Next Article in Journal
Climate Change Impacts on Gaseous Hydrogen (H2) Potential Produced by Photovoltaic Electrolysis for Stand-Alone or Grid Applications in Europe
Previous Article in Journal
Deep-Learning-Based Flow Prediction for CO2 Storage in Shale–Sandstone Formations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hybrid Anionic Electrolytes for the High Performance of Aqueous Zinc-Ion Hybrid Supercapacitors

1
Biomass New Materials Research Center, College of Architectural Engineering, Yunnan Agricultural University, Kunming 650201, China
2
Department of Science and Technology, Yunnan Agricultural University, Kunming 650201, China
3
Yunnan International Joint R&D Center of Smart Agriculture and Water Security, Yunnan Agricultural University, Kunming 650201, China
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(1), 248; https://doi.org/10.3390/en16010248
Submission received: 5 December 2022 / Revised: 16 December 2022 / Accepted: 21 December 2022 / Published: 26 December 2022
(This article belongs to the Section D: Energy Storage and Application)

Abstract

:
Aqueous zinc-ion hybrid supercapacitors (AZHSs) are promising candidates for powering mobile devices due to their intrinsically high safety, the high theoretical capacity of zinc anodes, and the wide range of sources of raw materials for activated carbon (AC) cathodes. Here, we report that there is a synergistic effect between the anions of an AZHS electrolyte, which can significantly improve the specific capacity and rate capability of an AC cathode. The results showed that the specific capacities of the AC cathode//2 M ZnSO4(aq)//Zn anode energy storage system were 115 and 41 mAh g−1 at 0.1 and 5 A g−1 current densities, respectively. The specific capacity at a 0.1 A g−1 current density was enhanced to 136 mAh g−1 by doping 0.5% ZnCl2 and 0.5% Zn(CF3SO3)2 in the 2 M ZnSO4 electrolyte. The specific capacity at a 5 Ag−1 current density was enhanced to 69 mAh g−1 by doping 1% ZnCl2 and 0.5% Zn(CF3SO3)2 in the 2 M ZnSO4 electrolyte. In addition, the co-doped electrolyte increased the energy consumption of the binding of the AC surface groups with H+ and inhibited the precipitation of Zn4SO4(OH)6·5H2O. This provides an important perspective for improving the performance of AZHSs.

Graphical Abstract

1. Introduction

The commitment of countries to solve the problem of global warming has led to a boom in research and practice on carbon emission mitigation, especially in the promotion and application of electric vehicles and electric agricultural machinery that replace oil-fired combustion with electricity [1,2,3]. The most critical technology for these electricity-powered mechanical devices is the energy storage system [4,5,6], so there is an urgent requirement to develop high-performance, safe, and inexpensive energy storage devices. Aqueous zinc-ion hybrid supercapacitors (AZHSs) inherit the high-power function of supercapacitors (SCs) and many advantages of some batteries [7] and have become one of the most promising candidates for promoting the application of electric vehicles and electric agricultural machinery. The zinc anodes of AZHSs have the advantages of a high theoretical electrical capacity (820 mAh g−1), a low redox potential (0.76 V compared to standard hydrogen electrodes), abundant resources, and low prices [8]. High-performance cathode materials such as B/N co-doped porous carbon [9], carbon nanotubes (CNT) [10,11], graphene (GO) [12], and commercial activated carbon (AC) [13,14] have been widely studied and applied. However, the development of practical AZHSs requires the consideration of a suitable electrolyte, as a suitable electrolyte can inhibit the dendrite growth of the Zn anode [15], thereby improving the cycling performance of the device. More importantly, the precise matching of the electrolyte ion size to the aperture of the carbon material facilitates its fast insertion/de-insertion inside the cathode, enhancing the energy density and output power of the device [16]. Therefore, the development of high-performance electrolytes is of great importance for the popularization of AZHSs.
In terms of electrolyte development, the high safety, relaxed preparation conditions, and low cost have been the eye-catching advantages of water-based electrolytes [17]. However, the AC cathodes of AZHSs still suffer from low energy density and the surface generation of side reactants [18]. Current aqueous electrolytes such as ZnSO4 and Zn(CF3SO3)2 are not able to provide the ideal energy density for AZHSs [19], nor can they solve the problem of the precipitation of side reactants on the surface of the cathode material [20]. Therefore, it is necessary to prepare a high energy density or inhibit the side reactants of high-performance aqueous electrolytes for AZHSs using new strategies. At present, many researchers consider the development of hybrid electrolytes, based on the synergistic effects between the ions and ions and between the ions and groups, an effective strategy to enhance the electrochemical performance of ZHSs [21,22], which have made significant progress in the development of organic electrolytes [23], hydrogel electrolytes [24], and organic–hydrogel electrolytes [25]. However, applying the concept of the synergistic effect between the ions to the development of high-performance aqueous electrolytes is an extremely interesting research idea. Further, different anions of AZHS electrolytes have different physical adsorption processes inside the AC cathode due to the different sizes and ion adsorption reaction kinetics between the anions, and the mixing of multiple anions is highly likely to have synergistic or antagonistic effects [26] that affect the electrochemical performance of the device. Here, 2 M ZnSO4, 2 M ZnCl2, and 2 M Zn(CF3SO3)2 aqueous solutions were used as electrolytes, AC was used as a cathode, and Zn was used as an anode to determine whether a 2 M ZnSO4 electrolyte was suitable as a mother liquor. Subsequently, we focused on the most suitable addition of ZnCl2 and Zn(CF3SO3)2 in the 2 M ZnSO4 electrolyte to determine whether there is an interaction between the SO42−, Cl, and CF3SO3 anions.
We report that this hybrid anionic electrolyte enhanced the specific capacity and rate capability of the AC cathode through the synergistic effect between the SO42−, Cl, and CF3SO3. Last but not least, this hybrid anionic electrolyte with the coexistence of SO42−, Cl, and CF3SO3 increased the energy consumption of the AC cathode surface groups for binding H+, reduced the consumption of H+ in the electrolyte during the discharge of AZHSs, and made it difficult to form a local alkaline environment on the AC cathode surface, which inhibited the precipitation of Zn4SO4(OH)6·5H2O, thus avoiding damage to the precision structure of the device due to the growth of dendrites on the surface of the AC cathode. The experimental results showed that the electrochemical performance of the 2 M ZnSO4 electrolyte was better than those of the 2 M ZnCl2 and 2 M Zn(CF3SO3)2 electrolytes. The specific capacities of the assembled AC cathode//2 M ZnSO4(aq)//Zn anode energy storage devices were 115 and 41 mAh g−1 at 0.1 and 5 Ag−1 current densities, respectively. Due to the synergistic effect between the anions, the 2 M ZnSO4 + 0.01 M ZnCl2, 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2, and 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.01 M ZnCl2 electrolytes had high energy densities of 133, 124, and 136 mAh g−1 at a current densities of 0.1 A g−1, respectively. The prepared electrolyte of 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2 achieved a specific capacity of 69 mAh g−1 at a current density of 5 A g−1. Therefore, the concept of anion co-doping opens a new direction for the development of AZHS electrolytes and provides an important reference for the development of hybrid aqueous electrolytes.

