang, H.; Gao, Y.; Li, B. Molten salt synthesis of nano-sized Li4Ti5O12 doped with Fe2O3 for use as anode material in the lithium-ion battery. RSC Adv. 2014, 4, 60327–60333. [Google Scholar] [CrossRef]
  83. Xue, X.; Yan, H.; Fu, Y. Preparation of pure and metal-doped Li4Ti5O12 composites and their lithium-storage performances for lithium-ion batteries. Solid State Ion. 2019, 335, 1–6. [Google Scholar] [CrossRef]
  84. Kim, H.K.; Jegal, J.-P.; Kim, J.-Y.; Yoon, S.-B.; Roh, K.C.; Kim, K.-B. In situ fabrication of lithium titanium oxide by microwave-assisted alkalization for high-rate lithium-ion batteries. J. Mater. Chem. A 2013, 1, 14849–14852. [Google Scholar] [CrossRef]
  85. He, N.; Wang, B.; Huang, J. Preparation and electrochemical performance of monodisperse Li4Ti5O12 hollow spheres. J. Solid State Electrochem. 2010, 14, 1241–1246. [Google Scholar] [CrossRef]
  86. Yu, L.; Wu, H.B.; Lou, X.W. Mesoporous Li4Ti5O12 hollow spheres with enhanced lithium storage capability. Adv. Mater. 2013, 25, 2296–2300. [Google Scholar] [CrossRef] [PubMed]
  87. Kawade, U.V.; Jayswal, M.S.; Ambalkar, A.A.; Kadam, S.R.; Panmand, R.P.; Ambekar, J.D.; Kulkarni, M.V.; Kale, B.B. Surface modified Li4Ti5O12 by paper templated approach for enhanced interfacial Li+ charge transfer in Li-ion batteries. RSC Adv. 2018, 8, 38391–38399. [Google Scholar] [CrossRef]
  88. Hermawan, A.; Wibowo, A.; Asri, L.A.T.W.; Shu Yin, S.; Purwasasmita, B.S. Improved ionic conductivity of porous Li4Ti5O12 synthesized by sol-gel method using eggshell membrane as soft template. Mater. Res. Express 2019, 6, 075030. [Google Scholar] [CrossRef]
  89. Kanamura, K.; Chiba, T.; Dokko, K. Preparation of Li4Ti5O12 spherical particles for rechargeable lithium batteries. J. Eur. Ceram. Soc. 2006, 26, 577–581. [Google Scholar] [CrossRef]
  90. Kim, D.H.; Ahn, Y.S.; Kim, J. Polyol-mediated synthesis of Li4Ti5O12 nanoparticles and its electrochemical properties. J. Electrochem. Commun. 2005, 7, 1340–1344. [Google Scholar] [CrossRef]
  91. Kataoka, K.; Takahashi, Y.; Kijima, N.; Akimoto, J.; Ohshima, K. Single crystal growth and structure refinement of Li4Ti5O12. J. Phys. Chem. Solids 2008, 69, 1454–1456. [Google Scholar] [CrossRef]
  92. Wang, G.; Xu, J.; Wen, M.; Cai, R.; Ran, R.; Shao, Z. Influence of high-energy ball milling of precursor on the morphology and electrochemical performance of Li4Ti5O12–Ball-milling time. Solid State Ion. 2008, 179, 946–950. [Google Scholar] [CrossRef]
  93. Han, S.-W.; Jeong, J.; Yoon, D.-H. Effects of high-energy milling on the solid-state synthesis of pure nano-sized Li4Ti5O12 for high power lithium battery applications. Appl. Phys. A 2014, 114, 925–930. [Google Scholar] [CrossRef]
  94. Li, Y.; Xie, H.; Li, J.; Wang, J. Mechanochemical synthesis and electrochemical performances of Li4Ti5O12 anode materials for lithium-ion batteries. Adv. Mater. Res. 2012, 512–515, 1660–1663. [Google Scholar]
  95. Lu, H.W.; Zeng, W.; Li, Y.S.; Fu, Z.W. Fabrication and electrochemical properties of three-dimensional net architectures of anatase TiO2 and spinel Li4Ti5O12 nanofibers. J. Power Sources 2007, 164, 874–879. [Google Scholar] [CrossRef]
  96. Castano, N.; Cortes, M.A.; Garcia, E.; Martinez, H.V. Synthesis and morphological characterization of Li-Ti/PVP fibers as precursors for Li4Ti5O12 towards its future use as anode materials in Li-ion batteries by means of electrospinning. IOP Conf. Ser. Mater. Sci. Eng. 2018, 437, 012016. [Google Scholar] [CrossRef]
  97. Prakash, A.S.; Manikandan, P.; Ramesha, K.; Sathiya, M.; Tarascon, J.M.; Shukla, A.K. Solution–combustion synthesized nanocrystalline Li4Ti5O12 as high-rate performance Li-ion battery anode. Chem. Mater. 2010, 22, 2857–2863. [Google Scholar] [CrossRef]
  98. Cai, R.; Yu, X.; Liu, X.; Shao, Z. Li4Ti5O12/Sn composite anodes for lithium-ion batteries: Synthesis and electrochemical performance. J. Power Sources 2010, 195, 8244–8250. [Google Scholar] [CrossRef]
  99. Yuan, T.; Cai, R.; Wang, K.; Ran, R.; Liu, S.; Shao, Z. Combustion synthesis of high-performance Li4Ti5O12 for secondary Li-ion battery. Ceram. Int. 2009, 35, 1757–1768. [Google Scholar] [CrossRef]
  100. De Sloovere, D.; Marchal, W.; Ulu, F.; Vranken, T.; Verheijen, M.; Van Bael, M.K.; Hardy, A. Combustion synthesis as a low temperature route to Li4Ti5O12 based powders for lithium ion battery anodes. RSC Adv. 2017, 7, 18745–18754. [Google Scholar] [CrossRef]
  101. Yuan, T.; Wang, K.; Cai, R.; Ran, R.; Shao, Z. Cellulose-assisted combustion synthesis of Li4Ti5O12 adopting anatase TiO2 solid as raw material with high electrochemical performance. J. Alloys Compd. 2009, 477, 665–672. [Google Scholar] [CrossRef]
  102. Yuan, T.; Cai, R.; Gu, P.; Shao, Z. Synthesis of lithium insertion material Li4Ti5O12 from rutile TiO2 via surface activation. J. Power Sources 2010, 195, 2883–2887. [Google Scholar] [CrossRef]
  103. Raja, M.W.; Mahanty, S.; Kundu, M.; Basu, R.N. Synthesis of nanocrystalline Li4Ti5O12 by a novel aqueous combustion technique. J. Alloys Compd. 2009, 468, 258–262. [Google Scholar] [CrossRef]
  104. Lee, S.S.; Byun, K.-T.; Park, J.P.; Kim, S.K.; Kwak, H.-Y.; Shim, I.-W. Preparation of nanoparticles by a simple sonochemical method. Dalton Trans. 2007, 37, 4182–4184. [Google Scholar] [CrossRef]
  105. Ghosh, S.; Mitra, S.; Barpanda, P. Sonochemical synthesis of nanostructured spinel Li4Ti5O12 negative insertion material for Li-ion and Na-ion batteries. Electrochim. Acta 2016, 222, 898–903. [Google Scholar] [CrossRef]
  106. Ni, H.; Song, W.-L.; Fan, L.-Z. A strategy for scalable synthesis of Li4Ti5O12/reduced graphene oxide toward high rate lithium-ion batteries. Electrochem. Commun. 2014, 40, 1–4. [Google Scholar] [CrossRef]
  107. Jin, Y.-H.; Min, K.-M.; Shim, H.-W.; Seo, S.-D.; Hwang, I.-S.; Park, K.-S.; Kim, D.W. Facile synthesis of nano- for high-rate Li-ion battery anodes. Nanoscale Res. Lett. 2012, 7, 10. [Google Scholar] [CrossRef]
  108. Mao, S.; Huang, X.; Chang, J.; Cui, S.; Zhou, G.; Chen, J. One-step continuous synthesis of spherical Li4Ti5O12/graphene composite as an ultra-long cycle life lithium-ion battery anode. NPG Asia Mater. 2015, 7, e224. [Google Scholar] [CrossRef]
  109. Tang, B.; Li, A.; Tong, Y.; Song, H.; Chen, X.; Zhou, J.; Ma, Z. Carbon-coated Li4Ti5O12 tablets derived from metal-organic frameworks as anode material for lithium-ion batteries. J. Alloys Compd. 2017, 708, 6–13. [Google Scholar] [CrossRef]
  110. 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, 5980–5984. [Google Scholar] [CrossRef]
  111. Wen, Z.; Gu, Z.; Huang, S.; Yang, J.; Lin, Z.; Yamamoto, O. Research on spray-dried lithium titanate as electrode materials for lithium ion batteries. J. Power Sources 2005, 146, 670–673. [Google Scholar] [CrossRef]
  112. Wu, X.; Xinghua Liang, X.; Zhang, X.; Lan, X.; Li, L.; Suo Gai, Q. Structural evolution of plasma sprayed amorphous Li4Ti5O12 electrode and ceramic/polymer composite electrolyte during electrochemical cycle of quasi-solid-state lithium battery. J. Adv. Ceram. 2021, 10, 347–354. [Google Scholar] [CrossRef]
  113. Yin, S.Y.; Song, L.; Wang, X.Y.; Zhang, M.F.; Zhang, K.L.; Zhang, Y.X. Synthesis of spinel Li4Ti5O12 anode material by a modified rheological phase reaction. Electrochim. Acta 2009, 54, 5629–5633. [Google Scholar] [CrossRef]
  114. Nugroho, A.; Yoon, D.; Joo, O.-S.; Chung, K.Y.; Kim, J. Continuous synthesis of Li4Ti5O12 nanoparticles in supercritical fluids and their electrochemical performance for anode in Li-ion batteries. Chem. Eng. J. 2014, 258, 357–366. [Google Scholar] [CrossRef]
  115. Doi, T.; Iriyama, Y.; Abe, T.; Ogumi, Z. Electrochemical insertion and extraction of lithium ion at uniform nanosized Li4/3Ti5/3O4 particles prepared by a spray pyrolysis method. Chem. Mater. 2005, 17, 1580–1582. [Google Scholar] [CrossRef]
  116. Naoi, K.; Ishimoto, S.; Isobe, Y.; Aoyagi, S. High-rate nano-crystalline Li4Ti5O12 attached on carbon nano-fibers for hybrid supercapacitors. J. Power Sources 2010, 195, 6250–6254. [Google Scholar] [CrossRef]
  117. Bresser, D.; Paillard, E.; Copley, M.; Bishop, P.; Winter, M.; Passerini, S. The importance of “going nano” for high power battery materials. J. Power Sources 2012, 219, 217–222. [Google Scholar] [CrossRef]
  118. Naoi, K. Nanohybrid capacitor: The next generation electrochemical capacitors. Fuel Cells 2010, 10, 825–833. [Google Scholar] [CrossRef]
  119. Liu, Y.; Liu, J.Y.; Hou, M.Y.; Fan, L.; Wang, Y.; Xia, Y. Carbon-coated Li4Ti5O12 nanoparticles with high electrochemical performance as anode material in sodium-ion batteries. J. Mater. Chem. A 2017, 5, 10902–10908. [Google Scholar] [CrossRef]
  120. Jiang, C.H.; Hosono, E.; Ichihara, M.; Honma, I.; Zhou, H.S. Synthesis of nanocrystalline Li4Ti5O12 by chemical lithiation of anatase nanocrystals and postannealing. J. Electrochem. Soc. 2008, 155, A553–A556. [Google Scholar] [CrossRef]
  121. Lu, J.; Nan, C.; Peng, Q.; Li, Y. Single crystalline lithium titanate nanostructure with enhanced rate performance for lithium ion battery. J. Power Sources 2012, 202, 246–252. [Google Scholar] [CrossRef]
  122. Song, K.; Seo, D.-H.; Jo, M.R.; Kim, Y.-I.; Kang, K.; Kang, Y.-M. Tailored oxygen framework of Li4Ti5O12 nanorods for high-power Li ion battery. J. Phys. Chem. Lett. 2014, 5, 1368–1373. [Google Scholar] [CrossRef] [PubMed]
  123. Li, Y.; Pan, G.L.; Liu, J.W.; Gao, X.P. Preparation of Li4Ti5O12 nanorods as anode materials for lithium-ion batteries. J. Electrochem. Soc. 2009, 156, 495–499. [Google Scholar] [CrossRef]
  124. Zhou, Q.; Liu, L.; Tan, J.L.; Tan, Z.; Huang, Z.; Wang, X. Synthesis of lithium titanate nanorods as anode materials for lithium and sodium ion batteries with superior electrochemical performance. J. Power Sources 2015, 283, 243–250. [Google Scholar] [CrossRef]
  125. Priyono, B.; Herwono, M.F.; Syahrial, A.Z.; Nugraha, M.R.; Faizah; Subhan, A. Enhancing lithium titanite (Li4Ti5O12) nanorods performance with graphite and nano tin as anode for lithium-ion batteries. In AIP Conference Proceedings; AIP Publishing: College Park, MD, USA, 2020; Volume 2232, p. 030008. [Google Scholar]
  126. Luo, H.; Shen, L.; Rui, K.; Li, H.; Zhang, X. Carbon coated Li4Ti5O12 nanorods as superior anode material for high rate lithium ion batteries. J. Alloys Compd. 2013, 572, 37–42. [Google Scholar] [CrossRef]
  127. Wang, W.; Guo, Y.; Liu, L.; Wang, S.; Yang, X.; Guo, H. Gold coating for a high performance nanorod aggregates anode in lithium-ion batteries. J. Power Sources 2014, 245, 624–629. [Google Scholar] [CrossRef]
  128. Li, Y.; Song, J.; Tian, Q. Li4Ti5O12 nanowires intertwined with carbon nanotubes for ultra-long life and conductive additive-free anodes of lithium-ion batteries. Mater. Lett. 2023, 348, 134698. [Google Scholar] [CrossRef]
  129. Hu, G.; Wu, J.; Du, K.; Peng, Z.; Jia, M.; Yang, H.; Cao, Y. Surface-fluorinated Li4Ti5O12 nanowires/reduced graphene oxide composite as a high-rate anode material for lithium ion batteries. Appl. Surf. Sci. 2019, 479, 158–166. [Google Scholar] [CrossRef]
  130. Li, J.R.; Tang, Z.L.; Zhang, Z.T. Controllable formation and electrochemical properties of one-dimensional nanostructured spinel Li4Ti5O12. Electrochem. Commun. 2005, 7, 894–899. [Google Scholar] [CrossRef]
  131. Bachtiar, A.R.; Syahrial, A.Z.; Nugraha, M.R.; Faizah; Subhan, A.; Priyono, B. Enhancing performance of Li4Ti5O12 nanowire with addition of graphite and ZnO nanoparticle as anode for lithium-ion batteries. In AIP Conference Proceedings; AIP Publishing: College Park, MD, USA, 2020; Volume 2232, p. 030003. [Google Scholar]
  132. Kim, T.-T.; Yu, C.-Y.; Yoon, C.S.; Kim, S.-J.; Sun, Y.-K.; Myung, S.-T. Carbon-coated Li4Ti5O12 nanowires showing high rate capability as an anode material for rechargeable sodium batteries. Nano Energy 2015, 12, 725–734. [Google Scholar] [CrossRef]
  133. Liu, H.; Tank, K.; Song, K.; van Aken, P.A.; Yu, Y.; Maier, J. Tiny Li4Ti5O12 nanoparticles embedded in carbon nanofibers as high-capacity and long-life anode materials for both Li-ion and Na-ion batteries. Phys. Chem. Chem. Phys. 2013, 15, 20813–20818. [Google Scholar] [CrossRef] [PubMed]
  134. Jia, X.; Lu, Y.; Wei, F. Confined growth of Li4Ti5O12 nanoparticles in nitrogen-doped mesoporous graphene fibers for high-performance lithium-ion battery anodes. Nano Res. 2016, 9, 230–239. [Google Scholar] [CrossRef]
  135. Xu, H.; Hu, X.; Sun, Y.; Luo, W.; Chen, C.; Liu, Y.; Huang, Y. Highly porous Li4Ti5O12/C nanofibers for ultrafast electrochemical energy storage. Nano Energy 2014, 10, 163–171. [Google Scholar] [CrossRef]
  136. Bian, M.; Yang, Y.; Tian, L. Carbon-free Li4Ti5O12 porous nanofibers as high-rate and ultralong-life anode materials for lithium-ion batteries. J. Phys. Chem. Solids 2018, 113, 11–16. [Google Scholar] [CrossRef]
  137. Jo, M.R.; Jung, Y.S.; Kang, Y.-M. Tailored Li4Ti5O12 nanofibers with outstanding kinetics for lithium rechargeable batteries. Nanoscale 2012, 4, 6870–6875. [Google Scholar] [CrossRef]
  138. Zhao, Y.; Li, J.; Li, Z.; Yang, K.; Gao, F. Pr-modified Li4Ti5O12 nanofibers as an anode material for lithium-ion batteries with outstanding cycling performance and rate performance. Ionics 2017, 23, 597–605. [Google Scholar] [CrossRef]
  139. Ji, X.; Lu, Q.; Guo, E.; Li, D.; Yao, L.; Liu, H.; Li, X. Bamboo-shaped Zn2+-doped Li4Ti5O12 nanofibers: One-step controllable synthesis and high-performance lithium-ion batteries. J. Electrochem. Soc. 2018, 165, A534–A541. [Google Scholar] [CrossRef]
  140. Xu, H.; Hu, X.; Luo, W.; Sun, Y.; Yang, Z.; Hu, C.; Huang, Y. Electrospun conformal Li4Ti5O12/C fibers for high-rate lithium-ion batteries. ChemElectroChem 2014, 1, 611–616. [Google Scholar] [CrossRef]
  141. Yagi, S.; Morinaga, T.; Togo, M.; Tsuda, H.; Shio, S.; Nakahira, A. Ion-exchange synthesis of Li4Ti5O12 nanotubes and nanoparticles for high-rate Li-ion batteries. Mater. Trans. 2016, 57, 42–45. [Google Scholar] [CrossRef]
  142. Jiang, Y.-M.; Wang, K.-X.; Wu, X.-Y.; Zhang, H.-J.; Bartlett, B.M.; Chen, J.-S. Li4Ti5O12/TiO2 hollow spheres composed nanoflakes with preferentially exposed Li4Ti5O12 (011) facets for high-rate lithium ion batteries. ACS Appl. Mater. Interfaces 2014, 6, 19791–19796. [Google Scholar] [CrossRef]
  143. Zhang, X.; Xu, W.; Li, X.; Zhong, X.; Liu, W.; Lin, Y.; Xia, R. Li4Ti5O12/Ti4O7/carbon nanotubes composite anode material for lithium-ion batteries. Micro Nano Lett. 2018, 13, 915–918. [Google Scholar] [CrossRef]
  144. Wang, H.; Li, S.; Yang, Y.; Yu, W.; Ma, Q.; Dong, X.; Wang, J.; Liu, G. Electrochemical characteristics of Li4Ti5O12/Ag composite nanobelts prepared via electrospinning. Russ. J. Phys. Chem. A 2019, 93, 144–150. [Google Scholar] [CrossRef]
  145. Qin, W.; Chen, Y.; An, J.; Zhang, J.; Wen, X. High-loaded nanobelt-array/nanobelt-microsphere multilayer Li4Ti5O12 self-supported on Ti foils for high-performance lithium ion batteries. Electrochim. Acta 2022, 419, 140407. [Google Scholar] [CrossRef]
  146. Tang, Y.F.; Yang, L.; Qiu, Z.; Huang, J.S. Preparation and electrochemical lithium storage of flower-like spinel Li4Ti5O12 consisting of nanosheets. Electrochem. Commun. 2008, 10, 1513–1516. [Google Scholar] [CrossRef]
  147. Hong, Z.S.; Lan, T.B.; Xiao, F.Y.; Zhang, H.X.; Wei, M.D. Ultrafine Li4Ti5O12 nanosheets as a high performance anode for Li-ion battery. Funct. Mater. Lett. 2011, 4, 389–393. [Google Scholar] [CrossRef]
  148. Wu, L.; Leng, X.; Liu, Y.; Wei, S.; Li, C.; Wang, G.; Lian, J.; Jiang, Q.; Nie, A.; Zhang, T.-Y. A strategy for synthesis of nanosheets consisting of alternating spinel Li4Ti5O12 and rutile TiO2 lamellas for high-rate anodes of lithium-ion batteries. ACS Appl. Mater. Interfaces 2017, 9, 4649–4657. [Google Scholar] [CrossRef] [PubMed]
  149. Xu, G.B.; Tian, Y.; Wei, X.L. Free-standing electrodes composed of carbon-coated Li4Ti5O12 nanosheets and reduced graphene oxide for advanced sodium ion batteries. J. Power Sources 2017, 337, 180–188. [Google Scholar] [CrossRef]
  150. Lai, C.; Dou, Y.Y.; Li, X.; Gao, X.P. Improvement of the high rate capability of hierarchical structured Li4Ti5O12 induced by the pseudocapacitive effect. J. Power Sources 2010, 195, 3676–3679. [Google Scholar] [CrossRef]
  151. Sha, Y.; Zhao, B.; Ran, R.; Cai, R.; Shao, Z. Synthesis of well-crystalized Li4Ti5O12 nanoplates for lithium-ion batteries with outstanding rate capability and cycling stability. J. Mater. Chem. A 2013, 1, 13233–13243. [Google Scholar] [CrossRef]
  152. Liu, J.; Wei, X.; Liu, X.-W. Two-dimensional wavelike spinel lithium titanate for fast lithium storage. Sci. Rep. 2015, 5, 9782. [Google Scholar] [CrossRef]
  153. Salvatore, K.L.; Vila, M.N.; Renderos, G.; Li, W.; Housel, L.M.; Tong, X.; McGuire, S.C.; Gan, J.; Paltis, A.; Lee, K.; et al. Probing the physicochemical behavior of variously doped Li4Ti5O12 nanoflowers. ACS Phys. Chem. Au 2022, 2, 331–345. [Google Scholar] [CrossRef] [PubMed]
  154. Lv, Y.; Zhang, H.; Cao, G.P.; Wang, B.Y.; Wang, X.D. Phenol-formaldehyde resin-assisted synthesis of pure porous Li4Ti5O12 for rate capability improvement. Mater. Res. Bull. 2011, 46, 2312–2316. [Google Scholar] [CrossRef]
  155. Kang, E.; Jung, Y.S.; Kim, G.-H.; Chun, J.; Wiesner, U.; Dillon, A.C.; Kim, J.K.; Lee, J. Highly improved rate capability for a lithium-ion battery nano-Li4Ti5O12 negative electrode via carbon-coated mesoporous uniform pores with a simple self-assembly method. Adv. Funct. Mater. 2011, 21, 4349–4357. [Google Scholar] [CrossRef]
  156. Saxena, S.; Sil, A. Nanoporous Li4Ti5O12 material for the electrode of lithium ion battery. IETE Techn. Rev. 2016, 33, 60–63. [Google Scholar] [CrossRef]
  157. Shao, D.; He, J.; Luo, Y.; Liu, W.; Yu, X.; Fang, Y. Synthesis and electrochemical performance of nanoporous Li4Ti5O12 anode material for the lithium-ion batteries. J. Solid State Electrochem. 2012, 16, 2047–2053. [Google Scholar] [CrossRef]
