Next Article in Journal
Impedimetric Early Sensing of Volatile Organic Compounds Released from Li-Ion Batteries at Elevated Temperatures
Previous Article in Journal
Micro-Sized MoS6@15%Li7P3S11 Composite Enables Stable All-Solid-State Battery with High Capacity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Developing a Se Quantum Dots@ CoFeOx Composite Nanomaterial as a Highly Active and Stable Cathode Material for Rechargeable Zinc–Air Batteries

1
Tianjin Key Laboratory of Composite and Functional Materials, Key Laboratory of Advanced Ceramics and Machining Technology (Ministry of Education), School of Materials Science and Engineering, Tianjin University, Tianjin 300350, China
2
State Key Laboratory of Marine Resource Utilization in South China Sea, School of Materials Science and Engineering, Hainan University, Haikou 570228, China
3
Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Binhai New City, Fuzhou 350207, China
*
Author to whom correspondence should be addressed.
Batteries 2023, 9(11), 561; https://doi.org/10.3390/batteries9110561
Submission received: 6 October 2023 / Revised: 25 October 2023 / Accepted: 27 October 2023 / Published: 17 November 2023

Abstract

:
With the urgent demand for clean energy, rechargeable zinc–air batteries (ZABs) are attracting increasing attention. Precious-metal-based electrocatalysts (e.g., commercial Pt/C and IrO2) are reported to be highly active towards the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER). Nevertheless, the limited catalytic kinetics, along with the scarcity of noble metals, still hinder the practical applications of ZABs. Consequently, it is of great importance to explore efficient bifunctional ORR/OER electrocatalysts with abundant reserves. Although iron oxides are considered to have some of the greatest potential as catalysts among the metal oxides, owing to their excellent redox properties, lower toxicity, simple preparation, and natural abundance, their poor electrical conductivity and high agglomeration still limit their development. In this work, we report a special Se quantum dots@ CoFeOx (Se-FeOx-Co) composite material, which exhibits outstanding bifunctional catalytic properties. And the potential gap between ORR and OER is low at 0.87 V. In addition, the ZAB based on Se-FeOx-Co achieves a satisfactory open-circuit voltage (1.46 V) along with an operation durability over 800 min. This research explores an effective strategy to fabricate iron oxide-based bifunctional catalysts, which contributes to the future design of related materials.

1. Introduction

In recent years, the overuse of fossil resources has brought about increasingly serious energy crises and environmental pollution. To meet global energy demands, it is strongly necessary to explore clean renewable energy. Electrochemical energy storage devices with sufficient energy output, stable performance, and high energy density are drawing increasing attention. Among the various devices, batteries (including metal-ion batteries and metal–air batteries) are regarded as the most promising energy technologies [1,2]. Among the various kinds of batteries, rechargeable zinc–air batteries (ZABs) are receiving great interest owing to many advantages, such as high energy density, environmental friendliness, and the abundance of Zn raw materials [3,4,5]. A reversible air–cathodic electrochemical reaction (O2 + 2H2O + 4e ↔4OH) proceeds in the catalysts’ surfaces, while an anodic electrochemical reaction (Zn + 4OH−2e→Zn (OH)42−, Zn (OH)42−↔ZnO + H2O + 2OH) occurs in Zn plates in the alkaline electrolyte. However, one of the main issues is the suitable selection and design of the bifunctional catalysts as the air cathode. The sluggish kinetics of the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) of catalysts would lead to a large potential gap and low round-trip efficiency. As a result, the further development of ZABs is hindered [6,7]. According to the reaction mechanism, the theoretical voltage of ZAB is determined to be 1.65 V. However, the actual working voltage is generally lower than 1.2 V, with a much higher charging voltage (>2 V), and the current round-trip energy efficiency is usually less than 60%. Precious metal catalysts (e.g., commercial Pt/C and IrO2) prove to be highly active catalysts for ORR and OER [8,9]. The limited reserves still preclude the large-scale applications of ZABs [10]. Consequently, rationally exploring bifunctional ORR/OER electrocatalysts with high activity and abundant reserves is necessary.
To date, non-noble transitional metal-based materials (such as Fe, Co, and Mn), especially sulfide, oxide, phosphide, nitride, and selenide, have been reported to be viable candidates to replace noble metal electrocatalysts because of their remarkable catalytic abilities [11]. As previously reported, Zou and co-workers synthesized a novel catalyst with Co sites anchored with a CoP electrocatalyst for oxygen electrocatalysis in ZAB, exhibiting a high power density of 139.8 mW cm−2 and rechargeability of 132 h [12]. Zhang et al. prepared a N-doped carbon-tentacles-modified metal (Fe or Co) Se nanomaterial through a scalable as well as facile solid-state strategy, which exhibited superior activity and durability thanks to abundant active sites [13]. Xiao et al. designed Co3O4 nanoparticles with atomically dispersed FeN2, and the FeN2 was determined to be the real catalytic active site in oxygen electrocatalysis [14]. Ren et al. made a rational design for the heterojunction structure of FeS2/NiS2, which has a potential gap of 0.92 V when utilized in a liquid ZAB [15]. Among these, Co, Fe, and Ni-based oxide catalysts have been attracting more and more extensive attention because of their high nontoxicity [16,17,18,19]. In particular, ABO3-type perovskites and the AB2O4-type spinel oxides have been widely used as bifunctional catalysts in metal–air batteries (A and B are different metal elements) [20,21,22,23]. It is worth mentioning that metal oxides can provide abundant oxidized metal sites on the catalyst surface and are easy to tune and affect the electronic structure, thereby improving the electrocatalysis performance [24].
Although Co oxides and Ni oxides have been widely reported as bifunctional catalysts in previous works in the literature, iron oxides have been less studied. Iron oxides, as a kind of well-known metal oxide, have attracted numerous research efforts in the scientific and commercial fields because of their excellent redox properties, lower toxicity, simple preparation, and natural abundance [25]. Unfortunately, iron oxides have an inherently poor electrical conductivity and high agglomeration, which would lead to rapid recession and poor cycling durability [26]. This issue creates poor ORR and OER performance in catalysts, so iron-based oxides are still far below expectation. Hence, various strategies have been put forward to improve the unsatisfactory activity or stability of iron oxides, including exposing active sites, reducing their nanoscale, and so on [27,28,29]. In recent years, designing composite materials with highly conductive additives has been widely reported to significantly improve their dispersion and catalytic activity [30,31,32]. For example, Wang et al. reported short carbon nanotubes supporting nickel–iron oxides, where the carbon composite confers a high electrochemical surface area and satisfactory electron transfer ability [31]. Li et al. synthesized heterostructural FeOx/Fe nanoparticles based on a nitrogen-doped carbon framework, which displayed high ORR activity and good durability due to the high electroconductivity and proper hydrophilicity of the carbon framework [33]. However, it is more desirable if the intrinsic conductivity and stability of iron oxide can be substantially boosted, which may result in a breakthrough in catalytic performance. Quantum dots, as a novel 0D material, are widely considered to be one of the most fascinating materials because of their specific electrical, optical, and chemical properties. Due to the small nanometer scale, quantum dots have some specific quantum tunneling effects and quantum size effects, which are widely used in the field of catalysis, optical, and chemical properties. Owing to their small nanometer scale, QDs have specific quantum tunneling and quantum size effects, which are widely harnessed in the fields of catalysis, optical sensing, and solar-harvesting devices [34,35]. The introduction of QDs into composites can improve their electrocatalytic properties by enhancing their electrical conductivity and energy conversion efficiency [36,37].
Herein, we report a facile one-pot hydrothermal process for synthesizing Co-doped iron oxide nanosheets loaded with Se QDs. The Co precursor ratio was optimized to enhance the ORR and OER performances. The catalytic activity and stability sharply increased when Co was successfully incorporated into the original Fe2O3 lattice. In particular, the special Se QD@ CoFeOx (Se-FeOx-Co) composite material exhibited remarkable catalytic behavior for the alkaline ORR and OER. In detail, the half-wave potential of ORR was 0.64 V. For OER, it required a reduced overpotential (280 mV) to reach 10 mA cm−2. Moreover, when Se-FeOx-Co served as an air cathode in a ZAB in a liquid electrolyte, the optimized material showed good catalytic activity and stability for at least 800 min. Our results not only explore an effective strategy to optimize the bifunctional performance of iron oxides but also contribute to the future design of energy-related electrocatalysts.

