Next Article in Journal
Preparation of Hydroxyapatite-Titanium Dioxide Composite from Eggshell by Hydrothermal Method: Characterization and Antibacterial Activity
Previous Article in Journal
Syntheses and Crystal Structures of Three Chiral Oxazolidinones with Different Ring Conformations
Previous Article in Special Issue
Preparation of Flower-like Nickel-Based Bimetallic Organic Framework Electrodes for High-Efficiency Hybrid Supercapacitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Study of Polymorphic MnO2 in Rechargeable Aqueous Zinc Batteries

1
Wenzhou Key Lab of Advanced Energy Storage and Conversion, Zhejiang Province Key Lab of Leather Engineering, College of Chemistry and Materials Engineering, Wenzhou University, Wenzhou 325035, China
2
Zhejiang Engineering Research Center for Electrochemical Energy Materials and Devices, Institute of New Materials and Industrial Technologies, Wenzhou University, Wenzhou 325035, China
*
Authors to whom correspondence should be addressed.
Crystals 2022, 12(11), 1600; https://doi.org/10.3390/cryst12111600
Submission received: 17 September 2022 / Revised: 30 October 2022 / Accepted: 7 November 2022 / Published: 10 November 2022
(This article belongs to the Special Issue Structure Property Relationship of Energy Storage Materials)

Abstract

:
Manganese dioxide is regarded as a promising energy functional material due to its open tunnel structure with enormous applications in energy storage and catalysis. In this paper, α-MnO2 with a 2 × 2 tunnel structure and β-MnO2 with a 1 × 1 tunnel structure were hydrothermally synthesized, which possess characteristic tunnel structures formed by the interconnected unit structure of [MnO6] octahedrons. With regards to their different tunnel dimensions, the specific mechanism of ion intercalation in these two phases and the effect on their performance as aqueous Zn-MnO2 battery cathodes are explored and compared. Comprehensive analyses illustrate that both α-MnO2 and β-MnO2 provide decent capacity in the aqueous battery system, but their intrinsic stability is poor due to the structural instability upon cycling. At the same time, experiments show that α-MnO2 has a better rate performance than β-MnO2 under larger currents, thus implying that the former has a broader application in this aqueous battery system.

1. Introduction

At present, the existing secondary battery market is populated by lithium-ion batteries (LIBs). However, drawbacks such as the insecurity of active lithium metal and the limitations of lithium resources hinder their application in the large energy storage devices of LIBs [1,2,3,4]. There is thus an urgent need to develop alternative, new secondary batteries. Aqueous zinc batteries have attracted a lot of attention because of their good safety, high theoretical capacity, and low cost [5,6,7]. When a given anode is zinc metal, the open-circuit voltage and zinc-ion storage capacity are almost entirely dependent on the cathode material of the aqueous zinc batteries [5,8,9,10]. The ion and electron transport properties and redox reaction kinetics of the cathode and cathode–electrolyte interface will affect the gravimetric energy and power density of zinc batteries [11]. Therefore, the design and development of cathode materials with a large storage capacity and robust crystalline structure has become a major challenge in the field [12]. So far, the cathode materials that have been developed and utilized mainly include vanadium-based materials, Prussian blue with its analogs, and manganese-based materials [13,14,15,16,17,18]. For example, vanadium-based materials have received extensive attention as promising cathode materials for multivalent batteries due to their huge theoretical capacity, multifunctional crystal structure, abundant resources, and low cost. Commonly used vanadium-based materials include vanadium oxide, metal vanadate, vanadium phosphate, etc. [19,20,21]. However, the structural instability and chemical dissolution of vanadium lead to fast capacity fading which means poor cycling stability [22]. As the tenth most abundant element in the earth’s crust, manganese is a typical transition metal element and a valence-variable metallic element, which is easily oxidized into manganese oxide or manganese hydroxide minerals. Manganese-based oxides have long been used as electrode materials for energy storage and conversion applications because of their good electrochemical performance, low cost, natural abundance, and low toxicity [23]. Common manganese-based oxides are mainly: MnO, MnO2, Mn2O3, Mn3O4, etc. [24,25,26]. As a commonly used cathode material, manganese dioxide has many different crystal forms, such as α-MnO2 [27], γ-MnO2 [28], β-MnO2 [29,30], todorokite-MnO2, [31,32], and so on. They are all connected by the unit structure of MnO2 octahedrons in various ways to form the rich polymorphism [32]. The microstructure and morphology of MnO2 will directly affect the reaction mechanism and electrochemical performance of aqueous Zn-MnO2 batteries [33]. Among various forms of MnO2, α-MnO2 with a 2 × 2 tunnel structure, β-MnO2 with a 1 × 1 tunnel, and δ-MnO2 with a layer structure are widely explored as electrode candidates for aqueous Zn-MnO2 batteries [34,35,36].
In this work, we synthesized 2 × 2 tunnel-structued α-MnO2, 1 × 1 tunnel-structued β-MnO2, and layer-structured δ-MnO2 by a simple hydrothermal method. The energy storage behaviors of the three samples were systematically compared by electrochemical analysis. The experimental results show that α-MnO2 exhibits better energy storage performance than β-MnO2 and δ-MnO2, which could be understood by considering both their tunnel dimension differences and structural stability under battery cycling conditions.

