Superior-Performance Aqueous Zinc Ion Battery Based on Structural

10 hours ago - https://doi.org/10.1021/acs.jpcc.9b05535. Copyright © 2019 American Chemical Society. RIGHTS & PERMISSIONS. Subscribed Access ...
0 downloads 0 Views 947KB Size
Subscriber access provided by TRINITY COLL

C: Energy Conversion and Storage; Energy and Charge Transport

Superior-Performance Aqueous Zinc Ion Battery Based on Structural Transformation of MnO by Rare Earth Doping 2

Jianwei Wang, Xiaolei Sun, Hongyang Zhao, Lingling Xu, Jiale Xia, Meng Luo, Yaodong Yang, and Yaping Du J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.9b05535 • Publication Date (Web): 23 Aug 2019 Downloaded from pubs.acs.org on August 27, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Superior-Performance Aqueous Zinc Ion Battery Based on Structural Transformation of MnO2 by Rare Earth Doping Jianwei Wang,a Xiaolei Sun,b Hongyang Zhao,a Lingling Xu,a Jiale Xia,a Meng Luo,a Yaodong Yang,a and Yaping Du*b a

Frontier Institute of Science and Technology, Xi’an Jiaotong University, Xi’an, Shaanxi

710054, China. b

School of Materials Science and Engineering & National Institute for Advanced Materials, Key

Laboratory of Advanced Energy Materials Chemistry, Tianjin Key Lab for Rare Earth Materials and Applications, Centre for Rare Earth and Inorganic Functional Materials, Nankai University, Tianjin, 300350, China. Corresponding Author *E-mail: [email protected]

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 20

ABSTRACT: Doping with heteroatoms is being used as an effective way to change electronic structure of electrode materials for advanced storage systems. Herein, β-MnO2 and rare earth (cerium) doped MnO2 cathode materials have been successfully prepared for aqueous zinc ion batteries. Cerium doping induced structural transformation of MnO2 from β- to α-phase, along with the evident improvement of conductivity, stability and reversibility. Compared with the undoped β-MnO2, the doping cathode possessed excellent cycling stability at a high rate of 5 C and higher rate capability. CV curves, EIS, GITT, SEM and TEM measurements demonstrated that cerium doping caused fast Zn2+ diffusion and excellent electrochemical stability.

ACS Paragon Plus Environment

2

Page 3 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

INTRODUCTION Rechargeable batteries have been extensively used in mobile digital devices, power tools, electric vehicles, and stationary energy storage systems. Compared with other rechargeable batteries, lithium-ion batteries dominate the field of energy storage systems on the market.1-5 Despite impressive achievements, lithium-ion batteries have suffered from high cost and safety concerns for large-scale applications.6,7 In this regard, aqueous rechargeable batteries have received tremendous attentions.8 Compared with organic electrolytes (1-10 mS cm-1), aqueous electrolytes offer much higher ionic conductivity (1 S cm-1).9,10 Importantly, the aqueous batteries are safe, cheap, and easy to scale up.11 Among rechargeable multivalent ion batteries such as calcium ion batteries, magnesium ion batteries, aluminum ion batteries and zinc ion batteries, aqueous zinc ion batteries (ZIBs) are regarded as the most promising alternatives for grid-scale energy storage in the near future.12-14 Zinc (Zn) has the advantages of low cost ($ 2 Kg-1) and high natural abundance globally (75 ppm in earth’s crust). The electrolyte without any organic solvents would be used to improve safety and decrease the environmental impact. In addition, zinc metal possesses good theoretical gravimetric capacity (820 mAh g-1), high volumetric capacity (5855 mAh cm-3) and relatively low redox potential (-0.76 V vs SHE). Although ZIBs are very attractive, their development is still in infancy and hampered by the limited selection of cathode materials. Therefore, it is a big challenge to find suitable cathode materials, which could not only offer superior and stable capacity but also retain structural stability during zinc ion insertion/extraction process. Manganese dioxide (MnO2) has been one of the most widely used cathode materials in ZIBs owing to its high theoretical capacity (308 mAh g-1), the low cost and less toxicity.15-17 But weak ionic and electronic conductivity, low ion diffusion coefficient and unstable structure result in a

