Next Article in Journal
Recent Developments on 1,3-Dipolar Cycloaddition Reactions by Catalysis in Green Solvents
Previous Article in Journal
The Influence of the Support on the Activity of Mn–Fe Catalysts Used for the Selective Catalytic Reduction of NOx with Ammonia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Manganese Oxide Nanorods Decorated Table Sugar Derived Carbon as Efficient Bifunctional Catalyst in Rechargeable Zn-Air Batteries

1
Material Science and Engineering Research Group, Faculty of Mechanical and Aerospace Engineering, Institut Teknologi Bandung, Jl. Ganesha 10, Bandung 40132, Indonesia
2
Department of Materials Science and Engineering, National University of Singapore, Singapore 117574, Singapore
3
Institute of Materials Research and Engineering (IMRE), A*STAR (Agency for Science, Technology and Research), 2 Fusionopolis Way, Innovis #08-03, Singapore 138634, Singapore
4
Research Center for Nanoscience and Nanotechnology, Institut Teknologi Bandung, Jl. Ganesha 10, Bandung 40132, Indonesia
5
National Center for Sustainable Transportation Technology, Institut Teknologi Bandung, Jl. Ganesha 10, Bandung 40132, Indonesia
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(1), 64; https://doi.org/10.3390/catal10010064
Submission received: 10 December 2019 / Revised: 24 December 2019 / Accepted: 30 December 2019 / Published: 1 January 2020
(This article belongs to the Section Electrocatalysis)

Abstract

:
Despite its commercial success as a primary battery, Zn-air battery is struggling to sustain a reasonable cycling performance mainly because of the lack of robust bifunctional electrocatalysts which smoothen the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) taking place on its air-cathode. Composites of carbon/manganese oxide have emerged as a potential solution with high catalytic performance; however, the use of non-renewable carbon sources with tedious and non-scalable synthetic methods notably compromised the merit of being low cost. In this work, high quantity of carbon is produced from renewable source of readily available table sugar by a facile room temperature dehydration process, on which manganese oxide nanorods are grown to yield an electrocatalyst of MnOx@AC-S with high oxygen bifunctional catalytic activities. A Zn-air battery with the MnOx@AC-S composite catalyst in its air-cathode delivers a peak power density of 116 mW cm−2 and relatively stable cycling performance over 215 discharge and charge cycles. With decent performance and high synthetic yield achieved for the MnOx@AC-S catalyst form a renewable source, this research sheds light on the advancement of low-cost yet efficient electrocatalyst for the industrialization of rechargeable Zn-air battery.

Graphical Abstract

1. Introduction

Energy storage devices are expected to store more energy in the given mass or volume, a continuous push in acquiring high specific energy or energy density. Zn-air batteries have gained special attention because of the higher specific energy as compared to the most of the currently available batteries [1]. For instance, its specific energy, 1086 Wh kg−1, is more than doubled than that of Li-ion batteries, 460 Wh kg−1 [2]. In addition, the good safety and low production cost of Zn-air batteries make them a favorable choice for commercial applications [3]. Nevertheless, it remains challenging to extend the cycle life of Zn-air batteries when recharged electrically [4]. The root cause here is mainly attributed to the sluggish kinetics of oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) at the air cathode, hindering the development of rechargeable Zn-air batteries, particularly for heavy duty purposes [5,6].
Noble-metal catalysts, particularly platinum-based catalysts, are the most widely used ORR catalysts because of its excellent performance in speeding up the kinetics of ORR. Despite that, platinum-based catalysts are expensive, rare, and may contribute to the short lifetime of rechargeable of Zn-air batteries due to its relatively poor performance in catalyzing the OER [7]. Alternatively, manganese oxides have become an object of interest because of its favorable catalytic activity, low cost, abundance and eco-friendliness [8,9]. However, powder of manganese oxides shows low catalytic performance on the air cathode of Zn-air battery because of its poor electrical conductivity [10]. Thus, conductive carbon supports are often incorporated to improve electronic conduction in the catalytic air-electrode [10,11,12]. Furthermore, it has been shown that the growing of manganese oxides on the carbon supports was able to further improve the stability and power density of the Zn-air batteries [4,10,13]. “Advanced” carbons such as graphene [14,15] or carbon nanotube [16] are often the substrate of choices, raising cost-concerns for implementation with the need of large-scale production. Additionally, carbon source for mass production of carbon black mainly originates from petroleum combustion process, which are deemed to be unclean and non-renewable. In response, research had begun to utilize biomass as a renewable and low-cost carbon source, including hydrocarbon-rich biomasses such as eggplants [17], coconuts [18], soybeans [19], and peanut shells [20], etc., which exhibited high catalytic activity for ORR and OER upon their carbonization into amorphous carbon.
Pyrolysis is one of the most commonly used methods for producing amorphous carbon from biomass, conducted at elevated temperature under inert atmosphere [21]. In contrast, dehydration is a facile method to produce carbon that is less explored [22]. In a dehydration process producing carbons, water molecules are removed from the substrate molecules (e.g., sucrose) in the form of steam by autogenous heat, leaving behind a black mass of carbon with high porosity. For instance, the dehydration reaction with concentrated sulfuric acid (H2SO4) gives amorphous carbon, as demonstrated in a carbon black snake experiment [22]. However, its application as electrocatalyst, particularly for metal-air batteries, is yet to be studied [21].
Herein, we demonstrate dehydration as a facile method for producing amorphous carbon from biomass. Table sugar is chosen as the source of biomass because of its availability and high carbon content. Previously, table sugar has been used as carbon source using pyrolysis method to obtain N-doped graphene and N-doped CNT [23,24]. Synthesis of carbon by dehydration has been demonstrated to fabricate an electrode for supercapacitor with promising electrochemical performance because of its high degree of porosities and specific surface area [25]. The growth of manganese oxides on the carbon surface upon the chemical activation of dehydrated sugar derived carbon are expected to increase the contact area and facilitate the electron transfer at manganese oxide/carbon interface [4,26]. Electrocatalyst fabricated from such method is assembled in a rechargeable Zn-air battery showing a peak power density of 116 mW cm−2 and a stable charge-discharge performance for more than 215 cycles. These results are comparable to the Zn-air batteries using similar catalysts from notably more tedious and expensive processes [4,13,27,28,29,30].

