Next Article in Journal
Calcium Phosphate Cements as Carriers of Functional Substances for the Treatment of Bone Tissue
Next Article in Special Issue
New Materials Used for the Development of Anion-Selective Electrodes—A Review
Previous Article in Journal
Interfacial Modification and Bending Performance of 3D Orthogonal Woven Composites with Basalt Filament Yarns
Previous Article in Special Issue
A Novel POP-Ni Catalyst Derived from PBTP for Ambient Fixation of CO2 into Cyclic Carbonates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Three-Dimensional Flower-like MoS2 Nanosheets Grown on Graphite as High-Performance Anode Materials for Fast-Charging Lithium-Ion Batteries

1
Korea Institute of Energy Research (KIER), Daejeon 34129, Republic of Korea
2
Graduate School of Energy Science and Technology (GEST), Chungnam National University, Daejeon 34134, Republic of Korea
3
Department of Advanced Energy Technologies and System Engineering, Korea University of Science and Technology (UST), Daejeon 34113, Republic of Korea
4
Computational Science and Engineering Laboratory, Korea Institute of Energy Research (KIER), Daejeon 34129, Republic of Korea
5
Department of Physics, Chungnam National University, Daejeon 34134, Republic of Korea
6
Department of Chemistry, Mokpo National University, Muan-gun 58554, Republic of Korea
7
Department of Energy Systems, Soonchunhyang University, Asan 31538, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2023, 16(11), 4016; https://doi.org/10.3390/ma16114016
Submission received: 20 April 2023 / Revised: 15 May 2023 / Accepted: 24 May 2023 / Published: 27 May 2023
(This article belongs to the Special Issue Novel Nanomaterials for Energy Storage and Catalysis)

Abstract

:
The demand for fast-charging lithium-ion batteries (LIBs) with long cycle life is growing rapidly due to the increasing use of electric vehicles (EVs) and energy storage systems (ESSs). Meeting this demand requires the development of advanced anode materials with improved rate capabilities and cycling stability. Graphite is a widely used anode material for LIBs due to its stable cycling performance and high reversibility. However, the sluggish kinetics and lithium plating on the graphite anode during high-rate charging conditions hinder the development of fast-charging LIBs. In this work, we report on a facile hydrothermal method to achieve three-dimensional (3D) flower-like MoS2 nanosheets grown on the surface of graphite as anode materials with high capacity and high power for LIBs. The composite of artificial graphite decorated with varying amounts of MoS2 nanosheets, denoted as MoS2@AG composites, deliver excellent rate performance and cycling stability. The 20−MoS2@AG composite exhibits high reversible cycle stability (~463 mAh g−1 at 200 mA g−1 after 100 cycles), excellent rate capability, and a stable cycle life at the high current density of 1200 mA g−1 over 300 cycles. We demonstrate that the MoS2-nanosheets-decorated graphite composites synthesized via a simple method have significant potential for the development of fast-charging LIBs with improved rate capabilities and interfacial kinetics.

1. Introduction

Among various energy storage technologies, lithium-ion batteries (LIBs) have been widely investigated as power sources for portable electronics and electric vehicles (EVs) due to their high energy density and long lifespan [1,2,3,4]. The rapid expansion of the global EV and energy storage systems (ESSs) market has led to significant demand for fast-charging battery technology that can support the high power and long cycle life requirements of these applications. Graphite, a commercial anode material used in LIBs, is still considered the most promising material because of its excellent cycle reversibility, stable cycle life, and superior electronic conductivity. Despite these advantages, due to its low theoretical capacity (~372 mAh g−1) and limited rate capability under fast charging conditions, graphite cannot meet the growing performance requirements of LIBs. The slow kinetics of lithium intercalation at the graphite–electrolyte interface during rapid charging conditions can lead to an undesirable anode voltage drop below 0 V vs. Li/Li+, resulting in the formation of lithium plating on the graphite surface. This phenomenon can cause capacity decay, dendrite growth, short circuits, and even serious safety issues [5,6,7,8]. To overcome this challenge, various approaches have been proposed to develop high-power lithium-ion battery (LIB) anodes with improved interfacial kinetics [9,10,11,12,13,14,15]. Among these approaches, surface modification using functional materials has emerged as an effective strategy to enhance the rate capability and cycle stability of graphite anodes [7,8,12,13,14,15].
Transition metal sulfides (TMDs) have been extensively investigated in various energy storage applications, such as batteries, supercapacitors, and electrocatalysts [16,17,18]. Among various TMDs, molybdenum disulfide (MoS2), a typical two-dimensional (2D) layered material, has emerged as a promising anode material for LIBs due to its high theoretical capacity (~670 mAh g−1), which is much higher than that of graphite [19]. Furthermore, the interlayer spacing of MoS2 (~0.62 nm) is larger than that of graphite (0.34 nm), providing ample space for fast Li+ diffusion paths [20,21]. However, MoS2 suffers from rapid capacity decay and inferior rate capability due to its low intrinsic electronic conductivity, large volume variation during cycling, and the production of polysulfide dissolution during the charge and discharge process [22,23,24,25,26].
Herein, we report a facile and efficient approach to synthesizing a 3D hierarchical MoS2/artificial graphite (MoS2@AG) composite for use as an anode material in LIBs. The composite was prepared using a hydrothermal reaction to decorate 3D hierarchically aligned flower-like MoS2 nanosheets directly on the surface of graphite. The resulting composite exhibits improved specific capacity and rate capability compared to commercial graphite anodes. The graphite matrix provides a conductive pathway for fast electron transfer through the electrode, while the 3D hierarchically aligned MoS2 nanosheets form a stable interface with the graphite, effectively preventing structural degradation and providing excellent electron transport. Moreover, the 3D hierarchical morphology of the MoS2 nanosheets enhances the electrode/electrolyte contact area, facilitating charge transfer kinetics. As a result, the MoS2-decorated graphite composite demonstrates exceptional rate capability, achieving charging times of less than 20 min for approximately 84% of its capacity. The composite also exhibits superior cycle stability under fast charging conditions compared to commercial graphite anodes. These results highlight the potential of the MoS2/graphite composite as a promising candidate for high-energy and high-power LIBs.

2. Materials and Methods

2.1. Preparation of MoS2- and MoS2-Decorated Graphite Composites

The MoS2-decorated graphite composite was synthesized via a facile hydrothermal reaction. To prepare the precursor solution, hexaammonium molybdate tetrahydrate ((NH4)6Mo7O24⋅4H2O) and thiourea (NH2CSNH2), both purchased from Sigma Aldrich, were dissolved in a molar ratio of 1:14 in 50 mL of water dispersion solution containing artificial graphite (Hitachi Chemical Co., Japan) at a concentration of 0.0342 g mL−1. The mixture was vigorously stirred for 1 h. Subsequently, the mixture was transferred into a Teflon-lined stainless autoclave and reacted at 220 °C for 16 h. After cooling naturally, the black precipitates were collected by centrifugation, washed with distilled water and ethanol, and dried at 80 °C for 24 h. The amount of thiourea and (NH4)6Mo7O24⋅4H2O added to the graphite solution varied to achieve mass ratios of 2.5:100, 5:100, 10:100, 20:100, or 30:100 and the corresponding products were denoted as 2.5−MoS2@AG, 5−MoS2@AG, 10−MoS2@AG, 20−MoS2@AG, or 30−MoS2@AG. The preparation process for the MoS2-decorated graphite composite is illustrated in Figure 1a. The synthesis procedure for a pristine flower-like MoS2 was similar to that of MoS2-decorated graphite composite but without the presence of graphite.