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 20

rapid decline during cycling. Various forms of MnO2 polymorphs include α-MnO2, β-MnO2, γMnO2, δ-MnO2, λ-MnO2 and todorokite. Among them, the most stable phase is β-MnO2 with narrow [11] tunnel structure. However, only limited research on the β-MnO2 polymorph can be found as cathode materials for ZIBs. Jaekook Kim’s groups have demonstrated that β-MnO2 cathode exhibited a sharp initial capacity decrease and poor electrochemical performance at high rate conditions,18 due to low conductivity and poor stability upon cycling. In the field of supercapacitors, many studies have also proved that doping with trace heteroatoms (Cr, Co, Ni, Ce, Cu, Au, etc.) can enhance intrinsic conductivity and improve capacitive performance of MnO2.19-22 More importantly, theoretical studies indicated that β-MnO2 could accommodate guest cations to improve the electronic and ionic transport, thus change the electrochemical activity.23,24 Therefore, doping would be an effective strategy to heighten electrochemical performance for ZIBs by anisotropically changing the lattice parameters. Compared with transition metal ions, rare earth ions possess unique 4f electron configuration, which will greatly affect the structure and electrochemical properties of electrode materials. To the best of our knowledge, the application of MnO2 with doping cerium (Ce) in aqueous zinc ion batteries has not been reported yet. Herein, Ce doped MnO2 nanorod-like electrode material was synthesized by one-step hydrothermal method. Cerium doping induced structural transformation of MnO2 from original β- to novel α-phase, along with the appearance of [22] tunnel structure. After doping Ce, the electrode material showed excellent cycling stability at a high rate of 5 C (1C = 308 mA g-1) and higher rate capability compared with un-doped β-MnO2. The electrode was further investigated systematically through electrochemical impedance spectroscopy (EIS), continuous cyclic voltammetry (CV) curves, galvanostatic intermittent titration technique (GITT) measurements

ACS Paragon Plus Environment

4

Page 5 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

and so forth. The results reveal that Ce doped MnO2 cathode has outstanding electrochemical stability and faster Zn2+ diffusion than un-doped β-MnO2. RESULTS AND DISCUSSION

Figure 1. (a) X-ray diffraction patterns of β-MnO2, 0.1-0.3 mmol Ce doped MnO2; (b) Rietveld refined XRD pattern of 0.1 mmol Ce doped MnO2. Experimental data, calculated results, allowed Bragg diffraction positions and difference are marked with black crosses, red line, vertical bars and baby blue curve, respectively; (c) crystallographic structure. Characterization of Electrode Materials. Different molar Ce doped MnO2 were prepared via a facile hydrothermal method and the detailed procedures were provided in supporting information section. The initial molar cerium doped MnO2 was marked as 0.1 mmol Ce doping, 0.2 mmol Ce doping and 0.3 mmol Ce doping, respectively. Figure 1a showed XRD patterns of MnO2 with various amounts of cerium. Without doping Ce, the sample with MnSO4 as precursor was clarified to be a pure β-MnO2 phase. All of diffraction peaks could be perfectly assigned to

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 20

β-MnO2 (PDF#01-0799) with a space group of P42/mnm (no. 136). While all the doping products with MnSO4 as precursor were identified to be a pure α-MnO2 phase, whose characteristic peaks were well-indexed to α-MnO2 (PDF#44-0141) with a space group of I4/m (no. 87). Compared to the corresponding standard values, no impurity phases were showed, which demonstrated that structural transformation was due to cerium doping into crystal lattice of MnO2. In order to obtain the crystal parameters of doping, the powder XRD data of 0.1 mmol Ce doping were refined by GASA software and its crystal structure was obtained and investigated by Rietveld method.25 The refined profiles matched well with the experimental data (Figure 1b) and indicated the formation of α-MnO2 phase, further confirming that doping Ce results in the transformation of phase. Figure 1c revealed that β-MnO2 has a [11] tunnel structure with edge and corner-sharing octahedral [MnO6] units along its c-axis and the size of tunnel was 2.3  2.3 Å2. While 0.1 mmol Ce doping possessed [11] and [22] tunnel structure, the different sizes of tunnel were calculated to be 2.9 x 2.9 Å2 [11] and 4.6  4.6 Å2 [22]. For the standard α-MnO2 (PDF#44-0141), the sizes of the tunnel were 2.7  2.7 Å2 [11] and 4.6  4.6 Å2 [22],26,27 respectively. The results suggested that the doping of cerium causes [11] tunnel to become larger and [22] tunnel to remain nearly unchanged. In order to further prove that cerium doping causes phase transition and cerium element exists in the novel phase simultaneously, pure α-MnO2 was synthesized and X-ray photoelectron spectroscopy (XPS) analysis was carried out (Figure S1). The wide energy spectra showed that α-MnO2 contained Mn and O elements, while besides Mn and O elements, 0.1 mmol Ce doping also contained Ce element. For the pure α-MnO2, the Mn 2p spectrum showed two major peaks centered at 640.5 eV for Mn 2p3/2 and 652.3 eV for Mn 2p1/2. The O 1s spectrum showed one single peak centered at 528.1 eV. However, three major peaks (Mn 2p3/2, Mn 2p1/2 and O 1s) for