2. Results and Discussion

2.1. Morphological and Structural Characterization

Chemical activation is often required to enlarge the surface area of the carbon [31]. Representative SEM images of the carbon synthesized from dehydration of sugar before and after the chemical activation are shown in Figure 1a,b. Difference of morphology can be clearly seen from the comparison between the non-activated (C-S) and activated carbon-derived sugar (AC-S). The porous structure of AC-S (Figure 1b) which is further confirmed by TEM imaging in Figure 2a indicates that the activation process is effective. High production yield of activated carbon at around 22% of original mass of the table sugar can be obtained. Change in the surface condition is also observed upon the introduction of manganese oxide on the prepared activated carbon (MnOx@AC-S). Significant presence of nanostructures deposited on the carbon surface can be seen during SEM imaging (Figure 1c), which is absent in the samples without the addition of manganese oxides.
TEM image shows the detailed morphology of the deposited nanoparticles in the form of nanorod structure (Figure 2b). This is agreeable with a previous report where a similar co-precipitation method was used to yield nanorod-structured MnO2 [32]. Both SEM and TEM images clearly suggest that the deposited nanoparticles are not agglomerated. In addition, they are distributed across the surface of carbon materials. This observation is supported by the EDS mapping, showing uniform distribution of manganese throughout the sample (Figure S1). Such morphology and distribution prevent blocking the access to the pores and conductive surface of the carbon support. Consequently, the diffusions of oxygen and electrolyte into the active sites are facilitated, with the large conductive surface area providing a facile charge transfer at the catalyst/electrolyte interfaces.
XRD patterns of both samples are shown in Figure 3a. The broad peaks at 12–37° and 37–50° are the amorphous carbon and associated to the (002) and (100) of its planes [33,34]. The presence of these two peaks in the XRD patterns of the samples before and after manganese oxide deposition suggests that no carbon structure change occurs during the manganese oxide growth. The additional peaks at 36–40° and 65–68° for MnOx@AC-S approves the presence of α-MnO2 [32,35]. These peaks are broad too, indicating poor crystalline/amorphous nature of α-MnO2 on the carbon surface to make a composite catalyst with rich catalytic active sites.
The surface elemental composition of MnOx@AC-S sample was probed by X-ray photoelectron spectroscopy (Figure 3b,c). The Mn2p spectrum shows two prominent peaks for the spin orbit doublet of Mn 2p3/2 and Mn 2p1/2, respectively (Figure 3b) [36]. The prominent peak of Mn 2p1/2 is centered at 653.75 eV, while the Mn 2p3/2 peak can be deconvoluted into three peaks located at 641.63, 643.31, and 646.22 eV. Those three peaks are attributed to the presence of Mn(III), Mn(IV), and Mn(II) oxide phase, respectively [37,38,39]. The presence of MnOx in our sample is further confirmed by the obtained O1s signal. The O1s spectrum can be deconvoluted into three peaks (Figure 3c) located at 530.20, 531.53, and 532.50 eV, corresponding to O-Mn, O-C, and O-H bond, respectively [37,40]. The presence of hydrate in manganese oxides has been reported to improve the diffusion of electrolyte ions and the ORR performance of the catalyst [41,42].

