Air-Stable and High-Voltage Layered P3-Type Cathode for Sodium

Jun 11, 2019 - The cathode and anode electrodes were separated by a Whatman glass ... group of R3m. The lattice parameters of the P3-NNMM were determi...
0 downloads 0 Views 4MB Size
Research Article Cite This: ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

www.acsami.org

Air-Stable and High-Voltage Layered P3-Type Cathode for SodiumIon Full Battery Ya-Nan Zhou,†,‡,§ Peng-Fei Wang,‡,§ Xu-Dong Zhang,‡ Lin-Bo Huang,‡ Wen-Peng Wang,‡ Ya-Xia Yin,‡ Sailong Xu,*,† and Yu-Guo Guo*,‡ †

State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, China CAS Key Laboratory of Molecular Nanostructure and Nanotechnology, CAS Research/Education Center for Excellence in Molecular Sciences, Beijing National Laboratory for Molecular Sciences (BNLMS), Institute of Chemistry, Chinese Academy of Sciences (CAS), Beijing 100190, China

Downloaded via BUFFALO STATE on July 23, 2019 at 13:40:35 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: The development of highly efficient and stable cathodes for sodium-ion batteries (SIBs) is strategically critical to achieving large-scale electrical energy storage. Creating air-stable and high-voltage layered cathodes for sodium-ion full batteries still remains a challenge. Herein, we describe a rational design and preparation of a stable P3-Na2/3Ni1/4Mg1/12Mn2/3O2 cathode. The cathode displays a satisfactory working voltage of 3.6 V and excellent cyclic stability over 100 cycles at a 1 C rate without obvious capacity fading. The results of ex situ X-ray diffraction (XRD) demonstrate that the P3-type structure is well retained even when charged to 4.4 V. Furthermore, the structural characterization by XRD Rietveld refinement, scanning electron microscopy, and electrochemical testing certifies that the cathode maintains its structure commendably even when soaked in water for 12 h. In particular, the P3- Na2/3Ni1/4Mg1/12Mn2/3O2∥hard carbon full battery exhibits a desired competitively high voltage of 3.45 V and an attractive energy density of up to 412.2 W h kg−1 based on the cathode. The comprehensive results achieved by the specially designed strategy provide guidance toward the exploration of stable cathodes in the application of SIBs as modern energy-storage devices. KEYWORDS: Layered P3 cathode, Air stability, High voltage, Sodium-ion full battery, Electrochemistry voltage is lower than that of O3-type LiCoO2 (∼3.9 V vs Li+/ Li), which is caused possibly by the larger ionic size and lower Lewis acidity of Na+ than Li+.4,11−13 A lower operating voltage generally results in a lower energy density. Therefore, it is of great importance to achieve reversible Na-ion extraction at high operating voltage and outstanding structural stability for the advancement of practical SIBs. In comparison, layered oxides with P-type structures are attractive cathode candidates due to their potential highvoltage and special crystal characteristics. In light of the difference in calcinating temperature, P-type cathodes can be categorized into P2-type and P3-type. In general, the P2 structure is synthesized after high-temperature calcination (around 900 °C), whereas the P3 structure is generated under low-temperature calcination (around 700 °C). This suggests that less energy consumption is needed for the P3-type cathode. Note that the P3-type cathode shows much better

1. INTRODUCTION Sodium-ion batteries (SIBs) have been well considered as a promising technology for large-scale stationary energy storage owing to sodium’s low cost and huge abundant resources as well as its similar electrochemical mechanism to lithium.1−3 Since the first creation of the SIBs, it has still been challenging to advance cathode materials toward practical applications due to the large dependence of energy density of SIBs on the cathode materials.4 Significant efforts have been devoted to exploring cathode materials. Among those cathode materials such as polyanion compounds, Prussian blue analogues, and layered oxides, layered transition-metal oxides NaxTMO2 (TM = Ni, Co, Mn, etc.) are very promising thanks to the abundant and cost-saving material sources expected for the nextgeneration SIBs.5−8 According to the definition by Delmas et al.,9 NaxTMO2 with a layer structure can be classified into O3type and P-type. In O3-type cathodes, they deliver much higher reversible capacity; however, this type of layered oxide typically has lower working potential than its lithium-ion counterparts. Such an example is O3-type NaCoO2, which has an operating voltage of ∼3.0 V (vs Na+/Na).10 The operating © 2019 American Chemical Society

Received: April 26, 2019 Accepted: June 11, 2019 Published: June 11, 2019 24184

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

Research Article

ACS Applied Materials & Interfaces

Figure 1. P3-NNMM material: (a) XRD pattern and Rietveld refinement profile; (b) crystal structure; (c) SEM image, (d) TEM image, and (e) high-resolution TEM image, as well as (f) the TEM/EDS mapping images of the elements of Na, Ni, Mg, Mn, and O.

and the P3-type cathode is apt to undergo a P3−O′3 irreversible phase transition at high voltage around 4.25 V. Recent studies have clarified that the introduction of inactive Mg2+ in the TMO2 layer can allow more Na+ preservation into the prismatic sites to stabilize the overall charge balance of the cathode, thereby resulting in very stable electrochemistry upon extended cycling.25,26 Despite these encouraging improvements, it is still challenging to design a P3 cathode having an air-stable and outstanding stable structure in the high-voltage range against the unfavorable phase transformation. Herein, we design a stable layered P3Na2/3Ni1/4Mg1/12Mn2/3O2 (denoted as P3-NNMM) material for the SIB cathode, which integrates the Na+/vacancy ordered superlattice and inactive Mg2+ substitution, thus delivering a high operating voltage window of up to 4.4 V and stability against water. The P3-NNMM cathode displays a competitive energy density of 436 W h kg−1 at 0.1 C, superior cycling stability (78% capacity retention over 100 cycles at 1 C), and remarkable rate performance (a high discharge capacity of 62.8 mA h g−1 even at 20 C). Furthermore, the cathode displays outstanding compatibility with the hard carbon (HC) anode in the Na-ion full battery, which exhibits an excellent energy density of 412.2 W h kg−1 based on the mass of the cathode.