  158. Wang, H.; Wang, L.; Lin, J.; Yang, J.; Feng Wu, F.; Li, L.; Chen, R. Structural and electrochemical characteristics of hierarchical Li4Ti5O12 as high-rate anode material for lithium-ion batteries. Electrochim. Acta 2021, 368, 137470. [Google Scholar] [CrossRef]
  159. Liu, J.; Song, K.; van Aken, P.A.; Maier, J.; Yu, Y. Self-supported Li4Ti5O12-C nanotube arrays as high-rate and long-life anode materials for flexible Li-ion batteries. Nano Lett. 2014, 14, 2597–2603. [Google Scholar] [CrossRef] [PubMed]
  160. Qin, W.; Chen, Y.; An, J.; Wen, X. Self-supported Li4Ti5O12 nanobelt array anode with long-life and improved low-temperature performance for flexible lithium-ion batteries. Ceram. Int. 2022, 48, 22194–22203. [Google Scholar] [CrossRef]
  161. Liu, J.; Wei, A.; Pan, G.; Shen, S.; Xiao, Z.; Zhao, Y.; Xia, X. Self-supported hierarchical porous Li4Ti5O12/carbon arrays for boosted lithium ion storage. J. Energy Chem. 2021, 54, 754–760. [Google Scholar] [CrossRef]
  162. Pawlitzek, F.; Althues, H.; Schumm, B.; Kaskel, S. Nanostructued networks for energy stotage: Vertically aligned carbon nanotubes (VACNT) as current collectors for high-power Li4Ti5O12(LTO)//LiMn2O4(LMO) lithium ion batteries. Batteries 2017, 3, 37. [Google Scholar] [CrossRef]
  163. Shen, L.; Uchaker, E.; Zhang, X.; Cao, G. Hydrogenated Li4Ti5O12 nanowire arrays for high rate lithium ion batteries. Adv. Mater. 2012, 24, 6502–6506. [Google Scholar] [CrossRef]
  164. Xia, Q.; Jabeen, N.; Savilov, S.V.; Aldoshind, S.M.; Xia, H. Black mesoporous Li4Ti5O12–δ nanowall arrays with improved rate performance as advanced 3D anodes for microbatteries. J. Mater. Chem. A 2016, 4, 17543–17551. [Google Scholar] [CrossRef]
  165. Jiang, C.; Ding, W.; Wu, H.; Yu, Z.; Ma, L.; Zou, Z. Hierarchical Li4Ti5O12 nanosheet arrays anchoring on carbon fiber cloth as ultra-stable free-standing anode of Li-ion battery. Ceram. Int. 2018, 44, 3040–3047. [Google Scholar] [CrossRef]
  166. Kim, S.D.; Rana, K.; Ahn, J.-H. Additive-free synthesis of Li4Ti5O12 nanowire arrays on freestanding ultrathin graphite as a hybrid anode for flexible lithium ion batteries. J. Mater. Chem. A 2016, 4, 19197–19206. [Google Scholar] [CrossRef]
  167. Sorensen, E.M.; Barry, S.J.; Jung, H.K.; Rondinelli, J.R.; Vaughey, J.T.; Poeppelmeier, K.R. Three-dimensionally ordered macroporous Li4Ti5O12: Effect of wall structure on electrochemical properties. Chem. Mater. 2006, 18, 482–489. [Google Scholar] [CrossRef]
  168. Li, N.; Zhou, G.; Li, F.; Wen, L.; Cheng, H.-M. A self-standing and flexible electrode of Li4Ti5O12 nanosheets with a N-doped carbon coating for high rate lithium ion batteries. Adv. Funct. Mater. 2013, 23, 5429–5435. [Google Scholar] [CrossRef]
  169. Chen, S.; Xin, Y.; Zhou, Y.; Ma, Y.; Zhou, H.; Qi, L. Self-supported Li4Ti5O12 nanosheet arrays for lithium ion batteries with excellent rate capability and ultralong cycle life. Energy Environ. Sci. 2014, 7, 1924–1930. [Google Scholar] [CrossRef]
  170. Julien, C.M.; Massot, M.; Zaghib, K. Structural studies of Li4/3Me5/3O4 (Me = Ti, Mn) electrode materials: Local structure and electrochemical aspects. J. Power Sources 2004, 136, 72–79. [Google Scholar] [CrossRef]
  171. Leonidov, I.A.; Leonidova, O.N.; Perelyaeva, A.L.; Samigullina, R.F.; Kovyazina, S.A.; Patrakeev, M.V. Structure, ionic conduction, and phase transformations in lithium titanate Li4Ti5O12. Phys. Solid State 2003, 45, 2183–2188. [Google Scholar] [CrossRef]
  172. Julien, C.M.; Massot, M. Lattice vibrations of materials for lithium rechargeable batteries. I. Lithium manganese oxide spinel. Mater. Sci. Eng. B 2003, 97, 217–230. [Google Scholar] [CrossRef]
  173. Pelegov, D.V.; Slautin, B.N.; Gorshkov, V.S.; Zelenovskiy, P.S.; Kiselev, E.A.; Kholkin, A.L.; Shur, V.Y. Raman spectroscopy “big data” and local heterogeneity of solid state synthesized lithium titanate. J. Power Sources 2017, 346, 143–150. [Google Scholar] [CrossRef]
  174. Mosa, J.; Aparicio, M.; Tadanaga, K.; Hayashi, A.; Tatsumisago, M. Li4Ti5O12 thin-film electrodes by in-situ synthesis of lithium alkoxide for Li-ion microbatteries. Electrochim. Acta 2014, 149, 293–299. [Google Scholar] [CrossRef]
  175. Tran, M.V.; Huynh, N.L.T.; Nguyen, T.T.; Ha, D.T.C.; Le, P.M.L. Facile solution route to synthesize nanostructure Li4Ti5O12 for high rate Li-ion battery. J. Nanomater. 2016, 2016, 4261069. [Google Scholar] [CrossRef]
  176. Zhang, Q.; Lu, H.; Zhong, H.; Yan, X.; Ouyang, C.; Zhang, L. W6+ and Br codoped Li4Ti5O12 anode with super rate performance for Li-ion batteries. J. Mater. Chem. A 2015, 3, 13706–13716. [Google Scholar] [CrossRef]
  177. Pelegov, D.V.; Nasara, R.N.; Tu, C.; Li, S. Defects in Li4Ti5O12 induced by carbon deposition: An analysis of unidentified bands in Raman spectra. Phys. Chem. Chem. Phys. 2019, 21, 20757–20763. [Google Scholar] [CrossRef]
  178. Kang, C.-Y.; Krajewski, M.; Lin, J.-Y. Impact of titanium precursors on formation and electrochemical properties of Li4Ti5O12 anode materials for lithium-ion batteries. J. Solid State Electrochem. 2021, 25, 575–582. [Google Scholar] [CrossRef]
  179. Li, X.; Hu, H.; Huang, S.; Yu, G.; Gao, L.; Liu, H.; Yu, Y. Nano-sized Li4Ti5O12 anode material with excellent performance prepared by solid state reaction: The effect of precursor size and morphology. Electrochim. Acta 2013, 112, 356–363. [Google Scholar] [CrossRef]
  180. Lin, C.; Fan, X.; Xin, Y.; Cheng, F.; Lai, M.O.; Zhou, H.; Lu, L. Monodispersed mesoporous Li4Ti5O12 submicrospheres as anode materials for lithium-ion batteries: Morphology and electrochemical performances. Nanoscale 2014, 6, 6651–6660. [Google Scholar] [CrossRef]
  181. Lin, Y.-S.; Duh, J.-G.; Tsai, M.-C.; Lee, C.-Y. Self-assembled synthesis of monodispersed mesoporous Li4Ti5O12 beads and their applications in secondary lithium-ion batteries. Electrochim. Acta 2012, 83, 47–52. [Google Scholar] [CrossRef]
  182. Pawlitzek, F.; Pampel, J.; Schmuck, M.; Althues, H.; Schumm, B.; Kaskel, S. High-power lithium ion batteries based on preorganized necklace type Li4Ti5O12/VACNT nano-composites. J. Power Sources 2016, 325, 1–6. [Google Scholar] [CrossRef]
  183. Tang, Y.; Tan, X.; Hou, G.; Zheng, G. Nanocrystalline Li4Ti5O12-coated TiO2 nanotube arrays as three-dimensional anode for lithium-ion batteries. Electrochim. Acta 2014, 117, 172–178. [Google Scholar]
  184. Wang, G.; Wang, H.; Ma, G.; Du, X.; Du, L.; Jing, P.; Wang, Y.; Wu, K.; Wu, H.; Wang, Q.; et al. Investigation on process mechanism of a novel energy-saving synthesis for Li4Ti5O12 anode material. J. Energy Chem. 2022, 70, 266–275. [Google Scholar] [CrossRef]
  185. Zhang, B.; Yu, Y.; Liu, Y.; Huang, Z.-D.; He, Y.; Kim, J.-K. Percolation threshold of graphene nanosheets as conductive additives in Li4Ti5O12 anodes of Li-ion batteries. Nanoscale 2013, 5, 2100–2106. [Google Scholar] [CrossRef]
  186. Li, Z.; Ding, F.; Zhao, Y.; Wang, Y.; Li, J.; Yang, K. synthesis and electrochemical performance of Li4Ti5O12 submirospheres coated with TiN as anode materials for lithium-ion batteries. Ceram. Int. 2016, 42, 15464–15470. [Google Scholar] [CrossRef]
  187. Lin, C.-Y.; Duh, J.-G. Porous Li4Ti5O12 anode material synthesized by one-step solid-state method for electrochemical properties enhancement. J. Alloys Compd. 2011, 509, 3682–3685. [Google Scholar] [CrossRef]
  188. Haetge, J.; Hartman, P.; Brezenski, K.; Janek, J.; Brezenski, T. Ordered large-pore mesoporous Li4Ti5O12 spinel thin film electrodes with nanocrystalline framework for high rate rechargeable lithium batteries: Relationships among charge storage, electrical conductivity, and nanoscale structure. Chem. Mater. 2011, 23, 4384–4393. [Google Scholar] [CrossRef]
  189. Nowack, L.V.; Waser, O.; Yarema, O.; Wood, V. Rapid, microwave-assisted synthesis of battery-grade lithium titanate (LTO). RSC Adv. 2013, 3, 15618–15621. [Google Scholar] [CrossRef]
  190. Nguyen Huynh, L.T.; Duy Ha, C.T.; Nguyen, V.D.; Nguyen, D.Q.; Phung Le, M.L.; Tran, V.M. Structure and electrochemical properties of Li4Ti5O12 prepared via low-temperature precipitation. J. Chem. 2019, 2019, 1727859. [Google Scholar]
  191. Mahmoud, A.; Saadoune, I.; Lippens, P.; Chamas, M. The design and study of new Li-ion full cells of LiCo2/3Ni1/6Mn1/6O2 positive electrode paired with MnSn2 and Li4Ti5O12 negative electrodes. Solid State Ion. 2017, 300, 175–181. [Google Scholar] [CrossRef]
  192. Wang, L.; Zhang, Y.M.; Scofield, M.E.; Yue, S.Y.; McBean, C.; Marschilok, A.C.; Takeuchi, K.J.; Takeuchi, E.S.; Wong, S.S. Enhanced performance of “flower-like” Li4Ti5O12 motifs as anode materials for high-rate lithium-ion batteries. ChemSusChem 2015, 8, 3304–3313. [Google Scholar] [CrossRef]
  193. Wang, J.; Yang, Z.; Li, W.; Zhong, X.; Gu, L.; Yu, Y. Nitridation Br-doped Li4Ti5O12 anode for high rate lithium ion batteries. J. Power Sources 2014, 266, 323–331. [Google Scholar] [CrossRef]
  194. Hong, H.-J.; Ban, G.; Lee, S.-M.; Park, I.-S.; Lee, Y.-J. Synthesis of 3D-structured Li4Ti5O12 from titanium(IV) oxysulfate (TiOSO4) solution as a highly sustainable anode material for lithium-ion batteries. J. Alloys Compd. 2020, 844, 156203. [Google Scholar] [CrossRef]
  195. Spitler, T.; Prochazka, J.; Kavan, L.; Graetzel, M.; Sugnaux, F. High Performance Lithium Titanium Spinel Li4Ti5O12 for Electrode Material. U.S. Patent 7,547,490 B2, 16 June 2009. [Google Scholar]
  196. Han, S.Y.; Kim, I.Y.; Jo, K.Y.; Hwang, S.J. Solvothermal-assisted hybridization between reduced graphene oxide and lithium metal oxides: A facile route to graphene-based composite materials. J. Phys. Chem. C 2012, 116, 7269–7279. [Google Scholar] [CrossRef]
  197. Jung, H.G.; Myung, S.T.; Yoon, C.S.; Son, S.B.; Oh, K.H.; Amine, K.; Scrosati, B.; Sun, Y.K. Microscale spherical carbon-coated Li4Ti5O12 as ultra high power anode material for lithium batteries. Energy Environ. Sci. 2011, 4, 1345–1351. [Google Scholar] [CrossRef]
  198. Wu, F.; Wang, Z.; Li, X.; Wu, L.; Wang, X.; Zhang, X.; Wang, Z.; Xiong, X.; Guo, H. Preparation and characterization of spinel Li4Ti5O12 anode material from industrial titanyl sulfate solution. J. Alloys Compd. 2011, 509, 596–601. [Google Scholar] [CrossRef]
  199. Tsai, M.C.; Tsai, T.L.; Lin, C.T.; Chung, R.J.; Sheu, H.S.; Chiu, H.T.; Lee, C.Y. Tailor made Mie scattering color filters made by size-tunable titanium dioxide particles. J. Phys. Chem. C 2008, 112, 2697–2702. [Google Scholar] [CrossRef]
  200. Zhang, Q.; Zhang, S.; Ning, F.; Lu, X.; Liu, Y.; Nie, L.; Ouyang, C.; Zhang, L. Calcium doping of lithium titanium oxide nanospheres: A combined first-principles and experimental study. Energy Technol. 2017, 5, 539–543. [Google Scholar] [CrossRef]
  201. Zhang, F.; Yi, F.; Gao, A.; Shu, D.; Sun, Z.; Mao, J.; Zhou, X.; Zhu, Z.; Sun, Y. Interfacial electrostatic self-assembly in water-in-oil microemulsion assisted synthesis of Li4Ti5O12/graphene for lithium-ion-batteries. J. Alloys Compd. 2020, 819, 153018. [Google Scholar] [CrossRef]
  202. Liao, J.-Y.; Xiao, X.; Higgins, D.; Lee, D.; Hassan, F.; Chen, Z. Hierarchical Li4Ti5O12-TiO2 composite microsphere consisting of nanocrystals for high power Li-ion batteries. Electrochim. Acta 2013, 108, 104–111. [Google Scholar] [CrossRef]
  203. Yu, S.-H.; Pucci, A.; Herntrich, T.; Willinger, M.-G.; Baek, S.-H.; Sungac, Y.-E.; Pinna, N. Surfactant-free nonaqueous synthesis of lithium titanium oxide (LTO) nanostructures for lithium ion battery applications. J. Mater. Chem. 2011, 21, 806–810. [Google Scholar] [CrossRef]
  204. Wu, L.; Kan, S.R.; Lu, S.G.; Zhang, X.J.; Jin, W.H. Effect of particle size and agglomeration of TiO2 on synthesis and electrochemical properties of Li4Ti5O12. Trans. Nonferrous Met. Soc. China 2007, 17, s117–s212. [Google Scholar]
  205. Yuan, T.; Cai, R.; Shao, Z. Different effect of the atmospheres on the phase formation and performance of Li4Ti5O12 prepared from ball-milling-assisted solid-phase reaction with pristine and carbon precoated TiO2 as starting materials. J. Phys. Chem. C 2011, 115, 4943–4952. [Google Scholar] [CrossRef]
  206. Xu, R.; Li, J.; Tang, Z.; Zhang, Z. Li4Ti5O12 heat treated under nitrogen ambient with outstanding rate capabilities. J. Nanomater. 2011, 2011, 635416. [Google Scholar] [CrossRef]
  207. Mahmoud, A.; Amarilla, J.M.; Lasri, K.; Saadoune, I. Influence of the synthesis method on the electrochemical properties of the Li4Ti5O12 spinel in Li-half and Li-ion full-cells. A systematic comparison. Electrochim. Acta 2013, 93, 163–172. [Google Scholar] [CrossRef]
  208. Chen, X.; Guan, X.; Li, L.; Li, G. Defective mesoporous Li4Ti5O12−y: An advanced anode with anomalous capacity and cycling stability at a high rate of 20C. J. Power Sources 2012, 210, 297–302. [Google Scholar] [CrossRef]
  209. Wang, Y.Q.; Gu, L.; Guo, Y.G.; Li, H.; He, X.Q.; Tsukimoto, S.; Ikuhara, Y.; Wan, L.J. Rutile-TiO2 nanocoating for a high-rate Li4Ti5O12 anode of a lithium-ion battery. J. Am. Chem. Soc. 2012, 134, 7874–7879. [Google Scholar] [CrossRef] [PubMed]
  210. Chiu, H.; Demopoulos, G.P. A novel green approach to synthesis of nanostructured Li4Ti5O12 anode material. ECS Trans. 2013, 50, 119–128. [Google Scholar] [CrossRef]
  211. Demopoulos, G.; Chiu, H.; Zaghib, K.; Guerfi, A. Layered and Spinel Lithium Titanates and Processes for Preparing the Same. World Patent WO 2014/056111 A1, 17 April 2014. [Google Scholar]
  212. Feng, X.; Zou, H.; Xiang, H.; Guo, X.; Zhou, T.; Wu, Y.; Xu, W.; Yan, P.; Wang, C.; Zhang, J.-G.; et al. Ultrathin Li4Ti5O12 nanosheets as anode materials for lithium and sodium storage. ACS Appl. Mater. Interfaces 2016, 8, 16718–16726. [Google Scholar] [CrossRef]
  213. Suzuki, S.; Kozawa, T.; Murakami, T.; Naito, M. Mechanochemical-hydrothermal synthesis of layered lithium titanate hydrate nanotubes at room temperature and their conversion to Li4Ti5O12. Mater. Res. Bull. 2017, 90, 218–223. [Google Scholar] [CrossRef]
  214. Liu, J.; Li, X.; Geng, D.; Li, Y.; Wang, D.; Li, R.; Sun, X.; Cai, M.; Verbrugge, M.W. Microwave-assisted hydrothermal synthesis of nanostructured spinel Li4Ti5O12 as anode materials for lithium ion batteries. Electrochim. Acta 2012, 63, 100–104. [Google Scholar] [CrossRef]
  215. Xu, G.B.; Li, W.; Yang, L.W.; Wei, X.L.; Ding, J.W.; Zhong, J.X.; Chu, P.K. Highly-crystalline ultrathin Li4Ti5O12 nanosheets decorated with silver nanocrystals as a high-performance anode material for lithium ion batteries. J. Power Sources 2015, 276, 247–254. [Google Scholar] [CrossRef]
  216. Feng, X.-Y.; Li, X.; Tang, M.; Gan, A.; Hu, Y.-Y. Enhanced rate performance of Li4Ti5O12 anodes with bridged grain boundaries. J. Power Sources 2017, 354, 172–178. [Google Scholar] [CrossRef]
  217. Yuan, T.; Cai, R.; Ran, R.; Zhou, Y.; Shao, Z. A mechanism study of synthesis of Li4Ti5O12 from TiO2 anatase. J. Alloys Compd. 2010, 505, 367–373. [Google Scholar] [CrossRef]
  218. Lin, C.; Lai, M.O.; Li, L.; Zhou, H.; Xin, Y. Structure and high performance of Ni2+ doped Li4Ti5O12 for lithium ion battery. J. Power Sources 2013, 244, 272–279. [Google Scholar] [CrossRef]
  219. Hu, X.; Li, Z.; Yang, K.; Huai, Y.; Deng, Z. Effects of carbon source and carbon content on electrochemical performances of Li4Ti5O12/C prepared by one-step solid-state reaction. Electrochim. Acta 2011, 56, 5046–5053. [Google Scholar] [CrossRef]
  220. Shen, Y.; Søndergaard, M.; Christensen, M.; Birgisson, S.; Iversen, B.B. Solid state formation mechanism of Li4Ti5O12 from an anatase TiO2 source. Chem. Mater. 2014, 26, 3679–3686. [Google Scholar] [CrossRef]
  221. Harrison, M.R.; Edwards, P.P.; Goodenough, J.B. The superconductor-semiconductor transition in the Li1+xTi2–xO4 spinel system. Philos. Mag. B 1985, 52, 679–699. [Google Scholar] [CrossRef]
  222. Jhan, Y.R.; Duh, J.G. Synthesis of entanglement structure in nanosized Li4Ti5O12/multi-walled carbon nanotubes composite anode material for Li-ion batteries by ball-milling-assisted solid-state reaction. J. Power Sources 2012, 198, 294–297. [Google Scholar] [CrossRef]
  223. Wang, Y.; Zou, W.; Dai, X.; Feng, L.; Zhang, H.; Zhou, A.; Li, J. Solid-state synthesis of graphite carbon-coated Li4Ti5O12 anode for lithium ion batteries. Ionics 2014, 20, 1377–1383. [Google Scholar] [CrossRef]
  224. Yi, T.F.; Yang, S.Y.; Li, X.Y.; Yao, J.H.; Zhu, Y.R.; Zhu, R.S. Sub-micrometric Li4−xNaxTi5O12 (0 ≤ x ≤ 0.2) spinel as anode material exhibiting high-rate capability. J. Power Sources 2014, 246, 505–511. [Google Scholar] [CrossRef]
  225. Yang, L.X.; Gao, L.J. Li4Ti5O12/C composite electrode material synthesized involving conductive carbon precursor for Li-ion battery. J. Alloys Compd. 2009, 485, 93–97. [Google Scholar] [CrossRef]
  226. Kim, J.; Cho, J. Spinel Li4Ti5O12 nanowires for high-rate Li-ion intercalation electrode. Electrochem. Solid-State Lett. 2007, 10, A81–A83. [Google Scholar] [CrossRef]
  227. Han, S.-W.; Ryu, J.H.; Jeong, J.; Yoon, D.-H. Solid-state synthesis of Li4Ti5O12 for high power lithium-ion battery applications. J. Alloys Compd. 2013, 570, 144–149. [Google Scholar] [CrossRef]
  228. Yao, W.; Zhuang, W.; Ji, X.; Chen, J.; Lu, X.; Wang, C. Solid-state synthesis of Li4Ti5O12 whiskers from TiO2-B. Mater. Res. Bull. 2016, 75, 204–210. [Google Scholar] [CrossRef]
  229. Cheng, L.; Yan, J.; Zhu, G.-N.; Luo, J.-Y.; Wang, C.-X.; Xia, Y.-Y. General synthesis of carbon-coated nanostructure Li4Ti5O12 as a high rate electrode material for Li-ion intercalation. J. Mater. Chem. 2010, 20, 595–602. [Google Scholar] [CrossRef]
  230. Wang, D.; Wu, X.; Zhang, Y.; Wang, J.; Yan, P.; Zhang, C.; He, D. The influence of the TiO2 particle size on the properties of Li4Ti5O12 anode material for lithium-ion battery. Ceram. Int. 2014, 40, 3799–3804. [Google Scholar] [CrossRef]
  231. Ohtake, T.; Iijima, K. Li4Ti5O12 synthesis with high specific surface area and single phase. J. Mater. Sci. Chem. Eng. 2015, 3, 68–73. [Google Scholar]
  232. Ma, G.; Cheng, M. Preparation and properties of Li4Ti5O12/C composites. Integr. Ferroelectr. 2019, 548, 42–49. [Google Scholar] [CrossRef]