2.2. Electrocatalytic Activity and Performance of Zn-air Batteries

Catalytic activity of AC-S and MnOx@AC-S during oxygen reduction reaction (ORR) in 0.1 M KOH was assessed by rotating disk electrode (RDE) method in a three-electrode setup. Linear scan voltammetry (LSV) during ORR are constructed to obtain the onset potential and limiting current density (Figure 4a). The LSV scans were also performed on commercial Pt/C and Vulcan XC-72 carbon black as the benchmark for the ORR catalyst. MnOx@AC-S sample shows an onset potential of 0.914 V, which is 0.1 V lower compared to Pt/C (1.035 V) but higher than that of AC-S (0.854 V) and Vulcan XC-72 carbon black (0.817 V). The onset potential is defined as overpotential at a current density of 0.1 mA cm−2 [9]. The relatively early onset potential and high limiting current density of AC-S and MnOx@AC-S indicate the high ORR catalytic activities of the activated carbon obtained from dehydration process [43,44]. Samples with such activated carbons show notably higher limiting current density than that of commercial carbon black, and they are comparable to that of Pt/C.
The catalytic activity of the catalysts was further studied under various rotation rate (Figure 4b,c). The catalytic mechanism during ORR is examined using Koutecky–Levich plots. The plots were constructed at several potentials at various rotating speeds. The number of electron transfer per oxygen molecule during ORR can be obtained from the least-square-fitted slopes on the basis of Koutecky–Levich equation. From the calculation, Pt/C and MnOx@AC-S catalysts have an electron transfer number of 4.00, indicating both catalysts undergo the desirable four electron process during the reduction of oxygen. In the absence of MnOx, AC-S, and Vulcan XC-72 carbon black catalysts show a small number of electron transfer (3.12 and 2.83, respectively), implying a two-electron pathway with a slower kinetics and production of detrimental peroxide species.
The RDE technique was also employed to examine the catalytic activity of the catalyst during oxygen evolution reaction (OER) in 1 M KOH. An increase in OER catalytic activity of MnOx@AC-S sample as compared to Pt/C, AC-S and Vulcan XC-72 carbon black are shown in Figure 4d. Despite having excellent ORR catalytic property, Pt tends to have low OER activity because of the formation of oxide layer that reduces its overall conductivity [2]. At current density of 10 mA cm−2, MnOx@AC-S catalyst exhibits overpotential of 500 mV, 38 and 80 mV lower than that of AC-S and Pt/C catalyst, respectively. The superiority of MnOx@AC-S can be attributed to the presence of amorphous manganese oxide which had been reported to exhibit better OER activity than that of Pt/C and the pristine carbon catalysts [4,7].
Potential application of AC-S and MnOx@AC-S as bifunctional catalyst for both ORR and OER was demonstrated in rechargeable Zn-air batteries. The catalyst ink was coated on carbon paper and employed as an air cathode in rechargeable Zn-air batteries. The performance of the Zn-air batteries with AC-S and MnOx@AC-S catalysts is compared to the batteries with Pt/C catalyst. According to the polarization curve in Figure 5a, lower discharge performance is obtained from the batteries with MnOx@AC-S catalyst as compared to the battery with Pt/C catalyst, but it is better than that of AC-S catalyst, matching the trend observed in RDE study. At a discharge voltage of 1 V, the Zn-air battery with MnOx@AC-S catalyst is able to deliver current density up to 81 mA cm−1, about 1.5 times increase of the current density as compared to the battery with AC-S catalyst. Peak power density was measured by extracting the data of galvanodynamic discharge experiment. Zn-air battery with Pt/C, MnOx@AC-S, and AC-S catalysts delivers a power density of 126, 116, and 105 mW cm−2, respectively. The battery with MnOx@AC-S catalyst has an improved peak power density as compared to the reported Zn-air batteries with MnOx catalyst directly grown on carbon paper [4], MnO2/LaNiO3/CNT [13], MnO2/eggplant derived activated carbon [45] and MnO2 nanotube [30] (Supplementary Materials Table S1). An increase in charging performance of Zn-air battery using MnOx@AC-S catalyst was observed as compared to the other prepared batteries. At a charging potential of 2 V, the battery was able to accept current density of 40 mA cm−2. This value is higher than that of Pt/C and AC-S (both at 33 mA cm−2) or commercial carbon black (29 mA cm−2). This number is also better than the Zn-air batteries with MnO2/N-CNT catalyst that had high crystallinity of manganese oxide [29]. Poorly crystalline/amorphous manganese oxide has been reported to exhibit an improved catalytic activity than the high crystalline forms [7,46]. Despite that, the Zn-air battery with MnOx@AC-S shows a higher charge voltage as compared to the battery with Pt/C+RuO2 catalyst, which can be attributed to the superior OER performance of RuO2 [47].
EIS measurement was performed at voltage of 0.9 V to evaluate the resistance of the cells during discharging. Nyquist plots were constructed out of those EIS measurements, in which solution resistance (Rs), charge transfer resistance (Rct), and resistance at the electrode/electrolyte interface (Rint) can be evaluated. All batteries show similar Rs which can be depicted from the intercept at the X axis (3.5–3.6 Ω). However, variation in the diameter of semicircle on the Nyquist plot of each sample indicates the difference in the Rint and Rct for each battery. Accordingly, the Rint and Rct value of the battery with Pt/C catalyst is the lowest, followed by the battery with MnOx@AC-S and AC-S catalyst respectively. The low Rint and Rct of the cell with Pt/C catalyst is attributed to the high ORR activity of Pt/C which is the key process to improve the discharge performance of Zn-air batteries.
The stability of Zn-air batteries with MnOx@AC-S, AC-S and Pt/C during continuous discharge and charge processes was evaluated by recurrent galvanostatic pulses test. The test was performed at current density of 10 mA cm−2 for the period of 20 min (Figure 5d). The Zn-air battery with MnOx@AC-S catalyst shows a promising stability and rechargeability as compared to its counterparts with Pt/C and AC-S catalysts. Our previous work had opted for using lower charging current because of the poor OER activity of amorphous manganese oxide [4]. However, the cyclability obtained from this experiment is still in line with previous report despite using higher charging current. Cell with MnOx@AC-S catalyst shows an initial discharge and charge voltage of 1.13 V and 2.23 V respectively, which is better than that of AC-S catalyst (1.05 V and 2.29 V). The large overpotential during charging and discharging of the battery with AC-S catalyst results in the poor performance of the cell during the stability test. Especially, the high overpotential during charging leads to a large positive voltage that is running through the carbon paper. This process gradually corrodes the air cathode and oxidize the electrocatalyst which results in the low stability of the battery.
Despite having a good initial discharge and charge voltage, the cell with Pt/C catalyst does not exhibit a stable discharge/charge voltage during the stability test. The cell with Pt/C catalyst ceases its operation at 141 cycles because of the low stability of Pt/C catalyst over repeated charging. Previously, Pt/C has been reported to form passivation layer under positive voltage, leading to the formation of poorly conductive oxide on its surface and reducing the overall conductivity [2]. On the other hand, the cell with MnOx@AC-S catalyst exhibits the highest cyclability (up to 215 cycles) out of all prepared batteries. The improved stability of the cell with MnOx@AC-S catalyst can be attributed to its superior OER activity as compared to Pt/C and AC-S catalysts. In addition, manganese oxides with firm attachment to its carbon support had been reported for good stability as bifunctional catalyst in rechargeable Zn-air batteries [4].