2. Methods and Methods

2.1. Preparation of AZHS Electrodes

An AC cathode was prepared by mixing an AC powder, acetylene black, and polyvinylidene fluoride (PVDF) dispersion in a mass ratio of 8:1:1 and then coated on a stainless-steel mesh (the prepared electrodes needed to be dried in an oven at 60 °C for 8 h). The stainless-steel mesh size was 1 cm × 2 cm, and the load capacity was 1 mg cm−2 [15]. The thickness of the zinc foil was 0.1 mm. It was purchased from Tianjin Anuohe New Energy Technology Co., Ltd. in China, polished, and directly used as an anode electrode. The energy storage device with an AC cathode//2 M ZnSO4(aq)//Zn anode system was assembled with a hybrid anionic electrolyte, AC cathode, Zn anode, and electrochemical cell shell [27].

2.2. Formulation of AZHS Electrolytes

Three electrolytes (2 M ZnSO4, 2 M ZnCl2, and 2 M Zn(CF3SO3)2) were prepared to evaluate the effect of electrolyte type on the electrochemical performance of AZHSs in order to show the effect of the hybrid anionic electrolyte on the electrochemical performance of AZHSs doped with 0.005 M, 0.01 M, 0.02 M, 0.03 M, and 0.05 M ZnCl2 or Zn(CF3SO3)2 electrolyte in the 2 M ZnSO4 electrolyte. Subsequently, ZnCl2 and Zn(CF3SO3)2 were co-doped in the 2 M ZnSO4 electrolyte to observe the effects of anion interactions on the electrochemical properties of AZHSs.

2.3. Material Characterization

Field emission scanning electron microscopy (SEM, JSM-7800F, Japan) was used to characterize the morphologies of the Zn anodes and AC cathodes. X-ray diffractograms of Cu Kα irradiation (λ = 1.54178 Å) were recorded with an X-ray diffractometer (XRD, Rigaku Ulitma IV, Japan) to observe the surface material structures of the Zn anodes and AC cathodes. An X-ray photoelectron spectrometer (XPS, Thermo Scientific ESCALAB 250Xi) was used for the elemental analyses of the Zn anode and AC cathode.

2.4. Electrochemical Performance Testing of AZHS Electrodes

AZHSs were tested for electrochemical properties, including the cyclic voltammetry (CV), constant current charge/discharge technique (GCD), electrochemical impedance spectroscopy (EIS), and Tafel curve measurements. The CV and EIS tests were performed on a CHI 660 electrochemical workstation with an AC cathode as the working electrode and Zn as the counter and reference electrodes. GCD tests were performed on a Sunway High Performance Battery Testing System with a voltage range of 0.2–1.8 V, and the specific capacities of AZHSs were evaluated at current densities of 0.1, 1, and 5 A g−1. In order to show that AZHSs are stable, they were charged and discharged for 200 cycles at a current density of 1 A g−1. In order to exclude the effect of the Zn2+ concentration on the performance of AZHSs, a hybrid anionic electrolyte with the same Zn2+ concentration and a ZnSO4 solution were prepared and charged and discharged for 1000 cycles at a current density of 5 A g−1. Tafel measurements were performed on a CHI 660 electrochemical workstation with a working electrode with an AC cathode, original Zn foil, used Zn anode, counter electrode of Zn, and reference electrode of saturated glycerol. The specific capacity (C, mAh g−1) for the ZHS device was precisely determined through the discharge curve (GD) by the equation below [9]:
C = 2 I Vdt / 3.6 Vm

