Open access peer-reviewed chapter

TiO2 Nanoparticles Prepared by Sol-Gel Method for Anode Application in Lithium-Ion Batteries

Written By

Wafaa Nachit, S. Touhtouh, Z. Ramzi, K. Benkhouja and R. Yazami

Submitted: 11 June 2018 Reviewed: 10 September 2018 Published: 30 April 2019

DOI: 10.5772/intechopen.81406

From the Edited Volume

Lithium-ion Batteries - Thin Film for Energy Materials and Devices

Edited by Mitsunobu Sato, Li Lu and Hiroki Nagai

Chapter metrics overview

1,179 Chapter Downloads

View Full Metrics

Abstract

TiO 2 nanoparticles are prepared via sol-gel method using titanium tetraisopropoxide (TTIP) as a precursor under refluxing and controlled pH. It is found that pure anatase phase is obtained with pH 10. Further characterization studies are carried out on pure nanoparticle anatase phase by XRD, SEM, and transmission electron microscope (TEM). Their electrochemical performances as anode material in lithium-ion batteries are investigated by cyclic voltammetry, galvanostatic cycling, and rate capability measurements. A high discharge capacity of 321 mAh/g (vs. 335 mAh/g theoretical) is achieved at 1C rate. After the first galvanostatic charge/discharge cycle, voltage profiles show plateaus at 1.75 and 1.95 V versus Li at discharge and charge, respectively. High Coulombic efficiency (>99%) is maintained after 300 cycles, which makes anatase TiO 2 nanoparticles prepared by sol-gel method, a very promising material for anode application in lithium rechargeable batteries.

Keywords

  • TiO2 nanoparticles
  • anode
  • lithium-ion batteries
  • electrochemical characteristics

1. Introduction

Nanostructured titania (NST) TiO 2 is a very promising alternative anode material compared with the commercialized lithium titanate L i 4 Ti 15 O 12 and graphite materials for lithium-ion battery application in particular when fast charging is required [1, 2, 3]. Fast charging while keeping long battery operational life is one of the most challenging tasks in lithium-ion batteries nowadays. Both mobile electronics and electro-mobility sectors need fast charging to compensate for relatively low energy density which impacts the operation time and driving range. One serious option to cope with charging time limitations resides in developing new materials able to store large amounts of lithium in a very short time. Nanomaterials, including NST, are among the most promising in this regard.

Good performances of NST result from their outstanding characteristics such as:

  1. A low volumetric change upon lithium-ion intercalation and de-intercalation providing a high crystal structure stability and long cycle life

  2. A fast diffusivity of lithium ions within the Ti O 2 structure providing fast charging and high power discharge

  3. A practical low operating voltage

  4. An enhanced safety

Several methods have been used for the synthesis of NST, including hydrothermal [4, 5, 6, 7, 8, 9, 10], sol-gel [5, 6, 7, 8, 9, 10, 11], hydrolysis [12, 13], atomic layer deposition [14], and anode deposition [15], for the most reported in the literature. According to the synthesis method and to the post-thermal treatment conditions, NST of different morphologies are obtained such as nanorods [4, 17], nanotubes [5, 6, 8, 9, 14, 17], nanofibers/nanowires [11, 12, 13, 14, 15, 16, 17, 18, 19, 20], nanoparticles [6, 7, 10], nanocomposites [12, 21], nanofilms, and nanosheets [15, 22].

Several polymorphs of Ti O 2 are known such as tetragonal rutile and anatase, orthorhombic brookite, and monoclinic Ti O 2 (B) [23]. Anatase and Ti O 2 (B) have open-channel structure and show high cycle capacity compared to rutile Ti O 2 [24]. Anatase has edge-sharing Ti O 6 octahedra so that planar double chains are formed with a tetragonal body-centered space group (I41/amd) [25].