2. Experimental Methods

2.1. Sample Synthesis

In this paper, the hydrothermal synthesis method is used to prepare α-MnO2, β-MnO2, and δ-MnO2. All chemicals were all purchased from Aladdin and used directly without any further treatment. The specific operation dosage is as follows:
The synthesis of α-MnO2: 23.5 mmol KMnO4 (3.725 g) and 9.38 mmol MnSO4·H2O (1.585 g) were dissolved in 150 mL of deionized water, stirred for 10 min to fully dissolve, mixed (slowly pour MnSO4·H2O into KMnO4) and set aside. In a 500 mL hydrothermal reactor, the reaction was stirred at room temperature for 30 min and finally placed in an oven at 160 °C for 12 h to obtain anhydrous α- MnO2.
The synthesis of β-MnO2: After dissolving 7.2 mmol KMnO4 (1.138 g) and 12 mmol MnSO4·H2O (2.028 g) in 40 mL of deionized water, respectively, the solution was fully stirred for 10 min. Then, the two (slowly pouring KMnO4 into MnSO4·H2O) were mixed into a 250 mL hydrothermal reactor, the reaction was stirred at room temperature for 30 min and then placed in an oven at 160 °C for 12 h to obtain anhydrous β-MnO2.
The synthesis of δ-MnO2: 100 mL 0.3 M Mn(Cl)2.4H2O solution and 1.58 g KMnO4 were added into 100 mL 3.0 M NaOH, the mix solution was left in an ice bath for 15 min until the color turned to brown. Then, the mixture was moved from the ice bath, aged over 7 days, and finally deionized water and ethanol were used to wash the product during the filtration until the filtrate was nearly neutral.

2.2. Sample Characterization

The morphology and microstructure were observed by field emission scanning electron microscopy (SEM, FEI, Nova 200 NanoSEM) and by a transmission electron microscope (TEM, JEOL, JEM-2100F). X-ray powder diffraction (XRD, Bruker, D8ADVANCE) was carried out to illustrate the crystallographic structure.

2.3. Electrochemical Measurement

A slurry was prepared for cathodes with 70 wt% active materials, 20 wt% Ketjen black, and 10 wt% polyvinylidene fluoride binder in N-methyl-2-pyrrolidone (NMP). The mixed slurry was casted on stainless steel foil, and dried at 80 °C for 12 h in a vacuum oven. CR2032 coin-type cells were manufactured to estimate the electrochemical performance of pristine α-MnO2, β-MnO2, and δ-MnO2 cathodes in aqueous Zn-MnO2 batteries, with zinc metals as the anode. The loading of active materials on the cathodes (r = 12 mm) was 1.5 mg cm−2. 2M ZnSO4 was used as electrolyte, and the separator was glass fiber. All the electrochemical tests were performed between 1.0 and 1.8 V. Cycling performance and Galvanostatic charge–dicharge were tested in Neware. EIS (0.01 Hz to 100,000 Hz) and CV profiles were measured by electrochemical workstation (CHI660E, Shanghai Chenhua, China).

