Gyeonghee Leea,
Xiao Zhanga,
Hongbo Zhanga,
Chakrapani V. Varanasi*b and
Jie Liu*a
aDepartment of Chemistry, Duke University, Durham, North Carolina 27708, USA
bArmy Research Office, Durham, North Carolina 27703, USA. E-mail: chakrapani.v.varanasi.civ@mail.mil; j.liu@duke.edu
First published on 15th June 2016
In sodium ion batteries, the ease of insertion and extraction of sodium ions in the electrode materials is one of the key parameters for the overall performance. In this article, the electrochemical sodium ion insertion in layered titanium hydrogeno phosphates (TiP) has been studied. In this material, the interlayer spacing and the particle morphology can be controlled by the choice of synthesis methods. Both nanostructured TiP and the coarse grained bulk counterpart were synthesized and properties were compared. While the specific capacity of nanostructured TiP materials was found to be not sensitive to the interlayer spacing, the specific capacity of coarse grained bulk TiP materials was significantly increased as the interlayer spacing was increased with the intercalation of water molecules in the layered host structure. These results indicate that interlayer spacing may not be the primary factor for Na-ion diffusion in nanostructured materials, where many interstitials are available for Na-ion diffusion. It is shown that nanostructured TiP materials can deliver excellent rate capability, and long term cycle stability with stable reversible capacity without the need of interlayer spacing expansion. The electrochemical properties of nanostructured materials were further enhanced when prepared as composites with carbon nanotubes that enhance the overall conductivity of the electrode materials.
Titanium hydrogeno phosphates, of a formula Ti(HPO4)2·xH2O (TiP–xH2O), are built on a two-dimensional structure consisting of PO3(OH) tetrahedra and TiO6 octahedra.12,13 They have been proposed as possible Li- and Na-ion conductors through both cation exchange reactions and direct cation insertion into the structure.13 The interlayer spacing of TiP–xH2O varies with the amounts of water in the structure.14,15 Recently, several groups reported the beneficial roles of water intercalation in layered structured electrode materials. The common conclusions from the earlier work is that water intercalation not only expands the interlayer spacing but also screens the strong interaction between elements in host layers and the charge carriers especially those with larger size or higher charge such as Na+ and Mg2+.10,16,17 Therefore, water intercalation could improve the kinetics of charge mobility in the electrode as well as charge storage capacity and long term stability.10,16,17 In this regard, the amenable interlayer properties with water intercalation provide sound rationale to explore applicability of TiP–xH2O as electrode materials for Na-ion storage. Their Na-ion conduction and storage behaviors, however, have not been demonstrated. To the best of our knowledge, this is the first study on Na-ion storage in materials based on transition metal hydrogeno phosphates.
In this work, we investigated the reversible insertion of Na ions in Ti(HPO4)2·xH2O (x = 0, 1), denoted as TiP–xH2O. The interlayer spacing of TiP–xH2O was tuned with the amounts of water intercalation and the degree of crystallinity of samples was controlled by the synthesis routes. The expanded interlayer spacing with water intercalation incorporation enables better Na-ion diffusion in the bulk hydrated TiP–H2O materials. However, this beneficial effect of larger interlayer spacing was less pronounced when the size of TiP–xH2O particles was reduced into nanoscale. Both hydrated and dehydrated nanostructured TiP–xH2O displayed high specific capacity regardless the interlayer spacing. These results imply that ion diffusion in such nanostructured materials is likely independent of interlayer spacing varied with water intercalation. As literatures have shown that CNTs can be used to fabricate composites with improved mechanical properties,18,19 we chose CNTs as a stable and conductive substrate for the composite electrode. The electrochemical properties of nanostructured TiP–xH2O materials were further enhanced when prepared as composites with CNTs, exhibiting excellent rate capability, and long cycle stability.
:
20
:
10 followed by casting the mixture paste on aluminum foil. The areal density of electrode was controlled to be between 2.7–3.5 mg cm−2. The cell assembly was done in a pure argon filled-glove box. 1 M NaClO4 in ethylene carbonate (EC) and propylene carbonate (PC) (1/1 = v/v) was used as an electrolyte. A glass microfiber filter (grade GF/F; Whatman, U.S.) was used as a separator. Galvanostatic charge–discharge measurements were performed at room temperature using MacPile Battery Analyzer. Cells were tested at various current rates (0.1, 0.2, 0.5, 1.0, 2.0, 5.0C) in the voltage range of 1.5–2.8 V. Cyclic voltammetry (CV) at 0.1 mV s−1 and electrochemical impedance spectroscopy (EIS) measurements were conducted after 50 cycles of charge–discharge in a frequency range of 100 kHz to 0.01 Hz with an AC voltage amplitude of 10 mV (Bio-logic SP300 instrument). To perform ex situ XRD, Swagelok-type cells were used. Swagelok-type cells in charged or discharged states were disassembled in the glove box. Electrodes were rinsed with PC to remove the residual electrolyte salts.