Li-ion diffusion in anatase occurs along the octahedral interstitial sites up to L i 0.5 Ti O 2 composition. Then the tetragonal structure reversibly transforms to the orthorhombic phase [26, 27, 28, 29, 30]. Anatase Ti O 2 anode material reversibly uptakes 0.5 Li per Ti O 2 formula with lithium insertion/extraction taking place at relatively low voltage and showing fast kinetics [29, 30]. However, 0.5 Li per Ti O 2 corresponds to half of theoretical capacity of 335 mAh/g [31, 32, 33]. Enhancing practical capacity in NST should make this material more attractive for anode application in lithium-ion batteries [34].

When wet chemistry route is used to prepare NST, the medium pH is of a significant importance in the crystal structure and morphology of the final product [32]. Syntheses of TiO2 anatase were mostly performed in acidic media usually at pH greater than 4.5 [33]. To the authors’ knowledge, there has been no report on sol-gel preparation of NST in alkaline pH.

In this chapter, we show that unlike in acidic pH, which lead to mixed phases, pure anatase NST is obtained via sol-gel at pH 10. NST prepared at different pH are analyzed by XRD, transmission electron microscope (TEM), and SEM. The electrochemical properties of pure anatase phase NST are investigated by cyclic voltammetry and galvanostatic cycling, including rate capability tests in view of its application as anode material in lithium-ion batteries.

Advertisement

2. Experimental

2.1 Sample preparation

NST are prepared via sol-gel method using titanium tetraisopropoxide (TTIP), distilled water, and 2-propanol. Hydrochloric acid and sodium hydroxide are mixed in 24 mL of 2-propanol to fix the pH to 1, 4, 10, and 13 and to control the hydrolysis process of the solution. 24 mL of TTIP is then slowly added to the mixture, followed by refluxing at 70°C to obtain the sol. 10 mL of distilled water is added to the solution and left for 1 h under stir heating.

The obtained gel is dried at 70°C in air and then calcined at 500°C in air to produce NST powders.

2.2 Characterization

NST are characterized by X-ray diffractometry (XRD) using a PR-XPERT diffractometer (PW3064) with CuKα radiation (λ = 1.5406 Å), operating at 45 kV and 40 mA in 2θ angular range of 10–70°. The morphology and particle size were examined by transmission electron microscope (TEM, JEOL, JEM-2010) and by field-emission scanning electron microscopy (FESEM, JEOL JSM-7600F, 5 kV).

2.3 Electrochemical measurements

The anode is prepared by mixing NST prepared at pH 10 as active material, acetylene carbon black as electronic conductor, and polyvinylidene fluoride (PVDF) as a binder in the weight ratio of 80:10:10 with N-methyl-2-pyrrolidone (NMP) to form a slurry. The slurry was coated on a copper foil by doctor blade and then vacuum dried at 110°C for 6 h. The anode loading is ~0.8 mg cm−2. Lithium half-cells are assembled in a glove box under ultra-dry argon atmosphere (<1 ppm H 2 O and O 2 ). Lithium metal was used as the counter and reference electrode. The electrolyte consisted of 1.0 M LiP F 6 solution in ethylene carbonate (EC) and diethyl carbonate (DEC) at a volumetric ratio of 1:1. The separator is microporous polyethylene. Li/NST half-cells are investigated by cyclic voltammetry at 0.2 mV/s sweeping rate and by galvanostatic cycling at 1C–20 C rates, all between 1.0 V and 3.0 V versus Li.

Advertisement

3. Results and discussion

3.1 Physical characterizations

Figure 1 shows the XRD patterns of TiO 2 nanoparticles prepared at pH 1, 4, 10, and 13 and calcined at 500°C. X-ray patterns of anatase, rutile, and brookite are shown for reference.

Figure 1.

XRD patterns of titania nanoparticles made at different pH. Anatase, rutile, and brookite spectra are also shown for reference.