  233. Purwanto, A.; Muzayanha, S.U.; Yudha, C.S.; Widiyandari, H.; Jumari, A.; Dyartanti, E.R.; Nizam, M.; Putra, M.I. High performance of salt-modified–LTO anode in LiFePO4 battery. Appl. Sci. 2020, 10, 7135. [Google Scholar] [CrossRef]
  234. Bai, X.; Li, T.; Bai, Y.-J. Capacity degradation of Li4Ti5O12 during long-term cycling in terms of composition and structure. Dalton Trans. 2020, 49, 10003–10010. [Google Scholar] [CrossRef]
  235. Allen, J.L.; Jow, T.R.; Wolfenstine, J. Low temperature performance of nanophase Li4Ti5O12. J. Power Sources 2006, 159, 1340–1345. [Google Scholar] [CrossRef]
  236. Yao, X.L.; Xie, S.; Nian, H.Q.; Chen, C.H. Spinel Li4Ti5O12 as a reversible anode material down to 0 V. J. Alloys Compd. 2008, 465, 375–379. [Google Scholar] [CrossRef]
  237. Wang, Z.; Wang, Z.; Peng, W.; Guo, H.; Li, X. An improved solid-state reaction to synthesize Zr-doped Li4Ti5O12 anode material and its application in LiMn2O4/Li4Ti5O12 full-cell. Ceram. Int. 2014, 40, 10053–10059. [Google Scholar] [CrossRef]
  238. Zaghib, K.; Simoneau, M.; Armand, M.; Gauthier, M. Electrochemical study of Li4Ti5O12 as negative electrode for Li-ion polymer rechargeable batteries. J. Power Sources 1999, 81–82, 300–305. [Google Scholar] [CrossRef]
  239. Berbenni, V.; Milanese, C.; Bruni, G.; Marini, A. Mechano-thermally activated solid-state synthesis of Li4Ti5O12 spinel from Li2CO3-TiO2 mixtures. Z. Naturforsch. B Chem. Sci. 2010, 65, 23–26. [Google Scholar] [CrossRef]
  240. Liu, W.; Zhang, J.; Wang, Q.; Xie, X.; Lou, Y.; Xia, B. The effects of Li2CO3 particle size on the properties of lithium titanate as anode material for lithium-ion batteries. Ionics 2014, 20, 1553–1560. [Google Scholar] [CrossRef]
  241. Wang, Y.; Ren, Y.; Dai, X.; Yan, X.; Huang, B.; Li, J. Electrochemical performance of ZnO-coated Li4Ti5O12 composite electrodes for lithium-ion batteries with the voltage ranging from 3 to 0.01 V. R. Soc. Open Sci. 2018, 5, 180762. [Google Scholar] [CrossRef] [PubMed]
  242. Huang, S.; Wen, Z.Y.; Zhu, X.J.; Gu, Z.H. Preparation and electrochemical performance of Ag doped Li4Ti5O12. Electrochem. Commun. 2004, 6, 1093–1097. [Google Scholar] [CrossRef]
  243. Zhou, T.P.; Feng, X.Y.; Guo, X.; Wu, W.W.; Cheng, S.; Xiang, H.F. Solid-state synthesis and electrochemical performance of Ce-doped Li4Ti5O12 anode materials for lithium-ion batteries. Electrochim. Acta 2015, 174, 369–375. [Google Scholar] [CrossRef]
  244. Becker, D.; Haberkorn, R.; Kickelbick, G. Mechanochemical induced structure transformations in lithium titanates: A detailed PXRD and 6Li MAS NMR study. Inorganics 2018, 6, 117. [Google Scholar] [CrossRef]
  245. Huang, Z.; Wang, D.; Lin, Y.; Wu, X.; Yan, P.; Zhang, C.; He, D. Enhancing the high-rate performance of Li4Ti5O12 anode material for lithium-ion battery by a wet ball milling assisted solid-state reaction and ultra-high speed nano-pulverization. J. Power Sources 2014, 266, 60–65. [Google Scholar] [CrossRef]
  246. Natalia, V.; Gustami, A.P.; Rahmawati, F.; Lestari, W.W.; Purwanto, A. Lithium titanate (LTO) synthesis through solid state reaction and its performance for LiFePO4/LTO battery. J. Math. Fund. Sci. 2018, 50, 290–302. [Google Scholar] [CrossRef]
  247. Xiao, H.; Huang, X.; Ren, Y.; Ding, X.; Zhou, S. Fabrication of Li4Ti5O12@CN composite with enhanced rate properties. Front. Chem. 2019, 7, 432. [Google Scholar] [CrossRef]
  248. Krajewski, M.; Michalska, M.; Hamankiewicz, B.; Ziolkowska, D.; Korona, K.P.; Jasinski, J.B.; Maria Kaminska, M.; Lipinska, L.; Czerwinski, A. Li4Ti5O12 modified with Ag nanoparticles as an advanced anode material in lithium-ion batteries. J. Power Sources 2014, 245, 764–771. [Google Scholar] [CrossRef]
  249. Kim, S.; Alauzun, J.G.; Louvain, N.; Brun, N.; Stievano, L.; Boury, B.; Monconduit, L.; Mutin, P.H. Alginic acid aquagel as a template and carbon source in the synthesis of Li4Ti5O12/C nanocomposites for application as anodes in Li-ion batteries. RSC Adv. 2018, 8, 32558–32564. [Google Scholar] [CrossRef]
  250. Yi, T.-F.; Xie, Y.; Zhu, Y.-R.; Zhu, R.-S.; Shen, H. Structural and thermodynamical stability of Li4Ti5O12 anode material for lithium-ion battery. J. Power Sources 2013, 222, 448–454. [Google Scholar] [CrossRef]
  251. Qiao, Y.; Hu, X.; Liu, Y.; Huang, Y. Li4Ti5O12 nanocrystallites for high-rate lithium-ion batteries synthesized by a rapid microwave-assisted solid-state process. Electrochim. Acta 2012, 63, 118–123. [Google Scholar] [CrossRef]
  252. Shi, L.; Hu, X.; Huang, Y. Fast microwave-assisted synthesis of Nb-doped Li4Ti5O12for high-rate lithium-ion batteries. J. Nanopart. Res. 2014, 16, 2332. [Google Scholar] [CrossRef]
  253. Ohtake, T. Single phase Li4Ti5O12 synthesis for nanoparticles by two steps sintering. J. Mater. Sci. Chem. Eng. 2015, 3, 5–10. [Google Scholar]
  254. Bach, S.; Pereira-Ramos, J.P.; Baffier, N. Electrochemical behaviour of a lithium titanium spinel compound synthesized via a sol-gel process. J. Mater. Chem. 1998, 8, 251–253. [Google Scholar] [CrossRef]
  255. Venkateswarlu, M.; Chen, C.H.; Do, J.S.; Lin, C.W.; Chou, T.C.; Hwang, B.J. Electrochemical properties of nano-sized Li4Ti5O12 powders synthesized by a sol–gel process and characterized by X-ray absorption spectroscopy. J. Power Sources 2005, 146, 204–208. [Google Scholar] [CrossRef]
  256. Rho, Y.H.; Kanamura, K. Li+ ion diffusion in Li4Ti5O12 thin film electrode prepared by PVP sol-gel method. J. Solis State Chem. 2004, 177, 2094–2100. [Google Scholar] [CrossRef]
  257. Rho, Y.H.; Kanamura, K. Preparation of Li4/3Ti5/3O4 thin film electrodes by a PVP sol-gel coating method and their electrochemical properties. J. Electrochem. Soc. 2004, 151, A106–A110. [Google Scholar] [CrossRef]
  258. Gao, J.; Jiang, C.; Wan, C. Synthesis and characterization of spherical La-doped nanocrystalline Li4/3Ti5/3O4/C compound for Lithium-ion batteries. J. Electrochem. Soc. 2010, 157, K39–K42. [Google Scholar] [CrossRef]
  259. Wang, G.J.; Gao, J.; Fu, L.J.; Zhao, N.H.; Wu, Y.P.; Takamura, T. Preparation and characteristic of carbon-coated Li4/3Ti5/3O4 anode material. J. Power Sources 2007, 174, 1109–1112. [Google Scholar] [CrossRef]
  260. Liu, D.Q.; Lai, Q.Y.; Hao, Y.J. Study on synthesis and mechanism of Li4/3Ti5/3O4 by sol-gel method. Chin. J. Inorg. Chem. 2004, 20, 829–832. [Google Scholar]
  261. Xiang, H.F.; Tian, B.B.; Lian, P.C.; Li, Z.; Wang, H. Sol-gel synthesis and electrochemical performance of Li4Ti5O12/graphene composite anode for lithium-ion batteries. J. Alloys Compd. 2011, 509, 7205–7209. [Google Scholar] [CrossRef]
  262. Fawwaz, T.A.; Retna, D.P.; Riesma, T.; Ade, U.H.; Sri, R.; Damish; Hanif, Y.; Oka, P.A.; Nendar, H.; Yelvia, D.; et al. Synthesis of lithium titanium oxide (Li4Ti5O12) through sol-gel- method and the effect of graphene addition in lithium-ion battery anodes. Defect Diff. Forum 2022, 417, 227–240. [Google Scholar]
  263. Zhang, C.; Zhang, Y.; Wang, J.; Wang, D.; He, D.; Xia, Y. Li4Ti5O12 prepared by a modified citric acid sol-gel method for lithium-ion battery. J. Power Sources 2013, 236, 118–125. [Google Scholar] [CrossRef]
  264. Supriadi, C.P.; Syahrial, A.Z.; Subhan, A. The effect of sol-gel derived TiO2 crystallite size to Li4Ti5O12 anode performance in lithium-ion battery. Ionics 2020, 26, 5907–5914. [Google Scholar] [CrossRef]
  265. Livage, J. Vanadium pentoxide gels. Chem. Mater. 1991, 3, 578–593. [Google Scholar] [CrossRef]
  266. Julien, C.; El-Farh, L.; Rangan, S.; Massot, M. Studies of LiNi1–yCoyO2 cathode materials prepared by a citric acid-assisted sol-gel method for lithium batteries. J. Sol-Gel Sci. Technol. 1999, 15, 63–72. [Google Scholar] [CrossRef]
  267. Luo, G.; He, J.; Song, X.; Huang, X.; Yu, X.; Fang, Y.; Chen, D. Bamboo carbon assisted sol-gel synthesis of Li4Ti5O12 anode material with enhanced electrochemical activity for lithium ion battery. J. Alloys Compd. 2015, 621, 268–768. [Google Scholar] [CrossRef]
  268. Hao, Y.-J.; Lai, Q.-Y.; Liu, D.-Q.; Xu, Z.-U.; Ji, X.-Y. Synthesis by citric acid sol-gel method and electrochemical properties of Li4Ti5O12 anode material for lithium-ion battery. Mater. Chem. Phys. 2005, 94, 382–387. [Google Scholar] [CrossRef]
  269. Yang, G.; Su, Z.; Fang, H.; Yao, Y.; Li, Y.; Yang, B.; Ma, W. Synthesis and performance of Li4Ti5O12/C with little inert carbon. Electrochim. Acta 2013, 93, 158–162. [Google Scholar] [CrossRef]
  270. Kuo, Y.-C.; Lin, J.-Y. One-pot sol-gel synthesis of Li4Ti5O12/C anode materials for high-performance Li-ion batteries. Electrochim. Acta 2014, 142, 43–50. [Google Scholar] [CrossRef]
  271. Kurajica, S. A brief review on the use of chelation agents in sol-gel Synthesis with emphasis on β-diketones and β-ketoesters. Chem. Biochem. Eng. Q. 2019, 33, 295–301. [Google Scholar] [CrossRef]
  272. Kirillov, S.A.; Romanova, I.V.; Lisnycha, T.V.; Potapenko, A.V. High-rate electrochemical performance of Li4Ti5O12 obtained from TiCl4 by means of a citric acid aided route. Electrochim. Acta 2018, 286, 163–171. [Google Scholar] [CrossRef]
  273. Wang, J.; Liu, X.-M.; Yang, H.; Shen, X. Characterization and electrochemical properties of carbon-coated Li4Ti5O12 prepared by a citric acid sol–gel method. J. Alloys Compd. 2011, 509, 712–718. [Google Scholar] [CrossRef]
  274. Purwamargapratala, Y.; Sujatno, A.; Sabayu, Y.L.; Kartini, E. Synthesis of Li4Ti5O12 (LTO) by sol-gel method for lithium ion battery anode. IOP Conf. Ser. Mater. Sci. Eng. 2019, 553, 012062. [Google Scholar] [CrossRef]
  275. Stenina, I.A.; Il’in, A.B.; Yaroslavtsev, A.B. Synthesis and ionic conductivity of Li4Ti5O12. Inorg. Mater. 2015, 51, 62–67. [Google Scholar] [CrossRef]
  276. Mahmoud, A.; Amarilla, J.M.; Saadoune, I. Effect of thermal treatment used in the sol-gel synthesis method of Li4Ti5O12 spinel on its electrochemical properties as anode for lithium ion batteries. Electrochim. Acta 2015, 163, 213–222. [Google Scholar] [CrossRef]
  277. Zhou, W.; Shao, Z.; Jin, W. Synthesis of nanocrystalline conducting composite oxides based on a non-ion selective combined complexing process for functional applications. J. Alloys Compd. 2006, 426, 368–374. [Google Scholar] [CrossRef]
  278. Rho, Y.H.; Kanamura, K. Preparation of Li4Ti5O12 thin film electrode with PVP sol-gel for a rechargeable lithium microbattery. J. Surf. Sci. Soc. Jpn. 2003, 24, 423–428. [Google Scholar] [CrossRef]
  279. Fang, W.; Cheng, X.; Zuo, P.; Ma, Y.; Liao, L.; Yin, G. Hydrothermal-assisted sol-gel synthesis of Li4Ti5O12/C nano-composite for high-energy lithium-ion batteries. Solid State Ion. 2013, 244, 52–56. [Google Scholar] [CrossRef]
  280. Chang, L.-J.; Luo, D.-H.; Zhang, H.-L.; Qi, X.-W.; Wang, Z.-Y.; Liu, Y.-G.; Zhai, Y.-C. Synthesis and performance of Li4Ti5O12 anode materials using the PVP-assisted combustion method. Chin. Chem. Lett. 2014, 25, 1569–1572. [Google Scholar] [CrossRef]
  281. Lee, J.-M.; Jun, Y.-D.; Kim, D.-W.; Lee, Y.-H.; Oh, S.-G. Effects of PVP on the formation of silver–polystyrene heterogeneous nanocomposite particles in novel preparation route involving polyol process: Molecular weight and concentration of PVP. Mater. Chem. Phys. 2009, 114, 549–555. [Google Scholar] [CrossRef]
  282. Ma, G.; Cheng, M. Preparation and performance of Li4Ti5O12 by sol-gel method. Integr. Ferroelectr. 2020, 209, 119–124. [Google Scholar] [CrossRef]
  283. Xia, Y.; Sun, B.; Wei, Y.; Tao, B.; Zhao, Y. Simple sol-gel method synthesis of 3-dimension Li4Ti5O12-TiO2 nanostructures using butterfly wings as biotemplates for high-rate performance lithium-ion batteries. J. Alloys Compd. 2017, 705, 58–63. [Google Scholar] [CrossRef]
  284. Qiu, C.; Yuan, Z.; Liu, L.; Ye, N.; Liu, J. Sol-gel preparation and electrochemical properties of La-doped anode material for lithium-ion batteries. J. Solid State Electrochem. 2013, 17, 841–847. [Google Scholar] [CrossRef]
  285. Kavan, L.; Grätzel, M. Facile synthesis of nanocrystalline Li4Ti5O12 (spinel) exhibiting fast Li insertion. Electrochem. Solid State Lett. 2002, 5, A39–A42. [Google Scholar] [CrossRef]
  286. Pershina, S.V.; Antonov, B.D.; Farlenkov, A.S. Optimization of technology for synthesis of Li4Ti5O12 anode materials for lithium-ion batteries. Russ. J. Appl. Chem. 2021, 94, 30–37. [Google Scholar] [CrossRef]
  287. Wang, J.; Liu, X.; Yang, H. Synthesis and electrochemical properties of highly dispersed Li4Ti5O12 nanocrystalline for lithium secondary batteries. Trans. Nonferrous Met. Soc. China 2012, 22, 613–620. [Google Scholar] [CrossRef]
  288. Zhang, N.; Liu, Z.; Yang, T.; Liao, C.; Wang, Z.; Sun, K. Facile preparation of nanocrystalline Li4Ti5O12 and its high electrochemical performance as anode material for lithium-ion batteries. Electrochem. Commun. 2011, 13, 654–656. [Google Scholar] [CrossRef]
  289. Morsy, S.M.I. Role of surfactants in nanotechnology and their applications. Int. J. Curr. Microbiol. Appl. Sci. 2014, 3, 237–260. [Google Scholar]
  290. Dokan, F.K.; Sahan, H.; Ozdemir, B.; Ozdemir, N.; Patat, S. Synthesis and characterization of spinel Li4Ti5O12 anode material by CTAB assisted sol-gel method. Acta Phys. Pol. A 2014, 125, 648–649. [Google Scholar] [CrossRef]
  291. Khomane, R.B.; Prakash, A.S.; Ramesha, K.; Sathiya, M. CTAB-assisted sol-gel synthesis of Li4Ti5O12 and its performance as anode material for Li- ion batteries. Mater. Res. Bull. 2011, 46, 1139–1142. [Google Scholar] [CrossRef]
  292. Chen, J.; Yang, S.; Fang, S.; Hirani, S.; Tachibana, K. Synthesis of hierarchical mesoporous nest-like Li4Ti5O12 for high-rate lithium ion batteries. J. Power Sources 2012, 200, 59–66. [Google Scholar] [CrossRef]
  293. Li, D.; Zhao, W.; Cao, L.; Gao, Y.; Liu, Y.; Wang, W.; Qi, T. Mixed-surfactant-assisted synthesis of dual-phase Li4Ti5O12-TiO2 hierarchical microspheres as high-performance anode materials for Li-ion batteries. ChemSusChem 2019, 12, 4412–4420. [Google Scholar] [CrossRef] [PubMed]
  294. Akintola, T.; Shellikeri, A.; Akintola, T.; Zheng, J.P. The influence of Li4Ti5O12 preparation method on lithium-ion capacitor performance. Batteries 2021, 7, 33. [Google Scholar] [CrossRef]
  295. Meyer, F.; Hempelmann, R.; Mathur, S.; Veith, M. Microemulsion mediated sol-gel synthesis of nano-scaled MAl2O4 (M=Co, Ni, Cu) spinels from single-source heterobimetallic alkoxide precursors. J. Mater. Chem. 1999, 9, 1755–1763. [Google Scholar] [CrossRef]
  296. Hasegawa, G.; Kanamori, K.; Kiyomura, T.; Kurata, H.; Nakanishi, K.; Abe, T. Hierarchically porous Li4Ti5O12 anode materials for Li- and Na-ion batteries: Effects of nanoarchitectural design and temperature dependence of the rate capability. Adv. Energy Mater. 2015, 5, 1400730. [Google Scholar] [CrossRef]
  297. Wang, D.; Liu, H.; Li, M.; Wang, X.; Bai, S.; Shi, Y.; Tian, J.; Shan, Z.; Meng, M.; Liu, P.; et al. Nanosheet-assembled hierarchical Li4Ti5O12 microspheres for high-volumetric-density and high-rate Li-ion battery anode. Energy Storage Mater. 2019, 21, 361–371. [Google Scholar] [CrossRef]
  298. Erdas, A.; Ozcan, S.; Guler, M.O.; Akbulut, H. Sol-gel synthesis of nanocomposite Cu–Li4Ti5O12 structures for ultrahigh rate Li-ion batteries. Acta Phys. Pol. A 2015, 127, 1026–1028. [Google Scholar] [CrossRef]
  299. Liu, Z.; Huang, Y.; Wang, X.; Zhang, Y.; Ding, J.; Guo, Y. Synthesis of Li4Ti5O12/V2O5 nanocomposites for lithium-ion batteries by one-pot co-precipitation method. J. Mater. Sci. Mater. Electron. 2021, 32, 12134–12138. [Google Scholar] [CrossRef]
  300. Zhang, Y.; Zhang, C.; Lin, Y.; Xiong, D.-B.; Wang, D.; Wu, X.; He, D. Influence of Sc3+ doping in B-site on electrochemical performance of Li4Ti5O12 anode materials for lithium-ion battery. J. Power Sources 2014, 250, 50–57. [Google Scholar] [CrossRef]
  301. Wei, G.; Rambo, C.R.; Guo, Y.; Ning, Z.; Guo, S.; Zhao, M.; Huang, Z.; Zhang, C.; He, D. Graphene coated La3+/Sc3+ co-doped Li4Ti5O12 anodes for enhanced Li-ion battery performance. Mater. Lett. 2017, 193, 179–182. [Google Scholar] [CrossRef]
  302. Alias, N.; Kufian, M.; Teo, L.; Majid, S.; Arof, A. Synthesis and characterization of Li4Ti5O12. J. Alloys Compd. 2009, 486, 645–648. [Google Scholar] [CrossRef]
  303. Llaín-Jiménez, H.A.; Buchberger, D.A.; Winkowska-Struzik, M.; Ratyński, M.; Krajewski, M.; Boczar, M.; Hamankiewicz, B.; Czerwiński, A. Correlation between lithium titanium oxide powder morphology and high rate performance in lithium-ion batteries. Batteries 2022, 8, 168. [Google Scholar] [CrossRef]
  304. Rajoba, S.J.; Shirsat, A.N.; Tyagi, D.; Jadhav, L.D.; Kalubarme, R.S.; Wani, B.N.; Varma, S. Sol-gel assisted spinel Li4Ti5O12 and its performance and stability as anode for long life Li-ion battery. Mater. Lett. 2021, 285, 129134. [Google Scholar] [CrossRef]
  305. Najihah, A.I.; Priyono, S.; Imam Supardi, Z.A.; Subhan, A.; Prihandoko, B. Synthesized of Li4Ti5O12 materials via sol-gel method to lithium ion battery anodes. IOP Conf. Ser. J. Phys. 2020, 1491, 012038. [Google Scholar] [CrossRef]
  306. Priyono, S.; Nuroniah, I.; Subhan, A.; Sanjaya, E.; Prihandoko, B. Synthesis and characterization of Li4Ti5O12 with sol gel method as a lithium-ion battery anode material. J. Sains Mater. Indones. 2019, 20, 67–74. [Google Scholar] [CrossRef]
  307. Chang, C.-M.; Chen, Y.-C.; Ma, W.-L.; Wang, P.-H.; Lee, C.-F.; Chen, H.-S.; Chen-Yang, Y.W. Sol-gel synthesis of low carbon content and low surface area Li4Ti5O12/carbon black composite as high-rate anode materials for lithium ion battery. RSC Adv. 2015, 5, 74381–74390. [Google Scholar] [CrossRef]
  308. Shen, L.; Yuan, C.; Luo, H.; Zhang, X.; Xu, K.; Zhang, F. In situ growth of Li4Ti5O12 on multi-walled carbon nanotubes: Novel coaxial nanocables for high rate lithium ion batteries. J. Mater. Chem. 2011, 21, 761–767. [Google Scholar] [CrossRef]
  309. Zhong, Z. Synthesis of Mo4+ substituted spinel Li4Ti5–xMoxO12. Electrochem. Solid-State Lett. 2007, 10, A267–A269. [Google Scholar] [CrossRef]
  310. Zhang, A.; Zheng, Z.-M.; Cheng, F.-Y.; Liang, Z.; Chen, J. Preparation of Li4Ti5O12 submicrospheres and their application as anode materials of rechargeable lithium-ion batteries. Sci. China Chem. 2011, 54, 936–940. [Google Scholar] [CrossRef]