2.2. Characterizations

Powder X-ray diffraction (XRD) was measured on an X-ray diffractometer (Rigaku SmartLab High Resolution, Tokyo, Japan) using Cu-Kα radiation (λ = 1.5406 Å). Raman spectroscopy measurements were taken using a Raman spectrometer (HORIBA, LabRAM HR Evolution, Lyon, France) with an excitation wavelength of 514 nm. X-ray photoelectron microscopy (XPS) analysis was conducted using a Thermo VG Scientific Sigma Probe instrument with a micro-focused monochromatized Al Ka X-ray source (1486.6 eV). Field-emission scanning electron microscopy (FE-SEM) images were acquired using a FEI NovaNano SEM 450, and transmission electron microscopy (TEM) with energy-dispersive X-ray spectroscopy (EDS) was performed using a JEOL F200 instrument to observe the surface morphologies and microstructures. The specific surface areas of pristine graphite (AG) and x−MoS2@AG composites were obtained by using the Brunauer–Emmett–Teller (BET) method, and pore size distribution was calculated using Barrett–Joyner–Halenda (BJH) method (BELSORP Max, Microtrac MRB, Osaka, Japan). Thermogravimetric analysis (TGA) was carried out on a TGA2 (Mettler Toledo) at a heating rate of 10 °C min−1 from 30 to 700 °C under an air atmosphere.

2.3. Electrochemical Measurements

The electrochemical properties of the MoS2@AG composites were evaluated using a CR2032 coin-type half-cell assembled in an argon-filled glove box. A slurry was prepared by mixing the active material (MoS2@AG composites), conducting agent (Super P), and polyvinylidene fluoride (PVDF) binder at a weight ratio of 8:1:1 in a solvent of N-methyl-2-pyrrolidone (NMP). The resulting slurry was coated onto a copper foil and dried in a vacuum at 120 °C for 12 h. The average loading density of the anode electrodes was about 2 mg cm−2. The anode was composed of the MoS2@AG-composite-coated copper foil, while a lithium metal foil was used as the counter/reference electrode. A polypropylene membrane was used as the separator. The electrolyte solution comprised a 1M LiPF6 solution in a mixture of ethylene carbonate (EC)/ethylene methyl carbonate (EMC)/diethyl carbonate (DEC) (=3:4:3, v/v/v) containing 1% vinylene carbonate (VC). The electrochemical cell was fabricated using 160 μL of electrolyte. The galvanostatic charge and discharge curves of the cells were obtained using a battery testing system (Biologic, BCS) in a voltage range from 0.01 to 2.5 V vs. Li/Li+ with various current densities. The electrochemical impedance spectroscopy (EIS) measurements were performed at an AC amplitude of 10 mV over the frequency range from 200 kHz to 50 mHz (Biologic, VSP) and the equivalent circuit fitting was conducted using ZView Software 3.2 (Scribner Associates, Inc., Southern Pines, NC, USA).