ACS Paragon Plus Environment

6

Page 7 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Ce doped MnO2 shifted to higher energy compared to those of pure α-MnO2, confirming the strong chemical interaction between Ce and MnO2. Additionally, a peak at 902.4 eV is clearly observed, which is the characteristic peak of Ce(III).28,29 As shown in Figure S2 and Table S1, the surface average oxidation state (AOS) of manganese is estimated for three samples respectively. The AOS increases in the order of α-MnO2 (3.53)﹤0.1 mmol Ce doping (3.75)﹤βMnO2 (3.98). Compared with undoped β-MnO2, the lower AOS of 0.1 mmol Ce doping is attributed to the phase transition of MnO2. However, in the same phase state (α-phase), cerium doping significantly improves the AOS of MnO2.

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 20

Figure 2. Characterization of 0.1 mmol Ce doping. (a) TEM image; (b, c) HRTEM images; (d) HAADF-TEM image; (e) EDS elemental mappings. The morphologies of different molar Ce doped MnO2 were characterized with transmission electron microscope (TEM) and scanning electron microscope (SEM). As shown in Figure S3, βMnO2 without Ce doping was made up of nanorods and agglomerates, which displayed different sizes. After cerium was doped into MnO2, all doping samples exhibited uniform nanorod-like morphology. SEM images (Figure S4) further revealed β-MnO2 possessed heterogeneous morphology and all the doping samples had uniform nanorod-like structure. The results of TEM and SEM images showed that the length of nanorods gradually decreased with the increase of cerium content. It can be noted that the amount of Ce doping significantly affects the morphology and size of final materials. To determine the amount of Ce in doped samples, TEMenergy dispersive X-ray spectroscopy (EDS) was employed to detect. The EDS spectra are shown in Figure S5 and contents of different elements are provided in Table S2. Crystal structure of 0.1 mmol Ce doping was analysed by high-resolution TEM (HRTEM), high-angle annular dark-field (HAADF) TEM and EDS elemental mappings. As shown in Figure 2a, the average diameter and length of nanorods was 30 and 300 nm, respectively. HRTEM images (Figure 2b and 2c) showed the lattice fringe with d space of 0.76 and 0.50 nm, corresponding to two tetragonal crystal planes (110) and (200), respectively. Meanwhile, HAADF image was also recorded (Figure 2d). In order to observe the distribution of cerium in MnO2, EDS elemental mappings was conducted, as displayed in Figure 2e. The results confirmed homogeneous distribution of Mn, O and Ce components in 0.1 mmol Ce doping. The intensity of Ce signal was obviously weaker than that of Mn, which was highly consistent with the amount of doping cerium.

ACS Paragon Plus Environment

8

Page 9 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Electrochemical Performance. The electrochemical properties were investigated using zinc as anode, 2 M ZnSO4 and 0.1 M MnSO4 as electrolyte, glass fibre as separator and prepared materials as cathode. The cyclic voltammetry (CV) curves of β-MnO2 and 0.1 mmol Ce doping electrodes within the potential range of 1.0-1.8 V at a scan rate of 0.1 mV s-1 were given in Figure 3a. β-MnO2 electrode revealed two cathodic peaks at around 1.32 and 1.13 V, corresponding to Zn-ion insertion into the host material and the continuous reduction reaction of Mn (Ⅳ).30 Upon Zn-ion release, a wide peak at around 1.64 V was observed, which indicated that Mn(Ⅲ)/Mn(Ⅱ) states were oxidized to Mn(Ⅳ) state.31 And then, 0.1 mmol Ce doping electrode displayed similar CV profiles. Alternatively, the cathodic peaks obviously shifted to