3. Materials and Methods

3.1. Synthesis of Sugar-Derived Carbon Decorated Manganese Oxide Nanorods

Sucrose was obtained in the form of common white table sugar. Concentrated sulfuric acid (H2SO4, 95–98%, Pudak Scientific, Bandung, Indonesia) was used as the dehydrating agent. 40 g of table sugar was put into a beaker, followed by pouring of H2SO4 into the beaker until the liquid covers all of the sugar. The mixture was stirred until the sugar turned to a black solid carbon. The resulting carbon is then left to cool naturally for 3 h and washed with distilled water to remove any lingering acid on the carbon surface. The carbon was then grinded, washed again with distilled water, and then left to dry in a 60 °C oven. The chemical activation of the sugar-derived carbon was performed as the following. Total of 4.5 g of the carbon powder was immersed into 40 mL of 6 M potassium hydroxide (KOH, Merck, Darmstadt, Germany), and left naturally for 24 h in a beaker glass. The mixture was then filtered and dried in an oven at temperature of 60 °C. The dried carbon was then heat-treated in N2 environment at 800 °C for 1 h with a ramp of 7 °C min−1. To neutralize the lingering unreacted KOH, 2 M of hydrochloric acid (HCl, 37%, Honeywell, Bangkok, Thailand) and distilled water were used to wash the carbon. The resulting powder was then dried in an oven at a temperature of 60 °C for 1 h. The sample is denoted as AC-S.
Manganese oxides were introduced onto the activated sugar carbon by using the following method. Total of 0.25 g of AC-S was mixed with 10 mL of 0.05 M sulfuric acid and 10 mL of 0.05 M potassium permanganate (KMnO4, Merck, Darmstadt, Germany. The mixture was stirred for 30 min, until the color changes from dark purple to dark brown. It was then washed with distilled water and left in a 150 °C oven for 24 h. This sample is denoted as MnOx@AC-S.

3.2. Material and Electrochemical Characterizations

The crystal structure of manganese oxide was studied using an X-ray diffractometer (Bruker D8 Advance, Massachusetts, United States). The morphology was studied using SEM equipped with EDS detector (SU3500 Hitachi) and TEM (HT7700 Hitachi, Tokyo, Japan). XPS data was collected on a Thermo Scientific VG ES-CALAB 200i-XL spectrometer (Massachusetts, United States) with monochromatized Al Kα (hλ = 1484.6 eV).
The electrocatalytic activities of catalysts toward the ORR and OER were assessed with a three-electrode half-cell setup comprising a glassy carbon rotating disk electrode (Metroohm RDE, diameter: 5 mm), platinum counter electrode and Ag/AgCl (3 M KCl) reference electrode which were connected to a potentiostat (Metroohm Autolab PGSTAT302 N). For each catalyst, the ink was made by dispersing the powder of catalyst (9 mg) and Vulcan XC-72 carbon black (2.25 mg) into H2O:IPA:Nafion (ratio: 2.5:1.0:0.094) solution. The ink was drop-casted onto the glassy carbon RDE to obtain the mass loading of 0.2 mg cm−2. The ORR properties of the catalyst was investigated in 0.1 M KOH with continuous purging of O2 gas, while the OER activity was measured in 1 M KOH with continuous purging of N2 gas. Linear sweeping voltammetry (LSV) at scan rate of 5 mV s−1 was used in both studies. Potential scale was calibrated to reversible hydrogen electrode (RHE). The number of electron transferred during ORR, (n) can be extracted from the slope of Koutecky–Levich plots on the basis of Koutecky–Levich equation (Equations (1) and (2)):
1/i = 1/iK + 1/idl = 1/iK + 1/0.62nFACO2DO22/3v−1/6ω1/2
Slope = (0.62nFACO2DO22/3v−1/6)−1
where F is the Faraday constant, A is the geometric area of the glassy carbon electrode, CO2 is the concentration of the dissolved oxygen in the electrolyte at partial pressure of 1 atm, DO2 is the diffusion coefficient of oxygen in the electrolyte, v is the kinematic viscosity of electrolyte, DO2, CO2, and v in 0.1 M KOH are about 1.88 × 10−5 cm2 s−1, 1.21 × 10−6 mol cm−3, and 0.01 cm2 s−1, respectively [48].
Zn-air battery was assembled by using a homemade Zn-air cell [49]. A polished zinc plate was used as anode. The electrocatalysts were drop-casted onto a 2 × 2 cm2 Teflon-coated carbon paper, with a mass loading of ~1.25 mg cm2. Titanium mesh was employed as current collector. Ethanol and Nafion 117 was used as dispersant and conductive binder, respectively. The electrolyte (6 M KOH and 0.15 M zinc oxide (ZnO, Merck)) was injected into the cell after assembly. The effective area of the air cathode was approximately about 0.785 cm2.
All battery performances were assessed by using potentiostat galvanostat machine (VersaSTAT 3). Polarization curves for charge and discharge were obtained from galvanodynamic measurements. EIS measurements were performed at a constant voltage of 0.9 V, with frequency ranging from 50 kHz to 0.2 Hz, and an amplitude of 20 mV. Stability test were performed by using recurrent galvanic pulse testing, with current density of 10 mA cm−2 for 10 min of discharging and 10 min of discharging per cycle.

4. Conclusions

We report a simple method for synthesizing a high yield carbon-based bifunctional electrocatalyst for rechargeable Zn-air battery. Facile dehydration of table sugar and the following activation process is chosen to produce large amount of carbon support which can be used to deposit manganese oxide nanorods. The presence of poorly crystalline manganese oxide further improves the bifunctional catalytic activity of the carbon during ORR and OER, which contribute to the enhancement of cycling stability of the obtained rechargeable Zn-air battery. The catalytic activity of each catalyst and the corresponding Zn-air battery performance are summarized in Table 1. This study contributes to the advancement of synthesizing efficient and large-scale bifunctional catalyst, especially for rechargeable Zn-air batteries application.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/1/64/s1. Figure S1: EDS mapping of Mn in MnOx@AC-S. Figure S2: ORR activity of Pt/C catalyst. Figure S3: ORR activity of Vulcan XC-72 catalyst. S4. ORR and OER stability of MnOx@AC-S catalyst. S5. Performance of rechargeable Zn-air batteries in literature.