rate performance than the P2 structure owing to its larger sodium interlayer spacing. A typical P3 structure is characterized by the ABBCCA oxygen stacking per unit cell, with Na+ sandwiched between the TMO2 slabs through occupying capacious prismatic sites. This special structure provides open prismatic channels for sodium-ion transport during the sodiation/desodiation and low Na-ion diffusion barrier as well as higher operating voltage.14,15 Generally, the larger the spacing of the Na layer is, the easier the Na+ ions exchange with H+ ions by up-taking H2O from the atmosphere in P3-type oxides,16,17 which is, however, problematic during the handling of the materials and battery assembly and thus limits the practical applications of P3-type layered oxides. From previous studies,18−20 it is well recognized that the uptake of water could be inhibited greatly by inducing a very strong interlayer interaction, such as constructing a superlattice ordering of Na+/vacancy between the transition-metal layers at a specific Na content of x = 2/3 in the part NaxTMO2 materials. Therefore, the design of a special P3-type Na2/3TMO2 structure with Na+/vacancy ordered superlattice could realize outstanding air-stable performance. From the viewpoint of designing TMO2 layers, Ni1/3/Mn2/3based materials are promising owing to the redox reaction of Ni2+/Ni3+/Ni4+. Ni4+ provides higher working voltage than other transition metals, while Mn4+ is devoted to stabilizing the structure during Na+ extraction/insertion.21,22 However, the irreversible phase transition occurs inevitably during the redox reaction of Ni2+/Ni4+, which affects the cycling stability within the high-voltage window in the sort of NaxNi1/3/Mn2/3O2 cathodes, thus limiting the further development of high-energy cathodes.23,24 For instance, the P2-type cathode suffers from a P2−O2 irreversible phase transition at a high voltage of 4.25 V,

2. EXPERIMENTAL SECTION 2.1. Materials Fabrication. The P3-Na2/3Ni1/4Mg1/12Mn2/3O2 (P3-NNMM) material was prepared using a typical sol−gel method. An aqueous solution with a suitable amount of citric acid (AR, 99.5%) was added to a saline solution of NaNO3 (Sinopharm Chemical Reagent Co., Ltd., China, AR, 99.0%, 5 wt % excess), Mg(NO3)2 (Sinopharm Chemical Reagent Co., Ltd., China, AR, 99.0%), Ni(NO3)2 (Sinopharm Chemical Reagent Co., Ltd., China, AR, 98.0%), 50% Mn(NO3)2 (Sinopharm Chemical Reagent Co., Ltd., 24185

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

Research Article

ACS Applied Materials & Interfaces

Figure 2. (a) Charge/discharge profiles of the P3-NNMM and P3-NNM cathodes. (b) Long-term cycling performances and Coulombic efficiencies of the P3-NNMM and P3-NNM cathodes at 0.1 C (1 C = 133 mA g−1) rate in the voltage range of 2.0−4.4 V. (c) Cycling performance of P3-NNMM with Coulombic efficiency at 1 C (initial three cycles: activation processes in 0.1 C). (d) CV profiles of P3-NNMM between 2.0 and 4.4 V at different scan rates. (e) Peak current Ip as a function of the square root of scan rate v1/2 for P3-NNMM. (f) Rate performance of the P3NNMM and P3-NNM cathodes. Electrochemical performance of Na-ion full batteries: (g) initial charge/discharge voltage profiles at 0.1 C rate based on cathode (inset) and capacity retention during 100 cycles at 1 C rate (initial three cycles: activation processes in 0.1 C). (h) Rate capability tests at various discharge current densities from 0.5 to 10 C. (i) Comparison of operating voltage and specific capacity (based on the cathodes) between our and other Na-ion full batteries reported recently. China, AR, 98.0%) with a stoichiometric ratio in Na2/3Ni1/4Mg1/12Mn2/3O2 under strong stirring. Water was then evaporated by heating at 70 °C for 6 h to acquire a clear and viscous gel. The obtained gel was dried at 120 °C overnight to give rise to a precursor. The precursor was ground to a new powder and then annealed at 450 °C for 6 h. Finally, the resultant product was calcined up to 700 °C for 12 h in air. The calcined materials were put immediately into a glove box filled with argon to avoid any damage. For comparison, a P3-Na2/3Ni1/3Mn2/3O2 (P3-NNM) cathode was prepared via the same method without Mg(NO3)2. 2.2. Structure and Morphology Characterizations. X-ray diffraction (XRD) was performed using a Bruker D8 Advance diffractometer with Cu Kα (λ = 1.5418 Å) radiation. The Rietveld refinement of X-ray data was performed using TOPAS software based on the Rietveld method. The samples for ex situ XRD characterization were prepared in a glove box filled with Ar. Scanning electron microscope (SEM) images were taken using a Hitachi SU-8020. The high-resolution transmission electron microscopy (HRTEM) images and energy-dispersive spectroscopy (EDS) maps were obtained by TEM (JEM-2100F) with an accelerating voltage of 200 kV. The X-ray photoelectron spectra (XPS) of the materials were recorded on an ESCALAB 250Xi (Thermo Scientific) spectrometer equipped with an Al Kα achromatic X-ray source. 2.3. Electrochemical Measurements. For the electrochemical measurement, the CR2032 coin-type or Swagelok-type (for ex situ XRD characterization) cells were assembled in a glove box filled with Ar (H2O, O2 < 0.1 ppm). The working electrodes were prepared by