3. Results and Discussion

The X-ray powder diffraction (XRD) patterns (Figure 1a) show that the as-synthesized samples can be indexed with α-MnO2 (JCPDS file #42-1348) and β-MnO2 (JCPDS file #24-0735) and δ-MnO2 (JCPDS file # 80-1089), individually. Figure 1b,c depict the ball and stick theoretical models of α-MnO2 and β-MnO2, respectively. β-MnO2 has no cations in 1 × 1 tunnel frames owing to its smaller space, while α-MnO2 with a 2 × 2 tunnel structure usually possesses large cations (pink spheres) to avoid structure collapse. Figure 1d shows the model of layered δ-MnO2. Typically, the MnO2 synthesized by a simple hydrothermal method are generally one-dimensional structures, such as nanowires and nanorods. The SEM images and TEM images of the obtained samples are similar to those recorded in the related literature [3,37,38], as shown in Figure 2 and Figure 3. Figure 2a shows α-MnO2 nanowires with a length of around 1.5 μm and a width of around 100 nm, and Figure 2b shows the β-MnO2 with a length of around 1.5 μm and a width of around 200 nm, indicating that β-MnO2 has a larger size. The α-MnO2 nanowires are mainly irregular, but some will gather in a flower shape, while β-MnO2 nanorods are uniformly dispersed. SEM images show layered δ-MnO2 mainly presenting irregular sheet shapes in Figure 2c.
The microscopic morphology of the α-MnO2 nanowires, β-MnO2 nanorods, and layered δ-MnO2 were further characterized in Figure 3. In most cases, the two ends of α-MnO2 are irregular sections, and β-MnO2 has the regular tips of two ends showing a shuttle-like shape, while δ-MnO2 are mainly distributed in sheets (Figure 3a,d,g). The high-resolution transmission electron microscope (HRTEM) images of α-MnO2, β-MnO2 and δ-MnO2 are shown in Figure 3c,f,g, respectively. HRTEM images show that both α-MnO2 nanowires and β-MnO2 nanorods grow along their individual [001] directions and are uniformly dispersed. The typical characteristic crystal planes of α-MnO2, β-MnO2, and δ-MnO2 are shown in Figure 3c,f,i, respectively. The (110) interplanar spacing of α-MnO2 is 6.95 Å, while the (110) interplanar spacing of β-MnO2 is 3.28 Å, which is much smaller than the (110) crystal plane of α-MnO2, indicating that the tunnel size of β-MnO2 is smaller than that of α-MnO2, which is in line with the existing literature reports [27,39,40]. At the same time, the edge of α-MnO2 exhibits a certain degree of lattice disorder and a relatively thick amorphous layer; on the contrary, the edge of β-MnO2 exhibits a regular crystalline layer. The above observations can further show that β-MnO2 has better thermodynamic stability despite its smaller tunnel size. The crystal lattice of δ-MnO2 was observed in the HRTEM image (Figure 3i), corresponding to the planes of (−111) and (002), for which the spacing is 0.24 and 0.34 nm, which takes no account of the direaction of the crystalline domins.
The galvanostatic charge–discharge (GCD) curves of α-MnO2, β-MnO2, and δ-MnO2 were obtained at the rate of 0.2 C (1 C = 300 mA h g−1), and the three MnO2 with different crystal forms exhibit completely different performances in the initial stage (Figure 4).
All MnO2 samples exhibit similar charge–discharge profiles with a gradient plateau corresponding with previously reported α-MnO2, β-MnO2, and δ-MnO2 electrodes (Figure 4a–c) [27,39,41,42]. The MnO2 samples have similar first discharge profiles with different plateau voltages. The single plateau voltage of α-MnO2 in the first discharge profiles is around 1.25 V (Figure 4a), while those of β-MnO2 and δ-MnO2 are around 1.3 V (Figure 4b) 1.32 V (Figure 4c), respectively. The first charge profiles of all MnO2 samples present two identical plateaus at 1.55 V and 1.6 V. Through the first cycle of the discharge–charge reaction, the δ-MnO2||Zn battery shows a smaller overpotential than the other two MnO2||Zn batteries, indicating that the α-MnO2||Zn and β-MnO2||Zn batteries need a larger applied voltage in the initial reaction. In other words, it is more difficult for α-MnO2 and β-MnO2 to react. However, in the subsequent cycles, the overpotential plateaus of α-MnO2||Zn batteries in charge–discharge profiles gradually tend to be consistent with β-MnO2||Zn batteries, further demonstrating the reason for the different cycle stability of the two batteries. It is notable that the structure of α-MnO2 remains good, but its initial specific capacity is poor.