3. Results and Discussion

Figure 1a illustrates the one-step hydrothermal synthesis process for forming 3D hierarchical flower-like MoS2 nanosheets directly grown on the surface of graphite. Figure 1b–d show representative SEM, TEM, and high-resolution TEM (HR-TEM) images of the MoS2@AG composites, respectively. The x−MoS2@AG composites with different weight percentages of sulfur and molybdenum sources relative to graphite were synthesized. In this process, MoS2 nanosheets were spontaneously nucleated and grown on the surface of graphite. Ultimately, a 3D composite of few-layer MoS2 nanosheets with a low loading on graphite was obtained. The unique structure of the MoS2-decorated graphite composite is anticipated to significantly enhance the capacity and rate capability of LIBs. Graphite acts as a support for MoS2 nucleation, thereby providing an excellent electron transfer channel. Additionally, 3D hierarchically aligned flower-like MoS2 nanosheets grown on the surface of graphite improve the rate characteristics by facilitating lithium diffusion and increase the specific capacity by offering abundant exposed active sites.
The morphologies of pristine graphite and MoS2-decorated graphite composites were examined by SEM and SEM-backscattered electron (SEM-BSE) images to investigate the effect of the amounts of thiourea and (NH4)6Mo7O24⋅4H2O on the formation process (Figure 2 and Figure S1). Three-dimensionally grown plate-like MoS2 nanosheets were observed on the surface of 2.5−MoS2@AG and 5−MoS2@AG composites, as shown in Figure 2a–c and Figure 2d–f, respectively. In contrast, flower-shaped MoS2 nanosheets were observed on the surfaces of 10−MoS2@AG and 20−MoS2@AG composites (Figure 2g–l). The increase in the amounts of thiourea and (NH4)6Mo7O24⋅4H2O provided more nucleation sites, resulting in the formation of more flower-like MoS2 nanosheets. Figure S2 shows the enlarged SEM images of the flower-like MoS2 nanosheets. Furthermore, Figure 2c,f,i,l show the SEM-BSE images corresponding to Figure 2b,e,h,k, respectively. The brighter areas indicate higher average atomic numbers, confirming the successful formation of MoS2 on the graphite surface.
The specific surface area and pore structure of pristine graphite (AG) and 20−MoS2@AG composite were examined by the N2 adsorption and desorption isotherm measurements (Figure S3). The Brunauer–Emmett–Teller (BET) specific surface area of the 20−MoS2@AG composite was calculated to be 2.98 m2 g−1, which is almost 2.6 times larger than that of pristine AG (1.15 m2 g−1). As shown in the inset of Figure S3, the mean pore diameters of AG and 20−MoS2@AG were 29.28 and 21.48 nm, respectively, and 20−MoS2@AG showed narrower average pores. The increased specific surface area and decreased pore diameter of the 20−MoS2@AG composite can be attributed to the formation of 3D hierarchically aligned MoS2 nanosheets on the surface of graphite. These structural features are expected to improve the rate characteristic by facilitating lithium diffusion and increase the specific capacity by offering abundant exposed active sites.
The amount of MoS2 grown on the graphite surface was calculated through thermogravimetric analysis (TGA), as shown in Figure S4. In the case of the pristine MoS2, the weight loss from 300 to 500 °C is indicative of the oxidation of MoS2 to MoO3. As for the 20−MoS2@AG composites, the TGA profiles display a two-step weight decrease, which can be attributed to the consecutive oxidations of MoS2 and carbon, respectively. The MoS2 content in the 20−MoS2@AG was estimated to be <5 wt% through the TGA results.
The low-magnification TEM image (Figure 3a) reveals the MoS2 nanosheets directly grown on the graphite surface. High-resolution TEM (HR-TEM) images in Figure 3b,c show that the few-layered MoS2 nanosheets are grown on the surface of graphite. The MoS2 nanosheets consist of 9–14 layers, and the interlayer distance (002) of MoS2 was found to be 0.64~0.67 nm, which is slightly larger than that of conventional bulk MoS2 (0.62 nm). This expanded interlayer spacing could contribute to enhanced kinetics and a low energy barrier for ion intercalation [25,26,27]. Furthermore, high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images of MoS2-decorated graphite are shown in Figure 3d and energy dispersive spectroscopic (EDS) mapping images of MoS2-decorated graphite reveal the existence of C, Mo, and S elements in the composite, as shown in Figure 3e–g.
Figure 4a shows the powder XRD pattern of AG, 2.5−MoS2@AG, 5−MoS2@AG, 10−MoS2@AG, and 20−MoS2@AG. The diffraction peaks observed in all MoS2-decorated graphite composites at 26.44°, 42.30°, 44.48°, and 54.58° were consistent with the (002), (100), (101), and (004) planes of hexagonal graphite (JCPDS 41-1487). Additionally, the (002) diffraction peak shifted to a lower angle compared to that of pristine graphite (26.58°), indicating an increased interlayer spacing (Figure S5). The three-dimensionally aligned few-layer MoS2 nanosheets grown on the surface of graphite are expected to provide more space for Li+ ion transport and reduce the kinetic barriers for their movement. Figure 4b shows an enlarged XRD pattern of the low-angle region in Figure 4a, and the peaks at 14.22° and 33.24° match well with the (002) and (101) planes of hexagonal MoS2 (JCPDS 37-1492). The average interlayer spacing (002) of MoS2 calculated from Bragg’s law is about 0.622 nm, which is slightly longer than that of highly crystalline MoS2 (0.615 nm).
The Raman spectra of MoS2-nanosheets-decorated graphite are presented in Figure 4c, showing two sharp peaks at 380 and 407 cm−1 attributed to the E12g and A1g vibration modes of MoS2, respectively [28]. Two characteristic bands were also observed at 1358 and 1580 cm−1, corresponding to the D band and G band of graphite, respectively.
The XPS measurements were used to analyze the elemental composition of MoS2-decorated graphite. The XPS survey spectra in Figure 4d show that MoS2-decorated graphite contains Mo, S, C, and O elements. In the high-resolution Mo 3d spectrum (Figure 4e), two peaks are observed at 229.7 and 232.8 eV, which can be attributed to the Mo 3d5/2 and Mo 3d3/2 binding energies, respectively, and are characteristic peaks of Mo4+ in MoS2. The peaks at 232.9 and 236.1 eV are related to the Mo 3d5/2 and 3d3/2 of Mo6+ (typical of the Mo-O bond). Furthermore, a small S 2s peak is displayed at 226.7 eV. The high-resolution S 2p spectrum is consistent with S 2p1/2 at 163.6 eV and S 2p3/2 at 162.5 eV, respectively (Figure 4f) [22].
The electrochemical performance of the MoS2-decorated graphite composite was evaluated as an LIB anode material. Figure 5a shows the galvanostatic charge and discharge (GCD) profiles of the 20−MoS2@AG electrode during the initial first cycle in a voltage range of 0.01–2.5 V vs. Li/Li+ at a current density of 35 mA g−1. The 20−MoS2@AG electrode exhibited an initial charge capacity of 494.1 mAh g−1 with a coulombic efficiency of 86.1%. The capacity of 20−MoS2@AG is higher than that of pristine graphite (373.3 mAh g−1) due to the formation of MoS2 nanosheets on the graphite surface. However, the initial coulombic efficiency is slightly lower than that of graphite (~91.2%), indicating that electrochemically active MoS2 nanosheets contribute to the total capacity of the 20−MoS2@AG composite [12,14].
Figure 5b shows the differential voltage profile of the 20−MoS2@AG. In the first cycle, the cathodic peaks at 0.08 V and 0.19 V in the discharge (lithiation) process are attributed to the gradual intercalation of Li+ into the interlayers of graphite [15,29]. Additional discharge contributions corresponding to the MoS2 nanosheets on the graphite surface were observed at higher voltages of ~0.68 and 1.18 V vs. Li/Li+. The peak at 1.18 V is ascribed to Li+ insertion into the structure of MoS2 to form LiXMoS2. Another peak at 0.68 V corresponds to the reduction/decomposition of MoS2 to Mo metal particles and Li2S through a conversion reaction (MoS2 + 4Li+ → Mo + 2Li2S), and the formation of the solid electrolyte interphase (SEI) layers [25,30,31]. MoS2 reacts with Li+ ions at a higher operating voltage compared to graphite. Therefore, the lithiated 20−MoS2@AG electrode can lower the Li+ adsorption energy on the surface of graphite and form a stable interface between MoS2 and graphite, promoting Li+ migration during cycling [12,14,29]. Additionally, the slightly higher anodic potential in the MoS2@AG composite would prevent the formation of lithium dendrites during repeated cycling, which typically occurs in conventional graphite anodes [29]. In the subsequent reverse anodic scan, two peaks at 1.63 V and 2.23 V are associated with the incomplete oxidation of Mo metal into MoS2 and the partial de-lithiation of Li2S into S, respectively. In the next cycle, the reduction peaks at 0.68 V and 1.18 V are disappeared and two new peaks at 1.1 V and 1.95 V are observed, assigned to the following reactions: 2Li+ + S + 2e → Li2S and MoS2 + xLi+ + xe → LixMoS2, respectively [25,32]. These peak shifts could be explained by the structural change and phase transformation during the first cycling [28]. Even at the 100th cycle, it was confirmed that two weak peaks at 1.3 V and 1.88 V were still observed (Figure S6a,b). As shown in Figure S6c,d, the pristine MoS2 electrode shows a similar differential voltage profile to the 20−MoS2@AG electrode. This means that the 20−MoS2@AG composite electrode has an energy storage mechanism combined with those of graphite and MoS2. Moreover, the presence of two main pairs of redox peaks during cycling indicated the stable cycle stability and reversibility of the 20−MoS2@AG electrode.