3. Results and Discussion

3.1. Comparison of the Performances of Different Electrolytes

As shown in Figure 1a, three AZHS energy storage systems were assembled using a Zn anode and an AC cathode with 2 M ZnSO4, 2 M ZnCl2, and 2 M Zn(CF3SO3)2 as the electrolytes. As shown in Figure 1b, the 2 M ZnSO4 electrolyte exhibited a slightly larger peak current density and CV curve integration area than the 2 M Zn(CF3SO3)2 electrolyte, indicating a better energy storage performance. In addition, the scan curve of the 2 M ZnCl2 electrolyte was abnormal, which may be due to the irreversible etching of the stainless-steel mesh by the ZnCl2 electrolyte, resulting in the inability of AZHSs to store energy properly. To further illustrate the specific capacity and stability of the three AZHSs, we demonstrated the rate capabilities and cycling performances, and the results are shown in Figure 1c–d. The 2 M ZnSO4 electrolyte in Figure 1c exhibited a higher specific capacity of 115 mAh g−1 at a current density of 0.1 A g−1, which was higher than that of the 2 M Zn(CF3SO3)2 electrolyte, at 89 mAh g−1. The capacitance retention rate of the 2 M ZnSO4 electrolyte with a 0.1–1 A g−1 current density change was 61.7%, which was slightly lower than the 64.5% of the 2 M Zn(CF3SO3)2 electrolyte. In addition, the electrolyte voltage of the 2 M ZnCl2 was unable reach 1.8 V. The specific capacity of the 2 M ZnSO4 electrolyte in Figure 1d at a current density of 5 A g−1 was 41 mAh g−1, which was higher than the 37 mAh g−1 of the 2 M Zn(CF3SO3)2 electrolyte. The capacitance retention of the 2 M ZnSO4 electrolyte was 98% after 200 cycles of charge and discharge, which is higher than the 96% of the 2 M Zn(CF3SO3)2 electrolyte.

3.2. Characterization Testing of Zinc Anodes

We observed the electrode states in the AC cathode//2 M ZnSO4(aq) + additive//Zn anode energy storage system by a characterization technique. First, we analyzed the reactions occurring on the zinc anode by electron microscopy scanning (SEM), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS). The micromorphology and phase analysis results of the zinc foil anode after charging and discharging in different electrolytes are shown in Figure 2. Compared to the pristine zinc foil in Figure 2a, the zinc foil anode after 200 cycles of charging and discharging in the electrolyte exhibited a relatively rough surface (Figure 2b–e). Subsequently, we analyzed the phase composition of the zinc foil anodes with XRD tests. As shown in Figure 2f, the most dominant phase component of the zinc foil anode was always Zn (JCPDS#87-0713), but other diffraction peaks appeared on the surface of the zinc foil anode, which indicated the formation of Zn4SO4(OH)6·5H2O [28] (JCPDS#39-0688) on the surface of the zinc metal. At the same time, the XPS pattern in Figure S1 (see Supplementary Materials) also confirmed the presence of the side reactants.

3.3. AC Characterization Test of AC Cathode

The AZHSs were disassembled, and the AC cathodes were removed, cleaned using pure water for 2–3 min, and dried under vacuum at a temperature of 60 °C for 6 h before a series of characterization tests were performed. From the SEM image in Figure 3a, it can be observed that the AC particles were irregular, the surface was relatively rough, and most of the particles were 5–10 μm in diameter. Due to the faster ion insertion reaction kinetics of H+, it preferentially undergoes coordination/decoordination reactions with AC cathode surface groups, resulting in Zn2+ reacting with OH to form Zn4SO4(OH)6·5H2O [29]. The continuous accumulation of these side reactants led to the formation of the lamellar dendrites on the surface of the AC cathode in Figure 3b–d. Compared with Figure 3b–d, we found that the morphology of the AC cathode in the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 + 0.02 M ZnCl2 electrolyte in Figure 3e was not significantly changed. As shown in the XRD pattern of Figure 3f, the intensity of the AC cathode impurity peaks in the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 + 0.02 M ZnCl2 electrolyte was significantly weaker than that of the other electrolytes, which once again proved that the electrolyte can effectively inhibit the precipitation of side reactants. The following equation has been reported for the deposition reaction process of Zn4SO4(OH)6·5H2O [15,30]:
{AC}……O + H+ + e → {AC}……OH (discharge process)
{AC}……OH → {AC}……O + H+ + e (charge process)
4Zn2+ + 6OH + SO42− + xH2O = Zn4SO4(OH)6·xH2O↓
The above equation shows that the binding of the AC cathode surface groups with H+ is the main reason for the precipitation of Zn4SO4(OH)6·5H2O. To investigate the effect of the electrolyte on the occurrence of coordination/non-coordination reactions between the AC cathode surface groups and H+, we tested the surface overpotential using Tafel curves to determine the ease of electron transfer of the AC cathode surface groups. As shown in Figure 3g,h, the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 + 0.02 M ZnCl2 electrolyte (1) increased the energy consumption of the H+ release from the AC cathode surface groups in the charging state, making it difficult to release H+ and (2) increased the combined energy consumption of H+ with the AC cathode surface groups in the discharging state, making it difficult to form a local alkaline environment on the AC cathode surface and inhibiting the precipitation of dendrites.

