Next Article in Journal
Regioselective Hydroxylation of Naringin Dihydrochalcone to Produce Neoeriocitrin Dihydrochalcone by CYP102A1 (BM3) Mutants
Next Article in Special Issue
Catalyst Special Issue on Catalytic Reactors Design for Industrial Applications
Previous Article in Journal
Conversion of Xylose to Furfural over Lignin-Based Activated Carbon-Supported Iron Catalysts
Previous Article in Special Issue
Numerical Investigation of Ventilation Air Methane Catalytic Combustion in Circular Straight and Helical Coil Channels with Twisted Tape Insert in Catalytic-Monolith Reactors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Evaluation of Nanoflower Structured Manganese Oxide Electrocatalyst for Oxygen Reduction in Alkaline Media

1
Chemical Engineering Department, Universiti Teknologi PETRONAS, Bandar Seri Iskandar 32610, Perak, Malaysia
2
HiCoE-Center for Biofuel and Biochemical Research, Institute of Self-Sustainable Building, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Perak Darul Ridzuan, Malaysia
3
Advanced Membrane Technology Research Centre (AMTEC), Universiti Teknologi Malaysia (UTM), UTM Johor Bahru 81310, Johor, Malaysia
4
Center of Excellence in Process and Energy Systems Engineering, Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(8), 822; https://doi.org/10.3390/catal10080822
Submission received: 27 May 2020 / Revised: 19 June 2020 / Accepted: 23 June 2020 / Published: 23 July 2020
(This article belongs to the Special Issue Catalytic Reactors Design for Industrial Applications)

Abstract

:
An electrochemical nanoflowers manganese oxide (MnO2) catalyst has gained much interest due to its high stability and high specific surface area. However, there are a lack of insightful studies of electrocatalyst performance in nanoflower MnO2. This study assesses the electrocatalytic performances of nanoflower structure MnO2 for both oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) in a zinc–air battery as a bifunctional electrocatalyst. The prepared catalyst was characterized in term of morphology, crystallinity, and total surface area. Cyclic voltammetry and linear sweep voltammetry were used to evaluate the electrochemical behaviors of the as-prepared nanoflower-like MnO2. The discharge performance test for zinc–air battery with a MnO2 catalyst was also conducted. The results show that the MnO2 prepared at dwell times of 2, 4 and 6 h were nanoflowers, nanoflower mixed with nanowires, and nanowires with corresponding specific surface areas of 52.4, 34.9 and 32.4 g/cm2, respectively. The nanoflower-like MnO2 catalyst exhibits a better electrocatalytic performance towards both ORR and OER compared to the nanowires. The number of electrons transferred for the MnO2 with nanoflower, nanoflower mixed with nanowires, and nanowire structures is 3.68, 3.31 and 3.00, respectively. The as-prepared MnO2 nanoflower-like structure exhibits the best discharge performance of 31% higher than the nanowires and reaches up to 30% of the theoretical discharge capacity of the zinc–air battery.

1. Introduction

Renewable energy sources have broadly attracted attention to supply global energy demand due to the excess utilization of petroleum-based fuels [1,2]. Nevertheless, the efficient utilization of renewable energy sources requires safe and cost-effective electricity storage systems. Zinc–air batteries are considered as the most promising alternative energy storage due to several advantages such as high theoretical specific energy density with a flat constant discharge voltage, the low reactivity of zinc, environmental safety and quick refueling with fresh zinc powder and granules. Moreover, the use of highly abundant and free oxygen as the reactant at the cathode does not require a heavy casing to keep it inside which generally makes the battery heavy and space consuming [3,4,5,6].
However, the large overpotential (∆V) between the oxygen evolution reaction (OER) and the oxygen reduction reaction (ORR) reduces the life cycle and, thus, limiting the performance of the secondary zinc–air batteries [7,8,9,10,11,12]. The development of efficient and stable bifunctional catalysts towards the OER and ORR is critical to support the technology developments.
Enormous researches have been conducted towards the development of high-performance batteries with low-cost materials [3,6]. Currently, the most used electrocatalyst for ORR and OER reactions is platinum. However, the high cost and susceptibility to catalyst poisoning has limited their commercialization in zinc–air batteries [13,14,15,16,17,18,19,20,21]. Hence, the search for an economical and viable catalyst with high catalytic performance as a potential component in zinc–air batteries has become the key for electrochemistry researchers.
The new development of effective and low-cost materials has been introduced for ORR such as nitrogen-doped graphene, which was synthesized by electrochemical exfoliation of commercial graphite in a Na2SO4 electrolyte with the addition of CaCO3 as a separator of newly exfoliated FL-graphene sheets. Exfoliated FL-graphene was later infused with a suspension of green algae which worked as a nitrogen carrier. Impregnated FL-graphene was carbonized at a high temperature under the flow of nitrogen. This nitrogen-doped graphene has shown significant results for catalytic activity in ORR, which can be employed in a Zn–air battery [22].
Manganese oxide (MnO2) is one of the most promising electrocatalysts since it has shown excellent electrocatalytic performances for OER and ORR under alkaline conditions [23,24,25,26,27]. Meanwhile, it is abundant in natural ores, low in toxicity, low cost and environmentally friendly. The catalytic activity of MnO2 in its native form is low compared to the platinum and its group metals due to its weak O2 binding ability and limited electronic conductivity [15,28,29]. However, the electrocatalytic performances of MnO2 for ORR and OER is a highly structure sensitive. Therefore, the performance of MnO2 catalysts can still be improved by modifying their morphologies, Mn valence state, preparation methods, crystalline phases and structures [30,31,32].
The nanoflower MnO2 structure has gained much interest due to the high stability and high surface area to volume ratio [33]. Several studies have shown that the nanoflower structure exhibits the large surface area and can offer a significant improvement in ORR activities [34]. However, there are a lack of insightful studies of OER electrocatalyst properties on nanoflower MnO2, as well as its electrocatalytic performances. Hence, this study evaluates the electrocatalytic performances of nanoflower structure MnO2 for both ORR and OER in the zinc–air battery as a bifunctional electrocatalyst.