The sample prepared at pH 1 consists mostly of the rutile phase. All other samples are anatase with brookite impurities at pH 4 and 13. TiO2 nanoparticles synthesized at pH 10 consist of highly crystalline anatase phase (00-004-0477). Table 1 summarizes the XRD data of TiO2 nanoparticles synthesized at pH 10 including diffraction peak 2θ, interlayer spacing “d,” and hkl indexes. Table 2 gives the crystallographic characteristics of TiO 2 anatase phase prepared at pH 10 as derived from the XRD data Rietveld refinement. Average crystallite sizes of nanoparticles synthesized at different pH are reported in Table 3 using Scherrer’s equation. The particle size is in ~20–30 nm range with no obvious trends versus pH.

2θ (º) d (nm) (hkl)
25.31 0.3518 (101)
36.98 0.2430 (103)
37.82 0.2378 (004)
38.61 0.2331 (112)
48.05 0.1893 (200)
53.92 0.1700 (105)
55.09 0.1637 (221)
62.13 0.1494 (213)
62.70 0.1481 (204)
62.80 0.1479 (116)

Table 1.

Indexation of XRD peaks on TiO2 nanoparticles obtained at pH 10.

Phase α β γ Lattice parameters
Anatase
(Tetragonal I41)
α = β = γ = 90° a = b = 3.783 c = 9.5190

Table 2.

Characteristics of anatase crystal structure.

pH Phase D (nm)
1 Anatase-rutile 32.0
4 Anatase-brookite 21.3
10 Anatase 26.2
13 Anatase-brookite 21.3

Table 3.

Crystallite size of TiO2 nanoparticles synthesized at different pH.

Figure 2 shows TEM images of Ti O 2 nanoparticles prepared at pH 10 at different magnifications. Figure 2a and b shows primary particles with relatively narrow size distribution. High-resolution TEM of Figure 2c shows the polycrystalline nature of the TiO 2 nanoparticles with well-resolved layers such as those assigned to the orientation corresponding to an interlayer of about 0.36 nm as indicated by an arrow.

Figure 2.

TEM images at different magnifications of TiO2 nanoparticles prepared at pH 10. (a, b) Image analysis enables size distribution to be achieved. HR-TEM image of (c) shows polycrystalline structure with some preferred layer orientation such as with interlayer spacing of about 3.6 nm.

TEM image analysis allows particle-size distribution to be determined as shown in Figure 3. The particle-size distribution is built on a discrete set of clearly identifiable grains, and, therefore, it is not necessarily representative of the overall description of the sample. This is especially true for grain sizes close to the distribution tails where the probability of finding a very large or a very small grain under the TEM objective is rather low. The average size of about 17 nm is slightly lower than the one estimated from the XRD data.

Figure 3.

Grain size distribution for the TiO2 at pH 10 from the TEM image analysis.

SEM images of NST pH 10 are displayed in Figure 4a and b. High magnification Figure 4a shows primary particles with narrow size distribution aggregate into secondary particles of about 150–600 nm in size, whereas lower magnification Figure 4b shows agglomerated particles remain basically submicron in size.

Figure 4.

SEM images of TiO2 nanoparticles synthesized at pH 10.

3.2 Electrochemical characterizations

Figure 1ashows cyclic voltammograms of the Li/NST of pH 10 half-cell at 0.2 mV/s sweeping rate for 5 cycles. In the first cycle, two reduction peaks appear at 2.15 V and at 1.75 V, respectively, whereas only one re-oxidation peak is present at 2.10 V. In the following 4 cycles, only the reduction peak at 1.75 V and re-oxidation peak at 2.10 V appeared. The first reduction peak at 2.15 V accounts for about 50% of the total discharge capacity of about 320 mAh/g, which is close to theoretical capacity of 335 mAh/g of TiO 2 corresponding to 1 Li/ TiO 2 ratio. The second reduction peak at 1.75 V accounts for the second half of the total capacity and is matched by the re-oxidation peak at 2.1 V. This situation remained almost unchanged during the four following cycles, that is, the capacity of the reduction peak at 1.75 V matches the one of the re-oxidation peaks at 2.1 V of about 140 mAh/g.