3.4. The 2 M ZnSO4 Electrolyte Doped with ZnCl2

First, we tested the electrochemical performance of the 2 M ZnSO4 electrolyte doped with ZnCl2. Figure 4a compares the CV curves of five hybrid anionic electrolytes, which showed approximately rectangular curves, which is a typical electric double-layer energy storage feature. Numbers 1, 2, 3, 4, and 5 of Figure 4b correspond to the 2 M ZnSO4 + 0.005 M ZnCl2, 2 M ZnSO4 + 0.01 M ZnCl2, 2 M ZnSO4 + 0.02 M ZnCl2, 2 M ZnSO4 + 0.03 M ZnCl2, and 2 M ZnSO4 + 0.05 M ZnCl2 electrolytes, respectively. The data show that the electrochemical performance of the 2 M ZnSO4 + 0.01 M ZnCl2 electrolyte was excellent, and the specific capacities at current densities of 0.1, 1, and 5 A g−1 were 134, 80, and 59 mAh g−1, which were 16.5, 12.7, and 43.9% higher than those of the 2 M ZnSO4 electrolyte, respectively. The cyclic test in Figure 4c showed that the 2 M ZnSO4 + 0.01 M ZnCl2 electrolyte had good cyclic stability, with a capacitance retention of 97% after 200 cycles at a current density of 1 A g−1. To further clarify that the electrochemical performance enhancement of AZHSs was mainly affected by mixed anions, we performed 1000 charge–discharge cycles for the 2 M ZnSO4 + 0.01 M ZnCl2 and 2.01 M ZnSO4 electrolytes. The specific capacity of the 2 M ZnSO4 + 0.01 M ZnCl2 electrolyte in Figure 4d was 1.44 times higher than that of the 2.01 M ZnSO4 electrolyte. The capacitance retention of the 2 M ZnSO4 + 0.01 M ZnCl2 electrolyte was 97.4%, which was higher than that of the 2.01 M ZnSO4 electrolyte, at 94.1%. Electrochemical impedance spectroscopy (EIS) is used to further investigate the electron/ion transport properties inside the electrolyte. Figure 4e shows the Nyquist curves in the high-frequency region, with the first intercept of the Nyquist curve along the real axis being the electrolyte resistance (Re). It can be seen that with the continuous addition of ZnCl2, the Re values of the 2 M ZnSO4 electrolyte were 9.5, 8.8, 7.8, 7.2, 7.5, and 7.5 Ω. The observed decrease in Re values may be due to the synergistic effect between the SO42− and Cl, which reduced the internal resistance and improved the conductivity, which in turn increased the specific capacity of AZHSs at a current density of 0.1 A g−1 [31]. Figure 4f shows the Nyquist curves in the frequency range of 0.01 Hz–1000 KHz. The data indicate that the synergistic effect of SO42− and Cl can significantly reduce the equivalent series resistance (ESR) values. However, with the continuous addition of ZnCl2, the electrolyte showed a regular trend of continuous increases in ESR values and a continuous decrease in the surface ion diffusion rate.

3.5. The 2 M ZnSO4 Electrolyte Doped with Zn(CF3SO3)2

Next, we tested the electrochemical performance of the 2 M ZnSO4 electrolyte doped with Zn(CF3SO3)2. Figure 5a compares the CV curves of five hybrid anionic electrolytes, which showed approximately rectangular curves, which is a typical electric double-layer energy storage feature. Numbers 1, 2, 3, 4, and 5 of Figure 5b correspond to the 2 M ZnSO4 + 0.005 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.03 M Zn(CF3SO3)2, and 2 M ZnSO4 + 0.05 M Zn(CF3SO3)2 electrolytes, respectively. The data show that the electrochemical performance of the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 electrolyte was excellent, and the specific capacities at current densities of 0.1, 1, and 5 A g−1 were 134, 80, and 59 mAh g−1, which were 7.8, 23.9, and 63.4% higher than those of the 2 M ZnSO4 electrolyte, respectively. The cycling test in Figure 5c showed that the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 electrolyte had good cycling stability, with a capacitance retention of 93.4% after 200 charge–discharge cycles at a current density of 1 A g−1. To further clarify that the specific capacity enhancement of AZHSs was mainly influenced by the mixed anions, we performed 1000 charge–discharge cycles for the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 and 2.02 M ZnSO4 electrolytes. As shown in Figure 5d, the specific capacity of the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 electrolyte was 1.56 times higher than that of the 2.02 M ZnSO4 electrolyte. The capacitance retention of the 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 electrolyte was 99.5%, which was higher than that of the 2.02 M ZnSO4 electrolyte at 95.1%. Figure 5e shows the Nyquist curves in the high-frequency region. It can be seen that, with the continuous addition of Zn(CF3SO3)2, the Re values of the 2 M ZnSO4 electrolyte were 9.5, 10.5, 11.5, 13.6, 11.8, and 12.2 Ω, respectively. The observed increase in the Re value may be attributed to the larger resistance value of Zn(CF3SO3)2 (Figure S2 please see the supplementary materials). Figure 5f shows the Nyquist curves in the frequency range of 0.01 Hz–1000 KHz. The data suggest that the synergistic effect of SO42− and CF3SO3 can significantly reduce the equivalent series resistance (ESR) values. With the continuous addition of Zn(CF3SO3)2, the electrolyte showed an irregular trend of ESR value changes and continuous decreases in the surface ion diffusion rate.