mixing the active material with Super P carbon black and polyvinylidene fluoride in a weight ratio of 7:2:1. The slurry was pasted on aluminum foil with an active material loading of 2−2.5 mg cm−2 and then dried at 70 °C under vacuum overnight. Na foil and hard carbon were used as the negative electrode for the half battery and full battery, respectively. In the fabrication of the Na-ion full battery, the mass ratio of cathode to hard carbon (Kureha, Japan) anode was about 1.2−1.3:1. The electrolyte was composed of 1 M NaPF6 in EC (ethylene carbonate)/DEC (diethyl carbonate) (1:1) with 5% FEC (fluoroethylene carbonate). The cathode and anode electrodes were separated by a Whatman glass fiber separator soaked with the electrolyte. The electrochemical measurements were carried out using an Arbin battery test system. Different cutoff voltage ranges of 2.0−4.4 and 2.0−4.25 V versus Na/Na+ were used to evaluate the sodium-storage performance of half and full batteries, respectively. Cyclic voltammetry measurements (CV) were performed using a Princeton electrochemical workstation.

3. RESULTS AND DISCUSSION The P3-NNMM was synthesized by a typical sol−gel method. The crystal structure was determined by combining XRD and Rietveld refinement. Figure 1a shows that all the XRD reflection peaks are indexed to a trigonal lattice with a space group of R3m. The lattice parameters of the P3-NNMM were determined to be a = b = 2.8915(6) Å, c = 16.796(4) Å, and V = 121.6(2) Å3, together with a convergence factor Rwp of 24186

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

Research Article

ACS Applied Materials & Interfaces

Figure 3. Comparison between the as as-synthesized P3-NNMM and soaked-in-water P3-NNMM: (a) XRD pattern, (b) lattice parameters, and (c) SEM images. (d) Charge/discharge curves (inset) and the cycling performance of the soaked-in-water P3-NNMM in the voltage range between 2.0 and 4.4 V at 0.1 C.

eV is observed, which corresponds to the Ni2+ state (Figure S3b).32,33 The Mg 1s peak at 1302.5 eV is also clearly resolved, reflecting the existence of Mg2+ (Figure S3c). The galvanostatic charge/discharge of the P3-NNMM electrode was measured in sodium-half batteries at room temperature. For comparison, a P3-NNM cathode without Mg doping was prepared via the same method without Mg(NO3)2. Figure 2a shows that the charge/discharge curves of the P3NNMM cathode are smoother than those of the P3-NNM cathode during the initial five cycles, even when the operating voltage is over 4.25 V. The smooth profiles strongly suggest a possibility that the irreversible phase transition of P3−O′3 is suppressed effectively, which will be illustrated by the later ex situ XRD analysis. The voltage plateau is observed between 3.5 and 4.0 V during the charge/discharge process, which well corresponds to the reversible oxidation/reduction peaks at 3.9/ 3.8 V observed in the CV curves (Figure S4). Both of them are well related to the presence of ordered intermediate phases caused by Na ordering rearrangement from the aforementioned Na+/vacancy ordering that was observed in Figure S1.28,34 Furthermore, the reversible storage capacity is calculated to be 125 mA h g−1. The value strongly suggests the approximate 0.47 Na+ de-/intercalation with the sequential reactions from Ni2+ to Ni4+ in which the reversible oxidation/ reduction peaks are clearly visible at 3.3/3.2, 3.4/3.3, 3.7/3.6, and 4.3/4.0 V (Figure S4). Note that the initial Coulombic efficiency of P3-NNMM is 95.4%, with a remarkable average voltage of 3.6 V, both of which are attractive to assemble Naion full batteries.35,36 In addition, the P3-NNMM cathode exhibits a superior cycling stability to the P3-NNM cathode (Figure 2b), which is assigned to the prevention of the irreversible phase transition. More importantly, the P3NNMM delivers a good structure stability with a capacity retention of 78% after undergoing 100 cycles and a Coulombic

7.15% (detailed crystallographic data on refined P3-NNMM are listed in Table S1). Note that two superstructural peaks at ∼27.7 and ∼29.3° are observed, as shown in the enlarged XRD pattern in Figure S1. Similar to the P3-NNMM cathode, the analogous superstructure also appears in the P3-NNM cathode (Figure S2). The superstructure originates from the ordering of in-plane Na-vacancy, which is expected to induce a very strong interlayer interaction and to prevent water from inserting into the layer of Na, as well as to promise the airstable performance.18,27,28 Figure 1b shows that the typical crystal structure of the P3-NNMM possesses a pattern of ABBCCA oxygen stacking; that is, three TMO2 sheets per unit share one face with one TMO6 octahedron and also share three edges with TMO6 octahedra of the next layer, thus offering only one sodium accommodation as the prismatic site.29,30 These special sites of the Na layer occupied afford an open trigonal prismatic path for the Na ion, resulting in a low Na-ion diffusion barrier without intervention of the intermediate and realizing an outstanding electrochemical kinetics in Na-ion extraction/insertion. The morphology of P3-NNMM was observed by SEM. Figure 1c shows that the P3-NNMM material has a typical nanoplate-like morphology. The nanoplate sizes range from 180 to 300 nm. The TEM image confirms the nanoplate-like morphology (Figure 1d). The high-resolution TEM visualization further reveals that the P3-NNMM nanoplate is highly crystalline (Figure 1e). A lattice plane with an interplanar spacing of 2.15 Å is well determined, which is attributed to the (104) plane that is also shown by the XRD pattern in Figure 1a. In addition, the TEM/EDS images reflect a uniform distribution of the elements of Na, Ni, Mg, Mn, and O (Figure 1f). Furthermore, the oxidation states of the species in the P3NNMM were probed by using XPS (Figure S3). The Mn 2p3/2 peak centered at 641.9 eV is visible, demonstrating the existence of Mn4+ (Figure S3a).31 The Ni 2p3/2 peak at 854.0 24187

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

Research Article

ACS Applied Materials & Interfaces

Figure 4. Ex situ XRD patterns of the (a) P3-NNM, (b) P3-NNMM electrode at various discharge and charge states, and (c) the roadmap of the recorded XRD patterns of P3-NNMM electrode corresponding to the crystal structural evolution in the initial cycle.

