Next Article in Journal
Redox Conversions of 5-Methyl-6-nitro-7-oxo-4,7-dihydro-1,2,4triazolo[1,5-a]pyrimidinide L-Arginine Monohydrate as a Promising Antiviral Drug
Next Article in Special Issue
Novel Silver-Plated Nickel-Coated Graphite Powder with Excellent Heat and Humidity Resistance: Facile Preparation and Performance Investigation
Previous Article in Journal
An Efficient Synthesis of 2-CF3-3-Benzylindoles
Previous Article in Special Issue
Hybrid Structures of Sisal Fiber Derived Interconnected Carbon Nanosheets/MoS2/Polyaniline as Advanced Electrode Materials in Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Catalysis of SnO2/Reduced Graphene Oxide for VO2+/VO2+ and V2+/V3+ Redox Reactions

School of Chemical Engineering, North China University of Science and Technology, Tangshan 063009, China
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(16), 5085; https://doi.org/10.3390/molecules26165085
Submission received: 29 June 2021 / Revised: 18 August 2021 / Accepted: 18 August 2021 / Published: 22 August 2021
(This article belongs to the Special Issue Design and Application Based on Versatile Nano-Composites)

Abstract

:
In spite of their low cost, high activity, and diversity, metal oxide catalysts have not been widely applied in vanadium redox reactions due to their poor conductivity and low surface area. Herein, SnO2/reduced graphene oxide (SnO2/rGO) composite was prepared by a sol–gel method followed by high-temperature carbonization. SnO2/rGO shows better electrochemical catalysis for both redox reactions of VO2+/VO2+ and V2+/V3+ couples as compared to SnO2 and graphene oxide. This is attributed to the fact that reduced graphene oxide is employed as carbon support featuring excellent conductivity and a large surface area, which offers fast electron transfer and a large reaction place towards vanadium redox reaction. Moreover, SnO2 has excellent electrochemical activity and wettability, which also boost the electrochemical kinetics of redox reaction. In brief, the electrochemical properties for vanadium redox reactions are boosted in terms of diffusion, charge transfer, and electron transport processes systematically. Next, SnO2/rGO can increase the energy storage performance of cells, including higher discharge electrolyte utilization and lower electrochemical polarization. At 150 mA cm−2, the energy efficiency of a modified cell is 69.8%, which is increased by 5.7% compared with a pristine one. This work provides a promising method to develop composite catalysts of carbon materials and metal oxide for vanadium redox reactions.

1. Introduction

With the development of industrial processes, energy consumption is increasing day by day. However, energy consumption brings environmental pollution [1,2,3,4]. In order to solve this problem and achieve “carbon neutrality”, humankind began to explore renewable energy [5,6,7,8,9]. However, this type of energy is not always available at all times, and in most cases, due to the oscillation of energy availability in its cycle and peak demand, supply and demand are not synchronized [10,11,12,13,14,15]. Energy storage equipment is needed to store and transform the electric energy generated by renewable energy sources to realize its large-scale application [16,17,18,19,20]. Among many energy storage devices, a vanadium redox flow battery (VRFB) has attracted much attention because of its unique performance [21,22,23]. The main components of a VRFB are electrodes, electrolytes, and proton exchange membranes [24,25,26]. The performance of an electrode has a direct impact on the overall performance of the battery. Carbon-based materials have become the most widely used electrode material in a VRFB due to their large specific surface area and good corrosion resistance. However, the electrochemical performance of carbon-based materials is poor, which is not conducive to their large-scale utilization. In order to improve the performance of carbon-based materials and expand the use of materials, researchers have modified them. Common electrode modification methods mainly include intrinsic treatment and the introduction of catalysts.
There are several categories of catalysts, including carbon materials, metal, metal oxide, metal nitride, metal boride, and metal carbide. Firstly, carbon materials (CNTs, graphene, carbon sheet, carbon nanofiber, etc.) have advantages including a low cost, obvious structure, and a larger surface area, and have been widely used as a catalyst in VRFB [27,28,29]. For example, Yan et al. [30] developed multi-walled carbon nanotubes as a catalyst for VO2+/VO2+ redox reaction. In addition, some biomass-based carbon materials derived from glucose, fish scales, and kiwi fruit have abundant groups, heteroatoms, and a large surface area, which also show certain electrochemical catalysts for a vanadium redox reaction [31,32]. Metal has a unique atomic structure and high conductivity, and exhibits intrinsic catalysis for a vanadium redox reaction. Metal catalysts mainly include Au, Pt, Ir, Cu, etc. Zeng et al. [33] used a copper nanoparticle to decorate graphite felt with in situ electrodeposition. The copper nanoparticle was an excellent catalyst for a V3+/V2+ redox reaction, and improved the energy efficiency of the VRFB. However, most of the metal catalysts have disadvantages, including a high price and easy access to hydrogen evolution. Moreover, metal oxides have attracted more attention due to the advantages of a low cost, high stability, and diversity. In recent decades, many metal oxides including ZrO2, TiO2, CeO2, WO3, SnO2, etc., have been developed and applied in VRFB [34,35,36,37,38,39]. For instance, Chen et al. [35] modified graphite felt with a ZrO2 nanoparticle by thermal decomposition. ZrO2 showed good wettability and acted as active sites for both redox reactions, resulting in faster mass and charge transfer processes. However, due to the low conductivity, the intrinsic catalysis of metal oxide is not efficiently used. Other metal compounds have been developed as catalysts for VRFB. For example, the Ti3C2Tx MXene spheres have metallic behavior and a high electrical conductivity, which shows high catalysis for a V3+/V2+ redox reaction [40]. Moreover, metal borides such as TiB2 and ZrB2 exhibit high conductivity for a V3+/V2+ redox reaction [41,42]. It is noteworthy that carbon materials can be used as support to not only improve conductivity of metal oxide but also to increase the surface area of metal oxide. Therefore, it is an efficient way to obtain high-performance catalysts by compositing carbon materials and metal oxide. For example, Li et al. [43] synthesized a polydopamine-Mn3O4 composite as a catalyst for a VO2+/VO2+ redox reaction. The synergistic effect of polydopamine and Mn3O4 improved the electrochemical performance of the VRFB.
SnO2 is an amphoteric oxide, which can exist stably in the strong acidic environment of a VRFB electrolyte [39,44]. Due to the wide band gap and unique performance of SnO2, it is widely used in optics, sensors, and other fields [45]. A large number of experiments have proved that SnO2 has a good catalytic effect on V3+/V2+ and VO2+/VO2+ redox reactions. Ha et al. [39] prepared in situ SnO2 nanoparticles and modified carbon felt electrodes by the hydrothermal method. SnO2 reduced the activation potential of the reaction and increased the reaction rate. The SnO2 modified electrode significantly improves the stability of the battery, which is very important for the VRFB. Graphene has good electrical conductivity, high stability, and a large surface area, making it an excellent electrocatalyst carbon support. Walsh et al. [46] studied the synergy between graphene and Mn3O4. Due to the interaction between graphene support and the Mn3O4 catalyst, the catalytic performance of the composite material is higher than that of any component alone.
In this article, we used SnO2 as the electrocatalyst and reduced graphene oxide as the carbon support, respectively. The large surface area of reduced graphene oxide not only makes the SnO2 nanoparticle more dispersive, but also provides a larger reaction place for redox reaction. In addition, the high conductivity of reduced graphene oxide boosts the transport of electrons for redox reaction. Therefore, it is inferred that the synergetic effect between reduced graphene oxide and SnO2 is more conducive to a vanadium redox reaction. The catalytic performance of the SnO2/reduced graphene oxide composite catalyst for V3+/V2+ and VO2+/VO2+ reactions is much higher than using SnO2 and reduced graphene oxide alone as a catalyst. Due to the synergy between SnO2 and reduced graphene oxide, the composite catalyst has good electrochemical activity and stability.

