Simin
Dai‡
a,
Xinyan
Zhuang‡
a,
Hongrun
Jin‡
a,
Ruixuan
Yang
b,
Yan
Wang
a,
Bei
Qi
a,
Wenhuan
Guo
a,
Kefeng
Xie
c,
Zhimi
Hu
a,
Meilin
Liu
d and
Liang
Huang
*a
aWuhan National Laboratory for Optoelectronics, School of Optical and Electronic Information, Huazhong University of Science and Technology, Wuhan, 430074, China. E-mail: huangliang421@hust.edu.cn
bSchool of Materials Science and Engineering, Xi'an University of Technology, Xi'an 710048, China
cSchool of Chemistry and Chemical Engineering, Lanzhou Jiaotong University, Lanzhou, 730000, China
dSchool of Materials Science and Engineering, Georgia Institute of Technology, 771 Ferst Drive, Atlanta, Georgia 30332-0245, USA
First published on 10th October 2024
Unfavorable proton intercalation leading to the generation and shedding of side reaction products is still a major challenge for the performance of manganese-based aqueous zinc-ion batteries (AZIBs). In this study, we present a porous oxygen-deficient MnO2 (Od-MnO2) synthesized through n-butyllithium reduction treatment to induce preferential Zn2+ intercalation, thereby effectively mitigating the adverse consequences of proton intercalation for high-performance AZIBs. Remarkably, Od-MnO2 as a cathode material for AZIBs exhibits a specific capacity of 341 mA h g−1 at 0.1 A g−1 and 139 mA h g−1 at 5 A g−1, and outstanding long-term stability with a capacity retention of 85.4% for over 1200 cycles at 1 A g−1. Moreover, the Zn/Od-MnO2 pouch cell displays decent durability with a capacity retention of ∼90% for over 200 cycles at 1C. Our study opens new opportunities for the rational design of high-performance cathode materials by regulating the electronic structure and optimizing the energy storage process for rechargeable AZIBs.
To date, manganese-based materials, vanadium-based materials, and Prussian blue analogues stand as the most widely studied cathode materials for ZIBs.19,23–26 Out of these, MnO2 stands out for its high theoretical capacity (616 mA h g−1, contributed capacity of two electron transfer), high operating voltage, polymorphic structures, low cost, and eco-friendliness.27–29 Presently, there are four main reaction mechanisms in Zn-MnO2 batteries: (1) Zn2+ intercalation, (2) H+/Zn2+ co-intercalation, (3) chemical conversion reaction, and (4) deposition–dissolution mechanism.30–34 Among these mechanisms, H+/Zn2+ co-intercalation is the most typical. From the kinetic point of view, proton intercalation is easier than that of Zn2+, but the variation of H+ concentration leads to the formation and disappearance of side reaction products such as Zn4SO4(OH)6·nH2O (ZSH) or Znx(OTf)y(OH)2x−y·nH2O (ZOTH) on the surface of the cathode. This phenomenon is prone to electrolyte depletion and increased interfacial impedance, and in practice, the shedding of ZSH or ZOTH decreases the reversibility of the electrochemical behavior, resulting in faster capacity degradation and a shorter cycle life (Fig. 1a).30,35–39 It is generally believed that increasing the proportion of Zn2+ intercalation and decreasing the proportion of H+ intercalation can effectively reduce the accumulation of by-products and thus enhance the electrode's cycling stability. It is well known that oxygen defects can serve as shallow donors to regulate the electronic structure of metal oxides (especially in charge density distribution, metal valence state, and energy band structure). The optimized charge density distribution is beneficial for adjusting the interaction between metal oxides and transport ions. Furthermore, a change in the electronic structure promotes a deeper degree of electron delocalization, thereby boosting electron activity and enhancing material conductivity.40–42 Up to now, however, there is rare study to discuss the modulation of defect sites on ion selective intercalation, which will be critical for the rational. design of electrode materials.
In this work, we found that oxygen-deficient sites have stronger interactions with Zn2+ and may induce the preferential intercalation of Zn2+. In this way, an increase in the proportion of Zn2+ intercalation corresponds to a decrease in the proportion of H+ intercalation, which inhibits the generation and shedding of by-products and the accompanying cathode dissolution problems, and thus increases the reversibility of the electrochemical behavior (Fig. 1b). Then, δ-MnO2 with abundant oxygen defects (named Od-MnO2) was rationally designed and synthesized by the molten salt method followed by n-butyllithium treatment. The Od-MnO2 based cathode for AZIBs exhibits an outstanding rate capability (341 mA h g−1 at 0.1 A g−1 and 94 mA h g−1 at 10 A g−1) and remarkable cycling stability with a capacity of 275 mA h g−1 after 200 cycles at a low current density of 0.2 A g−1 (94.4% capacity retention), and retains 190.8 mA h g−1 even after 1200 cycles at 1 A g−1 (85.4% capacity retention).