Author Contributions

Conceptualization, A.S., Z.L., and Y.Z.; methodology, M.A.M., Y.M., A.W., and A.S.; validation, A.W., Y.Z., Z.L., and A.S.; formal analysis, M.A.M., Y.M., B.P., J.J.H., A.W., Y.Z., Z.L., and A.S.; investigation, M.A.M., Y.M., J.J.H., and B.P.; resources, A.W., Y.Z., Z.L., and A.S.; data curation, M.A.M., Y.M., B.P., J.J.H., and A.S.; writing—original draft preparation, M.A.M., B.P., J.J.H., and A.S.; writing—review and editing, M.A.M., Y.M., B.P., J.J.H., A.W., Y.Z., Z.L., and A.S.; visualization, M.A.M., B.P., and J.J.H.; supervision, Z.L. and A.S.; project administration, A.W. and A.S.; funding acquisition, B.P. and A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Research, Community Service and Innovation Program (P3MI) Institut Teknologi Bandung, Grant 2019.

Acknowledgments

We gratefully acknowledge the financial support from Research, Community Service and Innovation Program (P3MI) Institut Teknologi Bandung, Grant 2019.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Li, Y.; Lu, J. Metal–air batteries: Will they be the future electrochemical energy storage device of choice? ACS Energy Lett. 2017, 2, 1370–1377. [Google Scholar] [CrossRef]
  2. Li, Y.; Dai, H. Recent advances in zinc–air batteries. Chem. Soc. Rev. 2014, 43, 5257–5275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Stevens, P.; Toussaint, G.; Caillon, G.; Viaud, P.; Vinatier, P.; Cantau, C.; Fichet, O.; Sarrazin, C.; Mallouki, M. Development of a lithium air rechargeable battery. ECS Trans. 2010, 28, 1–12. [Google Scholar]
  4. Sumboja, A.; Ge, X.; Goh, F.T.; Li, B.; Geng, D.; Hor, T.A.; Zong, Y.; Liu, Z. Manganese oxide catalyst grown on carbon paper as an air cathode for high-performance rechargeable zinc–air batteries. ChemPlusChem 2015, 80, 1341–1346. [Google Scholar] [CrossRef]
  5. Pei, P.; Wang, K.; Ma, Z. Technologies for extending zinc–air battery’s cyclelife: A review. Appl. Energy 2014, 128, 315–324. [Google Scholar] [CrossRef]
  6. Guan, C.; Sumboja, A.; Zang, W.; Qian, Y.; Zhang, H.; Liu, X.; Liu, Z.; Zhao, D.; Pennycook, S.J.; Wang, J. Decorating Co/CoNx nanoparticles in nitrogen-doped carbon nanoarrays for flexible and rechargeable zinc-air batteries. Energy Storage Mater. 2019, 16, 243–250. [Google Scholar] [CrossRef]
  7. Cheng, F.; Su, Y.; Liang, J.; Tao, Z.; Chen, J. MnO2-based nanostructures as catalysts for electrochemical oxygen reduction in alkaline media. Chem. Mater. 2009, 22, 898–905. [Google Scholar] [CrossRef]
  8. Xiao, W.; Wang, D.; Lou, X.W. Shape-controlled synthesis of MnO2 nanostructures with enhanced electrocatalytic activity for oxygen reduction. J. Phys. Chem. C 2009, 114, 1694–1700. [Google Scholar] [CrossRef]
  9. Lübke, M.; Sumboja, A.; McCafferty, L.; Armer, C.F.; Handoko, A.D.; Du, Y.; McColl, K.; Cora, F.; Brett, D.; Liu, Z.; et al. Transition-Metal-Doped α-MnO2 Nanorods as Bifunctional Catalysts for Efficient Oxygen Reduction and Evolution Reactions. ChemistrySelect 2018, 3, 2613–2622. [Google Scholar] [CrossRef]
  10. Li, P.-C.; Hu, C.-C.; Lee, T.-C.; Chang, W.-S.; Wang, T.H. Synthesis and characterization of carbon black/manganese oxide air cathodes for zinc-air batteries. J. Power Sources 2014, 269, 88–97. [Google Scholar] [CrossRef]
  11. Yang, C.-C.; Hsu, S.-T.; Chien, W.-C.; Shih, M.C.; Chiu, S.-J.; Lee, K.-T.; Wang, C.L. Electrochemical properties of air electrodes based on MnO2 catalysts supported on binary carbons. Int. J. Hydrogen Energy 2006, 31, 2076–2087. [Google Scholar] [CrossRef]
  12. Sharma, R.K.; Oh, H.-S.; Shul, Y.-G.; Kim, H. Growth and characterization of carbon-supported MnO2 nanorods for supercapacitor electrode. Phys. B 2008, 403, 1763–1769. [Google Scholar] [CrossRef]
  13. Ma, H.; Wang, B. A bifunctional electrocatalyst α-MnO 2-LaNiO 3/carbon nanotube composite for rechargeable zinc–air batteries. RSC Adv. 2014, 4, 46084–46092. [Google Scholar] [CrossRef]
  14. Sumboja, A.; Lübke, M.; Wang, Y.; An, T.; Zong, Y.; Liu, Z. All-Solid-State, Foldable, and Rechargeable Zn-Air Batteries Based on Manganese Oxide Grown on Graphene-Coated Carbon Cloth Air Cathode. Adv. Energy Mater. 2017, 7, 1700927. [Google Scholar] [CrossRef]
  15. Qu, L.; Liu, Y.; Baek, J.-B.; Dai, L. Nitrogen-doped graphene as efficient metal-free electrocatalyst for oxygen reduction in fuel cells. ACS Nano 2010, 4, 1321–1326. [Google Scholar] [CrossRef] [PubMed]
  16. Gong, K.; Du, F.; Xia, Z.; Durstock, M.; Dai, L. Nitrogen-doped carbon nanotube arrays with high electrocatalytic activity for oxygen reduction. Science 2009, 323, 760–764. [Google Scholar] [CrossRef] [Green Version]
  17. Li, B.; Geng, D.; Lee, X.S.; Ge, X.; Chai, J.; Wang, Z.; Zhang, J.; Liu, Z.; Hor, T.A.; Zong, Y. Eggplant-derived microporous carbon sheets: Towards mass production of efficient bifunctional oxygen electrocatalysts at low cost for rechargeable Zn–air batteries. Chem. Commun. 2015, 51, 8841–8844. [Google Scholar] [CrossRef]
  18. Borghei, M.; Laocharoen, N.; Kibena-Põldsepp, E.; Johansson, L.-S.; Campbell, J.; Kauppinen, E.; Tammeveski, K.; Rojas, O.J. Porous N, P-doped carbon from coconut shells with high electrocatalytic activity for oxygen reduction: Alternative to Pt-C for alkaline fuel cells. Appl. Catal. B 2017, 204, 394–402. [Google Scholar] [CrossRef]
  19. Liu, F.; Peng, H.; Qiao, X.; Fu, Z.; Huang, P.; Liao, S. High-performance doped carbon electrocatalyst derived from soybean biomass and promoted by zinc chloride. Int. J. Hydrogen Energy 2014, 39, 10128–10134. [Google Scholar] [CrossRef]