3.6. The 2 M ZnSO4 Electrolyte co-Doped with ZnCl2 and Zn(CF3SO3)2

Finally, Zn(CF3SO3)2 and ZnCl2 were co-doped into a 2 M ZnSO4 electrolyte and subsequently subjected to a series of electrochemical performance tests. Figure 6a compares the CV curves of five hybrid anionic electrolytes, which showed approximately rectangular curves, which is a typical electric double-layer energy storage feature. Numbers 1, 2, 3, 4, and 5 of Figure 6b correspond to the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.005 M ZnCl2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.01 M ZnCl2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2, 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 + 0.01 M ZnCl2, and 2 M ZnSO4 + 0.03 M Zn(CF3SO3)2 + 0.01 M ZnCl2 electrolytes. The data show that the highest specific capacity at a current density of 0.1 A g−1 belonged to the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.01 M ZnCl2 electrolyte, with a specific capacity of 136 mAh g−1, which was 18.3% higher than that of the 2 M ZnSO4 electrolyte. The highest specific capacity at a current density of 1 A g−1 belonged to the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.005 M ZnCl2 electrolyte, with a specific capacity of 89 mAh g−1, which was 25.4% higher than that of the 2 M Zn SO4 electrolyte. The highest specific capacity at a current density of 5 A g−1 belonged to the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2 electrolyte, with a specific capacity of 69 mAh g−1, which was 68.3% higher than that of the 2 M ZnSO4 electrolyte. The cycling test in Figure 6c showed that the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.005 M ZnCl2 electrolyte had good cycling stability, with a capacitance retention of 96.8% after 200 cycles at a current density of 1 A g−1. To further clarify that the capacity enhancement of AZHSs was mainly influenced by the mixed anions, we performed 1000 charge–discharge cycles for the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2 and 2.03 M ZnSO4 electrolytes. As shown in Figure 6d, the specific capacity of the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2 electrolyte was 1.56 times higher than that of the 2.03 M ZnSO4 electrolyte. The capacitance retention of the 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2 electrolyte was 99.1%, which was higher than that of the 2.03 M ZnSO4 electrolyte (87.3%). Figure 6e shows the Nyquist curves in the high-frequency region. It can be seen that the co-doping of Zn(CF3SO3)2 and ZnCl2 led to an increase in the Re values of the 2 M ZnSO4 electrolyte, with Re values of 9.5, 10.8, 12.2, 9.9, 10.4, and 13.3 Ω. Figure 6f shows the Nyquist curves in the frequency range of 0.01 Hz–1000 KHz. The data suggest that the co-doping of Zn(CF3SO3)2 and ZnCl2 gave rise to the complex electron/ion transport properties of the 2 M ZnSO4 electrolyte, as evidenced by the extreme instability of the charge transfer/interface resistance and surface ion diffusion rate. In summary, we have reason to believe that there is a complex electrochemical synergistic process between the SO42−, Cl, and CF3SO3 in the hybrid anionic electrolyte that can improve the electrochemical performance of AZHSs and inhibit the dendritic growth of the AC cathode.

4. Conclusions

The performance of AZHSs is limited by the electrochemical properties of the electrolyte. It is necessary to prepare high energy density or inhibit the side reactants of high-performance aqueous electrolytes for AZHSs. We report an aqueous hybrid anionic electrolyte based on ZnSO4, ZnCl2, and Zn(CF3SO3)2 and demonstrate that there is a synergistic effect between the SO42−, Cl, and CF3SO3 that can significantly enhance the electrochemical performance of AZHSs. In addition, we found that the coexistence of SO42−, Cl, and CF3SO3 in the electrolytes of AZHSs can effectively inhibit the precipitation of Zn4SO4(OH)6·5H2O by increasing the coordination/uncoordinated energy consumption of AC surface groups with H+. The results show that the specific capacities of the AC cathode//2 M ZnSO4(aq)//Zn anode energy storage system were 115 and 41 mAh g−1 at 0.1 and 5 A g−1 current densities. The specific capacity at the 0.1 Ag−1 current density was enhanced to 136 mAh g−1 by doping 0.5% ZnCl2 and 0.5% Zn(CF3SO3)2 in the 2 M ZnSO4 electrolyte. The specific capacity at the 5 Ag−1 current density was enhanced to 69 mAh g−1 by doping 1% ZnCl2 and 0.5% Zn(CF3SO3)2 in the 2 M ZnSO4 electrolyte. More importantly, this concept of anionic co-doping will guide the further development of the most suitable electrolytes for AZHSs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en16010248/s1, Figure S1; Figure S2.

Author Contributions

Conceptualization, B.X. and J.H.; methodology, B.X. and S.L.; writing—original draft preparation, B.X. and Y.S.; writing—review and editing, J.H. and J.L.; funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support from the National Natural Science Foundation of China (12064050); the key R&D Plan of Yunnan Province (202203AC100004); the Major Project of Science and Technology of Yunnan Province (202002AE090010); and the Scientific Research Fund of the Yunnan Provincial Department of Education (2022Y291).

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.