During the first discharge process, the specific capacities of α-MnO2, β-MnO2, and δ-MnO2 are 310 mA h g−1, 224 mA h g−1, and 173 mA h g−1, respectively, indicating that the larger tunnel size of α-MnO2 is beneficial for charge storage. It is worth mentioning that the specific capacity of α-MnO2 decays rapidly during 1–3 cycles and the declining trend becomes slower in the subsequent cycling process (Figure 4e). However, the specific capacity of β-MnO2 has an upward trend in the first few cycles, and a high specific capacity up to 290 mA h g−1 is obtained at the 4th cycle. The specific capacity declines rapidly and then slows down in the following cycling process. In addition, the cycling process of the β-MnO2 cathode is aborted after 90 cycles due to its poor structure stability. It is easy to be seen that δ-MnO2 shows the best capacity retention owing to its layered structure. Cyclic comparisons manifest that α-MnO2 has better performance benefitting from its larger tunnel size, and the structural stability of α-MnO2 is significantly better than that of β-MnO2. Moreover, the activation process of β-MnO2 in the initial stage of the cycling results in the material pulverizing rapidly, further leading to the battery breakdown after 90 cycles. Meanwhile, α-MnO2 still maintains the constant current charge and discharge state until after 100 cycles. Figure 4d depicts the rate performance of α-MnO2, β-MnO2 and δ-MnO2 electrodes. At the small current density of 0.1 C, β-MnO2 delivers a higher specific capacity than α-MnO2 and δ-MnO2. When the current density increases to 0.2 C, the specific capacity of α-MnO2 exceeds that of β-MnO2 and δ-MnO2, while at the high rate of 1 C, β-MnO2 provides almost no capacity. Notably, when the charge–discharge rate goes back to 0.1 C, the capacity retention of α-MnO2 and β-MnO2 was around 50%, yet δ-MnO2 still retained about 70%, whereas δ-MnO2 maintains good capacity retention. The overall capacity of δ-MnO2 decreases slightly during charging/discharging and this can be attributed to its layered structure. However, a higher capacity of α-MnO2 during the entire process attracts more attention.
The CV curves of different MnO2 cathodes at a scan rate of 0.1 mV s−1 between the voltage range of 1.0–1.8 V were shown in Figure 5a–c. In the first cycle, there is one obvious peak at 1.14 V for α-MnO2 and δ-MnO2, and one at 1.21 V for the β-MnO2. These peaks can be attributed to the intercalation of H+ into the different lattice positions of MnO2 tunnel structure, with the reaction of Mn4+ to Mn3+ at the same time. During the following cycle, two reduction peaks were viewed at 1.22 V and 1.36 V for α-MnO2, β-MnO2 and δ-MnO2. In addition, one obvious oxidation peak was observed around 1.6 V and 1.67 V for these MnO2 samples. With the cycling progress, the intensity of the redox peaks of α-MnO2 decrease slightly, while the peaks of β-MnO2 increase apparently, indicting increased polarization of the battery for α-MnO2 and the activation reaction for β-MnO2. By comparison, α-MnO2 and β-MnO2 present a smaller enclosed region of the CV profiles in the following process, yet δ-MnO2 is likely to remain unchanged, demonstrating that the tunnel structure has a lower capacity retention than the layer structure. Figure 5d illustrates the AC impedance spectroscopy curves of different MnO2 electrodes. According to the EIS curves of α-MnO2, β-MnO2 and δ-MnO2, apparently, all the Nyquist plots have a similar shape with a semicircle in the high–medium frequency region and a straight line in the low-frequency region, demonstrating that the electrochemical behavior of the three samples is controlled by both charge transfer and ion diffusion. The EIS data can be well fitted by a typical equivalent electrical circuit (the inset of Figure 5d), where R1 is the ohmic resistance, R2 is the faradic charge-transfer resistance, and Wo1 is the Warburg impedance. Notably, the lower the values of these parameters, the better the performance of the electrode. Figure 5d shows that the R2 value of α-MnO2 is much smaller than that of β-MnO2 and δ-MnO2, indicating a higher electric conductivity of α-MnO2. The higher slope of δ-MnO2 indicates the better Zn-ion diffusion in the internal electrodes owing to its layered structure.