2. Materials and Methods

2.1. Reagents and Materials

Cobaltous nitrate hexahydrate (Co (NO3)2·6H2O, 95%) and ammonium iron (II) sulfate hexahydrate (NH4Fe (SO4)2·6H2O, 98%) were produced by the Adamas Reagent Co., Ltd., Shanghai, China. Zinc acetate dihydrate (Zn (CH3COO)2·2H2O)) was obtained from Shanghai Macklin Biochemical Co., Ltd., Shanghai, China. Hydrazine hydrate (N2H4· H2O, AR, 99%) was purchased from the Yuanli Chemical Technology Co., Ltd., Tianjin, China. 5 wt.% Nafion solution was obtained from Sigma-Aldrich Co., Ltd., St. Shanghai, China. Potassium hydroxide (KOH, 95%) was obtained from Jiangtian Chemical Technology Co., Ltd., Tianjin, China.

2.2. Synthesis of Se-FeOx-Co Composite Materials

Firstly, 1.4 mmol Co (NO3)2· 6H2O and 4.2 mmol NH4Fe (SO4)2·6H2O were dissolved in 25 mL of deionized water in a clean beaker. After 30 min magnetic stirring, 0.11 g of NH4Cl, 0.4 g of NH4F, 0.31 g of SeO2, and 5 mL of hydrazine hydrate were added. Then, the mixing solution was transferred into a 50 mL PTFE-lined stainless autoclave and placed in an oven at 180 °C for 12 h. The as-synthesized precipitation was obtained via a centrifugation process. And the product was washed with DI water three times. Then, the products were dried in an oven at 60 °C for 12 h. According to the different addition amounts of Co (NO3)2· 6H2O (0.7 and 1.4 mmol), the products were named Se-FeOx-Co-1 and Se-FeOx-Co-2, respectively. In addition, the sample to which no Co precursor was added was named Se-FeOx.

2.3. Materials’ Characterization

X-ray diffraction (XRD) (Cu Kα radiation, D8 Advanced, 5 min−1) was used to measure the crystal structural information of the samples. The morphologies and characteristic structures of Se-FeOx-Co series materials were determined using a JSM-7800F field emission scanning electron microscope (SEM). Before characterization, the dried powder samples were added into absolute ethyl alcohol. After ultrasonic dispersion, the black solution was added dropwise to clean and smooth pieces of Al foil. After drying under room temperature, those pieces of Al foil were transferred to the vacuum box of the SEM device. In addition, the above-mentioned solution was further dropped on a copper mesh that supported an ultra-thin carbon film and then treated using a vacuum oven. The interior structure of the Se-FeOx-Co series materials was then explored via high-resolution transmission electron microscopy (HRTEM), which was carried out with a JEM-2100F device. The Raman spectra of the samples in the range of 100–1600 cm−1 were obtained using a multichannel Raman system (Horiba Jobin Yvon, Kyoto, Japan) with a 532 nm laser. The valence states and chemical information of the samples were measured with an energy-dispersive X-ray photoelectron spectroscope (XPS, Axis Supra, Kratos Analytical Ltd., Manchester, UK).

2.4. Electrochemical Measurements

Homogeneous inks from the ORR and OER tests were obtained by ultrasonically dispersing 7 mg of pre-synthesized sample, 3 mg of black carbon, and 35 µL of Nafion in 965 µL of isopropanol for 1 h. Then, the inks were drop-casted on electrodes: a rotating ring-disk electrode (RRDE, with a diameter of 5 mm) for the ORR test and a glassy carbon electrode (with diameter of 5 mm) for the OER test. The catalyst loading was calculated to be 2.8 mg cm−2.
All of the electrochemical performance tests (including linear sweep voltammetry (LSV) curves, cyclic voltammetry (CV), and chronoamperometry (CA)) in this study were performed at room temperature on a CS310X multi-channel electrochemical workstation (Corrtest). For the ORR test, the RRDE electrode modified by a catalyst served as the working electrode. The counter electrode was a carbon rod, and the reference electrode was Hg/HgO. All the ORR measurements were tested in a 0.1 M KOH solution. The reversible hydrogen electrode (ERHE) and the measuring potential (ESCE) can be converted via ERHE = EHg/HgOθ + 0.059pH + ESCE. Before each CV and LSV test, pure O2 was purged into the alkaline electrolyte for more than 1h in order to make the electrolyte saturated. During the LSV tests, the RRDE measurements of the samples were conducted via changing rotation rates (400, 625, 900, 1225, and 1600 rpm). In addition, electrocatalytic performances were compared at 1600 rpm in the O2-saturated electrolyte. During the OER test, the working electrode was a catalyst-modified glassy carbon electrode. The counter electrode was a carbon rod and the reference electrode was Hg/HgO. All the OER measurements were tested in a 1 M KOH solution. The pure N2 was saturated in the electrolyte for eliminating the effect of oxygen in the OER process.
The electrochemical performance of rechargeable Zn-air batteries was tested in a two-electrode system and was analyzed using a multichannel potential station (LAND CT2001A). Carbon cloths were dipped in dilute HNO3 solution and deionized water under ultrasonic conditions for 30min several times, and then washed with deionized water and absolute ethyl alcohol. The above-mentioned ink from the catalysts was dropped uniformly onto a hydrophilic carbon cloth (with an area of 1 × 1 cm2). And the loading of catalysts on the cathode was calculated to be 2.8 mg cm−2. Moreover, the polished Zn plate as the anode and 6 M KOH + 0.2 M Zn (CH3COO)2 as the liquid electrolyte were constitutive of the rechargeable liquid ZABs. All batteries were tested at a current density of 10 mA cm−2.