As depicted in the schematic diagram of synthesis (Fig. 2a), MnO2 was first synthesized by the molten salt method and then treated with n-butyllithium to obtain Od-MnO2 nanosheets (detailed synthesis process is described in the ESI†). The X-ray diffraction (XRD) patterns of MnO2 and Od-MnO2 are shown in Fig. S2a and S2b (ESI†) which indicate that these two samples exhibit similar crystallographic structures of layered δ-MnO2 (JCPDS no. 52-0556).43,44 Compared with MnO2, the broader and weaker XRD peak features of Od-MnO2 indicate its poor crystallinity, which may be attributed to oxygen defects. The scanning electron microscopy (SEM) images (Fig. 2b and Fig. S3, ESI†) show that both MnO2 and Od-MnO2 exhibit an ultrathin two-dimensional nanosheet morphology, which can effectively shorten ion diffusion paths. The transmission electron microscopy (TEM) images (Fig. 2c and Fig. S4a, ESI†) show that numerous nanopores with sizes of 2–10 nm (Fig. S4b, ESI†) are uniformly dispersed on the Od-MnO2 nanosheet. In the high-resolution transmission electron microscopy (HRTEM) image (Fig. 2d), lattice fringes with a d-spacing of 0.24 nm can be clearly observed, corresponding to the (012) crystal plane of δ-MnO2. The HRTEM magnification and inverse fast Fourier transformation (FFT) diagrams related to the lattice spacing calculation of Od-MnO2 are shown in Fig. S5a–c, ESI.† The corresponding selected area electron diffraction (SAED) pattern (Fig. S6, ESI†) presents typical polycrystalline diffraction rings indexed to the (012), (015) and (110) crystal planes, matching well with the XRD results. The nitrogen adsorption–desorption measurement (Fig. 2e, Fig. S7a and b, ESI†) indicates that Od-MnO2 exhibits a higher Brunauer–Emmett–Teller (BET) specific surface area (150.5 m2 g−1) than MnO2 (92.2 m2 g−1), with nanopores mainly ranging from 2 to 10 nm, providing copious ion diffusion channels and active sites and promoting electrolyte penetration. The presence of abundant oxygen defects is verified by X-ray photoelectron spectroscopy (XPS) and electron paramagnetic resonance (EPR) spectroscopy. The high-resolution O 1s spectrum can be fitted into three peaks at 529.8, 531.4 and 532.9 eV assigned to the Mn–O–Mn bond, oxygen defects (Mn–O–H) and H–O–H bond, respectively (Fig. 2f).10,45–48 The ratio of oxygen defects (calculated from the integrated area in the fitted spectra) for Od-MnO2 (41.5%) is substantially higher than that of pristine MnO2 (18.8%), demonstrating the enhanced concentration of oxygen defect sites after n-butyllithium treatment. The Mn 2p spectrum in Fig. S8a (ESI†) demonstrates the negative shift of the Mn 2p peak position of Od-MnO2 compared to pristine MnO2, suggesting a decrease in Mn valence due to the introduction of oxygen defects. Moreover, the high-resolution Mn 2p spectrum of Od-MnO2 reveals the co-existence of Mn3+ (641.6 eV and 653.1 eV) and Mn4+ (642.8 eV and 654.5 eV) deconvoluted from the Mn 2p3/2 and Mn 2p1/2 orbit, respectively (Fig. S8a, ESI†).49,50 Meanwhile, as shown in Fig. S8b (ESI†), the Mn 3s multistate splitting energy (ΔE) in Od-MnO2 (5.01 eV) is larger than that of MnO2 (4.72 eV), further demonstrating that oxygen defects lower the Mn valence. The EPR spectrum (Fig. 2g) exhibits a stronger electron spin resonance signal at g = 2.0 for Od-MnO2 compared to MnO2, which proves the presence of rich oxygen defects in Od-MnO2.