4. Conclusions

In this work, the characterization of the microscopic morphology and structure of α-MnO2 nanowires, β-MnO2 nanorods, and layered δ-MnO2 was carried out using HRTEM and XRD, etc. Compared with 2 × 2 tunnel-structured α-MnO2, 1 × 1 tunnel-structured β-MnO2 possesses better thermodynamic stability and a more regular morphological structure. However, when used as the cathode of zinc ion batteries, α-MnO2 with a larger tunnel size is superior in both cycling and rate performance compared with β-MnO2, although layered δ-MnO2 exhibits higher capacity retention and poor initial capacity. This shows that the larger tunnel structure of α-MnO2 can accommodate more carriers and has a broader application prospect. This work provides electrochemical insights into the energy storage of representative Zn-MnO2 aqueous batteries and points to future strategies to enhance cycling stability, such as face modification and modified electrolytes to suppress Mn dissolution. Furthermore, the electrochemical behaviors of α-MnO2, β-MnO2, and δ-MnO2 have been compared, which may motivate the development of more sustainable systems other than Zn-MnO2 batteries in which the performance of aqueous electrolytes is enhanced.

Author Contributions

Conceptualization, Y.Y. and H.J.; methodology, K.Y. and Y.Y.; formal analysis, K.Y., X.L., X.H. and W.S.; writing—original draft, K.Y.; writing—review and editing, Y.Y. and S.W.; supervision, Y.Y.; project administration, Y.Y.; funding acquisition Y.Y., H.J. and S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 52002287, 52272088 and 52072273, Zhejiang Ten Thousand Plan Program (2019R52042), and National Talent Introduction Demonstration Base (2021ZY1007).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, T.; Nam, G.; Liu, K.; Wang, J.-H.; Zhao, B.; Ding, Y.; Soule, L.; Avdeev, M.; Luo, Z.; Zhang, W.; et al. A niobium oxide with a shear structure and planar defects for high-power lithium ion batteries. Energy Environ. Sci. 2022, 15, 254–264. [Google Scholar] [CrossRef]
  2. Oh, H.S.; Jeong, H.M.; Park, J.H.; Ock, I.-W.; Kang, J.K. Hierarchical Si hydrogel architecture with conductive polyaniline channels on sulfonated-graphene for high-performance Li ion battery anodes having a robust cycle life. J. Mater. Chem. A 2015, 3, 10238–10242. [Google Scholar] [CrossRef]
  3. Yuan, Y.; Sharpe, R.; He, K.; Li, C.; Saray, M.T.; Liu, T.; Yao, W.; Cheng, M.; Jin, H.; Wang, S.; et al. Understanding intercalation chemistry for sustainable aqueous zinc–manganese dioxide batteries. Nat. Sustain. 2022, 5, 890–898. [Google Scholar] [CrossRef]
  4. Or, T.; Gourley, S.W.D.; Kaliyappan, K.; Yu, A.; Chen, Z. Recycling of mixed cathode lithium-ion batteries for electric vehicles: Current status and future outlook. Carbon Energy 2020, 2, 6–43. [Google Scholar] [CrossRef] [Green Version]
  5. Cao, J.; Zhang, D.; Gu, C.; Wang, X.; Wang, S.; Zhang, X.; Qin, J.; Wu, Z.S. Manipulating crystallographic orientation of zinc deposition for dendrite-free zinc ion batteries. Adv. Energy Mater. 2021, 11, 2101299. [Google Scholar] [CrossRef]
  6. Cresce, A.; Xu, K. Aqueous lithium-ion batteries. Carbon Energy 2021, 3, 721–751. [Google Scholar] [CrossRef]
  7. Yang, D.; Chen, D.; Jiang, Y.; Ang, E.H.; Feng, Y.; Rui, X.; Yu, Y. Carbon-based materials for all-solid-state zinc–air batteries. Carbon Energy 2020, 3, 50–65. [Google Scholar] [CrossRef]
  8. He, X.; Zhang, Y.; Wang, J.; Li, J.; Yu, L.; Zhou, F.; Li, J.; Shen, X.; Wang, X.; Wang, S.; et al. Biomass-derived Fe2N@NCNTs from bioaccumulation as an efficient electrocatalyst for oxygen reduction and Zn–Air battery. ACS Sustain. Chem. Eng. 2022, 10, 9105–9112. [Google Scholar] [CrossRef]
  9. Javed, M.S.; Mateen, A.; Ali, S.; Zhang, X.; Hussain, I.; Imran, M.; Shah, S.S.A.; Han, W. The Emergence of 2D MXenes Based Zn-Ion Batteries: Recent Development and Prospects. Small 2022, 18, e2201989. [Google Scholar] [CrossRef]