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 20

Figure 3. (a) CV curves at a scan rate of 0.1 mV s-1; (b) Charge-discharge voltage profiles at 0.5 C; (c) Cycling performances at 5 C; (d) Rate capability at various current densities. higher potentials (1.38 and 1.23 V) and the anodic peak clearly shifted to lower potential (1.57 V) and became narrower simultaneously. The voltage differences of 0.1 mmol Ce doping between the center of anodic and cathodic peaks were significantly smaller than those of βMnO2, suggesting a better reversibility of insertion and extraction of zinc ions in MnO2 electrode with doping cerium. Figure 3b exhibited the galvanostatic charge and discharge profiles of two electrode materials with multiple redox peaks at the current density of 0.5 C (1C = 308 mA g-1), which was in excellent agreement with CV curves respectively. Differential capacity (dQ/dV) plots were calculated from charge and discharge curves of three samples at 0.5 C (Figure S6). Compared with the voltage gap of 0.41 V for β-MnO2, 0.1 mmol Ce doping displayed a significant decrease to 0.27 V, indicating that severe polarization was effectively alleviated and the phase transformation of doping Ce caused outstanding redox kinetics. The presence of two plateaus during the discharge process was indicative of different embedding mechanisms.8 To further investigate the zinc insertion processes, three samples were analyzed at a high rate of 5 C for 100 cycles with the same parameters. As shown in Figure 3c and Figure S7, the cycling stability of aqueous ZIB was improved significantly by doping Ce in MnO2. Compared with the cycling performance of different molar Ce doped MnO2 cathodes, 0.1 mmol Ce doping was chosen to study the effects of doping in cells. The initial discharge capacity and Coloumbic efficiency of β-MnO2 were 128 mAh g-1 and 70%, respectively. Compared with β-MnO2, 0.1 mmol Ce doping displayed higher initial capacity and Coloumbic efficiency of 134 mAh g-1 and 82%. More importantly, after 100 cycles, the capacity retention of 0.1 mmol Ce doping was almost twice higher than that of β-MnO2 at such high current density. However, compared with

ACS Paragon Plus Environment

10

Page 11 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

α-MnO2, the capacity retention of 0.1 mmol Ce doping has been slightly improved, which proves that tunnel dimensions affect the diffusion numbers. The long cycling performance of 0.1 mmol Ce doping was measured at 2 C for 400 cycles (Figure S8). The capacity decay rate was only 0.065% per cycle, indicating excelllent long-term cycling stability at the high current density. Furthermore, it can be observed from Figure 3d that the two electrodes offered similar discharge capacity at 0.5 C. When the rate was located in the range of 1-5 C, 0.1 mmol Ce doping electrode had high discharge capacities, indicating the better rate capability. All above electrochemical results demonstrated that the doping of Ce in MnO2 might enhance the structural stability of electrode material and be sufficient to mitigate manganese dissolution. Electrochemical Kinetics. To understand the electrochemical kinetics of 0.1 mmol Ce doping and β-MnO2 electrode, CV measurements at different scan rates were conducted (Figure 4a and Figure S9). With the increase of scan rate, the CV curves of β-MnO2 electrode changed dramatically, while the CV curves of 0.1 mmol Ce doping electrode still maintained their shapes, indicating the exceptional stability of the electrode material. Capacitive processes and diffusioncontrolled redox reaction could be quantitatively separated by a simple method:32,33 i = aνb, where i represents current, a and b are adjustable parameters, ν represents scan rate. Thus, log(i) = log(a) + b·log(ν), when the b value is 0.5, it indicates that the capacity is totally controlled by semi-infinite diffusion. While b value is 1, it indicates a complete capacitive behavior. Furthermore, the b values of three redox peaks were determined to be 0.65, 0.90 and 0.81 respectively (Figure 4b), suggesting the process was partially controlled by capacitive behavior.9 As is known, at a certain scan rate, the current is divided into capacitor-like and diffusioncontrolled processes,34 as follow: i = k1ν+ k2ν1/2, in this equation, k1ν and k2ν1/2 refer to capacitive process and diffusion-controlled process, respectively. As shown in Figure 4c, 41.62%