P3-NNMM material in water for 12 h followed by drying at 70 °C overnight.48 The P3-NNMM cathode soaked in water still shows a crystal structure almost identical to that of the assynthesized P3-NNMM cathode in terms of peak location (Figure 3a). In particular, the peaks located at ∼27.7 and ∼29.3° are still visible (Figure S10), which strongly suggests that the superstructure of the as-synthesized P3-NNMM cathode is well preserved against soaking for 12 h. Moreover, the comparison of the lattice parameters was performed between the as-synthesized P3-NNMM and the P3-NNMM soaked in water (Figure 3b) on the basis of XRD refinement (Figure S11). The detailed comparison is shown in Table S2, clearly reflecting that the lattice parameters (such as a, c, and V) of the P3-NNMM soaked in water are almost identical to those of the as-synthesized P3-NNMM. In addition, the nanoplate-like morphology of the as- synthesized P3-NNMM cathode is well maintained (Figure 3c). Therefore, the electrochemical performance of the P3-NNMM soaked in water was examined. Figure 3d displays that the charge/ discharge profiles of the P3-NNMM soaked in water are quite similar to those of the as-synthesized P3-NNMM, informing the approximate capacities of both electrodes. In particular, the P3-NNMM soaked in water exhibits an initial discharge capacity of 128.3 mA h g−1 and a capacity retention of 85.5% after 50 cycles (Figure 3d), both of which are comparable with those of the as-synthesized P3-NNMM (a first discharge capacity of 125 mA h g−1 and 87% capacity retention after going 50 cycles at 0.1 C), as shown in Figure 2b. In addition, a similar aging experiment was also performed in the P3-NNM cathode. As shown in Figure S12, the structure (Figure S12a) and electrochemical performance (Figure S12b) can be maintained well in the P3-NNM cathode after the aging experiment. However, one can clearly see a long voltage plateau about 4.25 V, corresponding to the irreversible phase transition of P3−O′3 and in turn leading to the large volume change, which limits the reversible cycle. Therefore, the asobtained similarities of P3-NNMM strongly mean a strong air stability against water and outstanding structure stability. In order to understand the Na intercalation/deintercalation mechanism of P3-NNM and P3-NNMM, the structural evolution was probed by ex situ XRD at different charge/ discharge states during the first cycle.49 Figure 4a,b shows that, when the cathode is charged, the diffraction peaks of (003) and (006) are shifted to lower angles, whereas the peaks of (015),

efficiency of more than 99% after the completion of the continuous cycles even at a high rate of 1 C (Figure 2c). To evaluate the Na+ diffusion kinetics of the P3-NNMM, the CV testing was performed at various scan rates between 2.0 and 4.4 V (Figure 2d). The diffusion coefficients of Na+ were calculated to be 5.18 × 10−10 cm2 s−1 in the oxidation reaction and 5.48 × 10−10 cm2 s−1 in the reduction reaction (Figure 2e). The values strongly suggest the fast Na+ mobility in P3NNMM when compared with most P2- and O3-type oxides. The rapid kinetics is further corroborated by the result of rate performance. Figure 2f shows that the P3-NNMM electrode displays highly improved rate capability compared with the P3NNM electrode at different current densities ranging from 0.1 to 20 C. Note that, even at an ultrahigh rate of 20 C, the available capacity of the former electrode reaches about 50% capacity retention of 0.1 C, indeed suggesting an excellent high rate performance. The performance of the P3-NNMM-assembled sodium-ion full battery was evaluated by using P3-NNMM as the cathode material and hard carbon as the anode material. The hard carbon was precycled to realize a stable solid electrolyte interphase before the assembly of the P3-NNMM∥HC full battery, as shown in Figures S5−S7.37−39 Figure 2g (inset) displays that the Na-ion full battery demonstrates a remarkably high initial specific discharge capacity of 119.6 mA h g−1, with a desired competitively high operating voltage of 3.45 V. The energy density was thus determined to be as high as 412.2 W h kg−1 on the basis of the cathode. In particular, there is no obvious capacity fading in cycling curves or no noticeable diversification in the potential and area of the redox peak (vs Na+/Na) in the CV cycles (Figure S8). After undergoing 100 cycles at 1 C, a capacity retention of 80% is still maintained (Figure 2g). All the obtained results strongly suggest a good cycling stability of the sodium-ion full battery. Furthermore, even at a high a rate of 10 C, the P3-NNMM∥HC full battery exhibits an energy density of up to 153.6 W h kg−1, reflecting a strong tolerance of sodiation/desodiation (Figure 2h and Figure S9). Note that the average voltage delivered in the P3NNMM∥HC full battery is closely comparable to those of the Na-ion full batteries published recently (Figure 2i).40−47 The poor air stability of layered oxides is largely due to the Na+ exchanged with H+ by up-taking H2O from the atmosphere. Therefore, to evaluate the air stability of P3NNMM, an aging experiment was carried out by soaking the 24188

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

Research Article

ACS Applied Materials & Interfaces

recently. Our strategy can be easily extended to explore other high-voltage air-stable materials used as promising cathodes in large-scale application for future practical SIBs.

