Next Article in Journal
Dielectric, Ferroelectric, and Magnetic Properties of Sm-Doped BiFeO3 Ceramics Prepared by a Modified Solid-State-Reaction Method
Next Article in Special Issue
Polymer Electrode Materials for Sodium-ion Batteries
Previous Article in Journal
Thermal and Mechanical Properties of Expanded Graphite/Paraffin Gypsum-Based Composite Material Reinforced by Carbon Fiber
Previous Article in Special Issue
High Capacity and High Efficiency Maple Tree-Biomass-Derived Hard Carbon as an Anode Material for Sodium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of F-Doping on the Electrochemical Performance of Na2Ti3O7 as an Anode for Sodium-Ion Batteries

1
College of Chemistry and Chemical Engineering, Henan Polytechnic University, Jiaozuo 54000, China
2
School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
3
CAS Center for Excellence in Nanoscience, Beijing Institute of Nanoenergy and Nanosystems, Chinese Academy of Sciences, Beijing 100083, China
4
School of Nanoscience and Technology, University of Chinese Academy of Sciences, Beijing 100049, China
5
Center on Nanoenergy Research, School of Physical Science and Technology, Guangxi University, Nanning 530004, China
6
Medical College, Henan Polytechnic University, Jiaozuo 454000, China
*
Authors to whom correspondence should be addressed.
The authors equally contributed to this work.
Materials 2018, 11(11), 2206; https://doi.org/10.3390/ma11112206
Submission received: 10 October 2018 / Revised: 2 November 2018 / Accepted: 5 November 2018 / Published: 7 November 2018
(This article belongs to the Special Issue Electrode Materials for High Performance Sodium-ion Batteries)

Abstract

:
The effects of fluorine (F) doping on the phase, crystal structure, and electrochemical performance of Na2Ti3O7 are studied by X-ray diffraction (XRD), scanning electron microscopy (SEM), and electrochemical measurements. F-doping does not change the crystal structure of NTO, although it has an effect on the morphology of the resultant product. As an anode material for sodium-ion batteries, the specific capacity of Na2Ti3O7 exhibits a 30% increase with F-doping owing to the improved sodium ion diffusion coefficient. F-doped Na2Ti3O7 also displays an enhanced rate capability and favourable cycling performance for more than 800 cycles.

1. Introduction

At present, lithium-ion batteries (LIBs) are indispensable for power electronics [1,2]. In the coming decades, LIBs will be used on an increasing scale in the field of utility grid and a future “energy internet” [2]. However, the limited reserves and cost issues of lithium present great challenges for grid application of LIBs. Sodium-ion batteries (SIBs) have recently attracted interest because of their advantages such as low cost, environmentally friendliness, and the availability of rich sodium resources [3,4]. Various SIB electrode materials have been developed recently, including sodium metal phosphate [5], Na0.44MnO2 [6], Na0.67Ni0.23Mg0.1Mn0.67O2 [7,8], Na3[Ti2P2O10F] [9,10], NaMnFe2(PO4)3 [11], V2O5 [12], NiCo2O4 [13], Na3V2(PO4)3 [14], Sb2O4 [15], Na2V6O16 [16], NaFeF3 [17], and Na2Ti3O7 [18,19], as well as some hard carbon materials [20,21]. Among these materials, layered Na2Ti3O7 is one of the most ideal choices as negative active materials for SIBs, with favorable electrochemical performance at a relatively low intercalation voltage of approximately 0.3 V as compared to Na+/Na and a low activation energy (~0.186 eV) for Na ion insertion [22]. Although the capacity of NTO reaches 200 mAh g−1 [23,24], Na2Ti3O7 does not show a better rate capability or longer cycle life due to the large ionic radius of sodium ions (1.02 Å), which makes their diffusion much more difficult [25]. Moreover, the rapid capacity decay during cycling has been ascribed to the decomposition of Na2Ti3O7 in electrolyte [26]. Therefore, improvement of the rate performance of Na2Ti3O7 remains a great challenge [27]. Introducing additional vacancies to accommodate Na+ is favorable for reversible and fast ion intercalation and deintercalation. Since fluorine (F) is the most electronegative element, F-doping has demonstrated a certain influence on the crystal structure and stability of NaxV2O2(PO4)2F, leading to improved electrochemical performance [28].
Here, we studied the effects of F-doping on the electrochemical performance of Na2Ti3O7. The results demonstrated a 30% increase in the specific capacity of Na2Ti3O7 through F-doping thanks to the improved Na+ diffusion coefficient. Better rate capability and cycle performance were also observed.