2. Experimental

2.1. Preparation of Materials

As shown in Figure 1, the sample was prepared by the sol–gel method. First, 1.4977 g of SnCl2·2H2O was weighed and placed in a beaker. Then, 13.27 mL of an anhydrous ethanol/water mixed solution (volume ratio: 15:1) was added and stirred in a water bath at 65 °C for 40 min to gradually form a gel. After the gel was formed, it was aged in air for 24 h to obtain a precursor of tin sol. Then, 50 mg of graphene oxide (GO), the tin sol precursor, and 10 mL of anhydrous alcohol were mixed and ultrasonically dispersed for 1 h. Next, the sample was placed in a drying oven to dry adequately at 80 °C. The Ar was put into the tube furnace before calcining for 2 h to exhaust the air. Then, the sample was calcined at 500 °C for 2 h. The obtained sample was named as SnO2/rGO.

2.2. Characterization of Materials

The crystal phases of the samples were studied using the D8 Advance A25 Instrument (Bruker, Berlin, Germany). The microscopic morphology of the sample was observed by a scanning electron microscope (SEM, JSM-IT100). The internal morphology and crystal lattice of the sample were observed by a transmission electron microscope (TEM, JEOL JEM-2100F). The 3H-2000PM1 specific surface area and porous analyzer was used to analyze the materials surface area. The X-ray photoelectron spectroscopy (XPS) of the samples was studied using the Thermo Scientific Escalab 250Xi instrument (Thermo Fisher Scientific, Waltham, MA, USA) to determine the elemental composition of the samples.

2.3. Electrochemical Measurements

First, 10 mg of the catalyst was thoroughly mixed with 5 mL of N,N-dimethylformamide (DMF) to obtain a dispersion. The dispersion solution was then put into an ultrasonic cleaning machine and dispersed continuously for 3 h, until the catalyst was evenly dispersed in DMF. The dispersion was then dropped on the glassy carbon electrode (GCE) and left to stand at room temperature for 4 h.
Cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) measurements were performed by an electrochemical workstation (CHI660E, Shanghai Chenghua, Shanghai, China). In this three-electrode system, platinum electrode (1 × 1 cm2) was the counter electrode, glassy carbon electrode was the working electrode, and the calomel electrode was the reference electrode. Positive and negative CV and EIS tests were accomplished in 1.6 M VO2+ + 3.0 M H2SO4 and 1.6 M V3+ + 3.0 M H2SO4, respectively. The voltage scanning range of the CV test for a V2+/V3+ reaction was from −0.75 to −0.1 V, and the VO2+/VO2+ reaction was from 0.2 to 1.5 V. The frequency range for EIS was 1 to 1 × 106 Hz. The EIS of the VO2+/VO2+ and V2+/V3+ reactions were tested at the polarization of 0.85 V and −0.45 V, respectively.

2.4. Charge–Discharge Tests

The catalyst modification cell and the pristine cell were assembled to compare the VRFB’s performance. CT2001A was used for charge–discharge tests. Graphite felt (3 × 3 cm2) was activated in a muffle furnace at 400 °C for 10 h. The activated graphite felt was then dipped into the dispersion solution. The dispersion solution was composed of 3 mg of the catalyst and 10 mL of DMF. After the graphite felt completely absorbed the dispersion solution, it was taken out and dried at 80 °C. The above steps were repeated until the dispersion solution was completely absorbed in order to obtain the SnO2/rGO modify graphite felt electrode.
The SnO2/rGO modify graphite felt was, respectively, used as the positive and negative electrodes of the modified cell. The active graphite felt was used as the pristine cell. The graphite felt was placed in the electrolyte of 0.8 M V3+ + 0.8 M VO2+ + 3.0 M H2SO4 to fully absorb the electrolyte. The rate performance was tested at 50, 75, 100, 125, and 150 mA cm−2 current densities.