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 20

fraction of total capacity came from capacitive process at the scan rate of 0.1 mV s-1, leading to high-rate capability of 0.1 mmol Ce doping electrode. In order to further investigate the effect of

Figure 4. Characterizations of the 0.1 mmol Ce doping cathode. (a) CV curves at different scan rates; (b) log(current) versus log(scan rate) plots of three peaks; (c) capacitive fraction (shown by the shaded area) calculated at a scan rate of 0.1 mV s-1; (d) EIS Nyquist plots and an equivalent circuit; (e, f) GITT curves and diffusivity versus state of discharge. doping cerium on the charge transport and ion diffusion kinetics, EIS measurements were conducted by using a symmetric cell with two identical electrodes for a state of charge (SOC) at 0%, which can provide more accurate impedance spectra.35,36 From Nyquist plots and equivalent circuit (Figure 4d), the charge-transfer resistance (Table S3) of 0.1 mmol Ce doping (Rct = 43.82 Ω) was much lower than that of the β-MnO2 (Rct = 53.46 Ω), revealing that doping cerium accelerated charge-transfer process. In the low frequency region, Z' values as a function of ω-1/2 were shown in Figure S10a and the slope was correlated with solid state ion diffusivity.37 The

ACS Paragon Plus Environment

12

Page 13 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

slope of 0.1 mmol Ce doping (6.08) was smaller than that of the β-MnO2 (7.36), indicating a better diffusion behavior of the doping electrode. Bode plots of two samples were also used to estimate the Zn2+ diffusion kinetics in electrode materials. Relevant reports have demonstrated that the ion diffusion was related to the phase angle in the low-frequency region and the smaller the phase angle, the faster the ion diffusion.38-40 The cerium doped electrode showed a smaller phase angle (Figure S10b), suggesting a faster ion diffusion process. Additionally, galvanostatic intermittent titration technique (GITT) was used to calculate chemical diffusion coefficient of Zn2+ in two electrode materials. The Zn2+ diffusivity (DGITT) was obtained (Figure 4e and 4f) and the detailed process was provided in the supporting information section. Compared to β-MnO2 electrode, Zn2+ diffusion coefficient could be effectively increased by doping Ce ions into the MnO2 structure. The diffusion coefficient of 0.1 mmol Ce doping cathode during the entire Zn intercalation process was determined to be 10-10 to 10-13 cm2 s-1, which was about 10-102 higher than the β-MnO2 cathode. Obviously, doping led to a boost in Zn2+ diffusivity, which was in accordance with EIS studies. The results clearly showed that larger [11] and [22] tunnel structure was conducive to fast and reversible Zn2+ migration, thus resulting in high-rate capability. Structural Stability. In order to verify the structural stability of 0.1 mmol Ce doping electrode material, SEM, TEM and corresponding EDS characterizations were carried out after 20 cycles at the current density of 1 C. The SEM and TEM images (Figure 5a and 5b) showed that the nanorod-like morphology was kept without a noticeable change, even after 20 cycles. And then, the TEM-EDS elemental mappings as seen in Figure 5c and 5d exhibited elemental cerium still uniformly retained in MnO2. Therefore, the results clearly indicated that cerium

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 20

doped cathode material still maintained high stability after insertion/extraction of Zn2+ many times.

Figure 5. 0.1 mmol Ce doping cathode after 20 cycles at 1 C. (a) SEM image; (b) TEM image; (c-d) TEM-EDS elemental mappings. CONCLUSIONS In summary, the MnO2 electrode material with rare earth (Ce) doping was successfully synthesized by one-step hydrothermal method and used as cathode for aqueous zinc ion battery for the first time. Cerium doping not only adjusted the surface morphology but also induced structural transformation of MnO2. After doping Ce, the electrode could deliver a high capacity of 134 mAh g-1 at 5 C along with outstanding cycling stability and rate capability. Moreover, we also demonstrated that cerium doped cathode material possessed excellent electrochemical stability and reversibility, low charge-transfer resistance and high ion diffusivity. Doping rare