2. Preparation and Characterization

2.1. The Preparation of Na2Ti3O7 and F-Doping Na2Ti3O7 Samples

All reagents in this study were purchased and used directly. Titanium (IV) oxide, anatase (99.6%) was purchased from Alfa Aesar (Shanghai, China) and anhydrous Na2CO3 (99.8%) was bought from Sinopharm Chemical Regent Co., Ltd. (Shanghai, China). A solid-phase method was used to prepare Na2Ti3O7. Titanium oxide and sodium carbonate were mixed well at a molar ratio of 1:3. A 5% excess of Na2CO3 was added to prevent the composition of the Na2Ti16O13 impurity. The mixture was ground for 120 min. Then, the mixture powder was pressed to thin pellets, which were sintered at 800 °C for 10 h in a muffle furnace. After the muffle furnace was cooled to room temperature, the product, Na2Ti3O7 (NTO), was obtained. Na2Ti3O7Fx (NTOFx) was prepared by the same procedure, except that the calculated amount of NaF was added and the stoichiometric ratio of titanium and sodium was retained. The Na2Ti3O7 samples with different F-doping amounts were labeled as Na2Ti3O7Fx (x = 0.1, 0.2, 0.3 and 0.4).

2.2. Phase Analysis and Morphology Characterization

The synthesized Na2Ti3O7 and Na2Ti3O7Fx samples were characterized by an X-ray diffraction analyzer (PANalytical X’Pert3 Powder, Malvern Panalytical, Almelo, The Netherlands) with Cu/Kα (λ = 1.54178 Å) radiation in the 2 theta range from 5° to 80°. The morphologies of NTO and NTOF0.3 were measured on a scanning electron microscope (SU8020) (HITACHI, Tokyo, Japan). High-resolution TEM (HRTEM) was tested on an FEI Tecnai, Model G2 F20S-Twin (Brno, Czech), at a working voltage of 200 kV.

2.3. Electrochemical Test

For the preparation of the anode electrode, 70 wt% active material (NTO or NTOFs), 20 wt% acetylene black (Alfa Aesar, Shanghai, China), 10 wt% polyvinylidene fluoride (PVDF, KE JING, Hefei, China) as a binder, and N-methylpyrrolodone (NMP, Alfa Aesar, Shanghai, China) as a solvent were mixed to form a slurry. Then the slurry was painted on cleaned carbon-coated aluminum foil and dried at 120 °C in a vacuum oven for 10 h. Coin-type cells were used to test the electrochemical performances. All cells were assembled in a glove box filled with argon in which oxygen and water contents were less than 0.1 ppm. The glass fiber 1822-047 membrane (Whatman, Shanghai, China) was used as the separator, 1 M NaClO4 in diethyl carbonate (DEC)/dimethyl carbonate (DMC) (1:1 volume ratio) was used as the electrolyte (MJS, Nanjing, China), and metal sodium (Innochem, Beijing, China)was used as the counter electrode. Cyclic voltammetry experiments were carried out with scan rates at a certain range from 0.02 mV s−1 to 5 mV s−1, with a voltage window between 0.01 V and 2.5 V (versus Na/Na+) via CHI 640 E (B15536). Electrochemical impedance spectroscopy (EIS) was tested on an Autolab PGSTAT 302N instrument (Herisau, Switzerland) in a frequency range from 100 KHz to 100 mHz with a bias voltage of 5 mV.