(012), and (101) move to higher angles. The apparent shifts verify the Na+ extraction from the cathode in the charge profile combined with the expansion along the c axis and contraction along the a axis, which are due to the fact that the electrostatic repulsion between TMO2 slabs and the valence state of transition metals are both increased when Na+ is extracted, respectively.50 Note that, with the P3-NNM electrode charged above 4.2 V (Figure 4a), the reduction of a and increase of c lattice parameters appear clear gradually, and the interlayer is expanded. As a result, new hydrated peaks appears at 15.5 and 25.5°, which are attributed to the electrolyte solvent molecules and/or water molecules embedded in the sodium slabs of NaxTMO2 due to the expanded layer spacing, respectively.51−53 Moreover, a similar phenomenon occurs in the P3-NNMM cathode upon ∼0.4 Na+ extraction with the electrode charged above 4.0 V (Figure 4b). In addition, a characteristic peak at 19.9−20°, corresponding to the O′3 phase (Figure 4a), does not appear in the P3-NNMM electrode (Figure 4b) over the whole high-voltage region (4.25−4.4 V), which is indeed in agreement with the result of smooth curves rather than a long platform around 4.25 V in Figure 2a, indicating excellent high-voltage structural stability of the P3-NNMM.48,51 In the discharge process, the hydrated peak disappears, while the electrode undergoes a completely opposite process. Meanwhile, the TM ions are aligned along the c axis to form prismatic Na sites in the P3 structure. As a result, the layered P3 structure of both cathodes was recovered. Furthermore, the analysis by ex situ XRD of the P3-NNMM electrode is presented in a 3D visual map clearly and shows a good reversibility during electrochemical process (Figure 4c). In addition, the schematic diagram of the charge/discharge in Figure 4c reflects a solid-solution reaction of the P3 phase throughout the charge/discharge process for P3-NNMM.54 This result suggests that P3-NNMM is free from the P3−O′3 phase transition and the P3-stacked structure during cycling was retained, thereby guaranteeing the structural stability and the long cycle life.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.9b07299.



X-ray photoelectron spectra (XPS) of P3-NNMM; cyclic voltammograms of P3-NNMM; P3-NNMM∥HC Naion full batteries; long cycling performance of hard carbon; Rietveld refinement patterns of P3-NNMM and P3 NNMM soaked in water (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (S.X.). *E-mail: [email protected] (Y.-G.G.). ORCID

Peng-Fei Wang: 0000-0001-9882-5059 Ya-Xia Yin: 0000-0002-0983-9916 Sailong Xu: 0000-0002-1999-9323 Yu-Guo Guo: 0000-0003-0322-8476 Author Contributions §

Y.-N.Z. and P.-F.W. contributed equally to this work.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Y.-N.Z. and P.-F.W. contributed equally to this work. This work was supported by the Basic Science Center Project of National Natural Science Foundation of China (Grant no. 51788104), the National Natural Science Foundation of China (Grants No. 51772301, 21521005, U1607128, and 21773264), the ″Transformational Technologies for Clean Energy and Demonstration″, and Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDA 21070300), as well as the Fundamental Research Funds for the Central Universities (XK1802-6).

4. CONCLUSIONS In summary, we rationally designed a competitive P3-type cathode for the advanced sodium-ion full battery. The cathode obtained guaranteed the high structural stability and electrochemical utilization for efficient sodium storage, especially at the high operating voltage. The ex situ XRD patterns showed a reversible single-phase process by avoiding any unfavorable phase transition during the sodium insertion/extraction and even when increasing the charge voltage of the cell to 4.4 V, which not only provided a remarkable capacity of 125 mA h g−1 but also exhibited the long cycle stability with a good capacity retention of 78% after 100 cycles at 1 C. Moreover, the large trigonal prismatic sites occupied by Na+ in this P3 structure offered open diffusion paths for the sodium ion, which lead to the fast Na mobility (5.18 × 10−10/5.48 × 10−10 cm2 s−1) and outstanding rate performance (a capacity retention of 50% even at a high current density of 20 C). Most importantly, the cathode obtained indeed resisted considerable water, and its good structure stability, elucidated by XRD and SEM, and especially sodium-storage performance remained. In particular, the P3-NNMM∥HC full battery delivered a competitively high operating voltage of 3.45 V and energy density of up to 412.2 W h kg−1 based on the mass of the cathode, both of which are quite comparable to and outperform those of most of the Na-ion full batteries reported



REFERENCES

(1) Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Research Development on Sodium-Ion Batteries. Chem. Rev. 2014, 114, 11636−11682. (2) You, Y.; Manthiram, A. Progress in High-Voltage Cathode Materials for Rechargeable Sodium-Ion Batteries. Adv. Energy Mater. 2018, 8, 1701785. (3) Yan, Z.; Tang, L.; Huang, Y.; Hua, W.; Wang, Y.; Liu, R.; Gu, Q.; Indris, S.; Chou, S. L.; Huang, Y.; Wu, M.; Dou, S. X. A Hydrostable Cathode Material Based on the Layered P2@P3 Composite that Shows Redox Behavior for Copper in High-Rate and Long-Cycling Sodium-Ion Batteries. Angew. Chem., Int. Ed. 2019, 58, 1412−1416. (4) Kubota, K.; Kumakura, S.; Yoda, Y.; Kuroki, K.; Komaba, S. Electrochemistry and Solid-State Chemistry of NaMeO2 (Me = 3d Transition Metals). Adv. Energy Mater. 2018, 8, 1703415. (5) Wang, X.; Hu, P.; Niu, C.; Meng, J.; Xu, X.; Wei, X.; Tang, C.; Luo, W.; Zhou, L.; An, Q.; Mai, L. New-type K0.7Fe0.5Mn0.5O2 cathode with an expanded and stabilized interlayer structure for high-capacity sodium-ion batteries. Nano Energy 2017, 35, 71−78. (6) Deng, J.; Luo, W.-B.; Chou, S.-L.; Liu, H.-K.; Dou, S.-X. SodiumIon Batteries: From Academic Research to Practical Commercialization. Adv. Energy Mater. 2018, 8, 1701428. 24189