2. Results

2.1. Morphological Study of the As-Prepared Catalysts

Figure 1 shows the SEM images for the synthesized MnO2 catalysts with different dwell times. It shows that the rod-like MnO2 structure was formed at a longer dwell time. This phenomenon can be explained by the Ostwald Ripening process during formation of the catalyst [23]. The nanoflower structure was produced in a dwell time of 2 h, nanoflower mixed with nanowires was produced in a dwell time of 4 h, and nanowires were produced in a dwell time of 6 h. The obtained nanoflower structure of the MnO2 particles exhibits irregular shape with sizes ranging from 1 to 3 µm and all clumping together. Meanwhile, the nanowire morphology developed from one into a few microns’ length, about 50 nm in diameter.
Figure 2 displays N2 adsorption/desorption isotherms of all the catalysts, showing (from the slopes of the linear region of the curves) that the nanoflower MnO2 has a higher specific surface area (52.4 cm2/g) followed by nanoflower mixed with nanowires (34.9 cm2/g) and nanowires (32.4 cm2/g). The results indicate that the nanoflower structure exhibits a higher surface area compared to nanowires. Therefore, more abundant nanoflower structures are present on the surface of the catalyst. This general statement suggests that the higher surface is mainly responsible for higher catalytic performance for all types of catalysts, regardless of the type of metals and support [34]. The surface area varies from material to material and depends on preparation methods. In this study, a higher surface area was observed in nanoflower MnO2, which is a significant factor for high catalytic performance.

2.2. Electrochemical Properties of the As-Prepared Catalysts

ORR is a complex and multi-electron reaction. The ORR can proceed via two different reaction pathways, namely, a direct four-electron pathway, as shown in Equation (1), or a two-electron pathway, as exhibited in Equations (2) and (3) [35].
O 2 + 2 H 2 O + 4 e 4 OH
O 2 + H 2 O + 2 e HO 2 + OH
HO 2 + H 2 O + 2 e 3 OH
Figure 3 shows all the samples have a small but observable peak around −0.4 to −0.5 V vs. Ag/AgCl. This position of the reduction peak is close to the reduction potential of four-electron ORR. The findings indicate that all the samples tend to go for the four-electron ORR that does not produce hydrogen peroxide (HO2). A further reduction of HO2 to OH was also not observed at around −1.0 V. Moreover, the carbon Vulcan XC 72 has, overall, a higher reduction peak compared to all the MnO2 catalysts, which can be ascribed due to the low electrical conductivity of the MnO2.

2.3. Oxygen Reduction Reaction Performance

To further investigate the number of electron transferred and ORR performance of the catalyst, a linear sweep voltammetry (LSV) was conducted to obtain the polarization curves at different rotational speed in 0.1-M oxygen-saturated KOH for all the samples. Figure 4 demonstrates the polarization curves of all samples at 1600 rpm. It clearly shows that the onset potential and terminal current density of the prepared catalysts are both strongly influenced by the nanostructure. The carbon Vulcan XC 72/nanoflower-like MnO2 has the highest terminal current density (−6.65 mA/cm2) at a potential of −1.0 V and a higher onset potential (−0.13 V) at 1 mA/cm2. The carbon Vulcan XC 72/nanoflower-like MnO2 catalyst has the highest performance followed by carbon Vulcan XC 72 mixed with major nanoflower and minor nanowires, carbon Vulcan XC 72 mixed with major nanowires and minor nanoflower, pure carbon Vulcan XC 72 and carbon Vulcan XC 72/commercial MnO2. This trend indicates that the nanoflower-like is better than the nanowire-like structure due to its higher specific surface area that provides more abundant active sites for the ORR.
The number of electrons transferred per oxygen molecule for oxygen reduction was further evaluated by using the Koutecky–Levich (K-L) equation to study the ORR mechanism of the synthesized catalysts. The K-L equation is shown accordingly in Equations (4) and (5):
1 j = 1 j k + 1 B ω 1 / 2
B = 0.2 n F D O 2 2 / 3 v 1 / 6 C O 2
Figure 5 shows the K-L plot for carbon Vulcan XC 72/nanoflower-like MnO2, carbon Vulcan XC 72/nanoflower–nanowire-like MnO2, and carbon Vulcan XC 72/nanowire-like MnO2 catalysts. It shows that as the potential increase, the gradient of the K-L plot which also refers as the number of electrons transferred decreases. The linear and parallel behavior of these plots indicate the first-order kinetics with respect to molecular oxygen for the ORR [29].
The number of electrons transferred for ORR reaction obtained from the K-L plot for the Vulcan XC 72/nanoflower-like MnO2, carbon Vulcan XC 72/nanoflower-nanowire-like MnO2 and Vulcan XC 72/nanowire-like MnO2 catalysts are 3.68, 3.6 and 2.98 at −1.0 V, −0.8 V and −0.6 V, respectively. This result indicates that the synthesized nanoflower-like MnO2 exhibits excellent electrocatalytic performance and stability at higher potentials. Meanwhile, the carbon Vulcan XC 72, mixed with major nanoflower and minor nanowires, and the carbon Vulcan XC 72, mixed with major nanowires and minor nanoflower, exhibit a similar trend as the carbon Vulcan XC 72/nanoflower-like MnO2 but differ in the number of electrons transferred.
When comparing the K-L Plot of all samples at a fixed potential of −1.0 V and by obtaining the number of electrons transferred from the gradient of the K-L plot, the number of electrons transferred for carbon VXC 72 mixed with nanoflower MnO2, carbon Vulcan XC 72 mixed with nanoflower mixed with nanowires, and carbon VXC 72 mixed with nanowires is 3.68, 3.31 and 3.00, respectively.
In order to suppress the evolution of hydrogen peroxide, the desired ORR pathway is the four-electron pathway in which the number of electrons transferred experimentally must be between two and four to show that the electrocatalyst promotes some of the ORR pathways to occur in the four-electron pathway. The results show that MnO2 nanoflower has the best electrocatalyst performance towards the ORR, as the number of transferred electrons is close to four, suggesting the likelihood of promoting the ORR reaction toward the four-electron pathway. Meanwhile, the electrocatalytic performance for carbon Vulcan XC 72 mixed with major nanoflower and minor nanowires and carbon Vulcan XC 72 mixed with major nanowires and minor nanoflower are also within the acceptable range of two to four.