3. Results and Discussion

3.1. Structural and Composition Characterization

As shown in Figure 1, the Na2Ti3O7 samples with F-doping amounts of 0, 10%, 20%, 30%, and 40% are hereafter referred to as Na2Ti3O7, Na2Ti3O7F0.1, Na2Ti3O7F0.2, Na2Ti3O7F0.3, Na2Ti3O7F0.1, and Na2Ti3O7F0.4 (x = 0.1, 0.2, 0.3, and 0.4), respectively. As the doping amount of F increases, the intensity of the diffraction peak first increases and then decreases. Several very sharp peaks indicate the high crystallinity of the samples.
The X-ray diffraction patterns of the NTO and NTOF0.3 samples are shown in Figure 2a. All the diffraction peaks correspond to monoclinic Na2Ti3O7 with a (PDF No. 31-1329) P121 space group. No impurity phases were observed, revealing that F-doping does not change the crystal structure of NTO. Figure 2b,c shows the morphology of the prepared NTO and NTOF0.3. It is indicated that the size of NTO particles is several micrometres. Most of the NTOF0.3 particles show nanorod morphologies (Figure 2c,d), indicating that F-doping has an effect on the morphology of the obtained product.
Figure 3a shows a TEM image of NTOF0.3. The diameter of the nanorod is approximately 100 nm, and the length is several micrometres. Figure 3b displays a high-resolution TEM (HRTEM) image of a single nanorod. The interplanar spacing of the ordered stripes marked in Figure 3b is about 0.84 nm, which corresponds to the (001) lattice plane of Na2Ti3O7. The selected area electron diffraction (SAED) pattern further indicates that the nanorod is a monoclinic Na2Ti3O7. Na2Ti3O7 nanorods grow along the (010) direction. To further examine the distribution of the F elements, energy-dispersive X-ray spectrometry (EDX) mapping analysis was employed. The results demonstrate that Na, O, Ti, and F elements are uniformly distributed in Figure 3d, which indicates that F was doped into Na2Ti3O7.