The data shown above are confirmed during the two first galvanostatic cycles of the Li/NST cell displayed in Figure 5b. The first discharge profile shows two semi-plateaus at about 2.15 and 1.75 V, respectively, whereas the charge profile shows a main plateau at about 2.0 V. The first discharge semi-plateau accounts for nearly half of the total discharge capacity. The second half discharge capacity is matched by the charge capacity.

Figure 5.

Electrochemical tests performed on Li/TiO2 (pH 10) half-cells. (a) Cyclic voltammograms (CVs) at a scan rate of 0.2 mV/s for the first 5 cycles within a potential range from 1 to 3.0 V (vs. Li/Li+), (b) discharge-charge capacity of the first and second cycles at a specific current of 175 mA/g (1C), and (c) cycle capacity profile at 1C rate for 50 cycles.

Cyclic voltammetry and galvanostatic data suggest a two-step electrode reaction process:

  1. Irreversible formation of L i 0.5 Ti O 2 at ~ 2.15 V average voltage.

  2. Reversible formation of L i 1 Ti O 2 from L i 0.5 Ti O 2 at ~ 1.75 V. During re-oxidation L i 0.5 Ti O 2 is formed from L i 1 Ti O 2 at about 2.1 V.

The formation of lithium-rich L i 1 Ti O 2 is supported by both theoretical and experimental data [30, 31, 32, 33, 34, 35]. In the anatase phase, orthorhombic L i 0.5 Ti O 2 and distorted rock-salt structure Li1TiO2 in which lithium cation occupy the octahedral sites have been stabilized due to kinetic effects. Metastability has been reported in Li-poor Ti O 2 [30]. Accordingly, our data are consistent with reversible lithium intercalation in the stoichiometry range 0.5 < x < 1 in anatase L i x Ti O 2 , which corresponds to a theoretical capacity of 167.5 mAh/g. The cycle capacity profile in Figure 5c shows a plateau at about 120 mAh/g at 350 mA/g rate. A lower capacity may be due to limitation in lithium uptake to L i 0.85 Ti O 2 and L i 0.7 Ti O 2 as reported in [36, 37].

The intercalation mechanism of Li ions into anatase has been studied in a variety of experimental and theoretical works [38]. The lithium-ion intercalation can be described by the following Eq. (1):

Ti O 2 + x Li + + x e L i x Ti O 2 E1

Ti O 2 maximum theoretical capacity is 335 mAh/g [39] which corresponds to x = 1 and to the complete reduction of Ti 4 + to Ti 3 + . This makes Ti O 2 a highly competitive alternative to graphite anodes [40].

Practical capacity of anatase Ti O 2 is about 200 mAh/g associated with 0.6 Li mole reversible intercalation at 1.78 V versus Li+/Li [41]. It has been demonstrated that L i 1 Ti O 2 composition can be obtained from anatase Ti O 2 at high temperature [30] or with particle size below 7 nm [42].

The discharge-charge profiles at current rates ranging from 1.0C to 20C and corresponding rate capability profiles are depicted in Figure 6a and b, respectively. It is worth noticing that after 10 cycles at 20C rate where capacity decreased to about 50 mAh/g, the capacity at 1C rate was recovered during the following 10 cycles to about 137 mAh/g. This indicates a high stability of the anatase Ti O 2 nanoparticles based anode toward high rate cycling.

Figure 6.

Rate capability (1C–20C) tests performed on Li/TiO2 (pH 10) half-cells. (a) Discharge-charge voltage profiles, (b) discharge-charge capacity profile, and (c) cycle capacity and Coulombic efficiency profiles at 2C rate.

Figure 6c shows the cycle capacity profile at 2C rate for 300 cycles together with Coulombic efficiency. The latter stabilizes close to unity after the first 5 cycles, whereas the capacity gradually dropped to 82.4 mAh/g after 300 cycles indicating the high capacity retention and good cycling stability.