FT-IR spectra also confirmed the structure according the interlayer spacing (Fig. 1B and S1†). The hydrated samples, bulk-TiP–H2O, nano-TiP–H2O, and nano-TiP–H2O–CNT shows characteristic bands of water intercalation. The narrow bands at 3556 cm−1 and 3480 cm−1 are assigned to vibrational mode of water molecule.21 The narrow peak at 1613 cm−1 corresponds to the bending mode of water molecule.21,22 A broad band around 3100 cm−1 was observed in dehydrated samples, bulk-TiP, nano-TiP, and nano-TiP–CNT, instead of the characteristic bands of water intercalation. This band corresponds to the P–OH stretching mode of the hydrogen bond.23
Fig. 1C shows TGA curves of bulk-TiP–H2O, nano-TiP–H2O, and nano-TiP–H2O–CNT. The weight loss around 280 °C is due to the dehydration process with the water intercalation removal. The weight loss above 400 °C corresponds to the dehydroxylation, accompanied with conversion into the pyrophosphate, TiP2O7. The total weight loss of ∼15% indicated the removal of two moles of water per mole of TiP–xH2O. The additional weight loss of ∼5% after 600 °C observed in nano-TiP–H2O–CNT suggested that the TiP–CNT composite samples contain ∼5 wt% of CNTs. Notably, both dehydration and dehydroxylation processes occurred at a lower temperature in nanostructured samples. For example, the dehydration started at ∼200 °C and ∼100 °C for bulk-TiP–H2O and nano-TiP–H2O, respectively. This result indicate that nanostructured TiP–xH2O is more defective and more disordered. This is consistent with our previous observation, where the weight loss attributed to the conversion reaction started earlier in amorphous Ni(OH)2 than crystalline Ni(OH)2.24
The morphology of bulk and nanostructured TiP–xH2O samples were characterized using SEM and TEM (Fig. 2). As shown in Fig. 2A, bulk-TiP–H2O samples had large plates with the size around 5 μm obtained after hydrothermal synthesis. The morphology of nanostructured nano-TiP–H2O prepared using the ethanol mediated solvothermal synthesis was found to significantly differ from that of bulk counterparts. The nano-TiP–H2O particles consist of flakes with the size of 200–300 nm as shown in Fig. 2C. This small flake-like morphology enabled the feasibility to form uniform composites with CNTs. These TiP–xH2O nanoflakes appear to grow along the CNTs as observed in both SEM and TEM images (Fig. 2E and G). The dehydrated TiP–xH2O samples maintained the same morphology of their hydrated precursors (Fig. 2B, D, F and H).
![]() | ||
| Fig. 2 SEM images of (A) bulk-TiP–H2O, (B) bulk-TiP, (C) nano-TiP–H2O, (D) nano-TiP, (E) nano-TiP–H2O–CNT, and (F) nano-TiP–CNT, and TEM images of (G) nano-TiP–H2O–CNT, and (H) nano-TiP–CNT. | ||
The electrochemical properties of different TiP–xH2O materials were evaluated using coin-type cells and they showed significant differences in the properties. Fig. 3 shows the comparisons of charge–discharge profiles of TiP–xH2O materials in Na-ion half cells at 0.1C, respectively. The plateau extends over 1.8–2.4 V indicated Na ions insertion operating the Ti3+/Ti4+ redox pair for both bulk and nanostructured TiP–xH2O materials. As shown in Fig. S2A,† a pair of broad redox peaks was observed in TiP–xH2O materials, corresponding to the Ti3+/Ti4+ redox pair. It is consistent with the phase transition observed in charge–discharge profiles. For bulk TiP materials, bulk-TiP–H2O shows the smaller potential gap between anodic and cathodic peaks, indicating that the better electrochemical reversibility of bulk-TiP–H2O compared to bulk-TiP materials. This result support the idea that the larger interlayer spacing by water intercalation can improve Na-ion mobility in the electrode materials. Na-ion storage capacity according to the interlayer spacing, however, largely depends on the microstructure of TiP–xH2O materials. In the case of bulk TiP–xH2O materials, the hydrated phase, bulk-TiP–H2O exhibited higher specific capacity of 56 mA h g−1 than the dehydrated phase, bulk-TiP (31 mA h g−1). The active material density is slightly higher in dehydrated TiP than hydrated phase, TiP–H2O due to the absence of water intercalation in the crystal structure.25 Thus, considering this active material density difference, the specific capacity increase by ∼55% with the water intercalation incorporation is significant. This capacity increase can be explained by the expanded interlayer spacing by the water intercalation. The larger interlayer spacing can lead to the faster Na-ion diffusion and better active material utilization during the sodium insertion.26,27
![]() | ||
| Fig. 3 Comparisons of galvanostatic charge–discharge curves of bulk-TiP–H2O, bulk-TiP, nano-TiP–H2O, and nano-TiP. | ||