3.2. Electrochemical Performance

In Figure 4a,b, the cell with an NTOF0.3 electrode exhibits the best electrochemical performance. Therefore, we mainly focused on the investigation of NTOF0.3. Figure 5a,b exhibits the charge and discharge curves of the NTO and NTOF0.3 electrodes at a current density of 20 mA g−1. The initial reversible discharge and charge capacities of NTO and NTOF0.3 are 233.4 and 109.7 mAh g−1; 246.3 and 120.6 mAh g−1, respectively. The initial coulombic efficiency is only 46.7% and 48.9% but it reaches almost 100% in subsequent cycles. The coulombic efficiency of the initial cycle may originate from the irreversible formation of a solid electrolyte interphase (SEI) film [29,30,31,32]. Due to the formation of the passivating layer on the surface and the reactive Ti–O that leads to electrolyte decomposition, the irreversible capacity loss in the titanium-based Na electrode is usually serious. However, after the initial cycle, the low irreversible capacity loss can be suppressed and thus high efficiency can be achieved when active Ti–O is passivated [33]. There is a sloping voltage plateau at approximately 0.46 V during the charge process, while a plateau is obtained at about 0.63 V during the discharge process. Furthermore, Figure 5c shows the rate capabilities of NTO and NTOF0.3 electrodes at current densities ranging from 20 mA g−1 to 500 mA g−1. It can be seen that the specific capacity of the NTOF0.3 electrode is much higher than that of the NTO electrode. Figure 5d displays the cycling performance of NTO and NTOF0.3 electrodes. Both electrodes can run stably for more than 800 cycles at 100 mA g−1. The discharge specific capacity of the NTOF0.3 electrode is about 30% higher than that of the NTO electrode. Figure 5e,f displays the Nyquist plots of NTO and NTOF0.3 electrodes, respectively. All the plots show a depressed semicircle in the high-frequency region and a sloping line in the low-frequency region. It is believed that the former corresponds to the charge–transfer resistance (RCT) while the latter corresponds to the Warburg diffusion process. It can be seen that the NTOF0.3 electrode exhibits smaller semicircles (72.1 Ω) at high and medium frequencies when compared to those of the NTO electrode (147.7 Ω), indicating that F-doping can decrease the charge–transfer resistance. Compared with the resistances of the electrodes before and after the initial cycle, both the resistances of NTO and NTOF0.3 after the first cycle are significantly decreased. This indicates that an activation process occurs during the cycling.
To reveal why the NTOF0.3 electrode exhibits enhanced performance compared with the NTO electrode, the electrochemical kinetics of Na+ deintercalation and intercalation processes in Na2Ti3O7 were studied. Figure 6a,b shows the Cyclic voltammetry (CV) curves of the NTO and NTOF0.3 electrodes. A cathodic peak at 0.16 V and an anodic peak at 0.68 V versus Na/Na+ were observed, corresponding to typical Na+ insertion/extraction in the NTO lattice, which are in consistent with the charge and discharge curves. Significantly, a redox pair developed at 0.68 V and gradually augmented with the scanning rate. This redox may be ascribed to the storage of sodium with low-valence-state titanium atoms. These two redox peaks were assigned to the redox couple of Ti4+/Ti3+ over the discharge and charge processes. The reaction procedures can be depicted by Formula (1) [34]:
Na2Ti3O7 + x Na+ + x e → Na2+xTi3O7 (0 < x < 3.5)
The redox peaks of NTOF0.3 demonstrate its outstanding kinetic property. Figure 6c,d reveals the dependence of the logarithm of peak currents (log i) on the logarithm of the scan rates (log v). Furthermore, the reaction kinetics can be revealed by the formula of i = avb, which can also be expressed as log i = b × log v + log a, where i is the peak current, a and b denote related parameters, and ν represents the sweep rate [35,36]. When the value of b is close to 0.5, a battery behavior dominates the process; when the b value approaches 1.0, it shows the behavior of a capacitor. In Figure 6c, the log i versus the log v shows a linear relationship. Thus, the b values of the two peaks at 0.25 V and 0.16 V can be calculated as 0.58 and 0.45, respectively, while the values of b of the two redox peaks for NTOF0.3 are 0.47 and 0.46, respectively. Therefore, these results indicate that a mixed process exists in the NTO and NTOF0.3 anodes, although a diffusion-controlled process via a capacitive Na+ storage mechanism should be more dominant.
According to Equation (2) [37]:
I P = 2.69 × 10 5 A n 3 2 C 0 D 1 2 v 1 2
where Ip represents peak current (A), A is the electrode area (cm2), n is the number of electrons transferred, C0 is the concentration of Na+ ion (cm2 s−1), D is the diffusion coefficient of Na+ (mol cm−3), and v is the sweeping rate (V s−1) in CV. Thus, the calculated diffusion coefficients of Na ions in NTO and NTOF0.3 electrodes are 7.73 × 10−9 cm2 s−1 and 1.7 × 10−8 cm2 s−1, respectively. This result is consistent with the EIS results. These results indicate that F-doping is favourable for improving electrochemical performance.

4. Conclusions

In summary, F-doped Na2Ti3O7 nanorods were synthesized successfully by a solid-phase method. The effects of F-doping on the phase, morphology, and electrochemical performance of Na2Ti3O7 were investigated. F-doping does not change the crystal structure of NTO, although it has an effect on the morphology of the resultant product. The specific capacity of Na2Ti3O7 displays a 30% increase by F-doping due to the improved Na+ diffusion coefficient. F-doping can the charge–transfer resistance. The obtained material also shows a better rate capability and cycling performance for more than 800 cycles. F-doped Na2Ti3O7 nanorods are a promising anode for sodium-ion batteries (SIBs). This work provides a strategy for improving the electrochemical performance of the electrode materials.

Author Contributions

C.S. and C.Z. designed this work; L.L. and Q.Z. carried out the synthesis and electrochemical experiments and XRD measurements; Z.C., X.C. and Y.G. performed the SEM and TEM; L.L. wrote the paper; all the authors participated in the analysis of the experimental data and discussions of the results, as well as preparation of the paper.