References

  1. Chen, S.G.; Zhang, M.F.; Zou, P.M.; Sun, B.Y.; Tao, S.W. Historical development and novel concepts on electrolytes for aqueous rechargeable batteries. Energy Environ. Sci. 2022, 15, 1805–1839. [Google Scholar] [CrossRef]
  2. Gao, M.; Wang, W.-K.; Zheng, Y.-M.; Zhao, Q.-B.; Yu, H.-Q. Hierarchically porous biochar for supercapacitor and electrochemical H2O2 production. Chem. Eng. J. 2020, 402, 126171. [Google Scholar] [CrossRef]
  3. Fu, M.; Lv, R.; Lei, Y.; Terrones, M. Ultralight Flexible Electrodes of Nitrogen-Doped Carbon Macrotube Sponges for High-Performance Supercapacitors. Small 2021, 17, 2004827. [Google Scholar] [CrossRef] [PubMed]
  4. Sun, Q.; Cao, Z.; Wang, S.; Sun, L.; Zhou, L.; Xue, H.; Ming, J. Bio-inspired heteroatom-doped hollow aurilave-like structured carbon for high-performance sodium-ion batteries and supercapacitors. J. Power Sources 2020, 461, 228128. [Google Scholar] [CrossRef]
  5. Xu, X.L.; Chen, Y.; Zheng, D.; Ruan, P.C.; Cai, Y.H.; Dai, X.J.; Cao, X.H. Ultra-Fast and Scalable Saline Immersion Strategy Enabling Uniform Zn Nucleation and Deposition for High-Performance Zn-Ion Batteries. Small 2021, 17, e2101901. [Google Scholar] [CrossRef]
  6. Zhang, Q.; Luan, J.Y.; Tang, Y.G.; Ji, X.B.; Wang, H.Y. Interfacial Design of Dendrite-Free Zinc Anodes for Aqueous Zinc-Ion Batteries. Angew. Chem. Int. Edit. 2020, 59, 13180–13191. [Google Scholar] [CrossRef]
  7. Wang, F.; Borodin, O.; Gao, T.; Fan, X.; Sun, W.; Han, F.; Wang, C. Highly reversible zinc metal anode for aqueous batteries. Nat. Mater. 2018, 17, 543–549. [Google Scholar] [CrossRef]
  8. Getie, F.A.; Ayele, D.W.; Habtu, N.G.; Yihun, F.A.; Yemata, T.A. Development of electrolytes for rechargeable zinc-air batteries: Current progress, challenges, and future outlooks. SN Appl. Sci. 2022, 4, 270. [Google Scholar] [CrossRef]
  9. Lu, Y.Y.; Li, Z.W.; Bai, Z.Y.; Mi, H.Y.; Ji, C.C.; Pang, H.; Qiu, J.S. High energy-power Zn-ion hybrid supercapacitors enabled by layered B/N co-doped carbon cathode. Nano Energy 2019, 66, 104132. [Google Scholar] [CrossRef]
  10. Tian, Y.H.; Amal, R.; Wang, D.W. An Aqueous Metali-ion Capacitor with Oxidized Carbon Nanotubes and Metalicn Zinc Electrodes. Front. Energy Res. 2016, 4, 34. [Google Scholar] [CrossRef]
  11. Sun, G.; Yang, H.; Zhang, G.; Gao, J.; Jin, X.; Zhao, Y.; Qu, L. A capacity recoverable zinc-ion micro-supercapacitor. Energy Environ. Sci. 2018, 11, 3367–3374. [Google Scholar] [CrossRef]
  12. Wu, S.L.; Chen, Y.T.; Jiao, T.P.; Zhou, J.; Cheng, J.Y.; Liu, B.; Zhang, W.J. An Aqueous Zn-Ion Hybrid Supercapacitor with High Energy Density and Ultrastability up to 80000 Cycles. Adv. Energy Mater. 2019, 9, 1902915. [Google Scholar] [CrossRef]
  13. Wang, H.; Wang, M.; Tang, Y. A novel zinc-ion hybrid supercapacitor for long-life and low-cost energy storage applications. Energy Storage Mater. 2018, 13, 1–7. [Google Scholar] [CrossRef]
  14. Chen, S.; Ma, L.; Zhang, K.; Kamruzzaman, M.; Zhi, C.; Zapien, J.A. A flexible solid-state zinc ion hybrid supercapacitor based on co-polymer derived hollow carbon spheres. J. Mater. Chem. A 2019, 7, 7784–7790. [Google Scholar] [CrossRef]
  15. Dong, L.B.; Ma, X.P.; Li, Y.; Zhao, L.; Liu, W.B.; Cheng, J.Y.; Kang, F.Y. Extremely safe, high-rate and ultralong-life zinc-ion hybrid supercapacitors. Energy Storage Mater. 2018, 13, 96–102. [Google Scholar] [CrossRef]
  16. Rashidi, N.A.; Chai, Y.H.; Ismail, I.S.; Othman, M.F.H.; Yusup, S. Biomass as activated carbon precursor and potential in supercapacitor applications. Biomass Convers. Biorefinery 2022, 1–15. [Google Scholar] [CrossRef]
  17. Zhang, T.S.; Tang, Y.; Guo, S.; Cao, X.X.; Pan, A.Q.; Fang, G.Z.; Liang, S.Q. Fundamentals and perspectives in developing zinc-ion battery electrolytes: A comprehensive review. Energy Environ. 2020, 13, 4625–4665. [Google Scholar] [CrossRef]
  18. Gong, X.; Chen, J.; Lee, P.S. Zinc-Ion Hybrid Supercapacitors: Progress and Future Perspective. Batter. Supercaps 2021, 4, 1529–1546. [Google Scholar] [CrossRef]
  19. Wan, F.; Zhang, L.L.; Dai, X.; Wang, X.Y.; Niu, Z.Q.; Chen, J. Aqueous rechargeable zinc/sodium vanadate batteries with enhanced performance from simultaneous insertion of dual carriers. Nat. Commun. 2018, 9, 1656. [Google Scholar] [CrossRef] [Green Version]