  20. Yang, L.; Zeng, X.; Wang, D.; Cao, D. Biomass-derived FeNi alloy and nitrogen-codoped porous carbons as highly efficient oxygen reduction and evolution bifunctional electrocatalysts for rechargeable Zn-air battery. Energy Storage Mater. 2018, 12, 277–283. [Google Scholar] [CrossRef]
  21. Borghei, M.; Lehtonen, J.; Liu, L.; Rojas, O.J. Advanced Biomass-Derived Electrocatalysts for the Oxygen Reduction Reaction. Adv. Mater. 2018, 30, 1703691. [Google Scholar] [CrossRef] [PubMed]
  22. Hargreaves, N.J.; Cooper, S.J. Nanographite synthesized from acidified sucrose microemulsions under ambient conditions. Cryst. Growth Des. 2016, 16, 3133–3142. [Google Scholar] [CrossRef] [Green Version]
  23. Pan, F.; Jin, J.; Fu, X.; Liu, Q.; Zhang, J. Advanced oxygen reduction electrocatalyst based on nitrogen-doped graphene derived from edible sugar and urea. ACS Appl. Mater. Interfaces 2013, 5, 11108–11114. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Z.; Zhang, F.; Jin, C.; Luo, Y.; Sui, J.; Gong, H.; Yang, R. La2O3-NCNTs hybrids in-situ derived from LaNi0. 9Fe0. 1O3-C composites as novel robust bifunctional oxygen electrocatalysts. Carbon 2017, 115, 261–270. [Google Scholar] [CrossRef]
  25. Oyedotun, K.O.; Barzegar, F.; Mirghni, A.A.; Khaleed, A.A.; Masikhwa, T.M.; Manyala, N. Examination of High-Porosity Activated Carbon Obtained from Dehydration of White Sugar for Electrochemical Capacitor Applications. ACS Sustain. Chem. Eng. 2019, 7, 537–546. [Google Scholar] [CrossRef]
  26. Stoerzinger, K.A.; Risch, M.; Han, B.; Shao-Horn, Y. Recent insights into manganese oxides in catalyzing oxygen reduction kinetics. ACS Catal. 2015, 5, 6021–6031. [Google Scholar] [CrossRef] [Green Version]
  27. Li, Y.; Gong, M.; Liang, Y.; Feng, J.; Kim, J.-E.; Wang, H.; Hong, G.; Zhang, B.; Dai, H. Advanced zinc-air batteries based on high-performance hybrid electrocatalysts. Nat. Commun. 2013, 4, 1805. [Google Scholar] [CrossRef] [Green Version]
  28. Lee, D.U.; Choi, J.Y.; Feng, K.; Park, H.W.; Chen, Z. Advanced Extremely Durable 3D Bifunctional Air Electrodes for Rechargeable Zinc-Air Batteries. Adv. Energy Mater. 2014, 4, 1301389. [Google Scholar] [CrossRef]
  29. Chen, Z.; Yu, A.; Ahmed, R.; Wang, H.; Li, H.; Chen, Z. Manganese dioxide nanotube and nitrogen-doped carbon nanotube based composite bifunctional catalyst for rechargeable zinc-air battery. Electrochim. Acta 2012, 69, 295–300. [Google Scholar] [CrossRef]
  30. Du, G.; Liu, X.; Zong, Y.; Hor, T.A.; Yu, A.; Liu, Z. Co3O4 nanoparticle-modified MnO2 nanotube bifunctional oxygen cathode catalysts for rechargeable zinc–air batteries. Nanoscale 2013, 5, 4657–4661. [Google Scholar] [CrossRef]
  31. Wang, J.; Nie, P.; Ding, B.; Dong, S.; Hao, X.; Dou, H.; Zhang, X. Biomass derived carbon for energy storage devices. J. Mater. Chem. A 2017, 5, 2411–2428. [Google Scholar] [CrossRef]
  32. Wang, X.; Li, Y. Selected-control hydrothermal synthesis of α-and β-MnO2 single crystal nanowires. J. Am. Chem. Soc. 2002, 124, 2880–2881. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, X.-Y.; Huang, M.; Ma, H.-L.; Zhang, Z.-Q.; Gao, J.-M.; Zhu, Y.-L.; Han, X.-J.; Guo, X.-Y. Preparation of a carbon-based solid acid catalyst by sulfonating activated carbon in a chemical reduction process. Molecules 2010, 15, 7188–7196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Liu, X.; Chen, C.; Cheng, Q.; Zou, L.; Zou, Z.; Yang, H. Binary Nitrogen Precursor-Derived Porous Fe-N-S/C Catalyst for Efficient Oxygen Reduction Reaction in a Zn-Air Battery. Catalysts 2018, 8, 158. [Google Scholar] [CrossRef] [Green Version]
  35. Cheng, F.; Zhao, J.; Song, W.; Li, C.; Ma, H.; Chen, J.; Shen, P. Facile controlled synthesis of MnO2 nanostructures of novel shapes and their application in batteries. Inorg. Chem. 2006, 45, 2038–2044. [Google Scholar] [CrossRef]
  36. Sumboja, A.; Foo, C.Y.; Wang, X.; Lee, P.S. Large Areal Mass, Flexible and Free-Standing Reduced Graphene Oxide/Manganese Dioxide Paper for Asymmetric Supercapacitor Device. Adv. Mater. 2013, 25, 2809–2815. [Google Scholar] [CrossRef]
  37. Xiao, W.; Hu, D.; Peng, C.; Chen, G.Z. Interfacial Synthesis: Amphiphilic Monomers Assisted Ultrarefining of Mesoporous Manganese Oxide Nanoparticles and the Electrochemical Implications. ACS Appl. Mater. Interfaces 2011, 3, 3120–3129. [Google Scholar] [CrossRef]
  38. Lv, P.; Feng, Y.Y.; Li, Y.; Feng, W. Carbon fabric-aligned carbon nanotube/MnO2/conducting polymers ternary composite electrodes with high utilization and mass loading of MnO2 for super-capacitors. J. Power Sources 2012, 220, 160–168. [Google Scholar] [CrossRef]
  39. Shen, B.; Liu, T.; Zhao, N.; Yang, X.; Deng, L. Iron-doped Mn-Ce/TiO2 catalyst for low temperature selective catalytic reduction of NO with NH3. J. Environ. Sci. 2010, 22, 1447–1454. [Google Scholar] [CrossRef]
  40. Wu, Z.-S.; Ren, W.; Wang, D.-W.; Li, F.; Liu, B.; Cheng, H.-M. High-Energy MnO2 Nanowire/Graphene and Graphene Asymmetric Electrochemical Capacitors. ACS Nano 2010, 4, 5835–5842. [Google Scholar] [CrossRef]
  41. Ghaemi, M.; Ataherian, F.; Zolfaghari, A.; Jafari, S.M. Charge storage mechanism of sonochemically prepared MnO2 as supercapacitor electrode: Effects of physisorbed water and proton conduction. Electrochim. Acta 2008, 53, 4607–4614. [Google Scholar] [CrossRef]
  42. Zhao, Z.; Schlexer Lamoureux, P.; Kulkarni, A.; Bajdich, M. Trends in Oxygen Electrocatalysis of 3d-Layered (Oxy)(Hydro) Oxides. ChemCatChem 2019, 11, 3423–3431. [Google Scholar] [CrossRef]