ACS Paragon Plus Environment

14

Page 15 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

earth ions in electrode materials would be an effective strategy to improve the electrochemical performance in the fields of energy storage and conversion. ASSOCIATED CONTENT Supporting Information Synthetic procedures, materials and methods, XPS spectra, TEM and SEM images and electrochemical performance. AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS We gratefully acknowledge the support from the China National Funds for Excellent Young Scientists (21522106), the National Key R&D Program of China (2017YFA0208000). REFERENCES (1)

Schon, T. B.; McAllister, B. T.; Li, P. F.; Seferos, D. S. The Rise of Organic Electrode Materials for Energy Storage. Chem. Soc. Rev. 2016, 45, 6345-6404.

(2)

Feng, X.; Ouyang, M.; Liu, X.; Lu, L.; Xia, Y.; He, X. Thermal Runaway Mechanism of Lithium Ion Battery for Electric Vehicles: A Review. Energy Storage Mater. 2018, 10, 246-267.

(3)

Hannan, M. A.; Lipu, M. S. H.; Hussain, A.; Mohamed, A. A Review of Lithium-ion Battery State of Charge Estimation and Management System in Electric Vehicle

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 20

Applications: Challenges and Recommendations. Renew. Sust. Energ. Rev. 2017, 78, 834854. (4)

Wang, J.; Zhao, H.; Xu, L.; Yang, Y.; He, G.; Du, Y. Three-Electron Redox Enabled Dithiocarboxylate Electrode for Superior Lithium Storage Performance. ACS Appl. Mater. Interfaces 2018, 10, 35469-35476.

(5)

Xu, J. J.; Wang, Z. L.; Xu, D.; Meng, F. Z.; Zhang, X. B. 3D Ordered Macroporous LaFeO3 as Efficient Electrocatalyst for Li-O2 Batteries with Enhanced Rate Capability and Cyclic Performance. Energy Environ. Sci. 2014, 7, 2213-2219.

(6)

Lu, L.; Han, X.; Li, J.; Hua, J.; Ouyang, M. A Review on the Key Issues for Lithium-ion Battery Management in Electric Vehicles. J. Power Sources 2013, 226, 272-288.

(7)

Nayak, P. K.; Erickson, E. M.; Schipper, F.; Penki, T. R.; Munichandraiah, N.; Adelhelm, P.; Sclar, H.; Amalraj, F.; Markovsky, B.; Aurbach, D. Review on Challenges and Recent Advances in the Electrochemical Performance of High Capacity Li- and Mn-Rich Cathode Materials for Li-Ion Batteries. Adv. Energy Mater. 2018, 8, 1702397.

(8)

Pan, H.; Shao, Y.; Yan, P.; Cheng, Y.; Han, K. S.; Nie, Z.; Wang, C.; Yang, J.; Li, X.; Bhattacharya, P.; Mueller, K. T.; Liu, J. Reversible Aqueous Zinc/Manganese Oxide Energy Storage from Conversion Reactions. Nat. Energy 2016, 1, 16039.

(9)

Xia, C.; Guo, J.; Lei, Y.; Liang, H.; Zhao, C.; Alshareef, H. N. Rechargeable Aqueous Zinc-Ion Battery Based on Porous Framework Zinc Pyrovanadate Intercalation Cathode. Adv. Mater. 2018, 30, 1705580.

(10) Kundu, D.; Adams, B. D.; Duffort, V.; Vajargah, S. H.; Nazar, L. F. A High-capacity and Long-life Aqueous Rechargeable Zinc Battery Using a Metal Oxide Intercalation Cathode. Nat. Energy 2016, 1, 16119. (11) Yesibolati, N.; Umirov, N.; Koishybay, A.; Omarova, M.; Kurmanbayeva, I.; Zhang, Y.; Zhao, Y.; Bakenov, Z. High Performance Zn/LiFePO4 Aqueous Rechargeable Battery for Large Scale Applications. Electrochim. Acta 2015, 152, 505-511. (12) Lee, B.; Yoon, C. S.; Lee, H. R.; Chung, K. Y.; Cho, B. W.; Oh, S. H. Electrochemicallyinduced Reversible Transition from the Tunneled to Layered Polymorphs of Manganese Dioxide. Sci. Rep. 2014, 4, 6066. (13) Fang, G.; Zhou, J.; Pan, A.; Liang, S. Recent Advances in Aqueous Zinc-Ion Batteries. ACS Energy Lett. 2018, 3, 2480-2501.