Figure 5c and Figure S7a show the cycle stabilities of pristine AG, pristine MoS2, 5−MoS2@AG, 10−MoS2@AG, 20−MoS2@AG, and 30−MoS2@AG electrodes, with their initial charge (de-lithiation) capacities at a current density of 200 mA g−1 being 337.1, 812.0, 407.0, 418.0, 448.0, and 523.2 mAh g−1, respectively. The highest specific capacity of the 30-MoS2@AG can be ascribed to the greater loading mass of MoS2. The pristine flower-like MoS2 nanosheets electrode exhibited a rapid capacity decay during initial cycles, likely due to its low conductivity and severe structural damage caused by volume changes, leading to the pulverization of the active material [33]. In contrast, the 20−MoS2@AG composite electrode displayed excellent cycle performance with an average coulombic efficiency of 99.3% and delivered the highest reversible capacity of 463.2 mAh g−1 after 100 cycles. The reversible capacity of the 20−MoS2@AG electrode is much larger than the theoretical capacity of 386.9 mAh g−1 (C20−MoS2@AG = Cgraphite × wt% of graphite (95%) + CMoS2 × wt% of MoS2 (5%) = 372 × 0.95 + 670 × 0.05 = 386.9 mAh g−1). This can be attributed to the formation of 3D hierarchical MoS2 nanosheets on the graphite surface, resulting in abundant active sites for Li+ diffusion, expanded interlayer spacing, and a large electrode/electrolyte contact area [25,31].
The galvanostatic charge and discharge curves of pristine MoS2 and 20−MoS2@AG electrodes at current densities of 35 and 200 mA g−1 are presented in Figure S8. Interestingly, both pristine MoS2 nanosheets and x−MoS2@AG composite electrodes exhibited a capacity climbing phenomenon during cycles, which is likely due to the increasing electrochemically accessible surface area, resulting from the gradual appearance of cracks on the (002) basal planes of MoS2 [33,34,35,36,37].
The rate capabilities of pristine AG, pristine MoS2, 5−MoS2@AG, 10−MoS2@AG, 20−MoS2@AG, and 30−MoS2@AG electrodes were evaluated at current densities of 200, 400, 800, 1200, and 2400 mA g−1 (1C = ~400 mA g−1), as shown in Figure 5d and Figure S7b. Among them, the 20−MoS2@AG electrode demonstrated the best rate performance compared to pristine AG and pristine MoS2. At current densities of 200, 400, 800, 1200, and 2400 mA g−1, the average charge capacities of 20−MoS2@AG delivered reversible capacities of 441.7, 431.4, 413.1, 371.0, and 255.8 mAh g−1, respectively, which were superior to those of pristine MoS2 (699.5, 349.2, 198.8, 139.7, and 84.1 mAh g−1) and pristine AG (365.5, 342.7, 256.8, 172.6, and 92.5 mAh g−1). The 20−MoS2@AG electrode displayed a high reversible capacity of 371.0 mAh g−1 at a high current density of 1200 mA g−1. Notably, the 20−MoS2@AG electrode maintained a high-capacity value even under high-rate conditions, retaining approximately 84.0% of the capacity compared to that at a relatively low current density of 200 mA g−1 (Table S1). Even when the current density recovered to 200 mA g−1 after 50 cycles, the capacity of the 20−MoS2@AG electrode remained at 439.5 mAh g−1, indicating excellent reversibility and rate cycle stability. On the other hand, the 30-MoS2@AG electrode with a higher amount of MoS2 showed more capacity fading as the current density increased (71.0% retention of the capacity at 1200 mA g−1). These results showed that the higher the MoS2 content, the lower the rate characteristics due to its low intrinsic electronic conductivity and large volume change upon cycling. Therefore, it was determined that the 20−MoS2@AG composite was the most optimal condition. Additionally, all MoS2-decorated graphite electrodes exhibited enhanced rate performance and reversibility. This can be attributed to the combination of graphite’s excellent conductivity and the flower-like MoS2 nanosheets grown on its surface, which increases the contact/accessible area with the electrolyte, provides more reaction sites, and facilitates Li+ diffusion, thereby contributing to the excellent electrochemical performance [26].
A long-term cycling performance at a high current density is essential for practical applications of LIBs [25]. Figure 5e illustrates the cycle performance of graphite and 20−MoS2@AG electrodes at a high current density of 1200 mA g−1 over 300 cycles. The 20−MoS2@AG electrode exhibits an initial capacity of ~400 mAh g−1, higher than that of AG (~248 mAh g−1), and shows stable cycle performance over 300 cycles. The significant improvement in the overall performance of x−MoS2@AG composites compared to pristine graphite is mainly due to the surface decoration with MoS2 nanosheets. Furthermore, the 20−MoS2@AG electrode exhibited stable cycle characteristics over 400 cycles at a high current density of 1200 mA g−1, while the 30−MoS2@AG electrode showed a rapid decrease in capacity from around the 315th cycle (Figure S7c).
The electrochemical performance of our MoS2@AG as anode materials of LIBs was compared with those of the previously reported transition metal sulfides (MSx) or transition metal oxides (MOx)-based carbon composites (Table S2) [38,39,40,41,42]. Upon comparing the specific capacity of our MoS2@AG composite anode to that of carbon composites incorporating transition metal sulfides or transition metal oxides, we observed that the capacity of our composite was relatively lower than carbon-based composites containing more than 50 wt% of transition metal sulfides or oxides. However, in comparison to anode materials that contained small amounts (<10 wt%) of transition metal sulfides or oxides, our composite demonstrated superior capacity characteristics under not only low current density conditions, but also high current density conditions. These findings are expected to provide the MoS2@AG composite with a small amount of MoS2 surface modification on the graphite surface as a potential anode material for high-performance and cost-effective lithium-ion batteries.
Electrochemical impedance spectroscopy (EIS) was used to obtain further insight into the electrochemical reaction kinetics of AG, pristine MoS2 nanosheets, and 20−MoS2@AG electrodes (Figure 5f). EIS plots were analyzed based on the fitting results obtained using an equivalent circuit model (inset of Figure 5f). The Nyquist plots of the electrodes can be divided into four components. The intercept at the real axis in the high-frequency region represents the internal resistance (Rs), a semicircle in the high-frequency region corresponds to the surface film resistance (Rf), a semicircle in the medium-frequency region represents charge transfer resistance (Rct), and a straight line in the low-frequency region represents the Warburg impedance (Zw) related to the diffusion resistance of Li+ within the bulk of the electrode materials [15,21,29]. CPE represents a constant phase element, corresponding to the charge-transfer reaction. The values for the resistance parameters obtained from fitting using the equivalent circuit are summarized in Tables S3 and S4. The diameter of the semicircle (Rf + Rct) of the 20−MoS2@AG electrode (128 Ω) in the high/medium frequency is similar to that of AG (101 Ω). Moreover, the slopes of the straight lines at low frequencies for AG and 20−MoS2@AG are larger than those for pristine MoS2 nanosheets, indicating that the MoS2@AG electrode has a faster diffusion rate at the interface. The equivalent circuit fitting result confirms that both Warburg resistance (ZwR (Ω)) and Warburg time constant (ZwT (s)) values of AG and MoS2@AG electrodes are about 2–5 times smaller than those of MoS2. The ion diffusion coefficient is inversely proportional to the inverse ratio values of ZwT and ZwR2 [43]. In addition, the Nyquist plots of the pristine AG, pristine MoS2, and 20−MoS2@AG electrodes were obtained after two cycles at a current density of 35 mA g−1, as shown in Figure S9. The pristine AG and 20−MoS2@AG after two cycles show a decrease in both surface film and charge-transfer resistance (Rf + Rct) showing 101 → 76.3 Ω for AG and 128 → 76.8 Ω for MoS2@AG. This observation is attributed to the improved infiltration of electrolyte into electrode materials and charge transfer kinetics. However, the Rf + Rct of cycled MoS2 electrode increases slightly (128 → 139 Ω), which corresponds mainly to the rapid capacity decay due to the pulverization of active materials.
Examining the changes in surface and thickness of the electrode before and after cycling provides a visual means to understand the mechanism underlying the improved performance of the MoS2 nanosheets decorated on the graphite composite in comparison to pristine graphite. In light of this, we investigated the structural stability of the 20−MoS2@AG electrode before and after cycling by performing SEM measurements, as shown in Figure S10. Prior to cycling, the MoS2@AG electrode had a rough surface covered with numerous nanosheets and its thickness was approximately 36.2 μm. After 300 cycles at a current density of 200 mA g−1, the 3D hierarchically aligned MoS2 nanosheets synthesized on the surface of the graphite had transformed into nanoparticles and aggregated with each other, but some MoS2 nanosheets still maintained their original shape. In addition, we observed that the thickness of the 20−MoS2@AG electrode after cycling was about 36.5 μm, which was almost unchanged compared to the initial thickness of 36.2 μm before cycling. These ex situ SEM images provide evidence that the 20−MoS2@AG electrode maintains a stable structure without significant volume changes even after long-term cycling, indicating that high reversible capacity can be maintained even after prolonged cycling.