Advertisement

4. Conclusion

TiO2 nanoparticles are synthesized by simple sol-gel method at different pH values and found to have particle size of 10–30 nm. XRD measurements show that the sample prepared at pH 10 crystallized to pure anatase phase. The TEM and the SEM images confirm the formation of nanoparticles with diameters of 10–30 nm. The Ti O 2 nanoparticles prepared at pH 10 exhibit initial capacity of 321 mAh/g and a remarkable high Coulombic efficiency close to unity after 300 cycles at 2C rate. However, the first cycle efficiency is relatively low perhaps due to inherent metastability of Li-poor phases at the early stage of lithium intercalation. More work has currently been performed to improve the first cycle efficiency, therefore making anatase nanoparticles based anode more practical in rechargeable lithium batteries with fast charge and high lithium storage capacity. Among possible routes considered to improve the NTO initial cycle efficiency is the incorporation of a few amounts of lithium during sol-gel process so as to achieve a pre-lithiated phase making it easier to add more lithium during the first cathodic process.

References

  1. 1. Wang C, Wu LX, Wang H, Zuo WH, Li YY, Liu JP. Fabrication and shell optimization of synergistic TiO2-MoO3 core-shell nanowire array anode for high energy and power density lithium-ion batteries. Advanced Functional Materials. 2015;15(23):3524-3533
  2. 2. Wei W, Oltean G, Tai CW, Edstrom K, Bjorefors F, Nyholm L. High energy and power density TiO2 nanotube electrodes for 3D Li-ion microbatteries. Journal of Materials Chemistry. 2013;1(28):8160-8169
  3. 3. Ortiz GF, Hanzu I, Lavela P, Knauth P, Tirado JL, Djenizian T. Nanoarchitectured TiO2/SnO: A future negative electrode for high power density Li-Ion microbatteries. Chemistry of Materials. 2010;22(5):1926-1932
  4. 4. Liu YB, Ding TL, Shen DL, Dou J, Wei MD. Hierarchically structural TiO2 nanorods composed of rutile core and anatase shell as a durable anode material for lithium-ion intercalation. Journal of Electroanalytical Chemistry. 2017;804:87-91
  5. 5. Fang J, Liu WW, Yu F, Qin FR, Wang MR, Zhang K, et al. Fe, S co-doped anatase TiO2 nanotubes as anodes with improved electrochemical performance for lithium ion batteries. RSC Advances. 2016;6(74):70133-70,140
  6. 6. Tang YK, Liu L, Zhao HY, Jia DZ, Xie XL, Zhang Y, et al. Anatase/rutile titania anchored carbon nanotube porous nanocomposites as superior anodes for lithium ion batteries. CrystEngComm. 2016;18(24):4489-4494
  7. 7. Sanad MMS, Rashad MM, Powers K. Enhancement of the electrochemical performance of hydrothermally prepared anatase nanoparticles for optimal use as high capacity anode materials in lithium ion batteries (LIBs). Applied Physics A: Materials Science and Processing. 2015;118(2):665-674
  8. 8. Choi MG, Lee YG, Song SW, Kim KM. Anode properties of titanium oxide nanotube and graphite composites for lithium-ion batteries. Electrochimica Acta. 2010;55(20):5975-5983
  9. 9. Xu JW, Jia CH, Cao B, Zhang WF. Electrochemical properties of anatase TiO2 nanotubes as an anode material for lithium-ion batteries. Electrochimica Acta. 2007;52(28):8044-8047