ACS Paragon Plus Environment

16

Page 17 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(14) Wang, F.; Hu, E.; Sun, W.; Gao, T.; Ji, X.; Fan, X.; Han, F.; Yang, X. Q.; Xu, K.; Wang, C. Rechargeable Aqueous Zn2+-Battery with High Power Density and Long Cycle-life. Energy Environ. Sci. 2018, 11, 3168-3175. (15) Song, M.; Tan, H.; Chao, D.; Fan, H. J. Recent Advances in Zn-Ion Batteries. Adv. Funct. Mater. 2018, 28, 1802564. (16) Zhang, K.; Han, X.; Hu, Z.; Zhang, X.; Tao, Z.; Chen, J. Nanostructured Mn-based Oxides for Electrochemical Energy Storage and Conversion. Chem. Soc. Rev. 2015, 44, 699-728. (17) Wu, B.; Zhang, G.; Yan, M.; Xiong, T.; He, P.; He, L.; Xu, X.; Mai, L. Graphene ScrollCoated α-MnO2 Nanowires as High-Performance Cathode Materials for Aqueous Zn-Ion Battery. Small 2018, 14, 1703850. (18) Islam, S.; Alfaruqi, M. H.; Mathew, V.; Song, J.; Kim, S.; Kim, S.; Jo, J.; Baboo, J. P.; Pham, D. T.; Putro, D. Y.; Sun, Y. K.; Kim, J. Facile Synthesis and the Exploration of the Zinc Storage Mechanism of β-MnO2 Nanorods with Exposed (101) Planes as a Novel Cathode Material for High Performance Eco-friendly Zinc-ion Batteries. J. Mater. Chem. A 2017, 5, 23299-23309. (19) Kang, J.; Hirata, A.; Kang, L.; Zhang, X.; Hou, Y.; Chen, L.; Li, C.; Fujita, T.; Akagi, K.; Chen, M. Enhanced Supercapacitor Performance of MnO2 by Atomic Doping. Angew. Chem. Int. Ed. 2013, 52, 1664-1667. (20) Chen, K.; Pan, W.; Xue, D. Phase Transformation of Ce3+-Doped MnO2 for Pseudocapacitive Electrode Materials. J. Phys. Chem. C 2016, 120, 20077-20081. (21) Wang, G.; Ma, Z.; Zhang, G.; Li, C.; Shao, G. Cerium-doped Porous K-birnessite Manganese Oxides Microspheres as Pseudocapacitor Electrode Material with Improved Electrochemical Capacitance. Electrochim. Acta 2015, 182, 1070-1077. (22) Dong, R.; Ye, Q.; Kuang, L.; Lu, X.; Zhang, Y.; Zhang, X.; Tan, G.; Wen, Y.; Wang, F. Enhanced Supercapacitor Performance of Mn3O4 Nanocrystals by Doping Transition-metal Ions. ACS Appl. Mater. Interfaces 2013, 5, 9508-9516. (23) Wang, D.; Liu, L. M.; Zhao, S. J.; Li, B. H.; Liu, H.; Lang, X. F. β-MnO2 as a Cathode Material for Lithium Ion Batteries from First Principles Calculations. Phys. Chem. Chem. Phys. 2013, 15, 9075-9083. (24) Fang, C. M.; de Wijs, G. A. Local Structure and Chemical Bonding of Protonated LixMn2O4 Spinels from First Principles. Chem. Mater. 2006, 18, 1169-1173.