Funding

This research is supported by the National Natural Science Foundation of China (No. 51672029, 51372271, U1361119, 61503124), the National Key R & D Project from Ministry of Science and Technology, China (2016YFA0202702), the Natural Science Foundation of Henan Province (General Program) (No. 162300410119, 162300410115), and the Special Medical Project of Henan Polytechnic University (NSFRF1621).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yoshino, A. The birth of the lithium-ion battery. Angew. Chem. Int. Ed. Engl. 2012, 51, 5798–5800. [Google Scholar] [CrossRef]
  2. Scrosati, B.; Garche, J. Lithium batteries: Status, prospects and future. J. Power Sources 2010, 195, 2419–2430. [Google Scholar] [CrossRef]
  3. Liu, C.; Li, F.; Ma, L.P.; Cheng, H.M. Advanced materials for energy storage. Adv. Mater. 2010, 22, E28–E62. [Google Scholar] [CrossRef]
  4. Slater, M.D.; Kim, D.; Lee, E.; Johnson, C.S. Sodium-Ion Batteries. Adv. Funct. Mater. 2013, 23, 947–958. [Google Scholar] [CrossRef]
  5. Lee, K.T.; Ramesh, T.N.; Nan, F.; Botton, G.; Nazar, L.F. Topochemical Synthesis of Sodium Metal Phosphate Olivines for Sodium-Ion Batteries. Chem. Mater. 2011, 23, 3593–3600. [Google Scholar] [CrossRef]
  6. Li, Z.; Young, D.; Xiang, K.; Carter, W.C.; Chiang, Y.-M. Towards High Power High Energy Aqueous Sodium-Ion Batteries: The NaTi2(PO4)3/Na0.44MnO2 System. Adv. Eng. Mater. 2013, 3, 290–294. [Google Scholar] [CrossRef]
  7. Hou, H.D.; Gan, B.H.; Gong, Y.D.; Chen, N.; Sun, C.W. P2-Type Na0.66Ni0.23Mg0.1Mn0.67O2 as a High Performance Cathode for Sodium-Ion Battery. Inorg. Chem. 2016, 55, 9033–9037. [Google Scholar] [CrossRef]
  8. Hou, H.D.; Xu, Q.K.; Pang, Y.K.; Li, L.; Wang, J.L.; Zhang, C.; Sun, C.W. Efficient Storing Energy Harvested by Triboelectric Nanogenerators Using a Safe and Durable All-Solid-State Sodium-Ion Battery. Adv. Sci. 2017, 4, 1700072. [Google Scholar] [CrossRef] [Green Version]
  9. Ma, Z.H.; Wang, Y.S.; Sun, C.W.; Alonso, J.A.; Fernández-Díaz, M.T.; Chen, L.Q. Experimental Visualization of the Diffusion Pathway of Sodium Ions in the Na3[Ti2P2O10F] Anode for Sodium-Ion Battery. Sci. Rep. 2014, 4, 7231. [Google Scholar] [CrossRef]
  10. Ma, Z.H.; Sun, C.W.; Lv, Y.C.; Wang, Y.S.; Kim, Y.; Chen, L.Q. A New Oxyfluorinated Titanium Phosphate Anode for a High-Energy Lithium-Ion Battery. ACS Appl. Mater. Interfaces 2015, 7, 1270–1274. [Google Scholar] [CrossRef]
  11. Trad, K.; Carlier, D.; Croguennec, L.; Wattiaux, A.; Ben Amara, M.; Delmas, C. NaMnFe2(PO4)3 Alluaudite Phase: Synthesis, Structure, and Electrochemical Properties As Positive Electrode in Lithium and Sodium Batteries. Chem. Mater. 2010, 22, 5554–5562. [Google Scholar] [CrossRef]
  12. Raju, V.; Rains, J.; Gates, C.; Luo, W.; Wang, X.; Stickle, W.F.; Stucky, G.D.; Ji, X. Superior cathode of sodium-ion batteries: orthorhombic V2O5 nanoparticles generated in nanoporous carbon by ambient hydrolysis deposition. Nano Lett. 2014, 14, 4119–4124. [Google Scholar] [CrossRef]