  10. Chodankar, N.R.; Patil, S.J.; Lee, S.; Lee, J.; Hwang, S.K.; Shinde, P.A.; Bagal, I.V.; Karekar, S.V.; Seeta Rama Raju, G.; Shanmugam Ranjith, K.; et al. High energy superstable hybrid capacitor with a self-regulated Zn/electrolyte interface and 3D graphene-like carbon cathode. InfoMat 2022, 4, e12344. [Google Scholar] [CrossRef]
  11. Zhang, K.; Zhang, Y.; Zhang, Q.; Liang, Z.; Gu, L.; Guo, W.; Zhu, B.; Guo, S.; Zou, R. Metal-organic framework-derived Fe/Cu-substituted Co nanoparticles embedded in CNTs-grafted carbon polyhedron for Zn-air batteries. Carbon Energy 2020, 2, 283–293. [Google Scholar] [CrossRef] [Green Version]
  12. Guo, Y.; Hong, X.; Su, Y.; Luo, W.; Yu, R.; Wu, J.; Hensen, E.J.M.; Mai, L.; Cao, Y. Sub-Nanometer Confined Ions and Solvent Molecules Intercalation Capacitance in Microslits of 2D Materials. Small 2021, 17, e2104649. [Google Scholar] [CrossRef] [PubMed]
  13. Du, Z.; Gu, J.; Cao, Z.; Wang, H.; Zhao, Q.; Ye, Y.; Li, B.; Chen, W.; Liu, C.; Yang, S. 2D non-Van Der Waals transition-metal chalcogenide layers derived from vanadium-based MAX phase for ultrafast Zinc storage. Adv. Energy Mater. 2022, 12, 2200943. [Google Scholar] [CrossRef]
  14. Pan, Q.; Dong, R.; Lv, H.; Sun, X.; Song, Y.; Liu, X.-X. Fundamental understanding of the proton and zinc storage in vanadium oxide for aqueous zinc-ion batteries. Chem. Eng. J. 2021, 419, 129491. [Google Scholar] [CrossRef]
  15. Zeng, Y.; Lu, X.F.; Zhang, S.L.; Luan, D.; Li, S.; Lou, X.W.D. Construction of Co-Mn prussian blue analog hollow spheres for efficient aqueous Zn-ion batteries. Angew. Chem. Int. Ed. Engl. 2021, 60, 22189–22194. [Google Scholar] [CrossRef] [PubMed]
  16. Ma, K.; Li, Q.; Hong, C.; Yang, G.; Wang, C. Bi doping-enhanced reversible-phase transition of alpha-MnO2 raising the cycle capability of aqueous Zn-Mn batteries. ACS Appl. Mater. Interfaces 2021, 13, 55208–55217. [Google Scholar] [CrossRef]
  17. Patil, S.J.; Chodankar, N.R.; Hwang, S.-K.; Raju, G.S.R.; Ranjith, K.S.; Huh, Y.S.; Han, Y.-K. Ultra-stable flexible Zn-ion capacitor with pseudocapacitive 2D layered niobium oxyphosphides. Energy Storage Mater. 2022, 45, 1040–1051. [Google Scholar] [CrossRef]
  18. Javed, M.S.; Lei, H.; Wang, Z.; Liu, B.-t.; Cai, X.; Mai, W. 2D V2O5 nanosheets as a binder-free high-energy cathode for ultrafast aqueous and flexible Zn-ion batteries. Nano Energy 2020, 70, 104573. [Google Scholar] [CrossRef]
  19. Liu, Z.; Pulletikurthi, G.; Endres, F. A prussian blue/zinc secondary battery with a bio-ionic liquid-water mixture as electrolyte. ACS Appl. Mater. Interfaces 2016, 8, 12158–12164. [Google Scholar] [CrossRef]
  20. Liu, C.; Neale, Z.; Zheng, J.; Jia, X.; Huang, J.; Yan, M.; Tian, M.; Wang, M.; Yang, J.; Cao, G. Expanded hydrated vanadate for high-performance aqueous zinc-ion batteries. Energy Environ. Sci. 2019, 12, 2273–2285. [Google Scholar] [CrossRef]
  21. Jenkins, T.; Alarco, J.A.; Cowie, B.; Mackinnon, I.D.R. Validating the electronic structure of vanadium phosphate cathode materials. ACS Appl. Mater. Interfaces 2021, 13, 45505–45520. [Google Scholar] [CrossRef] [PubMed]
  22. Kim, S.J.; Wu, D.; Sadique, N.; Quilty, C.D.; Wu, L.; Marschilok, A.C.; Takeuchi, K.J.; Takeuchi, E.S.; Zhu, Y. Unraveling the dissolution-mediated reaction mechanism of alpha-MnO2 cathodes for aqueous Zn-ion batteries. Small 2020, 16, e2005406. [Google Scholar] [CrossRef]
  23. Ren, Y.; Meng, F.; Zhang, S.; Ping, B.; Li, H.; Yin, B.; Ma, T. CNT@MnO2 composite ink toward a flexible 3D printed micro-zinc-ion battery. Carbon Energy 2022, 4, 446–457. [Google Scholar] [CrossRef]
  24. Liu, C.; Zhang, C.; Song, H.; Zhang, C.; Liu, Y.; Nan, X.; Cao, G. Mesocrystal MnO cubes as anode for Li-ion capacitors. Nano Energy 2016, 22, 290–300. [Google Scholar] [CrossRef] [Green Version]
  25. He, K.; Yuan, Y.; Yao, W.; You, K.; Dahbi, M.; Alami, J.; Amine, K.; Shahbazian-Yassar, R.; Lu, J. Atomistic insights of irreversible Li(+) intercalation in MnO2 electrode. Angew. Chem. Int. Ed. Engl. 2022, 61, e202113420. [Google Scholar] [PubMed]