2.4. Oxygen Evolution Reaction

As the MnO2 has potential as a bifunctional catalyst towards OER, the OER performance of the nanostructure MnO2 was further investigated. An LSV from 0.2 V to 1.0 V was conducted to obtain the polarization curves for the OER on rotating disk electrode as shown in Figure 6. The results confirm that the nanoflower-like MnO2 has the lowest onset potential. The onset potential and the overpotential of the samples are summarized in Table 1.
Table 1 shows that the nanoflower-like MnO2 has the lowest onset potential. As overpotential is defined as the difference between the theoretical half-reaction reduction potential of OER and the onset potential (which is 0.3244 V), the trend of the overpotential is the same as the onset potential. At lower overpotential, the energy lost during the reaction in the electrode is also lower. As shown in Table 1, the pure carbon VXC 72 has the highest overpotential (energy lost) during OER, while the nanoflower MnO2 has the least energy lost. The difference in overpotential between MnO2 and pure carbon VXC 72 can be explained by the electrocatalyst behaviors of the MnO2, where the hydroxide ions can be adsorbed on the surface of MnO2 due to the valence state of MnO2. Such a phenomenon can be observed in Figure 6, which shows that the sharp increment in the current density (indicating the OER for MnO2 samples and pure carbon VXC 72) has significant differences. The electrocatalyst behaviors of MnO2 increases the rate of reaction and alters the pathway for the OER. Therefore, a higher gradient of the faradaic region from 0.7 to 1.0 V can be observed for the MnO2 samples. Overall, the results suggest that the nanoflower structure has a better nanostructure with a higher stability and better electrocatalyst performance at a high potential.

2.5. Zinc–Air Battery Discharge Performance

During discharge, at the anode, the zinc electrode reacts with hydroxide ion ( OH ) and oxidize to form zincate ion Zn ( OH ) 4 2 , shown in Equation (6). Consequently, the precipitation of zinc oxide (ZnO) takes place when the dissolved zincate ion saturates and reaches its solubility, as shown in Equation (7).
Zn + 4 OH 4 Zn ( OH ) 4 2 + 2 e
Zn ( OH ) 4 2 ZnO + H 2 O + 2 OH
At the cathode, the oxygen reduction reaction (ORR) occurs via four-electron (Equation (1)) or two-electron (Equations (2) and (3)) pathways. ORR consumes oxygen ( O 2 ) and water ( H 2 O ) and simultaneously produces a hydroxide ion. The hydroxide ion and water then transfers across the cell. Normally, the ORR kinetics are sluggish, and an ORR electrocatalyst is required to enhance ORR activity and battery performance. In alkaline solution, platinum (Pt) is an excellent ORR electrocatalyst, having excellent stability and showing a four-electron pathway. Nonetheless, its high cost hinders its practical applications. MnO2 is a good alternative because it is significantly cheaper and more abundant than platinum.
The performances of the synthesized nanostructure MnO2 as a catalyst in the cathode are presented in Figure 7. The galvanostatic discharge profiles indicate that the nanostructure MnO2 follows the expected electrochemical behavior from the electrochemical characterization. In all cases, the batteries exhibited flat discharge profiles, having a voltage plateau at around 1.25 V. The Vulcan XC 72/nanoflower-like MnO2 has the highest specific discharge capacity of up to 240 mAh/g zinc corresponding to a 31% higher performance than the nanowire-like MnO2 in terms of discharge capacity. However, the theoretical specific discharge capacity of a zinc–air battery is 819 mAh/g, suggesting that the nanoflower-like MnO2 only achieves 30% of the theoretical value, while the nanowire-like MnO2 only achieves up to 20%. Such a phenomenon might be contributed to by the passivation that occurred on the zinc anode that usually exists in a zinc–air coin cell battery.