3. Results and Discussion

Figure 2a–c show the SEM images of GO, SnO2, and SnO2/rGO, respectively. As shown in Figure 2a, GO is curly. Figure 2b shows the SnO2 particles have obvious agglomeration. It can be seen from Figure 2c that SnO2 is evenly distributed on the reduced graphene oxide sheet. As shown in Figure 2d–f, the morphology of the SnO2/rGO composite was further characterized by TEM. The low magnification TEM figure shows that a large amount of SnO2 is loaded on the reduced graphene oxide sheet. There is a large amount of aggregation, which can be seen more clearly from the medium magnification TEM figure. From the high magnification TEM figure, it can be further concluded that the obvious lattice fringes are 0.34 nm, which is the (110) surface of SnO2. Figure 2g, h display the EDX spectra of GO and SnO2/rGO, respectively. Absolutely, SnO2/rGO have C and O. Compared with Figure 2g, there is Sn in Figure 2h. It can be concluded that there is Sn in SnO2/rGO.
Figure 2i shows the X-ray diffraction pattern of the samples. The crystal structure of the samples was studied by XRD. There is a characteristic peak at 12.7° for graphene oxide, which proves that the purity of graphene oxide is very high [47]. The observed pattern of SnO2 is consistent with the standard value (JCPDS: 00-021-1250), and there is no characteristic peak of impurity. The reduced graphene oxide crystallization peaks in the XRD pattern of SnO2/rGO are lack. The order of the reduced graphene oxide plane is disrupted by the SnO2 particles, or reduced graphene oxide sheets are inserted into the SnO2 lattice [48]. SnO2/rGO has the advantages of two materials. It has both the conductivity of rGO and the catalytic performance of SnO2, which can improve the electrochemical catalytic performance.
Figure 3a, b show the N2 adsorption and desorption isotherms of GO and SnO2/rGO. The isotherms of GO and SnO2/rGO are type IV, indicating that the two are mesopores materials. Additionally, the two isotherms exhibit type H3 loops, which identified with plate-like porous aggregates, as these originate from slit-shaped interlayer pores. Based on the multilayer adsorption theory, the specific surface areas of GO and SnO2/rGO were computed to be 44.3 m2 g−1 and 208.9 m2 g−1, respectively. Pore distribution has been assessed by means of the Barrett–Joyner–Halenda (BJH) method, and the results have been shown in Figure 3c,d. It can be seen that GO and SnO2/rGO contain a large number of mesoporous of ~4 nm. The pore width is not affected by heat treatment. The increase of mesoporous can explain why SnO2/rGO has a larger surface area, which can provide a larger reaction place for electrode reaction.
The element composition of SnO2/rGO was further analyzed by XPS. The XPS spectrum survey of SnO2/rGO is shown in Figure 4a. It shows that the peaks in the energy spectrum of SnO2/rGO correspond to C 1s, O 1s, and Sn 3d, indicating that SnO2/rGO consists of C, O, and Sn elements. Figure 4b–d show the high-resolution XPS spectra of C 1s, O 1s, and Sn 3d, respectively. There are three peaks centered at 284.4, 285.6, and 288.6 eV, which correspond to C-C, C-O, and O-C=O bonds, respectively. These C 1s peaks related to oxygen-containing groups are very weak. The peaks at 531.1 and 532.3 eV shown in Figure 4c correspond to Sn-O and C-O bonds. These binding energies are characteristic of oxygen ionization on the SnO2 surface. Figure 4d shows two different peaks with binding energies of 487.0 and 495.4 eV, corresponding to Sn 3d5/2 and Sn 3d3/2, respectively. This indicates the presence of Sn4+ and the formation of SnO2 [49]. It can be concluded that SnO2/rGO is composed of reduced graphene oxide and SnO2.
CV measurements were used to evaluate the effect of the samples on the electrocatalysis of a VO2+/VO2+ redox reaction. Figure 5a shows the CV curve of bare GCE. It indicates bare GCE without a catalysis for a VO2+/VO2+ redox reaction. It is beneficial to compare the electrochemical activity of the catalysts. The poor peak shape of SnO2 is due to the poor electrical conductivity of SnO2, which reduces the electrochemical reversibility of the electrode. The peak shape of GO is better than that of SnO2, indicating the advantages of GO. The electrochemical performance of SnO2/rGO composite is better than those of both GO and SnO2. The composite has the synergistic effect of the good electrical conductivity of rGO and the good catalytic property of SnO2. Moreover, it can be seen that the peak current of the SnO2/rGO electrode (oxidation: 2.59 mA, reduction: 1.19 mA) is the largest, which again shows that SnO2/rGO has a good electrochemical activity for a VO2+/VO2+ redox reaction.
Figure 5b,c show the CVs of GO and SnO2/rGO. As shown in the figure, with the change of the scan rate, the redox peak current also moves up and down. At a larger scan rate, the difference between the oxidation peak potential and the reduction peak potential gradually increases. Figure 5d depicts the relationship between the redox peak current and the square root of the scan rate. It is clear that there is a proportional relationship between the peak current and the square root of the scan rate, indicating that the reaction is controlled by the diffusion process [50]. The slope of SnO2/rGO is higher than that of GO, which may be due to the fact that SnO2/rGO has a better catalytic effect, so that the concentration difference on the electrode surface is larger.
To further explore the electrocatalytic property of the samples for redox reaction, we performed an EIS test. Figure 5e displays the Nyquist plots of the positive electrodes of SnO2, GO, and SnO2/rGO. In the Nyquist diagram, a semicircle appears at the high frequency region and a line appears at the low frequency region. This means that the VO2+/VO2+ electrode reaction is jointly governed by charge and mass transfer processes [51,52]. Figure 5f shows a simplified equivalent circuit diagram conforming to a Nyquist diagram.
The corresponding fitting electrochemical parameters of three samples are shown in Table 1. The ohmic resistance (Rs) of SnO2/rGO and GO is smaller than that of SnO2. In addition, the charge transfer resistance (Rct) indicates that SnO2/rGO < GO < SnO2, which means that SnO2/rGO has the lowest Rct. The order of electrical double layer capacitance (Qm) is SnO2 < SnO2/rGO < GO. GO has the highest Qm because of a large number of hydrophilic groups on its surface. However, after calcination at high temperature, the hydrophilic groups of the composite are reduced and the Qm is reduced. SnO2 has the lowest Qm because it has no hydrophilic groups on its surface. The order of diffusion capacitance (Qt) is SnO2 < GO < SnO2/rGO. This may be due to the comprehensive effect of electrocatalysis and surface changes. The increase of Qm can promote the charge migration of a VO2+/VO2+ reaction, and the increase of Qt can accelerate the diffusion of VO2+/VO2+ ions on the electrode.
To explore the influence of the samples on the electrocatalysis of the V2+/V3+ reaction, we carried out a CV test. Figure 6a displays the CV curves of GO and SnO2/rGO in the 1.6 M V3+ + 3.0 M H2SO4, which indicate that SnO2/rGO has a higher redox peak current than GO. The oxidation peak current of SnO2/rGO can reach 1.56 mA, which is more than twice as high as that of GO (0.73 mA). The reduction peak current of SnO2/rGO (2.82 mA) is 1.55 times higher than that of GO (1.82 mA). The results indicate that SnO2/rGO also has superior electrocatalysis towards a V2+/V3+ redox reaction. This is consistent with the results of the positive electrode. The EIS measurement for the negative couple was carried out to study the electrocatalytic activity. Figure 6b is the Nyquist diagram of the V2+/V3+ reaction process of GO and SnO2/rGO. It can be seen that the Nyquist diagram is similar to a semicircle at the high frequency region and a line in the low frequency region. This means the V2+/V3+ reaction process is under co-control of the charge and mass transfer processes. Compared with GO, the smaller semicircle and higher slope of SnO2/rGO mean lower charge transfer resistance and faster diffusion performance, which comes from the special structure of the SnO2/rGO composite. It is the same as the positive EIS measurement analysis results.
Figure 7 shows the SEM images of pristine graphite felt and graphite felt modified with SnO2/rGO. It is seen from Figure 7a,b that pristine graphite felt presents a clean and smooth surface. The diameter of the carbon fiber is 10–15 μm. It is observed in Figure 7c,d that graphite felt is uniformly coated with the SnO2/rGO catalyst, which offers a larger reaction place and active sites for electrode reaction.
Figure 8a displays the discharge capacity of the pristine and SnO2/rGO modified cells at a different current density. It is observed that as the current density increases, the difference in discharge capacity between the two cells progressively increases. This result reflects that SnO2/rGO has a good effect on increasing discharge capacity at a high current density. As a result, SnO2/rGO can increase the reaction place and active site for electrode reaction, which increases the electrolyte utilization. In Figure 8b, the SnO2/rGO cell exhibits a slightly lower than current efficiency (CE) as compared to the pristine cell. This is owing to a longer charge–discharge time of the SnO2/rGO cell, and a more severe active substance penetration between two electrolytes in the SnO2/rGO cell. In Figure 8c, there is an inversely proportional relation between voltage efficiency (VE) and current density. VE can reflect the cell polarization, including ohmic polarization, electrochemical polarization, and concentration polarization. SnO2/rGO can reduce the electrochemical polarization and concentration polarization of the cell. The ohmic resistance of the two cells is comparable, as those two cells are comprised of the same components. The cell with SnO2/rGO has a higher VE than the pristine one at 50–150 mA cm−2. At 150 mA cm−2, the modified cell shows an increase of 6.2% in VE compared to the pristine one (65.7%). This suggests that the SnO2/rGO cell exhibits lower electrochemical polarization than the pristine one. As shown in Figure 8d, the energy efficiency (EE) is jointly affected by CE and VE. As the current density rises, the EE of both cells reduces. The SnO2/rGO cell presents a greater EE than the pristine one. This indicates that the cell using SnO2/rGO has a higher energy storage capacity. At 150 mA cm−2, the EE of the modified cell is 69.8%, which is higher than that of the pristine cell (64.1%). These results indicate that the SnO2/rGO modified cells can reduce the electrochemical polarization and have better energy storage performance. Figure 8e, f display the charge–discharge curves of two cells. It is seen from the comparison that the curves of the SnO2/rGO modified cell exhibit a lower charge voltage and a higher discharge voltage than those of the pristine one. The reason is that the SnO2/rGO composite material significantly boosts the redox reaction and reduces the polarization of the cell, suggesting that the SnO2/rGO composite raises the energy density of the cell.

4. Conclusions

In this work, an easy sol–gel approach was employed to a composite reduced graphene oxide, and SnO2 was used as a bifunctional synergistic catalyst. The reduced graphene oxide as carbon support provides an adequate reaction place and fast electron transport due to its large surface area and fine conductivity. Moreover, SnO2 has the intrinsic catalysis for a vanadium redox reaction. The strong interaction between the reduced graphene oxide and SnO2 results in synergistical catalysis for redox reactions of VO2+/VO2+ and V2+/V3+, which systematically boosts the electrode reaction from diffusion, charge transfer, and electron transmission processes. Furthermore, SnO2/rGO increases the electrolyte utilization of the cell and decreases the electrochemical polarization, resulting in higher discharge capacity and energy efficiency of the modified cell. At 150 mA cm−2, the energy efficiency of the modified cell is increased to 69.8%, as compared to 64.1% in the pristine one.

Author Contributions

Conceptualization, Y.L. (Yongguang Liu), Y.J. and Y.L. (Yanrong Lv); Data curation, Z.H.; Formal analysis, Y.J.; Methodology, Y.L. (Yongguang Liu), Y.J. and Y.L. (Yanrong Lv); Resources, Z.H.; Software, Y.L. (Yongguang Liu), L.D. and L.W.; Writing—original draft, Y.L. (Yongguang Liu); Writing—review & editing, Y.J., Y.L. (Yanrong Lv) and Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by National Natural Science Foundation of China (No. 51872090, 51772097); Hebei Natural Science Fund for Distinguished Young Scholar (No. E2019209433); Youth Talent Program of Hebei Provincial Education Department (No. BJ2018020); and Natural Science Foundation of Hebei Province (No. E2020209151).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability Statement

Samples of the compounds are NOT available from the authors.

References

  1. Saloux, E.; Candanedo, J.A. Model-based predictive control to minimize primary energy use in a solar district heating system with seasonal thermal energy storage. Appl. Energ. 2021, 291, 116840. [Google Scholar] [CrossRef]
  2. Wu, Y.H.; Tian, Z.N.; Yuan, S.F.; Qi, Z.Y.; Feng, Y.R.; Wang, Y.F.; Huang, R.; Zhao, Y.L.; Sun, J.H.; Zhao, W.; et al. Solar-driven self-powered alkaline seawater electrolysis via multifunctional earth-abundant heterostructures. Chem. Eng. J. 2021, 411, 127329. [Google Scholar] [CrossRef]
  3. Zhang, S.Y.; Huang, Z.Q.; Wang, H.L.; Liu, R.K.; Cheng, C.; Guo, Z.Q.; Yu, X.Y.; He, G.C.; Fu, W. Separation of wolframite ore by froth flotation using a novel “crab” structure sebacoyl hydroxamic acid collector without Pb(NO3)2 activation. Power Technol. 2021, 389, 96–103. [Google Scholar] [CrossRef]
  4. Huang, Z.Q.; Zhang, S.Y.; Wang, H.L.; Liu, R.K.; Cheng, C.; Liu, Z.W.; Guo, Z.Q.; Yu, X.Y.; He, G.C.; Ai, G.H.; et al. “Umbrella” structure trisiloxane surfactant: Synthesis and application for reverse flotation of phosphorite ore in phosphate fertilizer production. J. Agric. Food Chem. 2020, 68, 11114–11120. [Google Scholar] [CrossRef] [PubMed]
  5. Kasaeian, A.; Bellos, E.; Shamaeizadeh, A.; Tzivanidis, C. Solar-driven polygeneration systems: Recent progress and outlook. Appl. Energy 2020, 264, 114764–114774. [Google Scholar] [CrossRef]
  6. Khan, A.; Senthil, R.A.; Pan, J.; Osman, S.; Sun, Y.; Shu, X. A new biomass derived rod-like porous carbon from tea-waste as inexpensive and sustainable energy material for advanced supercapacitor application. Electrochim. Acta 2020, 335, 135588–135597. [Google Scholar] [CrossRef]
  7. Kim, H.Y.; Joo, S.H. Recent advances in nanostructured intermetallic electrocatalysts for renewable energy conversion reactions. J. Mater. Chem. A 2020, 8, 8195–8217. [Google Scholar] [CrossRef]
  8. Li, B.; Xue, J.; Han, C.; Liu, N.; Ma, K.X.; Zhang, R.C.; Wu, X.W.; Dai, L.; Wang, L.; He, Z.X. A hafnium oxide-coated dendrite-free zinc anode for rechargeable aqueous zinc-ion batteries. J. Colloid Interface Sci. 2021, 599, 467–475. [Google Scholar] [CrossRef]
  9. Kou, Z.Y.; Lu, Y.; Miao, C.; Li, J.Q.; Liu, C.J.; Xiao, W. High-performance sandwiched hybrid solid electrolytes by coating polymer layers for all-solid-state lithium-ion batteries. Rare Met. 2021, 40, 3175–3184. [Google Scholar] [CrossRef]
  10. Duburg, J.C.; Azizi, K.; Primdahl, S.; Hjuler, H.A.; Zanzola, E.; Schmidt, T.J.; Gubler, L. Composite polybenzimidazole membrane with high capacity retention for vanadium redox flow batteries. Molecules 2021, 26, 1679. [Google Scholar] [CrossRef] [PubMed]
  11. Delgado, S.; Arevalo, M.D.; Pastor, E.; Garcia, G. Electrochemical reduction of carbon dioxide on graphene-based catalysts. Molecules 2021, 26, 572. [Google Scholar] [CrossRef]
  12. Kim, S.; Kim, G.; Manthiram, A. A bifunctional hybrid electrocatalyst for oxygen reduction and oxygen evolution reactions: Nano-Co3O4-deposited La0.5Sr0.5MnO3 via infiltration. Molecules 2021, 26, 277. [Google Scholar] [CrossRef] [PubMed]
  13. Tang, F.; Gao, J.Y.; Ruan, Q.Y.; Wu, X.W.; Wu, X.S.; Zhang, T.; Liu, Z.X.; Xiang, Y.H.; He, Z.Q.; Wu, X.M. Graphene-wrapped MnO/C composites by MOFs-derived as cathode material for aqueous zinc ion batteries. Electrochim. Acta 2020, 353, 136570. [Google Scholar] [CrossRef]
  14. Gao, J.W.; Xie, X.S.; Liang, S.Q.; Lu, B.A.; Zhou, J. Inorganic colloidal electrolyte for highly robust zinc-ion batteries. Nano-Micro Lett. 2021, 13, 69. [Google Scholar] [CrossRef] [PubMed]
  15. Shan, L.T.; Wang, Y.R.; Liang, S.Q.; Tang, B.Y.; Yang, Y.Q.; Wang, Z.Q.; Lu, B.A.; Zhou, J. Interfacial adsorption-insertion mechanism induced by phase boundary toward better aqueous Zn-ion battery. Infomat 2021. [Google Scholar] [CrossRef]
  16. Chang, T.C.; Liu, Y.H.; Chen, M.L.; Tseng, C.C.; Lin, Y.S.; Huang, S.L. Cerium/ascorbic acid/iodine active species for redox flow energy storage battery. Molecules 2021, 26, 3443. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, L.M.; Zhang, X.; Wang, J.; Seveno, D.; Fransaer, J.; Locquet, J.P.; Seo, J.W. Carbon nanotube fibers decorated with MnO2 for wire-shaped supercapacitor. Molecules 2021, 26, 3479. [Google Scholar] [CrossRef]
  18. Moghaddam, M.; Sepp, S.; Wiberg, C.; Bertei, A.; Rucci, A.; Peljo, P. Thermodynamics, charge transfer and practical considerations of solid boosters in redox flow batteries. Molecules 2021, 26, 2111. [Google Scholar] [CrossRef]
  19. Wang, T.; Li, C.; Xie, X.; Lu, B.; He, Z.; Liang, S.; Zhou, J. Anode materials for aqueous zinc ion batteries: Mechanisms, properties, and perspectives. Acs Nano 2020, 14, 16321–16347. [Google Scholar] [CrossRef]
  20. Liu, N.; Li, B.; He, Z.X.; Dai, L.; Wang, H.Y.; Wang, L. Recent advances and perspectives on vanadium- and manganese-based cathode materials for aqueous zinc ion batteries. J. Energy Chem. 2021, 59, 134–159. [Google Scholar] [CrossRef]
  21. Baldinelli, A.; Barelli, L.; Bidini, G.; Discepoli, G. Economics of innovative high capacity-to-power energy storage technologies pointing at 100% renewable micro-grids. J Energy Storage 2020, 28, 101198. [Google Scholar] [CrossRef]
  22. Busacca, C.; Di Blasi, O.; Giacoppo, G.; Briguglio, N.; Antonucci, V.; Di Blasi, A. High performance electrospun nickel manganite on carbon nanofibers electrode for vanadium redox flow battery. Electrochim. Acta 2020, 355, 136755. [Google Scholar] [CrossRef]
  23. Chung, Y.; Noh, C.; Kwon, Y. Role of borate functionalized carbon nanotube catalyst for the performance improvement of vanadium redox flow battery. J. Power Sources 2019, 438, 227063. [Google Scholar] [CrossRef]
  24. Divya, K.; Rana, D.; Saraswathi, M.; Nagendran, A. Custom-made sulfonated poly (vinylidene fluoride-co-hexafluoropropylene) nanocomposite membranes for vanadium redox flow battery applications. Polym. Test 2020, 90, 106685. [Google Scholar] [CrossRef]
  25. Hu, G.J.; Jing, M.H.; Wang, D.W.; Sun, Z.H.; Xu, C.; Ren, W.C.; Cheng, H.M.; Yan, C.W.; Fan, X.Z.; Li, F. A gradient bi-functional graphene-based modified electrode for vanadium redox flow batteries. Energy Storage Mater. 2018, 13, 66–71. [Google Scholar] [CrossRef]
  26. Lv, Y.R.; Han, C.; Zhu, Y.; Zhang, T.; Yao, S.; He, Z.X.; Dai, L.; Wang, L. Recent advances in metals and metal oxides as catalysts for vanadium redox flow battery: Properties, structures, and perspectives. J. Mater. Sci. Technol. 2021, 75, 96–109. [Google Scholar] [CrossRef]
  27. Deng, Q.; Tian, Y.; Ding, P.; Yue, J.P.; Zeng, X.X.; Yin, Y.X.; Wu, X.W.; Lu, X.Y.; Guo, Y.G. Porous lamellar carbon assembled from Bacillus mycoides as high-performance electrode materials for vanadium redox flow batteries. J. Power Sources 2020, 450, 227633. [Google Scholar] [CrossRef]
  28. Xia, L.; Zhang, Q.F.; Wu, C.; Liu, Y.R.; Ding, M.; Ye, J.Y.; Cheng, Y.H.; Jia, C.K. Graphene coated carbon felt as a high-performance electrode for all vanadium redox flow batteries. Surf. Coat Tech. 2019, 358, 153–158. [Google Scholar] [CrossRef]
  29. Zhao, Y.; Ding, Y.; Li, Y.; Peng, L.; Byon, H.R.; Goodenough, J.B.; Yu, G. A chemistry and material perspective on lithium redox flow batteries towards high-density electrical energy storage. Chem. Soc. Rev. 2015, 44, 7968–7996. [Google Scholar] [CrossRef] [Green Version]
  30. Li, W.Y.; Liu, J.G.; Yan, C.W. Multi-walled carbon nanotubes used as an electrode reaction catalyst for VO2+/VO2+ for a vanadium redox flow battery. Carbon 2011, 49, 3463–3470. [Google Scholar] [CrossRef]
  31. He, Z.X.; Cheng, G.; Jiang, Y.Q.; Li, Y.H.; Zhu, J.; Meng, W.; Zhou, H.Z.; Dai, L.; Wang, L. Novel 2D porous carbon nanosheet derived from biomass: Ultrahigh porosity and excellent performances toward V2+/V3+ redox reaction for vanadium redox flow battery. Int. J. Hydrog. Energy 2020, 45, 3959–3970. [Google Scholar] [CrossRef]
  32. Cheng, D.; Tian, M.; Wang, B.; Zhang, J.; Chen, J.; Feng, X.; He, Z.; Dai, L.; Wang, L. One-step activation of high-graphitization N-doped porous biomass carbon as advanced catalyst for vanadium redox flow battery. J. Colloid Interface Sci. 2020, 572, 216–226. [Google Scholar] [CrossRef]
  33. Wei, L.; Zhao, T.S.; Zeng, L.; Zhou, X.L.; Zeng, Y.K. Copper nanoparticle-deposited graphite felt electrodes for all vanadium redox flow batteries. Appl. Energy 2016, 180, 386–391. [Google Scholar] [CrossRef]
  34. Jiang, Y.; Cheng, G.; Li, Y.; He, Z.; Zhu, J.; Meng, W.; Dai, L.; Wang, L. Promoting vanadium redox flow battery performance by ultra-uniform ZrO2@C from metal-organic framework. Chem. Eng. J. 2021, 415, 129014. [Google Scholar] [CrossRef]
  35. Zhou, H.; Shen, Y.; Xi, J.; Qiu, X.; Chen, L. ZrO2-nanoparticle-modified graphite felt: Bifunctional effects on vanadium flow batteries. ACS Appl. Mater. Inter. 2016, 8, 15369–15378. [Google Scholar] [CrossRef]
  36. Vazquez-Galvan, J.; Flox, C.; Fabrega, C.; Ventosa, E.; Parra, A.; Andreu, T.; Morante, J.R. Hydrogen-treated rutile TiO2 shell in graphite-core structure as a negative electrode for high-performance vanadium redox flow batteries. ChemSusChem 2017, 10, 2089–2098. [Google Scholar] [CrossRef] [PubMed]
  37. Bayeh, A.W.; Lin, G.-Y.; Chang, Y.-C.; Kabtamu, D.M.; Chen, G.-C.; Chen, H.-Y.; Wang, K.-C.; Wang, Y.-M.; Chiang, T.-C.; Huang, H.-C.; et al. Oxygen-vacancy-rich cubic CeO2 nanowires as catalysts for vanadium redox flow batteries. ACS Sustain. Chem. Eng. 2020, 8, 16757–16765. [Google Scholar] [CrossRef]
  38. Bayeh, A.W.; Kabtamu, D.M.; Chang, Y.-C.; Chen, G.-C.; Chen, H.-Y.; Liu, T.-R.; Wondimu, T.H.; Wang, K.-C.; Wang, C.-H. Hydrogen-treated defect-rich W18O49 nanowires modified graphite felt as high-performance electrode for vanadium redox flow battery. ACS Appl. Energy Mater. 2019, 2, 2541–2551. [Google Scholar] [CrossRef]
  39. Mehboob, S.; Ali, G.; Shin, H.-J.; Hwang, J.; Abbas, S.; Chung, K.Y.; Ha, H.Y. Enhancing the performance of all-vanadium redox flow batteries by decorating carbon felt electrodes with SnO2 nanoparticles. Appl. Energy 2018, 229, 910–921. [Google Scholar] [CrossRef]
  40. Wei, L.; Xiong, C.; Jiang, H.R.; Fan, X.Z.; Zhao, T.S. Highly catalytic hollow Ti3C2Tx MXene spheres decorated graphite felt electrode for vanadium redox flow batteries. Energy Storage Mater. 2020, 25, 885–892. [Google Scholar] [CrossRef]
  41. Xue, J.; Jiang, Y.; Zhang, Z.; Zhang, T.; He, Z. A novel catalyst of titanium boride toward V3+/V2+ redox reaction for vanadium redox flow battery. J. Alloy. Compd. 2021, 875, 159915. [Google Scholar] [CrossRef]
  42. Tz, A.; Yj, A.; Zz, A.; Jing, X.A.; Yl, A.; Yl, A.; Zc, B.; Zx, B.; Zl, B.; Lei, D. Zirconium boride as a novel negative catalyst for vanadium redox flow battery. Ceram. Int. 2021, 47, 20276–20285. [Google Scholar]
  43. Ji, Y.; Li, J.L.; Li, S.F.Y. Synergistic effect of the bifunctional polydopamin Mn3O4 composite electrocatalyst for vanadium redox flow batteries. J. Mater. Chem. A 2017, 5, 15154–15166. [Google Scholar] [CrossRef]
  44. Zhang, R.; Li, K.; Ren, S.; Chen, J.; Feng, X.; Jiang, Y.; He, Z.; Dai, L.; Wang, L. Sb-doped SnO2 nanoparticle-modified carbon paper as a superior electrode for a vanadium redox flow battery. Appl. Surf. Sci. 2020, 526, 146685–146695. [Google Scholar] [CrossRef]
  45. Choi, M.S.; Mirzaei, A.; Na, H.G.; Kim, S.; Kim, D.E.; Lee, K.H.; Jin, C.; Choi, S.W. Facile and fast decoration of SnO2 nanowires with Pd embedded SnO2−x nanoparticles for selective NO2 gas sensing. Sens. Actuat. B Chem. 2021, 340, 129984. [Google Scholar] [CrossRef]
  46. Ejigu, A.; Edwards, M.; Walsh, D.A. Synergistic catalyst-support interactions in a graphene-Mn3O4 electrocatalyst for vanadium redox flow batteries. ACS Catal. 2015, 5, 7122–7130. [Google Scholar] [CrossRef] [Green Version]
  47. Xu, S.; Liu, J.; Xue, Y.; Wu, T.; Zhang, Z. Appropriate conditions for preparing few-layered graphene oxide and reduced graphene oxide. Fuller. Nanotub. Carbon Nanostruct. 2017, 25, 40–46. [Google Scholar] [CrossRef]
  48. Yu, X.F.; Wu, Q.B.; Zhang, H.Y.; Zeng, G.X.; Li, W.W.; Qian, Y.N.; Li, Y.; Yang, G.Q.; Chen, M.Y. Investigation on synthesis, stability, and thermal conductivity properties of water-based SnO2/reduced graphene oxide nanofluids. Materials 2018, 11, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Li, G.D.; Shen, Y.B.; Zhou, P.F.; Hao, F.L.; Fang, P.; Wei, D.Z.; Meng, D.; San, X.G. Design and application of highly responsive and selective rGO-SnO2 nanocomposites for NO2 monitoring. Mater. Charact. 2020, 163, 110284. [Google Scholar] [CrossRef]
  50. Li, K.; Jiang, Y.Q.; Zhang, R.C.; Ren, S.Z.; Feng, X.J.; Xue, J.; Zhang, T.X.; Zhang, Z.X.; He, Z.X.; Dai, L.; et al. Oxygen vacancy and size controlling endow tin dioxide with remarked electrocatalytic performances towards vanadium redox reactions. Colloid Surf. A 2020, 602, 125073. [Google Scholar] [CrossRef]
  51. Han, P.X.; Wang, H.B.; Liu, Z.H.; Chen, X.A.; Ma, W.; Yao, J.H.; Zhu, Y.W.; Cui, G.L. Graphene oxide nanoplatelets as excellent electrochemical active materials for VO2+/VO2+ and V2+/V3+ redox couples for a vanadium redox flow battery. Carbon 2011, 49, 693–700. [Google Scholar] [CrossRef]