  26. Liu, C.; Li, Q.; Sun, H.; Wang, Z.; Gong, W.; Cong, S.; Yao, Y.; Zhao, Z. MOF-derived vertically stacked Mn2O3@C flakes for fiber-shaped zinc-ion batteries. J. Mater. Chem. A 2020, 8, 24031–24039. [Google Scholar] [CrossRef]
  27. Gao, X.; Wu, H.; Li, W.; Tian, Y.; Zhang, Y.; Wu, H.; Yang, L.; Zou, G.; Hou, H.; Ji, X. H(+) -insertion boosted alpha-MnO2 for an aqueous Zn-ion battery. Small 2020, 16, e1905842. [Google Scholar] [CrossRef]
  28. Zhang, L.; Liu, Q.; Wang, Y.; Xu, C.; Bi, J.; Mu, D.; Wu, B.; Wu, F. Modifying γ-MnO2 to enhance the electrochemical performance of lithium-sulfur batteries. Chem. Eng. J. 2021, 421, 129782. [Google Scholar] [CrossRef]
  29. Ding, S.; Zhang, M.; Qin, R.; Fang, J.; Ren, H.; Yi, H.; Liu, L.; Zhao, W.; Li, Y.; Yao, L.; et al. Oxygen-deficient beta-MnO2@graphene oxide cathode for high-rate and long-life aqueous zinc ion batteries. Nanomicro Lett. 2021, 13, 173. [Google Scholar]
  30. Liu, M.; Zhao, Q.; Liu, H.; Yang, J.; Chen, X.; Yang, L.; Cui, Y.; Huang, W.; Zhao, W.; Song, A.; et al. Tuning phase evolution of β-MnO2 during microwave hydrothermal synthesis for high-performance aqueous Zn ion battery. Nano Energy 2019, 64, 103942. [Google Scholar] [CrossRef]
  31. Byles, B.W.; Cullen, D.A.; More, K.L.; Pomerantseva, E. Tunnel structured manganese oxide nanowires as redox active electrodes for hybrid capacitive deionization. Nano Energy 2018, 44, 476–488. [Google Scholar] [CrossRef]
  32. Yuan, Y.; Liu, C.; Byles, B.W.; Yao, W.; Song, B.; Cheng, M.; Huang, Z.; Amine, K.; Pomerantseva, E.; Shahbazian-Yassar, R.; et al. Ordering heterogeneity of [MnO6] octahedra in tunnel-structured MnO2 and its influence on ion storage. Joule 2019, 3, 471–484. [Google Scholar] [CrossRef] [Green Version]
  33. Gupta, G.; Selvakumar, K.; Lakshminarasimhan, N.; Senthil Kumar, S.M.; Mamlouk, M. The effects of morphology, microstructure and mixed-valent states of MnO2 on the oxygen evolution reaction activity in alkaline anion exchange membrane water electrolysis. J. Power Sources 2020, 461, 228131. [Google Scholar] [CrossRef]
  34. Gu, Y.; Min, Y.; Li, L.; Lian, Y.; Sun, H.; Wang, D.; Rummeli, M.H.; Guo, J.; Zhong, J.; Xu, L.; et al. Crystal splintering of β-MnO2 induced by interstitial Ru doping toward reversible oxygen conversion. Chem. Mater. 2021, 33, 4135–4145. [Google Scholar] [CrossRef]
  35. Yao, W.; Odegard, G.M.; Huang, Z.; Yuan, Y.; Asayesh-Ardakani, H.; Sharifi-Asl, S.; Cheng, M.; Song, B.; Deivanayagam, R.; Long, F.; et al. Cations controlled growth of β-MnO2 crystals with tunable facets for electrochemical energy storage. Nano Energy 2018, 48, 301–311. [Google Scholar] [CrossRef]
  36. Chao, D.; Zhou, W.; Ye, C.; Zhang, Q.; Chen, Y.; Gu, L.; Davey, K.; Qiao, S.Z. An electrolytic Zn-MnO2 battery for high-voltage and scalable energy storage. Angew. Chem. Int. Ed. Engl. 2019, 58, 7823–7828. [Google Scholar] [CrossRef]
  37. Yuan, Y.; Zhan, C.; He, K.; Chen, H.; Yao, W.; Sharifi-Asl, S.; Song, B.; Yang, Z.; Nie, A.; Luo, X.; et al. The influence of large cations on the electrochemical properties of tunnel-structured metal oxides. Nat. Commun. 2016, 7, 13374. [Google Scholar] [CrossRef]
  38. Gu, Y.; Yan, G.; Lian, Y.; Qi, P.; Mu, Q.; Zhang, C.; Deng, Z.; Peng, Y. MnIII-enriched α-MnO2 nanowires as efficient bifunctional oxygen catalysts for rechargeable Zn-air batteries. Energy Storage Mater. 2019, 23, 252–260. [Google Scholar] [CrossRef]
  39. Chen, H.; Dai, C.; Xiao, F.; Yang, Q.; Cai, S.; Xu, M.; Fan, H.J.; Bao, S.J. Reunderstanding the reaction mechanism of aqueous Zn-Mn batteries with sulfate electrolytes: Role of the zinc sulfate hydroxide. Adv. Mater. 2022, 34, e2109092. [Google Scholar] [CrossRef]
  40. Zhao, Y.; Zhang, P.; Liang, J.; Xia, X.; Ren, L.; Song, L.; Liu, W.; Sun, X. Uncovering sulfur doping effect in MnO2 nanosheets as an efficient cathode for aqueous zinc ion battery. Energy Storage Mater. 2022, 47, 424–433. [Google Scholar] [CrossRef]