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

Research Article

ACS Applied Materials & Interfaces

0.05, 0.1) Na-Ion Cathodes with Enhanced Stability and Rate Capability. Chem. Mater. 2016, 28, 5087−5094. (26) Wang, P.-F.; You, Y.; Yin, Y.-X.; Wang, Y.-S.; Wan, L.-J.; Gu, L.; Guo, Y.-G. Suppressing the P2-O2 Phase Transition of Na0.67Mn0.67Ni0.33O2 by Magnesium Substitution for Improved Sodium-Ion Batteries. Angew. Chem., Int. Ed. 2016, 55, 7445−7449. (27) Wang, Y.; Xiao, R.; Hu, Y. S.; Avdeev, M.; Chen, L. P2Na0.6[Cr0.6Ti0.4]O2 Cation-Disordered Electrode for High-Rate Symmetric Rechargeable Sodium-Ion Batteries. Nat. Commun. 2015, 6, 6954. (28) Wang, P.-F.; Yao, H.-R.; Liu, X.-Y.; Yin, Y.-X.; Zhang, J.-N.; Wen, Y.; Yu, X.; Gu, L.; Guo, Y.-G. Na+/Vacancy Disordering Promises High-Rate Na-Ion Batteries. Sci. Adv. 2018, 4, No. eaar6018. (29) Kim, J.; Kwon, J.; Kim, M.; Do, J.; Lee, D.; Han, H. LowDielectric-Constant Polyimide Aerogel Composite Films with Low Water Uptake. Polym. J. 2016, 48, 829−834. (30) Komaba, S.; Yabuuchi, N.; Nakayama, T.; Ogata, A.; Ishikawa, T.; Nakai, I. Study on the Reversible Electrode Reaction of Na1‑xNi0.5Mn0.5O2 for a Rechargeable Sodium-Ion Battery. Inorg. Chem. 2012, 51, 6211−6220. (31) Wang, P.-F.; You, Y.; Yin, Y.-X.; Guo, Y.-G. An O3-type NaNi0.5Mn0.5O2 Cathode for Sodium-Ion Batteries with Improved Rate Performance and Cycling Stability. J. Mater. Chem. A 2016, 4, 17660−17664. (32) Chen, H.; Wu, Z.; Zheng, Z.; Chen, T.; Guo, X.; Li, J.; Zhong, B. Tuning the Component Ratio and Corresponding Sodium Storage Properties of Layer-Tunnel Hybrid Na0.6Mn1‑xNixO2 Cathode by a Simple Cationic Ni2+ Doping Strategy. Electrochim. Acta 2018, 273, 63−70. (33) Li, Z. Y.; Zhang, J.; Gao, R.; Zhang, H.; Hu, Z.; Liu, X. Unveiling the Role of Co in Improving the High-Rate Capability and Cycling Performance of Layered Na0.7Mn0.7Ni0.3‑xCoxO2 Cathode Materials for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 15439−15448. (34) Tie, D.; Gao, G.; Xia, F.; Yue, R.; Wang, Q.; Qi, R.; Wang, B.; Zhao, Y. Modulating the Interlayer Spacing and Na+/Vacancy Disordering of P2-Na0.67MnO2 for Fast Diffusion and High-Rate Sodium Storage. ACS Appl. Mater. Interfaces 2019, 11, 6978−6985. (35) Deng, J.; Luo, W.-B.; Lu, X.; Yao, Q.; Wang, Z.; Liu, H.-K.; Zhou, H.; Dou, S.-X. High Energy Density Sodium-Ion Battery with Industrially Feasible and Air-Stable O3-Type Layered Oxide Cathode. Adv. Energy Mater. 2018, 8, 1701610. (36) Ren, W.; Zhu, Z.; An, Q.; Mai, L. Emerging Prototype SodiumIon Full Cells with Nanostructured Electrode Materials. Small 2017, 13, 1604181. (37) Xiao, Y.; Wang, P.-F.; Yin, Y.-X.; Zhu, Y.-F.; Niu, Y.-B.; Zhang, X.-D.; Zhang, J.; Yu, X.; Guo, X.-D.; Zhong, B.-H.; Guo, Y.-G. Exposing {010} Active Facets by Multiple-Layer Oriented Stacking Nanosheets for High-Performance Capacitive Sodium-Ion Oxide Cathode. Adv. Mater. 2018, 30, No. e1803765. (38) Bai, P.; He, Y.; Xiong, P.; Zhao, X.; Xu, K.; Xu, Y. Long Cycle Life and High Rate Sodium-Ion Chemistry for Hard Carbon Anodes. Energy Storage Mater. 2018, 13, 274−282. (39) Dou, X.; Hasa, I.; Saurel, D.; Vaalma, C.; Wu, L.; Buchholz, D.; Bresser, D.; Komaba, S.; Passerini, S. Hard Carbons for Sodium-Ion Batteries: Structure, Analysis, Sustainability, and Electrochemistry. Mater. Today 2019, 23, 87−104. (40) Zhong, Y.; Xia, X.; Zhan, J.; Wang, Y.; Wang, X.; Tu, J. Monolayer Titanium Carbide Hollow Sphere Arrays Formed via an Atomic Layer Deposition Assisted Method and Their Excellent HighTemperature Supercapacitor Performance. J. Mater. Chem. A 2016, 4, 18717−18722. (41) Zeng, Z.; Jiang, X.; Li, R.; Yuan, D.; Ai, X.; Yang, H.; Cao, Y. A Safer Sodium-Ion Battery Based on Nonflammable Organic Phosphate Electrolyte. Adv. Sci. 2016, 3, 1600066. (42) Yuan, Z.; Si, L.; Zhu, X. Three-Dimensional Hard Carbon Matrix for Sodium-Ion Battery Anode with Superior-Rate Performance and Ultralong Cycle Life. J. Mater. Chem. A 2015, 3, 23403− 23411.