  311. Wang, X.-Y.; Li, Y.-J.; Xu, C.; Kong, L.; Li, L. Synthesis and characterization of Li4Ti5O12 via a hydrolysis process from TiCl4 aqueous solution. Rare Met. 2014, 33, 459–465. [Google Scholar] [CrossRef]
  312. Li, D.; Shen, G.; Zhao, W.; Gao, Y.; Hui, Z.; Liu, Y.; Yi, L.; Wang, W.; Cao, L.; Liu, Y.; et al. Synthesis of Li4Ti5O12 with theoretical capacity in Li2CO3-ammonia ball milling system. Mater. Res. Bull. 2019, 114, 177–183. [Google Scholar] [CrossRef]
  313. Wang, R.; Cao, X.; Zhao, D.; Zhu, L.; Xie, L.; Liu, J.; Liu, Y. Wet-chemistry synthesis of Li4Ti5O12 as anode materials rendering high-rate Li-ion storage. Int. J. Energy Res. 2020, 44, 4211–4223. [Google Scholar] [CrossRef]
  314. Xu, C.; Xue, L.; Zhang, W.; Fan, X.; Yan, Y.; Li, Q.; Huang, Y.; Zhang, W. Hydrothermal synthesis of Li4Ti5O12/TiO2 nanocomposite as high performance anode material for Li-ion batteries. Electrochim. Acta 2014, 147, 506–512. [Google Scholar] [CrossRef]
  315. Xie, L.-L.; Xu, Y.-D.; Zhang, J.J.; Cao, X.-Y.; Wang, B.; Yan, X.-Y.; Qu, L.-B. Facile Synthesis and characterization of Li4Ti5O12 as anode material for lithium ion batteries. Int. J. Electrochem. Sci. 2013, 8, 1701–1712. [Google Scholar] [CrossRef]
  316. Shin, J.W.; Hong, C.H.; Yoon, D.H. Effects of TiO2 starting materials on the solid-state formation of Li4Ti5O12. J. Am. Ceram. Soc. 2012, 95, 1894–1900. [Google Scholar] [CrossRef]
  317. Hong, C.H.; Noviyanto, A.; Ryu, J.H.; Kim, J.; Yoon, D.H. Effects of the starting materials and mechanochemical activation on the properties of solid-state reacted Li4Ti5O12 for lithium ion batteries. Ceram. Int. 2012, 38, 301–310. [Google Scholar] [CrossRef]
  318. Li, H.; Shen, L.; Zhang, X.; Wang, J.; Nie, P.; Che, Q.; Ding, B. Nitrogen-doped carbon coated Li4Ti5O12 nanocomposite: Superior anode materials for rechargeable lithium ion batteries. J. Power Sources 2013, 221, 122–127. [Google Scholar] [CrossRef]
  319. Veljković, I.; Poleti, D.; Karanović, L.; Zdujić, M.; Branković, G. Solid state synthesis of extra phase-pure Li4Ti5O12 spinel. Sci. Sinter. 2011, 4, 343–351. [Google Scholar] [CrossRef]
  320. Guerfi, A.; Yuichi, A.; Charest, P.; Mamoru, S.; Zaghib, K. Li4Ti5O12 material prepared by mechanochemical activation: Structure and electrochemical performance. ECS Meet. Abstr. 2006, MA2006-02, 206. [Google Scholar] [CrossRef]
  321. Priyono, B.; Winowatan, P.W.; Syahrial, A.Z.; Faizah; Subhan, A. Optimizing the performance of Li4Ti5O12/LTO by addition of silicon nanoparticles in half cell lithium_ion battery anode. IOP Conf. Ser. Earth Environ. Sci. 2018, 105, 012121. [Google Scholar] [CrossRef]
  322. Liu, W.; Zhang, J.; Wang, Q.; Xie, X.; Lou, Y.; Han, X.; Xia, B. Microsized TiO2 activated by high-energy ball milling as starting material for the preparation of Li4Ti5O12 anode material. Powder Technol. 2013, 247, 204–210. [Google Scholar] [CrossRef]
  323. Jia, P.; Shao, Z.; Liu, K. Synthesis and electrochemical performance of Li4Ti5O12 by high temperature ball milling method. Mater. Lett. 2014, 125, 218–220. [Google Scholar] [CrossRef]
  324. Jia, P.; Shao, Z.; Liu, K. Pretreatments-assisted high temperature ball milling route to Li4Ti5O12 and its electrochemical performance. Mater. Lett. 2014, 130, 71–74. [Google Scholar] [CrossRef]
  325. Michalska, M.; Krajewski, M.; Ziolkowska, D.; Hamankiewicz, B.; Andrzejczuk, M.; Lipinska, L.; Korona, K.P.; Czerwinski, A. Influence of milling time in solid-state synthesis on structure, morphology and electrochemical properties of Li4Ti5O12 of spinel structure. Powder Technol. 2014, 266, 372–377. [Google Scholar] [CrossRef]
  326. Yan, G.; Fang, H.; Zhao, H.; Li, G.; Yang, Y.; Li, L. Ball milling-assisted sol–gel route to Li4Ti5O12 and its electrochemical properties. J. Alloys Compd. 2009, 470, 544–547. [Google Scholar] [CrossRef]
  327. HKim, H.J.; Oh, M.H.; Son, W.K.; Kim, T.I.; Park, S.G. Novel synthesis method and electrochemical characteristics of lithium titanium oxide as anode material for high power device. In Proceedings of the 2006 IEEE 8th International Conference on Properties & Applications of Dielectric Materials, Bali, Indonesia, 26–30 June 2006; pp. 464–467. [Google Scholar]
  328. Wang, D.; Zhang, C.; Zhang, Y.; Wang, J.; He, D. Synthesis and electrochemical properties of La-doped Li4Ti5O12 as anode material for Li-ion battery. Ceram. Int. 2013, 39, 5145–5149. [Google Scholar] [CrossRef]
  329. Jang, I.-S.; Kang, S.-H.; Kang, Y.C.; Roh, K.C.; Chun, J. Facile synthesis of surface fluorinated-Li4Ti5O12/carbon nanotube nanocomposites for a high-rate capability anode of lithium-ion batteries. Appl. Surf. Sci. 2022, 605, 154710. [Google Scholar] [CrossRef]
  330. Ye, Z.; Zhong, F.; Chen, Y.; Zou, Z.; Jiang, C. Unique CNTs-chained Li4Ti5O12 nanoparticles as excellent high rate anode materials for Li-ion capacitors. Ceram. Int. 2022, 48, 20237–20244. [Google Scholar] [CrossRef]
  331. Chauque, S.; Oliva, F.Y.; Visintin, A.; Barraco, D.; Leiva, E.P.M.; Cámara, O.R. Lithium titanate as anode material for lithium ion batteries: Synthesis, post-treatment and its electrochemical response. J. Electroanal. Chem. 2017, 799, 142–155. [Google Scholar] [CrossRef]
  332. Zhang, C.; Shao, D.; Yu, J.; Zhang, L.; Huang, X.; Xu, D.; Yu, X. Synthesis and electrochemical performance of cubic Co-doped Li4Ti5O12 anode material for high-performance lithium-ion batteries. J. Electroanal. Chem. 2016, 776, 188–192. [Google Scholar] [CrossRef]
  333. Cheng, L.; Liu, H.J.; Zhang, J.J.; Xiong, H.M.; Xia, Y.Y. Nanosized Li4Ti5O12 prepared by molten salt method as an electrode material for hybrid electrochemical supercapacitors. J. Electrochem. Soc. 2006, 153, A1472–A1477. [Google Scholar] [CrossRef]
  334. Nithya, V.D.; Sharmila, S.; Vediappan, K.; Lee, C.W.; Vasylechko, L.; Kalai Selvan, R. Electrical and electrochemical properties of molten-saltsynthesized 0.05 mol Zr- and Si-doped Li4Ti5O12 microcrystals. J. Appl. Electrochem. 2014, 44, 647–654. [Google Scholar] [CrossRef]
  335. Nithya, V.D.; Kalai Selvan, R.; Kumaran, V.; Sharmila, S.; Lee, C.W. Molten salt synthesis and characterization of Li4Ti5−xMnxO12 (x = 0.0, 0.05 and 0.1) as anodes for Li-ion batteries. Appl. Surf. Sci. 2012, 261, 515–519. [Google Scholar] [CrossRef]
  336. Guo, Q.; Wang, Q.; Chen, G.; Shen, M.; Li, B. Molten salt synthesis of different ionic radii metallic compounds doped lithium titanate used in Li-ion battery anodes. Mater. Trans. 2017, 58, 383–389. [Google Scholar] [CrossRef]
  337. Rahman, M.; Wang, J.Z.; Hassan, M.F.; Wexler, D.; Liu, H.K. Amorphous carbon coated high grain boundary density dual phase Li4Ti5O12-TiO2: A nanocomposite anode material for Li-ion batteries. Adv. Energy. Mater. 2011, 1, 212–220. [Google Scholar] [CrossRef]
  338. Sharmila, S.; Senthilkumar, B.; Nithya, V.D.; Vediappan, K.; Lee, C.W.; Kalai Selvan, R. Electrical and electrochemical properties of molten salt-synthesized Li4Ti5–xO12SnxO12 (x = 0.0, 0.05 and 0.1) as anodes for Li-ion batteries. J. Phys. Chem. Solids 2013, 74, 1515–1521. [Google Scholar] [CrossRef]
  339. Rahman, M.M.; Wang, J.-Z.; Hassan, M.F.; Chou, S.; Wexler, D.; Liu, H.-K. Basic molten salt process—A new route for synthesis of nanocrystalline Li4Ti5O12–TiO2 anode material for Li-ion batteries using eutectic mixture of LiNO3–LiOH–Li2O2. J. Power Sources 2010, 195, 4297–4303. [Google Scholar] [CrossRef]
  340. Julien, C.; Camacho-Lopez, M.A.; Lemal, M.; Mohan, T.; Chitra, S.; Kalyani, P.; Gopukumar, S. Combustion synthesis and properties of substituted lithium cobalt oxides in lithium batteries. Solid State Ion. 2000, 135, 241–248. [Google Scholar] [CrossRef]
  341. Sha, Y.; Yuan, T.; Zhao, B.; Cai, R.; Wang, H.; Shao, Z. Solid lithium electrolyte-Li4Ti5O12 composites as anodes of lithium-ion batteries showing high-rate performance. J. Power Sources 2013, 231, 177–185. [Google Scholar] [CrossRef]
  342. Li, X.; Lin, H.; Cui, W.; Xiao, Q.; Zhao, J. Fast solution-combustion synthesis of nitrogen-modified Li4Ti5O12 nanomaterials with Improved electrochemical performance. ACS Appl. Mater. Interfaces 2014, 6, 7895–7901. [Google Scholar] [CrossRef]
  343. Shi, W.D.; Song, S.Y.; Zhang, H.J. Hydrothermal synthesis strategies of inorganic semiconducting nanostructures. Chem. Soc. Rev. 2013, 42, 5714–5743. [Google Scholar] [CrossRef]
  344. Li, B.; Han, C.; He, Y.-B.; Yang, C.; Du, H.; Yang, Q.-H.; Kang, F. Facile synthesis of Li4Ti5O12/C composite with super rate performance. Energy Environ. Sci. 2012, 5, 9595–9602. [Google Scholar] [CrossRef]
  345. Wu, H.-Y.; Hon, M.-H.; Kuan, C.-Y.; Leu, I.-C. Hydrothermal synthesis of nanosheets as anode materials for lithium ion batteries. RSC Adv. 2015, 5, 35224–35229. [Google Scholar] [CrossRef]
  346. Li, N.; Mei, T.; Zhu, Y.C.; Wang, L.; Liang, J.; Zhang, X.; Qian, Y.; Tang, K. Hydrothermal synthesis of layered Li1.81H0.19Ti2O5·xH2O nanosheets and their transformation to single-crystalline Li4Ti5O12 nanosheets as the anode materials for Li-ion batteries. CrystEngComm 2012, 14, 6435–6440. [Google Scholar] [CrossRef]
  347. Fattakhova, D.; Krtil, P. Electrochemical activity of hydrothermally synthesized Li-Ti-O cubic oxides toward Li insertion. J. Electrochem. Soc. 2002, 149, A1224–A1229. [Google Scholar] [CrossRef]
  348. Wu, S.C.; Guo, Y.X.; Zhou, J.H.; Zhao, J.; Ding, X.; Hu, J. Effect of heat-treatment temperature on the structure and properties of Li4Ti5O12 nanorods prepared by the hydrothermal ion exchange method. J. Inorg. Mater. 2011, 26, 123–128. [Google Scholar] [CrossRef]
  349. Zhang, Z.; Cao, L.; Huang, J.; Zhou, S.; Huang, Y.; Cai, Y. Hydrothermal synthesis of Zn-doped Li4Ti5O12 with improved high rate properties for lithium ion batteries. Ceram. Int. 2013, 39, 6139–6143. [Google Scholar] [CrossRef]
  350. Kim, K.-T.; Kim, S.-J.; Myung, S.-T. Fast Li+ transfer achieved by carbon coating on Li4Ti5O12 nanobelts. ECS Meet. Abstr. 2013, MA2013-02, 393. [Google Scholar] [CrossRef]
  351. Shi, Y.; Gao, J.; Abruna, H.D.; Liu, H.-K.; Li, H. Rapid synthesis of Li4Ti5O12/graphene composite with superior rate capability by a microwave-assisted hydrothermal method. Nano Energy 2014, 8, 297–304. [Google Scholar] [CrossRef]
  352. Ding, M.; Liu, H.; Zhu, J.; Zhao, X.; Pang, L.; Qin, Y.; Deng, L. Constructing of hierarchical yolk-shell structure Li4Ti5O12-SnO2 composites for high rate lithium ion batteries. Appl. Surf. Sci. 2018, 448, 389–399. [Google Scholar] [CrossRef]
  353. He, Y.; Muhetaer, A.; Li, J.; Wang, F.; Liu, C.; Li, Q.; Xu, D. Ultrathin Li4Ti5O12 nanosheet based hierarchical microspheres for high-rate and long-life Li-ion batteries. Adv. Energy Mater. 2017, 7, 1700950. [Google Scholar] [CrossRef]
  354. Zhang, Z.; Cao, L.; Huang, J.; Wang, D.; Wu, J.; Cai, Y. Hydrothermal synthesis of Li4Ti5O12 microsphere with high capacity as anode material for lithium ion batteries. Ceram. Int. 2013, 39, 2695–2698. [Google Scholar] [CrossRef]
  355. Zhang, Y.; Zhang, F.; Li, J.; Yang, Q. Hydrothermal synthesis and electrochemical performances of Ag-coated nanoflake-like Li4Ti5O12 as an anode material for lithium-ion batteries. Int. J. Electrochem. Sci. 2021, 16, 21035. [Google Scholar] [CrossRef]
  356. Prasad, G.V.; Reddy, T.M.; Narayana, A.L.; Hussain, O.M.; Gopal, T.V.; Shaikshavali, P. Construction of the embedded Li4Ti5O12-MWCNTs nanocomposite electrode for diverse applications in electrochemical sensing and rechargeable battery. J. Inorg. Organomet. Polymers Mater. 2023, 33, 1261–1279. [Google Scholar] [CrossRef]
  357. Wang, M.; He, Y.; Hong, W.; Zhang, S.Y.; Yang, C.X.; Yan, C. Three-dimensional network microstructure design of the Li4Ti5O12/rGO nanocomposite as an anode material for high-performance lithium-ion batteries. J. Phys. Chem. C 2023, 127, 10025–10037. [Google Scholar] [CrossRef]
  358. Fang, W.; Dong, E.; Zhang, Y.; Yang, L.; Zhang, L.; Zhang, H.; Wang, Y.; Che, G.; Yin, G. Self-assembled Li4Ti5O12/rGO nanocomposite anode for high power lithium-ion batteries. Inorg. Chem. Commun. 2022, 144, 109753. [Google Scholar] [CrossRef]
  359. Sha, Y.; Xu, X.; Li, L.; Cai, R.; Shao, Z. Hierarchical carbon-coated acanthosphere-like Li4Ti5O12 microspheres for high-power lithium-ion batteries. J. Power Sources 2016, 314, 18–27. [Google Scholar] [CrossRef]
  360. Wang, M.; Fang, P.G.; Chen, Y.; Leng, X.Y.; Yan, Y.; Yang, S.B.; Xu, P.; Yan, C. Synthesis of highly stable LTO/rGO/SnO2 nanocomposite via in situ electrostatic self-assembly for high-performance lithium-ion batteries. Adv. Func. Mater. 2023, 33, 2213902. [Google Scholar] [CrossRef]
  361. Zhu, M.; Deng, X.; Li, W.; He, M.; Xiong, D.; Feng, Z. Porous layered SnO2-LTO@C as a high-performance anode material for lithium-ion batteries. Ionics 2022, 28, 3699–3708. [Google Scholar] [CrossRef]
  362. Fang, W.; Ma, Y.; Zuo, P.; Cheng, X.; Yin, G. Nano-Li4Ti5O12 pore microspheres: A high power electrode material for lithium ion batteries. Int. J. Electrochem. Sci. 2013, 8, 1949–1956. [Google Scholar] [CrossRef]
  363. Kim, S.; Fang, S.H.; Zhang, Z.X.; Chen, J.Z.; Yang, L.; Penner-Hahn, J.E.; Deb, A. The electrochemical and local structural analysis of the mesoporous Li4Ti5O12 anode. J. Power Sources 2014, 268, 294–300. [Google Scholar] [CrossRef]
  364. Wu, F.; Li, X.; Wang, Z.; Guo, H. Petal-like Li4Ti5O12–TiO2 nanosheets as high-performance anode materials for Li-ion batteries. Nanoscale 2013, 5, 6936–6943. [Google Scholar] [CrossRef] [PubMed]
  365. Chen, C.; Huang, Y.; Zhang, H.; Wang, X.; Li, G.; Wang, Y.; Jiao, L.; Yuan, H. Small amount of reduced graphene oxide modified Li4Ti5O12 nanoparticles for ultrafast high-power lithium ion battery. J. Power Sources 2015, 278, 693–702. [Google Scholar] [CrossRef]
  366. Yi, T.-F.; Yang, S.-Y.; Zhu, Y.-R.; Ye, M.-F.; Xie, Y.; Zhu, R.-S. Enhanced rate performance of Li4Ti5O12 anode material by ethanol-assisted hydrothermal synthesis for lithium-ion battery. Ceram. Int. 2014, 40, 9853–9858. [Google Scholar] [CrossRef]
  367. Zhang, W.; Li, J.; Guan, Y.; Jin, Y.; Zhu, W.; Guo, X.; Qiu, X. Nano-Li4Ti5O12 with high rate performance synthesized by a glycerol assisted hydrothermal method. J. Power Sources 2013, 243, 661–667. [Google Scholar] [CrossRef]
  368. Li, X.-P.; Mao, J. Sol-hydrothermal synthesis of Li4Ti5O12/rutile-TiO2 composite as high rate anode material for lithium ion batteries. Ceram. Int. 2014, 40, 13553–13558. [Google Scholar] [CrossRef]
  369. Zhang, Y.; Zhang, Y.; Huang, L.; Zhou, Z.; Wang, J.; Liu, H.; Wu, H. Hierarchical carambola-like Li4Ti5O12-TiO2 composites as advanced anode materials for lithium-ion batteries. Electrochim. Acta 2016, 195, 124–133. [Google Scholar] [CrossRef]
  370. Chou, S.-L.; Wang, J.-Z.; Liu, H.-K.; Dou, S.-X. Rapid synthesis of Li4Ti5O12 microspheres as anode materials and its binder effect for lithium-ion battery. J. Phys. Chem. C 2011, 115, 16220–16227. [Google Scholar] [CrossRef]
  371. Hui, Y.; Cao, L.; Xu, Z.; Huang, J.; Ouyang, H.; Li, J. Mesoporous Li4Ti5O12 nanoparticles synthesized by a microwave-assisted hydrothermal method for high rate lithium-ion batteries. J. Electroanal. Chem. 2016, 763, 45–50. [Google Scholar] [CrossRef]
  372. Cheng, J.; Che, R.; Liang, C.; Liu, J.; Wang, M.; Xu, J. Hierarchical hollow Li4Ti5O12 urchin-like microspheres with ultra-high specific surface area for high rate lithium ion batteries. Nano Res. 2014, 7, 1043–1053. [Google Scholar] [CrossRef]
  373. Li, J.; Huang, S.; Xu, S.; Lan, L.; Lu, L.; Li, S. Synthesis of spherical silver-coated Li4Ti5O12 anode material by a sol-gel-assisted hydrothermal method. Nanoscale Res. Lett. 2017, 12, 576. [Google Scholar] [CrossRef]
  374. Pryono, B.; Murti, P.B.; Syahrial, A.Z.; Subhan, A. Optimizing the performance of Li4Ti5O12 anode synthesized from TiO2 xerogel and LiOH with hydrothermal-ball mill method by using acetylene black. In AIP Conference Proceedings; AIP Publishing: College Park, MD, USA, 2017; Volume 1826, p. 020005. [Google Scholar]
  375. Gao, L.; Wang, L.; Dai, S.; Cao, M.; Zhong, Z.; Shen, Y.; Wang, M. Li4Ti5O12-TiO2 nanowire arrays constructed with stacked nanocrystals for high-rate lithium and sodium ion batteries. J. Power Sources 2017, 344, 223–232. [Google Scholar] [CrossRef]
  376. Ma, G.; Cheng, M. Hydrothermal method preparing lithium ion battery anode material Li4Ti5O12 using metatitanic acid. Ferroelectrics 2018, 536, 181–186. [Google Scholar] [CrossRef]
  377. Nugroho, A.; Kim, S.J.; Chung, K.Y.; Cho, B.-W.; Lee, Y.-W.; Kim, J. Facile synthesis of nanosized Li4Ti5O12 in supercritical water. Electrochem. Commun. 2011, 13, 650–653. [Google Scholar] [CrossRef]
  378. Nugroho, A.; Kim, S.J.; Chung, K.Y.; Kim, J. Synthesis of Li4Ti5O12 in supercritical water for Li-ion batteries: Reaction mechanism and high-rate performance. Electrochim. Acta 2012, 78, 623–632. [Google Scholar] [CrossRef]
  379. Hayashi, H.; Nakamura, T.; Ebina, T. Hydrothermal synthesis of Li4Ti5O12 nanoparticles using a supercritical flow reaction system. J. Ceram. Soc. Jpn 2014, 122, 78–82. [Google Scholar] [CrossRef]
  380. Nugroho, A.; Kim, S.J.; Chang, W.; Chung, K.Y. Facile synthesis of hierarchical mesoporous Li4Ti5O12 microspheres in supercritical methanol. J. Power Sources 2013, 244, 164–169. [Google Scholar] [CrossRef]
  381. Laumann, A.; Bremholm, M.; Hald, P.; Holzapfel, M.; Fehr, K.T.; Iversen, B.B. Rapid green continuous flow supercritical synthesis of high performance Li4Ti5O12 nanocrystals for Li ion battery applications. J. Electrochem. Soc. 2012, 159, A166–A171. [Google Scholar] [CrossRef]
  382. Mathew, V.; Lim, J.; Gim, J.; Kim, D.; Moon, J.; Kang, J.; Kim, J. Optimized Li4Ti5O12 nanoparticles by solvothermal route for Li-ion batteries. J. Nanosci. Nanotechnol. 2011, 11, 7294–7298. [Google Scholar] [CrossRef]
  383. Li, N.; Liang, J.; Wei, D.; Zhu, Y.; Qian, Y. Solvothermal synthesis of micro-/nanoscale Cu/Li4Ti5O12 composites for high-rate Li-ion batteries. Electrochim. Acta 2014, 123, 346–352. [Google Scholar] [CrossRef]