3. Results and Discussion

Figure 1a schematically illustrates the preparation of the Se-FeOx-Co catalysts. Briefly, the SeO2 precursor decomposed to form Se QDs and O2 during the one-pot hydrothermal process, and O2 combined with the Fe3+ precursor to form iron oxide. Subsequently, numerous Co2+ ions were doped into the crystal lattice of the iron oxide. Figure 1b shows that the XRD pattern of Se-FeOx had clear peaks corresponding to elemental Se (PDF# 06-0362) and Fe2O3 (PDF# 39-1346). With the addition of the Co2+ precursor, the relative peaks of crystallographic Fe2O3 disappeared, indicating that Co affected the lattice formation of the samples. No heterophase was observed in the Se-FeOx-Co samples, suggesting the successful incorporation of Co ions into Se-FeOx [38]. The morphologies of the Se-FeOx-Co series are shown in Figure 1c,d, Figures S1 and S2. Se-FeOx shows a typical nanoflower structure. As the Co content increased, the 3D flower structure tended to deform in Se-FeOx-Co-1. The Se-FeOx-Co-2 sample exhibited a hierarchical structure consisting of two-dimensional (2D) sheets ~100 nm in length. According to previous reports, the unique 2D sheet structure may significantly improve the exposure of the metal centers and is conducive to wetting the electrolyte [39,40]. In addition, the microscopic structure of the Se-FeOx-Co-2 sample was further explored using HRTEM. As shown in Figure 1e, Se was observed as QDs with sizes of <2 nm. Notably, Se-FeOx-Co-2 showed that the (311) plane of crystallographic Fe2O3 was coated with an amorphous structure. Compared to Se-FeOx, the crystallinity of Se-FeOx-Co-2 was much lower, which indicates that Co was successfully incorporated into the original Fe2O3 lattice. The EDS elemental mapping analysis in Figure 1f demonstrates that the Co, Fe, and O atoms are homogeneously dispersed in Se-FeOx-Co-2, whereas the Se atoms in the region are mainly aggregated as well-distributed small particles. The atomic contents of Co, Fe, Se, and O were further measured (Figure 1g) and found to be 13.96%, 38.81%, 2.26%, and 44.97%, respectively. Therefore, the distribution of Fe, Co, and O signals is quite uniform throughout the Se-FeOx-Co-2 sample, illustrating the successful doping of Co into the Se-FeOx matrix. The low content and intensity of Se were due to the small number of Se QDs.
To further explore the structures of the as-synthesized Se-FeOx and Se-FeOx-Co-2 samples, Raman spectra of the two samples were acquired (Figure 2a). On the one hand, the peaks at 216 cm−1 belong to the A1g mode of Fe2O3. On the other hand, the peaks at 283 cm−1 could be attributed to the 2Eg mode of Fe2O3. The peaks at 390 and 594 cm−1 in Se-FeOx are assigned to the Eg modes of Fe2O3. The wide peak at ca. 1300 cm−1 is assigned to the second-order scattering of Fe2O3 [41,42]. Therefore, the existence of Fe2O3 is further suggested for all the samples [43]. In addition, no noticeable peaks are assigned to the Se–O stretching vibration in the Se-FeOx and Se-FeOx-Co-2 samples, indicating the highly stable existence of Se QDs in air with the formation of Se-Se bonds [44]. Notably, the peaks of the two samples slightly differed. Compared with the Se-FeOx sample, the Eg modes of Fe2O3 in the Se-FeOx-Co-2 sample slightly changed, and a prominent peak assigned to the Eg mode of O–Co–O bending vibration appeared at 474 cm−1. The structural and valence properties were further investigated using XPS, as shown in Figure 2b–f. The full XPS spectrum in Figure 2b shows typical peaks of Co 2s, Co 2p, Fe 2p, O1s, Fe 3s, Fe 3p, Se 3s, Se 3d, and Se 3p in the Se-FeOx-Co-2 sample, indicating the presence of Se, Co, Fe, and O elements. In Figure 2c, the Se 3d spectra of Se-FeOx and Se-FeOx-Co-2 display two characteristic peaks at 54.6 and 55.4 eV belonging to the Se0 3d5/2 and Se0 3d3/2 orbits of Se QDs [45,46]. When the Co precursor was added, the characteristic peaks of Se in Se-FeOx-Co-2 hardly changed. Thus, we inferred that the doping of Co exerted nearly no influence on the formation of Se QDs. Moreover, the O 1s and Fe 2p spectra show clear differences between the Se-FeOx and Se-FeOx-Co-2 samples. In Figure 2d, the high-resolution spectrum of elemental oxygen for the catalyst is deconvoluted into peaks at 529.6, 531.2, and 533.2 eV, which correspond to the metal-O bond (Fe-O bond or Co-O bond), oxygen vacancies, and -OH groups (oxygen in the surface-absorbed water molecule) [38,47,48]. In particular, the Se-FeOx-Co-2 sample had a higher concentration of oxygen vacancies than the Se-FeOx sample, suggesting that Co ions were successfully incorporated into the FeOx lattice, thus breaking the bonds between Fe and O atoms. Oxygen vacancies, which are generated by Co doping, are favorable for improving electrical conductivity [49,50,51]. As shown in Figure 2e, two main peaks at 724.8 and 710.9 eV belong to Fe 2p 1/2 and Fe 2p 3/2, respectively. The fitted curves of Fe 2p 3/2 are deconvoluted into two peaks at 713.3 and 710.6 eV, which are ascribed to Fe2+ and Fe3+, respectively [52]. Moreover, the wide peak at 718.6 eV is assigned to the shake-up satellite peak of Fe3+ [53,54]. Compared to the Se-FeOx sample, the Fe 2p spectrum of Se-FeOx-Co-2 exhibits a positive shift with the successful replacement of Fe by Co, indicating a rich electron state [55]. Furthermore, compared with Se-FeOx-Co-1, the Co 2p peak of Se-FeOx-Co-2 also displays a slightly opposite shift (Figure S3), indicating electron acceptance [56]. These shifted binding energies demonstrate a strong electronic interaction between Fe and Co, which promotes the transformation of valence electrons and improves the intrinsic electrocatalytic activity [57]. As shown in Figure 2f, this Co 2p 1/2 peak can be divided into 787.7, 780.4, and 782.3 eV, corresponding to the satellite peaks of Co3+ and Co2+, respectively [58]. The high Co3+/Co2+ in the Se-FeOx-Co-2 sample demonstrates that the Co valence in Se-FeOx-Co-2 mainly exists as Co3+. In conclusion, the incorporated Co effectively transformed more electrons into Fe, which was beneficial for increasing catalytic performance.
After characterizing the structural properties of the different catalysts, the electrocatalytic performances were evaluated in the KOH solution. The ORR and OER performances of those Se-FeOx-based catalysts were measured using LSV curves in a three-electrode system. Regarding the CV test of the Se-FeOx and Se-FeOx-Co-2 samples, Figure 3a displays distinct cathodic ORR peaks within the same voltage window at room temperature in 0.1 M KOH solution with an O2-saturated environment. The clear oxygen reduction peaks of Se-FeOx and Se-FeOx-Co-2 under an O2 atmosphere suggest their catalytic activity for oxygen conversion [59]. A peak at 0.64 V versus RHE is observed, which is higher than the Se-FeOx peak at 0.52 V in Se-FeOx. This demonstrates the prominent ORR activity for Se-FeOx-Co-2 [60]. In addition, the ORR peak current of Se-FeOx-Co-2 reaches 0.12 mA cm−2, which is better than that of Se-FeOx (with a peak current of 0.06 mA cm−2) [61]. As shown in Figure 3b, the ORR performance of the samples was compared at 1600 rpm. Se-FeOx-Co-2 had a high onset potential (Eonset = 0.76 V) and high half-wave potential (E1/2 = 0.64 V). However, the onset potential of Se-FeOx was 0.66 V, and the half-wave potential was 0.52 V. In addition, the onset potential of Se-FeOx-Co-1 was 0.70 V, with a half-wave potential of 0.59 V. Therefore, the Se-FeOx-Co-2 showed the best ORR performance. Further, Se-FeOx-Co-2 also exhibited excellent OER characteristics among the three catalysts. As shown in Figure 3c, it required an overpotential of only 280 mV to reach a current density of 10 mA·cm−2 for the Se-FeOx-Co-2. By contrast, Se-FeOx and Se-FeOx-Co-1 showed larger overpotentials of 400 and 290 mV to reach 10 mA·cm−2, respectively. The relative performance comparison of all catalysts is summarized in Figure 3d and Figure S4. It can be seen that the potential gaps of Se-FeOx, Se-FeOx-Co-1, and Se-FeOx-Co-2 are 1.11, 0.93, and 0.87 V, which implied that the as-prepared Se-FeOx-Co-2 showed significantly enhanced catalytic activity towards ORR and OER. Moreover, the lowered potential gap contributes to the charging/discharging performance of the ZAB devices. Compared with some reported references, the Se-FeOx-Co-2 material shows excellent performance in the same alkaline environment (Table S1) [62,63,64,65,66,67,68]. In addition, the stability of ORR and OER were analyzed using chronoamperometry. As shown in Figure 3e,f, after a 14 h measurement of the ORR process, 91.5% of the initial current density was maintained in Se-FeOx-Co-2, which was superior to Se-FeOx (only 59.2% after 10 h). In addition, Se-FeOx-Co-2 still showed long-term stability for 20 h in the OER process, which was much longer than that of Se-FeOx in the same conditions. Therefore, the Se-FeOx-Co-2 outperformed the other samples in both activity and stability. The enhanced electrocatalytic performance of Se-FeOx-Co-2 might be owing to the following reasons: First, the loosely stacked nanoflakes of Se-FeOx-Co-2 help to improve the electrochemical surface area, and therefore, more active sites are exposed [69,70]. Second, the Co dopants significantly enhance the conductivity of the FeOx and, thus, facilitate the charge and electron transfer during the electrochemical reaction, which contributes to improving the activity [71,72]. Third, as evidenced by the XPS results, an electronic interaction occurs after introducing Co into FeOx, and therefore, the synergistic effect between different metals significantly improves the electrocatalytic performance.
To further evaluate the practical properties of the Se-FeOx-Co-2 catalyst in energy conversion and storage devices, a rechargeable ZAB was assembled. A schematic of the operation of the ZABs is shown in Figure 4a. A ZAB assembled with the Se-FeOx catalyst was used as the control sample under the same conditions. As shown in Figure 4b,c, the ZAB with the Se-FeOx-Co-2 catalyst exhibited a stable open-circuit voltage far higher than that of Se-FeOx (1.16 V). As shown in Figure 4d, the Se-FeOx-Co-2-based ZAB delivered a dischargeable capacity of 391 mAh g−1, significantly larger than that of the Se-FeOx based ZAB (215 mAh g−1). This improved performance is related to the optimized oxygen catalysis properties of the Se-FeOx-Co-2 catalyst. Moreover, the battery, which consists of a Se-FeOx-Co-2 catalyst, exhibits a higher power density of 68 mW cm−2 than that of the Se-FeOx-based battery (only 54 mW cm−2) in Figure S5. The long-term cycling stability of the rechargeable ZAB with Se-FeOx-Co-2 and Se-FeOx was measured, as shown in Figure 4e. A slight voltage change was observed over 800 min during discharge/charge cycling, consistent with the good electrocatalytic stability of Se-FeOx-Co-2. The Se-FeOx-based ZAB showed poor cycling performance and could not operate normally after 570 min (with a low discharging platform of 0.5 V). The above results also confirmed that Se-FeOx-Co-2, with high activity and stability, has significant potential for application as an electrode catalyst in ZABs.