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 20

(25) Toby, B. H. EXPGUI, a Graphical User Interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210-213. (26) Konarov, A.; Voronina, N.; Jo, J. H.; Bakenov, Z.; Sun, Y. K.; Myung, S. T. Present and Future Perspective on Electrode Materials for Rechargeable Zinc-Ion Batteries. ACS Energy Lett. 2018, 3, 2620-2640. (27) Yuan, Y.; Liu, C.; Byles, B. W.; Yao, W.; Song, B.; Cheng, M.; Huang, Z.; Amine, K.; Pomerantseva, E.; Shahbazian-Yassar, R.; Lu, J. Ordering Heterogeneity of [MnO6] Octahedra in Tunnel-Structured MnO2 and Its Influence on Ion Storage. Joule 2018, 3, 114. (28) Putla, S.; Amin, M. H.; Reddy, B. M.; Nafady, A.; Al Farhan, K. A.; Bhargava, S. K. MnOx Nanoparticle-Dispersed CeO2 Nanocubes: A Remarkable Heteronanostructured System with Unusual Structural Characteristics and Superior Catalytic Performance. ACS Appl. Mater. Interfaces 2015, 7, 16525-16535. (29) Wu, K.; Zhou, L.; Jia, C. J.; Sun, L. D.; Yan, C. H. Pt-Embedded-CeO2 Hollow Spheres for Enhancing CO Oxidation Performance. Mater. Chem. Front. 2017, 1, 1754-1763. (30) Alfaruqi, M. H.; Gim, J.; Kim, S.; Song, J.; Jo, J.; Kim, S.; Mathew, V.; Kim, J. Enhanced Reversible Divalent Zinc Storage in a Structurally Stable α-MnO2 Nanorod Electrode. J. Power Sources 2015, 288, 320-327. (31) Alfaruqi, M. H.; Mathew, V.; Gim, J.; Kim, S.; Song, J.; Baboo, J. P.; Choi, S. H.; Kim, J. Electrochemically Induced Structural Transformation in a γ-MnO2 Cathode of a High Capacity Zinc-Ion Battery System. Chem. Mater. 2015, 27, 3609-3620. (32) Wang, J.; Polleux, J.; Lim, J.; Dunn, B. Pseudocapacitive Contributions to Electrochemical Energy Storage in TiO2 (Anatase) Nanoparticles. J. Phys. Chem. C 2007, 111, 1492514931. (33) Simon, P.; Gogotsi, Y.; Dunn, B. Where Do Batteries End and Supercapacitors Begin? Science 2014, 343, 1210-1211. (34) Ming, F.; Liang, H.; Lei, Y.; Kandambeth, S.; Eddaoudi, M.; Alshareef, H. N. Layered MgxV2O5·nH2O as Cathode Material for High-Performance Aqueous Zinc Ion Batteries. ACS Energy Lett. 2018, 3, 2602-2609. (35) Sun, H.; Mei, L.; Liang, J.; Zhao, Z.; Lee, C.; Fei, H.; Ding, M.; Lau, J.; Li, M.; Wang, C.; Xu, X.; Hao, G.; Papandrea, B.; Shakir, I.; Dunn, B.; Huang, Y.; Duan, X. Three-

ACS Paragon Plus Environment

18

Page 19 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

dimensional Holey-graphene/niobia Composite Architectures for Ultrahigh-rate Energy Storage. Science 2017, 356, 599-604. (36) Ogihara, N.; Itou, Y.; Sasaki, T.; Takeuchi, Y. Impedance Spectroscopy Characterization of Porous Electrodes under Different Electrode Thickness Using a Symmetric Cell for High-Performance Lithium-Ion Batteries. J. Phys. Chem. C 2015, 119, 4612-4619. (37) Zhou, X.; Wu, G.; Wu, J.; Yang, H.; Wang, J.; Gao, G. Carbon Black Anchored Vanadium Oxide Nanobelts and Their Post-sintering Counterpart (V2O5 nanobelts) as High Performance Cathode Materials for Lithium Ion Batteries. Phys. Chem. Chem. Phys. 2014, 16, 3973-3982. (38) 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. (39) Wang, D. W.; Li, F.; Fang, H. T.; Liu, M.; Lu, G. Q.; Cheng, H. M. Effect of Pore Packing Defects in 2-D Ordered Mesoporous Carbons on Ionic Transport. J. Phys. Chem. B 2006, 110, 8570-8575. (40) Yu, M.; Zeng, Y.; Han, Y.; Cheng, X.; Zhao, W.; Liang, C.; Tong, Y.; Tang, H.; Lu, X. Valence-Optimized Vanadium Oxide Supercapacitor Electrodes Exhibit Ultrahigh Capacitance and Super-Long Cyclic Durability of 100 000 Cycles. Adv. Funct. Mater. 2015, 25, 3534-3540.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 20

TOC Graphic

ACS Paragon Plus Environment

20