  20. Pan, H.L.; Shao, Y.Y.; Yan, P.F.; Cheng, Y.W.; Han, K.S.; Nie, Z.M.; Liu, J. Reversible aqueous zinc/manganese oxide energy storage from conversion reactions. Nat. Energy 2016, 1, 16039. [Google Scholar] [CrossRef]
  21. Shen, Y.H.; Liu, B.; Liu, X.R.; Liu, J.; Ding, J.; Zhong, C.; Hu, W.B. Water-in-salt electrolyte for safe and high-energy aqueous battery. Energy Storage Mater. 2021, 34, 461–474. [Google Scholar] [CrossRef]
  22. Wang, G.Q.; Liu, J.Q.; Dong, W.N.; Yan, C.; Zhang, W. Nitrogen/sulfur co-doped porous carbon nanosheets and its electrochemical performance. Acta Phys. Sin. 2018, 67. [Google Scholar]
  23. Zhang, L.; Liu, Z.; Wang, G.; Feng, J.; Ma, Q. Developing high voltage Zn(TFSI)2/Pyr14TFSI/AN hybrid electrolyte for a carbon-based Zn-ion hybrid capacitor. Nanoscale 2021, 13, 17068–17076. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, L.; Song, L.; Feng, Y.; Cao, M.; Zhang, P.; Zhang, X.-F.; Yao, J. Zinc ion trapping in a cellulose hydrogel as a solid electrolyte for a safe and flexible supercapacitor. J. Mater. Chem. 2020, 8, 12314–12318. [Google Scholar] [CrossRef]
  25. Wang, H.; Li, X.; Jiang, D.; Wu, S.; Yi, W.; Sun, X.; Li, J. Organohydrogel electrolyte-based flexible zinc-ion hybrid supercapacitors with dendrite-free anode, broad temperature adaptability and ultralong cycling life. J. Power Sources 2022, 528, 231210. [Google Scholar] [CrossRef]
  26. Zhang, Y.J.; Chen, H.L.; Wang, S.J.; Zhao, X.; Kong, F.G. Regulatory pore structure of biomass-based carbon for supercapacitor applications. Microporous Mesoporous Mater. 2020, 297, 110032. [Google Scholar] [CrossRef]
  27. Nguyen, T.; Tran, T.; Zhao, M.; Geng, S.; Ivey, D.G. Ethylene Glycol as an Antifreeze Additive and Corrosion Inhibitor for Aqueous Zinc-Ion Batteries. Batter. Supercaps 2022, 5, e202100420. [Google Scholar]
  28. Chen, M.F.; Chen, J.Z.; Zhou, W.J.; Han, X.; Yao, Y.G.; Wong, C.P. Realizing an All-Round Hydrogel Electrolyte toward Environmentally Adaptive Dendrite-Free Aqueous Zn-MnO2 Batteries. Adv. Mater. 2021, 33, 2007559. [Google Scholar] [CrossRef]
  29. Tie, Z.W.; Liu, L.J.; Deng, S.Z.; Zhao, D.B.; Niu, Z.Q. Proton Insertion Chemistry of a Zinc-Organic Battery. Angew. Chem. Int. Edit. 2020, 59, 4920–4924. [Google Scholar] [CrossRef]
  30. Huang, J.H.; Wang, Z.; Hou, M.Y.; Dong, X.L.; Liu, Y.; Wang, Y.G.; Xia, Y.Y. Polyaniline-intercalated manganese dioxide nanolayers as a high-performance cathode material for an aqueous zinc-ion battery. Nat. Commun. 2018, 9, 2906. [Google Scholar] [CrossRef] [Green Version]
  31. Biesheuvel, P.M.; Bazant, M.Z. Nonlinear dynamics of capacitive charging and desalination by porous electrodes. Phys. Rev. E 2010, 81, 031502. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Digital photographs of the Zn anode, AC cathode, and 2 M ZnSO4, 2 M ZnCl2, and 2 M Zn(CF3SO3)2 electrolytes; (b) Comparison of three CV curves (5 mV s−1); (c) Rate capabilities of three electrolytes at 0.1 and 1 A g−1 current densities; (d) Cycling performances (5 A g−1) of 2 M ZnSO4 and 2 M Zn(CF3SO3)2 electrolytes.
Figure 1. (a) Digital photographs of the Zn anode, AC cathode, and 2 M ZnSO4, 2 M ZnCl2, and 2 M Zn(CF3SO3)2 electrolytes; (b) Comparison of three CV curves (5 mV s−1); (c) Rate capabilities of three electrolytes at 0.1 and 1 A g−1 current densities; (d) Cycling performances (5 A g−1) of 2 M ZnSO4 and 2 M Zn(CF3SO3)2 electrolytes.
Energies 16 00248 g001
Figure 2. (a) Zinc foil; (b) Zinc anode in ZnSO4 electrolyte; (c) Zinc anode in ZnSO4 + ZnCl2 electrolyte; (d) Zinc anode in ZnSO4 + Zn(CF3SO3)2 electrolyte; (e) Zinc anode in ZnSO4 + Zn(CF3SO3)2 + ZnCl2 electrolyte; (f) XRD patterns of five zinc anodes.
Figure 2. (a) Zinc foil; (b) Zinc anode in ZnSO4 electrolyte; (c) Zinc anode in ZnSO4 + ZnCl2 electrolyte; (d) Zinc anode in ZnSO4 + Zn(CF3SO3)2 electrolyte; (e) Zinc anode in ZnSO4 + Zn(CF3SO3)2 + ZnCl2 electrolyte; (f) XRD patterns of five zinc anodes.
Energies 16 00248 g002
Figure 3. (a) AC cathode; (b) AC cathode in ZnSO4 electrolyte; (c) AC cathode in ZnSO4 + ZnCl2 electrolyte; (d) AC cathode in ZnSO4 + Zn(CF3SO3)2 electrolyte; (e) AC cathode in ZnSO4 + Zn(CF3SO3)2 + ZnCl2 electrolyte; (f) XRD patterns of 5 AC cathodes; (g) AC cathode charge state overpotential; (h) AC cathode discharge state overpotential.
Figure 3. (a) AC cathode; (b) AC cathode in ZnSO4 electrolyte; (c) AC cathode in ZnSO4 + ZnCl2 electrolyte; (d) AC cathode in ZnSO4 + Zn(CF3SO3)2 electrolyte; (e) AC cathode in ZnSO4 + Zn(CF3SO3)2 + ZnCl2 electrolyte; (f) XRD patterns of 5 AC cathodes; (g) AC cathode charge state overpotential; (h) AC cathode discharge state overpotential.