4. Conclusions

In conclusion, Se QD@ CoFeOx composite materials were synthesized using a hydrothermal process. During the synthesis process, the SeO2 precursor used as the Se source decomposed to form Se QDs and iron oxide was generated simultaneously. At high temperatures and pressures, numerous Co2+ ions were doped into the crystal lattice of iron oxide. The nanostructure and valence state of the materials were investigated via SEM, TEM, HRTEM, XRD, Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS). All successfully demonstrated the insertion of cobalt ions into the iron oxide lattice, which had a positive effect on the exposure of active sites, fast charge and electron transfer, and electron interactions in the electrochemical process. Because the Co ions were successfully incorporated into the iron oxide lattice, the as-prepared catalyst showed high activity for oxygen catalysis, with a potential gap of only 0.87 V. In addition, it displayed a satisfactory open-circuit voltage (1.46 V) and long-term durability of over 800 min when fabricated into a liquid ZAB. Therefore, the Se QD@ CoFeOx catalyst is one of the most promising candidates for practical applications. Our research explores new prospects for the development of cathodic materials for ZABs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/batteries9110561/s1, Figure S1: SEM and TEM images of Se-FeOx material; Figure S2: SEM and TEM images of Se-FeOx-Co-1 material; Figure S3: The Co 2p spectra of Se-FeOx-Co-1 and Se-FeOx-Co-2 materials. Figure S4: The relative bar in the value of Ej=10-E1/2; Figure S5: The polarization and power density curves of Se-FeOx and Se-FeOx-Co-2 materials as cathode. Table S1. The ORR and OER performance comparison with reported references.

Author Contributions

Conceptualization, D.Z., Y.W., X.H. and W.H; methodology, D.Z. and X.H.; investigation, X.H. and W.H.; resources, X.H. and W.H.; writing—original draft preparation, D.Z.; writing—review and editing, X.H.; supervision, X.H. and W.H.; project administration, X.H. and W.H.; funding acquisition, X.H. and W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number U22A20119.