  13. Alcántara, R.; Jaraba, M.; Lavela, P.; Tirado, J.L. NiCo2O4 Spinel:  First Report on a Transition Metal Oxide for the Negative Electrode of Sodium-Ion Batteries. Chem. Mater. 2002, 14, 2847–2848. [Google Scholar] [CrossRef]
  14. Jian, Z.; Zhao, L.; Pan, H.; Hu, Y.-S.; Li, H.; Chen, W.; Chen, L. Carbon coated Na3V2(PO4)3 as novel electrode material for sodium ion batteries. Electrochem. Commun. 2012, 14, 86–89. [Google Scholar] [CrossRef]
  15. Sun, Q.; Ren, Q.-Q.; Li, H.; Fu, Z.-W. High capacity Sb2O4 thin film electrodes for rechargeable sodium battery. Electrochem. Commun. 2011, 13, 1462–1464. [Google Scholar] [CrossRef]
  16. Yuan, S.; Zhao, Y.; Wang, Q. Layered Na2V6O16 nanobelts as promising cathode and symmetric electrode for Na-ion batteries with high capacity. J. Alloys Compd. 2016, 688, 55–60. [Google Scholar] [CrossRef]
  17. Yamada, Y.; Doi, T.; Tanaka, I.; Okada, S.; Yamaki, J.-I. Liquid-phase synthesis of highly dispersed NaFeF3 particles and their electrochemical properties for sodium-ion batteries. J. Power Sources 2011, 196, 4837–4841. [Google Scholar] [CrossRef]
  18. Pan, H.; Lu, X.; Yu, X.; Hu, Y.-S.; Li, H.; Yang, X.-Q.; Chen, L. Sodium Storage and Transport Properties in Layered Na2Ti3O7 for Room-Temperature Sodium-Ion Batteries. Adv. Eng. Mater. 2013, 3, 1186–1194. [Google Scholar] [CrossRef]
  19. Rudola, A.; Saravanan, K.; Mason, C.W.; Balaya, P. Na2Ti3O7: an intercalation based anode for sodium-ion battery applications. J. Mater. Chem. A 2013, 1, 2653–2662. [Google Scholar] [CrossRef]
  20. Väli, R.; Jänes, A.; Thomberg, T.; Lust, E. D-Glucose Derived Nanospheric Hard Carbon Electrodes for Room-Temperature Sodium-Ion Batteries. J. Electrochem. Soc. 2016, 163, A1619–A1626. [Google Scholar] [CrossRef]
  21. Wang, K.; Jin, Y.; Sun, S.; Huang, Y.; Peng, J.; Luo, J.; Zhang, Q.; Qiu, Y.; Fang, C.; Han, J. Low-Cost and High-Performance Hard Carbon Anode Materials for Sodium-Ion Batteries. ACS Omega 2017, 2, 1687–1695. [Google Scholar] [CrossRef]
  22. Zukalová, M.; Pitňa Lásková, B.; Mocek, K.; Zukal, A.; Bouša, M.; Kavan, L. Electrochemical performance of sol-gel-made Na2Ti3O7 anode material for Na-ion batteries. J. Solid State Electrochem. 2018, 22, 2545–2552. [Google Scholar] [CrossRef]
  23. Kim, Y.; Ha, K.H.; Oh, S.M.; Lee, K.T. High-capacity anode materials for sodium-ion batteries. Chemistry 2014, 20, 11980–11992. [Google Scholar] [CrossRef]
  24. Araújo-Filho, A.A.; Silva, F.L.R.; Righi, A.; da Silva, M.B.; Silva, B.P.; Caetano, E.W.S.; Freire, V.N. Structural, electronic and optical properties of monoclinic Na2Ti3O7 from density functional theory calculations: A comparison with XRD and optical absorption measurements. J. Solid State Chem. 2017, 250, 68–74. [Google Scholar] [CrossRef]
  25. Kim, S.-W.; Seo, D.-H.; Ma, X.; Ceder, G.; Kang, K. Electrode Materials for Rechargeable Sodium-Ion Batteries: Potential Alternatives to Current Lithium-Ion Batteries. Adv. Eng. Mater. 2012, 2, 710–721. [Google Scholar] [CrossRef]
  26. Dynarowska, M.; Kotwiński, J.; Leszczynska, M.; Marzantowicz, M.; Krok, F. Ionic conductivity and structural properties of Na2Ti3O7 anode material. Solid State Ionics 2017, 301, 35–42. [Google Scholar] [CrossRef]