Interestingly, nanostructured TiP–xH2O materials showed considerably different Na-ion storage characteristics from those of bulk counterpart. Both hydrated phase, nano-TiP–H2O and dehydrated phase, nano-TiP displayed higher but similar specific capacity regardless the presence of water intercalation. The specific capacity values of nano-TiP–H2O and nano-TiP were 66 mA h g−1 and 71 mA h g−1, respectively. These results implied that interlayer spacing expansion is less effective for ion diffusion in nanostructured materials. This result is surprising but can be explained based on the fact that a lot of interstitials in nanostructured materials are available for Na ions to diffuse through and thus can participate in redox process primarily in both hydrated and dehydrated samples. Thus, interlayer spacing has negligible impacts on the overall performance. However, nanostructured TiP material showed better rate capability compared to bulk-TiP–xH2O materials with interlayer expansion (Fig. S2B and S3†). In case of bulk TiP–xH2O materials, it lost almost all capacity at 0.5C, indicating limited ion diffusion. The particle size reduction to nanoscale greatly shortens Na-ion diffusion distances, substantially enhancing the rate-dependent capacity retention.28,29 It is possible to expect nanostructured materials have higher surface area that can participate in surface redox reactions. To confirm this possibility, BET surface area of bulk-TiP–H2O and nano-TiP–H2O was measure and N2 adsorption–desorption isotherms of these samples were shown in Fig. S4.† However, the measured BET surface area of 338 cm3 g−1 for bulk-TiP–H2O was more than 2 times higher than 132 cm3 g−1 for nano-TiP–H2O. The higher surface area of bulk-TiP–H2O can be due to the well-developed two dimensional plate-like morphology as seen in the SEM image (Fig. 2A). Despite the higher surface area, bulk-TiP–H2O materials showed lower capacity than nano-TiP–H2O materials. These results, thus, support the major charge storage mechanism in layered TiP·xH2O materials is Na-ion diffusion and intercalation into the host lattice, with minor contributions from surface redox reactions.
Since nano-TiP–xH2O materials showed higher Na-ion storage capacity, this nano-TiP–xH2O material and CNTs composite was prepared for the improved conductivity, and their electrochemical properties were further studied. CV curves of nano-TiP–H2O–CNT and nano-TiP–CNT composites are displayed in Fig. 4A. The pairs of current peak positioned at 2.0 and 2.3 V for nano-TiP–H2O–CNT and 1.9 and 2.3 V for nano-TiP–CNT correspond to the redox pair of Ti3+/Ti4+. Fig. 4B and S3† display the representative charge–discharge profiles of nano-TiP–H2O–CNT and nano-TiP–CNT cells at various current rates of 0.1, 0.2, 0.5, 1.0, 2.0 and 5.0C. Both nano-TiP–H2O–CNT and nano-TiP–CNT composites exhibited excellent rate capability at various current rates (Fig. 4C and S3†). A large capacity drop after the first discharge was observed in all cells and the possible reasons for such a behavior will be discussed later in detail. The high capacity values of 92 mA h g−1 and 84 mA h g−1 were obtained from nano-TiP–H2O–CNT and nano-TiP–CNT, respectively, in the second discharge cycle at 0.1C. These capacity values indicated about 0.88 and 0.75 Na-ion insertion per unit formula based on the calculation using the equation n = (3.6MC)/F, where n is the number of moles of Na-ion inserted (mol), F is Faraday constant (C mol−1), C is the specific capacity (mA h g−1), and M is the molecular weight of TiP–xH2O (g mol−1).30 Note that all the capacity values for nano-TiP–xH2O–CNT materials are calculated based on the total mass including both nano-TiP–xH2O and CNT. The specific capacity values of nano-TiP–H2O–CNT and nano-TiP–CNT composites were 57 mA h g−1 and 56 mA h g−1, respectively, at a high current rate of 5.0C. This excellent rate performance can be attributed to the synergistic contributions from shortened Na-ion diffusion length due to the nanostructure and improved electrical conductivity provided by CNTs.
The long term cycle stability of nano-TiP–H2O, nano-TiP, nano-TiP–H2O–CNT, and nano-TiP–CNT was evaluated at 2.0C. As discussed earlier, the water intercalation does not significantly affect the specific capacity of these nanostructured samples. In the case of nano-TiP–xH2O cells without CNT, the capacity faded rapidly in 200 cycles (Fig. 4D and S5†). On the other hand, nano-TiP–xH2O and CNT composite cells showed outstanding cycle stability. The charge storage capacity became stabilized after a capacity drop in the first discharge, exhibiting a stable reversible capacity of approximately 50 mA h g−1. This reversible capacity was almost retained over 1000 cycles (Fig. 4D and S5†). The observed excellent long cycle stability is thought to be due to CNT networks in the sample. The CNT network can provide mechanical stability to maintain the integrity of the electrodes against the repeated volume expansion and shrinkage during charge and discharge. Our solvothermal synthesis condition produced significantly reduced size of TiP–xH2O nanoflakes, which enables direct deposition on CNTs as well as more efficient active materials utilization during charge and discharge cycles.