  384. Zhang, Y.-Q.; Zeng, M.; Wu, X.; Bai, Y.; Li, J. Solvothermal synthesis for dual-phase Li4Ti5O12/TiO2 composites for high-stability lithium storage. Mater. Res. Express 2020, 7, 015515. [Google Scholar] [CrossRef]
  385. Chen, J.Z.; Yang, L.; Fang, S.H.; Tang, Y.F. Synthesis of sawtooth-like Li4Ti5O12 nanosheets as anode materials for Li-ion batteries. Electrochim. Acta 2010, 55, 6596–6600. [Google Scholar] [CrossRef]
  386. Kageyama, H.; Oaki, Y.; Imai, H. Basicity-controlled synthesis of Li4Ti5O12 nanocrystals by a solvothermal method. RSC Adv. 2014, 4, 44124–44129. [Google Scholar] [CrossRef]
  387. Feckl, J.M.; Fominykh, K.; Döblinger, M.; Fattakhova-Rohlfing, D.; Bein, T. Nanoscale porous framework of lithium titanate for ultrafast lithium insertion. Angew. Chem. Int. Ed. 2012, 51, 7459–7463. [Google Scholar] [CrossRef]
  388. Kim, H.-K.; Roh, K.C.; Kang, K.; Kim, K.-B. Synthesis of nano-Li4Ti5O12 decorated on non-oxidized carbon nanotubes with enhanced rate capability for lithium-ion batteries. RSC Adv. 2013, 3, 14267–14272. [Google Scholar] [CrossRef]
  389. Gangaja, B.; Nair, S.; Santhanagopalan, D. Surface-engineered Li4Ti5O12 nanostructures for high-power Li-ion batteries. Nano-Micro Lett. 2020, 30, 12. [Google Scholar] [CrossRef]
  390. Santhoshkumar, P.; Shaji, N.; Sim, G.S.; Nanthagopal, M.; Lee, C.W. Towards environment-friendly and versatile energy storage devices: Design and preparation of mesoporous Li4Ti5O12-TiO2 nano-hybrid electrode materials. Appl. Surf. Sci. 2022, 583, 152490. [Google Scholar] [CrossRef]
  391. Ho, C.K.; Li, C.Y.V.; Chan, K.Y.; Yung, H.; Tay, Y.Y. Interfacing TiO2(B) nanofibers with Li4Ti5O12 towards highly reversible and durable TiO2-based anode for Li−ion batteries. Energy Technol. 2019, 7, 107–112. [Google Scholar] [CrossRef]
  392. Lu, S.; Shang, Y.; Zheng, W.; Huang, Y.; Wang, R.; Zeng, W.; Zhan, H.; Yang, Y.; Mei, J. TiO2(B) nanosheets modified Li4Ti5O12 microsphere anode for high-rate lithium-ion batteries. Nanotechnology 2022, 33, 245404. [Google Scholar] [CrossRef]
  393. Singhai, A.; Skandan, G. Nanostructured Li4Ti5O12 Powders and Method of Making the Same. US Patent 6,827,921, 7 December 2004. [Google Scholar]
  394. Meng, T.; Zeng, R.; Sun, Z.; Yi, F.; Shu, D.; Li, K.; Li, S.; Zhang, F.; Cheng, H.; He, C. Chitosan-confined synthesis of N-doped and carbon-coated Li4Ti5O12 nanoparticles with enhanced lithium storage for lithium-ion batteries. J. Electrochem. Soc. 2018, 165, A1046–A1053. [Google Scholar] [CrossRef]
  395. Yin, Y.; Luo, X.; Xu, B. In-situ self-assembly synthesis of low-cost, long-life, shape-controllable spherical Li4Ti5O12 anode material for Li-ion batteries. J. Alloys Compd. 2022, 904, 164026. [Google Scholar] [CrossRef]
  396. Dong, H.; Yin, Y.; Zhang, Z.; Yang, S. Synthesis and properties of Li4Ti5O12/C composite by a microwave-assisted method using PAM as both the template and the carbon source. Phys. Scr. 2012, 86, 055802. [Google Scholar] [CrossRef]
  397. Hao, X.; Bartlett, B.M. Li4Ti5O12 nanocrystals synthesized by carbon templating from solution precursors yield high performance thin film Li-ion battery electrodes. Adv. Energy Mater. 2013, 3, 753–761. [Google Scholar] [CrossRef]
  398. Lee, S.J.; Jung, K.H.; Park, B.G.; Kim, H.G.; Park, Y.J. Synthesis of Li4Ti5O12 thin film with inverse hemispheric structure. Bull. Korean Chem. Soc. 2010, 31, 360–364. [Google Scholar] [CrossRef]
  399. Liu, J.; Wei, A.X.; Chen, M.; Xia, X. Rational synthesis of Li4Ti5O12/N-C nanotube arrays as advanced high-rate electrodes for lithium-ion batteries. J. Mater. Chem. A 2018, 6, 3857–3863. [Google Scholar] [CrossRef]
  400. Singh, D.P.; Mulder, F.M.; Wagemaker, M. Templated spinel Li4Ti5O12 Li-ion battery electrodes combining high rates with high energy density. Electrochem. Commun. 2013, 35, 124–127. [Google Scholar] [CrossRef]
  401. Vertruyen, B.; Eshraghi, N.; Piffet, C.; Bodart, J.; Mahmound, A.; Boschini, F. Spray-drying of electrode materials for lithium- and sodium-ion batteries. Materials 2018, 11, 1076. [Google Scholar] [CrossRef]
  402. Nowack, L.V.; Bunjaku, T.; Wegner, K.; Pratsinis, S.E.; Luisier, M.; Wood, V. Design and fabrication of microspheres with hierarchical internal structure for tuning performance. Adv. Sci. 2015, 2, 1500078. [Google Scholar] [CrossRef]
  403. Wu, F.; Wang, Z.; Li, X.; Guo, H.; Yue, P.; Xiong, X.; He, Z.; Zhang, Q. Characterization of spherical-shaped Li4Ti5O12 prepared by spray-drying. Electrochim. Acta 2012, 78, 331–339. [Google Scholar] [CrossRef]
  404. Alaboina, P.K.; Ge, Y.; Uddin, M.-J.; Liu, Y.; Lee, D.; Park, S.; Zhang, X.; Cho, S.-J. Nanoscale porous lithium titanate anode for superior high temperature performance. ACS Appl. Mater. Interfaces 2016, 8, 12127–12133. [Google Scholar] [CrossRef] [PubMed]
  405. Dai, C.; Ye, J.; Zhao, S.; He, P.; Zhou, H. Fabrication of high-energy Li-ion cells with Li4Ti5O12 microspheres as anode and 0.5 Li2MnO3·0.5 LiNi0.4Co0.2Mn0.4O2 microspheres as cathode. Chem. Asian J. 2016, 11, 1273–1280. [Google Scholar] [CrossRef]
  406. Deng, L.; Yang, W.-H.; Zhou, S.-X.; Chen, J.-T. Effect of carbon nanotubes addition on electrochemical performance and thermal stability of Li4Ti5O12 anode in commercial LiMn2O4/Li4Ti5O12 full-cell. Chin. Chem. Lett. 2015, 26, 1529–1534. [Google Scholar] [CrossRef]
  407. Gao, J.; Jiang, C.; Wan, C. Influence of carbon additive on the properties of spherical Li4Ti5O12 and LiFePO4 materials for lithium-ion batteries. Ionics 2010, 16, 417–424. [Google Scholar] [CrossRef]
  408. He, Z.; Wang, Z.; Wu, F.; Guo, H.; Li, X.; Xiong, X. Spherical Li4Ti5O12 synthesized by spray-drying from a different kind of solution. J. Alloys Compd. 2012, 540, 39–45. [Google Scholar] [CrossRef]
  409. Hsiao, K.-C.; Liao, S.-C.; Chen, J.-M. Microstructure effect on the electrochemical property of Li4Ti5O12 as an anode material for lithium-ion batteries. Electrochim. Acta 2008, 53, 7242–7247. [Google Scholar] [CrossRef]
  410. Hsieh, C.-T.; Chen, I.-L.; Jiang, Y.-R.; Lin, J.-Y. Synthesis of spinel lithium titanate anodes incorporated with rutile titania nanocrystallites by spray-drying followed by calcination. Solid State Ion. 2011, 201, 60–67. [Google Scholar] [CrossRef]
  411. Jung, H.-G.; Kim, J.; Scrosati, B.; Sun, Y.-K. Micron-sized, carbon-coated Li4Ti5O12 as high power anode material for advanced lithium batteries. J. Power Sources 2011, 196, 7763–7766. [Google Scholar] [CrossRef]
  412. Kadoma, Y.; Chiba, Y.; Yoshikawa, D.; Mitobe, Y.; Kumagai, N.; Ui, K. Influence of the carbon source on the surface and electrochemical characteristics of lithium excess Li4.3Ti5O12 carbon composite. Electrochemistry 2012, 80, 759–761. [Google Scholar] [CrossRef]
  413. Lee, B.; Yoon, J.R. Synthesis of high-performance Li4Ti5O12 and its application to the asymmetric hybrid capacitor. Electron. Mater. Lett. 2013, 9, 871–873. [Google Scholar] [CrossRef]
  414. Li, C.; Li, G.; Wen, S.; Ren, R. Spray-drying synthesis and characterization of Li4Ti5O12 anode material for lithium ion batteries. J. Adv. Oxid. Technol. 2017, 20, 183. [Google Scholar] [CrossRef]
  415. Lu, X.; Gu, L.; Hu, Y.-S.; Chiu, H.-C.; Li, H.; Demopoulos, G.P.; Chen, L. New insight into the atomic-scale bulk and surface structure evolution of Li4Ti5O12 anode. J. Am. Chem. Soc. 2015, 137, 1581–1586. [Google Scholar] [CrossRef] [PubMed]
  416. Ren, J.; Ming, H.; Jia, Z.; Zhang, Y.; Ming, J.; Zhou, Q.; Zheng, J. High tap density Li4Ti5O12 microspheres: Synthetic conditions and advanced electrochemical performance. Energy Technol. 2017, 5, 1680–1686. [Google Scholar] [CrossRef]
  417. Ruan, D.; Kim, M.-S.; Yang, B.; Qin, J.; Kim, K.-B.; Lee, S.-H.; Liu, Q.; Tan, L.; Qiao, Z. 700 F hybrid capacitors cells composed of activated carbon and Li4Ti5O12 microspheres with ultra-long cycle life. J. Power Sources 2017, 366, 200–206. [Google Scholar] [CrossRef]
  418. Wu, F.; Li, X.; Wang, Z.; Guo, H.; He, Z.; Zhang, Q.; Xiong, X.; Yue, P. Low-temperature synthesis of nano-micron Li4Ti5O12 by an aqueous mixing technique and its excellent electrochemical performance. J. Power Sources 2012, 202, 374–379. [Google Scholar] [CrossRef]
  419. Yoshikawa, D.; Suzuki, N.; Kadoma, Y.; Ui, K.; Kumagai, N. Li excess Li4+xTi5–xO12–δ/C composite using spray-drying method and its electrode properties. Funct. Mater. Lett. 2012, 5, 1250001. [Google Scholar] [CrossRef]
  420. Zhang, Q.; Peng, W.; Wang, Z.; Li, X.; Xiong, X.; Guo, H.; Wang, Z.; Wu, F. Li4Ti5O12/reduced graphene oxide composite as a high rate capability material for lithium ion batteries. Solid State Ion. 2013, 236, 30–36. [Google Scholar] [CrossRef]
  421. Zheng, X.; Dong, L.; Dong, C. Easy synthesis of Li4Ti5O12/C microspheres containing nanoparticles as anode material for high-rate lithium batteries. Surf. Rev. Lett. 2014, 21, 1450023. [Google Scholar] [CrossRef]
  422. Zhu, G.-N.; Liu, H.-J.; Zhuang, J.-H.; Wang, C.-X.; Wang, Y.-G.; Xia, Y.-Y. Carbon-coated nano-sized Li4Ti5O12 nanoporous micro-sphere as anode material for high-rate lithium-ion batteries. Energy Environ. Sci. 2011, 4, 4016. [Google Scholar] [CrossRef]
  423. Zhu, W.; Zhuang, Z.; Yang, Y.; Zhang, R.; Lin, Z.; Lin, Y.; Huang, Z. Synthesis and electrochemical performance of hole-rich Li4Ti5O12 anode material for lithium-ion secondary batteries. J. Phys. Chem. Solids 2016, 93, 52–58. [Google Scholar] [CrossRef]
  424. Wen, S.; Li, G.; Ren, R.; Li, C. Preparation of spherical Li4Ti5O12 anode materials by spray drying. Mater. Lett. 2015, 148, 130–133. [Google Scholar] [CrossRef]
  425. Xu, G.; Quan, X.; Gao, H.; Li, J.; Cai, Y.; Cheng, X.; Guo, L. Facile spray drying route for large scale nitrogen-doped carbon-coated Li4Ti5O12 anode material in lithium-ion batteries. Solid State Ion. 2017, 304, 40–45. [Google Scholar] [CrossRef]
  426. Yoshikawa, D.; Kadoma, Y.; Kima, J.-M.; Ui, K.; Kumagai, N.; Kitamura, N.; Idemoto, Y. Spray-drying synthesized lithium-excess Li4+xTi5–xO1–δ and its electrochemical property as negative electrode material for Li-ion batteries. Electrochim. Acta 2010, 55, 1872–1879. [Google Scholar] [CrossRef]
  427. Lee, B.-G.; Lee, S.-H. Application of hybrid supercapacitor using granule Li4Ti5O12/activated carbon with variation of current density. J. Power Sources 2017, 343, 545–549. [Google Scholar] [CrossRef]
  428. Kumagai, N.; Yoshikawa, D.; Kadoma, Y.; Ui, K. Spray-drying synthesized lithium-excess Li4+xTi4.95−xNb0.05O12−δ and its electrochemical property as negative electrode material for Li-ion batteries. Electrochemistry 2010, 78, 754–756. [Google Scholar] [CrossRef]
  429. Liu, W.; Wang, Q.; Cao, C.; Han, X.; Zhang, J.; Xie, X.; Xia, B. Spray drying of spherical Li4Ti5O12/C powders using polyvinyl pyrrolidone as binder and carbon source. J. Alloys Compd. 2015, 621, 162–169. [Google Scholar] [CrossRef]
  430. Piffet, C.; Vertruyen, B.; Caes, S.; Thomassin, J.-M.; Broze, G.; Malherbe, C.; Boschini, F.; Cloots, R.; Mahmoud, A. Aqueous processing of flexible, free-standing Li4Ti5O12 electrodes for Li-ion batteries. Chem. Eng. J. 2020, 397, 125508. [Google Scholar] [CrossRef]
  431. Yuan, T.; Li, W.-T.; Zhang, W.; He, Y.-S.; Zhang, C.; Liao, X.-Z.; Ma, Z.-F. One-pot spray dried graphene sheets encapsulated nano-Li4Ti5O12 microspheres for a hybrid BatCap system. Ind. Eng. Chem. Res. 2014, 53, 10849–10857. [Google Scholar] [CrossRef]
  432. Wu, F.; Li, X.; Wang, Z.; Guo, H. Synthesis of chromium-doped lithium titanate microspheres as high-performance anode material for lithium ion batteries. Ceram. Int. 2014, 40, 13195–13204. [Google Scholar] [CrossRef]
  433. Jia, X.L.; Kan, Y.F.; Zhu, X.; Ning, G.Q.; Lu, Y.F.; Wei, F. Building flexible Li4Ti5O12/CNT lithium-ion battery anodes with superior rate performance and ultralong cycling stability. Nano Energy 2014, 10, 344–352. [Google Scholar] [CrossRef]
  434. Wang, Q.; Geng, J.; Yuan, C.; Kuai, L.; Geng, B. Mesoporous spherical Li4Ti5O12/TiO2 composites as an excellent anode material for lithium ion batteries. Electrochim. Acta 2016, 212, 41–46. [Google Scholar] [CrossRef]
  435. Park, J.-H.; Kang, S.-W.; Kwon, T.-S.; Park, H.S. Spray-drying assisted synthesis of a Li4Ti5O12/C composite for high rate performance lithium ion batteries. Ceram. Int. 2018, 44, 2683–2690. [Google Scholar] [CrossRef]
  436. Lee, G.-W.; Kim, M.-S.; Jeong, J.H.; Roh, H.-K.; Roh, K.C.; Kim, K.-B. Comparative study of Li4Ti5O12 composites prepared with pristine, oxidized, and surfactant-treated multiwalled carbon nanotubes for high-power hybrid supercapacitors. ChemElectroChem 2018, 5, 2357–2366. [Google Scholar] [CrossRef]
  437. Chien, W.-C.; Wu, Z.-H.; Hsieh, Y.-C.; Wu, Y.-S.; Wu, S.-H.; Yang, C.-C. Electrochemical performance of Li4Ti5O12 anode materials synthesized using a spray-drying method. Ceram. Int. 2020, 46, 26923–26935. [Google Scholar] [CrossRef]
  438. Hsieh, C.-T.; Lin, J.-Y. Influence of Li addition on charge-discharge behavior of spinel lithium titanate. J. Alloys Compd. 2010, 506, 231–236. [Google Scholar] [CrossRef]
  439. Jamin, C.; Jungers, T.; Piffet, C.; Mahmoud, A.; Cloots, R.; Vertruyen, B.; Boschini, F. Li4Ti5O12 powders by spray-drying: Influence of the solution concentration and particle size on the electrochemical properties. IOP Conf. Ser. J. Phys. 2018, 1081, 012001. [Google Scholar] [CrossRef]
  440. Lenggoro, I.W.; Hata, T.; Iskandar, F.; Lunden, M.M.; Okuyama, K. An experimental and modeling investigation of particle production by spray pyrolysis using a laminar flow aerosol reactor. J. Mater. Res. 2000, 15, 733–743. [Google Scholar] [CrossRef]
  441. Zhu, Y.; Choi, S.H.; Fan, X.; Shin, J.; Ma, Z.; Zachariah, M.R.; Choi, J.W.; Wang, C. Recent progress on spray pyrolysis for gigh performance electrode materials in lithium and sodium rechargeable batteries. Adv. Energy Mater. 2017, 7, 1601578. [Google Scholar] [CrossRef]
  442. Laine, R.M.; Baranwal, R.; Hinklin, T.; Treadwell, D.; Sutorik, A.; Bickmore, C.; Waldner, K.; Neo, S.S. Making nanosized oxide powders from precursors by flame spray pyrolysis. Key Eng. Mater. 1999, 159, 17–24. [Google Scholar] [CrossRef]
  443. Ogihara, T.; Yamada, M.; Fujita, A.; Akao, S.; Myoujin, K. Effect of organic acid on the electrochemical properties of Li4Ti5O12/C composite powders synthesized by spray pyrolysis. Mater. Res. Bull. 2011, 46, 796–800. [Google Scholar] [CrossRef]
  444. Karhunen, T.; Lähde, A.; Leskinen, J.; Büchel, R.; Waser, O.; Tapper, U.; Jokiniemi, J. Transition metal-doped lithium titanium oxide nanoparticles made using flame spray pyrolysis. ISRN Nanotechnol. 2011, 2011, 180821. [Google Scholar] [CrossRef]
  445. Waser, O.; Brenner, O.; Groehn, A.J.; Pratsinis, S.E. Process design for size-controlled flame spray synthesis of Li4Ti5O12 and electrochemical performance. Chem. Proc. Eng. 2017, 38, 51–66. [Google Scholar] [CrossRef]
  446. Ju, S.H.; Kang, Y.C. Effects of drying control chemical additive on properties of Li4Ti5O12 negative powders prepared by spray pyrolysis. J. Power Sources 2010, 195, 4327–4331. [Google Scholar] [CrossRef]
  447. Takahashi, M.; Tani, J.; Kido, H.; Hayashi, A.; Tadanaga, K.; Tatsumisago, M. Thin film electrode materials Li4Ti5O12 and LiCoO2 prepared by spray pyrolysis method. IOP Conf. Ser. Mater. Sci. Eng. 2011, 18, 122004. [Google Scholar] [CrossRef]
  448. Yamada, M.; Kodera, T.; Ogihara, T. Synthesis and electrochemical properties of C/Li4Ti5O12 powders by spray pyrolysis using aqueous solution of organic acid. Electrochemistry 2010, 78, 463–466. [Google Scholar] [CrossRef]
  449. Kim, M.H.; Kang, Y.C. Electrochemical properties of nanosized Li4Ti5O12 powders prepared by flame spray pyrolysis. Int. J. Electrochem. Sci. 2013, 8, 3379–3389. [Google Scholar] [CrossRef]
  450. Du, G.; Winton, B.R.; Hashim, I.M.; Sharma, N.; Konstantinov, K.; Reddy, M.V.; Guo, Z. Mass production of Li4Ti5O12 with a conductive network via in-situ spray pyrolysis as a long cycle life, high rate anode material for lithium ion batteries. RSC Adv. 2014, 4, 38568–38574. [Google Scholar] [CrossRef]
  451. Birrozzi, A.; Copley, M.; von Zamory, J.; Pasqualini, M.; Calcaterra, S.; Nobili, F.; Di Cicco, A.; Rajantie, H.; Briceno, M.; Bilbe, E.; et al. Scaling up “nano” Li4Ti5O12 for high-power lithium-ion anodes using large scale flame spray pyrolysis. J. Electrochem. Soc. 2015, 162, A2331–A2338. [Google Scholar] [CrossRef]
  452. Zhang, X.; Xu, H.; Zhao, Y.; Zhu, G.; Yu, A. A facile one-step spray pyrolysis method to synthesize spherical Li4Ti5O12 for lithium-ion battery. Mater. Lett. 2014, 129, 101–103. [Google Scholar] [CrossRef]
  453. Yang, K.M.; Ko, Y.N.; Yun, J.-Y.; Kang, Y.C. Preparation of Li4Ti5O12 yolk–shell powders by spray pyrolysis and their electrochemical properties. Chem. Asian J. 2014, 9, 443–446. [Google Scholar] [CrossRef]
  454. Kim, J.H.; Yoon, J.R. Preparation and characterization of Li4Ti5O12 synthesized using hydrogen titanate nanowire for hybrid super capacitor. J. Adv. Ceram. 2013, 2, 285–290. [Google Scholar] [CrossRef]
  455. Bhat, M.H.; Miura, A.; Vinatier, P.; Levasseur, A.; Rao, K.J. Microwave synthesis of lithium lanthanum titanate. Solid State Commun. 2003, 125, 557–562. [Google Scholar] [CrossRef]
  456. Li, J.; Jin, Y.-L.; Zhang, X.-G.; Yang, H. Microwave solid-state synthesis of spinel Li4Ti5O12 nanocrystallites as anode material for lithium-ion batteries. Solid State Ion. 2007, 178, 1590–1594. [Google Scholar] [CrossRef]
  457. Balaji, S.; Mutharasu, D.; Sankara Subramanian, N.; Ramanathan, K. A review on microwave synthesis of electrode materials for lithium-ion batteries. Ionics 2009, 15, 765–777. [Google Scholar] [CrossRef]
  458. Yang, L.H.; Dong, C.; Guo, J. Hybrid microwave synthesis and characterization of the compounds in the Li-Ti-O system. J. Power Sources 2008, 175, 575–580. [Google Scholar] [CrossRef]
  459. Liu, H.W.; Feng, C.Q.; Tang, H.; Song, L.; Zhang, K.L. Synthesis and electrochemical properties of indium doped spinel LiMn2O4. J. Mater. Sci. 2004, 15, 495–497. [Google Scholar] [CrossRef]