Data Availability Statement

The data presented in this study are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, S.; Li, S.; Lu, Y. Designing safer lithium-based batteries with nonflammable electrolytes: A review. eScience 2021, 1, 163–177. [Google Scholar] [CrossRef]
  2. Yuan, M.; Sun, Z.; Yang, H.; Wang, D.; Liu, Q.; Nan, C.; Li, H.; Sun, G.; Chen, S. Self-catalyzed rechargeable lithium-air battery by in situ metal ion doping of discharge products: A combined theoretical and experimental study. Energy Environ. Mater. 2023, 6, e12258. [Google Scholar] [CrossRef]
  3. Cheng, F.; Chen, J. Metal–air batteries: From oxygen reduction electrochemistry to cathode catalysts. Chem. Soc. Rev. 2012, 41, 2172–2192. [Google Scholar] [CrossRef]
  4. Niu, Y.; Gong, S.; Liu, X.; Xu, C.; Xu, M.; Sun, S.-G.; Chen, Z. Engineering iron-group bimetallic nanotubes as efficient bifunctional oxygen electrocatalysts for flexible Zn–air batteries. eScinece 2022, 2, 546–556. [Google Scholar] [CrossRef]
  5. Yu, H.; Fan, F.; He, C.; Zhou, M.; Ma, T.; Wang, Y.; Cheng, C. Sulfur-modulated FeNi nanoalloys as bifunctional oxygen electrode for efficient rechargeable aqueous Zn-air batteries. Sci. China Mater. 2022, 65, 3007–3016. [Google Scholar] [CrossRef]
  6. Xin, L.; Yang, F.; Rasouli, S.; Qiu, Y.; Li, Z.-F.; Uzunoglu, A.; Sun, C.-J.; Liu, Y.; Ferreira, P.; Li, W.; et al. Understanding Pt nanoparticle anchoring on graphene supports through surface functionalization. ACS Catal. 2016, 6, 2642–2653. [Google Scholar] [CrossRef]
  7. Liu, H.; Liu, Q.; Wang, Y.; Wang, Y.; Chou, S.; Hu, Z.; Zhang, Z. Bifunctional carbon-based cathode catalysts for zinc-air battery: A review. Chin. Chem. Lett. 2022, 33, 683–692. [Google Scholar] [CrossRef]
  8. Zhang, W.; Lai, W.; Cao, R. Energy-related small molecule activation reactions: Oxygen reduction and hydrogen and oxygen evolution reactions catalyzed by porphyrin- and corrole-based systems. Chem. Rev. 2017, 117, 3717–3797. [Google Scholar] [CrossRef]
  9. Wang, X.; Li, Z.; Qu, Y.; Yuan, T.; Wang, W.; Wu, Y.; Li, Y. Review of metal catalysts for oxygen reduction reaction: From nanoscale engineering to atomic design. Chem 2019, 5, 1486–1511. [Google Scholar] [CrossRef]
  10. Peng, X.; Zhang, L.; Chen, Z.; Zhong, L.; Zhao, D.; Chi, X.; Zhao, X.; Li, L.; Lu, X.; Leng, K.; et al. Hierarchically porous carbon plates derived from wood as bifunctional ORR/OER electrodes. Adv. Mater. 2019, 31, 1900341. [Google Scholar] [CrossRef]
  11. Yu, J.; Li, Z.; Liu, T.; Zhao, S.; Guan, D.; Chen, D.; Shao, Z.; Ni, M. Morphology control and electronic tailoring of CoxAy (A = P, S, Se) electrocatalysts for water splitting. Chem. Eng. J. 2023, 460, 141674. [Google Scholar] [CrossRef]
  12. Liu, B.; Wang, X.; Wang, R.; Zhang, G.; Xu, X.; Liu, J.; Sun, Z.; Liu, M.; Wang, C.; Meng, X.; et al. Activating and stabilizing Co sites in CoP for triggering oxygen electrocatalysis in zinc-air battery. Chem. Eng. J. 2023, 475, 146154. [Google Scholar] [CrossRef]
  13. Qi, P.; Chen, M.; Luo, T.; Zhao, C.; Lin, C.; Luo, H.; Zhang, D. Solid-state self-catalyzed growth of N-doped carbon tentacles on an M(Fe, Co)Se surface for rechargeable Zn–air batteries. Chem. Commun. 2023, 59, 5898–5901. [Google Scholar] [CrossRef]
  14. Gui, F.; Jin, Q.; Xiao, D.; Jin, Z.; Zhang, Y.; Cao, Y.; Yang, M.; Tan, Q.; Zhang, C.; Siahrostami, S.; et al. High-performance zinc–air batteries enabled by hybridizing atomically dispersed FeN2 with Co3O4 nanoparticles. J. Mater. Chem. A 2023, 11, 1312–1323. [Google Scholar] [CrossRef]
  15. Wu, L.; Li, J.; Shi, C.; Li, Y.; Mi, H.; Deng, L.; Zhang, Q.; He, C.; Ren, X. Rational design of the FeS2/NiS2 heterojunction interface structure to enhance the oxygen electrocatalytic performance for zinc–air batteries. J. Mater. Chem. A 2022, 10, 16627–16638. [Google Scholar] [CrossRef]
  16. Wang, J.-J.; Li, X.-P.; Cui, B.-F.; Zhang, Z.; Hu, X.-F.; Ding, J.; Deng, Y.-D.; Han, X.-P.; Hu, W.-B. A review of non-noble metal-based electrocatalysts for CO2 electroreduction. Rare Met. 2021, 40, 3019–3037. [Google Scholar] [CrossRef]
  17. Wang, Y.; Wu, X.; Jiang, X.; Wu, X.; Tang, Y.; Sun, D.; Fu, G. Citrulline-induced mesoporous CoS/CoO heterojunction nanorods triggering high-efficiency oxygen electrocatalysis in solid-state Zn-air batteries. Chem. Eng. J. 2022, 434, 134744. [Google Scholar] [CrossRef]
  18. Weng, C.; Lv, X.; Ren, J.; Wang, Y.; Tian, W.; Gao, L.; Wang, H.; Yuan, Z. Self-promoted electrocatalysts derived from surface reconstruction for rechargeable Zinc–air batteries. ACS Sustain. Chem. Eng. 2022, 10, 6456–6465. [Google Scholar] [CrossRef]
  19. Zhang, Z.; Liang, X.; Li, J.; Qian, J.; Liu, Y.; Yang, S.; Wang, Y.; Gao, D.; Xue, D. Interfacial engineering of NiO/NiCo2O4 porous nanofibers as efficient bifunctional catalysts for rechargeable Zinc–air batteries. ACS Appl. Mater. Interfaces 2020, 12, 21661–21669. [Google Scholar] [CrossRef]
  20. Agyeman, D.; Zheng, Y.; Lee, T.; Park, M.; Tamakloe, W.; Lee, G.; Jang, H.; Cho, K.; Kang, Y. Synergistic catalysis of the lattice oxygen and transition metal facilitating ORR and OER in perovskite catalysts for Li–O2 batteries. ACS Catal. 2021, 11, 424–434. [Google Scholar] [CrossRef]
  21. Huang, H.; Huang, A.; Liu, D.; Han, W.; Kuo, C.; Chen, H.; Li, L.; Pan, H.; Peng, S. Tailoring oxygen reduction reaction kinetics on perovskite oxides via oxygen vacancies for low-temperature and knittable zinc–air batteries. Adv. Mater. 2023, 35, 2303109. [Google Scholar] [CrossRef] [PubMed]