  43. Ma, Y.; Sumboja, A.; Zang, W.; Yin, S.; Wang, S.; Pennycook, S.J.; Kou, Z.; Liu, Z.; Li, X.; Wang, J. Flexible and Wearable All-Solid-State Al–Air Battery Based on Iron Carbide Encapsulated in Electrospun Porous Carbon Nanofibers. ACS Appl. Mater. Interfaces 2019, 11, 1988–1995. [Google Scholar] [CrossRef] [PubMed]
  44. Sumboja, A.; Chen, J.; Ma, Y.; Xu, Y.; Zong, Y.; Lee, P.S.; Liu, Z. Sulfur-Rich Colloidal Nickel Sulfides as Bifunctional Catalyst for All-Solid-State, Flexible and Rechargeable Zn-Air Batteries. ChemCatChem 2019, 11, 1205–1213. [Google Scholar] [CrossRef]
  45. Li, B.; Chai, J.; Ge, X.; An, T.; Lim, P.C.; Liu, Z.; Zong, Y. Sheet-on-Sheet Hierarchical Nanostructured C@ MnO2 for Zn-Air and Zn-MnO2 Batteries. ChemNanoMat 2017, 3, 401–405. [Google Scholar] [CrossRef]
  46. Meng, Y.; Song, W.; Huang, H.; Ren, Z.; Chen, S.-Y.; Suib, S.L. Structure–property relationship of bifunctional MnO2 nanostructures: Highly efficient, ultra-stable electrochemical water oxidation and oxygen reduction reaction catalysts identified in alkaline media. J. Am. Chem. Soc. 2014, 136, 11452–11464. [Google Scholar] [CrossRef]
  47. Ding, J.; Wang, P.; Ji, S.; Wang, H.; Linkov, V.; Wang, R. N-doped mesoporous FeNx/carbon as ORR and OER bifunctional electrocatalyst for rechargeable zinc-air batteries. Electrochim. Acta 2019, 296, 653–661. [Google Scholar] [CrossRef]
  48. Davis, R.; Horvath, G.; Tobias, C. The solubility and diffusion coefficient of oxygen in potassium hydroxide solutions. Electrochim. Acta 1967, 12, 287–297. [Google Scholar] [CrossRef]
  49. Sumboja, A.; Chen, J.; Zong, Y.; Lee, P.S.; Liu, Z. NiMn layered double hydroxides as efficient electrocatalysts for the oxygen evolution reaction and their application in rechargeable Zn–air batteries. Nanoscale 2017, 9, 774–780. [Google Scholar] [CrossRef]
Figure 1. SEM images of the carbon produced by dehydration of sugar: (a) prior to activation. (b) after activation with KOH and (c) after introduction of manganese oxide.
Figure 1. SEM images of the carbon produced by dehydration of sugar: (a) prior to activation. (b) after activation with KOH and (c) after introduction of manganese oxide.
Catalysts 10 00064 g001
Figure 2. TEM images of (a) activated carbon derived from table sugar (AC-S) and (b) MnOx grown on the activated carbon sugar (MnOx@AC-S), showing deposited MnOx nanoparticles in the form of nanorod which are not present in the activated carbon derived from table sugar (AC-S).
Figure 2. TEM images of (a) activated carbon derived from table sugar (AC-S) and (b) MnOx grown on the activated carbon sugar (MnOx@AC-S), showing deposited MnOx nanoparticles in the form of nanorod which are not present in the activated carbon derived from table sugar (AC-S).
Catalysts 10 00064 g002
Figure 3. (a) XRD pattern of AC-S and MnOx@AC-S. (b) Mn 2p XPS spectra and (c) O1s XPS spectra.
Figure 3. (a) XRD pattern of AC-S and MnOx@AC-S. (b) Mn 2p XPS spectra and (c) O1s XPS spectra.
Catalysts 10 00064 g003
Figure 4. (a) Oxygen reduction reaction (ORR) polarization curves of AC-S, MnOx@AC-S, Vulcan XC-72 carbon, and Pt/C at rotating speed of 1600 rpm. Linear scan voltammograms of (b) AC-S and (c) MnOx@AC-S at various rotating speeds and the corresponding Koutecky–Levich plots extracted at different potentials. (d) Oxygen evolution reaction (OER) polarization curves of AC-S, MnOx@AC-S, Vulcan XC-72 carbon, and Pt/C at rotating speed of 1600 rpm.
Figure 4. (a) Oxygen reduction reaction (ORR) polarization curves of AC-S, MnOx@AC-S, Vulcan XC-72 carbon, and Pt/C at rotating speed of 1600 rpm. Linear scan voltammograms of (b) AC-S and (c) MnOx@AC-S at various rotating speeds and the corresponding Koutecky–Levich plots extracted at different potentials. (d) Oxygen evolution reaction (OER) polarization curves of AC-S, MnOx@AC-S, Vulcan XC-72 carbon, and Pt/C at rotating speed of 1600 rpm.
Catalysts 10 00064 g004
Figure 5. Performance of rechargeable Zn-air batteries with AC-S, MnOx@AC-S and Pt/C as bifunctional catalysts: (a) Galvanodynamic discharge and charge polarization curves. (b) Power density curve constructed from galvanodynamic discharge curve. (c) Nyquist plot obtained from electrochemical impedance spectroscopy (EIS) measurements. (d) Cycling stability during galvanostatic pulse cycling test for 20 min per cycle.
Figure 5. Performance of rechargeable Zn-air batteries with AC-S, MnOx@AC-S and Pt/C as bifunctional catalysts: (a) Galvanodynamic discharge and charge polarization curves. (b) Power density curve constructed from galvanodynamic discharge curve. (c) Nyquist plot obtained from electrochemical impedance spectroscopy (EIS) measurements. (d) Cycling stability during galvanostatic pulse cycling test for 20 min per cycle.
Catalysts 10 00064 g005
Table 1. Electrocatalytic activity of the catalysts and the corresponding Zn-air battery performance.
Table 1. Electrocatalytic activity of the catalysts and the corresponding Zn-air battery performance.
ParameterAC-SMnOx@AC-SPt/CVulcan XC-72
ORR onset potential (V vs. RHE)0.8540.9141.0350.817
ORR limiting current density (mA cm−2)4.1804.3804.4053.947
Electron transfer number (n)3.124.004.002.83
OER overpotential @10 mA cm−2 (mV)538500580-
Peak power density (mW cm−2) of Zn-air battery with the corresponding catalyst105116126-
Maximum cycle (n) of rechargeable Zn-air battery with the corresponding catalyst66215141-