  460. Liu, H.; Feng, Y.; Wang, K.; Xie, J. Synthesis and electrochemical properties of Li4Ti5O12/C composite by the PVB rheological phase method. J. Phys. Chem. Solids 2008, 69, 2037–2040. [Google Scholar] [CrossRef]
  461. Wang, Z.; Xie, G.; Gao, L. Electrochemical characterization of Li4Ti5O12/C anode material prepared by starch-sol-assisted rheological phase method for Li-ion battery. J. Nanomater. 2012, 2012, 876197. [Google Scholar] [CrossRef]
  462. Abureden, S.; Hassan, F.M.; Lui, G.; Ahn, W.; Sy, S.; Yu, A.P.; Chen, Z.W. Multigrain electrospun nickel doped lithium titanate nanofibers with high power lithium ion storage. J. Mater. Chem. A 2016, 4, 12638–12647. [Google Scholar] [CrossRef]
  463. Zhu, N.; Liu, W.; Xue, M.; Xie, Z.; Zhao, D.; Zhang, M.; Chen, J.; Cao, T. Graphene as a conductive additive to enhance the high-rate capabilities of electrospun Li4Ti5O12 for lithium-ion batteries. Electrochim. Acta 2010, 55, 5813–5818. [Google Scholar] [CrossRef]
  464. Guo, B.; Li, Y.; Yao, Y.; Lin, Z.; Ji, L.; Xu, G.; Liang, Y.; Shi, Q.; Zhang, X. Electrospun Li4Ti5O12/C composites for lithium-ion batteries with high rate performance. Solid State Ion. 2011, 204–205, 61–65. [Google Scholar] [CrossRef]
  465. Wang, L.; Xiao, Q.; Li, Z.; Lei, G.; Zhang, P.; Wu, L. Synthesis of Li4Ti5O12 fibers as a high-rate electrode material for lithium-ion batteries. J. Solid State Electrochem. 2012, 16, 3307–3313. [Google Scholar] [CrossRef]
  466. Ji, X.; Li, D.; Lu, Q.; Guo, E.; Yao, L. Electrospinning preparation of one-dimensional Ce3+-doped Li4Ti5O12 sub-microbelts for high-performance lithium-ion batteries. J. Nanopart. Res. 2017, 19, 393. [Google Scholar] [CrossRef]
  467. Li, X.; Yang, K.; Gao, F.; Hu, C.; Liu, H.; Li, D.; Wang, L. Electrospinning synthesis of spinel Li4Ti5O12 and its characterization. IOP Conf. Ser. Mater. Sci. Eng. 2015, 87, 012098. [Google Scholar] [CrossRef]
  468. Park, H.; Song, T.; Han, H.; Paik, U. Electrospun Li4Ti5O12 nanofibers sheathed with conductive TiN/TiOxNy layer as an anode material for high power Li-ion batteries. J. Power Sources 2013, 244, 726–730. [Google Scholar] [CrossRef]
  469. Chen, C.H.; Feng, J.J.; Shao, W.Q.; Chen, S.O.; He, L.Z.; Li, Y.Q.; Ban, Y.Q.; Hei, F.; Guo, X.J. Electrochemical performance of Li4Ti5O12/TiO2/Li0.4TiO2/C composite prepared by the electrospinning method using as lithium battery anode material. Mater. Sci. Forum 2016, 852, 959–962. [Google Scholar] [CrossRef]
  470. Zou, H.L.; Xiang, H.F.; Liang, X.; Feng, X.Y.; Cheng, S.; Jin, Y.; Chen, C.H. Electrospun Li3.9Cr0.3Ti4.8O12 nanofibers as anode material for high-rate and low-temperature lithium-ion batteries. J. Alloys Compd. 2017, 701, 99–106. [Google Scholar] [CrossRef]
  471. Kim, J.-G.; Shi, D.; Park, M.-S.; Jeong, G.; Heo, Y.-U.; Seo, M.; Kim, Y.-J.; Kim, J.H.; Dou, S.X. Controlled Ag-driven superior rate-capability of Li4Ti5O12 anodes for lithium rechargeable batteries. Nano Res. 2013, 6, 365–372. [Google Scholar] [CrossRef]
  472. Zhang, Z.; Deng, X.; Sunarso, J.; Cai, R.; Chu, S.; Miao, J.; Zhou, W.; Shao, Z. Two-step fabrication of Li4Ti5O12-coated carbon nanofibers as a flexible film electrode for high-power lithium-ion batteries. ChemElectroChem 2017, 4, 2286–2292. [Google Scholar] [CrossRef]
  473. Zhou, Y.; Xiao, S.; Jiang, J.; Wu, R.; Niu, X.; Chen, J.S. In-situ construction of Li4Ti5O12/rutile TiO2 heterostructured nanorods for robust and high-power lithium storage. Nano Res. 2023, 16, 1513–1521. [Google Scholar] [CrossRef]
  474. Wang, J.; Shen, L.; Li, H.; Ding, B.; Nie, P.; Dou, H.; Zhang, X. Mesoporous Li4Ti5O12/carbon nanofibers for high-rate lithium-ion batteries. J. Alloys Compd. 2014, 587, 171–176. [Google Scholar] [CrossRef]
  475. Liu, C.; Wang, S.; Zhang, C.; Fu, H.; Nan, X.; Yang, Y. High power high safety battery with electrospun Li3V2(PO4)3 cathode and Li4Ti5O12 anode with 95% energy efficiency. Energy Storage Mater. 2016, 5, 93–102. [Google Scholar] [CrossRef]
  476. Wu, Y.; Reddy, M.V.; Chowdari, B.V.R.; Ramakrishna, S. Electrochemical studies on electrospun Li(Li1/3Ti5/3)O4 grains as an anode for Li-ion batteries. Electrochim. Acta 2012, 67, 33–40. [Google Scholar] [CrossRef]
  477. Kim, E.-K.; Choi, B.-H.; Jee, M.-J.; Kwon, Y.-J.; Seo, H.; Kim, Y.-J.; Kim, K.-B. Synthesis and electrochemical characteristics of Li4Ti5O12 nanofibers by hydrothermal method. J. Korean Ceram. Soc. 2010, 47, 627–632. [Google Scholar] [CrossRef]
  478. Kataoka, K.; Nagai, H.; Sotokawa, T.; Kumashiro, Y.; Ishibai, Y.; Akimoto, J. Ion-exchange synthesis and improved Li insertion property of lithiated H2Ti12O25 as a negative electrode material for lithium-ion batteries. J. Asian Ceram. Soc. 2016, 4, 75–80. [Google Scholar] [CrossRef]
  479. Zhang, J.; Zhang, F.; Li, J.; Cai, W.; Zhang, J.; Yu, L.; Jin, Z.; Zhang, Z. Preparation of Li4Ti5O12 by solution ion-exchange of sodium titanate nanotube and evaluation of electrochemical performance. J. Nanopart. Res. 2013, 15, 2005. [Google Scholar] [CrossRef]
  480. 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]
  481. Li, N.; Katase, T.; Zhu, Y.; Matsumoto, T.; Umemura, T.; Ikuhara, Y.; Ohta, H. Solid-liquid phase epitaxial growth of Li4Ti5O12 thin film. Appl. Phys. Express 2016, 9, 125501. [Google Scholar] [CrossRef]
  482. Wang, C.-L.; Liao, Y.C.; Hsu, F.C.; Tai, N.H.; Wu, M.K. Preparation and characterization of thin film Li4Ti5O12 electrodes by magnetron sputtering. J. Electrochem. Soc. 2005, 152, A653–A657. [Google Scholar] [CrossRef]
  483. Kumatani, A.; Shiraki, S.; Takagi, Y.; Suzuki, T.; Ohsawa, T.; Gao, X.; Ikuhara, Y.; Hitosugi, T. Epitaxial growth of Li4Ti5O12 thin films using RF magnetron sputtering. Jpn. J. Appl. Phys. 2014, 53, 058001. [Google Scholar] [CrossRef]
  484. Pagani, F.; Döbeli, M.; Battaglia, C. Lithium-ion transport in Li4Ti5O12 epitaxial thin films vs state of charge. Batter. Supercaps 2021, 4, 316–321. [Google Scholar] [CrossRef]
  485. Deng, J.Q.; Lu, Z.G.; Belharouak, I.; Amine, K.; Chung, C.Y. Preparation and electrochemical properties of Li4Ti5O12 thin film electrodes by pulsed laser deposition. J. Power Sources 2009, 193, 816–821. [Google Scholar] [CrossRef]
  486. Kumatani, A.; Ohsawa, T.; Shimizu, R.; Takagi, Y.; Shiraki, S.; Hitosugi, T. Growth processes of lithium titanate thin films deposited by using pulsed laser deposition. Appl. Phys. Lett. 2012, 101, 123103. [Google Scholar] [CrossRef]
  487. Yu, P.; Li, C.; Guo, X. Sodium storage and pseudocapacitance charge in textured Li4Ti5O12 thin films. J. Phys. Chem. C 2014, 118, 10616–10624. [Google Scholar] [CrossRef]
  488. Yu, X.; Wang, R.; He, Y.; Hu, Y.; Li, H.; Huang, X. Electrochromic behavior of transparent Li4Ti5O12/FTO electrode. Electrochem. Solid-State Lett. 2010, 13, J99–J101. [Google Scholar] [CrossRef]
  489. Deng, J.; Lu, Z.; Chung, C.Y.; Hanc, X.; Wang, Z.; Zhou, H. Electrochemical performance and kinetic behavior of lithium ion in Li4Ti5O12 thin film electrodes. Appl. Surf. Sci. 2014, 314, 936–941. [Google Scholar] [CrossRef]
  490. Zhao, M.; Lian, J.; Jia, Y.; Jin, K.; Xu, L.; Hu, Z.; Yang, X.; Kang, S. Investigation of the optical properties of LiTi2O4 and Li4Ti5O12 spinel films by spectroscopic ellipsometry. Opt. Mater. Express 2016, 6, 3366–3374. [Google Scholar] [CrossRef]
  491. Schichtel, P.; Geiß, M.; Leichtweiß, T.; Sann, J.; Weber, D.A.; Janek, J. On the impedance and phase transition of thin film all-solid-state batteries based on the Li4Ti5O12 system. J. Power Sources 2017, 360, 593–604. [Google Scholar] [CrossRef]
  492. Julien, C.M.; Mauger, A. Pulsed laser deposited films for energy storage and conversion: A review. Coatings 2018, 9, 386. [Google Scholar] [CrossRef]
  493. Rho, Y.H.; Kanamura, K.; Umegaki, T. Preparation of Li4/3Ti5/3O4 thin film anode with high electrochemical response for rechargeable lithium batteries by sol–gel method. Chem. Lett. 2001, 30, 1322–1323. [Google Scholar] [CrossRef]
  494. Rho, Y.H.; Kanamura, K.; Fujisaki, M.; Hamagami, J.; Suda, S.; Umegaki, T. Preparation of Li4Ti5O12 and LiCoO2 thin film electrodes from precursors obtained by sol-gel method. Solid State Ion. 2002, 151, 151–157. [Google Scholar] [CrossRef]
  495. Wude, F.; Berkemeier, F.; Schmitz, G. Lithium diffusion in sputter-deposited Li4Ti5O12 thin films. J. Power Sources 2012, 215, 109–115. [Google Scholar]
  496. Hirayama, M.; Kim, K.; Toujigamori, T.; Cho, W.; Kanno, R. Epitaxial growth and electrochemical properties of Li4Ti5O12 thin-film lithium battery anodes. Dalton Trans. 2011, 40, 2882–2887. [Google Scholar] [CrossRef]
  497. Kim, K.; Toujigamori, T.; Suzuki, K.; Taminato, S.; Tamura, K.; Mizuki, J.; Hirayama, M.; Kanno, R. Characterization of nano-sized epitaxial Li4Ti5O12 (110) film electrode for lithium batteries. Electrochemistry 2012, 80, 800–803. [Google Scholar] [CrossRef]
  498. Pfenninger, R.; Afyon, S.; Garbayo, I.; Struzik, M.; Rupp, J.L.M. Lithium titanate anode thin films for Li-ion solid state battery based on garnets. Adv. Funct. Mater. 2018, 28, 1800879. [Google Scholar] [CrossRef]
  499. Mesoraca, S.; Kleibeuker, J.E.; Prasad, B.; MacManus-Driscoll, J.L.; Blamire, M.G. Lithium outdiffusion in LiTi2O4 thin films grown by pulsed laser deposition. J. Cryst. Growth 2016, 454, 134–138. [Google Scholar] [CrossRef]
  500. Chopdekar, R.V.; Wong, F.J.; Takamura, Y.; Arenholz, E.; Suzuki, Y. Growth and characterization of superconducting spinel oxide LiTi2O4 thin films. Phys. C Supercond. 2009, 469, 1885–1891. [Google Scholar] [CrossRef]
  501. Pagani, F.; Stilp, E.; Pfenninger, R.; Reyes, E.R.; Remhof, A.; Balogh-Michels, Z.; Neels, A.; Sastre-Pellicer, J.; Stiefel, M.; Döbeli, M.; et al. Epitaxial thin films as a model system for Li-ion conductivity in Li4Ti5O12. ACS Appl. Mater. Interfaces 2018, 10, 44494–44500. [Google Scholar] [CrossRef] [PubMed]
  502. Cunha, D.M.; Hendriks, T.A.; Vasileiadis, A.; Vos, C.M.; Verhallen, T.; Singh, D.P.; Wagemaker, M.; Huijben, M. Doubling reversible capacities in epitaxial Li4Ti5O12 thin film anodes for microbatteries. ACS Appl. Energy Mater. 2019, 2, 3410–3418. [Google Scholar] [CrossRef]
  503. Daminabo, S.; Goel, S.; Grammatikos, S.; Yazdani Nezhad, H.; Thakur, V.K. Fused deposition modeling-based additive manufacturing (3D printing): Techniques for polymer material systems. Mater. Today Chem. 2020, 16, 100248. [Google Scholar] [CrossRef]
  504. Sun, K.; Wei, T.S.; Ahn, B.Y.; Seo, J.Y.; Dillon, S.J.; Lewis, J.A. 3D Printing of interdigitated Li-ion microbattery architectures. Adv. Mater. 2013, 25, 4539–4543. [Google Scholar] [CrossRef] [PubMed]
  505. Zhao, Y.; Liu, G.; Liu, L.; Jiang, Z. High-performance thin-film Li4Ti5O12 electrodes fabricated by using ink-jet printing technique and their electrochemical properties. J. Solid State Electrochem. 2009, 13, 705–711. [Google Scholar] [CrossRef]
  506. Chen, Q.; Xu, R.; He, Z.; Zhao, K.; Pan, L. Printing 3D gel polymer electrolyte in lithium-ion microbattery using stereolithography. J. Electrochem. Soc. 2017, 164, A1852–A1857. [Google Scholar] [CrossRef]
  507. Delannoy, P.E.; Riou, B.; Lestriez, B.; Guyomard, D.; Brousse, T.; Le Bideau, J. Toward fast and cost-effective ink-jet printing of solid electrolyte for lithium microbatteries. J. Power Sources 2015, 274, 1085–1090. [Google Scholar] [CrossRef]
  508. Zhou, L.; Ning, W.; Wu, C.; Zhang, D.; Wei, W.; Ma, J.; Li, C.; Chen, L. 3D-printed microelectrodes with a developed conductive network and hierarchical pores toward high areal capacity for microbatteries. Adv. Mater. Technol. 2018, 4, 1800402. [Google Scholar] [CrossRef]
  509. Wei, T.S.; Ahn, B.Y.; Grotto, J.; Lewis, J.A. 3D printing of customized Li-ion batteries with thick electrodes. Adv. Mater. 2018, 30, 1703027. [Google Scholar] [CrossRef]
  510. Wang, Y.; Chen, C.; Xie, H.; Gao, T.; Yao, Y.; Pastel, G.; Han, X.; Li, Y.; Zhao, J.; Fu, K.; et al. 3D-printed all-fiber Li-ion battery toward wearable energy storage. Adv. Funct. Mater. 2017, 27, 1703140. [Google Scholar] [CrossRef]
  511. Kohlmeyer, R.R.; Blake, A.J.; Hardin, J.O.; Carmona, E.A.; Carpena-Núñez, J.; Maruyama, B.; Daniel Berrigan, J.; Huang, H.; Durstock, M.F. Composite batteries: A simple yet universal approach to 3D printable lithium-ion battery electrodes. J. Mater. Chem. A 2016, 4, 16856–16884. [Google Scholar] [CrossRef]
  512. Ragones, H.; Menkin, S.; Kamir, Y.; Gladkikh, A.; Mukra, T.; Kosa, G.; Golodnitsky, D. Towards smart free form-factor 3D printable batteries. Sustain. Energy Fuels 2018, 2, 1542–1549. [Google Scholar] [CrossRef]
  513. Kim, H.; Choi, J.-M.; Kwon, M.-S.; Song, M.S.; Park, Y.; Doo, S.-G. Inkjet printed electrode for rechargeable lithium thin film battery. ECS Meet. Abstr. 2009, MA2009-02, 691. [Google Scholar] [CrossRef]
  514. Lawes, S.; Sun, Q.; Lushington, A.; Xiao, B.; Liu, Y.; Sun, X. Inkjet-printed silicon as high performance anodes for Li-ion batteries. Nano Energy 2017, 36, 313–321. [Google Scholar] [CrossRef]
  515. Viviani, P.; Gibertini, E.; Iervolino, F.; Levi, M.; Magagnin, L. Carbon additive effect on the electrochemical performances of inkjet printed thin-film Li4Ti5O12 electrodes. J. Manufact. Proc. 2021, 72, 411–418. [Google Scholar] [CrossRef]
  516. Zhao, E.; Qin, C.; Jung, H.-R.; Berdichevsky, G.; Nese, A.; Marder, S.; Yushin, G. Lithium titanate confined in carbon nanopores for asymmetric supercapacitors. ACS Nano 2016, 10, 3977–3984. [Google Scholar] [CrossRef]
  517. Zhu, J.; Chen, J.; Xu, H.; Sun, S.; Xu, Y.; Zhou, M.; Gao, X.; Sun, Z. Plasma-introduced oxygen defects confined in Li4Ti5O12 nanosheets for boosting lithium-ion diffusion. ACS Appl. Mater. Interfaces 2019, 11, 17384–17392. [Google Scholar] [CrossRef]
  518. Kitta, M.; Kojima, T.; Kataoka, R.; Yazawa, K.; Tada, K. Realizing the single-phase spinel-type sodium titanium oxide with the Li4Ti5O12-like structure for building stable sodium-ion batteries. ACS Appl. Mater. Interfaces 2020, 12, 9322–9331. [Google Scholar] [CrossRef]
  519. Ritala, M.; Leskela, M.; Rauhala, E.; Haussalo, P. Atomic layer epitaxy growth of TiN thin films. J. Electrochem. Soc. 1995, 142, 2731–2737. [Google Scholar] [CrossRef]
  520. Richard, M.N.; Fuller, E.W.; Dahn, J.R. the effect of ammonia reduction on the spinel electrode materials, LiMn2O4 and Li(Li1/3Mn5/3)O4. Solid State Ion. 1994, 73, 81–91. [Google Scholar] [CrossRef]
  521. Santucci, S.; Lozzi, L.; Passacantando, M.; Picozzi, P.; Alfonsetti, R.; Diamanti, R.; Moccia, G. Study by X-ray photoelectron spectroscopy and X-ray diffraction of the growth of TiN thin films obtained by nitridation of Ti layers. Thin Solid Films 1996, 290–291, 376–380. [Google Scholar] [CrossRef]
  522. Wan, Z.N.; Cai, R.; Jiang, S.M.; Shao, Z.P. Nitrogen- and TiN-modified Li4Ti5O12: One-step synthesis and electrochemical performance optimization. J. Mater. Chem. 2012, 22, 17773–17781. [Google Scholar] [CrossRef]
  523. Kahrisi, M.; Kashani, H.; Ghaffarinejad, A. Improving the cyclability and rate capability of Li4Ti5O12 anode material by Cu2+ and F co-doping. ChemistrySelect 2023, 8, e202204198. [Google Scholar] [CrossRef]
  524. Ali, B.; Muhammad, R.; Islam, M.; Anang, D.A.; Han, D.-S.; Moeez, I.; Chung, K.Y.; Cho, M.; Kim, J.Y.; Kim, M.G.; et al. Cd-doped Li4–xCdxTi5O12 (x = 0.20) as a high rate capable and stable anode material for lithium-ion batteries. ACS Appl. Energy Mater. 2023, 6, 4198–4210. [Google Scholar] [CrossRef]
  525. Noreochim, L.; Prabowo, R.S.; Widyastuti, W.; Susanti, D.; Subhan, A.; Idris, N.H. Enhanced high-rate capability of iodide-doped Li4Ti5O12 as an anode for lithium-ion batteries. Batteries 2023, 9, 38. [Google Scholar] [CrossRef]
  526. Qasim, K.F.; Mousa, M.A. Physicochemical properties of oriented crystalline assembled polyaniline/metal doped Li4Ti5O12 composites for Li-ion storage. J. Inorg. Organomet. Polym. 2023, 33, 2601–2617. [Google Scholar] [CrossRef]
  527. Yeo, S.; Raj, M.R.; Lee, G. Oxygen vacancy-modulated zeolitic Li4Ti5O12 microsphere anode for superior lithium-ion battery. Electrochim. Acta 2023, 441, 141809. [Google Scholar] [CrossRef]
  528. Tian, K.; Hui, X.; Wang, H.; Zhang, Z.; Zhang, L.; Wang, C.; Yin, L. In situ growth of Li4Ti5O12 nanoparticles on Ti3C2Tx MXene for efficient electron transfer as high-rate anode of Li ion battery. Electrochim. Acta 2022, 415, 140242. [Google Scholar] [CrossRef]
  529. Li, Y.; Zhang, W.; Lai, C.; Yang, T.; Chang, X.; Zhang, M.; Sheng, L.; Yang, Z.; Ye, D.; Huang, K.; et al. Ti3C2 MXene-derived Li4Ti5O12 nanoplates with in-situ formed carbon quantum dots for metal-ion battery anodes. J. Colloid Interface Sci. 2023, 629, 263–269. [Google Scholar] [CrossRef]
  530. Meng, Q.; Hao, Q.; Chen, F.; Wang, L.; Li, N.; Sun, X. Li4Ti5O12 hollow macrospheres combine high tap density and excellent low-temperature performance as anode materials for Li-ion batteries. Mater. Character. 2023, 203, 113089. [Google Scholar] [CrossRef]
  531. Liang, K.; Ren, Y. Research progress on lithium titanate as anode material for sodium-ion batteries. Mater. Rep. 2020, 34, 9041–9047. [Google Scholar]
  532. Doyle, M.; Newman, J. The use of mathematical modeling in the design of lithium/polymer battery systems. Electrochim. Acta 1995, 40, 2191–2196. [Google Scholar] [CrossRef]
  533. Christensen, J.; Srinivasan, V.; Newman, J. Optimization of lithium titanate electrodes for high-power cells. J. Electrochem. Soc. 2006, 153, A560–A565. [Google Scholar] [CrossRef]
  534. Kashkooli, A.G.; Lui, G.; Farhad, S.; Lee, D.U.; Feng, K.; Yu, A.; Chen, Z. Nano-particle size effect on the performance of Li4Ti5O12 spinel. Electrochim. Acta 2016, 196, 33–40. [Google Scholar] [CrossRef]
  535. Lim, J.; Choi, E.; Mathew, V.; Kim, D.; Ahn, D.; Gim, J.; Kang, S.-H.; Kim, J. Enhanced high-rate performance of Li4Ti5O12 nanoparticles for rechargeable Li-ion batteries. J. Electrochem. Soc. 2011, 158, A275–A280. [Google Scholar] [CrossRef]
Scheme 1. Graphical abstract giving a brief introduction about the focus of the review paper.