  27. Chen, J.; Zhou, X.; Mei, C.; Xu, J.; Wong, C.-P. Improving the sodiation performance of Na2Ti3O7 through Nb-doping. Electrochim. Acta 2017, 224, 446–451. [Google Scholar] [CrossRef]
  28. Xu, M.W.; Xiao, P.H.; Shannon, S.; Song, J.; Henkelman, G.; Goodenough, J.B. Theoretical and Experimental Study of Vanadium-Based Fluorophosphate Cathodes for Rechargeable Batteries. Chem. Mater. 2014, 26, 3089–3097. [Google Scholar] [CrossRef]
  29. Agubra, V.A.; Zuniga, L.; Flores, D.; Campos, H.; Villarreal, J.; Alcoutlabi, M. A comparative study on the performance of binary SnO2/NiO/C and Sn/C composite nanofibers as alternative anode materials for lithium ion batteries. Electrochim. Acta 2017, 224, 608–621. [Google Scholar] [CrossRef]
  30. Agubra, V.A.; Zuniga, L.; Flores, D.; Villareal, J.; Alcoutlabi, M. Composite Nanofibers as Advanced Materials for Li-ion, Li-O2 and Li-S Batteries. Electrochim. Acta 2016, 192, 529–550. [Google Scholar] [CrossRef]
  31. Ji, L.; Lin, Z.; Alcoutlabi, M.; Zhang, X. Recent developments in nanostructured anode materials for rechargeable lithium-ion batteries. Energy Environ. Sci. 2011, 4, 2682–2699. [Google Scholar] [CrossRef]
  32. Ji, L.; Meduri, P.; Agubra, V.; Xiao, X.; Alcoutlabi, M. Graphene-Based Nanocomposites for Energy Storage. Adv. Eng. Mater. 2016, 6, 1502159. [Google Scholar] [CrossRef] [Green Version]
  33. He, Y.-B.; Li, B.; Liu, M.; Zhang, C.; Lv, W.; Yang, C.; Li, J.; Du, H.; Zhang, B.; Yang, Q.-H.; et al. Gassing in Li4Ti5O12-based batteries and its remedy. Sci. Rep. 2012, 2, 913. [Google Scholar] [CrossRef] [Green Version]
  34. Song, T.; Chen, H.; Xu, Q.; Liu, H.; Wang, Y.G.; Xia, Y. Black Phosphorus Stabilizing Na2Ti3O7/C Each Other with an Improved Electrochemical Property for Sodium-Ion Storage. ACS Appl. Mater. Interfaces 2018, 10, 37163–37171. [Google Scholar] [CrossRef]
  35. Wang, L.; Dong, Y.; Zhao, K.; Luo, W.; Li, S.; Zhou, L.; Mai, L. Interconnected LiCuVO4 networks with in situ Cu generation as high-performance lithium-ion battery anode. Phys. Chem. Chem. Phys. 2017, 19, 13341–13347. [Google Scholar] [CrossRef]
  36. Kim, H.S.; Cook, J.B.; Lin, H.; Ko, J.S.; Tolbert, S.H.; Ozolins, V.; Dunn, B. Oxygen vacancies enhance pseudocapacitive charge storage properties of MoO3−x. Nat. Mater. 2017, 16, 454–460. [Google Scholar] [CrossRef]
  37. Dunn, B.; Kamath, H.; Tarascon, J.-M. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of NTO and NTOFx: (a) NTO; (b) NTOF0.1; (c) NTOF0.2; (d) NTOF0.3; and (e) NTOF0.4.
Figure 1. X-ray diffraction (XRD) patterns of NTO and NTOFx: (a) NTO; (b) NTOF0.1; (c) NTOF0.2; (d) NTOF0.3; and (e) NTOF0.4.
Materials 11 02206 g001
Figure 2. (a) XRD patterns of NTO and NTOF0.3; (b) morphology of NTO; (c,d) morphology of NTOF0.3.
Figure 2. (a) XRD patterns of NTO and NTOF0.3; (b) morphology of NTO; (c,d) morphology of NTOF0.3.
Materials 11 02206 g002
Figure 3. (a) TEM image and (b) high-resolution TEM (HRTEM) image of the as-prepared NTOF0.3; (c) selected area electron diffraction (SAED) patterns of NTOF0.3; (d) the opposite energy-dispersive X-ray spectrometry (EDX) elemental mappings of Na, Ti, O and F.
