Next Article in Journal
Implementation of Distributed Autonomous Control Based Battery Energy Storage System for Frequency Regulation
Next Article in Special Issue
Concept of a Dual Energy Storage System for Sustainable Energy Supply of Automated Guided Vehicles
Previous Article in Journal
Modeling, Control System Design and Preliminary Experimental Verification of a Hybrid Power Unit Suitable for Multirotor UAVs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Pot Synthesis of Bismuth Sulfide Nanostructures as an Active Electrode Material for Aqueous Hybrid Capacitors

Department of Process Engineering and Technology of Polymer and Carbon Materials, Faculty of Chemistry, Wrocław University of Science and Technology, Gdańska 7/9, 50–344 Wrocław, Poland
*
Author to whom correspondence should be addressed.
Submission received: 8 April 2021 / Revised: 26 April 2021 / Accepted: 30 April 2021 / Published: 6 May 2021
(This article belongs to the Special Issue Hybrid Energy Storage Based on Battery and Ultracapacitor)

Abstract

:
The high theoretical capacity of Bi2S3 shows high promise as a negative electrode material for energy storage devices. Herein, we investigate a facile, one-step chemical precipitation method using common organic solvents, such as acetone, ethanol, and isopropanol, for the synthesis of Bi2S3 nanostructures. The nanospherical Bi2S3 from acetone (Bi2S3-A) presents the most balanced electrochemical properties, exhibiting a high specific capacity of 181 mAh g−1 at 1 A g−1 and decent rate capability. Additionally, Bi2S3-A is used as a negative electrode in an aqueous hybrid system with an activated carbon positive electrode, demonstrating a capacitance of 86 F g−1, a specific energy of 7.6 Wh kg−1, and an initial capacity retention of 74% after 1000 cycles.

1. Introduction

The development of energy storage systems is currently flourishing due to the growing demand for systems that can deliver high amounts of energy. Lithium-ion batteries, electrochemical capacitors, supercapacitors, and their combination as supercapatteries are used in many portable devices, and it is necessary to develop these devices to deliver high power/energy densities in a relatively short period of time. Among the many materials that are used as electrode materials in the discussed devices, the most known are activated carbons (ACs) [1,2,3]. Other carbon materials, such as carbon nanofibers and nanotubes or graphene and reduced graphene oxide, have been increasingly considered for energy storage devices [4,5,6]. However, the capacitance of carbon materials is limited by the electric double layer (EDL) due to the specific surface area, which is a crucial factor [7,8,9]. In addition to carbon materials, materials with pseudocapacitive properties as transition metal oxides/hydroxides have been intensively considered for application in pseudocapacitors due to their fast redox reactions and high specific capacitance. Transition metal oxides and/or hydroxides are characterized by high resistance values, which are undesirable in energy storage devices [10,11]. To manage the current requirements, innovative materials with high specific capacitance and low resistance have been increasingly studied. A novel group of electrode materials are transition metal sulfides, which combine the most important features for being competitive in comparison to carbon materials and metal oxides/hydroxides. Transition metal sulfides are characterized by extremely high specific capacitance values and semiconducting properties. Among various transition metal sulfides, such as CuS, CoS, Ni2S3, MnS, and MoS2, Bi2S3 seems to be the most attractive material for energy storage and conversion devices [12,13,14,15,16,17]. The extraordinary features of bismuth sulfide include its unique electric properties resulting from its semiconductor behavior with a direct band gap of 1.3 eV and a high capacitance value; thus, Bi2S3 shows tremendous promise. Recently, a large amount of effort has been exerted to explore a new concept of electrochemical capacitors called hybrid capacitors based on electrodes with EDL or pseudocapacitive properties that are combined with transition metal compounds with battery-like responses [18,19,20]. Zong et al. [21] combined defect-rich bismuth sulfides with nitrogen-doped carbon nanofibers, and this material exhibited an improved electrochemical performance in a 6 M KOH electrolyte. Another study reported by Liang et al. [22] described an asymmetric hybrid capacitor based on a positive electrode of NiCo2S4@Ni(OH)2@PPy core-shell structures and a negative electrode of activated carbon, achieving a high energy density of 34.6 Wh kg−1 at a power density of 120 W kg−1. Current works indicate that the design and fabrication of hybrid capacitors show promising results; however, the synthesis of battery-like electrode materials using a facile approach remains a very large challenge.
Currently, various morphologies of Bi2S3 nanostructures, including nanorods, nanoribbons, hexagonal plates, lamellar platelets, spherical nanoparticles, or their mixtures, can be synthesized in a controlled manner [23]. Notably, the synthesis method can affect the morphology of Bi2S3 [24], while the simplicity of Bi2S3 synthesis plays an important role from an application point of view. In this paper, we propose using various solvents such as isopropanol, ethanol, acetone, and water for a one-pot synthesis of Bi2S3 that can be differentiated by its morphology, crystalline structure, and electrochemical performance. Furthermore, we assemble an aqueous hybrid capacitor with battery-like Bi2S3 and EDL AC electrodes and investigate its electrochemical performance, demonstrating the decent specific energy and cycling stability of the device.