Scheme 1. Graphical abstract giving a brief introduction about the focus of the review paper.
Micromachines 15 00310 sch001
Figure 2. Raman spectra of spinel Li4Ti5O12 nanoparticles (a) compared with patterns of rutile and anatase TiO2 (b).
Figure 2. Raman spectra of spinel Li4Ti5O12 nanoparticles (a) compared with patterns of rutile and anatase TiO2 (b).
Micromachines 15 00310 g002
Figure 3. XRD pattern of the intermediate Li1.81H0.19Ti2O5·xH2O (white powder) hydrothermal product. Reproduced with permission from [215]. Copyright 2015 Elsevier.
Figure 3. XRD pattern of the intermediate Li1.81H0.19Ti2O5·xH2O (white powder) hydrothermal product. Reproduced with permission from [215]. Copyright 2015 Elsevier.
Micromachines 15 00310 g003
Figure 4. Schematic representation of the formation process of Li4Ti5O12 with different anatase crystallite sizes. When the temperature increases above 500 °C, Li2COO3 starts to react with the anatase at the particle interface forming a Li-rich Li2TiO3 phase. This process occurs at a steady rate until all the Li2COO3 articles have completely reacted at about 650 °C. The reaction conditions are based on results from anatase with crystallite sizes of ~50 nm (a) and ~20 nm (b). Reproduced from [220] with permission. Copyright 2014 American Chemical Society.
Figure 4. Schematic representation of the formation process of Li4Ti5O12 with different anatase crystallite sizes. When the temperature increases above 500 °C, Li2COO3 starts to react with the anatase at the particle interface forming a Li-rich Li2TiO3 phase. This process occurs at a steady rate until all the Li2COO3 articles have completely reacted at about 650 °C. The reaction conditions are based on results from anatase with crystallite sizes of ~50 nm (a) and ~20 nm (b). Reproduced from [220] with permission. Copyright 2014 American Chemical Society.
Micromachines 15 00310 g004
Figure 5. Schematic comparison of the synthesis mechanism for the solid-state reaction (left) and cellulose-assisted glycine-nitrate combustion (right). The solid-phase reaction between TiO2 and lithium compounds to form Li–Ti–O spinel and the phase transition of TiO2 from anatase to “inert” rutile phase occur competitively during both synthesis processes. Reproduced with permission from [217]. Copyright 2010 Elsevier.
Figure 5. Schematic comparison of the synthesis mechanism for the solid-state reaction (left) and cellulose-assisted glycine-nitrate combustion (right). The solid-phase reaction between TiO2 and lithium compounds to form Li–Ti–O spinel and the phase transition of TiO2 from anatase to “inert” rutile phase occur competitively during both synthesis processes. Reproduced with permission from [217]. Copyright 2010 Elsevier.
Micromachines 15 00310 g005
Figure 6. Typical solid-state reaction for the synthesis of LTO whiskers using TiO2-B and CH3COOLi·2H2O (ratio fixed at 1:1.05). The two steps (diffusion and reaction) lead to a low solid-state reaction temperature of 650 °C and maintain the morphologies of whiskers. Reproduced from [228]. Copyright 2016 Elsevier.
Figure 6. Typical solid-state reaction for the synthesis of LTO whiskers using TiO2-B and CH3COOLi·2H2O (ratio fixed at 1:1.05). The two steps (diffusion and reaction) lead to a low solid-state reaction temperature of 650 °C and maintain the morphologies of whiskers. Reproduced from [228]. Copyright 2016 Elsevier.
Micromachines 15 00310 g006
Figure 7. LTO nanoparticles prepared by a two-step sintering. (ac) TEM images of the samples for the pre-sintering temperature of 400, 450, and 500 °C, respectively. (df) SEM images of the samples sintered at 750 °C with the pre-sintered temperature of 400, 450, and 500 °C, respectively. XRD patterns of the samples sintered at 700 °C (g) and at 750 °C (h) with the pre-sintered temperature of 400 °C (black line), 450 °C (blue line), and 500 °C (red line), which show ● Li4Ti5O12 and ▲ Li2TiO3. Reproduced from [249]. Copyright the terms of the Creative Commons Attribution 4.0 International License.
Figure 7. LTO nanoparticles prepared by a two-step sintering. (ac) TEM images of the samples for the pre-sintering temperature of 400, 450, and 500 °C, respectively. (df) SEM images of the samples sintered at 750 °C with the pre-sintered temperature of 400, 450, and 500 °C, respectively. XRD patterns of the samples sintered at 700 °C (g) and at 750 °C (h) with the pre-sintered temperature of 400 °C (black line), 450 °C (blue line), and 500 °C (red line), which show ● Li4Ti5O12 and ▲ Li2TiO3. Reproduced from [249]. Copyright the terms of the Creative Commons Attribution 4.0 International License.
Micromachines 15 00310 g007aMicromachines 15 00310 g007b
Figure 8. Typical 3-solution sol–gel method (A,B,C baths) for the preparation of LTO nanoparticles using tetrabutyl titanate and lithium acetate as Ti and Li raw materials and citric acid as complexing agent.
Figure 8. Typical 3-solution sol–gel method (A,B,C baths) for the preparation of LTO nanoparticles using tetrabutyl titanate and lithium acetate as Ti and Li raw materials and citric acid as complexing agent.
Micromachines 15 00310 g008
Figure 10. (a) SEM images of LTO obtained after 12, 18, 24, and 30 h of high-energy ball-milling process. (b) Rate capability as a function of synthesis time. (c) Cycleability tests at 1C rate. Reproduced with permission from [325]. Copyright 2014 Elsevier.
Figure 10. (a) SEM images of LTO obtained after 12, 18, 24, and 30 h of high-energy ball-milling process. (b) Rate capability as a function of synthesis time. (c) Cycleability tests at 1C rate. Reproduced with permission from [325]. Copyright 2014 Elsevier.
Micromachines 15 00310 g010
Figure 11. (a) Schematic illustration of the feasible morphology-tailoring process of the bare LTO and LTO@C microspheres via hydrothermal treatment. SEM images of the TiOx·H2O precursor (bd), the bare LTO (eg), and the LTO@C (hj) samples with different magnifications. Inset in (b) shows a photograph of an acanthosphere. Reproduced with permission from [359]. Copyright 2016 Elsevier.
Figure 11. (a) Schematic illustration of the feasible morphology-tailoring process of the bare LTO and LTO@C microspheres via hydrothermal treatment. SEM images of the TiOx·H2O precursor (bd), the bare LTO (eg), and the LTO@C (hj) samples with different magnifications. Inset in (b) shows a photograph of an acanthosphere. Reproduced with permission from [359]. Copyright 2016 Elsevier.
Micromachines 15 00310 g011
Figure 12. Hydrothermal synthesis of hollow structured LTO with the shell consisting of nanosheets prepared from AHTO. In a typical process of fabricating an LTO hollow structure, 4 mmol spherical AHTO colloids were dispersed in 20 mL, 0.2 mol L−1 LiOH solution. After stirring for 10 min, the suspension was transferred into a 30 mL Teflon-lined stainless-steel autoclave and then placed in an oven at 180 °C. TEM images present samples obtained for different hydrothermal reaction times: (a) 0 h, (b) 0.5 h, (c) 1.5 h, and (d) 10 h. (e) TEM image of a nanosheet from a hollow microsphere (inset is the SAED pattern taken from the nanosheet). (f) HRTEM image from the nanosheet. (g) XRD patterns of the products from different hydrothermal reaction stages and after calcination. The AHTO precursor is transformed to pure anatase TiO2 after calcination at 500 °C and the spinel LTO phase emerged after a 3 h reaction. Reproduced with permission from [52]. Copyright 2009 Elsevier.
Figure 12. Hydrothermal synthesis of hollow structured LTO with the shell consisting of nanosheets prepared from AHTO. In a typical process of fabricating an LTO hollow structure, 4 mmol spherical AHTO colloids were dispersed in 20 mL, 0.2 mol L−1 LiOH solution. After stirring for 10 min, the suspension was transferred into a 30 mL Teflon-lined stainless-steel autoclave and then placed in an oven at 180 °C. TEM images present samples obtained for different hydrothermal reaction times: (a) 0 h, (b) 0.5 h, (c) 1.5 h, and (d) 10 h. (e) TEM image of a nanosheet from a hollow microsphere (inset is the SAED pattern taken from the nanosheet). (f) HRTEM image from the nanosheet. (g) XRD patterns of the products from different hydrothermal reaction stages and after calcination. The AHTO precursor is transformed to pure anatase TiO2 after calcination at 500 °C and the spinel LTO phase emerged after a 3 h reaction. Reproduced with permission from [52]. Copyright 2009 Elsevier.
Micromachines 15 00310 g012
Figure 13. SEM images of the LTO-TiO2 samples obtained at different hydrothermal treatment times: (a) 12 h: flower-like nanosheets, (b) 15h: ultrathin nanosheets, (c) 18 h: uniform carambola-like frame, and (d) 24 h: urchin-like morphology. Reproduced with permission from [369]. Copyright 2016 Elsevier.
Figure 13. SEM images of the LTO-TiO2 samples obtained at different hydrothermal treatment times: (a) 12 h: flower-like nanosheets, (b) 15h: ultrathin nanosheets, (c) 18 h: uniform carambola-like frame, and (d) 24 h: urchin-like morphology. Reproduced with permission from [369]. Copyright 2016 Elsevier.
Micromachines 15 00310 g013
Figure 15. TEM images highlighting the growth mechanism of C-LTO nanocrystals from a carbon black template. (a) Carbon template at room temperature composed of 50 nm spheres, (b) precursor solution loaded with the carbon template, (c) crystalline domains (~10 nm in diameter) growing off of the carbon black template, (d). crystalline domains after annealing at 450 °C for 1 h, (e) final product as regular octahedra after annealing to 700 °C. Reproduced with permission from [397]. Copyright 2013 Wiley-VCH.
Figure 15. TEM images highlighting the growth mechanism of C-LTO nanocrystals from a carbon black template. (a) Carbon template at room temperature composed of 50 nm spheres, (b) precursor solution loaded with the carbon template, (c) crystalline domains (~10 nm in diameter) growing off of the carbon black template, (d). crystalline domains after annealing at 450 °C for 1 h, (e) final product as regular octahedra after annealing to 700 °C. Reproduced with permission from [397]. Copyright 2013 Wiley-VCH.
Micromachines 15 00310 g015
Figure 16. (A) Chart of the LTO spray-drying process. (B) Spray drying apparatus. The process consists of a sequence of four phases: (1) Atomization of the liquid to be dried (slurry), (2) Contact between hot gas and nebulized liquid, (3) Evaporation of solvent, and (4) Separation of the dried product from the drying medium. Reproduced from [403]. Copyright 2012 Elsevier.
Figure 16. (A) Chart of the LTO spray-drying process. (B) Spray drying apparatus. The process consists of a sequence of four phases: (1) Atomization of the liquid to be dried (slurry), (2) Contact between hot gas and nebulized liquid, (3) Evaporation of solvent, and (4) Separation of the dried product from the drying medium. Reproduced from [403]. Copyright 2012 Elsevier.
Micromachines 15 00310 g016
Figure 18. (A) Electrospinning synthesis route of pristine LTO and LTO/C fibers. (B) SEM images of (a) as-spun fibrous precursor, (b) LTO fibers, (c) 0.85 wt.%C/LTO, (d) 0.95 wt.% C/LTO, (e) 6.02 wt.% C/LTO, and (f) 9.90 wt.% C/LTO fibers. The insets show the corresponding high-magnification images. Reproduced with permission from [140]. Copyright 2014 Wiley.
Figure 18. (A) Electrospinning synthesis route of pristine LTO and LTO/C fibers. (B) SEM images of (a) as-spun fibrous precursor, (b) LTO fibers, (c) 0.85 wt.%C/LTO, (d) 0.95 wt.% C/LTO, (e) 6.02 wt.% C/LTO, and (f) 9.90 wt.% C/LTO fibers. The insets show the corresponding high-magnification images. Reproduced with permission from [140]. Copyright 2014 Wiley.
Micromachines 15 00310 g018
Figure 19. Electrospun LTO nanowires. (a) XRD pattern, (b) SEM image. (c) Cyclic voltammograms at various scanning rates. (d) Relationship between peak currents and square root of the scanning rate; Li-ion diffusion coefficient values are 6.33 × 10−10 and 4.25 × 10−10 cm2 s−1 for anodic and cathodic reactions, respectively. (e) Rate capability. (f) Cycling performance of LTO nanowires in Li-ion half cells. Reproduced with permission from [475]. Copyright 2016 Elsevier.
Figure 19. Electrospun LTO nanowires. (a) XRD pattern, (b) SEM image. (c) Cyclic voltammograms at various scanning rates. (d) Relationship between peak currents and square root of the scanning rate; Li-ion diffusion coefficient values are 6.33 × 10−10 and 4.25 × 10−10 cm2 s−1 for anodic and cathodic reactions, respectively. (e) Rate capability. (f) Cycling performance of LTO nanowires in Li-ion half cells. Reproduced with permission from [475]. Copyright 2016 Elsevier.
Micromachines 15 00310 g019
Table 1. Synthesis conditions of various types of nanosized LTO materials classified according to their dimensionality from 0D to 3D architectures.
Table 1. Synthesis conditions of various types of nanosized LTO materials classified according to their dimensionality from 0D to 3D architectures.
TypeParticleMethod (a)Reagents (b)Particule Size
(nm)
Ref.
0DnanoparticleCLA-TiO2+LiOH30 [119]
nanosphereHLTiN+LiOH+PVPdia. ~120[39]
1DnanorodHTA-TiO2+LiOHdia. ~200[122]
nanowireHT+IETiO2+NaOH+LiOHdia. 50–100[129]
nanofiberESLAA+TIP+PVPdia. ~500[135]
nanotubeHRTNBT+CTACl+LiOHouter 11/inner 6[54]
nanoflakeHTTBuT+EG+LiOH200–400[142]
nanobeltESTIP+LiNO3+DMF+PVPdia. 100 [144]
2DnanoplateHTTBuT+LiOH+BA300 × 8[152]
nanosheetEF+CLH0.68Ti1.83O4+LiOH400 × 2.5[168]
wave-likeHLLi+DMEA+TBuT400 × 10[152]
3DnanoflowersHTLiOH+H2O2+TNBT-[153]
nanoporousHTA-TiO2+LiOHpore dia. 2.6[55]
mesoporousSGTIP+Li+AC+PBI-b-PEOpore = 20[67]
hierarchicalUS+HTTBuT+LiOH+CTAB-[158]
nanotube arraysTPTiO2+LiOH-[159]
nanobelt arraysIE+HTTi foil+HCL+LIOH-[160]
nanoflake arraysTP+HTTIP+LiOH10–50[161]
nanosheet arraysHT+TDTi foil+LiOH~14[169]
nanowire arraysHT+IETi foil+LiOH90[163]
nanowall arraysHTTi foil+NaOH+LiOH10–20[164]
(a) Synthesis methods: CL = chemical lithiation; HL = hydrolysis; HT = hydrothermal, ES = electrospinning; TP = templating; IE = ion exchange; EF = exfoliation; CL = chemical lithiation. (b) Reagent abbreviations: PVP = poly(vinyl pyrrolidone); LAA = lithium acetylacetonate; TIP = titanium isopropoxide; TNBT = titanium n-butoxide; CTACl = cetyltrimethyl ammonium chloride; TBuT = tetrabutyl titanate; EG = ethylene glycol; DMEA = N,N-dimethylethanolamine; BA = benzyl alcohol-NH3·H2O; PVDF = polyvinylidene fluoride; DMF = N,N-dimethylformamide; CTAB = hexadecyl trimethylammonium bromide; A-TiO2 = anatase titanium dioxide.
Table 2. Frequencies and assignments of Raman active modes of spinel Li4Ti5O12.
Table 2. Frequencies and assignments of Raman active modes of spinel Li4Ti5O12.
Raman Shift (cm−1)Assignment
ObservedFrom Ref. [174]from Ref. [175]From Ref. [176]From Ref. [177]
234225232233232νb(O−Li−O)F2g1
265272271-271νb(O−Ti−O)F2g2
347340347348355νs(Li16dO6)F2g3
427420430419428νs(Li8aO4)Eg
517514520-6overtone-
674675671670672νs(TiO6)A1g
754-749748730νs(TiO6)A1g shoulder
Table 3. Starting materials employed for the synthesis of Li4Ti5O12.
Table 3. Starting materials employed for the synthesis of Li4Ti5O12.
SynthesisTi Source (a)Li SourceSolution/Additive (b)Particle Size (nm)Ref.
Solid-state reactionTiO2Li2CO3acetone>1000[185]
Solid-state reactionTiO2Li2CO3urea400[179]
HydrolysisTiNLiOH∙H2OH2O2+NH3120[39]
SolvothermalTBTLiOH∙H2Owater-ethanol (2:3)800–1300[186]
Spray pyrolysisTIPLi(CH3COO)2-propanol (IPA)120[115]
Ball millingA-TiO2Li2CO3ammonium salt230[93]
Spray dryingR-TiO2Li2CO3polyvinyl butyral1000[111]
HydrothermalTIPLiOH∙H2Oethanol-water200–400[110]
Wet chemistryTiOCl2+NH4OHLi(CH3COO)∙2H2Owater+oxalic acid5000[167]
CombustionTiO2LiOH∙H2Omethyl alcohol30–40[104]
CombustionTi(OH)4LiNO3HNO340–80[103]
CombustionR-TiO2Li2CO3cellulose/HNO3-[102]
CombustionA-TiO2LiNO3cellulose/glycine200–800[101]
Ball millingTIPLiNO3lactic acid/NH4NO3200–300[100]
Solid-state reactionTiCl4LiCl70 wt.% oxalic acid41 (c)[187]
Sol–gelTIPLi ethoxidePBI-b-PEO20–30[67]
Soft-templatingTBTLi(CH3COO)∙2H2O2-methoxyethanol18 (d)[188]
MicrowaveTIPdissolved Li metalbenzyl alcohol150 (d)[189]
PrecipitationTBTLiOHDI water10 (e)[190]
Sol–gelTiCl4LiOH∙H2Oethanol+TEA360[191]
HydrothermalTi foilLiOH∙H2OH2O2+DI water~12[192]
Solid-phaseH2TiO3Li2CO3-800[194]
(a) TBT = tetrabutyl titanate (Ti(OC4H9)4); TIP = titanium(IV)isopropoxide (Ti(OC3H7)4); PBI-b-PEO = poly(isobutylene)-b-polyethylene oxide; R-TiO2 = rutile titanium dioxide; A-TiO2 = anatase titanium dioxide. (b) PBI-b-PEO = poly(isobutylene-b-polyethylene oxide); TEA = triethanolamine. (c) Crystallite size according to the Scherrer equation. (d) Average diameter of open pores. (e) Length of nanorods as primary particles.
Table 4. Preparation of LTO anode materials via solid-state reaction.
Table 4. Preparation of LTO anode materials via solid-state reaction.
Ti SourceLi Source (a)Synthesis ProcedureParticle Size
(µm)
Ref.
anatase TiO2Li2CO3mixed in acetone; 900 °C/10 h in air~1.5[43]
anatase TiO2Li2CO3ball milled; 800 °C/5 h in air, N2, H2/Ar0.3–0.4[205]
anatase TiO2LiOAcintermediate phase Li2TiO30.1[231]
TiO(OH)2Li2TiO3Li:Ti = 0.85; 750–800 °C/6 h in air1–5[232]
TiO2+6% NaClLiOHLi:Ti = 0.8; 800 °C/12 h in O21.8–2.3[233]
Ti(C4H9O)4LiOH·H2Odissolved in ethyl alcohol 750 °C/5 h in air-[234]
rutile TiO2Li2CO3milling methanol; 800 °C/12 h in air-[45]
rutile TiO2Li2CO3ground with methanol; 800 °C/12 h in air0.35[235]
Ti foilLiOH30% (w/w) H2O2 aqueous solution; 130 °C/5 h1/0.012[192]
anatase TiO2Li2CO3ground with ethanol/3 h; 900 °C/5 h in air0.4–0.5[47]
anatase TiO2Li2CO3ball mill 400 rpm/4 h; 750 °C/12 h≈0.7[48]
TiO2Li2CO3ball mil/48 h with hexamethylene; 900 °C/24 h-[236]
TiO2Li2CO3phenolic resin+ethyl alcool slurry solution0.2–0.5[237]
H2TiO3Li2CO3direct mixing + sintering at 700 °C/6 h in Ar0.8[184]
(a) LiOAc = lithium acetate dihydrate (CH3COOLi·2H2O).