  22. Kumar, R.S.; Prabhakaran, S.; Ramakrishnan, S.; Karthikeyan, S.C.; Kim, A.R.; Kim, D.H.; Yoo, D.J. Developing outstanding bifunctional electrocatalysts for rechargeable Zn-air batteries using high-purity spinel-type ZnCo2Se4 nanoparticles. Small 2023, 19, 2207096. [Google Scholar] [CrossRef] [PubMed]
  23. Guan, D.; Shi, C.; Xu, H.; Gu, Y.; Zhong, J.; Sha, Y.; Hu, Z.; Ni, M.; Shao, Z. Simultaneously mastering operando strain and reconstruction effects via phase-segregation strategy for enhanced oxygen-evolving electrocatalysis. J. Energy Chem. 2023, 82, 572–580. [Google Scholar] [CrossRef]
  24. Li, H.; Askari, S.; Wang, J.; Wolff, N.; Behrens, M.; Kienle, L.; Benedikt, J. Nitrogen-doped NiCo2O4 nanowires on carbon paper as a self-supported air cathode for rechargeable Zn-air batteries. Int. J. Hydrog. Energy 2023, 48, 26107–26118. [Google Scholar] [CrossRef]
  25. Bo, L.; Nian, F.; Pu, L.; Li, P.; Hu, Y.; Tong, J. Simply embedding α-Fe2O3/Fe3O4 nanoparticles in N-doped graphitic carbon polyhedron layered arrays as excellent electrocatalyst for rechargeable Zn-air battery. Sep. Purif. Technol. 2023, 312, 123413. [Google Scholar] [CrossRef]
  26. Samanta, A.; Das, S.; Jana, S. Doping of Ni in α-Fe2O3 nanoclews to boost oxygen evolution electrocatalysis. ACS Sustain. Chem. Eng. 2019, 7, 12117–12124. [Google Scholar] [CrossRef]
  27. Niu, Y.; Yuan, Y.; Zhang, Q.; Chang, F.; Yang, L.; Chen, Z.; Bai, Z. Morphology-controlled synthesis of metal-organic frameworks derived lattice plane-altered iron oxide for efficient trifunctional electrocatalysts. Nano Energy 2021, 82, 105699. [Google Scholar] [CrossRef]
  28. Zhu, S.; Tian, H.; Wang, N.; Chen, B.; Mai, Y.; Feng, X. Patterning graphene surfaces with iron-oxide-embedded mesoporous polypyrrole and derived N-doped carbon of tunable pore size. Small 2018, 14, 1702755. [Google Scholar] [CrossRef]
  29. Hof, F.; Liu, M.; Valenti, G.; Picheau, E.; Paolucci, F.; Pénicaud, A. Size control of nanographene supported iron oxide nanoparticles enhances their electrocatalytic performance for the oxygen reduction and oxygen evolution reactions. J. Phys. Chem. C 2019, 123, 20774–20780. [Google Scholar] [CrossRef]
  30. Chen, M.; Huang, Z.; Ye, X.; Zhang, L.; Feng, J.; Wang, A. Caffeine derived graphene-wrapped Fe3C nanoparticles entrapped in hierarchically porous FeNC nanosheets for boosting oxygen reduction reaction. J. Colloid Interface Sci. 2023, 637, 216–224. [Google Scholar] [CrossRef]
  31. Wang, M.; Ji, S.; Wang, H.; Wang, X.; Linkov, V.; Wang, R. Foamed carbon-supported nickel-iron oxides interspersed with bamboo-like carbon nanotubes for high-performance rechargeable Zinc-air batteries. Small 2022, 18, 2204474. [Google Scholar] [CrossRef] [PubMed]
  32. Fan, Z.; Li, J.; Yang, W.; Fu, Q.; Sun, K.; Song, Y.; Wei, Z.; Liao, Q.; Zhu, X. Green and facile synthesis of iron oxide nanoparticle-embedded N-doped biocarbon as an efficient oxygen reduction electrocatalyst for microbial fuel cells. Chem. Eng. J. 2020, 385, 123393. [Google Scholar] [CrossRef]
  33. Xiao, Z.; Wu, C.; Wang, W.; Pan, L.; Zou, J.; Wang, L.; Zhang, X.; Li, G. Tailoring the hetero-structure of iron oxides in the framework of nitrogen doped carbon for the oxygen reduction reaction and zinc–air batteries. J. Mater. Chem. A 2020, 8, 25791–25804. [Google Scholar] [CrossRef]
  34. Rasal, A.S.; Yadav, S.; Yadav, A.; Kashale, A.A.; Manjunatha, S.T.; Altaee, A.; Chang, J.Y. Carbon quantum dots for energy applications: A review. ACS Appl. Nano Mater. 2021, 4, 6515–6541. [Google Scholar] [CrossRef]
  35. García de Arquer, F.P.; Talapin, D.V.; Klimov, V.I.; Arakawa, Y.; Bayer, M.; Sargent, E.H. Semiconductor quantum dots: Technological progress and future challenges. Science 2021, 373, eaaz8541. [Google Scholar] [CrossRef]
  36. Meng, D.; Wei, L.; Shi, J.; Jiang, Q.; Tang, J. A review of enhanced electrocatalytic composites hydrogen/oxygen evolution based on quantum dot. J. Ind. Eng. Chem. 2023, 121, 27–39. [Google Scholar] [CrossRef]
  37. Yu, Y.; Ma, T.; Huang, H. Semiconducting quantum dots for energy conversion and storage. Adv. Funct. Mater. 2023, 33, 2213770. [Google Scholar] [CrossRef]
  38. Qian, J.; Guo, X.; Wang, T.; Liu, P.; Zhang, H.; Gao, D. Bifunctional porous Co-doped NiO nanoflowers electrocatalysts for rechargeable zinc-air batteries. Appl. Catal. B Environ. 2019, 250, 71–77. [Google Scholar] [CrossRef]
  39. Lian, Y.; Yang, W.; Zhang, C.; Sun, H.; Deng, Z.; Xu, W.; Song, L.; Ouyang, Z.; Wang, Z.; Guo, J.; et al. Unpaired 3d electrons on atomically dispersed cobalt centres in coordination polymers regulate both oxygen reduction reaction (ORR) activity and selectivity for use in zinc–air batteries. Angew. Chem. Int. Ed. 2020, 59, 286–294. [Google Scholar] [CrossRef]
  40. Ramakrishnan, S.; Velusamy, D.B.; Sengodan, S.; Nagaraju, G.; Kim, D.H.; Kim, A.R.; Yoo, D.J. Rational design of multifunctional electrocatalyst: An approach towards efficient overall water splitting and rechargeable flexible solid-state zinc–air battery. Appl. Catal. B Environ. 2022, 300, 120752. [Google Scholar] [CrossRef]
  41. Li, Y.-S.; Church, J.S.; Woodhead, A.L. Infrared and Raman spectroscopic studies on iron oxide magnetic nano-particles and their surface modifications. J. Magn. Magn. Mater. 2012, 324, 1543–1550. [Google Scholar] [CrossRef]
  42. El Mendili, Y.; Bardeau, J.F.; Randrianantoandro, N.; Gourbil, A.; Greneche, J.M.; Mercier, A.M.; Grasset, F. New evidences of in situ laser irradiation effects on γ-Fe2O3 nanoparticles: A Raman spectroscopic study. J. Raman Spectrosc. 2011, 42, 239–242. [Google Scholar] [CrossRef]
  43. Zheng, H.; Bao, J.; Huang, Y.; Xiang, L.; Faheem; Ren, B.; Du, J.; Nadagouda, M.N.; Dionysiou, D.D. Efficient degradation of atrazine with porous sulfurized Fe2O3 as catalyst for peroxymonosulfate activation. Appl. Catal. B Environ. 2019, 259, 118056. [Google Scholar] [CrossRef]