The conductivity of different TiP–xH2O Na-ion half cells were measured by performing electrochemical impedance spectroscopy (EIS). Fig. 5 shows EIS results performed on stabilized cells after 20 cycles of charge and discharge. The Nyquist plot can be interpreted based on the equivalent circuit (inset of Fig. 5). R1 corresponds to the combination of ionic resistance of electrolyte, intrinsic resistance of the substrate, and the contact resistance at the interface between active material and the substrate.31 R1 appeared as the intercept at the real part in the high frequency region. The resistor, R2 paralleled with the constant phase elements, Q2, represents the charge transfer resistance corresponding to the diameter of semicircle.32 The ion diffusion in the host structure is described with the Warburg element (Zw).33 As presented, the simulated data (dotted line) from the equivalent circuit fit well the impedance data (symbol) for all TiP–xH2O samples. The resistance values obtained from fitting and Na-ion diffusion coefficient (DNa+) estimated using Fig. S6† are summarized in Table 1. As is shown, resistance originated from the intrinsic properties (R1) such as electrolyte and substrate is similar for all the electrodes. However, the charge transfer resistances (R2) of nanostructured materials was almost half those of bulk crystalline materials. This significantly reduced resistance is likely due to the shortened electron transport distance in nanostructured materials. More importantly, Na-ion storage properties of bulk-TiP–xH2O that depend on the interlayer spacing was also confirmed by EIS results. The hydrated bulk-TiP–H2O exhibited much smaller charge transfer resistance (1212 Ω) than dehydrated bulk-TiP (1622 Ω). In addition, the DNa+ calculated for bulk-TiP–H2O (7.5 × 10−12 cm2 s−1) is about 100 times higher than that of bulk-TiP (8.7 × 10−14 cm2 s−1). These results imply that Na-ion diffusion kinetics in such hydrated form can be enhanced due to the expanded interlayer spacing with the water intercalation. Meanwhile, the water intercalation can screen the interaction between elements in host layer and Na ions.17 However, the calculated DNa+ values for nano-TiP and nano-TiP–H2O materials are not affected by the water intercalation, supporting the idea that nano-TiP–xH2O could utilize interstitials for Na-ion storage rather than rely on insertion facilitated by the enlarged interlayer spacing. Moreover, nano-TiP–xH2O–CNT composites showed further decreased charge transfer resistance. This further reduced resistance is attributed to the fact that CNT network provides conducting pathways for electrons and thus enhances the charge transfer processes.
![]() | ||
| Fig. 5 Nyquist plots and equivalent circuit of bulk-TiP, bulk-TiP–H2O, nano-TiP, nano-TiP–H2O, nano-TiP–CNT, and nano-TiP–H2O–CNT. | ||
| Bulk | Nanostructured | |||||
|---|---|---|---|---|---|---|
| TiP | TiP–H2O | TiP | TiP–H2O | TiP–CNT | TiP–H2O–CNT | |
| R1 (Ω) | 26.11 | 11.58 | 12.1 | 26.9 | 39.08 | 39.12 |
| R2 (Ω) | 1622 | 1212 | 660.3 | 658 | 522.4 | 474.6 |
| DNa+ (cm2 s−1) | 8.7 × 10−14 | 7.5 × 10−12 | 1.0 × 10−12 | 1.4 × 10−12 | 1.6 × 10−11 | 3.7 × 10−11 |
To understand the origin of the large capacity drop after the first discharge cycle observed in all Na-ion half cells, ex situ XRD was performed on the electrode materials and resulting patterns are shown in Fig. 6. For this measurement, bulk sample, bulk-TiP–H2O was selected to clearly see the peak position change. As is shown, (002) obviously shifted from 11.6° to 10.6° after the cell was fully discharged, indicating the intercalation of Na-ion into TiP host. In the fully charged state, however, XRD patterns showed two peaks at 11.6° and 10.6°, respectively. This result suggested that the inserted Na ions are not completely extracted from TiP host when charging. Thus, active site is only partially available for the next discharge and charge cycles, resulting in large decrease in capacity. Importantly, however, the recovered (002) peak at 11.6° after the first cycle indicated that the water intercalation remains between layers, indicating the layered structure with the water intercalation can be maintained in the organic solvent-based electrolyte.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra08242d |
| This journal is © The Royal Society of Chemistry 2016 |