Table 5. Comparison of the electrochemical performance of nanostructured LTO materials synthesized by solid-state reaction with different morphologies. Cycle number is given in parentheses.
Table 5. Comparison of the electrochemical performance of nanostructured LTO materials synthesized by solid-state reaction with different morphologies. Cycle number is given in parentheses.
Type (a)Reagents (b)Sintering
(°C:h)
Particle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
NPsA-TiO2+Li2CO3800:240[email protected] A g−1 (500)[119]
NPsA-TiO2+LiOH+PAA800:8~300133@10C (50)[219]
MPsA-TiO2+LiOH800:12~2000114@10C (1065)[233]
AGsLiCl+TiCl4+OA800:10-70@10C (200)[187]
SpGsTiO2+Li2CO3800:5~30080@40C (5)[205]
Sub-µR-TiO2+Li2CO3800:3670035@30C (5)[235]
Sub-µTiO2+Li2CO3850:24~60093@5C (200)[224]
SpGsA-TiO2+Li2CO3750:12~70060@10C (500)[48]
NPsA-TiO2+Li2CO3800:3162139@10C (3)[227]
NPsTiCl4+H2C2O4+LiCl800:1020070@10C (100)[222]
NPsTiO2+Li2CO3800:20200–500170@1C (50)[248]
AGsA-TiO2+LiOAc800:3-130@10C (10)[249]
NPsA-TiO2+Li2CO3850:24~500212@1C (150)[250]
(a) NPs = nanoparticles; MPs = microparticles; AGs = aggregates; SpGs = spherical grains; Sub-µ = submicronic gains. (b) PAA = polyacrylate acid; OA = oxalic acid; LiOAc = lithium acetate dihydrate; A-TiO2 = anatase titanium dioxide; R-TiO2 rutile titanium dioxide.
Table 6. Examples of starting materials used in wet-chemical methods for the synthesis of LTO anodes (a).
Table 6. Examples of starting materials used in wet-chemical methods for the synthesis of LTO anodes (a).
Ti SourceLi SourceComplexing AgentSolventParticle Size
(nm)
Ref.
TBTLiOAc Triethanolamineethanol223–264[64]
TIPLiOEtPEGethanol4–5[285]
TBTLiOAcacetic acidisopropyl alcohol100[62]
TIPLiOHn-octanol+HOPacetonitrile450[89]
TETLi2CO3citric acidTiH4O4+HNO3300–650[286]
TBTLiOAc lauric acidalcohol65[287]
TBTLi2CO3oxalic acidalcohol200[63]
TBTLiOAcglacial acetic acidethanol+water400–800[267]
TBTLiOAcpropionic acidalcohol150 × 60[283]
TIPLiOAcP123 copolymerethanol90–110[288]
TBTLiOAcPVPisopropyl alcohol20–75[279]
(a) TBT = tetrabutyl titanate; LiOEt = lithium ethoxide; PEG = polyethylene glycol; TIP = titanium isopropoxide; HOP = hydroxypropyl cellulose; TET = tetraethoxytitanium; P123 = non-ionic surfactant (PEO)20(PPO)70(PEO)20 triblock copolymer; PVP = polyvinylpyrrolidone; LiOAc = lithium acetate (CH3COOLi·2H2O).
Table 7. Comparison of the electrochemical performance of nanostructured LTO materials synthesized by sol–gel route with different morphologies. Cycle number is given in parentheses.
Table 7. Comparison of the electrochemical performance of nanostructured LTO materials synthesized by sol–gel route with different morphologies. Cycle number is given in parentheses.
Type (a)Reagents (b)Sintering
(°C:h)
Particle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
NCsTBT+ LiOAc800:1065163@1C (50)[287]
NCsTBT+LiOAc700:12136157@10C (1)[303]
TFTBT+LiOAc65018 (pores)148@64C (3000)[188]
NCsTIP+LiOAc700:5100108@40C (250)[288]
NPsTIP+LiOH800:10450[email protected] (80)[89]
NSsTBT+LiOAc800:1150 × 6056@20C (80)[283]
NPsTBT+Li2CO3750:5200–500160@1C (1000)[263]
NPsTBT+LiOAc800:1230060@10C (50)[178]
NPsTIP+LiOAc800:12200[email protected] (30)[291]
NPsTBT+Li2CO3750:5~100145@5C (400)[301]
NPsTBT+Li2CO3800:20200[email protected] mA cm−2 (35)[63]
NPsTIP+LiOAc800:250–200[email protected] (200)[304]
NPsTBT+LiOAc+HCl900:2240302@1C (3)[305]
NPsTIP+LiOH+CA800:2-[email protected] (3)[306]
NPsTIP+LiOAc800:125–14214@1C (150)[307]
NCsTBT+LiOH+EtOH600:250–20090@40C (100)[308]
(a) NCs = nanocrystals; TF = thin film; NPs = nanoparticles; NSs = nanostructures; NCs = nanocables. (b) TBT = tetrabutyl titanate; LiOAc = lithium acetate dihydrate. TIP = titanate isopropoxide.
Table 8. Preparation of the LTO spinel phases by hydrolysis route.
Table 8. Preparation of the LTO spinel phases by hydrolysis route.
Ti SourceLi SourceSolutionProcessRef.
TBTLiOHDI water/ethanolstirring for 24 h[106]
TBT+ammoniaLiOHglycerol/waterformation of clear solution[314]
TBTLiOAcDI waterdrying at 90 °C for 5 h[313]
TBTLiOHDI waterslurry dried at 100 °C + grinding[315]
TBTLi metalDMEAH2 byproduct is used as fuel[152]
TiNLiOHDI water/H2O2stirring with ammonia[39]
Table 9. Comparison of the electrochemical performance of nanostructured LTO materials synthesized by mechano-chemical method with different morphologies. Cycle number is given in parentheses.
Table 9. Comparison of the electrochemical performance of nanostructured LTO materials synthesized by mechano-chemical method with different morphologies. Cycle number is given in parentheses.
Type (a)Reagents (b)Sintering
(°C:h)
Particle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
NPsA-TiO2+LiOAc+AG800:1050–400153@1C (1000)[316]
Nc1A-TiO2+Li2CO3+EtOH750:12<1 µm108@10C (500)[48]
Nc1A-TiO2+Li2CO3+EtOH750:12100–200260@88 A g−1 (160)[241]
NPsTiO2+Li2CO3+EA800:20200–500171@1C (50)[325]
C-NcsTiO2+Li2CO3+EA800:1050–400142@1C (1000)[318]
NPsR-TiO2+Li2CO3800:3~400150@4C (3)[328]
HEBMA-TiO2+LiOH+GL800:12150140@1C (1)[92]
HEBMA-TiO2+Li2CO3+PAA800:3146120@20C (3)[93]
NPsTOS+LiOH500:2.5~1 µm[email protected] A g−1 (50)[46]
NPsTBT+Li2CO3+NH3800:7200–500[email protected] (300)[312]
NPsTiCl4+LiOAc+CA800:10500141@30 mA g−1 (30)[326]
NPsA-TiO2+Li2CO3800:8620135@10C (10)[329]
MPsa-TiO2+Li2CO3+Et800:121 µm[email protected] (100)[330]
(a) NPs = nanoparticles; Nc1 = LTO-Li2TiO3 nanocomposite; Nc2 = LTO-ZnO nanocomposite; C-Ncs N-doped carbon LTO nanocomposites; HEBM = high-energy ball-milled particles; MPS = micron-sized particles. (b) A-TiO2 = anatase TiO2; R-TiO2 = rutile TiO2; a-TiO2 = amorphous TiO2; AG = acetyl glucosamine; EtOH = ethanol; EA = ethyl alcohol; GL = glycol; PAA = polycarboxylic acid ammonium salt; TOS = titanium(IV) oxide sulfate sulfuric acid hydrate; LiOAc = lithium acetate (CH3COOLi·2H2O).
Table 10. Electrochemical performances of nanostructured LTO materials synthesized by combustion method. Cycle number is given in parentheses.
Table 10. Electrochemical performances of nanostructured LTO materials synthesized by combustion method. Cycle number is given in parentheses.
Reagents (a)Chelate (b)Sintering
(°C:h)
Particle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
TiO2+LiOAcPVP800:8200–500142@5C (20)[280]
TiO2+LiNO3glycine750:240–80175@1C (1)[217]
TIP+LiNO3lactic acid300200–300132@1C (100)[100]
TBT+LiNO3glycine850:1750–100187@1C (50)[99]
R-TiO2+Li2CO3cotton fiber800:5-130@2C (50)[102]
TBT+LiNO3glycine800:550–350111@10C (150)[341]
TiNt+LiNO3glycine800:½50–100159@1C (200)[342]
TiNt+LiNO3glycine800<150[email protected] (100)[97]
(a) TIP = titanium(IV) isopropoxide (Ti(OC3H7)4); TBT = tetrabutyl titanate (Ti(OC4H9)4); R-TiO2 = rutile TiO2; TiNt = titanyl nitrate (TiO(NO3)2); LiOAc = lithium acetate (CH3COOLi·2H2O). (b) PVP = polyvinylpyrrolidone.
Table 11. Electrochemical performances of nanostructured LTO materials synthesized by hydrothermal method. Cycle number is given in parentheses.
Table 11. Electrochemical performances of nanostructured LTO materials synthesized by hydrothermal method. Cycle number is given in parentheses.
Reagents (a)HT Reaction
(°C:h)
Sintering
(°C:h)
Particle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
TIP+LiOH+EG170:36500:2200152@8C (100)[146]
TIP+LiOH+H2O2150:¼500:2~100[email protected] (30)[351]
TBT+LiOH180:12600:3200–300[email protected] (50)[50]
TBT+LiOH180:12550:6300–600150@20C (5)[345]
TBT+LiOH120500:66–11 (b)[email protected] (25)[54]
TBT/LiOH180:24500:10nanosheets115@5C (50)[354]
Ti(SO4)2+LiOH100:20800:2500[email protected] (70)[53]
TBT+LiOH180:3666:6nanoflakes[email protected] (100)[355]
H2TiO3+LiOH170:24--[email protected] (30)[376]
TBT+LiOEt150:10400–500<100140@2C (1)[195]
TBT+LiOH180:36700:6nanosheets144@ 2C (600)[149]
Na2Ti3O7+LiOH180:24450:1500 × 400.55@3 mA cm−2 (1000) (c)[145]
H2Ti3O7+LiOH150:12500:450 × 1095@5C (300)[129]
TBT+LiOAc170:721050:130–4080@60C (5)[362]
TBT+LiOAc+PVP160:24800:1020–75125@20C (50)[279]
TiO2(P25)+LiOH120:24400:4nanosheets126@20C (3000)[353]
(a) TIP = titanium(IV) isopropoxide (Ti(OC3H7)4); TBT = tetrabutyl titanate (Ti(OC4H9)4); LiOEt = lithium ethoxide (CH3CH2OLi); LiOAc = lithium acetate (CH3COOLi·2H2O); PVP = polyvinylpyrrolidone; EG = ethylene glycol. (b) LTO nanotube diameter. (c) Specific discharge capacity expressed in mAh cm−2.
Table 12. Electrochemical performances of nanostructured LTO materials synthesized by solvothermal method. Cycle number is given in parentheses.
Table 12. Electrochemical performances of nanostructured LTO materials synthesized by solvothermal method. Cycle number is given in parentheses.
Reagents (a)HT Reaction
(°C:h)
Sintering
(°C:h)
Particle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
TBT+LTB+P123170:20400:1035–55140@100C (1000)[387]
TIP+LiOH+EtOH180:10600:320125@1C (100)[383]
TBT+LIOH+EG180:16600:8100–200160@10C (1000)[200]
TBT+LiOAc+EtOH-550:20.5–1 µm92@20C (100)[60]
TIP+Li+BA250:48750:11–2 µm148@1C (200)[203]
TBT+LiOH+EtOH200:24600:3200[email protected] (100)[384]
TiO2+LiOH+EtOH180:16700:4~660106@10C (100)[180]
TBL+LiOH+EtOH200:12500:313105@10C (10)[386]
(a) TBT = tetrabutyl titanate (Ti(OC4H9)4); LTB = lithium tert-butoxide (LiOC(CH3)3); TIP = titanium(IV) isopropoxide (Ti(OC3H7)4); EtOH = ethanol; EG = ethylene glycol (C2H6O2); BA = benzyl alcohol (C7H8O); TBL = titanium(IV) bis(ammonium lactato)dihydroxide ([CH3CH(O-)CO2NH4]2Ti(OH)2).
Table 13. Electrochemical performances of nanostructured LTO materials synthesized by templating method. Cycle number is given in parentheses.
Table 13. Electrochemical performances of nanostructured LTO materials synthesized by templating method. Cycle number is given in parentheses.
Reagents (a)TemplateParticle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
TIP+Li+EtOH+Tolcarbon black~70160@1C (100)[397]
TIP+Li+EtOH+OAPI-b-PEOmesoporous115@10C (500)[155]
TiCl4+LiOAcPMMAmicroporous[email protected] mA cm−2[167]
TiCl4+LiOHZnO nanorodsnanotube arrays110@40C (200)[399]
R-TiO2+Li2CO3PFR~80060@20C (5)[154]
TiO2+LiOHZnO nanorodsnanotubes (b)144@10C (500)[159]
TBT+LiOAcNPGFfibers146@1C (100)[134]
TOT+LiOAc+EtOHwhatman paper250–30088@5C (1100)[87]
(a) TIP = titanium(IV) isopropoxide (Ti(OC3H7)4); TBT = tetrabutyl titanate (Ti(OC4H9)4); EtOH = ethanol; Tol = toluene; OA = oxalic acid; PI-b-PEO = polyisoprene-block-poly(ethylene oxide; PMMA = poly(methyl methacrylate); PFR = phenol–formaldehyde resin; NPGF = nitrogen-doped mesoporous graphene fibers; TOT = tetraethyl orthotitanate (Ti(OC2H5)4; R-TiO2 = rutile titanium dioxide. (b) nanotubes with diameter of 200–300 nm and shell thickness 20–30 nm.
Table 14. Electrochemical performances of nanostructured LTO materials synthesized by spray-drying method. Cycle number is given in parentheses.
Table 14. Electrochemical performances of nanostructured LTO materials synthesized by spray-drying method. Cycle number is given in parentheses.
Reagents (a)Sintering
(°C:h)
Particle Size
(µm)
Specific Capacity
(mAh g−1)
Ref.
TIP+LiOH+H2O2675:161–575@20C (100)[432]
TiO2 (P25)+LiOH750:10~5140@1C (200)[437]
TiO2+LiOH+PEG850:20.6120@4C (10)[430]
TBT+LiOH650:161–1.5100@20C (100)[403]
TIP+Li2CO3+OA800:120.2–0.4120@2C (200)[407]
A-TiO2+LiOH850:120.3–0.540@20C (2)[410]
A-TiO2+Li2CO3+APC850:50.2110@5C (100)[404]
A-TiO2+Li2CO3800:1210–20145@1C (250)[405]
n-TiO2+Li2CO3850:121.18[email protected] (1)[406]
TIP+H2O2+LiOH+NH3700:161–2135@5C (100)[408]
A-TiO2+Li2CO3850:810173@1C (400)[416]
TBT+LiOH+AA750:40.2192@2C (20)[424]
TIP+LiNO3+IsoP800:23.5–5.550@10C (10)[440]
(a) TIP = titanium(IV) isopropoxide (Ti(OC3H7)4); PEG = polyethylene glycol;TBT = tetrabutyl titanate (Ti(OC4H9)4); IsoP = isopropanol; APC = ammonium polycarbonate; n-TiO2 = nano titatium dioxide (20 nm); AA = acetic acid.
Table 15. Electrochemical performances of nanostructured LTO materials synthesized by spray-pyrolysis technique. Cycle number is given in parentheses.
Table 15. Electrochemical performances of nanostructured LTO materials synthesized by spray-pyrolysis technique. Cycle number is given in parentheses.
Reagents (a)Sintering
(°C:h)
Particle Size
(nm)
Specific Capacity
(mAh g−1)
Ref.
TBT+LiOAc+EtOH800:1800146@10C (500)[450]
LNT-EtOH+TIP-EHA-~10146@1C (450)[77]
Li-acac+TIP+EHA+xylene-~14100@25C (10)[445]
LTN+TIP in water800:12700[email protected] (70)[72]
Li-acac+TIP+EHAnone~1087@1C (200)[77]
LiOH+TIP+LA800~1000162@1C (100)[443]
THE+LiOAc+MeOHnone20140@ 1C (75)[451]
TBT+LNT+EGnone600[email protected] (50)[452]
TIP+LNT800:3700175@1C (100)[453]
TIP+LNT700:3300[email protected] (50)[449]
(a) TBT = tetrabutyl titanate (Ti(OC4H9)4); TIP = titanium(IV) isopropoxide (Ti(OC3H7)4); EtOH = ethanol; LNT = LiNO3; EHA = ethylhexanoic acid; Li-acac = lithium acetylacetonate; LA = lactic acid; THE = titanium (IV) 2-ethylhexanoate; MeOH = methanol; EG = ethylene glycol.
Table 16. Detail of the electrospinning synthesis and electrochemical performance of LTO nanofibers. Cycle number is given in parentheses.
Table 16. Detail of the electrospinning synthesis and electrochemical performance of LTO nanofibers. Cycle number is given in parentheses.
Ti Source (a)Li Source (b)Solvent (c)Polymer (d)Performance (e)
(mAh g−1)
Ref.
TIPLiOAcDMF+AAPVAc75@10C (1)[476]
TIPLiOAcEtOH+AAPVP+Pluronic123@5C (100)[474]
TIPLiAAcisoproxyl alcoholF127[email protected] (30)[95]
TiOAcLiOAcEtOHPVP165@1C (100)[133]
nano-LTO-DMFPVP115@2C (1)[464]
TBTLiOAcethanol+AAc+AAPAN139@5C (500)[474]
TIPLiAAcisoproxyl alcoholPVP120@10C (100)[140]
TIPLiOAcEtOH+AAPVP120@1C (200)[471]
TBTLiAAcEtOH+AAPVP60@25C (1)[468]
TIPLiOAcEtOH+AAPVP140@10C (30)[137]
TIPLAAcEtOH+AAPVP132@20C (300[135]
(a) TIP = Titanium(IV) isopropoxide, Ti[OCH(CH3)2]4; TBT = tetrabutyl titanate, Ti(C4H9O)4. (b) LiOAc = lithium acetate dihydrate, Li(CH3CO2)·2H2O; LiAAc = lithium acetylacetonate (LiC5H7O2). (c) DMF = N,N-dimethyl formamide; EtOH = ethanol; AA = acetic acid; AAc = acethylacetone C5H8O2. (d) PVAc = polyvinyl acetate; PVP = polyvinylpyrrolidone; PAN = polyacrylonitrile. (e) 1C = 175 mA g−1.
Table 17. Electrochemical performances of nanostructured LTO materials synthesized by using two-step ion exchange technique. Cycle number is given in parentheses.
Table 17. Electrochemical performances of nanostructured LTO materials synthesized by using two-step ion exchange technique. Cycle number is given in parentheses.
First StepSecond StepParticle (a)Specific Capacity
(mAh g−1)
Ref.
TiO2/NaOH/HNO3H-TiOx/LiOHNTs 7 nm110@1C (30)[130]
TiO2/Na2CO3/HClH2Ti3O7/LiNO31–2 µm190@5C (3)[478]
TiO2/NaOH/HClH2Ti2O5/LiOHNFs118@5C (4)[477]
Ti/NaOH+NaCl/HClH2Ti2O5/LiOHNW arrays125@30C (5000)[375]
-TiO2/NaOH/LiOH10–50 nm60@50C (1)[141]
TiO2/NaOH/HClH2Ti3O7/LiOHNWs[email protected] (50)[132]
Ti/NaOH/HClH-TiOx/LiOHarrays 7 µm0.2@1 mA cm−2 (1000) (b)[165]
-TiO2/NaOH/LiOH20–30 nm148@10C (200)[479]
TiO2/NaOH/HClH2Ti2O5/LiOHNWs 80 nm125@10C (100)[480]
(a) NT = nanotube; NF = nanofiber; NW = nanowire; (b) Areal specific capacity in mAh cm−2.
Table 18. Advantages and drawbacks of LTO synthesis methods.
Table 18. Advantages and drawbacks of LTO synthesis methods.
Synthesis MethodAdvantagesDisadvantages
Solid state reactionSimple method; use of starting materials at low cost; large-scale productionNeed for high-temperature treatment; formation of unwanted byproducts; difficult to adjust particle size, particle size distribution, and lack of stoichiometric control
Hydrothermal/solvothermalProduction of nanomaterials; well controlled through liquid-phase reactionExpensive method; time-consuming; not feasible for mass production; organic residues appear with organic solvents
Sol–gelLow-cost; offer products with a homogeneous distribution of uniform, submicron-size particles with good stoichiometric controlUse of chelating agent for the control of nanostructure and morphology; difficult to use for mass production
Solution-combustion Needs additional post-annealing treatment
MicrowaveFast synthesis route; low-temperature process; rapid volumetric heating, high reaction rate; small particle sizePurity and the crystallinity of the products are dependent upon the irradiation power and time; difficult to use for mass production
Spray pyrolysisDirect synthesis of powder from precursor solutions; products with high purity and homogeneity; industrially scalableDifficult to control the morphology and chemical composition
ElectrospinningSimple, versatile, and low-cost technology to produce well-defined fibers at the nanoscaleTwo-step procedure; expensive
TemplatingFormation of mesoporous structures; monodispersed productsRemoval of the templates after synthesis may damage the desired configurations; not suitable for large-scale production
Molten saltLow temperature and short time; suitable for small particle sizeNot suitable for large-scale production
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

Julien, C.M.; Mauger, A. Fabrication of Li4Ti5O12 (LTO) as Anode Material for Li-Ion Batteries. Micromachines 2024, 15, 310. https://doi.org/10.3390/mi15030310

AMA Style

Julien CM, Mauger A. Fabrication of Li4Ti5O12 (LTO) as Anode Material for Li-Ion Batteries. Micromachines. 2024; 15(3):310. https://doi.org/10.3390/mi15030310

Chicago/Turabian Style

Julien, Christian M., and Alain Mauger. 2024. "Fabrication of Li4Ti5O12 (LTO) as Anode Material for Li-Ion Batteries" Micromachines 15, no. 3: 310. https://doi.org/10.3390/mi15030310

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