(7) Guo, S.; Li, Q.; Liu, P.; Chen, M.; Zhou, H. Environmentally Stable Interface of Layered Oxide Cathodes for Sodium-Ion Batteries. Nat. Commun. 2017, 8, 135. (8) Jiang, Y.; Tan, S.; Wei, Q.; Dong, J.; Li, Q.; Xiong, F.; Sheng, J.; An, Q.; Mai, L. Pseudocapacitive layered birnessite sodium manganese dioxide for high-rate non-aqueous sodium ion capacitors. J. Mater. Chem. A 2018, 6, 12259−12266. (9) Delmas, C.; Fouassier, C.; Hagenmuller, P. Structural classification and properties of the layered oxides. Physica B+C 1980, 99, 81−85. (10) Zheng, L.; Li, J.; Obrovac, M. N. Crystal Structures and Electrochemical Performance of Air-Stable Na2/3Ni1/3−xCuxMn2/3O2 in Sodium Cells. Chem. Mater. 2017, 29, 1623−1631. (11) Kim, H.; Kim, H.; Ding, Z.; Lee, M. H.; Lim, K.; Yoon, G.; Kang, K. Recent Progress in Electrode Materials for Sodium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1600943. (12) Kundu, D.; Talaie, E.; Duffort, V.; Nazar, L. F. The Emerging Chemistry of Sodium Ion Batteries for Electrochemical Energy Storage. Angew. Chem., Int. Ed. 2015, 54, 3431−3448. (13) Pan, H.; Hu, Y.-S.; Chen, L. Room-Temperature Stationary Sodium-Ion Batteries For Large-Scale Electric Energy Storage. Energy Environ. Sci. 2013, 6, 2338−2360. (14) Rong, X.; Liu, J.; Hu, E.; Liu, Y.; Wang, Y.; Wu, J.; Yu, X.; Page, K.; Hu, Y.-S.; Yang, W.; Li, H.; Yang, X.-Q.; Chen, L.; Huang, X. Structure-Induced Reversible Anionic Redox Activity in Na Layered Oxide Cathode. Joule 2018, 2, 125−140. (15) Zhou, Y.-N.; Wang, P.-F.; Niu, Y.-B.; Li, Q.; Yu, X.; Yin, Y.-X.; Xu, S.; Guo, Y.-G. A P2/P3 Composite Layered Cathode for HighPerformance Na-Ion Full Batteries. Nano Energy 2019, 55, 143−150. (16) Monyoncho, E.; Bissessur, R. Unique Properties of α-NaFeO2: De-Intercalation of Sodium Via Hydrolysis and the Intercalation of Guest Molecules into the Extract Solution. Mater. Res. Bull. 2013, 48, 2678−2686. (17) Yao, H.-R.; Wang, P.-F.; Wang, Y.; Yu, X.; Yin, Y.-X.; Guo, Y.G. Excellent Comprehensive Performance of Na-Based Layered Oxide Benefiting from the Synergetic Contributions of Multimetal Ions. Adv. Energy Mater. 2017, 7, 1700189. (18) Lu, Z.; Dahn, J. R. Intercalation of Water in P2, T2 and O2 Structure Az[CoxNi1/3‑xMn2/3]O2. Chem. Mater. 2001, 13, 1252− 1257. (19) Yabuuchi, N.; Ikeuchi, I.; Kubota, K.; Komaba, S. Thermal Stability of NaxCrO2 for Rechargeable Sodium Batteries; Studies by High-Temperature Synchrotron X-ray Diffraction. ACS Appl. Mater. Interfaces 2016, 8, 32292−32299. (20) Guignard, M.; Didier, C.; Darriet, J.; Bordet, P.; Elkaim, E.; Delmas, C. P2-NaxVO2 System as Electrodes for Batteries and Electron-Correlated Materials. Nat. Mater. 2013, 12, 74−80. (21) Wang, P.-F.; Xin, H.; Zuo, T.-T.; Li, Q.; Yang, X.; Yin, Y.-X.; Gao, X.; Yu, X.; Guo, Y.-G. An Abnormal 3.7 Volt O3-Type SodiumIon Battery Cathode. Angew. Chem., Int. Ed. 2018, 57, 8178−8183. (22) Tapia-Ruiz, N.; Dose, W. M.; Sharma, N.; Chen, H.; Heath, J.; Somerville, J. W.; Maitra, U.; Islam, M. S.; Bruce, P. G. High Voltage Structural Evolution and Enhanced Na-Ion Diffusion in P2Na2/3Ni1/3−xMgxMn2/3O2 (0 ≤ x ≤ 0.2) Cathodes from Diffraction, Electrochemical and ab initio Studies. Energy Environ. Sci. 2018, 11, 1470−1479. (23) Liu, Y.; Fang, X.; Zhang, A.; Shen, C.; Liu, Q.; Enaya, H. A.; Zhou, C. Layered P2-Na2/3[Ni1/3Mn2/3]O2 as High-Voltage Cathode for Sodium-Ion Batteries: The Capacity Decay Mechanism and Al2O3 Surface Modification. Nano Energy 2016, 27, 27−34. (24) Wang, Q.-C.; Meng, J.-K.; Yue, X.-Y.; Qiu, Q.-Q.; Song, Y.; Wu, X.-J.; Fu, Z.-W.; Xia, Y.-Y.; Shadike, Z.; Wu, J.; Yang, X.-Q.; Zhou, Y.N. Tuning P2-Structured Cathode Material by Na-Site Mg Substitution for Na-Ion Batteries. J. Am. Chem. Soc. 2019, 141, 840−848. (25) Singh, G.; Tapia-Ruiz, N.; Lopez Del Amo, J. M.; Maitra, U.; Somerville, J. W.; Armstrong, A. R.; Martinez De Ilarduya, J.; Rojo, T.; Bruce, P. G. High Voltage Mg-Doped Na0.67Ni0.3−xMgxMn0.7O2 (x = 24190