To investigate the effect of oxygen defects on the reaction kinetics, a coin cell was assembled using pristine MnO2 or Od-MnO2 as the cathode (electrode preparation details are given in the ESI†), Zn foil as the anode, and 2 M Zn (CF3SO3)2 + 0.1 M MnSO4 solution as the electrolyte. All the electrochemical tests in this paper were done in a two-electrode configuration. Fig. 3a displays the cyclic voltammetry (CV) curves of Zn/MnO2 and Zn/Od-MnO2 batteries where the pristine MnO2 or Od-MnO2 electrode is discharged first and then charged in the potential range from 0.8 V to 1.9 V at a scan rate of 0.1 mV s−1, in which two pairs of redox peaks at 1.184/1.598 V and 1.36/1.651 V for MnO2 and 1.203/1.596 V and 1.364/1.646 V for Od-MnO2 corresponding to the insertion/extraction processes of Zn2+ and H+ can be observed, respectively.48,51 It is clear that Od-MnO2 presents higher peak current densities and smaller overpotential gaps than MnO2, suggesting the improved reaction kinetics due to oxygen defects. The basic coincidence of CV curves (Fig. S9†) proves the high reversibility of Od-MnO2. The galvanostatic charge/discharge (GCD) curves of Zn/MnO2 and Zn/Od-MnO2 batteries at a current density of 0.1 A g−1 are displayed in Fig. 3b, demonstrating two obvious charging and discharging voltage plateaus, in agreement with the CV results. In general, it is believed that the first plateau in the MnO2 discharge curve corresponds to the intercalation of H+, while the second plateau is mainly generated by the intercalation of Zn2+. It is known that the capacity contribution of Zn2+ of Od-MnO2 is improved (Fig. 3b), which suggests that oxygen defects can promote the process of Zn2+ intercalation. According to the reaction equation:31,41,52
xZn2+ + 2xe− + MnO2 → MnOOZnx | (1) |
To further analyze the electrochemical kinetics of MnO2 and Od-MnO2, the CV experiment at various scan rates from 0.1 to 0.5 mV s−1 was performed as illustrated in Fig. 3d and Fig. S11a (ESI†). The peak current (i) and scan rate (v) follow the relationship below:53,54
i = avb | (2) |
i = k1v + k2v1/2 | (3) |
To gain a comprehensive insight into the energy storage mechanism of Zn/Od-MnO2 batteries, ex situ measurements including XRD, Raman, XPS, and TEM elemental mapping were performed to analyze the changes in the chemical composition and electronic structure of Od-MnO2 electrodes. Fig. 4a shows the GCD profiles of the first and second cycles at 0.1 A g−1, along with the selected charge–discharge status for ex situ measurements. Ex situ XRD patterns of Od-MnO2 and MnO2 are displayed in Fig. 4b and Fig. S13 (ESI†). It is apparent that MnO2 shows obvious peaks of the Znx(OTf)y(OH)2x−y·nH2O (ZOTH) phase (Fig. S13, ESI†). While the ZOTH peaks of Od-MnO2 are much weaker than that of MnO2 during discharge, and the ZOTH peaks gradually decrease and almost disappear in Od-MnO2 when charging to 1.9 V (Fig. 4b), which further proves the preferential intercalation of Zn2+, thereby reducing the ZOTH produced by H+ intercalation.55,56 Moreover, the appearance/disappearance of the ZnMn2O4 phase (JCPDS no. 71-2499) indicates the reversible Zn2+ insertion/extraction.54,56 In addition, the morphology of Od-MnO2 nanosheets remains unchanged before and after cycling (Fig. S14, ESI†), which further demonstrates that zinc ions are reversibly intercalated/de-intercalated into Od-MnO2. The XRD characteristic peaks of Od-MnO2 remain (Fig. S15a†) and clear lattice fringes attributed to the (012) and (015) crystal planes can still be observed in HRTEM (Fig. S15b†) after discharge, proving that there are no obvious structural changes. Moreover, the high-resolution O 1s spectra (Fig. S16†) indicate that there are still abundant oxygen defects in Od-MnO2 after discharge, ensuring the effectiveness of the preferential Zn2+ intercalation induced by oxygen defects. In ex situ Raman spectra (Fig. 4e), a band of around 650 cm−1 assigned to the symmetric stretching vibration (Mn–O) of the MnO6 groups remains almost unchanged, suggesting the structural stability of Od-MnO2. Upon discharging and charging, the emergence and disappearance of a pair of bands between 300 and 400 cm−1 associated with Zn–O vibrations demonstrate the reversible insertion/extraction of Zn2+.51,57 The intercalation mechanism is further unveiled by ex situ XPS (Fig. 4d and e). The high-resolution Mn 2p analysis (Fig. 4d) reveals the change in the valence state of Mn during charging and discharging. The peak area of Mn3+ and Mn2+ apparently enlarges during discharge, implying the intercalation of Zn2+ into the Od-MnO2 electrode and the reduction of Mn4+.19 Then the peak area of Mn4+ gradually increases during charging along with the manganese which is oxidized to the initial state.57–60 The high-resolution Zn 2p spectra (Fig. 4e) reveal prominent peaks in the discharged state due to the intercalation of Zn2+ into the Od-MnO2 electrode, which weaken after charging caused by Zn2+ extraction.54,61 These phenomena are consistent with the elemental mapping results of Od-MnO2 in fully discharged/charged states (Fig. S14, ESI†), which shows obvious Zn signals together with Mn and O in the discharged state (Fig. S17a†) while showing very weak Zn signals in the charged state (Fig. S17b†), verifying successful Zn2+ insertion/extraction into Od-MnO2.