  44. Karuppasamy, P.; Karthika, A.; Senthilkumar, S.; Rajapandian, V. An efficient and highly sensitive amperometric quercetin sensor based on a lotus flower like SeO2-decorated rGO nanocomposite modified glassy carbon electrode. Electrocatalysis 2022, 13, 269–282. [Google Scholar] [CrossRef]
  45. Wu, Z.; Lin, X.; Zhang, J.; Chu, X.; Xu, J.; Li, J.; Liu, Y.; Yu, H.; Yan, L.; Zhang, L.; et al. Encapsulating selenium into mesoporous carbon for high-performance aqueous Cu-Se battery. Chem. Eng. J. 2023, 454, 140433. [Google Scholar] [CrossRef]
  46. Liu, K.; Niu, J.; Liu, L.; Tian, F.; Nie, H.; Liu, X.; Chen, K.; Zhao, R.; Sun, S.; Jiao, M.; et al. LUMO-mediated Se and HOMO-mediated Te nanozymes for selective redox biocatalysis. Nano Lett. 2023, 23, 5131–5140. [Google Scholar] [CrossRef]
  47. Sapna; Budhiraja, N.; Kumar, V.; Singh, S.K. Synergistic effect in structural and supercapacitor performance of well dispersed CoFe2O4/Co3O4 nano-hetrostructures. Ceram. Int. 2018, 44, 13806–13814. [Google Scholar] [CrossRef]
  48. Chen, H.; Liu, W.; Qin, Z. ZnO/ZnFe2O4 nanocomposite as a broad-spectrum photo-Fenton-like photocatalyst with near-infrared activity. Catal. Sci. Technol. 2017, 7, 2236–2244. [Google Scholar] [CrossRef]
  49. Li, C.; Lin, Y.; Li, X.; Li, Z.; Luo, P.; Jin, Y.; Li, Z. Effect of Co-doping concentration on α-Fe2O3/Graphene as anode materials for lithium ion batteries. Colloids Surf. A Physicochem. Eng. Asp. 2023, 660, 130681. [Google Scholar] [CrossRef]
  50. Liu, G.; Li, J.; Fu, J.; Jiang, G.; Lui, G.; Luo, D.; Deng, Y.P.; Zhang, J.; Cano, Z.P.; Yu, A.; et al. An oxygen-vacancy-rich semiconductor-supported bifunctional catalyst for efficient and stable zinc–air batteries. Adv. Mater. 2019, 31, 1806761. [Google Scholar] [CrossRef]
  51. Qin, J.; Liu, Z.; Wu, D.; Yang, J. Optimizing the electronic structure of cobalt via synergized oxygen vacancy and Co-N-C to boost reversible oxygen electrocatalysis for rechargeable Zn-air batteries. Appl. Catal. B Environ. 2020, 278, 119300. [Google Scholar] [CrossRef]
  52. Han, X.; Liu, C.; Tang, Y.; Meng, Q.; Zhou, W.; Chen, S.; Deng, S.; Wang, J. Unveiling the role of cobalt doping in optimizing ammonia electrosynthesis on iron–cobalt oxyhydroxide hollow nanocages. J. Mater. Chem. A 2023, 11, 14424–14431. [Google Scholar] [CrossRef]
  53. Li, M.; Tang, Y.; Shi, W.; Chen, F.; Shi, Y.; Gu, H. Design of visible-light-response core–shell Fe2O3/CuBi2O4 heterojunctions with enhanced photocatalytic activity towards the degradation of tetracycline: Z-scheme photocatalytic mechanism insight. Inorg. Chem. Front. 2018, 5, 3148–3154. [Google Scholar] [CrossRef]
  54. Zhu, H.; Dong, S.; Du, X.; Du, H.; Xia, J.; Liu, Q.; Luo, Y.; Guo, H.; Li, T. Defective CuO-rich CuFe2O4 nanofibers enable the efficient synergistic electrochemical reduction of nitrate to ammonia. Catal. Sci. Technol. 2022, 12, 4998–5002. [Google Scholar] [CrossRef]
  55. Luo, J.; Han, X.; Ge, J.; Wang, Y.; Zhao, X.; Zhang, F.; Lei, X. Cu9S5/Fe2O3 nanospheres as advanced negative electrode materials for high performance battery-like hybrid capacitors. ACS Appl. Energy Mater. 2022, 5, 7016–7025. [Google Scholar] [CrossRef]
  56. Yang, H.; Wang, B.; Li, H.; Ni, B.; Wang, K.; Zhang, Q.; Wang, X. Trimetallic sulfide mesoporous nanospheres as superior electrocatalysts for rechargeable Zn–air batteries. Adv. Energy Mater. 2018, 8, 1801839. [Google Scholar] [CrossRef]
  57. Guo, Y.; Wang, T.; Yang, Q.; Li, X.; Li, H.; Wang, Y.; Jiao, T.; Huang, Z.; Dong, B.; Zhang, W.; et al. Highly efficient electrochemical reduction of nitrogen to ammonia on surface termination modified Ti3C2Tx MXene nanosheets. ACS Nano 2020, 14, 9089–9097. [Google Scholar] [CrossRef]
  58. Zhang, G.; Wang, S.; Zeng, X.; Li, X.; Xiao, L.; Chen, K.; Lu, Q.; Xu, Q.; Weng, J.; Xu, J. Holey amorphous FeCoO-coated black phosphorus for robust polysulfide adsorption and catalytic conversion in lithium–sulfur batteries. J. Mater. Chem. A 2022, 10, 11676–11683. [Google Scholar] [CrossRef]
  59. Lee, K.; Zhang, L.; Lui, H.; Hui, R.; Shi, Z.; Zhang, J. Oxygen reduction reaction (ORR) catalyzed by carbon-supported cobalt polypyrrole (Co-PPy/C) electrocatalysts. Electrochim. Acta 2009, 54, 4704–4711. [Google Scholar] [CrossRef]
  60. Xu, C.; Guo, C.; Liu, J.; Hu, B.; Dai, J.; Wang, M.; Jin, R.; Luo, Z.; Li, H.; Chen, C. Accelerating the oxygen adsorption kinetics to regulate the oxygen reduction catalysis via Fe3C nanoparticles coupled with single Fe-N4 sites. Energy Storage Mater. 2022, 51, 149–158. [Google Scholar] [CrossRef]
  61. Gao, S.; Fan, B.; Feng, R.; Ye, C.; Wei, X.; Liu, J.; Bu, X. N-doped-carbon-coated Fe3O4 from metal-organic framework as efficient electrocatalyst for ORR. Nano Energy 2017, 40, 462–470. [Google Scholar] [CrossRef]
  62. Mao, S.; Wen, Z.; Huang, T.; Hou, Y.; Chen, J. High-performance bi-functional electrocatalysts of 3D crumpled graphene–cobalt oxide nanohybrids for oxygen reduction and evolution reactions. Energy Environ. Sci. 2014, 7, 609–616. [Google Scholar] [CrossRef]
  63. Du, S.; Ren, Z.; Qu, Y.; Wu, J.; Xi, W.; Zhu, J.; Fu, H. Co3O4 nanosheets as a high-performance catalyst for oxygen evolution proceeding via a double two-electron process. Chem. Commun. 2016, 52, 6705–6708. [Google Scholar] [CrossRef]
  64. Tian, G.; Zhao, M.; Yu, D.; Kong, X.; Huang, J.; Zhang, Q.; Wei, F. Nitrogen-doped graphene/carbon nanotube hybrids: In situ formation on bifunctional catalysts and their superior electrocatalytic activity for oxygen evolution/reduction reaction. Small 2014, 10, 2251–2259. [Google Scholar] [CrossRef] [PubMed]