3. Discussion

MnO2 can exist in various crystal structures, such as α-, β-, γ-, δ- and λ-MnO2, depending on the way in which the MnO6 octahedral unit shares their edges and corners. In δ-MnO2, all MnO6 octahedra are edge sharing, however, they are corner sharing in α-MnO2. δ-MnO2 has a layered structure and shows a nanoflower-like architecture, whilst α-MnO2 has a tunnel structure and exhibited architecture of nanowires [36]. The SEM images of the MnO2 catalysts, shown in Figure 1, indicate that the rod-like MnO2 structure was formed at longer dwell time, indicating that the catalyst formation follows the Ostwald Ripening process [23]. In this process, smaller particles gain thermodynamic stability by depositing on a larger particle to minimize the surface area to volume ratio. Hence, as the reaction time increased larger nanostructures were formed, leading to a smaller surface area. The N2 adsorption/desorption data suggest that nanostructure electro-catalysts improve electronic and catalytic properties due to the higher surface area, i.e., (52.4 cm2/g) compared to other counterparts such as nanoflower mixed with nanowires (34.9 cm2/g) and nanowires (32.4 cm2/g), offering more active reaction sites as proven from the highest anodic and cathodic current peaks of the Vulcan XC 72/nanoflower-like MnO2, which correlates well with the high performances in ORR and OER owing to its highest specific surface area. This claim is confirmed by the data of the linear sweep voltammetry curves in Figure 4.
The results on the number of electrons transferred for the ORR reaction obtained from the K-L plot for different MnO2 structures confirm that the synthesized nanoflower-like MnO2 exhibits excellent electrocatalytic performance and stability at higher potentials, as also proven by the K-L plot. It shows that the MnO2 nanoflower has the highest number of electrons transferred towards the four-electron pathway on the ORR. The results can be explained by the specific surface area of the synthesized MnO2. A higher specific surface area leads to more active sites for reaction, ascribing the superiority of the nanoflower-like MnO2 over the other structures, as also shown by the Onset potential and overpotential in Table 1.
Table 2 compares the performance of the MnO2 catalysts developed in this study with other MnO2 catalysts reported in the literature. The results obtained in the study are somewhat in line with earlier report. Despite showing a similar structure (nanoflower) and lower surface area, the number of electrons transferred for ORR is slightly higher than the reference [29]. It shows that the nanoflower-like MnO2 poses surface area close to the nanorod- and nanoflake-like catalysts developed via the electrosynthesis method [20]. Despite showing a lower surface area, it shows higher electrons transferred for ORR from K-L plot. It is also worth noting that there is consistency in the data of the literature on the positive impact of the nanoflower structure in enhancing surface area, but this does not always positively correlate with the number of electrons transferred for the ORR.

4. Materials and Methods

4.1. Chemical and Materials

Analytical grade potassium permanganate (KMnO4) (R&M Chemicals, London, UK) and Manganese (II) sulfate-i-hydrate (MnSO4·H2O) (Bendosen Laboratory Chemicals, Bendosen, Norway) were used to synthesize MnO2. KOH pellets (99%) and zinc sulphate (ZnSO4·7H2O) (CT Chemical Co., Ltd., Bangkok, Thailand), were used to prepare the electrolytes for the electrochemical characterization and battery, respectively. Carbon black (Vulcan® XC 72, Cabot Corporation, Massachusetts, United States) was used to prepare the working electrode for electrochemical characterization. For the zinc–air coin cell battery fabrication, 1-mm-thick nickel foam with a purity of 99.97% (Qijing Trading Co., Ltd., Bangkok, Thailand) was used as the current collector for the cathode. Carbon BP2000 (Black Pearls 2000, Cabot Corporation) and Polystyrene-co-butadiene binder (5%, Sigma-Aldrich, Missouri, United States) and toluene solvent (99.8%, QReC, Selangor, Malaysia) were used to prepare the hydrophilic catalyst layer of the cathode, while poly(tetrafluoroethylene) (PTFE powder, 1 μm, Sigma-Aldrich, Missouri, United States) was used as a binder to prepare the air diffusion layer of cathode. Carbopol (940) with a molecular weight of approximately 1450 monomer units, Whatman filter paper No.1 (Sigma-Aldrich, Missouri, United States) and poly (vinyl acetate) (PVAC) (TOA Paint Public Co., Ltd., Samutprakan, Thailand) were used in fabrication of separator for zinc–air coin cell. The standard CR2032 Stainless steel coin cell parts with 20 holes drilled at the coin cell top is used as the casing of the coin cell.

4.2. Synthesis of Nanoflower-Like MnO2 Catalyst

The nanoflower-like MnO2 was synthesized using hydrothermal synthesis. Separately, 0.1 g of MnSO4·H2O and 0.25 g of KMnO4 were dissolved in 25 mL of deionized water and the solutions were mixed and transferred into a 100-mL, Teflon-lined, stainless-steel autoclave and heated at 140 °C in an oil bath or oven for 2 h. The autoclave was cooled down to room temperature and then the excess solution was filtered. Lastly, the obtained MnO2 powder was washed with ethanol and water and dried in an oven at 60 °C for overnight. The synthesis was repeated with different reaction times for 4 and 6 h to obtain the different nanostructures of MnO2.

4.3. Characterization

The obtained samples were characterized by Scanning electron microscopy (SEM, JEOL-JSM-6480) to investigate the nanostructure of the MnO2. Brunauer–Emmett–Teller (BET) specific surface areas of the samples were determined from nitrogen adsorption isotherms that were performed on a BEL Japan BELSORP-mini II.

4.4. Electrochemical Measurement

All electrochemical measurements were carried out in a three-electrode configuration by a potentiostat/galvanostat with an impedance measurement unit (AMETEK, PAR VersaSTAT 3A). The standard three-electrode cell consists of platinum foil as the counter electrode, a Ag/AgCl electrode as a reference electrode, and a modified glassy carbon electrode (GCE) as the working electrode. The electrocatalytic activity of the as-prepared catalysts towards ORR was measured by cyclic voltammogram (CV) in a mixture of 0.1-M KOH at a scan rate of 50 mV s−1 from −1.8 to 1 V under an ambient atmosphere. For the ORR, Linear Sweep Voltammetry (LSV) was applied with a scan rate of 1 mV/s from 0.2 to −1 V, with the cathodic traces set at a rotational speed of 0, 400, 800, 1200, 1600 and 2000 RPM. For OER, the LSV was applied with a scan rate of 1 mV/s from 0.2 to 1 V, with the anodic traces set at rotational speeds of 0, 400, 800, 1200, 1600 and 2000 RPM.