Energies 16 00248 g003
Figure 4. (a) Comparison of five CV curves (5 mV s−1); (b) Numbers 1, 2, 3, 4, and 5 are 2 M ZnSO4 + 0.005 M ZnCl2, 2 M ZnSO4 + 0.01 M ZnCl2, 2 M ZnSO4 + 0.02 M ZnCl2, 2 M ZnSO4 + 0.03 M ZnCl2, and 2 M ZnSO4 + 0.05 M ZnCl2 electrolytes, respectively; (c) Cycling performance (1 A g−1) of hybrid anionic electrolytes; (d) Cycling performance (5 A g−1) of electrolytes with the same Zn2+ concentration; (e) Nyquist plot of hybrid anionic electrolyte in high-frequency region; (f) Nyquist plot of hybrid anionic electrolyte.
Figure 4. (a) Comparison of five CV curves (5 mV s−1); (b) Numbers 1, 2, 3, 4, and 5 are 2 M ZnSO4 + 0.005 M ZnCl2, 2 M ZnSO4 + 0.01 M ZnCl2, 2 M ZnSO4 + 0.02 M ZnCl2, 2 M ZnSO4 + 0.03 M ZnCl2, and 2 M ZnSO4 + 0.05 M ZnCl2 electrolytes, respectively; (c) Cycling performance (1 A g−1) of hybrid anionic electrolytes; (d) Cycling performance (5 A g−1) of electrolytes with the same Zn2+ concentration; (e) Nyquist plot of hybrid anionic electrolyte in high-frequency region; (f) Nyquist plot of hybrid anionic electrolyte.
Energies 16 00248 g004
Figure 5. (a) Comparison of five CV curves (5 mV s−1); (b) Numbers 1, 2, 3, 4, and 5 are 2 M ZnSO4 + 0.005 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.03 M Zn(CF3SO3)2, and 2 M ZnSO4 + 0.05 M Zn(CF3SO3)2 electrolytes, respectively; (c) Cycling performance (1 A g−1) of hybrid anionic electrolyte; (d) Cycling performance (5 A g−1) of electrolyte with the same Zn2+ concentration; (e) Nyquist plot of hybrid anionic electrolyte in high-frequency region; (f) Nyquist plot of hybrid anionic electrolyte.
Figure 5. (a) Comparison of five CV curves (5 mV s−1); (b) Numbers 1, 2, 3, 4, and 5 are 2 M ZnSO4 + 0.005 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2, 2 M ZnSO4 + 0.03 M Zn(CF3SO3)2, and 2 M ZnSO4 + 0.05 M Zn(CF3SO3)2 electrolytes, respectively; (c) Cycling performance (1 A g−1) of hybrid anionic electrolyte; (d) Cycling performance (5 A g−1) of electrolyte with the same Zn2+ concentration; (e) Nyquist plot of hybrid anionic electrolyte in high-frequency region; (f) Nyquist plot of hybrid anionic electrolyte.
Energies 16 00248 g005
Figure 6. (a) Comparison of five CV curves (5 mV s−1); (b) Numbers 1, 2, 3, 4, and 5 are 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.005 M ZnCl2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.01 M ZnCl2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2, 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 + 0.01 M ZnCl2, and 2 M ZnSO4 + 0.03 M Zn(CF3SO3)2 + 0.01 M ZnCl2 electrolytes; (c) Cycling performance (1 A g−1) of hybrid anionic electrolyte; (d) Cycling performance (5 A g−1) of electrolyte with the same Zn2+ concentration; (e) Nyquist plot of hybrid anionic electrolyte in high-frequency region; (f) Nyquist plot of hybrid anionic electrolyte.
Figure 6. (a) Comparison of five CV curves (5 mV s−1); (b) Numbers 1, 2, 3, 4, and 5 are 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.005 M ZnCl2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.01 M ZnCl2, 2 M ZnSO4 + 0.01 M Zn(CF3SO3)2 + 0.02 M ZnCl2, 2 M ZnSO4 + 0.02 M Zn(CF3SO3)2 + 0.01 M ZnCl2, and 2 M ZnSO4 + 0.03 M Zn(CF3SO3)2 + 0.01 M ZnCl2 electrolytes; (c) Cycling performance (1 A g−1) of hybrid anionic electrolyte; (d) Cycling performance (5 A g−1) of electrolyte with the same Zn2+ concentration; (e) Nyquist plot of hybrid anionic electrolyte in high-frequency region; (f) Nyquist plot of hybrid anionic electrolyte.
Energies 16 00248 g006
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

Xie, B.; He, J.; Sun, Y.; Li, S.; Li, J. Hybrid Anionic Electrolytes for the High Performance of Aqueous Zinc-Ion Hybrid Supercapacitors. Energies 2023, 16, 248. https://doi.org/10.3390/en16010248

AMA Style

Xie B, He J, Sun Y, Li S, Li J. Hybrid Anionic Electrolytes for the High Performance of Aqueous Zinc-Ion Hybrid Supercapacitors. Energies. 2023; 16(1):248. https://doi.org/10.3390/en16010248

Chicago/Turabian Style

Xie, Bin, Junjie He, Yuchen Sun, Senlin Li, and Jing Li. 2023. "Hybrid Anionic Electrolytes for the High Performance of Aqueous Zinc-Ion Hybrid Supercapacitors" Energies 16, no. 1: 248. https://doi.org/10.3390/en16010248

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