  65. Rasaki, S.A.; Chen, Z.; Shen, H.; Guo, H.; Thomas, T.; Yang, M. Cobalt nanoparticles modified single-walled titanium carbonitride nanotube derived from solid-solid separation for oxygen reduction reaction in alkaline solution. Electrocatalysis 2020, 11, 579–592. [Google Scholar] [CrossRef]
  66. Tang, K.; Hu, H.; Xiong, Y.; Chen, L.; Zhang, J.; Yuan, C.; Wu, M. Hydrophobization engineering of the air–cathode catalyst for improved oxygen diffusion towards efficient zinc–air batteries. Angew. Chem. Int. Ed. 2022, 61, e202202671. [Google Scholar] [CrossRef]
  67. Faisal, S.N.; Haque, E.; Noorbehesht, N.; Liu, H.; Islam, M.M.; Shabnam, L.; Roy, A.K.; Pourazadi, E.; Islam, M.S.; Harris, A.T.; et al. A quadrafunctional electrocatalyst of nickel/nickel oxide embedded N-graphene for oxygen reduction, oxygen evolution, hydrogen evolution and hydrogen peroxide oxidation reactions. Sustain. Energy Fuels 2018, 2, 2081–2089. [Google Scholar] [CrossRef]
  68. Shudo, Y.; Fukuda, M.; Islam, M.S.; Kuroiwa, K.; Sekine, Y.; Karim, M.R.; Hayami, S. 3D porous Ni/NiOx as a bifunctional oxygen electrocatalyst derived from freeze-dried Ni(OH)2. Nanoscale 2021, 13, 5530–5535. [Google Scholar] [CrossRef]
  69. Xiong, P.; Sun, B.; Sakai, N.; Ma, R.; Sasaki, T.; Wang, S.; Zhang, J.; Wang, G. 2D superlattices for efficient energy storage and conversion. Adv. Mater. 2020, 32, 1902654. [Google Scholar] [CrossRef]
  70. Cheng, Y.; Pang, K.; Wu, X.; Zhang, Z.; Xu, X.; Ren, J.; Huang, W.; Song, R. In situ hydrothermal synthesis MoS2/guar gum carbon nanoflowers as advanced electrocatalysts for electrocatalytic hydrogen evolution. ACS Sustain. Chem. Eng. 2018, 6, 8688–8696. [Google Scholar] [CrossRef]
  71. Kong, H.; Lv, C.; Yan, C.; Chen, G. Engineering mesoporous single crystals Co-doped Fe2O3 for high-performance lithium ion batteries. Inorg. Chem. 2017, 56, 7642–7649. [Google Scholar] [CrossRef] [PubMed]
  72. Hou, Y.; Zuo, F.; Dagg, A.; Feng, P. A three-dimensional branched cobalt-doped α-Fe2O3 Nanorod/MgFe2O4 heterojunction array as a flexible photoanode for efficient photoelectrochemical water oxidation. Angew. Chem. Int. Ed. 2013, 52, 1248–1252. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Schematic of the fabrication of Se-FeOx-Co. (b) XRD patterns of Se-FeOx, Se-FeOx-Co-1, and Se-FeOx-Co-2 samples. (c) SEM, (d) TEM, (e) HRTEM, and (f) EDS mapping images of Se-FeOx-Co-2; (g) relative atomic content.
Figure 1. (a) Schematic of the fabrication of Se-FeOx-Co. (b) XRD patterns of Se-FeOx, Se-FeOx-Co-1, and Se-FeOx-Co-2 samples. (c) SEM, (d) TEM, (e) HRTEM, and (f) EDS mapping images of Se-FeOx-Co-2; (g) relative atomic content.
Batteries 09 00561 g001
Figure 2. (a) Raman spectra of Se-FeOx and Se-FeOx-Co-2 samples. (b) XPS full survey spectra of those powders. High-resolution XPS spectra of (c) Se 3d, (d) O 1s, (e) Fe 2p, and (f) Co 2p for Se-FeOx and Se-FeOx-Co-2 samples.
Figure 2. (a) Raman spectra of Se-FeOx and Se-FeOx-Co-2 samples. (b) XPS full survey spectra of those powders. High-resolution XPS spectra of (c) Se 3d, (d) O 1s, (e) Fe 2p, and (f) Co 2p for Se-FeOx and Se-FeOx-Co-2 samples.
Batteries 09 00561 g002
Figure 3. (a) CV curves of Se-FeOx-Co and Se-FeOx-Co-2 catalyst at 100 mV s −1. (b) LSV curves of a series of Se-FeOx sample catalysts at 1600 rpm for ORR test. (c) LSV curves of catalysts for OER test. (d) Relative bar in the value of E1/2 and Ej=10 in a series of samples. (e) ORR stability of catalysts tested in half-wave potential, and (f) OER stability of catalysts at a current density of 10 mA cm−2.
Figure 3. (a) CV curves of Se-FeOx-Co and Se-FeOx-Co-2 catalyst at 100 mV s −1. (b) LSV curves of a series of Se-FeOx sample catalysts at 1600 rpm for ORR test. (c) LSV curves of catalysts for OER test. (d) Relative bar in the value of E1/2 and Ej=10 in a series of samples. (e) ORR stability of catalysts tested in half-wave potential, and (f) OER stability of catalysts at a current density of 10 mA cm−2.
Batteries 09 00561 g003
Figure 4. (a) Schematic of the operation of zinc–air battery (ZAB) based on the Se-FeOx-Co bifunctional catalyst. (b) Comparison of open-circuit voltage and (c) optical photograph of ZAB based on Se-FeOx-Co-2. (d) Discharging curves and (e) rechargeable ZAB cycling profiles at current density of 10 mA cm−2.
Figure 4. (a) Schematic of the operation of zinc–air battery (ZAB) based on the Se-FeOx-Co bifunctional catalyst. (b) Comparison of open-circuit voltage and (c) optical photograph of ZAB based on Se-FeOx-Co-2. (d) Discharging curves and (e) rechargeable ZAB cycling profiles at current density of 10 mA cm−2.
Batteries 09 00561 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, D.; Wang, Y.; Han, X.; Hu, W. Developing a Se Quantum Dots@ CoFeOx Composite Nanomaterial as a Highly Active and Stable Cathode Material for Rechargeable Zinc–Air Batteries. Batteries 2023, 9, 561. https://doi.org/10.3390/batteries9110561

AMA Style

Zhang D, Wang Y, Han X, Hu W. Developing a Se Quantum Dots@ CoFeOx Composite Nanomaterial as a Highly Active and Stable Cathode Material for Rechargeable Zinc–Air Batteries. Batteries. 2023; 9(11):561. https://doi.org/10.3390/batteries9110561

Chicago/Turabian Style

Zhang, Donghao, Yang Wang, Xiaopeng Han, and Wenbin Hu. 2023. "Developing a Se Quantum Dots@ CoFeOx Composite Nanomaterial as a Highly Active and Stable Cathode Material for Rechargeable Zinc–Air Batteries" Batteries 9, no. 11: 561. https://doi.org/10.3390/batteries9110561

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