Figure 3. (a) TEM image and (b) high-resolution TEM (HRTEM) image of the as-prepared NTOF0.3; (c) selected area electron diffraction (SAED) patterns of NTOF0.3; (d) the opposite energy-dispersive X-ray spectrometry (EDX) elemental mappings of Na, Ti, O and F.
Materials 11 02206 g003
Figure 4. (a) The specific capacity performance of NTO and NTOFx at different current densities and (b) the cycling property of NTO and NTOFx at a current density of 100 mA g−1.
Figure 4. (a) The specific capacity performance of NTO and NTOFx at different current densities and (b) the cycling property of NTO and NTOFx at a current density of 100 mA g−1.
Materials 11 02206 g004
Figure 5. Electrochemical performances of NTO and NTOF0.3 electrodes. (a) Charge–discharge profiles of NTO; (b) charge–discharge profiles of NTOF0.3; (c) rate capability of NTO and NTOF0.3 electrodes; (d) cycling performance of NTO and NTOF0.3 electrodes; Nyquist plots of (e) NTO electrode and (f) NTOF0.3 electrode before and after the initial cycle at a current density of 100 mA g−1, tested under open circuit voltage conditions with a bias voltage of 5 mV.
Figure 5. Electrochemical performances of NTO and NTOF0.3 electrodes. (a) Charge–discharge profiles of NTO; (b) charge–discharge profiles of NTOF0.3; (c) rate capability of NTO and NTOF0.3 electrodes; (d) cycling performance of NTO and NTOF0.3 electrodes; Nyquist plots of (e) NTO electrode and (f) NTOF0.3 electrode before and after the initial cycle at a current density of 100 mA g−1, tested under open circuit voltage conditions with a bias voltage of 5 mV.
Materials 11 02206 g005
Figure 6. (a,b) Cyclic voltammetry (CV) curves of the NTO and NTOF0.3 electrodes at different scan rates; (c,d) the corresponding linear relationship between the logarithm of peak current and the logarithm of the scan rate (log v) for the NTO electrode and NTOF0.3 electrode.
Figure 6. (a,b) Cyclic voltammetry (CV) curves of the NTO and NTOF0.3 electrodes at different scan rates; (c,d) the corresponding linear relationship between the logarithm of peak current and the logarithm of the scan rate (log v) for the NTO electrode and NTOF0.3 electrode.
Materials 11 02206 g006

Share and Cite

MDPI and ACS Style

Chen, Z.; Lu, L.; Gao, Y.; Zhang, Q.; Zhang, C.; Sun, C.; Chen, X. Effects of F-Doping on the Electrochemical Performance of Na2Ti3O7 as an Anode for Sodium-Ion Batteries. Materials 2018, 11, 2206. https://doi.org/10.3390/ma11112206

AMA Style

Chen Z, Lu L, Gao Y, Zhang Q, Zhang C, Sun C, Chen X. Effects of F-Doping on the Electrochemical Performance of Na2Ti3O7 as an Anode for Sodium-Ion Batteries. Materials. 2018; 11(11):2206. https://doi.org/10.3390/ma11112206

Chicago/Turabian Style

Chen, Zehua, Liang Lu, Yu Gao, Qixiang Zhang, Chuanxiang Zhang, Chunwen Sun, and Xingying Chen. 2018. "Effects of F-Doping on the Electrochemical Performance of Na2Ti3O7 as an Anode for Sodium-Ion Batteries" Materials 11, no. 11: 2206. https://doi.org/10.3390/ma11112206

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