4.5. Preparation of Working Electrode

In a typical procedure, 0.5 g of the nanoflower-like MnO2 catalysts were ultrasonically mixed in 1 mL 5% polyvinylidene fluoride (PVDF) solutions to form homogenous catalyst ink under room condition. Then, 0.05 mL of the catalyst ink was dripping onto the surface of the glassy carbon electrode (GCE) attached to the rotary machine using a micropipette and then dried at room temperature.

4.6. Electrode and Battery Fabrication

Zinc–air coin cell batteries were fabricated and used to evaluate the discharge capacity of batteries with different catalysts. The fabricated coin cell consisted of three main parts: cathode, anode, and separator. The cathode was synthesized by coating carbon ink solution on a nickel foam. One side of the nickel foam (exhibited hydrophilic properties for reaction occur in the electrolyte) was coated with carbon homogeneous ink solution which consisted of 30 wt.% MnO2, 70 wt.% carbon Vulcan XC 72, and 2.5 wt.% of (MnO2/carbon) dissolved in toluene. Another side of the nickel (exhibited hydrophobic properties as an air diffusion layer) was coated with carbon homogeneous ink solution consisted of 40 wt.% of carbon Vulcan XC 72, 40 wt.% Polytetrafluoroethylene (PTFE), 20% glucose and ethanol were used as the solvent. The anode was prepared by zinc electroplating in 500 cm3 of 1-M ZnSO4·7H2O solution on the surface of the nickel foam. Nickel foams with dimensions of 10 × 1 × 0.1 cm were placed vertically, in parallel at the opposite side of the zinc plate in the electrolyte. The electrodes were connected to a digital DC power supply (ATTEN APS 3005s) operated at galvanostatic mode at voltage of 1.6 V for 2 h for each side of the nickel foam. The separator was a solid polymer electrolyte (SPE) prepared homogeneously by mixing 0.6 g of Carbopol 940 and 15 g of PVAC in 40 mL of deionized water and 15 mL of 7-M KOH solution using a homogenizer. The prepared Carbopol solution was then transferred into a petri dish and dried at open air for 2 to 3 days. The obtained SPE should be around 0.1 to 0.2 mm. After the cathode, anode and separator were ready, a coin cell battery was set up according to the schematic diagram shown in Figure 8. The assembled coin-cell stack was then pressed under 100 kg/cm2 using hydraulic presser.

4.7. Battery Performance Test

The prepared coin cell was connected to a coin cell battery tester, Netware (TOB-CT-4008-5V10mA-164), and discharged at a discharge current of 5 mA until the potential of coin cell depleted below the cut-off voltage of 0.01 V. The specific discharge capacity was obtained by divide the mass of zinc electroplated on the nickel foam which required to measure and record every time before setting up the coin cell.

5. Conclusions

The nanoflower-like MnO2 catalyst structure was successfully prepared by the hydrothermal method. The formation of such structure was confirmed from the SEM shows image in which the dwell times of 2, 4 and 6 h resulting in nanoflowers, nanoflower mixed with nanowires and nanowires structures, respectively. BET results confirm that the nanoflower poses the highest specific surface area of 52.4 cm2/g followed by nanoflower mixed with nanowires of 34.9 cm2/g and nanowires of 32.4 cm2/g. The ORR and OER LSV polarization curve prove the hypothesis of the high performance of nanoflower MnO2 due to its large specific surface area. In terms of ORR, the number of electrons transferred for carbon VXC 72 mixed with nanoflower MnO2, carbon VXC 72 mixed with nanoflower mixed with nanowires and carbon VXC 72 mixed with nanowires are 3.68, 3.31 and 3.00, respectively. The synthesized MnO2 has an excellent coin cell battery performance on improving the discharge capacity of a zinc–air battery. Among all the samples, nanoflower MnO2 exhibits the best discharge performance, reaching up to 30% of the theoretical discharge capacity of the zinc–air battery.

Author Contributions

S.J.H. and M.R.B. conceived and designed the experiments; S.J.H. performed the experiments and wrote the paper, A.A. and M.A. helped to analyzed the data and format the paper; S.K. contributed in material supply and equipment support in their laboratory; M.F.R.H. and J.J. supported in software tools. All authors have read and agreed to the published version of the manuscript.

