Next Article in Journal
Luteolin 4′-Neohesperidoside Inhibits Clinically Isolated Resistant Bacteria In Vitro and In Vivo
Next Article in Special Issue
A Multifunctional Coating on Sulfur-Containing Carbon-Based Anode for High-Performance Sodium-Ion Batteries
Previous Article in Journal
Adsorption and Removal of Cr6+, Cu2+, Pb2+, and Zn2+ from Aqueous Solution by Magnetic Nano-Chitosan
Previous Article in Special Issue
Silicon/Graphite/Amorphous Carbon as Anode Materials for Lithium Secondary Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interlayer-Expanded MoS2 Enabled by Sandwiched Monolayer Carbon for High Performance Potassium Storage

1
Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
2
University of Chinese Academy of Sciences, Beijing 101400, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2608; https://doi.org/10.3390/molecules28062608
Submission received: 17 February 2023 / Revised: 5 March 2023 / Accepted: 6 March 2023 / Published: 13 March 2023
(This article belongs to the Special Issue Frontier in Lithium-Ion Battery)

Abstract

:
Potassium-ion batteries (PIBs) have aroused a large amount of interest recently due to the plentiful potassium resource, which may show cost benefits over lithium-ion batteries (LIBs). However, the huge volume expansion induced by the intercalation of large-sized potassium ions and the intrinsic sluggish kinetics of the anode hamper the application of PIBs. Herein, by rational design, nano-roses assembled from petals with a MoS2/monolayer carbon (C-MoS2) sandwiched structure were successfully synthesized. The interlayer distance of ultrathin C-MoS2 was expanded from original MoS2 of 6.2 to 9.6 Å due to the formation of the MoS2-carbon inter overlapped superstructure. This unique structure efficiently alleviates the mechanical strain, prevents the aggregation of MoS2, creates more active sites, facilitates electron transport, and enhances the specific capacity and K+ diffusion kinetics. As a result, the prepared C-MoS2-1 anode delivers a high reversible specific capacity (437 mAh g−1 at 0.1 A g−1) and satisfying rate performance (123 mAh g−1 at 6.4 A g−1). Therefore, this work provides new insights into the design of high-performance anode materials of PIBs.

Graphical Abstract

1. Introduction

The development of post lithium-ion batteries (LIBs) is important to deal with the increasing energy demands because of the scarcity of lithium resources. Potassium-ion batteries (PIBs) and sodium-ion batteries (SIBs) have recently aroused a large amount of interest in large-scale energy storage applications due to the abundance of potassium and sodium, which is expected to be more cost effective [1]. Compared with Na/Na+ (−2.73 V vs. standard hydrogen electrode (SHE)), the potential of K/K+ (−2.93 V vs. SHE) is close to that of Li/Li+ (−3.04 V vs. SHE), resulting in the increase in the operating voltage, which provides an opportunity to achieve the high energy density of PIBs. The smaller Stokes radius of K+ (3.6 Å) results from the weaker Lewis acidity than in the cases of Li+ (4.8 Å) and Na+ (4.6 Å) enables K+ to have a higher transport mobility in the electrolyte, which indicates that PIBs possess the potential for higher power density [2]. In addition, although it is possible potassium dendrites to occur that are similar to lithium dendrites in PIBs under the overcharged state, the lower melting point of potassium metal (63.4 °C) than that of lithium metal (180.5 °C) will lead to dendritic potassium, which would melt readily before creating a short-circuit, leading to a much better safety capability of PIBs. Furthermore, aluminum (Al) and potassium do not form alloy compounds, thus, Al foil can be used as the current collector in PIBs to replace copper (Cu) foil, which reduces both the weight as well as the economic cost of the electrode [3]. Therefore, these merits make PIBs one of the most promising alternatives to LIBs, particularly for large-scale energy storage applications.
Up to now, a variety of anode materials for PIBs have drawn special attention including carbonaceous materials (graphite, soft carbon, hard carbon, graphenes, carbon nanotubes) [4,5,6,7], alloy types materials (bismuth, stannum) [8,9], and conversion type materials (chalcogenide compounds) [10,11]. Up until now, graphite-based anodes have been extensively reported because of the low-costs, safety, and low voltage plateau for PIBs [12]. In spite of the above advantages, the theoretical capacity of graphite is only 279 mAh g−1, and the sluggish diffusion kinetics of K+ in graphite hinder the rate capability for PIBs [13]. Consideration of the fast potassiation/depotassiation kinetics and relatively high charge/discharge capacities of transition metal dichalcogenides (TMDs), TMDs have become a class of promising alternatives [14]. As a typical TMD, MoS2 has a two-dimensional (2D) layered structure in which sulfur atoms are interconnected with molybdenum atoms to form hexagons within the layers and weak van der Waals forces act between the layers. Due to the anisotropic structure for MoS2, permeable channels and substantial amounts of reaction sites facilitate the insertion/extraction of potassium ions. Ren et al. first studied bare MoS2 as a host material for K+ electrochemical intercalation, and the MoS2 delivered a reversible capacity of 65 mAh g−1 at 20 mA g−1 after 200 cycles [15]. The inferior cycle stability of MoS2 for PIBs is due to limited interlayer distance between the adjacent MoS2 monolayer and MoS2 tends to aggregate when undergoing volume change during charge/discharge cycles. Moreover, the intrinsically low electronic conductivity of MoS2 leads an to unsatisfactory rate performance because of the large polarization [16]. Thereby, new approaches and structures are necessary when seeking to further the improvement of MoS2 electrodes.
Recently, strategies of enlarging the interlayer distance of adjacent MoS2 monolayers, fabricating few-layer MoS2, and decorating MoS2 with carbon materials have been reported to improve the K+ storage performance. For example, Wang et al. reported few-layered 2H MoS2 nanosheets with an expansion of the interlayer spacing by liquid-phase exfoliation in water as the PIB anode, which delivered a high capacity of 203 mAh g−1 at 200 mA g−1 after 300 cycles and a low capacity decay of 0.02% per cycle over 1500 cycles [17]. Hu et al. synthesized a yolk-shell structured hollow porous carbon-sphere-confined MoS2 composite (MoS2@HPCS) as a PIB anode. The high reversible capacity of 254 mAh g−1 was obtained at 500 mA g−1 after 100 cycles and maintained 126 mAh g−1 over 500 cycles at a current density of 1000 mA g−1 [18].
However, the contact interface between MoS2 and carbon is limited in several reports. Interestingly, Jia and Cui et al. respectively reported the MoS2-based structure with expanded interplanar spacing by the intercalation of conductive carbon through in situ carbonation of organic compounds (oleylamine (OAm), ethylene glycol (EG), and dopamine) [16,19]. The carbon layer in direct contact with the MoS2 sheets results in maximizing the interface between the MoS2 layers and carbon layer, which has been considered beneficial for ion and electron transport [20]. Inspired by the strategies, positively charged poly diallyl dimethyl ammonium chloride (PDDA) has been reported to decorate metal–organic frameworks (MOFs) and reduced graphene oxide (rGO) to gain a positive charge surface [21]. A negatively charged molybdenum source can easily be adsorbed onto the PDDA. After hydrothermal and annealing treatment, the monolayer carbon derived from PDDA can intercalate between MoS2 layers to expand the interplanar spacing of MoS2.
Herein, we demonstrate a novel approach to fabricate nano-roses assembled from a MoS2 and carbon monolayer (C-MoS2) sandwiched structure nano-petals by PDDA-assisted hydrothermal and annealing treatment. The (002) interlayer spacing of C-MoS2 nano-petals was significantly expanded as large as 9.6 Å. Compared with the bare MoS2, the C-MoS2 helped to conquer the shortcomings in the following aspects. First, the enlarged interlayer distance of ultra-thin MoS2 is in favor of alleviating the mechanical stress during the charge/discharge cycles to maintain the structural integrity. Second, direct contact between MoS2 and the single-layer carbon sheet, on one hand, is beneficial for electron transport to allow for faster potassiation/depotassiation kinetics, and on the other hand, effectively prevents the aggregation of MoS2 to improve the cycling stability. Third, the few-layer MoS2 with expended distance and carbon derived from PDDA creates more active sites to improve the capacity and rate performance of C-MoS2. As a result, the C-MoS2 sandwiched structure well-addressed the poor cycling stability and low rate performance of MoS2 for PIBs. The C-MoS2-1 electrode exhibited high rate performance (123 mAh g−1 at 6.4 A g−1), which was 11 times higher than that of bare MoS2 (11 mAh g−1). C-MoS2-1 delivered a specific capacity of 437 mAh g−1 at 0.1 A g−1, which was 417 mAh g−1 higher than MoS2. At the high current density of 1.0 A g−1, the capacity of C-MoS2-1 was maintained over 273 mAh g−1 after 100 cycles, which was more than 10 times higher than that of bare MoS2 (25 mAh g−1).

2. Results and Discussion

The synthetic strategy of C-MoS2-1 involves the hydrothermal reaction and annealing treatment. First, poly dimethyl diallyl ammonium chloride (PDDA), as a conductive cationic polymer, can be strongly adsorbed on Mo7O246− anions by electrostatic attraction. In the subsequent hydrothermal process, Mo7O246− reacts with thiourea to generate MoS2. After the heat treatment, a C-MoS2 hybrid nanosheet is obtained from the crystallization of MoS2 and carbon degradation from PDDA, intercalating into MoS2 layers. The hybrid samples were named C-MoS2-X, and X is the gram weight of the PDDA aqueous solution added during sample preparation.
The crystallographic characteristics of the MoS2 and C-MoS2 samples were characterized by X-ray diffraction (XRD). As shown in Figure 1a, the bare MoS2 material showed diffraction peaks at 14.3°, 33.3°, and 59.1°, corresponding to the (002), (100), and (110) crystalline planes of 2H MoS2 (JCPDS#37-1492) [22]. According to Bragg’s formula (nλ = 2dsin θ), the distance of the (002) lattice plane can be calculated to be about 6.2 Å. Interestingly, as the content of PDDA increased, the sharp peak located at 14.3° gradually decreased and a dispersive peak at 9.2° gradually appeared, which indicates that the distance between the (002) crystal plane of MoS2 can be enlarged from about 6.2 Å to 9.6 Å by the intercalation of carbon [23,24]. Moreover, a small peak located at 18.0° gradually appeared as the content of PDDA increased, and the interplanar distance was calculated to be about 4.8 Å according to Bragg’s formula. This distance is approximately half of the distance of expended (002) lattice plane, which indicates the distance between MoS2 and the graphene layer (Figure 1b). The approximate diploid relation verifies the MoS2 and graphene monolayer inter overlapped stacked superstructure [25]. Interlayer expansion of C-MoS2 is very attractive due to improved reaction kinetics, alleviating the strain and improving the cycling stability during the charge/discharge process. Aside from peak shifting, the diffraction peaks of expanded (002) were much broader than that of bare MoS2, which could be attributed to the presence of carbon, inhibiting the growth of crystalline MoS2 [16].
Raman spectra was used to further characterize the composition and crystallinity information of the MoS2 and C-MoS2 samples. As shown in Figure 1c, two characteristic peaks located at around 384 and 413 cm−1 of bare MoS2 were attributed to the Mo-S in-plane E 2 g 1 and out-of-plane A1g vibrational modes [26]. As for C-MoS2, the A1g blue shifted to 401 cm−1, which illustrates the thickness of MoS2 decreased after adding PDDA. Compared to bare MoS2, the inter-peak separation of C-MoS2 between E 2 g 1 and A1g significantly decreased from 29 to 24 cm−1, further confirming the formation of few-layer MoS2 [27]. Few-layer MoS2 is advantageous for accelerating K+ diffusion and decreasing mechanical strain during K+ intercalation/deintercalation. Moreover, the characteristic D-band (sp3-hybridized amorphous carbon) and G-band (sp2-hybridized graphitic carbon) were observed at approximately 1355 and 1582 cm−1 from C-MoS2-0.5 and C-MoS2-1, respectively. The ratio of the relative intensities of the D- and G-bands (ID/IG) reflects the defect concentration of carbon materials [28]. The calculated ID/IG value of C-MoS2-1 (ID/IG = 1.295) was higher than that of C-MoS2-0.5 (ID/IG = 1.198), indicating a higher defect concentration of the former. These defects mainly result from PDDA derived amorphous carbon, which has been reported to improve electron conductivity [29].
The morphology and microstructure of C-MoS2-1 were investigated by scanning electron microscopy (SEM) and transmission electron microscopy (TEM) analyses (Figure 2). As shown in Figure 2a,b, it can be seen that flower-like MoS2 consists of nanosheets, and the diameter of flower-like structures is usually in the range of 400 to 600 nm. As presented in Figure 2c,d, the TEM images for C-MoS2-1 indicate that the interlayer spacing of MoS2 was about 10.0 Å, and the layer number of MoS2 nanosheets was almost in the range of 2 to 6, which is in accordance with the XRD and Raman observations. Therefore, the intercalation of monolayer carbon via organic PDDA molecule degradation would be a new strategy to fabricate few-layer C-MoS2 with expanded interlayer spacing. The expanded interlayer distance would boost the reaction kinetics, and few-layer C-MoS2 is in favor of alleviating the mechanical stress to maintain the structural integrity. As shown in Figure 2e–h, high-angle annular dark-field (HAADF) microscopy equipped with energy dispersive spectroscopy (EDS) revealed the homogeneous distribution of C, Mo, and S in C-MoS2-1.
The chemical states of C-MoS2-1 were tested by X-ray photoelectron spectroscopy (XPS) (Figure 3a–d). In the high-resolution C 1s spectra (Figure 3a), three peaks centering at 284.8, 286.4, and 288.4 eV correspond to C–C, C–O, and C–N bond, respectively [17]. In Figure 3b, the high-resolution S 2p spectra were deconvoluted into two separate peaks of S 2p1/2 and S 2p3/2 at 163.3 and 162.1 eV, respectively. In the case of the Mo 3d spectrum in Figure 3c, two strong peaks at 229.3 and 232.5 eV as well as two weak peaks at 226.5 and 235.6 eV belonged to S-Mo 3d5/2, S-Mo 3d3/2, S 2s, and Mo 3d5/2. The S 2p and Mo 3d spectra exhibit typical 2H-MoS2 features [30]. Figure 3d revealed the strong peak at 395.2 eV ascribed to Mo 3p3/2, and the existence of the pyridinic N and pyrrolic N at 398.0 and 400.2 eV, respectively, which has been reported to be beneficial for ion and electron transport [31]. The Brunauer–Emmett–Teller (BET) surface area of the C-MoS2-1 material was calculated to be 25.64 m2 g−1 (Figure 3e). The large specific surface area facilitates the contact of the electrolyte with a more active material, thus shortening the ion diffusion distance, enhancing K+ insertion/extraction within individual MoS2 monolayers, and accommodating volume change [32]. Moreover, C-MoS2-1 shows an IV isotherm with a hysteresis loop, which indicates the existence of mesopores in C-MoS2-1. Meanwhile, the pore size distribution of C-MoS2-1 mostly lies in 2 to 50 nm according to the Barrett–Joyner–Halenda (BJH) pore-size distribution curve (Figure 3f). The diffusion length for potassium ions would be shortened due to the mesoporous structure.
Cyclic voltammetry (CV) curves were recorded to precisely analyze the charge and discharge properties of the C-MoS2-1 electrode in PIBs (Figure 4a). The peak at 1.2 V was related to K+ intercalation into MoS2 layers to form KxMoS2 [33]. The weak and broad peak that appeared at about 0.8 V may be ascribed to the formation of SEI. When the potential decreased below 0.5 V, KxMoS2 subsequently underwent a conversion reaction from KxMoS2 to Mo and K2S, which is in contrast with the peak near 1.7 V for the formation of MoS2 in the anodic scan [34]. The CV curves of the C-MoS2-1 electrode overlapped each other except for the first cycle, indicating great reversibility. The discharge and charge processes of the C-MoS2-1 electrode in PIBs can be expressed as follows:
MoS2 + xK+ + xe = KxMoS2
KxMoS2 + (4 − x) K+ + (4 − x) e = Mo + 2K2S
Figure 4b shows the galvanostatic charge–discharge (GCD) profiles of C-MoS2-1 at a current density of 0.1 A g−1. The initial coulombic efficiency (ICE) of C-MoS2-1 was 57.7%, which was 0.5% higher than that of bare MoS2. The capacity loss at the first cycle was mainly due to the irreversible electrolyte reduction by the formation of the solid-electrolyte interphase (SEI) and the partially reversible redox reaction of MoS2 in the first discharge/charge cycle.
Electrochemical performances of C-MoS2 are exhibited compared to the bare MoS2 for PIBs in Figure 4c–g. The C-MoS2-1 electrode delivers a high reversible capacity of 413 mAh g−1 after 50 cycles at a current density of 0.1 A g−1. In contrast, the discharging capacities of the MoS2 and C-MoS2-0.5 electrodes were only 6 and 34 mAh g−1, respectively (Figure 4c). The increased capacity of C-MoS2-1 is mainly due to the increased spacing of the few-layer MoS2 and the addition of carbon, providing more reactive sites. As displayed in Figure 4d, C-MoS2-1 exhibited the best rate capability. At the stepwise augmenting current densities from 0.1, 0.2, 0.4, 0.8, 1.6, 3.2 to 6.4 A g−1, C-MoS2-1 delivered high specific capacities of 384, 368, 349, 312, 230, 167, and 123 mAh g−1. Nevertheless, the specific capacity of MoS2 decayed severely from 281 to 15 mAh g−1 under the same testing condition. When the current density switched back to 0.1 A g−1, the capacity of C-MoS2-1 remained at 437 mAh g−1, which was 417 mAh g−1 higher than that of the bare MoS2. The superior potassium storage properties of C-MoS2-1 not only result from the enlarged interlayer distance of few-layer MoS2, but also the improvement in the conductance. The variation in the polarization voltage of MoS2 and C-MoS2 at various current densities is an important manifestation of dynamic excellence. As shown in Figure 4e, at the stepwise current densities from 0.1, 0.2, 0.4, 0.8, 1.6, 3.2 to 6.4 A g−1, the polarization voltage of C-MoS2-1 was always smaller than that of MoS2 and C-MoS2-0.5. Moreover, the enhancive polarization voltage variation of C-MoS2-1 between 0.1 and 6.4 A g−1 was 0.83 V. In comparison, the polarization voltage increment of MoS2 and C-MoS2-0.5 was between 5.66 V and 4.33 V in the same case, which was 6.8 and 5.2 times as much as that of C-MoS2-1, respectively. The apparent superiority of C-MoS2-1 in the reaction kinetics may benefit from a unique MoS2-carbon inter overlapped superstructure. This structure efficiently facilitates electron transport and enhances the K+ diffusion kinetic, which greatly enhances the K+ transport rate in C-MoS2-1. In comparison with the other reported MoS2-based anodes for PIBs, C-MoS2-1 presented an outstanding rate performance (Figure 4f) [16,17,18,35,36,37,38,39]. At a current density of 1.0 A g−1, the superiority of C-MoS2-1 was more satisfactory (Figure 4g). The C-MoS2-1 electrode showed a reversible capacity of 273 mAh g−1 after 100 cycles, which was more than ten times higher than that of MoS2 and C-MoS2-0.5. It is worth mentioning that the specific capacity showed an increase followed by a decrease of C-MoS2-1 both at 0.1 A g−1 and 1.0 A g−1, which is speculated to be the gradually expanded interlayers of MoS2 during charge and discharge, resulting in enriched active sites for K+ storage, thus specific capacity increased at the beginning.
Electrochemical impedance spectroscopy (EIS) recorded a frequency range from 105 Hz to 10−2 Hz (Figure 5a). In the high-medium frequency region, it can be seen that the C-MoS2-1 electrode delivered the smallest diameter of the semicircle, representing decreased charge-transfer resistance because of the presence of carbon. In the low frequency region, the C-MoS2-1 electrode also presents a high slope angle than MoS2 and C-MoS2-0.5, suggesting a better K+ diffusion in the bulk electrode of C-MoS2-1 [40].
The galvanostatic intermittent titration technique (GITT) was used to further evaluate the diffusion kinetics at different potassiation and depotassiation states in the C-MoS2-1 electrodes (Figure 5b). The K+ coefficient (D) can be obtained according to the equation [41]:
D = 4 π τ ( V B m B SM B ) 2 ( Δ E s Δ E t ) 2
where τ is the relaxation time; S, mB, MB, and VB denote the geometric area of the electrode, mass, molar mass, and molar volume of the electrode material; ΔEt and ΔES represent the voltage change during the current pulse and steady-state process.
According to the GITT analysis, the D of the C-MoS2-1 electrode was calculated to almost range from 10−10 to 10−11 cm2 s−1 during the potassiation and depotassiation process (Figure 5c,d). In the discharge process, the D of K+ in the initial discharge was 1.2 × 10−10 cm2 s−1 and then increased gradually. This phenomenon was attributed to the fast insertion of K+ into the expanded interlayer MoS2 to form KxMoS2. When the discharge potential dropped to 0.7 V, the conversion reaction began and the D of K+ decreased gradually. Notably, the D rapidly decreased to 5.5 × 10−11 cm2 s−1 at the end of the conversion reaction, with a voltage-dependent feature at this stage [42]. Figure 5d shows the D of K+ in the charging process, where the D of K+ in the initial charge was 5.3 × 10−5 cm2 s−1 and then gradually decreased because of the slow conversion reaction of K2S with Mo. The D of K+ increased gradually when the charge potential lifted to about 1.2 V, which may also relate to the deintercalation of K+ from KxMoS2.

3. Materials and Methods

Materials synthesis: All chemical reagents used were without further purification in this work. First, 0, 0.5, 0.6, 0.8, and 1.0 g of a PDDA aqueous solution (20 wt%, Mw: 100,000–200,000, Aladdin, Beijing, China) were dispersed in 20 mL of deionized water. Then, 400 mg of (NH4)6Mo7O24·4H2O powder (AR, Hushi, Shanghai, China) was dissolved in 10 mL of deionized water followed by the dropwise addition of the PDDA aqueous solution to obtain a milk-white homogeneous suspension. After 30 min, 1.0 g thiourea (99%, Zhanyun, Shanghai, China) was dissolved in the obtained suspension. The synthesis time was the same for all samples. The as-prepared suspension was transferred into a Teflon-lined autoclave and maintained at 200 °C for 20 h. The black suspension was washed to neutral and freeze-dried to obtain the precursor. Finally, C-MoS2 was obtained by heat-treating the precursor at 800 °C for 2 h with a ramping rate of 3 °C min−1 in an argon flow (100 sccm) tube furnace. The samples were named MoS2, C-MoS2-0.5, and C-MoS2-1 according to the gram weight of the PDDA aqueous solution added during the preparation.
Material characterization: XRD patterns were taken using an X-ray diffractometer (D8 Advance DaVinci, Bruker, Bremen, Germany) with Cu Kα radiation (λ = 1.5406 Å). Raman spectra were collected on a micro-Raman spectrometer (inVia-reflex, Renishaw, Kingswood, UK) using a laser wavelength (532 nm) with a 100% filter. SEM images and TEM images were conducted using a scanning electron microscope (S4800, HITACHI, Tokyo, Japan) and transmission electron microscope (Talos F200x, Thermo Fisher, Waltham, MA, USA). XPS analysis was implemented by a photoelectron spectrometer (AXIS ULTRA DLD, Shimadzu, Kyoto, Japan). The specific surface area and pore size distribution of the sample were evaluated by an N2 adsorption/desorption analyzer (ASAP-2020M, Micrometric, GA, USA).
Electrochemical performance measurements: The slurry was prepared by mixing 80 wt% active material, 10 wt% carbon black, and 10 wt% polyvinylidene in N-methyl-2-pyrrolidone (NMP). Then, the blade was coated with a copper foil and dried at 120 °C in a vacuum oven for 12 h. The working electrodes were further fabricated by cutting into disks with a diameter of 12 mm. The electrolyte was 0.8 mol L−1 KPF6 in ethylene carbonate and diethyl carbonate (EC: DEC = 1:1 in volume). Pure potassium metal pieces were used as the counter electrode and the glass fiber mat was used as the separator. With these components, CR 2016 coin-type cells were assembled in an argon-filled glove box.
The CV and EIS tests were carried out on a 1470 E electrochemical workstation (Solartron Metrology, Bognor Regis, UK). EIS was measured over the frequency range from 105 Hz to 10−2 Hz with a sinusoidal voltage amplitude of 10 mV. The GCD and GITT tests were performed on a CT2001A measurement system (LAND, Wuhan, China) in the voltage range of 0.01–3.00 V (vs. K/K+) at room temperature.

4. Conclusions

In summary, an interesting induced growth of carbon on the interlayer of MoS2 (C-MoS2) was successfully achieved under the assistance of PDDA by the hydrothermal reaction and annealing treatment. The interlayer distance of few-layer C-MoS2 was expanded from 6.2 to 9.6 Å due to the formation of the MoS2-carbon inter overlapped superstructure. The enlarged interlayer distance of MoS2 is in favor of alleviating the mechanical stress during the charge and discharge cycles and maintaining the structural integrity. Benefiting from the direct contact between the single-layer MoS2 and carbon, the charge transfer rate was obviously enhanced, the resultant C-MoS2 sample exhibited faster potassiation and depotassiation kinetics. The MoS2 with expended distance and carbon derived from PDDA created more active sites to improve the capacity of C-MoS2. As a result, the C-MoS2 well addresses the poor cycling stability, low capacity, and low rate performance of MoS2 for PIBs. C-MoS2-1 exhibited excellent electrochemical properties with higher reversibility specific capacity (437 mAh g−1 at 0.1 A g−1), superior rate capability (123 mAh g−1 at 6.4 A g−1), and improved cycle stability (273 mAh g−1 after 100 cycles at 1.0 A g−1). This work indicates that C-MoS2 has great potential as a promising MoS2-based anode material for high performance PIBs.

Author Contributions

Conceptualization, H.H.; Methodology, Y.Z.; Validation, B.C. and H.X.; Writing—original draft preparation, Y.Z.; Writing—review and editing, H.D.; Visualization, L.Z.; Supervision, Q.W.; Funding acquisition, H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the High-quality Development Project of Ministry of Industry and Information Technology of People’s Republic of China (TC210H041) (to He, H.), the Hundred Talents Program, the National Natural Science Foundation of China (Grant No. 51872304) (to He, H.), and the Ningbo S&T Innovation 2025 Major Special Program (2018B10024; 2019B10044; 2020Z101; 2022Z022) (to He, H.).

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. Min, X.; Xiao, J.; Fang, M.; Wang, W.; Zhao, Y.; Liu, Y.; Abdelkader, A.M.; Xi, K.; Kumar, R.V.; Huang, Z. Potassium-ion batteries: Outlook on present and future technologies. Energy Environ. Sci. 2021, 14, 2186–2243. [Google Scholar] [CrossRef]
  2. Hosaka, T.; Kubota, K.; Hameed, A.S.; Komaba, S. Research Development on K-Ion Batteries. Chem. Rev. 2020, 120, 6358–6466. [Google Scholar] [CrossRef]
  3. Zhang, W.; Yin, J.; Wang, W.; Bayhan, Z.; Alshareef, H.N. Status of rechargeable potassium batteries. Nano Energy 2021, 83, 105792. [Google Scholar] [CrossRef]
  4. Zhang, Q.; Cheng, X.; Wang, C.; Rao, A.M.; Lu, B. Sulfur-assisted large-scale synthesis of graphene microspheres for superior potassium-ion batteries. Energy Environ. Sci. 2021, 14, 965–974. [Google Scholar] [CrossRef]
  5. Yuan, F.; Lei, Y.; Wang, H.; Li, X.; Hu, J.; Wei, Y.; Zhao, R.; Li, B.; Kang, F.; Zhai, D. Pseudo-capacitance reinforced modified graphite for fast-charging potassium-ion batteries. Carbon 2021, 185, 48–56. [Google Scholar] [CrossRef]
  6. Wang, B.; Yuan, F.; Yu, Q.; Li, W.; Sun, H.; Zhang, L.; Zhang, D.; Wang, Q.; Lai, F.; Wang, W. Amorphous carbon/graphite coupled polyhedral microframe with fast electronic channel and enhanced ion storage for potassium ion batteries. Energy Storage Mater. 2021, 38, 329–337. [Google Scholar] [CrossRef]
  7. Luo, P.; Zheng, C.; He, J.; Tu, X.; Sun, W.; Pan, H.; Zhou, Y.; Rui, X.; Zhang, B.; Huang, K. Structural Engineering in Graphite-Based Metal-Ion Batteries. Adv. Funct. Mater. 2021, 32, 2107277. [Google Scholar] [CrossRef]
  8. Cheng, X.; Sun, Y.; Li, D.; Yang, H.; Chen, F.; Huang, F.; Jiang, Y.; Wu, Y.; An, X.; Yu, Y. From 0D to 3D: Dimensional Control of Bismuth for Potassium Storage with Superb Kinetics and Cycling Stability. Adv. Energy Mater. 2021, 11, 2102263. [Google Scholar] [CrossRef]
  9. Huang, K.; Xing, Z.; Wang, L.; Wu, X.; Zhao, W.; Qi, X.; Wang, H.; Ju, Z. Direct synthesis of 3D hierarchically porous carbon/Sn composites via in situ generated NaCl crystals as templates for potassium-ion batteries anode. J. Mater. Chem. A 2018, 6, 434–442. [Google Scholar] [CrossRef]
  10. Zhang, D.; Chen, B.; Wang, S.; Shen, Y.; Wang, C.; Wang, Z.; Wang, L.; Cheng, Y. Boosting Reversibility of Conversion/Alloying Reactions for Sulfur-Rich Antimony-Based Sulfides with Extraordinary Potassium Storage Performance. ACS Mater. Lett. 2022, 4, 2604–2612. [Google Scholar] [CrossRef]
  11. Huang, M.; Xi, B.; Mi, L.; Zhang, Z.; Chen, W.; Feng, J.; Xiong, S. Rationally Designed Three-Layered TiO2@amorphous MoS3@Carbon Hierarchical Microspheres for Efficient Potassium Storage. Small 2022, 18, 2107819. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, S.; Kang, L.; Henzie, J.; Zhang, J.; Ha, J.; Amin, M.A.; Hossain, M.S.A.; Jun, S.C.; Yamauchi, Y. Recent Advances and Perspectives of Battery-Type Anode Materials for Potassium Ion Storage. ACS Nano 2021, 15, 18931–18973. [Google Scholar] [CrossRef] [PubMed]
  13. Luo, W.; Wan, J.; Ozdemir, B.; Bao, W.; Chen, Y.; Dai, J.; Lin, H.; Xu, Y.; Gu, F.; Barone, V.; et al. Potassium Ion Batteries with Graphitic Materials. Nano Lett. 2015, 15, 7671–7677. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, S.; Cui, X.; Jian, C.; Cheng, H.; Niu, M.; Yu, J.; Yan, J.; Huang, W. Transition Metal Dichalcogenides: Stacking-Engineered Heterostructures in Transition Metal Dichalcogenides. Adv. Mater. 2021, 33, 2170127. [Google Scholar] [CrossRef]
  15. Ren, X.; Zhao, Q.; McCulloch, W.D.; Wu, Y. MoS2 as a long-life host material for potassium ion intercalation. Nano Res. 2017, 10, 1313–1321. [Google Scholar] [CrossRef]
  16. Jia, B.; Yu, Q.; Zhao, Y.; Qin, M.; Wang, W.; Liu, Z.; Lao, C.; Liu, Y.; Wu, H.; Zhang, Z.; et al. Bamboo-Like Hollow Tubes with MoS2/N-Doped-C Interfaces Boost Potassium-Ion Storage. Adv. Funct. Mater. 2018, 28, 1803409. [Google Scholar] [CrossRef]
  17. Wang, H.; Niu, J.; Shi, J.; Lv, W.; Wang, H.; van Aken, P.A.; Zhang, Z.; Chen, R.; Huang, W. Facile Preparation of MoS2 Nanocomposites for Efficient Potassium-Ion Batteries by Grinding-Promoted Intercalation Exfoliation. Small 2021, 17, 2102263. [Google Scholar] [CrossRef]
  18. Hu, J.; Xie, Y.; Zhou, X.; Zhang, Z. Engineering Hollow Porous Carbon-Sphere-Confined MoS2 with Expanded (002) Planes for Boosting Potassium-Ion Storage. ACS Appl. Mater. Interfaces 2020, 12, 1232–1240. [Google Scholar] [CrossRef]
  19. Cui, Y.; Liu, W.; Feng, W.; Zhang, Y.; Du, Y.; Liu, S.; Wang, H.; Chen, M.; Zhou, J. Controlled Design of Well-Dispersed Ultrathin MoS2 Nanosheets inside Hollow Carbon Skeleton: Toward Fast Potassium Storage by Constructing Spacious “Houses” for K Ions. Adv. Funct. Mater. 2020, 30, 1908755. [Google Scholar] [CrossRef]
  20. Shi, Z.; Kang, W.; Xu, J.; Sun, Y.; Jiang, M.; Ng, T.; Xue, H.; Yu, D.; Zhang, W.; Lee, C. Hierarchical nanotubes assembled from MoS2-carbon monolayer sandwiched superstructure nanosheets for high-performance sodium ion batteries. Nano Energy 2016, 22, 27–37. [Google Scholar] [CrossRef]
  21. Li, Z.; Yin, L. Sandwich-like reduced graphene oxide wrapped MOF-derived ZnCo2O4–ZnO–C on nickel foam as anodes for high performance lithium ion batteries. J. Mater. Chem. A 2015, 3, 21569–21577. [Google Scholar] [CrossRef]
  22. Wang, J.; Sun, L.; Tan, H.; Xie, F.; Qu, Y.; Hu, J.; Gao, K.; Shi, X.; Wang, K.; Zhang, Y. Double-phase 1T/2H–MoS2 heterostructure loaded in N-doped carbon/CNT complex carbon for efficient and rapid lithium storage. Mater. Today Energy 2022, 29, 101103. [Google Scholar] [CrossRef]
  23. Liang, S.; Zhang, S.; Liu, Z.; Feng, J.; Jiang, Z.; Shi, M.; Chen, L.; Wei, T.; Fan, Z. Approaching the Theoretical Sodium Storage Capacity and Ultrahigh Rate of Layer-Expanded MoS2 by Interfacial Engineering on N-Doped Graphene. Adv. Energy Mater. 2021, 11, 2002600. [Google Scholar] [CrossRef]
  24. Saraf, M.; Natarajan, K.; Saini, A.K.; Mobin, S.M. Small biomolecule sensors based on an innovative MoS2-rGO heterostructure modified electrode platform: A binder-free approach. Dalton Trans. 2017, 46, 15848–15858. [Google Scholar] [CrossRef] [PubMed]
  25. Yu, X.; Zhao, G.; Liu, C.; Zhang, N. A MoS2 and Graphene Alternately Stacking van der Waals Heterostructure for Li+/Mg2+ Co-Intercalation. Adv. Funct. Mater. 2021, 31, 2103214. [Google Scholar] [CrossRef]
  26. Ganatra, R.; Zhang, Q. Few-Layer MoS2: A Promising Layered Semiconductor. ACS Nano 2014, 8, 4074–4099. [Google Scholar] [CrossRef]
  27. Li, X.; Li, J.; Ma, L.; Yu, C.; Ji, Z.; Pan, L.; Mai, W. Graphite Anode for Potassium Ion batteries: Current Status and Perspective. Energy Environ. Mater. 2021, 5, 458–469. [Google Scholar] [CrossRef]
  28. Liu, C.; Xie, H.; Sui, S.; Chen, B.; Ma, L.; Liu, E.; Zhao, N. Interface engineering of MoS2-based ternary hybrids towards reversible conversion of sodium storage. Mater. Today Energy 2022, 26, 100993. [Google Scholar] [CrossRef]
  29. Lim, H.; Yu, S.; Choi, W.; Kim, S.O. Hierarchically Designed Nitrogen-Doped MoS2/Silicon Oxycarbide Nanoscale Heterostructure as High-Performance Sodium-Ion Battery Anode. ACS Nano 2021, 15, 7409–7420. [Google Scholar] [CrossRef]
  30. Saraf, M.; Natarajan, K.; Mobin, S.M. Emerging Robust Heterostructure of MoS2-rGO for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2018, 10, 16588–16595. [Google Scholar] [CrossRef]
  31. Zhao, C.; Yu, C.; Qiu, B.; Zhou, S.; Zhang, M.; Huang, H.; Wang, B.; Zhao, J.; Sun, X.; Qiu, J. Ultrahigh Rate and Long-Life Sodium-Ion Batteries Enabled by Engineered Surface and Near-Surface Reactions. Adv. Mater. 2018, 30, 1702486. [Google Scholar] [CrossRef] [PubMed]
  32. Gan, Z.; Yin, J.; Xu, X.; Cheng, Y.; Yu, T. Nanostructure and Advanced Energy Storage: Elaborate Material Designs Lead to High-Rate Pseudocapacitive Ion Storage. ACS Nano 2022, 16, 5131–5152. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, L.; Chen, Z.; Chen, L.; Zhou, P.; Wang, J.; Yang, H.; Feng, Z.; Li, X.; Huang, J. The exfoliation of bulk MoS2 by a three-roller mill for high-performance potassium ion batteries. Appl. Surf. Sci. 2023, 615, 156253. [Google Scholar] [CrossRef]
  34. Li, X.; Qian, T.; Zai, J.; He, K.; Feng, Z.; Qian, X. Co stabilized metallic 1Td MoS2 monolayers: Bottom-up synthesis and enhanced capacitance with ultra-long cycling stability. Mater. Today Energy 2018, 7, 10–17. [Google Scholar] [CrossRef]
  35. Yao, K.; Xu, Z.; Ma, M.; Li, J.; Lu, F.; Huang, J. Densified Metallic MoS2/Graphene Enabling Fast Potassium-Ion Storage with Superior Gravimetric and Volumetric Capacities. Adv. Funct. Mater. 2020, 30, 2001484. [Google Scholar] [CrossRef]
  36. Zhou, Y.; Zhang, M.; Han, Q.; Liu, Y.; Wang, Y.; Sun, X.; Zhang, X.; Dong, C.; Jiang, F. Hierarchical 1 T-MoS2/MoOx@NC microspheres as advanced anode materials for potassium/sodium-ion batteries. Chem. Eng. J. 2022, 428, 131113. [Google Scholar] [CrossRef]
  37. Li, J.; Rui, B.; Wei, W.; Nie, P.; Chang, L.; Le, Z.; Liu, M.; Wang, H.; Wang, L.; Zhang, X. Nanosheets assembled layered MoS2/MXene as high performance anode materials for potassium ion batteries. J. Power Sources 2020, 449, 227481. [Google Scholar] [CrossRef]
  38. Chong, S.; Sun, L.; Shu, C.; Guo, S.; Liu, Y.; Wang, W.; Liu, H.K. Chemical bonding boosts nano-rose-like MoS2 anchored on reduced graphene oxide for superior potassium-ion storage. Nano Energy 2019, 63, 103868. [Google Scholar] [CrossRef]
  39. Xu, Y.; Bahmani, F.; Zhou, M.; Li, Y.; Zhang, C.; Liang, F.; Kazemi, S.H.; Kaiser, U.; Meng, G.; Lei, Y. Enhancing potassium-ion battery performance by defect and interlayer engineering. Nanoscale Horiz. 2019, 4, 202–207. [Google Scholar] [CrossRef] [PubMed]
  40. Vivier, V.; Orazem, M.E. Impedance Analysis of Electrochemical Systems. Chem. Rev. 2022, 122, 11131–11168. [Google Scholar] [CrossRef] [PubMed]
  41. Fang, G.; Wu, Z.; Zhou, J.; Zhu, C.; Cao, X.; Lin, T.; Chen, Y.; Wang, C.; Pan, A.; Liang, S. Observation of Pseudocapacitive Effect and Fast Ion Diffusion in Bimetallic Sulfides as an Advanced Sodium-Ion Battery Anode. Adv. Energy Mater. 2018, 8, 1703155. [Google Scholar] [CrossRef]
  42. Wang, P.; Sun, S.; Jiang, Y.; Cai, Q.; Zhang, Y.H.; Zhou, L.; Fang, S.; Liu, J.; Yu, Y. Hierarchical Microtubes Constructed by MoS2 Nanosheets with Enhanced Sodium Storage Performance. ACS Nano 2020, 14, 15577–15586. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Characterization of samples. (a) XRD patterns of the MoS2 and C-MoS2 samples. (b) Crystalline structure of C-MoS2-1. (c) Raman spectra of the MoS2 and C-MoS2 samples.
Figure 1. Characterization of samples. (a) XRD patterns of the MoS2 and C-MoS2 samples. (b) Crystalline structure of C-MoS2-1. (c) Raman spectra of the MoS2 and C-MoS2 samples.
Molecules 28 02608 g001
Figure 2. Morphology of samples. (a,b) SEM images of C-MoS2-1. (c) Low-resolution TEM of C-MoS2-1. (d) High-resolution TEM of C-MoS2-1. (eh) HAADF and EDS of C, Mo, and S for C-MoS2-1.
Figure 2. Morphology of samples. (a,b) SEM images of C-MoS2-1. (c) Low-resolution TEM of C-MoS2-1. (d) High-resolution TEM of C-MoS2-1. (eh) HAADF and EDS of C, Mo, and S for C-MoS2-1.
Molecules 28 02608 g002
Figure 3. Characterization of samples. (ad) High-resolution C 1s, S 2p, Mo 3d, and N 1s XPS spectra of C-MoS2-1. (e,f) N2 adsorption–desorption isotherm and BJH pore-size distribution curve of C-MoS2-1.
Figure 3. Characterization of samples. (ad) High-resolution C 1s, S 2p, Mo 3d, and N 1s XPS spectra of C-MoS2-1. (e,f) N2 adsorption–desorption isotherm and BJH pore-size distribution curve of C-MoS2-1.
Molecules 28 02608 g003
Figure 4. Electrochemical performance of samples. (a) CV profiles of C-MoS2-1. (b) GCD curves of C-MoS2-1 electrode at a current density of 0.1 A g−1. (c) Cycling performance of MoS2 and C-MoS2 electrodes at 0.1 A g−1. (d) Rate capability and (e) polarization voltage comparison under different current densities of MoS2 and C-MoS2 electrodes. (f) The comparison of the rate capability between C-MoS2-1.0 and the other reported MoS2-based anodes for PIBs. (g) Cycling performance of MoS2 and C-MoS2 at a current density of 1.0 A g−1.
Figure 4. Electrochemical performance of samples. (a) CV profiles of C-MoS2-1. (b) GCD curves of C-MoS2-1 electrode at a current density of 0.1 A g−1. (c) Cycling performance of MoS2 and C-MoS2 electrodes at 0.1 A g−1. (d) Rate capability and (e) polarization voltage comparison under different current densities of MoS2 and C-MoS2 electrodes. (f) The comparison of the rate capability between C-MoS2-1.0 and the other reported MoS2-based anodes for PIBs. (g) Cycling performance of MoS2 and C-MoS2 at a current density of 1.0 A g−1.
Molecules 28 02608 g004
Figure 5. Reaction kinetics of samples. (a) Nyquist plots of the MoS2 and C-MoS2 electrodes. (b) GITT test of C-MoS2-1 electrode. (c,d) K+ diffusion coefficients from GITT of C-MoS2-1 electrode under different voltage. Inset: Diffusion coefficient under different potassiation and depotassiation states.
Figure 5. Reaction kinetics of samples. (a) Nyquist plots of the MoS2 and C-MoS2 electrodes. (b) GITT test of C-MoS2-1 electrode. (c,d) K+ diffusion coefficients from GITT of C-MoS2-1 electrode under different voltage. Inset: Diffusion coefficient under different potassiation and depotassiation states.
Molecules 28 02608 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Y.; Zhu, L.; Xu, H.; Wu, Q.; Duan, H.; Chen, B.; He, H. Interlayer-Expanded MoS2 Enabled by Sandwiched Monolayer Carbon for High Performance Potassium Storage. Molecules 2023, 28, 2608. https://doi.org/10.3390/molecules28062608

AMA Style

Zhang Y, Zhu L, Xu H, Wu Q, Duan H, Chen B, He H. Interlayer-Expanded MoS2 Enabled by Sandwiched Monolayer Carbon for High Performance Potassium Storage. Molecules. 2023; 28(6):2608. https://doi.org/10.3390/molecules28062608

Chicago/Turabian Style

Zhang, Yuting, Lin Zhu, Hongqiang Xu, Qian Wu, Haojie Duan, Boshi Chen, and Haiyong He. 2023. "Interlayer-Expanded MoS2 Enabled by Sandwiched Monolayer Carbon for High Performance Potassium Storage" Molecules 28, no. 6: 2608. https://doi.org/10.3390/molecules28062608

Article Metrics

Back to TopTop