  10. 10. Oh SW, Park SH, Sun YK. Hydrothermal synthesis of nano-sized anatase TiO2 powders for lithium secondary anode materials. Journal of Power Sources. 2006;161(2):1314-1318
  11. 11. Lee S, Eom W, Park H, Han TH. High-temperature stable anatase titanium oxide nanofibers for lithium-ion battery anodes. ACS Applied Materials and Interfaces. 2017;9(30):25332-25,338
  12. 12. Liu JL, Feng HB, Jiang JB, Qian D, Li JH, Peng SJ, et al. Anatase-TiO2/CNTs nanocomposite as a superior high-rate anode material for lithium-ion batteries. Journal of Alloys and Compounds. 2014;603:144-148
  13. 13. Xu W, Wang ZY, Guo ZD, Liu Y, Zhou NZ, Niu B, et al. Nanoporous anatase TiO2/single-wall carbon nanohorns composite as superior anode for lithium ion batteries. Journal of Power Sources. 2013;232:193-198
  14. 14. Panda SK, Yoon Y, Jung HS, Yoon WS, Shin H. Nanoscale size effect of titania (anatase) nanotubes with uniform wall thickness as high performance anode for lithium-ion secondary battery. Journal of Power Sources. 2012;204:162-167
  15. 15. Wu MS, Wang MJ, Jow JJ, Yang WD, Hsieh CY, Tsai HM. Electrochemical fabrication of anatase TiO(2) nanostructure as an anode material for aqueous lithium-ion batteries. Journal of Power Sources. 2008;185(2):1420-1424
  16. 16. Kim J, Cho J. Rate characteristics of anatase TiO2 nanotubes and nanorods for lithium battery anode materials at room temperature. Journal of the Electrochemical Society. 2017;154(6):A542-A546
  17. 17. Zhang MH, Yin KB, Hood ZD, Bi ZH, Bridges CA, Dai S, et al. In situ TEM observation of the electrochemical lithiation of N-doped anatase TiO2 nanotubes as anodes for lithium-ion batteries. Journal of Materials Chemistry A. 2017;5(39):20651-20,657
  18. 18. Luo Y, Li J, Huang JG. Bioinspired hierarchical nanofibrous silver-nanoparticle/anatase-rutile-titania composite as an anode material for lithium-ion batteries. Langmuir. 2016;32(47):12338-12,343
  19. 19. Yang XJ, Teng DH, Liu BX, Yu YH, Yang XP. Nanosized anatase titanium dioxide loaded porous carbon nanofiber webs as anode materials for lithium-ion batteries. Electrochemistry Communications. 2011;139(10):1098-1101
  20. 20. Wu FX, Li XH, Wang ZX, Guo HJ, Wu L, Xiong XH, et al. A novel method to synthesize anatase TiO2 nanowires as an anode material for lithium-ion batteries. Journal of Alloys and Compounds. 2011;509(18):3711-3715
  21. 21. Lai C, Li CGRY, Dou Y, Gao XP. Mesoporous polyaniline or polypyrrole/anatase TiO2 nanocomposite as anode materials for lithium-ion batteries. Electrochimica Acta. 2010;55(15):4567-4572
  22. 22. Cheng XL, Hu M, Huang R, Jiang JS. HF-free synthesis of anatase TiO2 nanosheets with largely exposed and clean {001} facets and their enhanced rate performance as anodes of lithium-ion battery. ACS Applied Materials and Interfaces. 2014;6(2):19176-19,183
  23. 23. Roy P, Srivastava SK. Nanostructured anode materials for lithium ion batteries. Journal of Materials Chemistry A. 2015;3(6):2454-2484
  24. 24. Wang XD, Li ZD, Shi J, Yu YH. One-dimensional titanium dioxide nanomaterials: Nanowires, nanorods, and nanobelts. Chemical Reviews. 2014;114(19):9346-9384
  25. 25. Ortiz GF, Hanzu I, Djenizian T, Lavela P, Tirado JL, Knauth P. Alternative Li-ion battery electrode based on self-organized titania nanotubes. Chemistry of Materials. 2009;21(1):63-67