Funding

Authors acknowledge the Ministry of Higher Education Malaysia for providing Fundamental Research Grant and the APC (Cost Center: 015MA0-037, Reference Number: FRGS/1/2018/TK02/UTP/03/2).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ameen, M.; Azizan, M.T.; Yusup, S.; Ramli, A.; Shahbaz, M.; Aqsha, A. Process optimization of green diesel selectivity and understanding of reaction intermediates. Renew. Energy 2020, 149, 1092–1106. [Google Scholar] [CrossRef]
  2. Ameen, M.; Azizan, M.T.; Yusup, S.; Ramli, A.; Shahbaz, M.; Aqsha, A.; Kaur, H.; Wai, C.K. Parametric Studies on Hydrodeoxygenation of Rubber Seed Oil for Diesel Range Hydrocarbon Production. Energy Fuels 2020, 34, 4603–4617. [Google Scholar] [CrossRef]
  3. 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]
  4. Li, Y.; Dai, H. Recent advances in zinc–air batteries. Chem. Soc. Rev. 2014, 43, 5257–5275. [Google Scholar] [CrossRef] [Green Version]
  5. Yang, D.; Zhang, L.; Yan, X.; Yao, X. Recent Progress in Oxygen Electrocatalysts for Zinc-Air Batteries. Small Methods 2017, 1, 1700209. [Google Scholar] [CrossRef]
  6. Tijani, M.M.; Aqsha, A.; Yu, N.; Mahinpey, N. Determination of redox pathways of supported bimetallic oxygen carriers in a methane fuelled chemical looping combustion system. Fuel 2018, 233, 133–145. [Google Scholar] [CrossRef]
  7. Mainar, A.R.; Colmenares, L.C.; Leonet, O.; Alcaide, F.; Iruin, J.J.; Weinberger, S.; Hacker, V.; Iruin, E.; Urdanpilleta, I.; Blazquez, J.A. Manganese oxide catalysts for secondary zinc air batteries: From electrocatalytic activity to bifunctional air electrode performance. Electrochim. Acta 2016, 217, 80–91. [Google Scholar] [CrossRef] [Green Version]
  8. Mao, L.; Sotomura, T.; Nakatsu, K.; Koshiba, N.; Zhang, D.; Ohsaka, T. Electrochemical Characterization of Catalytic Activities of Manganese Oxides to Oxygen Reduction in Alkaline Aqueous Solution. J. Electrochem. Soc. 2002, 149, A504. [Google Scholar] [CrossRef]
  9. Su, H.-Y.; Gorlin, Y.; Man, I.C.; Calle-Vallejo, F.; Nørskov, J.K.; Jaramillo, T.F.; Rossmeisl, J. Identifying active surface phases for metal oxide electrocatalysts: A study of manganese oxide bi-functional catalysts for oxygen reduction and water oxidation catalysis. Phys. Chem. Chem. Phys. 2012, 14, 14010. [Google Scholar] [CrossRef]
  10. Tian, L.; Wang, J.; Wang, K.; Wo, H.; Wang, X.; Zhuang, W.; Li, T.; Du, X. Carbon-quantum-dots-embedded MnO2 nanoflower as an efficient electrocatalyst for oxygen evolution in alkaline media. Carbon 2019, 143, 457–466. [Google Scholar] [CrossRef]
  11. Yang, J.; Xu, J.J. Nanoporous amorphous manganese oxide as electrocatalyst for oxygen reduction in alkaline solutions. Electrochem. Commun. 2003, 5, 306–311. [Google Scholar] [CrossRef]
  12. Yeager, E. Dioxygen electrocatalysis: Mechanisms in relation to catalyst structure. J. Mol. Catal. 1986, 38, 5–25. [Google Scholar] [CrossRef]
  13. Calegaro, M.L.; Lima, F.H.B.; Ticianelli, E.A. Oxygen reduction reaction on nanosized manganese oxide particles dispersed on carbon in alkaline solutions. J. Power Sources 2006, 158, 735–739. [Google Scholar] [CrossRef]
  14. Cheng, F.; Chen, J. Metal–air batteries: From oxygen reduction electrochemistry to cathode catalysts. Chem. Soc. Rev. 2012, 41, 2172. [Google Scholar] [CrossRef] [PubMed]
  15. Cheng, F.; Su, Y.; Liang, J.; Tao, Z.; Chen, J. MnO2 -Based Nanostructures as Catalysts for Electrochemical Oxygen Reduction in Alkaline Media . Chem. Mater. 2010, 22, 898–905. [Google Scholar] [CrossRef]
  16. Christensen, P.A.; Hamnett, A.; Linares-Moya, D. Oxygen reduction and fuel oxidation in alkaline solution. Phys. Chem. Chem. Phys. 2011, 13, 5206. [Google Scholar] [CrossRef]
  17. Eftekhari, A. Tuning the electrocatalysts for oxygen evolution reaction. Mater. Today Energy 2017, 5, 37–57. [Google Scholar] [CrossRef]
  18. Sweth, J.A.; Geetha, A.; Ramamurthi, K. Morphological and structural analysis of manganese oxide nanoflowers prepared under different reaction conditions. Appl. Surf. Sci. 2018, 449, 228–232. [Google Scholar] [CrossRef]
  19. Liu, X.; Chen, C.; Zhao, Y.; Jia, B. A Review on the Synthesis of Manganese Oxide Nanomaterials and Their Applications on Lithium-Ion Batteries. J. Nanomater. 2013, 2013, 1–7. [Google Scholar] [CrossRef]
  20. Mahmudi, M.; Widiyastuti, W.; Nurlilasari, P.; Affandi, S.; Setyawan, H. Manganese dioxide nanoparticles synthesized by electrochemical method and its catalytic activity towards oxygen reduction reaction. J. Ceram. Soc. Jpn. 2018, 126, 906–913. [Google Scholar] [CrossRef] [Green Version]
  21. 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] [PubMed]
  22. Ilnicka, A.; Skorupska, M.; Romanowski, P.; Kamedulski, P.; Lukaszewicz, J.P. Improving the Performance of Zn-Air Batteries with N-Doped Electroexfoliated Graphene. Materials 2020, 13, 2115. [Google Scholar] [CrossRef]
  23. Haoran, Y.; Lifang, D.; Tao, L.; Yong, C. Hydrothermal Synthesis of Nanostructured Manganese Oxide as Cathodic Catalyst in a Microbial Fuel Cell Fed with Leachate. Sci. World J. 2014, 2014, 1–6. [Google Scholar] [CrossRef]
  24. Lima, F.H.B.; Calegaro, M.L.; Ticianelli, E.A. Electrocatalytic activity of manganese oxides prepared by thermal decomposition for oxygen reduction. Electrochim. Acta 2007, 52, 3732–3738. [Google Scholar] [CrossRef]
  25. Sim, H.; Lee, J.; Yu, T.; Lim, B. Manganese oxide with different composition and morphology as electrocatalyst for oxygen evolution reaction. Korean J. Chem. Eng. 2018, 35, 257–262. [Google Scholar] [CrossRef]
  26. Valim, R.B.; Santos, M.C.; Lanza, M.R.V.; Machado, S.A.S.; Lima, F.H.B.; Calegaro, M.L. Oxygen reduction reaction catalyzed by ε-MnO2: Influence of the crystalline structure on the reaction mechanism. Electrochim. Acta 2012, 85, 423–431. [Google Scholar] [CrossRef]
  27. Zhou, D.; Lü, X.; Liu, D. Electro-catalytic effect of manganese oxide on oxygen reduction at teflonbonded carbon electrode. Trans. Nonferrous Met. Soc. China 2006, 16, 217–222. [Google Scholar] [CrossRef]
  28. Giménez, S. Photoelectrochemical Solar Fuel Production; Springer International Publishing: Berlin, Germany, 2016; ISBN 978-3-319-29641-8. [Google Scholar]