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191

Research Article

ACS Applied Materials & Interfaces (43) Yu, C.-Y.; Park, J.-S.; Jung, H.-G.; Chung, K.-Y.; Aurbach, D.; Sun, Y.-K.; Myung, S.-T. NaCrO2 Cathode for High-Rate Sodium-Ion Batteries. Energy Environ. Sci. 2015, 8, 2019−2026. (44) Kim, J.; Seo, D.-H.; Kim, H.; Park, I.; Yoo, J.-K.; Jung, S.-K.; Park, Y.-U.; Goddard, W. A., III; Kang, K. Unexpected Discovery of Low-Cost Maricite NaFePO4 as a High-Performance Electrode for Na-Ion Batteries. Energy Environ. Sci. 2015, 8, 540−545. (45) Hwang, J. Y.; Oh, S. M.; Myung, S. T.; Chung, K. Y.; Belharouak, I.; Sun, Y. K. Radially Aligned Hierarchical Columnar Structure as a Cathode Material for High Energy Density Sodium-Ion Batteries. Nat. Commun. 2015, 6, 6865. (46) Oh, S.-M.; Myung, S.-T.; Hwang, J.-Y.; Scrosati, B.; Amine, K.; Sun, Y.-K. High Capacity O3-Type Na[Li0.05(Ni0.25Fe0.25Mn0.5)0.95]O2 Cathode for Sodium Ion Batteries. Chem. Mater. 2014, 26, 6165− 6171. (47) Ponrouch, A.; Dedryvère, R.; Monti, D.; Demet, A. E.; Ateba Mba, J. M.; Croguennec, L.; Masquelier, C.; Johansson, P.; Palacín, M. R. Towards High Energy Density Sodium Ion Batteries Through Electrolyte Optimization. Energy Environ. Sci. 2013, 6, 2361−2369. (48) Mu, L.; Xu, S.; Li, Y.; Hu, Y. S.; Li, H.; Chen, L.; Huang, X. Prototype Sodium-Ion Batteries Using an Air-Stable and Co/Ni-Free O3-Layered Metal Oxide Cathode. Adv. Mater. 2015, 27, 6928−6933. (49) Zhou, C.; Yang, L.; Zhou, C.; Lu, B.; Liu, J.; Ouyang, L.; Hu, R.; Liu, J.; Zhu, M. Co-Substitution Enhances the Rate Capability and Stabilizes the Cyclic Performance of O3-Type Cathode NaNi0.45‑xMn0.25Ti0.3CoxO2 for Sodium-Ion Storage at High Voltage. ACS Appl. Mater. Interfaces 2019, 11, 7906−7913. (50) Sathiya, M.; Thomas, J.; Batuk, D.; Pimenta, V.; Gopalan, R.; Tarascon, J.-M. Dual Stabilization and Sacrificial Effect of Na2CO3 for Increasing Capacities of Na-Ion Cells Based on P2-Nax MO2 Electrodes. Chem. Mater. 2017, 29, 5948−5956. (51) Li, Q.; Qiao, Y.; Guo, S.; Jiang, K.; Li, Q.; Wu, J.; Zhou, H. Both Cationic and Anionic Co-(de)intercalation into a Metal-Oxide Material. Joule 2018, 2, 1134−1145. (52) Wang, P.-F.; Yao, H.-R.; Liu, X.-Y.; Zhang, J.-N.; Gu, L.; Yu, X.Q.; Yin, Y.-X.; Guo, Y.-G. Ti-Substituted NaNi0.5Mn0.5‑xTixO2 Cathodes with Reversible O3-P3 Phase Transition for HighPerformance Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1700210. (53) Kalapsazova, M.; Ortiz, G. F.; Tirado, J. L.; Dolotko, O.; Zhecheva, E.; Nihtianova, D.; Mihaylov, L.; Stoyanova, R. P3-Type Layered Sodium-Deficient Nickel-Manganese Oxides: A Flexible Structural Matrix for Reversible Sodium and Lithium Intercalation. ChemPlusChem 2015, 80, 1642−1656. (54) Jiang, K.; Xu, S.; Guo, S.; Zhang, X.; Zhang, X.; Qiao, Y.; Fang, T.; Wang, P.; He, P.; Zhou, H. A Phase-Transition-Free Cathode for Sodium-Ion Batteries with Ultralong Cycle Life. Nano Energy 2018, 52, 88−94.

24191

DOI: 10.1021/acsami.9b07299 ACS Appl. Mater. Interfaces 2019, 11, 24184−24191