DFT calculations were done to unravel the underlying mechanism of improving electrochemical kinetics. The calculated partial density of states (PDOS) and charge density distribution of pristine MnO2 and Od-MnO2 are illustrated in Fig. 4f and g. Pristine MnO2 exhibits a significant bandgap at the Fermi level (Fig. 4f), indicating its semiconducting feature. The PDOS of Od-MnO2 (Fig. 4g) shifts to low energy levels and the bandgap narrows after introducing oxygen defects, corresponding to easier electron transition from the valence band to the conduction band, which effectively increases the conductivity and results in lower ohmic polarization and improved rate capability. Moreover, the charge density difference analysis (Fig. S18a and S18b, ESI†) reveals an altered charge density distribution with the introduction of oxygen defects, thereby accelerating electron transfer kinetics.40,62
The electrochemical performance was further elucidated with coin cells and pouch cells using pristine MnO2 or Od-MnO2 as a cathode, Zn foil as an anode, and 2 M Zn(CF3SO3)2 + 0.1 M MnSO4 solution as an electrolyte. Od-MnO2 displays remarkable rate capability, demonstrating high discharge capacities of 341, 306, 254, 217, 187, and 139 mA h g−1 at current densities of 0.1, 0.2, 0.5, 1, 2 and 5 A g−1, respectively (Fig. 5a and Fig. S19, ESI†). Even at an ultrahigh current density of 10 A g−1, Od-MnO2 can still deliver a capacity of 94 mA h g−1 As expected, Od-MnO2 exhibits considerably higher capacities than that of MnO2 at each current density. Upon reverting to a current density of up to 0.1 A g−1, 97.5% of the discharge capacity (332.4 mA h g−1) of Od-MnO2 is recovered. The superior rate performance could be attributed to the stabilization and excellent kinetics of Od-MnO2. To highlight the superiority and practical realization of the Od-MnO2 electrode, Ragone plots reflecting the relationship between the energy density and power density are shown in Fig. 5b. The Zn/Od-MnO2 battery delivers a maximum power density of 16 kW kg−1 and a maximum energy density of 532 W h kg−1 (based on the active mass of the Od-MnO2 electrode), which are better than those of most reported cathode materials for aqueous ZIBs, such as α-MnO2,63 CuHCF,64 Mn2O3,65 δ-MnO2,66 ZnHCF,67 β-MnO2,55 Zn0.25V2O5·nH2O,2 VS2,68 Todorokite,69 Od-Mn3O4@C,54 and K0.8Mn8O16,53 and is promising for energy storage applications. Furthermore, Od-MnO2 exhibits outstanding long-term cycling stability (Fig. 5c) with a discharge capacity of 186.7 mA h g−1 and a high-capacity retention of 85.4% for over 1200 cycles at a current density of 1 A g−1, while the Zn/MnO2 battery shows rapid capacity fading after 300 cycles. Remarkably, even at low current densities of 0.2A g−1 and 0.5A g−1, Od-MnO2 exhibits impressive durability for over 400 cycles (84.7% retention) and over 800 cycles (81.8% retention) with high depths of discharge (DODs), respectively, as shown in Fig. S20a and S20b (ESI†). To further evaluate the practicability of the Od-MnO2 cathode, pouch cells with a size of 4 cm × 4 cm were assembled. As can be seen in Fig. S21 (ESI†), the Zn/Od-MnO2 pouch cell presents similar charge–discharge curves to those in the coin cell, demonstrating the uniformity of electrochemical performance when the battery is scaled up. At a rate of 1C, the Zn/Od-MnO2 pouch cell exhibits a high reversibility of ≈100% CE and negligible capacity decay for over 100 cycles. Moreover, the specific capacity (based on the mass of the active materials of the cathode) after 500 cycles is 126.2 mA h g−1, about 69.8% retention compared with the maximum value (180.7 mA h g−1). These results underscore the potential of a Zn/Od-MnO2 system utilizing a mild aqueous electrolyte for high performance, long life, and environmentally friendly energy storage.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nr03100h |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2024 |