  29. Selvakumar, K.; Senthil Kumar, S.M.; Thangamuthu, R.; Ganesan, K.; Murugan, P.; Rajput, P.; Jha, S.N.; Bhattacharyya, D. Physiochemical Investigation of Shape-Designed MnO2 Nanostructures and Their Influence on Oxygen Reduction Reaction Activity in Alkaline Solution. J. Phys. Chem. C 2015, 119, 6604–6618. [Google Scholar] [CrossRef]
  30. Chen, Y.; Hong, Y.; Ma, Y.; Li, J. Synthesis and formation mechanism of urchin-like nano/micro-hybrid α-MnO2. J. Alloys Compd. 2010, 490, 331–335. [Google Scholar] [CrossRef]
  31. Hashem, A.M.; Abdel-Ghany, A.E.; El-Tawil, R.; Bhaskar, A.; Hunzinger, B.; Ehrenberg, H.; Mauger, A.; Julien, C.M. Urchin-like α-MnO2 formed by nanoneedles for high-performance lithium batteries. Ionics 2016, 22, 2263–2271. [Google Scholar] [CrossRef]
  32. Rossouw, M.H.; Liles, D.C.; Thackeray, M.; David, W.I.F.; Hull, S. Alpha manganese dioxide for lithium batteries: A structural and electrochemical study. Mater. Res. Bull. 1992, 27, 221–230. [Google Scholar] [CrossRef]
  33. Shende, P.; Kasture, P.; Gaud, R.S. Nanoflowers: The future trend of nanotechnology for multi-applications. Artif. Cellsnanomed. Biotechnol. 2018, 46, 413–422. [Google Scholar] [CrossRef] [PubMed]
  34. Tahir, M.; Pan, L.; Idrees, F.; Zhang, X.; Wang, L.; Zou, J.-J.; Wang, Z.L. Electrocatalytic oxygen evolution reaction for energy conversion and storage: A comprehensive review. Nano Energy 2017, 37, 136–157. [Google Scholar] [CrossRef]
  35. Poolnapol, L.; Kao-ian, W.; Somwangthanaroj, A.; Mahlendorf, F.; Nguyen, M.T.; Yonezawa, T.; Kheawhom, S. Silver Decorated Reduced Graphene Oxide as Electrocatalyst for Zinc–Air Batteries. Energies 2020, 13, 462. [Google Scholar] [CrossRef] [Green Version]
  36. Corpuz, R.D.; De Juan-Corpuz, L.M.; Nguyen, M.T.; Yonezawa, T.; Wu, H.-L.; Somwangthanaroj, A.; Kheawhom, S. Binder-Free α-MnO2 Nanowires on Carbon Cloth as Cathode Material for Zinc-Ion Batteries. Int. J. Mol. Sci. 2020, 21, 3113. [Google Scholar] [CrossRef] [PubMed]
  37. Kumar, N.; Dineshkumar, P.; Rameshbabu, R.; Sen, A. Facile size-controllable synthesis of single crystalline β-MnO2 nanorods under varying acidic strengths. RSC Adv. 2016, 6, 7448–7454. [Google Scholar] [CrossRef] [Green Version]
Figure 1. SEM images of synthesized MnO2 with 2 (a), 4 (b), and 6 h (c) dwell time.
Figure 1. SEM images of synthesized MnO2 with 2 (a), 4 (b), and 6 h (c) dwell time.
Catalysts 10 00822 g001
Figure 2. N2 adsorption/desorption isotherms of all the samples.
Figure 2. N2 adsorption/desorption isotherms of all the samples.
Catalysts 10 00822 g002
Figure 3. Cyclic voltammetry of pure carbon Vulcan XC 72, Nanoflower-like MnO2, nanoflower–nanowire-like MnO2, nanowire-like MnO2 in O2 saturated with 0.1-M KOH at a 50-mV/s scan rate.
Figure 3. Cyclic voltammetry of pure carbon Vulcan XC 72, Nanoflower-like MnO2, nanoflower–nanowire-like MnO2, nanowire-like MnO2 in O2 saturated with 0.1-M KOH at a 50-mV/s scan rate.
Catalysts 10 00822 g003
Figure 4. Linear sweep voltammetry curves of all prepared MnO2 catalysts.
Figure 4. Linear sweep voltammetry curves of all prepared MnO2 catalysts.
Catalysts 10 00822 g004
Figure 5. Koutecky–Levich plot for (a) nanoflower-like MnO2, (b) nanoflower–nanowire-like MnO2 and (c) nanowire-like MnO2/carbon Vulcan XC 72 electrode at the various potentials.
Figure 5. Koutecky–Levich plot for (a) nanoflower-like MnO2, (b) nanoflower–nanowire-like MnO2 and (c) nanowire-like MnO2/carbon Vulcan XC 72 electrode at the various potentials.
Catalysts 10 00822 g005
Figure 6. Linear sweep voltammetry curves for OER at 1600 rpm from 0.2 V to −1.0 V at the scan rate of 5 mV/s. (vxc72: Vulcan XC 72).
Figure 6. Linear sweep voltammetry curves for OER at 1600 rpm from 0.2 V to −1.0 V at the scan rate of 5 mV/s. (vxc72: Vulcan XC 72).
Catalysts 10 00822 g006
Figure 7. Galvanostatic discharge profiles of the Vulcan XC 72/commercial MnO2, Vulcan XC 72/nanoflower-like MnO2, Vulcan XC 72 mixed with major MnO2 and minor manganese wires and Vulcan XC 72 mixed with major nanowires and minor nanoflower at a constant discharge current of 5 mA. (vxc72: Vulcan XC 72).
Figure 7. Galvanostatic discharge profiles of the Vulcan XC 72/commercial MnO2, Vulcan XC 72/nanoflower-like MnO2, Vulcan XC 72 mixed with major MnO2 and minor manganese wires and Vulcan XC 72 mixed with major nanowires and minor nanoflower at a constant discharge current of 5 mA. (vxc72: Vulcan XC 72).
Catalysts 10 00822 g007
Figure 8. Schematic diagram of the coin cell battery.
Figure 8. Schematic diagram of the coin cell battery.
Catalysts 10 00822 g008
Table 1. Onset potential and overpotential for all the catalysts and pure carbon Vulcan XC 72.
Table 1. Onset potential and overpotential for all the catalysts and pure carbon Vulcan XC 72.
SampleOnset Potential at 5 mA/cm2, VOverpotential, V
NF-MnO2//Vulcan XC 72 30%0.890.57
NF/NW-MnO2//Vulcan XC 72 30%0.910.59
NW-MnO2//Vulcan XC 72 30%0.950.63
Vulcan XC 720.980.66
Table 2. Summary of the ORR and OER parameters of recently reported nanostructure manganese oxide catalyst.
Table 2. Summary of the ORR and OER parameters of recently reported nanostructure manganese oxide catalyst.
CatalystNanostructureElectrolyte and Reference ElectrodeSynthesizing MethodBET Surface Area (m2/g)Overpotential for ORR (mV) at −0.1 mA/cm2Electrons Transferred for ORR from K-L PlotCathodic Tafel Slope (mV dec−1)Overpotential for OER (mV) at Specific DensityAnodic Tafel Slope (mV dec−1)Ref.
α-MnO2Nanoflower0.1 M KOH and Ag/AgClHydrothermal52.4−0.5703.68-0.89 at 5 mA/cm2-This study
α-MnO2Nanoflower/Nanowires34.9−0.5903.31-0.91 at 5 mA/cm2-
α-MnO2Nanowires32.4−0.5303.00-0.95 at 5 mA/cm2-
α-MnO2Nanoflower0.1 M KOH and Ag/AgClHydrothermal68.3−0.3023.7---[15]
α-MnO2Nanowires40.1−0.5003.87---
α-MnO2Nanowires0.1 M KOH and SCEHydrothermal27.7−0.6163.565--[29]
α-MnO2Nanotubes21.1−0.5863.090--
α-MnO2Nanoparticles34.7−0.7362.390--
α-MnO2Nanorod24.8−0.6063.265--
α-MnO2Nanoflower32.4−0.8761.9115--
β-MnO2Nanorod0.1 M KOH and Ag/AgClHydrothermal37.9−0.75----[37]
α-MnO2Nanorod0.1 M KOH and Ag/AgClElectrosynthesis59.58−0.3512.23---[20]
α-MnO2Nanoflakes59.58−0.6511.75---
β-MnO2Nanorod0.1 M KOH and SCESolid state method5−0.5512.4-0.6 at 10 mA/cm2180.2[21]
δ-MnO2NanoflowerHydrothermal26−0.7011.7-0.75 at 10 mA/cm2188.6
Porous Mn2O3Nanoplates0.1 M KOH and Hg/HgOWet-chemical-----81[25]