  52. Yang, S.B.; Feng, X.L.; Zhi, L.J.; Cao, Q.A.; Maier, J.; Mullen, K. Nanographene-constructed hollow carbon spheres and their favorable electroactivity with respect to lithium storage. Adv. Mater. 2010, 22, 838. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Illustration of preparation procedure of SnO2/rGO by sol-gel method.
Figure 1. Illustration of preparation procedure of SnO2/rGO by sol-gel method.
Molecules 26 05085 g001
Figure 2. (a) SEM image of GO; (b) SEM image of SnO2; (c) SEM image of SnO2/rGO; (d,e) TEM images; (f) high-resolution TEM image of SnO2/rGO; (g) EDX spectra of GO; (h) EDX spectra of SnO2/rGO; (i) XRD patterns of all samples.
Figure 2. (a) SEM image of GO; (b) SEM image of SnO2; (c) SEM image of SnO2/rGO; (d,e) TEM images; (f) high-resolution TEM image of SnO2/rGO; (g) EDX spectra of GO; (h) EDX spectra of SnO2/rGO; (i) XRD patterns of all samples.
Molecules 26 05085 g002
Figure 3. (a) N2 adsorption–desorption isotherms of GO; (b) SnO2/rGO; (c) pore-size distributions of GO; (d) SnO2/rGO based on BJH model.
Figure 3. (a) N2 adsorption–desorption isotherms of GO; (b) SnO2/rGO; (c) pore-size distributions of GO; (d) SnO2/rGO based on BJH model.
Molecules 26 05085 g003
Figure 4. (a) XPS spectra of SnO2/rGO, and the high-resolution XPS spectra of (b) C 1s, (c) O 1s, and (d) Sn 3d.
Figure 4. (a) XPS spectra of SnO2/rGO, and the high-resolution XPS spectra of (b) C 1s, (c) O 1s, and (d) Sn 3d.
Molecules 26 05085 g004
Figure 5. (a) CV curves for SnO2, GO, and SnO2/rGO in 1.6 M VO2+ + 3.0 M H2SO4 at a scan rate of 10 mV s−1; (b) CVs for GO; (c) CVs for SnO2/rGO; (d) CVs at different scan rates (5–25 mV s−1) in 1.6 M VO2+ + 3.0 M H2SO4, and plots of the redox peak current versus the square root of the scan rate for GO and SnO2/rGO electrodes; (e) Nyquist plots for SnO2, GO, and SnO2/rGO in 1.6 M VO2+ + 3.0 M H2SO4 at 0.85 V; (f) simplified electrical equivalent circuit fitting with Nyquist plots.
Figure 5. (a) CV curves for SnO2, GO, and SnO2/rGO in 1.6 M VO2+ + 3.0 M H2SO4 at a scan rate of 10 mV s−1; (b) CVs for GO; (c) CVs for SnO2/rGO; (d) CVs at different scan rates (5–25 mV s−1) in 1.6 M VO2+ + 3.0 M H2SO4, and plots of the redox peak current versus the square root of the scan rate for GO and SnO2/rGO electrodes; (e) Nyquist plots for SnO2, GO, and SnO2/rGO in 1.6 M VO2+ + 3.0 M H2SO4 at 0.85 V; (f) simplified electrical equivalent circuit fitting with Nyquist plots.
Molecules 26 05085 g005
Figure 6. (a) CV curves; (b) CV curves at the scan rate of 10 mV s1 and Nyquist plots for GO and SnO2/rGO in 1.6 M V3+ + 3.0 M H2SO4 electrolyte.
Figure 6. (a) CV curves; (b) CV curves at the scan rate of 10 mV s1 and Nyquist plots for GO and SnO2/rGO in 1.6 M V3+ + 3.0 M H2SO4 electrolyte.
Molecules 26 05085 g006
Figure 7. (a,b) SEM images of pristine graphite felt; (c,d) modified graphite felt with SnO2/rGO.
Figure 7. (a,b) SEM images of pristine graphite felt; (c,d) modified graphite felt with SnO2/rGO.
Molecules 26 05085 g007
Figure 8. (a) Discharge capacity; (b) CE, (c) VE, and (d) EE of cells with and without SnO2/rGO; (e) charge–discharge curves of pristine cell; (f) SnO2/rGO modified cell at current density of 50–150 mA cm−2.
Figure 8. (a) Discharge capacity; (b) CE, (c) VE, and (d) EE of cells with and without SnO2/rGO; (e) charge–discharge curves of pristine cell; (f) SnO2/rGO modified cell at current density of 50–150 mA cm−2.
Molecules 26 05085 g008
Table 1. Corresponding fitting electrochemical parameters for SnO2, GO, and SnO2/rGO.
Table 1. Corresponding fitting electrochemical parameters for SnO2, GO, and SnO2/rGO.
SampleRsQmRctQt
Y0n0Y1n1
GO7.7716.43 × 10−30.2437.42.86 × 10−50.859
SnO28.9473.12 × 10−20.48177.53.08 × 10−50.856
SnO2/rGO7.0013.14 × 10−20.537.414.95 × 10−40.663
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Jiang, Y.; Lv, Y.; He, Z.; Dai, L.; Wang, L. Synergistic Catalysis of SnO2/Reduced Graphene Oxide for VO2+/VO2+ and V2+/V3+ Redox Reactions. Molecules 2021, 26, 5085. https://doi.org/10.3390/molecules26165085

AMA Style

Liu Y, Jiang Y, Lv Y, He Z, Dai L, Wang L. Synergistic Catalysis of SnO2/Reduced Graphene Oxide for VO2+/VO2+ and V2+/V3+ Redox Reactions. Molecules. 2021; 26(16):5085. https://doi.org/10.3390/molecules26165085

Chicago/Turabian Style

Liu, Yongguang, Yingqiao Jiang, Yanrong Lv, Zhangxing He, Lei Dai, and Ling Wang. 2021. "Synergistic Catalysis of SnO2/Reduced Graphene Oxide for VO2+/VO2+ and V2+/V3+ Redox Reactions" Molecules 26, no. 16: 5085. https://doi.org/10.3390/molecules26165085

Article Metrics

Back to TopTop