  26. 26. Sudant G, Baudrin E, Larcher D, Tarascon JM. Electrochemical lithium reactivity with nanotextured anatase-type TiO2. Journal of Materials Chemistry. 2005;15(12):1263-1269
  27. 27. Wagemaker M, Borghols WJH, van Eck ERH, Kentgens APM, Kearley GL, Mulder FM. The influence of size on phase morphology and Li-ion mobility in nanosized lithiated anatase TiO2. Chemistry--A European Journal. 2007;13(7):2023-2028
  28. 28. Wagemaker M, Van Well AA, Kearley GJ, Mulder FM. The life and times of lithium in anatase TiO2. Solid State Ionics. 2004;175(1):191-193
  29. 29. He YB, Liu M, Xu ZL, Zhang B, Li B, Kang F, et al. Li-ion reaction to improve the rate performance of nanoporous anatase TiO2 anodes. Energy Technology. 2013;1(11):668-674
  30. 30. Macklin WJ, Neat RJ. Performance of titanium dioxide-based cathodes in a lithium polymer electrolyte cell. Solid State Ionics. 1992;53:694-700
  31. 31. Zhu GN, Wang YG, Xia YY. Ti-based compounds as anode materials for Li-ion batteries. Energy and Environmental Science. 2012;5(5):6652-6667
  32. 32. Zhou ZY, Tian N, Li JT, Broadwell I, Sun SG. Nanomaterials of high surface energy with exceptional properties in catalysis and energy storage. Chemical Society Reviews. 2011;40(7):4167-4185
  33. 33. Bavykin DV, Friedrich JM, Walsh FC. Protonated titanates and TiO2 nanostructured materials: Synthesis, properties, and applications. Advanced Materials. 2006;18(2):2807-2824
  34. 34. Zakharova GS, Jahne C, Popa A, Taschner C, Gemming T, Leonhardt A, et al. Anatase nanotubes as an electrode material for lithium-ion batteries. Physical Chemistry C. 2012;116(15):8714-8720
  35. 35. Wagemaker M, Kentgens APM, Mulder FM. Equilibrium lithium transport between nanocrystalline phases in intercalated TiO2 anatase. Nature. 2002;418(6896):397-399
  36. 36. Hu YS, Kienle L, Guo YG, Maier J. High lithium electroactivity of nanometer-sized rutile TiO2. Advanced Materials. 2006;18(11):1421-1426
  37. 37. Murphy DW, Greenblatt M, Zahurak SM, Cava RJ, Waszczak JV, Hull GW, et al. Lithium insertion in anatase - a new route to the spinel LiTi2O4. Chemischer Informationsdienst. 1982;19(4-5):441-449
  38. 38. Wagemaker M, Kearley GJ, van Well AA, Mulka H, Mulder FM. Multiple li positions inside oxygen octahedra in lithiated TiO2 anatase. Journal of the American Chemical Society. 2003;125(3):840-848
  39. 39. Jiang C, Zhang J. Nanoengineering titania for high rate lithium storage: A review. Journal of Materials Science and Technology. 2013;29(2):97-122
  40. 40. Fröschl T, Hörmann U, Kubiak P, Kucerova G, Pfanzelt M, Weiss CK, et al. Nanoengineering titania for high rate lithium storage: A review. Journal of Royal Society of Chemistry. 2012;41(15):5313-5360
  41. 41. Kavan L, Kalbac M, Zukalova M, Exnar I, Lorenzen V, Nesper R, et al. Lithium storage in nanostructured TiO2 made by hydrothermal growth. Chemistry of Materials. 2004;16(3):477-485
  42. 42. Lafont U, Carta D, Mountjoy G, Chadwick AV, Kelder EM. In situ structural changes upon electrochemical lithium insertion in nanosized anatalse TiO2. Physical Chemistry C. 2010;114(2):1372-1378

Written By

Wafaa Nachit, S. Touhtouh, Z. Ramzi, K. Benkhouja and R. Yazami

Submitted: 11 June 2018 Reviewed: 10 September 2018 Published: 30 April 2019