Share and Cite

MDPI and ACS Style

Jing Han, S.; Ameen, M.; Hanifah, M.F.R.; Aqsha, A.; Bilad, M.R.; Jaafar, J.; Kheawhom, S. Catalytic Evaluation of Nanoflower Structured Manganese Oxide Electrocatalyst for Oxygen Reduction in Alkaline Media. Catalysts 2020, 10, 822. https://doi.org/10.3390/catal10080822

AMA Style

Jing Han S, Ameen M, Hanifah MFR, Aqsha A, Bilad MR, Jaafar J, Kheawhom S. Catalytic Evaluation of Nanoflower Structured Manganese Oxide Electrocatalyst for Oxygen Reduction in Alkaline Media. Catalysts. 2020; 10(8):822. https://doi.org/10.3390/catal10080822

Chicago/Turabian Style

Jing Han, Siow, Mariam Ameen, Mohamad Fahrul Radzi Hanifah, Aqsha Aqsha, Muhammad Roil Bilad, Juhana Jaafar, and Soorathep Kheawhom. 2020. "Catalytic Evaluation of Nanoflower Structured Manganese Oxide Electrocatalyst for Oxygen Reduction in Alkaline Media" Catalysts 10, no. 8: 822. https://doi.org/10.3390/catal10080822

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