4. Conclusions

In conclusion, we have successfully synthesized a 3D hierarchical MoS2@AG composite that exhibits high performance as an anode material for LIBs. The use of graphite as a substrate for the nucleation and growth of three-dimensionally aligned few-layer MoS2 nanosheets provided a conductive matrix and exposed surface area, resulting in a significant improvement in the rate capability and capacity of the composite. The hydrothermal reaction facilitated the growth of MoS2 nanosheets on the graphite surface, forming a 3D network with excellent structural stability during the charge/discharge process. The three-dimensionally grown few-layer MoS2 nanosheets on the surface of graphite facilitated the diffusion of Li+ and reduced the diffusion resistance, leading to improved rate performance. The 20−MoS2@AG electrode exhibited excellent cyclability and rate capability with outstanding stability over 300 cycles, even at a high current density of 1200 mA g−1. Notably, at this current density, the 20−MoS2@AG composite displayed an initial capacity of around 400 mAh g−1, which is approximately 60% higher than that of the graphite electrode. These results demonstrate that the MoS2-nanosheets-decorated graphite composite developed through a simple hydrothermal method holds great promise as a high-power anode material for LIBs in various practical applications.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ma16114016/s1, Figure S1: SEM images of pristine artificial graphite (AG) obtained at (a) low magnification (scale bar: 100 µm) and (b) enlarged magnification (scale bar: 30 µm); Figure S2: SEM images of the MoS2@AG composite obtained at various enlarged magnifications; Figure S3: N2 adsorption and desorption isotherms of pristine graphite (AG) and 20−MoS2@AG composite. The inset shows their pore size distribution curves; Figure S4: TGA curves of pristine MoS2, 5−MoS2@AG, 10−MoS2@AG, and 20−MoS2@AG composites with a heating rate of 10 °C min−1 in air; Figure S5: XRD patterns of MoS2@AG composites; Figure S6: The differential voltage profiles of (a,b) the 20−MoS2@AG electrode and (c,d) pristine MoS2; Figure S7: (a) Cycle performance at a current density of 200 mA g−1, (b) rate performance comparison of pristine AG, pristine MoS2, 20−MoS2@AG, and 30−MoS2@AG electrodes at different current densities ranging from 200 to 2400 mA g−1, and (c) cycle performance at a high current density of 1200 mA g−1 for pristine AG, 20−MoS2@AG, and 30−MoS2@AG electrodes; Figure S8: Galvanostatic charge and discharge curves of (a) pristine MoS2 and (b) 20−MoS2@AG electrodes; Figure S9: Nyquist plots of pristine AG, pristine MoS2, and 20−MoS2@AG obtained after 2 cycles at a current density of 35 mA g−1; Figure S10: Ex situ SEM images of the 20−MoS2@AG electrodes before and after 300 cycles at a current density of 200 mA g−1; (a–c) before cycling and (d–f) after cycling. Panels (c) and (f) of Figure S10 show cross-sectional SEM images of the electrode before and after cycling, respectively; Table S1: Comparison of the initial capacities and rate performance for pristine graphite, pristine MoS2, 5−MoS2@AG, 10−MoS2@AG, 20−MoS2@AG, and 30−MoS2@AG composites; Table S2: Comparison of LIB electrochemical performance for transition metal sulfides (MSx) and transition metal oxides (MOx)-based composites; Table S3: EIS fitting parameters for AG, MoS2, and 20−MoS2@AG before the cycling tests; Table S4: EIS fitting parameters for AG, MoS2, and 20−MoS2@AG after 2 cycles. References [38,39,40,41,42] are cited in the supplementary file.

Author Contributions

Conceptualization, Y.A.L., H.S., K.L. and H.Y.; methodology, Y.A.L., K.Y.J., C.-Y.Y. and H.Y.; investigation, Y.A.L., K.Y.J., J.Y., K.Y., W.J., K.-N.J., J.L. and M.H.R.; resources, H.Y.; data curation, Y.A.L., K.Y.J. and J.Y.; writing—original draft preparation, Y.A.L., K.Y.J., J.Y. and H.Y.; writing—review and editing, Y.A.L., K.Y.J., K.Y., K.-N.J., C.-Y.Y., Y.C., J.L., M.H.R., H.S., K.L. and H.Y.; visualization, Y.A.L.; supervision, H.S., K.L. and H.Y.; project administration, H.Y.; funding acquisition, H.S. and H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by framework of the research and development program of the Korea Institute of Energy Research (KIER, Project No. C3-2408, C3-8601, and C3-2420), the R&D Program for Forest Science Technology (Project No. 2020229B10-2222-AC01) provided by Korea Forest Service (Korea Forestry Promotion Institute), and Korea Institute of Planning and Evaluation for Technology in Food, Agriculture and Forestry (IPET) funded by Ministry of Agriculture, Food and Rural Affairs (MAFRA) (321077022HD020). This work was also supported by research fund of Chungnam National University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available upon request from the corresponding authors.

Acknowledgments

SEM analysis was supported by Go-Woon Lee and TEM analysis was supported by Byung-Sun Ahn, both from the Analytical Center of Energy Research (ACER) at KIER.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xie, J.; Lu, Y.-C. A retrospective on lithium-ion batteries. Nat. Commun. 2020, 11, 2499. [Google Scholar] [CrossRef] [PubMed]
  2. Goodenough, J.B.; Park, K.-S. The Li-Ion Rechargeable Battery: A Perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef]
  3. Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the development of advanced Li-ion batteries: A review. Energy Environ. Sci. 2011, 4, 3243–3262. [Google Scholar] [CrossRef]
  4. Tarascon, J.-M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef]
  5. Liu, Q.; Du, C.; Shen, B.; Zuo, P.; Cheng, X.; Ma, Y.; Yin, G.; Gao, Y. Understanding undesirable anode lithium plating issues in lithium-ion batteries. RSC Adv. 2016, 6, 88683–88700. [Google Scholar] [CrossRef]
  6. Waldmann, T.; Hogg, B.-I.; Wohlfahrt-Mehrens, M. Li plating as unwanted side reaction in commercial Li-ion cells—A review. J. Power Sources 2018, 384, 107–124. [Google Scholar] [CrossRef]
  7. Li, S.; Wang, K.; Zhang, G.; Li, S.; Xu, Y.; Zhang, X.; Zhang, X.; Zheng, S.; Sun, X.; Ma, Y. Fast Charging Anode Materials for Lithium-Ion Batteries: Current Status and Perspectives. Adv. Funct. Mater. 2022, 32, 2200796. [Google Scholar] [CrossRef]
  8. Li, L.; Zhang, D.; Deng, J.; Gou, Y.; Fang, J.; Cui, H.; Zhao, Y.; Cao, M. Carbon-based materials for fast charging lithium-ion batteries. Carbon 2021, 183, 721–734. [Google Scholar] [CrossRef]
  9. Cheng, Q.; Yuge, R.; Nakahara, K.; Tamura, N.; Miyamoto, S. KOH etched graphite for fast chargeable lithium-ion batteries. J. Power Sources 2015, 284, 258–263. [Google Scholar] [CrossRef]
  10. Chen, K.-H.; Goel, V.; Namkoong, M.J.; Wied, M.; Müller, S.; Wood, V.; Sakamoto, J.; Thornton, K.; Dasgupta, N.P. Enabling 6C Fast Charging of Li-Ion Batteries with Graphite/Hard Carbon Hybrid Anodes. Adv. Energy Mater. 2021, 11, 2003336. [Google Scholar] [CrossRef]
  11. Kim, N.; Chae, S.; Ma, J.; Ko, M.; Cho, J. Fast-charging high-energy lithium-ion batteries via implantation of amorphous silicon nanolayer in edge-plane activated graphite anodes. Nat. Commun. 2017, 8, 812. [Google Scholar] [CrossRef]
  12. Rhee, D.Y.; Kim, J.; Moon, J.; Park, M.S. Off-stoichiometric TiO2-x-decorated graphite anode for high-power lithium-ion batteries. J. Alloys Compd. 2020, 843, 156042. [Google Scholar] [CrossRef]
  13. Lee, S.-M.; Kim, J.; Moon, J.; Jung, K.-N.; Kim, J.H.; Park, G.-J.; Choi, J.-H.; Rhee, D.Y.; Kim, J.-S.; Park, M.-S. A cooperative biphasic MoOx–MoPx promoter enables a fast-charging lithium-ion battery. Nat. Commun. 2021, 12, 39. [Google Scholar] [CrossRef] [PubMed]
  14. Lee, J.W.; Kim, S.Y.; Rhee, D.Y.; Park, S.; Jung, J.Y.; Park, M.-S. Tailoring the Surface of Natural Graphite with Functional Metal Oxides via Facile Crystallization for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2022, 14, 29797–29805. [Google Scholar] [CrossRef] [PubMed]
  15. Yan, X.; Jiang, F.; Sun, X.; Du, R.; Zhang, M.; Kang, L.; Han, Q.; Du, W.; You, D.; Zhou, Y. A simple, low-cost and scale-up synthesis strategy of spherical-graphite/Fe2O3 composites as high-performance anode materials for half/full lithium ion batteries. J. Alloys Compd. 2020, 822, 153719. [Google Scholar] [CrossRef]
  16. Song, Y.; Wang, Z.; Yan, Y.; Zhao, W.; Bakenov, Z. NiCo2S4 nanoparticles embedded in nitrogen-doped carbon nanotubes networks as effective sulfur carriers for advanced Lithium–Sulfur batteries. Micro. Meso. Mater. 2021, 316, 110924. [Google Scholar] [CrossRef]
  17. Qi, J.-Q.; Huang, M.-Y.; Ruan, C.-Y.; Zhu, D.-D.; Zhu, L.; Wei, F.-X.; Sui, Y.-W.; Meng, Q.-K. Construction of CoNi2S4 nanocubes interlinked by few-layer Ti3C2Tx MXene with high performance for asymmetric supercapacitors. Rare Metals 2022, 41, 4116–4126. [Google Scholar] [CrossRef]
  18. Xiao, Y.; Shen, Y.; Su, D.; Zhang, S.; Yang, J.; Yan, D.; Fang, S.; Wang, X. Engineering Cu1.96S/Co9S8 with sulfur vacancy and heterostructure as an efficient bifunctional electrocatalyst for water splitting. J. Mater. Sci. Technol. 2023, 154, 1–8. [Google Scholar] [CrossRef]
  19. Yu, S.-H.; Zachman, M.J.; Kang, K.; Gao, H.; Huang, X.; Disalvo, F.J.; Park, J.; Kourkoutis, L.F.; Abruña, H.D. Atomic-Scale Visualization of Electrochemical Lithiation Processes in Monolayer MoS2 by Cryogenic Electron Microscopy. Adv. Energy Mater. 2019, 9, 1902773. [Google Scholar] [CrossRef]
  20. Zhan, J.; Wu, K.; Yu, X.; Yang, M.; Cao, X.; Lei, B.; Pan, D.; Jiang, H.; Wu, M. α-Fe2O3 Nanoparticles Decorated C@MoS2 Nanosheet Arrays with Expanded Spacing of (002) Plane for Ultrafast and High Li/Na-Ion Storage. Small 2019, 15, 1901083. [Google Scholar] [CrossRef]
  21. Tian, H.; Yu, M.; Liu, X.; Qian, J.; Qian, W.; Chen, Z.; Wu, Z. Plant-cell oriented few-layer MoS2/C as high performance anodes for lithium-ion batteries. Electrochim. Acta 2022, 424, 140685. [Google Scholar] [CrossRef]
  22. Wang, G.; Zhang, J.; Yang, S.; Wang, F.; Zhuang, X.; Müllen, K.; Feng, X. Vertically Aligned MoS2 Nanosheets Patterned on Electrochemically Exfoliated Graphene for High-Performance Lithium and Sodium Storage. Adv. Energy Mater. 2018, 8, 17702254. [Google Scholar]
  23. Tiwari, A.P.; Yoo, H.; Lee, J.; Kim, D.; Park, J.H.; Lee, H. Prevention of sulfur diffusion using MoS2-intercalated 3D-nanostructured graphite for high-performance lithium-ion batteries. Nanoscale 2015, 7, 11928–11933. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, B.; Zhang, Y.; Zhang, J.; Xia, R.; Chu, Y.; Zhou, J.; Yang, X.; Huang, J. Facile Synthesis of a MoS2 and Functionalized Graphene Heterostructure for Enhanced Lithium-Storage Performance. ACS Appl. Mater. Interfaces 2017, 9, 12907–12913. [Google Scholar] [CrossRef] [PubMed]
  25. Zhen, M.; Wang, J.; Guo, S.-Q.; Shen, B. Vertically aligned nanosheets with MoS2/N-doped-carbon interfaces enhance lithium-ion storage. Appl. Surf. Sci. 2019, 487, 285–294. [Google Scholar] [CrossRef]
  26. Hu, S.; Chen, W.; Zhou, J.; Yin, F.; Uchaker, E.; Zhang, Q.; Cao, G. Preparation of carbon coated MoS2 flower-like nanostructure with self-assembled nanosheets as high-performance lithium-ion battery anodes. J. Mater. Chem. A 2014, 2, 7862–7872. [Google Scholar] [CrossRef]
  27. Hu, Z.; Wang, L.; Zhang, K.; Wang, J.; Cheng, F.; Tao, Z.; Chen, J. MoS2 Nanoflowers with Expanded Interlayers as High-Performance Anodes for Sodium-Ion Batteries. Angew. Chem. Int. Ed. 2014, 53, 12794–12798. [Google Scholar] [CrossRef]
  28. Jiao, Y.; Mukhopadhyay, A.; Ma, Y.; Yang, L.; Hafez, A.M.; Zhu, H. Ion Transport Nanotube Assembled with Vertically Aligned Metallic MoS2 for High Rate Lithium-Ion Batteries. Adv. Energy Mater. 2018, 8, 1702779. [Google Scholar] [CrossRef]
  29. Xiang, F.; Hou, W.; Gu, X.; Wen, L.; Sun, Y.; Lu, W. One-pot synthesis of MnO-loaded mildly expanded graphite composites as high-performance lithium-ion battery anode materials. J. Alloys Compd. 2022, 897, 163202. [Google Scholar] [CrossRef]
  30. Hu, L.; Ren, Y.; Yang, H.; Xu, Q. Fabrication of 3D Hierarchical MoS2/Polyaniline and MoS2/C Architectures for Lithium-Ion Battery Applications. ACS Appl. Mater. Interfaces 2014, 6, 14644–14652. [Google Scholar] [CrossRef]
  31. Zhao, J.; Ren, H.; Gu, C.; Guan, W.; Song, X.; Huang, J. Synthesis of hierarchical molybdenum disulfide microplates consisting of numerous crosslinked nanosheets for lithium-ion batteries. J. Alloys Compd. 2019, 781, 174–185. [Google Scholar] [CrossRef]
  32. Luo, X.; Li, N.; Guo, X.; Wu, K. One-pot hydrothermal synthesis of MoS2 anchored corncob-derived carbon nanospheres for use as a high-capacity anode for reversible Li-ion battery. J. Solid State Chem. 2021, 296, 122020. [Google Scholar] [CrossRef]
  33. Teng, Y.; Zhao, H.; Zhang, Z.; Li, Z.; Xia, Q.; Zhang, Y.; Zhao, L.; Du, X.; Du, Z.; Lv, P.; et al. MoS2 Nanosheets Vertically Grown on Graphene Sheets for Lithium-Ion Battery Anodes. ACS Nano 2016, 10, 8526–8535. [Google Scholar] [CrossRef]
  34. Liu, M.; Li, N.; Wang, S.; Li, Y.; Liang, C.; Yu, K. 3D nanoflower-like MoS2 grown on wheat straw cellulose carbon for lithium-ion battery anode material. J. Alloys Compd. 2023, 933, 167689. [Google Scholar] [CrossRef]
  35. Liu, Y.; Zhang, L.; Zhao, Y.; Shen, T.; Yan, X.; Yu, C.; Wang, H.; Zeng, H. Novel plasma-engineered MoS2 nanosheets for superior lithium-ion batteries. J. Alloys Compd. 2019, 787, 996–1003. [Google Scholar] [CrossRef]
  36. Zhang, L.; Fan, W.; Tjiu, W.W.; Liu, T. 3D porous hybrids of defect-rich MoS2/graphene nanosheets with excellent electrochemical performance as anode materials for lithium ion batteries. RSC Adv. 2015, 5, 34777. [Google Scholar] [CrossRef]
  37. Chang, K.; Geng, D.; Li, X.; Yang, J.; Tang, Y.; Cai, M.; Li, R.; Sun, X. Ultrathin MoS2 /Nitrogen-Doped Graphene Nanosheets with Highly Reversible Lithium Storage. Adv. Energy Mater. 2013, 3, 839. [Google Scholar] [CrossRef]
  38. Gao, M.; Liu, B.; Zhang, X.; Zhang, Y.; Li, X.; Han, G. Ultrathin MoS2 nanosheets anchored on carbon nanofibers as free-standing flexible anode with stable lithium storage performance. J. Alloys Compd. 2022, 894, 162550. [Google Scholar] [CrossRef]
  39. Sun, H.; Wang, J.-G.; Zhang, X.; Li, C.; Liu, F.; Zhu, W.; Hua, W.; Li, Y.; Shao, M. Nanoconfined Construction of MoS2@C/MoS2 Core−Sheath Nanowires for Superior Rate and Durable Li-Ion Energy Storage. ACS Sustain. Chem. Eng. 2019, 7, 5346–5354. [Google Scholar] [CrossRef]
  40. Kim, H.; Nguyen, Q.H.; Kim, I.T.; Hur, J. Scalable synthesis of high-performance molybdenum diselenide-graphite nanocomposite anodes for lithium-ion batteries. Appl. Surf. Sci. 2019, 481, 1196–1205. [Google Scholar] [CrossRef]
  41. Liu, X.; Wang, Q.; Zhang, J.; Guan, H.; Zhang, C. One-Step Preparation of MoS2/Graphene Nanosheets via Solid-State Pan-Milling for High Rate Lithium-Ion Batteries. Ind. Eng. Chem. Res. 2020, 59, 16240–16248. [Google Scholar] [CrossRef]
  42. Choi, J.H.; Kim, M.-C.; Moon, S.-H.; Kim, H.; Kim, Y.-S.; Park, K.-W. Enhanced electrochemical performance of MoS2/graphene nanosheet nanocomposites. RSC Adv. 2020, 10, 19077. [Google Scholar] [CrossRef] [PubMed]
  43. Yoon, H.; Kim, H.-J.; Yoo, J.J.; Yoo, C.-Y.; Park, J.H.; Lee, Y.A.; Cho, W.K.; Han, Y.-K.; Kim, D.H. Pseudocapacitive slurry electrodes using redox-active quinone for high-performance flow capacitors: An atomic-level understanding of pore texture and capacitance enhancement. J. Mater. Chem. A 2015, 3, 23323–23332. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic illustration and microstructural analysis of one-step hydrothermal synthesis of the 3D hierarchically aligned flower-like MoS2 nanosheets grown directly on the surface of graphite. The MoS2@AG composite was formed through spontaneous nucleation and growth of few-layer MoS2 nanosheets on the surface of graphite. Representative images of the composite obtained by (b) SEM, (c) TEM, and (d) high-resolution TEM (HR-TEM), revealing the morphology and structure of the composite at different scales.
Figure 1. (a) Schematic illustration and microstructural analysis of one-step hydrothermal synthesis of the 3D hierarchically aligned flower-like MoS2 nanosheets grown directly on the surface of graphite. The MoS2@AG composite was formed through spontaneous nucleation and growth of few-layer MoS2 nanosheets on the surface of graphite. Representative images of the composite obtained by (b) SEM, (c) TEM, and (d) high-resolution TEM (HR-TEM), revealing the morphology and structure of the composite at different scales.
Materials 16 04016 g001
Figure 2. SEM and SEM-BSE images of MoS2-nanosheets-decorated graphite composites with varying amounts of thiourea and (NH4)6Mo7O24⋅4H2O: (ac) 2.5−MoS2@AG, (df) 5−MoS2@AG, (gi) 10−MoS2@AG, and (jl) 20−MoS2@AG. The images show vertically aligned plate-like MoS2 nanosheets on 2.5−MoS2@AG and 5−sMoS2@AG composites, and flower-shaped MoS2 nanosheets on 10−MoS2@AG and 20−MoS2@AG composites. The SEM-BSE images in (c,f,i,l) confirm the successful formation of MoS2 nanosheets on the graphite surface by brighter contrast indicating a higher atomic number.
Figure 2. SEM and SEM-BSE images of MoS2-nanosheets-decorated graphite composites with varying amounts of thiourea and (NH4)6Mo7O24⋅4H2O: (ac) 2.5−MoS2@AG, (df) 5−MoS2@AG, (gi) 10−MoS2@AG, and (jl) 20−MoS2@AG. The images show vertically aligned plate-like MoS2 nanosheets on 2.5−MoS2@AG and 5−sMoS2@AG composites, and flower-shaped MoS2 nanosheets on 10−MoS2@AG and 20−MoS2@AG composites. The SEM-BSE images in (c,f,i,l) confirm the successful formation of MoS2 nanosheets on the graphite surface by brighter contrast indicating a higher atomic number.
Materials 16 04016 g002
Figure 3. TEM and STEM images of MoS2-nanosheets-decorated graphite composites. (a) Low-magnification TEM image showing 3D architectural MoS2 nanosheets formed on the graphite surface. (b,c) HR-TEM images displaying few-layered MoS2 nanosheets grown on the graphite surface. (d) STEM-HAADF image and (eg) TEM-EDS mapping images indicating the distribution of C, Mo, and S elements in the composite.
Figure 3. TEM and STEM images of MoS2-nanosheets-decorated graphite composites. (a) Low-magnification TEM image showing 3D architectural MoS2 nanosheets formed on the graphite surface. (b,c) HR-TEM images displaying few-layered MoS2 nanosheets grown on the graphite surface. (d) STEM-HAADF image and (eg) TEM-EDS mapping images indicating the distribution of C, Mo, and S elements in the composite.
Materials 16 04016 g003
Figure 4. (a,b) XRD patterns, (c) Raman spectra, (d) XPS survey spectrum, and high-resolution XPS spectra of (e) Mo 3d and (f) S 2p for MoS2@AG composites.
Figure 4. (a,b) XRD patterns, (c) Raman spectra, (d) XPS survey spectrum, and high-resolution XPS spectra of (e) Mo 3d and (f) S 2p for MoS2@AG composites.
Materials 16 04016 g004
Figure 5. (a) Initial galvanostatic charge and discharge curves at a current density of 35 mA g−1 and (b) corresponding differential voltage profiles of the 20−MoS2@AG electrode, (c) cycle performance at a current density of 200 mA g−1, (d) rate performance comparison of pristine AG, pristine MoS2, and MoS2@AG composite electrodes at different current densities ranging from 200 to 2400 mA g−1, (e) cycle performance at a high current density of 1200 mA g−1 for pristine AG and 20−MoS2@AG electrodes, and (f) Nyquist plots from EIS data (symbols) and fitting results (solid lines) of pristine AG, pristine MoS2, and 20−MoS2@AG electrodes before the cycle.
Figure 5. (a) Initial galvanostatic charge and discharge curves at a current density of 35 mA g−1 and (b) corresponding differential voltage profiles of the 20−MoS2@AG electrode, (c) cycle performance at a current density of 200 mA g−1, (d) rate performance comparison of pristine AG, pristine MoS2, and MoS2@AG composite electrodes at different current densities ranging from 200 to 2400 mA g−1, (e) cycle performance at a high current density of 1200 mA g−1 for pristine AG and 20−MoS2@AG electrodes, and (f) Nyquist plots from EIS data (symbols) and fitting results (solid lines) of pristine AG, pristine MoS2, and 20−MoS2@AG electrodes before the cycle.
Materials 16 04016 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

Lee, Y.A.; Jang, K.Y.; Yoo, J.; Yim, K.; Jung, W.; Jung, K.-N.; Yoo, C.-Y.; Cho, Y.; Lee, J.; Ryu, M.H.; et al. Three-Dimensional Flower-like MoS2 Nanosheets Grown on Graphite as High-Performance Anode Materials for Fast-Charging Lithium-Ion Batteries. Materials 2023, 16, 4016. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16114016

AMA Style

Lee YA, Jang KY, Yoo J, Yim K, Jung W, Jung K-N, Yoo C-Y, Cho Y, Lee J, Ryu MH, et al. Three-Dimensional Flower-like MoS2 Nanosheets Grown on Graphite as High-Performance Anode Materials for Fast-Charging Lithium-Ion Batteries. Materials. 2023; 16(11):4016. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16114016

Chicago/Turabian Style

Lee, Yeong A., Kyu Yeon Jang, Jaeseop Yoo, Kanghoon Yim, Wonzee Jung, Kyu-Nam Jung, Chung-Yul Yoo, Younghyun Cho, Jinhong Lee, Myung Hyun Ryu, and et al. 2023. "Three-Dimensional Flower-like MoS2 Nanosheets Grown on Graphite as High-Performance Anode Materials for Fast-Charging Lithium-Ion Batteries" Materials 16, no. 11: 4016. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16114016

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