Share and Cite

MDPI and ACS Style

Marsudi, M.A.; Ma, Y.; Prakoso, B.; Hutani, J.J.; Wibowo, A.; Zong, Y.; Liu, Z.; Sumboja, A. Manganese Oxide Nanorods Decorated Table Sugar Derived Carbon as Efficient Bifunctional Catalyst in Rechargeable Zn-Air Batteries. Catalysts 2020, 10, 64. https://doi.org/10.3390/catal10010064

AMA Style

Marsudi MA, Ma Y, Prakoso B, Hutani JJ, Wibowo A, Zong Y, Liu Z, Sumboja A. Manganese Oxide Nanorods Decorated Table Sugar Derived Carbon as Efficient Bifunctional Catalyst in Rechargeable Zn-Air Batteries. Catalysts. 2020; 10(1):64. https://doi.org/10.3390/catal10010064

Chicago/Turabian Style

Marsudi, Maradhana Agung, Yuanyuan Ma, Bagas Prakoso, Jayadi Jaya Hutani, Arie Wibowo, Yun Zong, Zhaolin Liu, and Afriyanti Sumboja. 2020. "Manganese Oxide Nanorods Decorated Table Sugar Derived Carbon as Efficient Bifunctional Catalyst in Rechargeable Zn-Air Batteries" Catalysts 10, no. 1: 64. https://doi.org/10.3390/catal10010064

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