Next Article in Journal
Electronic, Magnetic and Optical Properties of Double Perovskite Compounds: A First Principle Approach
Previous Article in Journal
Detail Analysis for Energy Transfer and Pigment Assembling in C−Phycocyanin through Time−Resolved Spectroscopy and AlphaFold2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Optimization of a Carbon Paper/MnO2 Composite Current Collector for Manufacturing a High-Performance Li–S Battery Cathode

1
Tianjin Key Laboratory for Photoelectric Materials and Devices, School of Materials Science and Engineering, Tianjin University of Technology, Tianjin 300384, China
2
State Forestry and Grassland Administration Key Laboratory of Silviculture in Downstream Areas of the Yellow River, School of Forestry, Shandong Agricultural University, Taian 271018, China
3
Feicheng Science and Technology Innovation Research Center, Feicheng 271600, China
*
Authors to whom correspondence should be addressed.
Crystals 2022, 12(11), 1596; https://doi.org/10.3390/cryst12111596
Submission received: 26 October 2022 / Revised: 7 November 2022 / Accepted: 7 November 2022 / Published: 9 November 2022
(This article belongs to the Section Materials for Energy Applications)

Abstract

:
High theoretical energy density endows lithium–sulfur batteries to be a promising candidate of the secondary batteries. Numerous studies have been implemented relying on exploring efficient host materials or separator modifying layers to solve the problematic shuttling and insufficient conversion of soluble polysulfides, whereas few studies have focused on the modification of the cathode collector. Herein, a high-performance sulfur cathode is manufactured with carbon paper/MnO2 as the cathode collector and liquid lithium polysulfides as the electrode material. The interface of carbon paper/MnO2 is proposed to afford fast electronic transport, strong chemical adsorption, and effective electrocatalysis to confine the diffusion of lithium polysulfides and facilitate their conversion during the charge/discharge process. More importantly, with no conductive additives and binders assisting, the gravimetric energy density of the sulfur cathode could be largely improved. Specifically, lithium–sulfur batteries using carbon paper/MnO2 as a cathode collector could stably circulate for 200 cycles at 0.2 C with a capacity of 664 mAh g−1, which is higher than that of carbon paper as a cathode collector (486 mAh g−1). This work may provide a new perspective to enhance the electrochemical performance of lithium–sulfur batteries by optimizing the cathode collector.

1. Introduction

Commercial lithium-ion batteries have brought revolutionary changes to society in the past decades [1,2], but they are involved in a dilemma of achieving the expected high-energy density by employing active materials with a limited capacity [3,4]. High-theoretical-energy-density lithium–sulfur (Li–S) batteries (2600 Wh kg−1) show the competitive potential by using economical, eco-friendly, and high-capacity sulfur as a cathode [5,6]. However, some complex problems, such as the insulative property of S and Li2S, the volumetric fluctuation during the reaction process, and the unrestrained lithium polysulfides (LiPS) [7,8], have impeded the electrochemical performance of Li–S batteries.
To solve the above issues, researchers have concentrated on optimizing the sulfur cathode to improve the electrochemical properties of lithium–sulfur batteries. The applicable hosts are vital to guarantee a fast electrochemical reaction [9,10]. Carbon materials, including graphene, carbon nanotubes, carbon fibers and porous carbon, are the common candidates due to their inherently benign conductivity [11,12]. However, the intrinsically weak interaction of carbon matrixes with LiPS is seemingly insufficient to alleviate the adverse results caused by the shuttle effect, except for the exquisite structural design and heteroatomic doping [13,14]. Polar compounds, such as metal oxides [15,16], metal nitrides [17,18], metal sulfides, and carbides [19,20], have been extensively studied and have been demonstrated to effectively confine the LiPS by chemical adsorption and even accelerate their conversion by catalysis [21,22]. However, metal compounds always possess an inferior conductivity in contrast with carbon, which may induce large cell resistance and limit electrochemical availability and stability [23,24]. Constructing the sulfur composites with the hybrid hosts involving conductive carbon and polar metal compounds could be a wise strategy for improving the electrochemical performance of the sulfur cathode [25,26]. However, to exert the electrochemical performance of sulfur cathodes with these traditional composite sulfur cathode materials, the common conductive additives, binders, and cathode collector are necessary to provide the required functions [27,28], which would increase the proportion of the electrochemical inactive materials and compromise the enhancement of energy density to some extent [29,30]. Therefore, a multifunctional hybrid cathode collector equipped with the characteristics of a cathode collector and host material is ideal. On one hand, the multifunctional cathode collector dispensing with the extra accessory components could be an effective solution to balance the electrochemical capacity and energy density of Li–S batteries [31,32,33,34]. On the other hand, the relevant optimization could provide fast electron transfer, and facilitate the electrochemical reaction near the triple-phase interface. Carbon paper has shown a certain competitive advantage in fabricating the high-loading and flexible electrode with an excellent conductive framework [35,36], but its weak anchoring ability on soluble LiPS may lead to obvious capacity decay. MnO2 has been proven to act as the chemical adsorbent and electrocatalysis for LiPS to improve the electrochemical performance of Li–S batteries, by forming the active mediators (thiosulfate and polythionate) [37,38,39], as well as the valid chemical restriction. Thus, the hybrid cathode collector based on carbon paper and MnO2 could simultaneously combine the conductivity and chemical interaction during the electrochemical process of the Li–S batteries without the extra conductive additives and binders, which could be conductive to fabricate high-performance Li–S batteries.
Herein, the carbon paper/MnO2 cathode collector is prepared by a one-step hydrothermal method and assembled into Li–S batteries without any auxiliary modification. Liquid LiPS are employed as the active sulfur sources. Carbon paper could provide the benign electron pathway and sufficient space for the electrochemical conversion, as well as the growth sites for MnO2. The in situ formed MnO2 could restrict the LiPS by chemical adsorption, accelerate their conversion by forming active thiosulfate and polythionate, and further enhance the electrochemical performance of Li–S batteries. Moreover, the simple cathode without certain traditional components (such as conductive additives, and binders) may contribute to the battery fabrication and energy density. The obtained Li–S batteries deliver the reversible capacity of 664 mAh g−1 at 0.2 C after 200 cycles and show a good rate of performance.

2. Materials and Methods

2.1. Synthesis of Carbon Paper/MnO2

Carbon paper was purchased from Toray (Tokyo, Japan, TGP-H-060). MnO2 particles were decorated on carbon paper (marked as CP) through the hydrothermal reaction, and the corresponding content was adjusted by adding H2SO4 with different usages, as shown in Figure 1 [40]. Typically, KMnO4 (0.0015 mol) was dissolved in a solution containing H2SO4 (0.1, 0.3, and 0.5 mL) and water (15 mL) with continuous stirring for 30 min. A piece of carbon paper (2 × 3 cm2) and the reaction solution were placed in the hydrothermal reactor (25 mL), followed by heating treatment at 120 °C for 3 h. The fetched carbon paper/MnO2 hosts were washed with water and ethanol several times, dried at 60 °C overnight, and cut into discs (Φ = 10 mm), which were marked as MC-1, MC-3, and MC-5, according to the usage of H2SO4.

2.2. Synthesis of Li2S6 Catholyte

Li2S6 catholyte (0.5 mol L−1) was synthesized through a reaction between Li2S (114.88 mg) and sulfur (400.81 mg) in 1,3-dioxolane (DOL)/dimethoxyethane (DME) solution (v/v = 1/1, 5 mL) at 50 °C for 24 h, followed by the continual stirring for 12 h after adding 1M lithium bis(triflfluoromethylsulfonyl) imide (LiTFSI, Sigma-Aldrich, Darmstadt) and 2 wt% LiNO3 [41]. The prepared Li2S6 catholyte could be employed as the active materials’ resources to replace the common elemental sulfur.

2.3. Assembly of Lithium–Sulfur Battery

The test Li–S cells (CR2032) were assembled in an argon-filled glove box by adding Li2S6 catholyte (0.5 M, 10 μL) in the cathode collectors, and using lithium foils (Φ = 14 mm) as reference electrodes, and microporous polypropylene film (Celgard 2400, Φ = 16 mm) as a separator with a moderate amount of electrolyte. The sulfur loading of the cathode was approximately 1.2 mg cm−2. The electrolyte was composed of 1 M LiTFSI and 2 wt% LiNO3 in DOL/DME (1:1 by volume).

2.4. Structural Characterization

Field emission scanning electron microscopy (FESEM, SU8010, Hitachi Company, Tokyo, Japan) was employed to obtain the morphology information and elemental distribution. X-ray diffraction (XRD, X’pert PRO, PANalytical B.V., Almelo, The Netherlands) with an angle resolution of ±0.0001, Raman spectroscopy (HORIBA EVOLUTION, Kyoto, Japan), and X-ray photoelectron spectroscopy (XPS, ESCALAB 250Xi, Thermo Scientific, Massachusetts) were utilized to acquire the structural information. Thermogravimetry (TG) was adopted to confirm the actual content of MnO2 from the ambient temperature to 900 °C in air.

2.5. Electrochemical Characterization

A cyclic voltammetry (CV) measurement on an electrochemical workstation (CHI660D, Shanghai Chenhua Instruments Ltd., Shanghai) was conducted among 1.7–2.8 V (vs. Li/Li+) at an ambient temperature with a scanning rate of 0.1 mV s−1. Capacity tests on LAND systems (CT-2001A, Wuhan LAND Electronic Co. Ltd., Wuhan, China) were carried out within the potential window of 1.7–2.8 V (vs. Li/Li+) at an ambient temperature and various current densities (1 C = 1675 mAh g−1). The specific capacity was related to the elemental sulfur.

3. Results and Discussion

3.1. Structure and Morphology

The fabrication process of the 3D binder-free cathode collector is depicted in Figure 2a. Flexible CP could serve as the conductive framework and growing substrates of MnO2 during the hydrothermal reactions with the typical fiber-jointed structure (Figure 2b). The aggregation state and microstructures of the formed MnO2 on the CP could be regulated by adjusting the usage of H2SO4. After hydrothermal treatment in the aqueous solution containing H2SO4 and KMnO4, the grey CP is partly decorated with the black MnO2. The optical images show the obvious reduction of MnO2 on the CP with the increase of H2SO4. The MC-1 sample is black with a nearly full coverage of MnO2 (Figure 2c). Fibrous carbon could be hardly observed, and the MnO2 clusters are assembled by the nanoplates. The sea urchin-like MnO2 with close packing is loaded on CP in MC-3 (Figure 2d). The MC-5 sample also possesses the urchin-like MnO2 with low content and uniform distribution (Figure 2e). By continually increasing the amount of H2SO4, the matrix could be obtained containing CP and MnO2, but its flexibility drops sharply corresponding to the fragile structure during the battery assembly, which may impair the electrochemical performance of Li–S batteries. Thus, adding different amounts of H2SO4 significantly affects the generated MnO2 on CP.
The XRD patterns in Figure 3a show that the pristine CP possesses a high graphitization with the typical characteristic peaks of graphite (PDF#41-1487), which may signify the benign electric conductivity and facilitate the electrochemical conversion [42]. The diffraction peaks of graphite are negligible in MC-1 with the appearance of α-MnO2 (PDF#44-0141) [43,44], mainly owing to the excessive coverage of MnO2 on CP. Then, the graphite peak at 26.4° could be detected in MC-3, demonstrating that the increasing H2SO4 avails to reduce the immoderate loading of MnO2 on the conductive skeleton. The MC-5 sample displays the more obvious peaks of graphite and MnO2, which may simultaneously endow the electric conductivity and chemical interaction in Li–S batteries. Meanwhile, the particle sizes of MnO2 in thickness direction could be calculated according to the Scherrer formula. The values for MC-1, MC-3 and MC-5 are 25.0 nm, 23.0 nm and 56.1 nm, respectively, based on the lattice plane of (110) in MC-1, (211) in MC-3 and MC-5. The actual content of MnO2 in the prepared samples is confirmed by thermogravimetry. As shown in Figure 3b, some impurities exist in the pristine CP with a mass ratio of 3.26%. The residual of MC-1, MC-3, and MC-5 after full combustion includes the impurities and oxidation products of MnO2 at a high temperature [45]. The detailed content of MnO2 in MC-1, MC-3, and MC-5 could be 14.06%, 8.81%, and 3.44%, respectively, manifesting the reduction of MnO2 along with the increase of H2SO4, which is consistent with the results of SEM and XRD.
Raman spectrum is adopted to further explore the effect of the loaded MnO2 on CP. The pristine CP has the typical D peak, G peak, and 2D peak (Figure 3c), which are located at 1355 cm−1, 1598 cm−1, and 2689 cm−1, respectively. The corresponding ratio of ID/IG is calculated to be 1.17, which signifies some existing defects in the CP [46]. In contrast, MC-5 possesses the D peak, G peak, and 2D peak at 1349 cm−1, 1582 cm−1, and 2717 cm−1, respectively, with an ID/IG ratio of 0.25. The significant decrease of ID/IG by decorating MnO2 on CP indicates the enhanced graphitization with fewer defects, which may contribute to the electron transfer and the electrochemical reaction [47]. Furthermore, a small characteristic peak of Mn-O bond occurs at ~327.5 cm−1 in the inset figure [48]. The XPS spectrum in Figure 3d confirms the loaded MnO2 on CP with the characteristic peak of Mn4+ at 642.6 eV and the related three satellite peaks at 643.8 eV, 644.8 eV and 645.4 eV [37], while some Mn3+ at 641.6 eV exist in the prepared MnO2, which may be the intermediate reaction products on the MnO2 surface [49].

3.2. Electrochemical Performance

Controlling the usage of H2SO4 could regulate the microstructure of the prepared cathode, and the corresponding impact on the cycle performance of the Li–S batteries is shown in Figure 4a. The battery using MC-5 possesses the superior electrochemical performance at 0.2 C in contrast to the batteries with MC-1 and MC-3. The capacity could stabilize at 704 mAh g−1 after 50 cycles, which is higher than those of the MC-1/S (442 mAh g−1) and MC-3/S (506.5 mAh g−1). The corresponding coulombic efficiency could stabilize at about 100%, showing the benign reversible redox conversion. Therefore, increasing the content of H2SO4 is inclined to fabricate the satisfying collector with moderate MnO2 loading, which may be derived from the balance of conductivity and chemical interaction [18,50]. Nanoplate-like MnO2 with excessive coverage on carbon paper may possess limited active sites and impede the charge transfer during the redox processes, while the sea urchin-like MnO2 with good crystallinity and modest coverage could guarantee the sufficient contact with LiPS, and the unhindered charge transport. The emphasis is then focused on the comparison between MC-5 and CP in Li–S batteries.
The cyclic voltammograms curves in Figure 4b show the typical redox pairs of the Li–S batteries [51], except for the weak reduction peaks at ~2.3 V (vs. Li/Li+) in the initial cycle, owing to the usage of liquid LiPS as sulfur resources. Meanwhile, the second reduction peaks of the CP/S during the CV tests are smaller than those of the MC-5/S, indicating the unsatisfactory anchoring of LiPS and the insufficient redox conversion from LiPS to Li2S2/Li2S [52]. Decorating MnO2 on CP could inhibit the shuttle of LiPS and promote their transformation.
The cycle performance of Li–S batteries at 0.2 C (Figure 5a) further demonstrates the advantage of MC-5. The Li–S battery with MC-5 possesses a higher capacity and better cyclic stability in contrast with the battery using CP. The capacity of MC-5/S could reach 606 mAh g−1 (vs. 411 mAh g−1) in the first cycle and be maintained at 664 mAh g−1 (vs. 486 mAh g−1) after 200 cycles with a stable coulombic efficiency near 100%. The corresponding charge/discharge curves (Figure 5b) show two reduction plateaus after the first cycle in consistence with the results of the CV tests, which are associated with the electrochemical conversion from solid sulfur to soluble LiPS and solid Li2S2/Li2S. The shorter voltage platforms in the CP/S battery with smaller capacities indicate the limited LiPS restriction and inadequate electrochemical reaction compared to MC-5/S. The pristine CP with good conductivity endows the Li–S battery inferior capacity, while the composite collector with decorated MnO2 on CP could improve the electrochemical conversion in the Li–S battery. Thus, the prepared cathode collector could balance the conductivity and chemical interaction and further enhance the reliability of the Li–S battery [50].
The rate property of the Li–S batteries in Figure 5c provides more evidence of promoting the electrochemical performance by adopting the hybrid collector. MC-5 endows the Li–S battery with high capacity and stability at relatively small current densities, such as 0.1 C, 0.2 C, and 0.5 C. The capacity at 1.0 C is fairly close and stabilizes at ~200 mAh g−1, which may be ascribed to the inherently sluggish reaction kinetics of LiPS, especially at large currents [53]. As the current switches back to 0.1 C, the capacity of the Li–S battery with MC-5 could reach 933 mAh g−1, which is higher than that of the Li–S battery using CP (~400 mAh g−1), indicating the enhanced rate of performance by employing MC-5 as a cathode collector. The coulombic efficiency in different C-rates maintains a high level, indicating the highly reversible electrochemical reactions. The related charge–discharge curves in Figure 5d also show the advantage of MC-5 in facilitating the electrochemical transformation at different C-rates, with the long voltage platforms. Carbon paper has been reported to directly load with sulfur as the cathode of Li–S batteries [54]. The corresponding initial capacity is 480 mAh g−1 and maintains only 150 mAh g−1 after 100 cycles. MnO2/acetylene black composites have been reported as the efficient host in Li/dissolved polysulfide batteries by Sivakumar and co-workers [55]. The related Li–S battery delivers a high initial capacity of 1540 mAh g−1 at 0.1 C but fast capacity decay occurs (649 mAh g−1 after 50 cycles). In this work, MnO2 was loaded on carbon paper to serve as the current collector for Li–S batteries. The electrochemical results demonstrate that the fabricated MC-5 collector could cycle for 200 cycles with considerable capacity and improve the performance of the Li–S battery.

3.3. Mechanism Analysis

The adsorption experiments in Figure 5a reveal that the anchoring capability of MC-5 on LiPS is more prominent compared to CP, with fast and obvious color fading. According to the UV–Vis spectra in Figure 6a, the characteristic absorption peak of Li2S6 at 350 nm [56] is significantly weakened by using MC-5, while the related peak of the Li2S6 solution containing CP is almost unchanged. This indicates that the MC-5 could interact with the lithium polysulfide in the solution and reduce the content of Li2S6 in the solution. The tested XPS spectrum of Mn after adsorption implies the interaction between MC-5 and Li2S6 (Figure 6b). The binding shift of Mn4+ from 642.6 eV to 641.9 eV proves the specific chemical interaction [57], while the emerging peaks of Mn3+ at 641.0 eV and Mn2+ at 640.5 eV and 639.8 eV illustrate the catalysis of MnO2 on LiPS by forming thiosulfate and polythionate as active mediators to accelerate the electrochemical processes [37,58]. Meanwhile, the obtained SEM images of the cathode and anode after 200 cycles provide more information (Figure 6c–f). The cycled cathode with CP becomes dense (Figure 6c), and the fibrous structure vanishes due to the repeated dissolution/deposition of S and Li2S, which may be detrimental to the electrochemical stability. The corresponding Li surface is covered with swellings and gullies (Figure 6d), showing the uneven deposition behavior and excessive side-reaction with LiPS. Instead, the cycled cathode with MC-5 (Figure 6e) still possesses the porous structure, and the Li anode has a flat morphology (Figure 6f). These phenomena show that loading moderate MnO2 on carbon paper could inhibit LiPS shuttle and guarantee the electrochemical and structural stability of the Li–S battery.
The published literatures have demonstrated that MnO2 loaded on graphene [37], CNTs [59], carbon fibers [60], and designed carbon with an exquisite structure [61] could significantly enhance the electrochemical performance of Li–S batteries, along with loading elemental sulfur into the hosts and employing conductive additives and polymer binders. Developing a high-performance current collector without other accessories (binders, conductive additives, metal collectors) could increase the proportion of the active materials to fabricate high-energy-density Li–S batteries. Meanwhile, different kinds of active materials, such as solid elemental sulfur, lithium sulfides, and liquid LiPS, are all appropriate for the 3D current collector and even could match the Li-free anode to enhance the safety of Li–S batteries.

4. Conclusions

In summary, the hybrid cathode collector containing carbon paper and MnO2 is fabricated by a simple hydrothermal method. The loaded MnO2 could be flexibly regulated by altering the usage of H2SO4. Increasing H2SO4 is inclined to generate urchin-like MnO2 with low content and uniform distribution. The composite collector is proven to alleviate the shuttling of soluble LiPS and improve the electrochemical performance of the Li–S battery by virtue of the moderate chemical adsorption and catalysis. The corresponding Li–S battery could deliver 664 mAh g−1 at 0.2 C after 200 cycles with good stability. This strategy of designing a multifunctional cathode collector for Li–S batteries without the auxiliary components (conductive additives, binders, and metal collector) is a feasible way to manufacture high-performance Li–S batteries.

Author Contributions

Conceptualization, Z.P., H.Z. and L.Z.; methodology, Z.P.; formal analysis, B.D., D.S., X.S. and Y.M.; writing—original draft preparation, Z.P. and L.K.; writing—review and editing, L.K., H.Z. and L.Z.; supervision, L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant No. 21905289) and Young Elite Scientists Sponsorship Program by Tianjin (Grant No. TJSQNTJ-2017-05).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yoshino, A. The birth of the lithium-ion battery. Angew. Chem. Int. Ed. 2012, 51, 5798–5800. [Google Scholar] [CrossRef] [PubMed]
  2. Larcher, D.; Tarascon, J.M. Towards greener and more sustainable batteries for electrical energy storage. Nature 2015, 7, 19–29. [Google Scholar] [CrossRef] [PubMed]
  3. Tarascon, J.M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef] [PubMed]
  4. Gao, X.P.; Yang, H.X. Multi-electron reaction materials for high energy density batteries. Energy Environ. Sci. 2010, 3, 174–189. [Google Scholar] [CrossRef]
  5. Bruce, P.G.; Freunberger, S.A.; Hardwick, L.J.; Tarascon, J.M. Li–O2 and Li–S batteries with high energy storage. Nat. Mater. 2011, 11, 19–29. [Google Scholar] [CrossRef]
  6. Ji, X.; Lee, K.T.; Nazar, L.F. A highly ordered nanostructured carbon–sulphur cathode for lithium–sulphur batteries. Nat. Mater. 2009, 8, 500–506. [Google Scholar] [CrossRef]
  7. Manthiram, A.; Chung, S.H.; Zu, C.X. Lithium–sulfur batteries: Progress and prospects. Adv. Mater. 2015, 27, 1980–2006. [Google Scholar] [CrossRef]
  8. Yin, Y.X.; Xin, S.; Guo, Y.G.; Wan, L.J. Lithium–sulfur batteries: Electrochemistry, materials, and prospects. Angew. Chem. Int. Ed. 2013, 52, 13186–13200. [Google Scholar] [CrossRef]
  9. Wang, L.; Wang, Z.Y.; Wu, J.F.; Li, G.R.; Liu, S.; Gao, X.P. To effectively drive the conversion of sulfur with electroactive niobium tungsten oxide microspheres for lithium–sulfur battery. Nano Energy 2020, 77, 105173. [Google Scholar] [CrossRef]
  10. McNulty, D.; Landgraf, V.; Trabesomger, S. The importance of sulfur host structural preservation for lithium–sulfur battery performance. J. Mater. Chem. A 2020, 8, 26085–26097. [Google Scholar] [CrossRef]
  11. Manthiram, A.; Fu, Y.Z.; Chung, S.H.; Zu, C.X.; Su, Y.S. Rechargeable lithium–sulfur batteries. Chem. Rev. 2014, 114, 11751–11787. [Google Scholar] [CrossRef]
  12. She, Z.W.; Sun, Y.M.; Zhang, Q.F.; Cui, Y. Designing high-energy lithium–sulfur battery. Chem. Soc. Rev. 2016, 45, 5605–5634. [Google Scholar]
  13. Ci, H.N.; Wang, M.L.; Sun, Z.T.; Wei, C.H.; Cai, J.S.; Lu, C.; Cui, G.; Liu, Z.F.; Sun, J.Y. Direct insight into sulfiphilicity-lithiophilicity design of bifunctional heteroatom-doped graphene mediator toward durable Li–S batteries. J. Energy Chem. 2022, 66, 474–482. [Google Scholar] [CrossRef]
  14. Laverde, J.; Rosero-Navarro, N.C.; Miura, A.; Buitrago-Sierra, R.; Tadanaga, K.; Lópes, D. Impact of sulfur infiltration time and its content in an N-doped mesoporous carbon for application in Li–S batteries. Batteries 2022, 8, 58. [Google Scholar] [CrossRef]
  15. Liang, X.; Kwok, C.Y.; Lodi-Marzano, F.; Pang, Q.; Cuisinier, M.; Huang, H.; Hart, C.J.; Houtarde, D.; Kaup, K.; Sommer, H.; et al. Tuning transition metal oxide-sulfur interactions for long life lithium sulfur batteries: The “goldilocks” principle. Adv. Energy Mater. 2016, 6, 501636. [Google Scholar] [CrossRef]
  16. Tao, X.Y.; Wang, J.G.; Liu, C.; Wang, H.T.; Yao, H.B.; Zheng, G.Y.; She, Z.W.; Cai, Q.X.; Li, W.Y.; Zhou, G.M.; et al. Balancing surface adsorption and diffusion of lithium-polysulfides on nonconductive oxides for lithium–sulfur battery design. Nat. Commun. 2016, 7, 11203. [Google Scholar] [CrossRef] [Green Version]
  17. Deng, D.R.; Xue, F.; Jia, Y.J.; Ye, J.C.; Bai, C.D.; Zheng, M.S.; Dong, Q.F. Co4N nanosheet assembled mesoporous sphere as a matrix for ultrahigh sulfur content lithium–sulfur batteries. ACS Nano 2017, 11, 6031–6039. [Google Scholar] [CrossRef]
  18. Ma, L.L.; Yu, L.J.; Liu, J.C.; Su, Y.Q.; Li, S.; Zang, X.H.; Meng, T.; Zhang, S.H.; Song, J.J.; Wang, J.Y.; et al. Construction of Ti4O7/TiN/carbon microdisk sulfur host with strong polar N-Ti-O bond for ultralong life lithium–sulfur battery. Energy Storage Mater. 2022, 44, 180–188. [Google Scholar] [CrossRef]
  19. He, J.R.; Hartmann, G.; Lee, M.; Hwang, G.S.; Chen, Y.; Manthiram, A. Freestanding 1T MoS2/Graphene heterostructure as a highly efficient electrocatalyst for lithium polysulfides in Li–S batteries. Energy Environ. Sci. 2019, 12, 344–350. [Google Scholar] [CrossRef]
  20. Zhang, C.F.; Ma, Y.L.; Zhang, X.T.; Abdolhosseinzadeh, S.; Sheng, H.W.; Lan, W.; Pakdel, A.; Heier, J.; Nüesch, F. Two-dimensional transition metal carbides and nitrides (MXenes): Synthesis, properties, and electrochemical energy storage applications. Energy Environ. Mater. 2020, 3, 29–55. [Google Scholar] [CrossRef]
  21. Ng, S.F.; Lau, M.Y.L.; Ong, W.J. Lithium–sulfur battery cathode design: Tailoring metal-based nanostructures for robust polysulfide adsorption and catalytic conversion. Adv. Mater. 2021, 33, 2008654. [Google Scholar] [CrossRef] [PubMed]
  22. Zhuang, Z.H.; Kang, Q.; Wang, D.S.; Li, Y.D. Single-atom catalysis enables long-life, high-energy lithium–sulfur batteries. Nano Res. 2020, 13, 1856–1866. [Google Scholar] [CrossRef]
  23. Zhang, H.B.; Yu, L.; Chen, T.; Zhou, W.; Lou, X.W. Surface modulation of hierarchical MoS2 nanosheets by Ni single atoms for enhanced electrocatalytic hydrogen evolution. Adv. Funct. Mater. 2018, 28, 1807086. [Google Scholar] [CrossRef]
  24. Tang, Y.Q.; Huang, Y.; Luo, L.; Fan, D.L.; Lu, Y.C.; Manthiram, A. Self-supported MoO2/MoS2 nano-sheets embedded in a carbon cloth as a binder-free substrate for high-energy lithium–sulfur batteries. Electrochim. Acta 2021, 367, 137482. [Google Scholar] [CrossRef]
  25. Zhang, H.; Ono, L.K.; Tong, G.Q.; Liu, Y.Q.; Qi, Y.B. Long-life lithium–sulfur batteries with high areal capacity based on coaxial CNT@TiN-TiO2 sponge. Nat. Commun. 2021, 12, 4738. [Google Scholar] [CrossRef] [PubMed]
  26. He, B.; Li, W.C.; Zhang, Y.; Yu, X.F.; Zhang, B.S.; Li, F.; Lu, A.H. Paragenesis BN/CNTs hybrid as a monoclinic sulfur host for high rate and ultra-long life lithium–sulfur battery. J. Mater. Chem. A 2018, 6, 24194–24200. [Google Scholar] [CrossRef]
  27. Yuan, J.J.; Huang, Z.; Song, Y.Z.; Li, M.Y.; Li, H.Y. In-situ crosslinked binder for high-stability S cathodes with greatly enhanced conduction and polysulfides anchoring. Chem. Eng. J. 2021, 426, 128705. [Google Scholar] [CrossRef]
  28. Li, M.; Zhang, Y.N.; Bai, Z.Y.; Liu, W.W.; Liu, T.C.; Gim, J.; Jiang, G.P.; Yuan, Y.F.; Luo, D.; Yassar, R.S.; et al. A lithium–sulfur battery using a 2D current collector architecture with a large-sized sulfur host operated under high areal loading and low E/S ratio. Adv. Mater. 2018, 30, 1804271. [Google Scholar] [CrossRef]
  29. Xiao, P.T.; Sun, L.X.; Liao, D.K.; Agboola, P.; Imran, S.; Xu, Y. Unexpected effect of electrode architecture on high-performance lithium–sulfur batteries. ACS Appl. Mater. Interfaces 2018, 10, 33269–33275. [Google Scholar] [CrossRef]
  30. Yang, B.Z.; Xie, P.; Luo, Q.; Li, Z.W.; Zhou, C.Y.; Yin, Y.H.; Liu, X.B.; Li, Y.S. Binder-free ω-Li3V2O5 catalytic network with multi-polarization center assists lithium–sulfur batteries for enhanced kinetics behavior. Adv. Funct. Mater. 2022, 32, 2110665. [Google Scholar] [CrossRef]
  31. Chung, S.H.; Chang, C.H.; Manthiram, A. Hierarchical sulfur electrodes as a testing platform for understanding the high-loading capability of Li–S batteries. J. Power Sources 2016, 334, 179–190. [Google Scholar] [CrossRef]
  32. Yen, Y.J.; Chung, S.H. Lean-electrolyte lithium–sulfur electrochemical cells with high-loading carbon nanotube/nanofiber–polysulfide cathodes. Chem. Commun. 2021, 57, 2009–2012. [Google Scholar] [CrossRef]
  33. Yu, C.H.; Yen, Y.J.; Chung, S.H. Nanoporosity of carbon–sulfur nanocomposites toward the lithium–sulfur battery electrochemistry. Nanomaterials 2021, 11, 1518. [Google Scholar] [CrossRef]
  34. Chen, S.Y.; Chung, S.H. Advanced current collectors with carbon nanofoams for electrochemically stable lithium–sulfur cells. Nanomaterials 2021, 11, 2083. [Google Scholar] [CrossRef]
  35. Meng, L.; Li, Y.; Lin, Q.X.; Long, J.; Wang, Y.; Hu, J. Nitrogen and Oxygen dual self-doped flexible PPTA nanofiber carbon paper as an effective interlayer for lithium–sulfur batteries. ACS Appl. Energy Mater. 2021, 4, 8592–8603. [Google Scholar] [CrossRef]
  36. Chen, Y.Y.; Xu, P.; Liu, Q.B.; Yuan, D.; Zhu, S. Cobalt embedded in porous carbon fiber membranes for high-performance lithium–sulfur batteries. Carbon 2022, 187, 187–195. [Google Scholar] [CrossRef]
  37. Liang, X.; Hart, C.; Pang, Q.; Garsuch, A.; Weiss, T.; Nazar, L.F. A highly efficient polysulfide mediator for lithium–sulfur batteries. Nat. Commun. 2015, 6, 5682. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, X.; Yang, L.W.; Wang, Y.; Li, Q.; Chen, C.T.; Zhong, B.H.; Chen, Y.X.; Guo, X.D.; Wu, Z.G.; Liu, Y.; et al. Novel functional separator with self-assembled MnO2 layer via a simple and fast method in lithium–sulfur battery. J. Colloid Interface Sci. 2022, 606, 666–676. [Google Scholar] [CrossRef]
  39. Chen, Z.H.; Hu, Y.X.; Liu, W.; Yu, F.; Yu, X.F.; Mei, T.; Yu, L.; Wang, X.B. Three-dimensional engineering of sulfur/MnO2 composites for high-rate lithium–sulfur batteries. ACS Appl. Mater. Interfaces 2021, 13, 38394–38404. [Google Scholar] [CrossRef]
  40. Pang, Z.Y.; Ma, Y.; Zhou, Y.; Miao, X.Y.; Kong, L.L.; Song, D.W.; Shi, X.X.; Zhang, H.Z.; Zhang, L.Q. Tailoring 3D carbon foam using CNTs and MnO2 to fabricate stable lithium/dissolved lithium polysulfide batteries. Langmuir 2021, 37, 4016–4024. [Google Scholar] [CrossRef]
  41. Yoshie, Y.; Hori, K.; Mae, T.; Noda, S. High-energy density Li–S battery with positive electrode of lithium polysulfides held by carbon nanotube sponge. Carbon 2021, 182, 32–41. [Google Scholar] [CrossRef]
  42. Li, Z.Q.; Lu, C.J.; Xia, Z.P.; Zhou, Y.; Luo, Z. X-ray diffraction patterns of graphite and turbostratic carbon. Carbon 2007, 45, 1686–1695. [Google Scholar] [CrossRef]
  43. Pan, H.L.; Shao, Y.Y.; Yan, P.F.; Cheng, Y.W.; Han, K.S.; Nie, Z.M.; Wang, C.M.; Yang, J.H.; Li, X.L.; Bhattacharya, P.; et al. Reversible aqueous zinc/manganese oxide energy storage from conversion reactions. Nat. Energy 2016, 1, 16039. [Google Scholar] [CrossRef]
  44. Khan, Z.; Park, S.; Hwang, S.M.; Yang, J.; Ko, H. Hierarchical urchin-shaped α-MnO2 on graphene-coated carbon microfibers: A binder-free electrode for rechargeable aqueous Na–air battery. NPG Asia Mater. 2016, 8, 294. [Google Scholar] [CrossRef]
  45. Hu, J.; Wang, Z.; Fu, Y.; Lyu, L.; Lu, Z.; Zhou, L. In situ assembly of MnO2 nanosheets on sulfur-embedded multichannel carbon nanofiber composites as cathode for lithium–sulfur batteries. Sci. China Mater. 2020, 63, 728–738. [Google Scholar] [CrossRef] [Green Version]
  46. Ferrari, A.C.; Meyer, J.C.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselvo, K.S.; Roth, S.; et al. Raman spectrum of graphene and graphene layers. Phys. Rev. Lett. 2006, 97, 187401. [Google Scholar] [CrossRef] [Green Version]
  47. Binder, C.; Bendo, T.; Hammes, G.; Neves, G.; Binder, R.; Mello, J.D.; Klein, A.N. Structure and properties of in situ-generated two-dimensional turbostratic graphite nodules. Carbon 2017, 124, 685–692. [Google Scholar] [CrossRef]
  48. Cheng, S.; Yang, L.F.; Chen, D.C.; Ji, X.; Jiang, Z.J.; Ding, D.; Liu, M.L. Phase evolution of an alpha MnO2-based electrode for pseudo-capacitors probed by in operando Raman spectroscopy. Nano Energy 2014, 9, 161–167. [Google Scholar] [CrossRef]
  49. Nesbitt, H.W.; Banerjee, D. Interpretation of XPS Mn(2p) spectra of Mn oxyhydroxides and constraints on the mechanism of MnO2 precipitation. Am. Mineral. 1998, 83, 305–315. [Google Scholar] [CrossRef]
  50. Wang, L.; Li, G.R.; Liu, S.; Gao, X.P. Hollow molybdate microspheres as catalytic hosts for enhancing the electrochemical performance of sulfur cathode under high sulfur loading and lean electrolyte. Adv. Funct. Mater. 2021, 31, 2010693. [Google Scholar] [CrossRef]
  51. Zhao, C.X.; Li, X.Y.; Zhao, M.; Chen, Z.X.; Song, Y.W.; Chen, W.J.; Liu, J.N.; Wang, B.; Zhang, X.Q.; Chen, C.M.; et al. Semi-immobilized molecular electrocatalysts for high-performance lithium–sulfur batteries. J. Am. Chem. Soc. 2021, 143, 19865–19872. [Google Scholar] [CrossRef]
  52. Sun, Z.H.; Zhang, J.Q.; Yin, L.C.; Hu, G.J.; Fang, R.P.; Cheng, H.M.; Li, F. Conductive porous vanadium nitride/graphene composite as chemical anchor of polysulfides for lithium–sulfur batteries. Nat. Commun. 2017, 8, 14627. [Google Scholar] [CrossRef] [Green Version]
  53. He, J.R.; Bhargav, A.; Manthiram, A. High-energy-density, long-life lithium–sulfur batteries with practically necessary parameters enabled by low-cost Fe–Ni nanoalloy catalysts. ACS Nano 2021, 15, 8583–8591. [Google Scholar] [CrossRef]
  54. Wang, D.H.; Xie, D.; Yang, T.; Zhong, Y.; Wang, X.L.; Xia, X.H.; Gu, C.D.; Tu, J.P. Conversion from Li2SO4 to Li2S@C on carbon paper matrix: A novel integrated cathode for lithium-sulfur batteries. J. Power Sources 2016, 331, 475–480. [Google Scholar] [CrossRef]
  55. Radhika, G.; Subadevi, R.; Sivakumar, M. Sulfur nested with mixture of MnO2/AB composite as efficient host for high-performance Li–S batteries. J. Chem. Sci. 2020, 132, 53. [Google Scholar] [CrossRef]
  56. Zou, Q.L.; Lu, Y.C. Solvent-dictated lithium sulfur redox reactions: An Operando UV−vis spectroscopic study. J. Phys. Chem. Lett. 2016, 7, 1518–1525. [Google Scholar] [CrossRef]
  57. Rehman, S.; Tang, T.Y.; Ali, Z.S.; Huang, X.X.; Hou, Y.L. Integrated design of MnO2@carbon hollow nanoboxes to synergistically encapsulate polysulfides for empowering lithium sulfur batteries. Small 2017, 13, 1700087. [Google Scholar] [CrossRef]
  58. Cao, K.Z.; Liu, H.Q.; Li, Y.; Wang, Y.J.; Jiao, L.F. Encapsulating sulfur in δ-MnO2 at room temperature for Li–S battery cathode. Energy Storage Mater. 2017, 9, 78–84. [Google Scholar] [CrossRef]
  59. Wang, N.; Peng, S.K.; Cheng, X.; Wang, J.X.; Wang, C.; Qi, X.; Dai, S.L.; Yan, S.J. Construction of ultrathin MnO2 decorated graphene/carbon nanotube nanocomposites as efficient sulfur hosts for high-performance lithium–sulfur batteries. RSC Adv. 2019, 9, 6346–6355. [Google Scholar] [CrossRef] [Green Version]
  60. Li, Z.; Zhang, J.T.; Lou, X.W. Hollow carbon nanofibers filled with MnO2 nanosheets as efficient sulfur hosts for lithium–sulfur batteries. Angew. Chem. Int. Ed. 2015, 54, 12886–12890. [Google Scholar] [CrossRef]
  61. Chen, H.; Dong, W.D.; Xia, F.J.; Zhang, Y.J.; Yan, M.; Song, J.P.; Zou, W.; Liu, Y.; Hu, Z.Y.; Liu, J.; et al. Hollow nitrogen-doped carbon/sulfur@MnO2 nanocomposite with structural and chemical dual-encapsulation for lithium–sulfur battery. Chem. Eng. J. 2020, 381, 122746. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of preparing the composite cathode collector and Li–S batteries.
Figure 1. Schematic illustration of preparing the composite cathode collector and Li–S batteries.
Crystals 12 01596 g001
Figure 2. Schematic illustration of the fabrication of carbon paper/MnO2 (a), optical (in the green outline), and SEM images of CP (b), MC-1 (c), MC-3 (d), and MC-5 (e).
Figure 2. Schematic illustration of the fabrication of carbon paper/MnO2 (a), optical (in the green outline), and SEM images of CP (b), MC-1 (c), MC-3 (d), and MC-5 (e).
Crystals 12 01596 g002
Figure 3. X-ray diffraction patterns (a,b) and TG curves of CP, MC-1, MC-3, and MC-5. Raman spectra of CP and MC-5 (c). (d) Mn 2p3/2 spectra of the pristine MC-5.
Figure 3. X-ray diffraction patterns (a,b) and TG curves of CP, MC-1, MC-3, and MC-5. Raman spectra of CP and MC-5 (c). (d) Mn 2p3/2 spectra of the pristine MC-5.
Crystals 12 01596 g003
Figure 4. (a) Cycle performance of the Li–S batteries using MC-1, MC-3, and MC-5. Cyclic voltammograms of the Li–S batteries with CP and MC-5 as collectors (b).
Figure 4. (a) Cycle performance of the Li–S batteries using MC-1, MC-3, and MC-5. Cyclic voltammograms of the Li–S batteries with CP and MC-5 as collectors (b).
Crystals 12 01596 g004
Figure 5. Cycle performance (a) and rate property (c) of the CP and MC-5 as collectors in Li–S batteries with the corresponding charge/discharge profiles (b,d).
Figure 5. Cycle performance (a) and rate property (c) of the CP and MC-5 as collectors in Li–S batteries with the corresponding charge/discharge profiles (b,d).
Crystals 12 01596 g005
Figure 6. (a) Adsorption experiments of CP and MC-5 in Li2S6 solution including the digital photos and UV–Vis spectra, and (b) Mn 2p3/2 spectra of MC-5 after the adsorption experiment. SEM images of the cycled sulfur cathodes (c,e) and lithium anodes (d,f) from the Li–S batteries by using CP (c,d) and (e,f) MC-5 as collectors.
Figure 6. (a) Adsorption experiments of CP and MC-5 in Li2S6 solution including the digital photos and UV–Vis spectra, and (b) Mn 2p3/2 spectra of MC-5 after the adsorption experiment. SEM images of the cycled sulfur cathodes (c,e) and lithium anodes (d,f) from the Li–S batteries by using CP (c,d) and (e,f) MC-5 as collectors.
Crystals 12 01596 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pang, Z.; Kong, L.; Zhang, H.; Deng, B.; Song, D.; Shi, X.; Ma, Y.; Zhang, L. The Optimization of a Carbon Paper/MnO2 Composite Current Collector for Manufacturing a High-Performance Li–S Battery Cathode. Crystals 2022, 12, 1596. https://doi.org/10.3390/cryst12111596

AMA Style

Pang Z, Kong L, Zhang H, Deng B, Song D, Shi X, Ma Y, Zhang L. The Optimization of a Carbon Paper/MnO2 Composite Current Collector for Manufacturing a High-Performance Li–S Battery Cathode. Crystals. 2022; 12(11):1596. https://doi.org/10.3390/cryst12111596

Chicago/Turabian Style

Pang, Zhiyuan, Linglong Kong, Hongzhou Zhang, Bin Deng, Dawei Song, Xixi Shi, Yue Ma, and Lianqi Zhang. 2022. "The Optimization of a Carbon Paper/MnO2 Composite Current Collector for Manufacturing a High-Performance Li–S Battery Cathode" Crystals 12, no. 11: 1596. https://doi.org/10.3390/cryst12111596

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