2. Materials and Methods

2.1. Materials Synthesis

Bismuth sulfides were chemically precipitated at 75 °C [25]. The appropriate amount of bismuth nitrate hexahydrate (Bi(NO3)3∙6H2O) was dissolved in 200 mL of ethanol, isopropanol, acetone, and distilled water. The solvents were heated to a set temperature of 75 °C under mechanical stirring. Subsequently, the aqueous solution of ammonium sulfide ((NH4)2S) was slowly dropped until the dark brown/black colored solution appeared. The as-prepared bismuth sulfides were washed and centrifugated five times. Finally, the resulting materials were dried under vacuum at 60 °C for 24 h. The Bi2S3 samples were labelled according to the solvent used: ethanol (E), isopropanol (I), acetone (A), and distilled water (W): Bi2S3-E, Bi2S3-I, Bi2S3-A, and Bi2S3-W, respectively.

2.2. Characterization

Synthesized materials have been characterized by their porous structure by nitrogen gas sorption analysis (Quantachrome), the crystalline structure was determined using X-ray diffraction (XRD, Ultima IV Rigaku) measurements and chemical composition of the samples were analyzed using X-ray photoelectron spectroscopy (XPS, PHI 5000 VersaProbe). The morphologies of the bismuth sulfides were observed via field-emission scanning electron microscope (FESEM, Merlin Zeiss). Detailed information on the materials characterization, including porous structure parameters, XRD, FESEM, and XPS setup and analysis conditions are described in Supplementary Materials.

2.3. Electrochemical Measurements

Electrochemical performance of the electrodes’ were determined in a three-electrode setup (Swagelok™, France) in a 6 M KOH electrolyte using BioLogic VSP potentiostat-galvanostat with an Hg|HgO reference electrode and pitch-based activated carbon counter electrode. The electrochemical measurements included cyclic voltammetry (CV) and galvanostatic charge-discharge (GCD) experiments at different current densities and electrochemical impedance spectroscopy (EIS). The detailed procedure for the electrochemical characterization is included in Supplementary Materials.
The specific capacity was expressed in mAh per mass of active material in one electrode. The specific capacity values (Qs/mAh g−1) were calculated from the cyclic voltammetry curves and galvanostatic discharge profiles using Equations (1) and (2), respectively.
Q s = I d V 3.6 v m e l
Q s = I t 3.6 m e l
where I is the current (A), Δt is the discharge time (s), ν is the scan rate (V s−1), and mel is the mass of the active material in the electrode (g).
For the two-electrode Swagelok™ cell, a hybrid device with a Bi2S3 negatrode and pitch-based AC positrode (Bi2S3-A/AC) was assembled at a negative electrode:positive electrode mass ratio of 1:2 (4–5 and 9–10 mg, respectively). The gravimetric cell capacitance (Ccell) was calculated using Equation (3) [26]:
C c e l l = 2 I V d t M   ×   ( V i 2 V f 2 )  
where I is the discharge current (A), ∫Vdt is the integrated area under the galvanostatic discharge curve, M is the total mass of the electrodes (g), and ViVf is the operating discharge voltage from initial to final potential values. Additionally, electrochemical impedance spectroscopy (EIS) measurements in the frequency range of 400 kHz to 10 mHz at an AC amplitude of 5 mV were performed for the assembled device along with cycling stability tests at a current density of 1 A g−1.
The specific energy (E in Wh kg−1) and specific power delivered (P in kW kg−1) by a hybrid capacitor were calculated using Equations (4) and (5), respectively:
E = I V d t M × 3.6
P = 3.6 × E t

3. Results and Discussion

The porous structure of the synthesized Bi2S3 using various solvents was recognized by N2 sorption at 77 K. The mesoporous character of all bismuth sulfides are corroborated by the hysteresis loops (Figure 1). The BET surface area of obtained samples was relatively low, as suspected for pristine inorganic materials. The highest BET value of 32 m2 g−1 was noticed for Bi2S3-E, while the lowest value of 17 m2 g−1 for Bi2S3-I. The structural parameters of the obtained samples are included in Table 1. Moreover, among the four discussed samples, two groups with comparable porous structure parameters can be distinguished. The first group consists of the Bi2S3 samples synthesized using acetone and ethanol as solvents, while the second group consists of bismuth sulfides obtained in isopropanol and water environments. Notably, within both groups, it is clearly seen that the porous texture of Bi2S3 is more developed when the synthesis is carried out using acetone and ethanol as solvents. The VT of the samples ranges from 0.080 cm3 g−1 to 0.121 cm3 g−1 for Bi2S3-W and Bi2S3-A, respectively. Some differences are observed in the case of Vmes, which varies from 0.067 cm3 g−1 for Bi2S3-I to 0.088 cm3 g−1 for Bi2S3-A.
The XRD patterns (Figure 2) present different crystalline structures depending on the synthesis solvent. An amorphous phase is recorded for Bi2S3-A and Bi2S3-E. The XRD pattern of Bi2S3-I reveals the appearance of some characteristic of a primitive orthorhombic crystalline structure of Bi2S3 (JCPDS 17-0320). Results can suggest that when isopropanol is applied as the synthesis solvent, crystalline structure is not fully formed and isopropanol led to intermediate form between crystalline and amorphous phase. Moreover, the crystalline phase of Bi2S3 is clearly observed when water is the synthesis environment, and the diffraction peaks correlate with the Bi2S3 orthorhombic phase. The diffraction planes are assigned in agreement with the standard diffraction pattern of Bi2S3 (JCPDS 65–2435). The different crystalline structures of the obtained bismuth sulfides indicate that the kind of solvent strictly affects the crystalline structure. In addition to the influence of the solvent type, the formation of crystalline structure of bismuth sulfide can be also achieved by the Bi2S3 treatment at a growing temperature [27]. No impurities or Bi2O3 phases are observed in the materials synthesized via the proposed precipitation method with ammonium sulfide.
FESEM was employed to compare and contrast the surface morphologies of the bismuth sulfides synthesized using various solvents. The observations reveal discernible differences in the Bi2S3 morphologies (Figure 3). The use of distinct solvents leads to Bi2S3 with different particle shapes. Coral-like particles of Bi2S3 are obtained when ethanol or acetone is used as the solvent. However, the Bi2S3-E particles are significantly smaller than the Bi2S3-A particles and do not exceed 100 nm, whereas the Bi2S3-A particles can have a diameter of 250 nm. Moreover, the size of particles can affect the porous structure of Bi2S3. It was observed that the size of bismuth sulfide particles increases in the sequence Bi2S3-E < Bi2S3-A < Bi2S3-W < Bi2S3-I, which is in accordance with the decreasing BET surface area of bismuth sulfides.
To further analyze the surface chemical composition and oxidation state of the Bi2S3 samples synthesized in different solvents, XPS was performed, and the results are shown in Figure 4. The XPS survey spectra presented in Figure 4a explicitly reveal the presence of only Bi and S, suggesting the purity of the obtained bismuth sulfides. The Bi and S peak positions agree well with those in other papers [23,24]. The high-resolution Bi 4f and S 2p spectra of the obtained bismuth sulfides are shown in Figure 4b–e. The two strong peaks at binding energies of 158.9 and 164.2 eV correspond to the Bi 4f7/2 and Bi 4f5/2 in Bi2S3, respectively [28]. The bismuth sulfides synthesized using ethanol, isopropanol, and acetone as solvents are characterized by the nearly symmetrical shape of their peaks, which is attributed to the pure phase of the Bi2S3 surface. In the case of bismuth sulfide obtained in water, two extra signals, leading to some distortions on the shoulders of the Bi 4f7/2 and Bi 4f5/2 peaks, are observed. This feature can be attributed to the heterogeneous composition on the surface of Bi2S3. Moreover, the two peaks with lower intensity located between Bi 4f7/2 and Bi 4f5/2 can be ascribed to S 2p3/2 (161.1 eV) and S 2p1/2 (162.2 eV), confirming the valence state of sulfur as S2− [29].
The results of the CV measurements, presented in Figure 5, show distinguishable redox peaks related to the faradaic reactions of the Bi2S3 nanostructures. At a low scan rate, where the current response is diffusion controlled, two pairs of anodic/cathodic peaks appear [30]. The first pair is attributed to the reaction of Bi2S3 with OH- ions, and this pair originates from the change in the bismuth valence state (−0.6 and −0.45 V, Figure 5a) [23]. Moreover, Bi2S3 prepared from isopropanol presents a peak shift towards more negative values compared with the rest of the synthesized materials. The second pair is related to the adsorption of hydrogen in the lamellar structure of Bi2S3 (−0.85 V), while the anodic peak at −0.3 V is related to hydrogen oxidation during the electrochemical measurements. [25,31] As the scan rate increases to 100 mV s−1 (Figure 5b,c), the cathodic peaks become very wide, and a shift of both the anodic and cathodic redox peaks is observed, indicating the change in the internal resistance of the electrode and the quasi-reversible nature of the faradaic reaction. The faradaic reaction of bismuth sulfide is described according to the following Equation:
Bi2S3 + H2O +e ⇄ (Hads + Bi2S3) + OH
In accordance with the CV results, the GCD discharge curves of the Bi2S3 electrodes (Figure 6a) show a distinct plateau at a potential of approximately −0.6 V, indicating the battery-like behavior of the charge storage mechanism. At a current density of 0.5 A g−1, a smaller plateau, which is related to hydrogen adsorption and is observable in the CV measurements at a potential of −0.85 V, does not appear due to diffusion limitations [31]. At a higher current density of 10 A g−1, changes in the charging curve occur in the Bi2S3 electrodes as the slope becomes more capacitive in nature. The longest charge–discharge time among all materials at 0.5 A g−1 is observed for Bi2S3-I; however, under increasing current regimes, its charge storage capability decreases compared with other materials (Figure 6b). These results indicate that the formed crystalline structure and hexagonal nanoplatelet morphology lead to the highest specific capacity values, but in terms of rate capability, the more amorphous structure and spherical-like particles of Bi2S3-A are more favorable traits. These observations can be related to the degree of crystallinity of the material; for instance, more amorphous/enriched defects in the structure of a material exhibiting Nernstian features allows for the distribution of redox reactions over a wider potential range, thereby translating to improved electrochemical performance at higher scan rates [32].
The relationships between the specific capacity and scan rate are presented in Figure 7a, while the specific capacity vs. discharge current density relationship is shown in Figure 7b. The highest specific capacity at a scan rate of 1 mV s−1 is achieved by Bi2S3-I, with a value of 247 mAh g−1, while the lowest value of 169 mAh g−1 is recorded for Bi2S3-W. From the galvanostatic charge–discharge measurements, Bi2S3-I exhibits a specific capacity of 204 mAh g−1 at a current density of 0.5 A g−1, followed by Bi2S3-E, Bi2S3-A and Bi2S3-W, which present specific capacities of 189, 187, and 135 mAh g−1, respectively. The fully crystalline structure of Bi2S3-W and its nanorod-like morphology does not provide enough active sites for redox reactions; thus, its charge storage capability is the lowest among all materials. Moreover, bismuth sulfide prepared in acetone and water presents the best rate capability among all samples, maintaining 25 and 29% of the initial capacity at a high current density of 20 A g−1, respectively. Bi2S3-I exhibits the highest recorded capacity, while Bi2S3-A is a more rational choice for electrochemical applications, as its charge storage properties do not decrease as fast as the material obtained from isopropanol. Among other reported bismuth sulfides, Bi2S3 synthesized from organic solvents such as acetone, isopropanol, and ethanol exhibits higher specific capacities due to its stable electrochemical performance over a wide working potential window (Table 2).
For further evaluation of the application potential of bismuth sulfides, the material with the most balanced specific capacity and rate capability was selected (Bi2S3-A) as a negatrode for an asymmetric device with an activated carbon positrode. The assembled hybrid capacitor with a battery-type oxidation-reduction reaction and AC with an electrochemical double layer mechanism was investigated in an aqueous 6 M KOH electrolyte. As shown in Figure 8a, the CV curves present an irregular shape with visible peaks, which can be attributed to the combination of the faradaic and non-faradaic capacitance from the transition metal sulfide and carbon electrodes. The quasi-triangular shape of the GCD profile further confirms the capacitive nature of the hybrid capacitor (Figure 8b), indicating its reversible charge–discharge capability [35]. Furthermore, both the negatrode and positrode potential ranges are balanced while the cell is operating, with a clear distinction between the EDL-type GCD curve of the carbon electrode and the battery-like plot of Bi2S3-A. The assembled hybrid capacitor exhibits a specific capacitance of 104 F g−1 at scan rates of 1 mV s−1 and 86 F g−1 and a current density of 0.2 A g−1, as calculated from the CV and GCD measurements, respectively (Figure 8c). The rate capability at higher scan rates and current regimes is approximately 20–33% due to diffusion-controlled processes at the negatrode [18].
The EIS measurement results of the hybrid capacitor are presented in the form of Nyquist plots, Figure 8d. The small semicircle in the high-frequency region is attributed to the charge transfer resistance (Rct), which is related to the faradaic reactions of Bi2S3 [23]. The Rct value of the Bi2S3/AC capacitor is 0.25 Ω, implying fast charge kinetics. In the low-frequency region, the more inclined the plot is towards the -Z’’ axis, the lower the resistance attributed to the diffusion of ions at the electrode/electrolyte interface, which for the hybrid device shows certain distortions in the ion movement at both the electrode/electrolyte interfaces; thus, this resistance may explain the capacitive performance drop at higher current densities [36]. The investigated asymmetric system presents an equivalent circuit of the solution resistance and two parallel ladders, one for each electrode [37] (Figure 8d). The fitting of the impedance spectrum presents components such as charge transfer resistance, Warburg impedance related with diffusion of ions and constant phase element (CPE) attributed to the double layer capacitance at the electrodes interface and faradaic reactions of bismuth sulfide [38]. The long-term performance of the hybrid capacitor is an important factor in the evaluation of its applicability for energy storage devices (Figure 9a). The initial drop of 35% is related to the formation of a surface passivation layer and the side reactions related to bismuth sulfide [38]; however, after the 125th cycle, the assembled Bi2S3-A/AC capacitor shows greatly improved stability, retaining 74% of the initial capacitance after 1000 charge-discharge cycles.
The energy storage capability of the Bi2S3-A/AC hybrid capacitor was evaluated in the form of a Ragone plot (Figure 9b). The assembled device can reach a maximum specific energy density of 7.6 Wh kg−1 at a specific power density of 0.09 kW kg−1, which is in the medium range/slightly improved compared with the previously reported polyaniline-co-pyrrole/thermally reduced graphene oxide composite capacitor [39], Bi2S3/CoNi-LDH hybrid capacitor or polyimidazole/CuS/CNT composite [40], and on pair with commercially available devices [41]. However, due to the decreasing capacitive performance at higher current densities, the Bi2S3-A/AC hybrid capacitor shows approximately 1 Wh kg−1 of specific energy at a high power density of 5 kW kg−1.

4. Conclusions

In summary, we showed that the one-pot synthesis of bismuth sulfide nanostructures is a promising and economical approach for preparing negative electrodes for hybrid aqueous capacitors. The amorphous coral-like structure of Bi2S3 synthesized from an acetone solution provided active sites for faradaic reactions at the electrode/electrolyte interface, yielding high capacity values and reasonable rate capability. Moreover, a hybrid aqueous capacitor assembled with Bi2S3 as a negatrode and activated carbon as a positrode exhibited a specific energy density of 7.6 Wh kg−1 at a power density of 0.9 W kg−1. The current work provides directions for the successful design of Bi2S3 nanostructures to explore the use of high-performance anode materials in potential energy-related applications such as hybrid supercapacitors or aqueous lithium or sodium batteries.

Supplementary Materials

Author Contributions

Conceptualization, A.M. (Agata Moyseowicz) and A.M. (Adam Moyseowicz); methodology, A.M. (Agata Moyseowicz) and A.M. (Adam Moyseowicz); formal analysis, A.M. (Agata Moyseowicz); investigation, A.M. (Agata Moyseowicz) and A.M. (Adam Moyseowicz); resources, A.M. (Agata Moyseowicz); writing—original draft preparation, A.M. (Agata Moyseowicz) and A.M. (Adam Moyseowicz); writing—review and editing, A.M. (Agata Moyseowicz); visualization, A.M. (Agata Moyseowicz) and A.M. (Adam Moyseowicz); supervision, A.M. (Agata Moyseowicz); funding acquisition, A.M. (Agata Moyseowicz). All authors have read and agreed to the published version of the manuscript.

Funding

The research leading to the above results has received funding from the National Science Centre (Poland) under the grant agreement 2016/21/D/ST5/01642.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

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

References

  1. Conway, B.E. Electrochemical Supercapacitors: Scientific Fundamentals and Technological Applications; Springer: New York, NY, USA, 1999. [Google Scholar]
  2. Pandolfo, A.G.; Hollenkamp, A.F. Carbon properties and their role in supercapacitors. J. Power Sources 2006, 157, 11–27. [Google Scholar] [CrossRef]
  3. Kierzek, K.; Frąckowiak, E.; Lota, G.; Gryglewicz, G.; Machnikowski, J. Electrochemical capacitors based on highly porous carbons prepared by KOH activation. Electrochim. Acta 2004, 49, 515–523. [Google Scholar] [CrossRef]
  4. Ko, T.-H.; Hung, K.-H.; Tzeng, S.-S.; Shen, J.-W.; Hung, C.-H. Carbon nanofibers grown on activated carbon fiber fabrics as electrode of supercapacitors. Phys. Scr. 2007, 129, 80–84. [Google Scholar] [CrossRef]
  5. Śliwak, A.; Grzyb, B.; Ćwikła, J.; Gryglewicz, G. Influence of wet oxidation of herringbone carbon nanofibers on the pseudocapacitance effect. Carbon 2013, 64, 324–333. [Google Scholar] [CrossRef]
  6. Grzyb, B.; Gryglewicz, S.; Śliwak, A.; Diez, N.; Machnikowski, J.; Gryglewicz, G. Guanidine, amitrole and imidazole as nitrogen dopants for the synthesis of N-graphenes. RSC Adv. 2016, 6, 15782. [Google Scholar] [CrossRef]
  7. Śliwak, A.; Diez, N.; Miniach, E.; Gryglewicz, G. Nitrogen-containing chitosan-based carbon as an electrode material for high-performance supercapacitors. J. Appl. Electrochem. 2016, 46, 667–677. [Google Scholar] [CrossRef] [Green Version]
  8. Raymundo-Pinero, E.; Kierzek, K.; Machnikowski, J.; Beguin, F. Relationship between the nanoporous texture of activated carbons and their capacitance properties in different electrolytes. Carbon 2006, 44, 2498–2507. [Google Scholar] [CrossRef]
  9. Mirzaein, M.; Abbas, Q.; Hunt, M.R.C.; Hall, P. Pseudocapacitive effect of carbons doped with different functional groups as electrode materials for electrochemical capacitors. Energies 2020, 13, 5577. [Google Scholar] [CrossRef]
  10. Brousse, T.; Taberna, P.L.; Crosnier, O.; Dugas, R.; Guillement, P.; Scudeller, Y.; Zhou, Y.; Favier, F.; Bélanger, D.; Simon, P. Long-term cycling behavior of asymmetric activated carbon/MnO2 aqueous electrochemical supercapacitor. J. Power Sources 2007, 173, 633–641. [Google Scholar] [CrossRef]
  11. Śliwak, A.; Gryglewicz, G. High-voltage asymmetric supercapacitors based on carbon and manganese oxide/oxidized carbon nanofiber composite electrodes. Energy Technol. 2014, 2, 819. [Google Scholar] [CrossRef]
  12. Zhang, Z.; Huang, Z.; Ren, L.; Shen, Y.; Qi, X.; Zhong, J. One-pot synthesis of hierarchically nanostructured Ni3S2 dendrites as active materials for supercapacitors. Electrochim. Acta 2014, 149, 316–323. [Google Scholar] [CrossRef]
  13. Wei, W.; Mi, L.; Gao, Y.; Zheng, Z.; Chen, W.; Guan, X. Partial ion-exchange of nickel-sulfide-derived electrodes for high performance supercapacitors. Chem. Mater. 2014, 26, 3418–3426. [Google Scholar] [CrossRef]
  14. Huang, K.-J.; Zhang, J.Z.; Shi, G.W.; Liu, Y.-M. One-step hydrothermal synthesis of two-dimensional cobalt sulfide for high-performance supercapacitors. Mater. Lett. 2014, 131, 45–48. [Google Scholar] [CrossRef]
  15. Tie, J.; Han, J.; Diao, G.; Liu, J.; Xie, Z.; Cheng, G.; Sun, M.; Yu, L. Controllable synthesis of hierarchical nickel cobalt sulfide with enhanced electrochemical activity. Appl. Surf. Sci. 2018, 435, 187–194. [Google Scholar] [CrossRef]
  16. Yang, H.; Xie, J.; Li, C.M. Bi2S3 nanorods modified with Co(OH)2 ultrathin nanosheets to significantly improve its pseudocapacitance for high specific capacitance. RSC Adv. 2014, 4, 48666–48670. [Google Scholar] [CrossRef]
  17. Durga, I.K.; Rao, S.S.; Reddy, A.E.; Gopi, C.V.V.M.; Kim, H.-J. Achieving copper sulfide leaf like nanostructure electrode for high performance supercapacitor and quantum-dot sensitized solar cells. Appl. Surf. Sci. 2018, 435, 666–675. [Google Scholar] [CrossRef]
  18. Yu, X.; Zhou, J.; Li, Q.; Zhao, W.-N.; Zhao, S.; Chen, H.; Tao, K.; Han, L. Bi2S3 nanorod-stacked hollow microtubes self-assembled from bismuth-based metal–organic frameworks as advanced negative electrodes for hybrid supercapacitors. Dalton Trans. 2019, 48, 9057–9061. [Google Scholar] [CrossRef] [PubMed]
  19. Ock, I.W.; Choi, J.Q.; Jeong, H.M.; Kang, J.K. Synthesis of pseudocapacitive polymer chain anode and subnanoscale metal oxide cathode for aqueous hybrid capacitors enabling high energy and power densities along with long cycle life. Adv. Energy Mater. 2018, 8, 1702895. [Google Scholar] [CrossRef]
  20. Ding, J.; Hu, W.; Paek, W.; Mitlin, D. Review of hybrid ion capacitors: From aqueous to lithium to sodium. Chem. Rev. 2018, 118, 6457–6498. [Google Scholar] [CrossRef] [PubMed]
  21. Zong, W.; Lai, F.; He, G.; Feng, J.; Wang, W.; Lian, R.; Miao, Y.-E.; Wang, G.-C.; Parkin, I.P.; Liu, T. Sulfur-deficient bismuth sulfide/nitrogen-doped carbon nanofibers as advanced free-standing electrode for asymmetric supercapacitors. Small 2018, 14, 1801562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Liang, M.; Zhao, M.; Wang, H.; Shen, J.; Song, X. Enhanced cycling stability of hierarchical NiCo2S4@Ni(OH)2@PPy core–shell nanotube arrays for aqueous asymmetric supercapacitors. J. Mater. Chem. A 2018, 6, 2482–2493. [Google Scholar] [CrossRef]
  23. Moyseowicz, A.; Moyseowicz, A. Tailoring the morphology, crystalline structure, and electrochemical properties of nanostructured Bi2S3 using various solvent mixtures. Mater. Renew. Sustain. Energy 2020, 9, 11. [Google Scholar] [CrossRef]
  24. Miniach, E.; Gryglewicz, G. Solvent-controlled morphology of bismuth sulfide for supercapacitor applications. J. Mater. Sci. 2018, 53, 16511–16523. [Google Scholar] [CrossRef] [Green Version]
  25. Moyseowicz, A. Scalable one-pot synthesis of bismuth sulfide nanorods as an electrode active material for energy storage applications. J. Solid State Electrochem. 2019, 23, 1191–1199. [Google Scholar] [CrossRef] [Green Version]
  26. Laheäär, A.; Przygocki, P.; Abbas, Q.; Béguin, F. Appropriate methods for evaluating the efficiency and capacitive behavior of different types of supercapacitors. Electrochem. Commun. 2015, 60, 21–25. [Google Scholar] [CrossRef]
  27. Lu, J.; Wang, Z.; Zhang, Y.; Zhou, X. Hydrothermal synthesis of Bi2S3 nanorods from a single-source precursor and their promotional effect on the photocatalysis of TiO2. J. Nanomater. 2013, 125409. [Google Scholar] [CrossRef]
  28. Li, W.H. Synthesis and characterization of bismuth sulfide nanowires through microwave solvothermal technique. Mater. Lett. 2008, 62, 243–245. [Google Scholar] [CrossRef]
  29. Lou, W.; Chen, M.; Wang, X.; Liu, W. Novel single-source precursors approach to prepare highly uniform Bi2S3 and Sb2S3 nanorods via a solvothermal treatment. Chem. Mater. 2007, 19, 872–878. [Google Scholar] [CrossRef]
  30. Wang, B.; Aoki, K.J.; Chen, J.; Nishiumi, T. Slow scan voltammetry for diffusion-controlled currents in sodium alginate solutions. J. Electroanal. Chem. 2013, 70, 60–64. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, B.; Ye, X.; Hou, W.; Zhao, Y.; Xie, Y. Biomolecule-assisted synthesis and electrochemical hydrogen storage of Bi2S3 flowerlike patterns with well-aligned nanorods. J. Phys. Chem. B 2006, 110, 8978–8985. [Google Scholar] [CrossRef]
  32. Eftekhari, A.; Mohamedi, M. Tailoring pseudocapacitive materials from a mechanistic perspective. Mater. Today Energy 2017, 6, 211–229. [Google Scholar] [CrossRef]
  33. Yue, H.; Chen, S.; Li, P.; Zhu, C.; Yang, X.; Li, T.; Gao, Y. Lemongrass-like Bi2S3 as a high-performance anode material for lithium-ion batteries. Ionics 2019, 25, 3587–3592. [Google Scholar] [CrossRef]
  34. Duca, M.; Guerrini, E.; Colombo, A.; Trasatti, S. Activation of nickel for hydrogen evolution by spontaneous deposition of iridium. Electrocatal 2013, 4, 338–345. [Google Scholar] [CrossRef]
  35. Zhai, X.; Gao, J.; Xu, X.; Hong, W.; Wang, H.; Wu, F.; Liu, Y. 3D interconnected Bi2S3 nanosheets network directly grown on nickel foam as advanced performance binder-free electrode for hybrid asymmetric supercapacitor. J. Power Sources 2018, 396, 648–658. [Google Scholar] [CrossRef]
  36. Ramasamy, K.; Gupta, R.K.; Sims, H.; Palchoudhury, S.; Ivanov, S.; Gupta, A. Layered ternary sulfide CuSbS2 nanoplates for flexible solid-state supercapacitors. J. Mater. Chem. A 2015, 3, 13263–13274. [Google Scholar] [CrossRef] [Green Version]
  37. Fletcher, S.; Black, V.J.; Kirkpatrick, I. A universal equivalent circuit for carbon-based supercapacitors. J. Solid State Electrochem. 2014, 18, 1377–1387. [Google Scholar] [CrossRef] [Green Version]
  38. Chai, W.; Yin, W.; Wang, K.; Ye, W.; Li, X.; Tang, B.; Rui, Y. Bismuth sulfide–integrated carbon derived from organic ligands as a superior anode for sodium storage. Energy Technol. 2019, 7, 1900668. [Google Scholar] [CrossRef]
  39. Moyseowicz, A.; González, Z.; Menéndez, R.; Gryglewicz, G. Three-dimensional poly(aniline-co-pyrrole)/thermally reduced graphene oxide composite as a binder-free electrode for high-performance supercapacitors. Compos. B Eng. 2018, 145, 232–239. [Google Scholar] [CrossRef]
  40. Ravi, S.; Gopi, C.V.V.M.; Kim, H.J. Enhanced electrochemical capacitance of polyimidazole coated covellite CuS dispersed CNT composite materials for application in supercapacitors. Dalton Trans. 2016, 45, 12362–12371. [Google Scholar] [CrossRef]
  41. Yassine, M.; Fabris, D. Performance of commercially available supercapacitors. Energies 2017, 10, 1340. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Adsorption/desorption isotherms at 77 K for the bismuth sulfides.
Figure 1. Adsorption/desorption isotherms at 77 K for the bismuth sulfides.
Energies 14 02670 g001
Figure 2. XRD patterns of the synthesized materials.
Figure 2. XRD patterns of the synthesized materials.
Energies 14 02670 g002
Figure 3. FESEM images of the bismuth sulfides synthesized in ethanol (a), acetone (b), isopropanol (c), and distilled water (d).
Figure 3. FESEM images of the bismuth sulfides synthesized in ethanol (a), acetone (b), isopropanol (c), and distilled water (d).
Energies 14 02670 g003
Figure 4. XPS survey spectra (a), Bi 4f and S 2p regions of Bi2S3 (upper graph and lower graph, indicated within yellow color range, respectively), regions of Bi2S3 obtained in various solvents (be).
Figure 4. XPS survey spectra (a), Bi 4f and S 2p regions of Bi2S3 (upper graph and lower graph, indicated within yellow color range, respectively), regions of Bi2S3 obtained in various solvents (be).
Energies 14 02670 g004
Figure 5. Cyclic voltammograms of the bismuth sulfides synthesized in different solvents at 1, 10, and 100 mV s−1 (ac).
Figure 5. Cyclic voltammograms of the bismuth sulfides synthesized in different solvents at 1, 10, and 100 mV s−1 (ac).
Energies 14 02670 g005
Figure 6. Charge–discharge profiles of the bismuth sulfides recorded at 0.5 A g−1 (a) and 10 A g−1 (b).
Figure 6. Charge–discharge profiles of the bismuth sulfides recorded at 0.5 A g−1 (a) and 10 A g−1 (b).
Energies 14 02670 g006
Figure 7. Specific capacity versus scan rate (a) and discharge current density (b) of the bismuth sulfides.
Figure 7. Specific capacity versus scan rate (a) and discharge current density (b) of the bismuth sulfides.
Energies 14 02670 g007
Figure 8. CV curves at the different scan rates (a), GCD curves at a current density of 0.2 A g−1 (b), cell capacitance versus scan rate/discharge current density (c), and the Nyquist plot with an inset representing the high-frequency range and the equivalent circuit diagram (d) of the hybrid capacitor based on the Bi2S3-A and AC electrodes.
Figure 8. CV curves at the different scan rates (a), GCD curves at a current density of 0.2 A g−1 (b), cell capacitance versus scan rate/discharge current density (c), and the Nyquist plot with an inset representing the high-frequency range and the equivalent circuit diagram (d) of the hybrid capacitor based on the Bi2S3-A and AC electrodes.
Energies 14 02670 g008
Figure 9. Cycling stability at a current density of 1 A g−1 (a) and the Ragone plot of the specific energy–power density relationship of the hybrid capacitor and other various hybrid devices (b).
Figure 9. Cycling stability at a current density of 1 A g−1 (a) and the Ragone plot of the specific energy–power density relationship of the hybrid capacitor and other various hybrid devices (b).
Energies 14 02670 g009
Table 1. Obtained porous structure parameters based on the N2 isotherms for the bismuth sulfides synthesized with different solvents.
Table 1. Obtained porous structure parameters based on the N2 isotherms for the bismuth sulfides synthesized with different solvents.
SymbolSBET
m2 g−1
VT
cm3 g−1
Vmes
cm3 g−1
VDR
cm3 g−1
Bi2S3-A290.1210.0880.011
Bi2S3-E320.1040.0850.012
Bi2S3-I170.0830.0670.006
Bi2S3-W180.0800.0720.007
Table 2. Comparison of the electrochemical properties of the Bi2S3-based materials synthesized under different conditions.
Table 2. Comparison of the electrochemical properties of the Bi2S3-based materials synthesized under different conditions.
Sample NameSynthesis MethodMorphologyElectrolyteSpecific
Capacity *
Cycling
Stability
Ref.
Bi2S3Self-sacrificial template methodNanorod-stacked hollow microtubes1 M KOH148 mAh g−1 at 1 A g−191% after 3000 cycles[18]
dr-Bi2S3/S-NCNF **Hydrothermal methodNanofibers6 M KOH175 mAh g−1 at 1 A g−181% after 1000 cycles[21]
Bi2S3–75 °CChemical precipitationNanorods6 M KOH68 mAh g−1 at 1 A g−152% after 1000 cycles[25]
Bi2S3Hydrothermal methodNanoflowers6 M KOH40 mAh g−1 at 1 A g−1-[33]
Bi2S3Solvothermal methodNanorods1 M Na2SO453 mAh  g−1 at 1 A g−1-[34]
Bi2S3-AChemical precipitationCoral-like nanospheres6 M KOH181 mAh g−1 at 1 A g−174% after 1000 cyclesThis work
Bi2S3-IChemical precipitationHexagonal nanoplatelets6 M KOH193 mAh g−1 at 1 A g−1-This work
* Specific capacity was calculated according to the respective capacitance and operating voltage of given materials. ** dr-Bi2S3/S-NCNF—defect-rich bismuth sulfide/nitrogen-doped carbon nanofibers.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Moyseowicz, A.; Moyseowicz, A. One-Pot Synthesis of Bismuth Sulfide Nanostructures as an Active Electrode Material for Aqueous Hybrid Capacitors. Energies 2021, 14, 2670. https://0-doi-org.brum.beds.ac.uk/10.3390/en14092670

AMA Style

Moyseowicz A, Moyseowicz A. One-Pot Synthesis of Bismuth Sulfide Nanostructures as an Active Electrode Material for Aqueous Hybrid Capacitors. Energies. 2021; 14(9):2670. https://0-doi-org.brum.beds.ac.uk/10.3390/en14092670

Chicago/Turabian Style

Moyseowicz, Adam, and Agata Moyseowicz. 2021. "One-Pot Synthesis of Bismuth Sulfide Nanostructures as an Active Electrode Material for Aqueous Hybrid Capacitors" Energies 14, no. 9: 2670. https://0-doi-org.brum.beds.ac.uk/10.3390/en14092670

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