Temperature-controlled fabrication of hydrophilic manganese oxide microspheres as high-performance electrode materials for supercapacitors

Yuchen Wang Yaoyu Liu Ahsan Ejaz Kai Yan

Citation:  Yuchen Wang, Yaoyu Liu, Ahsan Ejaz, Kai Yan. Temperature-controlled fabrication of hydrophilic manganese oxide microspheres as high-performance electrode materials for supercapacitors[J]. Chinese Chemical Letters, 2023, 34(5): 107538. doi: 10.1016/j.cclet.2022.05.052 shu

Temperature-controlled fabrication of hydrophilic manganese oxide microspheres as high-performance electrode materials for supercapacitors

English

  • Nowadays, the growing energy crisis and environmental problem have accelerated the development of energy storage technologies [1,2]. Supercapacitors (SCs) are one of representative energy storage devices with the advantages of rapid charge/discharge rate, high specific power and long cycle life [3-6]. Manganese oxides (MnOx) with multiple valence states have been considered as promising candidates of electrode materials for SCs due to their large theoretical specific capacitance value, natural abundance and environmental friendliness [7-9]. Generally, supercapacitive performance of MnOx is characterized by their specific capacitance CMnOx and maximum power Pmax, which are governed by the Mn valence and conductivity according to [10,11]: CMnOx = Fn(OsmaxOsmin)/ΔV and Pmax = ΔV2/4R, respectively, where F is Faraday constant, n is the mole number of Mn per unit mass, Osmax is the highest Mn oxidation state, Osmin is the lowest Mn oxidation state, ΔV is the potential range and R is the internal resistance of MnOx. Except for Mn valence and conductivity, tunnel size also determines the supercapacitive performance of MnOx by affecting the transportation and intercalation of electrolyte ions [12]. Since these aforementioned intrinsic material properties of MnOx are highly dependent on their crystal structure [13,14], the modulation of MnOx crystal structure is crucial to achieve ideal supercapacitive performance.

    Numerous efforts have been carried out to investigate the effect of MnOx crystal structure on their supercapacitive performance. For example, Devaraj and Munichandraiah [15] fabricated nanostructured α, β, γ, δ, λ-MnO2 and compared their supercapacitive performance. Among these MnO2 materials, the wide tunnel size and large surface area are key factors for α-MnO2 to realize the outstanding maximum specific capacitance of 297 F/g. Hu et al. [16] derived Mn3O4 with different crystal structures/morphologies by adjusting the hydrothermal temperature and studied their electrochemical properties. Compared with other Mn3O4 materials, orthorhombic Mn3O4 nanofibers showed a relatively high specific capacitance of 235 F/g at 1 mV/s and excellent cycling performance of 80% capacitance retention after 10,000 cycles. To date, various crystal forms of MnOx have been explored to some extent, the influence of MnOx crystal structure on the supercapacitive performance remains elusive. Furthermore, although MnOx have displayed high specific capacitance values, their limited capacitance retention (< 90%) after long cycle times (> 10,000 times) cannot meet the requirement of practical utilization [17]. Therefore, it is scientifically significant to optimize MnOx crystal structure to achieve high specific capacitance and great cycling stability simultaneously.

    Previously, our group have successfully synthesized MnO and Mn2O3 microspheres for efficient photo-Fenton removal of fluoroquinolone antibiotics through a facile hydrothermal method [18]. Inspired by this work, we reported a temperature-controlled methodology to obtain MnO2, Mn3O4, MnO and Mn2O3 microspheres with different crystal structure. The structure-activity relationship of MnOx microspheres was revealed by systematically investigating the effect of temperature on their valence state, specific surface area, conductivity and morphology, which are vital to their supercapacitive performance. Among these MnOx materials, MnO2 synthesized at lowest temperature realized the best supercapacitive performance with the following benefits: (1) The high valence state of MnO2 microspheres determines high theoretical charge storage capacity of each Mn active site; (2) The microporous/mesoporous structure with high specific surface area could provide abundant active sites for energy storage; (3) The relatively stable Mn-O bonds suppress Jahn-Teller effect to achieve desirable electronic conductivity; (4) The enhanced wettability of MnO2 microspheres during the cycling measurement results in an excellent cyclability. This work provides a deep insight of excellent electrochemical performance of temperature-dependent MnOx which could be potentially expanded to other metal oxides in energy storage applications.

    Fig. S1 (Supporting information) clearly shows the temperature-controlled synthesis route of MnOx microspheres. First of all, MnO2 microspheres were fabricated through a facile hydrothermal process at 90 ℃. The formation process of MnO2 was elucidated as MnSO4 + (NH4)2S2O8 + 2H2O → MnO2 + (NH4)2SO4 + 2H2SO4 [19]. After that, as-prepared MnO2 microspheres were calcined at 280, 500 and 600 ℃ under different atmospheres to obtain Mn3O4, MnO and Mn2O3 microspheres, respectively. Fabrication and characterization details of MnOx are described in Supporting information.

    X-ray diffraction (XRD) patterns are presented in Fig. S2 (Supporting information) to confirm the structure of as-prepared MnOx materials. The characteristic peaks of MnO2, Mn3O4, MnO and Mn2O3 matched well with standard PDF cards #44–0142, #24–0734, #07–0230 and #41–1442, respectively, implying the successful preparation of MnOx materials. Based on the strongest peaks from MnO2 (210), Mn3O4 (211), MnO (200) and Mn2O3 (222), the corresponding average particle size was estimated as 17.2, 12.7, 23.9 and 31.3 nm using the Scherrer equation [20]. As summarized in Fig. S3 (Supporting information), MnOx materials synthesized at high temperature possessed large particle size. MnO2 and Mn3O4 with the relatively small particle size and low crystallinity were expected to exhibit large surface area [21], which would be verified in the following porosity analysis.

    The surface Mn valence and porosity information of MnOx materials were analyzed by X-ray photoelectron spectroscopy (Fig. S4 in Supporting information) and N2 adsorption-desorption isotherms (Fig. S5 in Supporting information), respectively. As shown in Fig. S4a, survey spectra indicated the existence of Mn and O elements in MnOx materials except the reference C peak. According to high resolution Mn 2p spectra (Fig. S4b), two distinct peaks of MnO2 at 654.1 eV and 642.1 eV were assigned to Mn 2p1/2 and Mn 2p3/2, respectively, which are consistent with the reported values [22]. By decreasing valence state of Mn, the peaks of Mn 2p1/2 and Mn 2p3/2 gradually shifted to low binding energy. For MnO, the Mn 2p1/2 peak at 653.5 eV and Mn 2p3/2 peak at 641.4 eV revealed that the valence state of Mn was +2 [23]. Fig. S5a displays N2 adsorption-desorption isotherms of MnOx materials. The isotherms with the combination of type Ⅰ and type Ⅳ features demonstrated the presence of micropores and mesopores [24], which were confirmed by the pore size distribution (Fig. S5b). The pore size distribution of MnOx materials was mainly centered at around 1.0 and 3.6 nm. Consequently, the Brunauer-Emmett-Teller (BET) specific surface area of MnO2, Mn3O4, MnO and Mn2O3 were calculated as 113.4, 83.5, 57.4 and 76.9 m2/g, respectively. The relatively high specific surface area of MnO2 and Mn3O4 were in good agreement with XRD results. As summarized in Table S1 (Supporting information), MnO2 possessed the largest BET specific surface area and total pore volume, which provide plenty of actives sites for energy storage.

    Electronic conductivity of MnOx materials were investigated by four-point probe measurements. As displayed in Fig. 1a, the electronic conductivity values of MnO2, Mn3O4, MnO and Mn2O3 were calibrated as 1.73 × 10−5, 4.15 × 10−7, 1.79 × 10−7 and 5.89 × 10−7 S/cm, respectively. It is reported that the electronic conductivity of MnOx materials was affected by Jahn-Teller effect [25], which was also determined by the Mn-O bond length in MnO6 octahedra [26]. According to the comparison of average Mn-O bonding distance between MnOx materials (Fig. 1b), MnO2 possessed the shortest average Mn-O bond length, indicating that the disruptive Jahn-Teller distortion is effectively suppressed in the crystal structure of MnO2 [26]. Therefore, MnO2 with high structural stability achieved desirable electronic conductivity, which is beneficial for electron transfer in energy storage processes.

    Figure 1

    Figure 1.  (a) Conductivity values of MnOx materials. (b) The comparison of average Mn-O bonding distances between MnOx materials.

    Transmission electron microscopy (TEM) was used to study morphologies and microstructure of MnOx materials. All MnOx materials showed spherical morphologies with an average diameter of approximately 2 µm (Figs. 2a, d, g and j), indicating that the calcination temperature has negligible effects on the size of porous MnOx microspheres. The magnified TEM image in Figs. 2b illustrated that MnO2 microspheres were assembled by slender nanoneedles. The high-resolution transmission electron microscopy (HR-TEM) image (Fig. 2c) on these nanoneedles clearly displayed well-defined lattice fringes with d-spacing of 0.24 nm, which is attributed to the crystal plane of (210) from MnO2. By increasing calcination temperature, orthorhombic MnO2 were phase-transformed to tetragonal Mn3O4, cubic MnO and cubic Mn2O3. Accordingly, the assembly unit of MnOx microspheres was changed from nanoneedles to nanowires (Fig. 2e), nanorods (Fig. 2h) and large-size irregular nanorods (Fig. 2k). Furthermore, distinct lattice fringes on these assembly units were observed from HR-TEM images (Figs. 2f, i and l) to further confirm the formation of Mn3O4, MnO and Mn2O3, respectively.

    Figure 2

    Figure 2.  TEM images of (a, b) MnO2, (d, e) Mn3O4, (g, h) MnO and (j, k) Mn2O3. HR-TEM images of (c) MnO2, (f) Mn3O4, (i) MnO and (l) Mn2O3.

    The supercapacitive performance of MnOx electrodes were investigated by cyclic voltammetry (CV), galvanostatic charge/discharge (GCD), electrochemical impedance spectroscopy (EIS) and cycling measurements. In Fig. 3a, CV curves of MnOx electrodes displayed distinct redox peaks, which originate from the reaction between MnOx and OH in alkaline electrolyte [27]:

    Figure 3

    Figure 3.  (a) CV curves of MnOx electrodes at 50 mV/s. (b) CV curves of MnO2 electrode at scan rates of 5, 10, 20, 30 and 50 mV/s. (c) Contribution ratio of capacitive and diffusion-controlled mechanism at different scan rates for the MnO2 electrode. (d) Specific capacitance values of MnOx electrodes at different specific currents. (e) Cycle stability of MnO2 electrode at 10 A/g (inset: GCD curves before and after cycling test).

    With the increase of Mn valence state, the positions of anodic peaks shifted to relatively high potential. Compared to other MnOx electrodes, the CV curve of the MnO2 electrode explicitly showed two pairs of redox peaks with highest specific current values, indicating the greatest charge storage ability. Afterwards, CV curves of MnO2 electrode from 5 mV/s to 50 mV/s (Fig. 3b) showed an unchangeable shape with negligible ohmic polarization, demonstrating an excellent electrochemical reversibility. Subsequently, the electrochemical kinetics of MnOx electrodes were studied according to the power-law relationship between the cathodic peak current (i) and the scan rate (ν): i = b [28]. The b-value of 0.5 and 1.0 correspond to the Faradaic diffusion-controlled and capacitive mechanism, respectively. By fitting in Fig. S6 (Supporting information), the b-value were derived as approximately 0.6, denoting that the charge storage mechanism of MnOx electrodes are mainly controlled by diffusion-controlled process. Specifically, the contribution of diffusion-controlled and capacitive mechanism were separated by applying the formula: i = k1ν + k2ν0.5 [28]. For the MnO2 electrode (Fig. 3c), the diffusion-controlled contribution was calculated as 77.6% at 5 mV/s, which is in accordance with the b-value analysis. With the increasing scan rate, the diffusion-controlled contribution decreased to 52.7% at 50 mV/s due to the inherent nature of battery-like redox reactions [29].

    Fig. S7 (Supporting information) presents GCD curves of MnOx electrodes at 2 A/g. The nonlinear shape of these GCD curves represented the faradaic processes in MnOx electrodes, which are in good agreement with above CV curves. On the basis of GCD curves at various specific currents (Fig. S8 in Supporting information), the specific capacitance values of MnOx electrodes were summarized in Fig. 3d. Among these MnOx electrodes, the MnO2 electrode realized the highest specific capacitance of 274.1 F/g, which is comparable or superior to relevant reported publications (Table S2 in Supporting information). Furthermore, the comparison of Nyquist plots of MnOx electrodes in Fig. S9 (Supporting information) was consistent with previous electrochemical analyses. After fitting the impedance data with an equivalent circuit (Table S3 in Supporting information), the MnO2 electrode exhibited the relatively low equivalent series resistance RESR (0.36 Ω) and charge transfer resistance Rct (8.29 Ω), which correspond to high electronic conductivity and rapid electrolyte diffusion into the micro/mesopores of MnO2 microspheres [30].

    Based on above mentioned material and electrochemical characterizations, the enhanced supercapacitive performance of MnO2 electrode can be attributed to following reasons: (1) The high Mn valence state of MnO2 microspheres enabled more hydroxide ions to interact with each Mn active site; (2) MnO2 microspheres with the high specific surface area (113.4 m2/g) and unique microporous/mesoporous structure provided abundant Mn active sites; (3) MnO2 with minimized Jahn-Teller distortion realized high electronic conductivity (1.73 × 10−5 S/cm), which is conducive to the fast electron transport for electrochemical kinetics.

    The cycling performance of MnO2 electrode at 10 A/g is illustrated in Fig. 3e. It is interesting that the capacitance retention of MnO2 electrode increased to 115% after 10,000 cycles. Electrochemical characterizations after cycling test were carried out to elucidate the unpredictable enhancement. Fig. S10 (Supporting information) displays CV curves of MnO2 electrode at initial and after cycle states. Compared to the initial state, the CV curve of cycled MnO2 electrode showed a pair of redox peaks with greater specific current at higher potential, illuminating that the charge storage processes became more preferential to proceed between relatively higher valence states of Mn. Meanwhile, GCD curves with higher charge/discharge platform in Fig. S11 (Supporting information) presented the same tendency. Furthermore, according to the b-value calculation from Figs. S12 and S13a (Supporting information), the charge storage mechanism changes from diffusion-controlled process to diffusion/capacitive-controlled processes during the cycling measurement. Concretely, as depicted in Fig. S13b (Supporting information), the capacitive contribution of MnO2 electrode was increased by 14.0%, 15.2%, 16.8%, 17.1% and 16.6% after 10,000 cycles at scan rates of 5, 10, 20, 30 and 50 mV/s, respectively. In other words, fast Faradaic redox reactions were proceeded at MnO2 electrode after long-term GCD measurements. Consequently, the cycled MnO2 electrode possessed much higher rate capability (75.2%) than the initial MnO2 electrode (47.4%) and retained the maximum specific capacitance, which were derived from Fig. S14 (Supporting information).

    Structural and morphological characterizations including XRD, SEM and contact angle measurements were performed to explain the enhanced rate capability of MnO2 electrode after cycling test. From XRD results in Fig. 4a, the appearance of characteristic peaks of Mn3O4 indicated the phase transformation from MnO2 to MnO2/Mn3O4 during the cycling test. Meanwhile, new characteristic peaks at around 40° were observed and attributed to MnO2 (111) and (002), revealing the exposure of more Mn active sites. Although the specific capacitance of Mn3O4 was lower than that of MnO2 at 1 A/g, the competing effect of phase transformation and increased active sites led to the invariant maximum specific capacitance of MnO2 electrode. Figs. 4b and c show SEM images of MnO2 electrode at initial and after cycle states. After cycling test, the spherical morphology of MnO2 materials was almost unchanged without any agglomeration. Afterwards, the wettability of initial and cycled MnO2 electrode was distinctly distinguished in Figs. 4d and e, respectively. In the 20 min measurement period, the contact angle of initial MnO2 electrode was kept constant at 120° whereas the contact angle of cycled MnO2 electrode decreased from 80° to 40°. According to Young's equation, the MnO2 electrode was changed from hydrophobicity to hydrophilicity with the electrolyte in cycling process. The enhanced wettability improves the interaction between electrolyte ions and Mn active sites, which enhances utilization of active sites of MnO2 electrode [31,32]. Moreover, another contrast experiment was designed to further verify the high wettability of cycled MnO2 electrode. After soaking the cycled MnO2 electrode in the electrolyte for 1 day, MnO2 spheres covered with shells were observed in Fig. S15 (Supporting information). Compared to EDS elemental mappings of initial MnO2 electrode (Figs. S16 and S17 in Supporting information), the shell was mainly composed of potassium ions, revealing that MnO2 spheres in the cycled electrode are more hydrophilic to the electrolyte. Therefore, the enhanced specific capacitance of MnO2 electrode during cycling test is mainly attributed to the intimate relationship between the electrode material and electrolyte.

    Figure 4

    Figure 4.  (a) XRD patterns of MnO2 electrode before and after cycling test. SEM images of MnO2 electrode (b) before and (c) after cycling test. Contact angle measurements with 6 mol/L KOH electrolyte of MnO2 electrode (d) before and (e) after cycling test at 0, 1 and 20 min.

    In conclusion, we have successfully fabricated MnO2, Mn3O4, MnO and Mn2O3 microspheres through a facile temperature-controlled method. Compared with other MnOx microspheres, MnO2 realized a higher Mn valence state of +4, a larger specific surface area of 113.4 m2/g and a greater electronic conductivity of 1.73 × 10−5 S/cm, which are beneficial for enhancing the supercapacitive performance. Accordingly, MnO2 possessed a relatively high specific capacitance of 274.1 F/g at 1 A/g and showed a remarkable capacitance retention of 115% after 10,000 cycles. This work provides an important guidance for understanding and optimizing the supercapacitive performance of temperature-dependent MnOx.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    This work was supported by Key-Area Research and Development Program of Guangdong Province (No. 2019B110209003), Guangdong Basic and Applied Basic Research Foundation (No. 2019B1515120058), the Scientific and Technological Planning Project of Guangzhou, China (No. 202206010145), National Natural Science Foundation of China (No. 22078374), National Key R & D Program of China (No. 2020YFC1807600), National Ten Thousand Talent Plan and Hundred Talent Plan (No. 201602) from Sun Yat-sen University.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2022.05.052.


    1. [1]

      L. Lin, J. Chen, D. Liu, et al., Adv. Energy Mater. 10 (2020) 2002621. doi: 10.1002/aenm.202002621

    2. [2]

      V. Dusastre, L. Martiradonna, Nat. Mater. 16 (2017) 15. doi: 10.1038/nmat4838

    3. [3]

      J. Feng, X. Zhang, Y. Xu, et al., Energy Storage Mater. 46 (2022) 278-288. doi: 10.1016/j.ensm.2022.01.024

    4. [4]

      C. Liu, Q. Li, K. Wang, Renew. Sustain. Energy Rev. 150 (2021) 111408. doi: 10.1016/j.rser.2021.111408

    5. [5]

      Y. Zhou, H. Qi, J. Yang, et al., Energy Environ. Sci. 14 (2021) 1854-1896. doi: 10.1039/d0ee03167d

    6. [6]

      K. Li, X. Liu, T. Zheng, et al., Chem. Eng. J. 370 (2019) 136-147. doi: 10.1016/j.cej.2019.03.190

    7. [7]

      A. Zhang, R. Gao, L. Hu, et al., Chem. Eng. J. 417 (2021) 129186. doi: 10.1016/j.cej.2021.129186

    8. [8]

      L. Shi, Y. Wang, P. Zou, et al., Chin. Chem. Lett. 29 (2018) 592-595. doi: 10.1016/j.cclet.2018.01.024

    9. [9]

      J. Kang, A. Hirata, L. Kang, et al., Angew. Chem. Int. Ed. 52 (2013) 1664-1667. doi: 10.1002/anie.201208993

    10. [10]

      P. Guillemet, T. Brousse, O. Crosnier, et al., Electrochim. Acta 67 (2012) 41-49. doi: 10.1016/j.electacta.2012.01.110

    11. [11]

      Y. Wu, C. Cao, Sci. China Mater. 61 (2018) 1517-1526. doi: 10.1007/s40843-018-9290-y

    12. [12]

      Y. Hu, Y. Wu, J. Wang, Adv. Mater. 30 (2018) 1802569. doi: 10.1002/adma.201802569

    13. [13]

      Y. Liu, X. Wu, Chin. Chem. Lett. 33 (2022) 1236-1244. doi: 10.1016/j.cclet.2021.08.081

    14. [14]

      L. Wei, Z. Wang, Y. Liu, et al., J. Hazard. Mater. 416 (2021) 126117. doi: 10.1016/j.jhazmat.2021.126117

    15. [15]

      S. Devaraj, N. Munichandraiah, J. Phys. Chem. C 112 (2008) 4406-4417. doi: 10.1021/jp7108785

    16. [16]

      Y. Hu, Y. Zhang, D. Yuan, et al., Nanoscale Horiz. 2 (2017) 326-332. doi: 10.1039/C7NH00078B

    17. [17]

      A. Zhang, R. Zhao, L. Hu, et al., Adv. Energy Mater. 11 (2021) 2101412. doi: 10.1002/aenm.202101412

    18. [18]

      A. Wang, H. Wang, H. Deng, et al., Appl. Catal. B Environ. 248 (2019) 298-308. doi: 10.1016/j.apcatb.2019.02.034

    19. [19]

      N. Li, X. Zhu, C. Zhang, et al., J. Alloy. Compd. 692 (2017) 26-33. doi: 10.1016/j.jallcom.2016.08.321

    20. [20]

      D. Gangwar, C. Rath, Phys. Chem. Chem. Phys. 22 (2020) 14236-14245. doi: 10.1039/d0cp02245d

    21. [21]

      T. Falk, S. Anke, H. Hajiyani, et al., Catal. Sci. Technol. 11 (2021) 7552-7562. doi: 10.1039/d1cy00944c

    22. [22]

      L. Mo, H. Zheng, J. Alloy. Compd. 788 (2019) 1162-1168. doi: 10.1016/j.jallcom.2019.02.321

    23. [23]

      Z. Cui, Q. Liu, C. Xu, et al., J. Mater. Chem. A 5 (2017) 21699-21708. doi: 10.1039/C7TA05986H

    24. [24]

      S. Liu, Y. Liang, W. Zhou, et al., J. Mater. Chem. A 6 (2018) 12046-12055. doi: 10.1039/C8TA02838A

    25. [25]

      L. Liu, Y. Luo, W. Tan, et al., J. Colloid Interface Sci. 482 (2016) 183-192. doi: 10.1016/j.jcis.2016.07.077

    26. [26]

      J.U. Choi, J. Kim, J.Y. Hwang, et al., Nano Energy 61 (2019) 284-294. doi: 10.1016/j.nanoen.2019.04.062

    27. [27]

      Q. Zhang, M.D. Levi, Q. Dou, et al., Adv. Energy Mater. 9 (2019) 1802707. doi: 10.1002/aenm.201802707

    28. [28]

      Y. Mao, Y. Chen, J. Qin, et al., Nano Energy 58 (2019) 192-201. doi: 10.1016/j.nanoen.2019.01.048

    29. [29]

      S.M. Ingole, S.T. Navale, Y.H. Navale, et al., J. Solid State Electrochem. 21 (2017) 1817-1826. doi: 10.1007/s10008-017-3557-8

    30. [30]

      G. Zhang, L. Ren, Z. Yan, et al., Chem. Commun. 53 (2017) 2950-2953. doi: 10.1039/C6CC10250F

    31. [31]

      Y. Wang, Y. Liu, Z. Chen, et al., Green Chem. Eng. 3 (2022) 55-63. doi: 10.1117/12.2645475

    32. [32]

      Y. Zhang, M. Park, H.Y. Kim, S.J. Park, J. Colloid Interface. Sci. 500 (2017) 155-163. doi: 10.1016/j.jcis.2017.04.022

  • Figure 1  (a) Conductivity values of MnOx materials. (b) The comparison of average Mn-O bonding distances between MnOx materials.

    Figure 2  TEM images of (a, b) MnO2, (d, e) Mn3O4, (g, h) MnO and (j, k) Mn2O3. HR-TEM images of (c) MnO2, (f) Mn3O4, (i) MnO and (l) Mn2O3.

    Figure 3  (a) CV curves of MnOx electrodes at 50 mV/s. (b) CV curves of MnO2 electrode at scan rates of 5, 10, 20, 30 and 50 mV/s. (c) Contribution ratio of capacitive and diffusion-controlled mechanism at different scan rates for the MnO2 electrode. (d) Specific capacitance values of MnOx electrodes at different specific currents. (e) Cycle stability of MnO2 electrode at 10 A/g (inset: GCD curves before and after cycling test).

    Figure 4  (a) XRD patterns of MnO2 electrode before and after cycling test. SEM images of MnO2 electrode (b) before and (c) after cycling test. Contact angle measurements with 6 mol/L KOH electrolyte of MnO2 electrode (d) before and (e) after cycling test at 0, 1 and 20 min.

  • 加载中
计量
  • PDF下载量:  11
  • 文章访问数:  1131
  • HTML全文浏览量:  41
文章相关
  • 发布日期:  2023-05-15
  • 收稿日期:  2022-05-09
  • 接受日期:  2022-05-17
  • 网络出版日期:  2022-05-20
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章