  41. Wu, D.; Housel, L.M.; Kim, S.J.; Sadique, N.; Quilty, C.D.; Wu, L.; Tappero, R.; Nicholas, S.L.; Ehrlich, S.; Zhu, Y.; et al. Quantitative temporally and spatially resolved X-ray fluorescence microprobe characterization of the manganese dissolution-deposition mechanism in aqueous Zn/α-MnO2 batteries. Energy Environ. Sci. 2020, 13, 4322–4333. [Google Scholar] [CrossRef]
  42. Islam, S.; Alfaruqi, M.H.; Mathew, V.; Song, J.; Kim, S.; Kim, S.; Jo, J.; Baboo, J.P.; Pham, D.T.; Putro, D.Y.; et al. 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. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray powder diffraction of α-MnO2, β-MnO2, and δ-MnO2. Ball and stick models for the theoretical structures of (b) α-MnO2 and (c) β-MnO2 phase. Yellow and red spheres are Mn and O atoms, respectively, while the pink spheres are K ions in the tunnel (b) (The theoretical model was obtained by CrystalMaker 10.8). (d) Structural illustration of the as-prepared Na ion and water molecule intercalated layered δ-MnO2. Copyright © 2019, American Chemical Society.
Figure 1. (a) X-ray powder diffraction of α-MnO2, β-MnO2, and δ-MnO2. Ball and stick models for the theoretical structures of (b) α-MnO2 and (c) β-MnO2 phase. Yellow and red spheres are Mn and O atoms, respectively, while the pink spheres are K ions in the tunnel (b) (The theoretical model was obtained by CrystalMaker 10.8). (d) Structural illustration of the as-prepared Na ion and water molecule intercalated layered δ-MnO2. Copyright © 2019, American Chemical Society.
Crystals 12 01600 g001
Figure 2. SEM images of (a) α-MnO2, (b) β-MnO2, and (c) δ-MnO2.
Figure 2. SEM images of (a) α-MnO2, (b) β-MnO2, and (c) δ-MnO2.
Crystals 12 01600 g002
Figure 3. TEM images of (ac) α-MnO2, (df) β-MnO2, and (gi) δ-MnO2.
Figure 3. TEM images of (ac) α-MnO2, (df) β-MnO2, and (gi) δ-MnO2.
Crystals 12 01600 g003
Figure 4. Charge−discharge profiles at cycles 1, 2, 3, 4, 5, and 6 for (a) α-MnO2, (b) β-MnO2 and (c) δ-MnO2. (d) Rate performance at various current densities (in the order of 0.1 C, 0.2 C, 0.5 C, 1 C and 0.1 C) of α-MnO2, β-MnO2 and δ-MnO2; (e) evolution of discharge capacity (mA h g−1) over 100 cycles in 0.2 C for different MnO2 phases.
Figure 4. Charge−discharge profiles at cycles 1, 2, 3, 4, 5, and 6 for (a) α-MnO2, (b) β-MnO2 and (c) δ-MnO2. (d) Rate performance at various current densities (in the order of 0.1 C, 0.2 C, 0.5 C, 1 C and 0.1 C) of α-MnO2, β-MnO2 and δ-MnO2; (e) evolution of discharge capacity (mA h g−1) over 100 cycles in 0.2 C for different MnO2 phases.
Crystals 12 01600 g004
Figure 5. The different cyclic voltammogram (CV) profiles of (a) α-MnO2, (b) β-MnO2, and (c) δ-MnO2 at a scan rate of 0.1 mV s−1 from 1.0 to 1.8 V; (d) the electrochemical impedance spectroscopy (EIS) of α-MnO2, β-MnO2, and δ-MnO2 electrodes.
Figure 5. The different cyclic voltammogram (CV) profiles of (a) α-MnO2, (b) β-MnO2, and (c) δ-MnO2 at a scan rate of 0.1 mV s−1 from 1.0 to 1.8 V; (d) the electrochemical impedance spectroscopy (EIS) of α-MnO2, β-MnO2, and δ-MnO2 electrodes.
Crystals 12 01600 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

You, K.; Yuan, Y.; Liao, X.; Song, W.; He, X.; Jin, H.; Wang, S. Electrochemical Study of Polymorphic MnO2 in Rechargeable Aqueous Zinc Batteries. Crystals 2022, 12, 1600. https://doi.org/10.3390/cryst12111600

AMA Style

You K, Yuan Y, Liao X, Song W, He X, Jin H, Wang S. Electrochemical Study of Polymorphic MnO2 in Rechargeable Aqueous Zinc Batteries. Crystals. 2022; 12(11):1600. https://doi.org/10.3390/cryst12111600

Chicago/Turabian Style

You, Kun, Yifei Yuan, Xiuxian Liao, Wenjun Song, Xuedong He, Huile Jin, and Shun Wang. 2022. "Electrochemical Study of Polymorphic MnO2 in Rechargeable Aqueous Zinc Batteries" Crystals 12, no. 11: 1600. https://doi.org/10.3390/cryst12111600

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop