Next Article in Journal
State-of-Health Prediction of Lithium-Ion Batteries Based on CNN-BiLSTM-AM
Next Article in Special Issue
Effects of Different Charging Currents and Temperatures on the Voltage Plateau Behavior of Li-Ion Batteries
Previous Article in Journal
On the Specific Capacity and Cycle Stability of Si@void@C Anodes: Effects of Particle Size and Charge/Discharge Protocol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Step Hydrothermal Reaction Induced Nitrogen-Doped MoS2/MXene Composites with Superior Lithium-Ion Storage

1
Joint International Laboratory on Environmental and Energy Frontier Materials, School of Environmental and Chemical Engineering, Shanghai University, Shanghai 200444, China
2
Centre for Clean Energy Technology, Faculty of Science, University of Technology Sydney, Broadway, Sydney, NSW 2007, Australia
*
Authors to whom correspondence should be addressed.
Batteries 2022, 8(10), 156; https://doi.org/10.3390/batteries8100156
Submission received: 15 August 2022 / Revised: 7 September 2022 / Accepted: 29 September 2022 / Published: 3 October 2022
(This article belongs to the Special Issue Lithium Batteries: Theory, Design, and Applications)

Abstract

:
MoS2, a typical transition metal dichalcogenide (TMDs), inheriting high theoretical capacity, open framework and unique electrochemical properties, is regarded as a promising electrode material. However, the low electronic conductivity and slow chemical kinetics of two-dimensional (2D) MoS2 lamellars, along with the large volume expansion during cycling hinder their application in Li-ion batteries. MXene inherits the strengths of excellent metallic conductivity, a low lithium-ion diffusion potential barrier and superior mechanical stability; however, its low reversible capacity and self-stacking problems as anode still need to be solved. Herein, the MXene Ti3C2Tx compound with MoS2 through a simple one-step hydrothermal reaction is introduced. The introduction of nitrogen-doped Ti3C2Tx can effectively restrain the volume change of MoS2 and ameliorate the electronic conductivity of the whole electrode, while MoS2 can alleviate the self-stacking of Ti3C2Tx during cycling. The as-prepared MoS2/MXene electrode delivers an initial discharge capacity of 1087 mA h g−1 with an initial Coulombic efficiency (ICE) of 81.6% at 100 mA g−1, and a specific discharge capacity of 731 mA h g−1 can be retained after 100 cycles. The excellent electrochemical performance demonstrates that nitrogen-doped MoS2/MXene can be a potential electrode material for Li-ion batteries.

1. Introduction

With the increasing energy demand, the exploration of new energy storage materials is pushing forward as well. Lithium-ion batteries (LIBs), as dominant power sources for portable electronic devices and electric/hybrid vehicles, are urgently required to develop a new generation of electrode materials to meet the growing demand. Anode materials play an important role in the practical application of lithium-ion batteries. Among the various anode materials, commercialized carbon materials have low specific capacity and poor compatibility with electrolytes, which limits their application in power battery materials, while other anode materials, including metal oxides (MOs), metal sulfides (MSs) and many other alloy-based anode materials (Sn, Si and Ti) with high specific capacities are good candidates for lithium storage [1,2,3,4,5,6].
Since the discovery of graphene in 2004, 2D materials composed of a single layer of atoms have caught significant concern. They possess unique electrical, optical as well as mechanical properties, such as high electroconductivity, good flexibility and strength, which are promising characteristics for the usage of energy storage systems. MXenes, as a new type of two-dimensional material, have attracted more and more attention as electrode material for energy storage. MXenes are 2D transition metal carbides or nitrides (Mn+1XnTx) deriving from ternary MAX phase (Mn+1AXn) in which M represents the early transition metals (Sc, Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, etc.); A symbolizes mainly group IIIA elements; n = 1–3; X is nitrogen and/or carbon; and Tx represents the surface terminations (OH, O or F) [7]. The metallic nature of the transition metal (e.g., Ti) atoms implanted between the layers of carbon atoms endows MXenes with excellent electroconductivity [8,9,10,11,12]. Additionally, it is facile for polar organic molecules and metal ions to intercalate into its host structure because of the open channels provided by MXenes, thus forming composites with other materials to enhance their energy storage properties. However, their application prospects still need to be further explored due to the shortcomings of easy oxidation, low reversible capacity and self-stacking.
Transition metal dichalcogenides (TMDs) with graphene-like layer structure and excellent physicochemical properties, such as ultra-thin layer spacing, huge specific surface area, abundant edge sites, good chemical stability and adjustable forbidden band width, have drawn wide attention in the fields of optoelectronic devices, photocatalysts, energy storage and supercapacitors, etc. Compared to their metal oxide (MOs) counterparts, TMDs are expected to exhibit better mechanical stability due to their smaller volume expansion, higher initial Coulomb efficiency (ICE) and great reversibility [13,14,15,16]. Furthermore, the discharge product of MSs (Li2S) has better electrical conductivity than that of MOs (Li2O), and the M–S bond in MSs is weaker than the M–O bond in MOs, which may be kinetically favorable for the conversion reaction [17]. As a result, more and more attention is being focused on the design and study of the electrochemical properties of MSs. As a representative of transition metal dichalcogenides, molybdenum disulfide (MoS2) with a layered S-Mo-S sandwich structure possesses abundant ion storage sites [18,19,20] and high lithium storage capacity (~670 mA h g−1) [21], all of which are regarded as superior properties for energy storage application. However, the low conductivity, huge volume changes and morphology collapse of MoS2 during electrochemical process hinder its application in practical energy storage. Interfacial modulation and combination with conductive substrates are effective strategies to optimize the electronic properties of MoS2 [22,23,24]. In addition, construction MoS2-based composite structures or introducing heteroatomic doping (N, S, P, etc.) to MoS2 can ameliorate the chemical reaction kinetics through interlayer interactions and increase the conductivity of MoS2 [25,26,27,28]. Chen et al. [29] synthesized 2D MoS2-on-MXene heterostructures as anode materials for lithium-ion batteries by in situ sulfidation of Mo2TiC2Tx MXene. As electrode for LIBs, a specific capacity of 509 mA h g−1 still can be realized after 100 cycles at a current density of 100 mA g−1. Hu et al. [30] prepared MoS2/MXene hybrids using freeze-drying and a subsequent thermal process. When it is the electrode for lithium-ion batteries, it remains at a capacity of 835.1 mA h g−1 after 110 cycles at 0.5 A g−1. However, these methods require harsh reaction conditions. Moreover, the synthetic steps are not particularly concise. In addition, many enlightening works have been conducted by researchers in heteroatom doping. Xu et al. [31] fabricated series 2D MoS2 @N-doped graphene composites by a straightforward solvothermal strategy. When as the anode for sodium ion batteries, a reversible specific capacity of 141 mA h g−1 under 12.8 A g−1 can be obtained. Li et al. [32] obtained MoS2/nitrogen-doped carbon nanorods by a polyethyleneimine (PEI)-guided hydrothermal method and subsequent calcination process. The corresponding electrode shows a reversible capacity of 397.3 mA h g−1 after 100 cycles at 0.1 A for sodium-ion batteries.
Based on the above experience, we generated MoS2 in situ grown on the MXene layers (MoS2/MXene) by using a facile hydrothermal reaction. On the one hand, the highly conductive network of the as-prepared MoS2/MXene composite can slow down the volume expansion of MoS2 during the cycling process, facilitate the diffusion of electrons and ions, and improves the electrochemical properties of the entire electrode. On the other hand, MoS2 can inhibit the re-stacking of MXene, and thus the two complement each other to make up for each other’s deficiencies. This synergistic effect and the nitrogen doping endow the MoS2/MXene composite with strong mechanical stability and electrochemical properties. Meanwhile, the as-prepared stable 3D conductive network can suppress the volume change of MoS2 during cycling. Based on all of the above advantages, excellent electrochemical performance can be anticipated. The obtained MoS2/MXene electrode exhibit excellent cycling stability: the initial charging/ discharging capacities of MoS2/MXene electrode at 100 mA g−1 are 887 and 1087 mA h g−1, respectively, with ICE of 81.6%. Meanwhile, a discharge capacity of 731 mA h g−1 with Coulombic efficiency of 99% can be maintained after 100 cycles.

2. Materials and Methods

The source/manufacturer of materials used in the study are shown in Table S1.

2.1. Preparation of MAX Precursor (Ti3AlC2)

Ti3AlC2 solid particles were fabricated by the molten salt technique. TiC, Ti, Al, NaCl and KCl were firstly added to the mortar in the molar ratio of 2:1:1.1:4:4 and ground for 30 min and then transferred it to a furnace calcinated at 1100 °C for 3 h in Ar atmosphere with a heating rate of 4 °C min−1. The reaction products were ground for 1 h to a fine powder at room temperature, and then the residual NaCl and KCl were removed by deionized water and dried overnight (80 °C). The final MAX precursor was stored in a desiccator before use.

2.2. Preparation of MXene from MAX Precursor (f-Ti3C2Tx)

Briefly, 2 g LiF was added to 20 mL (9 M) of hydrochloric acid under stirring for 30 min, followed by the slow addition of 2 g MAX precursor. The etching process occurred in a poly tetra fluoroethylene (PTFE) reactor for 24 h at the temperature of 37 °C with a controlled speed (400 rpm) and washed with deionized water (4000 rpm, 10 min) until the pH > 6. Then, ethanol pre-infused with argon was added for ultrasonic intercalation for 1 h. The precipitate was then added to the deionized water for centrifugation, and the upper layer of solution was stored at 5 °C.

2.3. Preparation of MoS2/MXene

A certain mass of MXene was dispersed in 20 mL of ethylene glycol (EG) and labeled as solution A. A certain mass (20, 60, 80, 100 mg) of MoCl5 was dispersed in 20 mL of EG labeled as solution B, which was added to solution A with vigorous stirring and protected by argon gas and stirred for 24 h. Thioacetamide (TAA) with equal mass of MoCl5 was added to above reaction solution and then stirred for 10 min. After that, it was transferred to Teflon-lined reactor at 160 °C for 12 h. When the reactor cooled at room temperature, the precipitate was collected by filtration. The precipitate was washed by deionized water and ethanol, and the product was dried overnight at 80 °C under vacuum. Finally, the dried samples were transferred to a tube reactor filled with argon and calcined at 220 °C for 4 h (at a heating rate of 5 °C min−1). At the same time, the N element released by the decomposition of thioacetamide was doped into the MoS2/MXene framework during the reaction. For comparison, the products with different mass ratios of MoCl5 and MXene were prepared (MoS2/MXene-x:1 (x = 2, 3, 4, 5). Note: The MoS2/MXene-4:1 marked as (MoS2/MXene), and the samples without MXene were also prepared (MoS2).

2.4. Characterization

The materials synthesized in this article were studied by scanning electron microscope (SEM) and transmission electron microscope (TEM) for their microscopic morphological characteristics and elemental distribution. X-ray diffraction (XRD) was used to characterize the components and crystal structure of the materials by collecting signals with 2θ range of 3–70 ° using a Cu-Kα radiation source (wavelength 0.154 nm). The nitrogen adsorption/desorption isotherms curves and the pore-size distribution of materials were determined by means of BET tests (BET is an acronym for three scientists (Brunauer, Emmett and Teller)). Fourier transform infrared spectroscopy (FTIR) was used to qualitatively analyze the material and complement the structure of the material. X-ray photoelectron spectroscopy (XPS) analyzed the chemical valence of the material surface. Inductively coupled plasma mass spectrometry (ICP-MS) quantified the proportion of elements in the synthesized material.

2.5. Electrochemical Measurement

The materials in this article were subsequently tested by assembling as CR 2032 button half-cells in Ar-filled glove box. The active material, acetylene black and Carboxymethyl Cellulose (CMC) were blended in deionized water at a mass ratio of 7:2:1 and applied to the copper foil and then dried under vacuum at 80 °C overnight. The average mass loading of the electrodes is about 1 mg cm−2. Celgard-2400 separator was utilized to separate electrodes, and a fresh lithium sheet was applied as the counter electrode. The electrolyte was 1 M LiPF6 dissolved in ethylene carbonate (EC) + diethyl carbonate (DEC) (1:1 Vol%). The voltage range was 0.01–3.0 V. The cycling and rate performance as well as Galvanostatic Intermittent Titration Technique (GITT) test were studied in the NEWARE battery test system. CHI 660E was used to test the cyclic voltammetry (CV) curve and electrochemical impedance spectroscopy (EIS) impedance analysis.

3. Results and Discussion

The schematic fabrication process of the MoS2/MXene material is illustrated in Figure 1a. Firstly, the MAX phase precursor (Ti3AlC2) was prepared by molten salt method and chemically etching Al in Ti3AlC2 to prepare Ti3C2Tx MXene nanosheets. The Ti3AlC2 shows typical micron-sized powders [33]. After etching treatment, the obtained MXene shows several thin flakes (Figures S1 and S2). Meanwhile, MoS2/MXene composites were obtained by a typical solvothermal reaction, in which MoS2 was in situ generated on the surface of MXene, and the N element released from decomposition of thioacetamide (TAA) can be doped into the MoS2/MXene framework during the reaction process. Figure 1b shows the SEM image of the pristine MoS2, which presents nanomicrosphere structures (100–200 nm) with tiny protrusions on the surface [34,35]. Figure 1c displays high-resolution transmission electron microscope image of the as-prepared MXene, and it can be seen that the layer spacing (d = 1.32 nm) of the few-layer MXene is identified with that previously reported work, indicating the successful synthesis of the material. Figure 1d exhibits the SEM image of the synthesized MoS2/MXene. After the introduction of MXene, the MoS2 nanomicrospheres were uniformly grown on the surface MXene, which offers a good transport pathway for electrons and ions. On the other hand, the layered lamellar structure of MXene wraps around MoS2, which can suppress the volume expansion of MoS2 during cycling process. Figure 1e–f displays the TEM image of MoS2/MXene, and we can observe that MXene tightly wraps around MoS2 to form a stable composite structure, and obvious folds can be observed around MoS2, implying that MoS2 is not simply attached to the MXene surface, and the inset selected area electron diffraction (SAED) shows obvious diffraction loops, further proving the presence of MoS2. Notably, the MXene surface of the composite is not smooth compared to the pure MXene sheet, which may be due to the slight oxidation of the MXene surface. Four layers of MXene can be observed by high resolution transmission electron microscope (HRTEM) in Figure 1g, and their layer spacing (d = 1.48 nm) is increased compared to the 1.32 nm of the less-layered MXene in Figure 1d, indicating the existence of ions insert between the MXene layers during the solvothermal process, which is beneficial for lithium storage. Figure 1h–l shows mapping images of MoS2/MXene. The distribution of Mo and S elements, and the surrounding uniformly distributed Ti, C and O elements correspond to MXene can be observed; meanwhile, the N element can be observed, which originates from the decomposition of thioacetamide.
Figure 2a illustrates the XRD plots of MoS2/MXene composites with different ratios, showing that the different MoS2 content does not significantly affect the XRD, and the (003) crystal plane of MoS2 at 14.4° is still retained. In addition, the (002) crystallographic plane of MXene in the composite is biased to a smaller angle, indicating an increase in the MXene layer spacing, which is identified with the TEM plots. Notably, the peak at around 2θ = 25° corresponds to TiO2, indicating a slight oxidation of MXene during the hydrothermal reaction. Meanwhile, the synthesized MAX and MoS2 are identified with the standard card, indicating the successful synthesis of them, and the (002) plane of the prepared MXene is shifted from 9.38° to a smaller 6.52° after etching, implying an expanded layer spacing (d = 1.36 nm) for the layered MXene calculated by the Bragg equation, which matches the HRTEM image in Figure 1c (Figures S3 and S4). Figure S5 and Figure 2b show Raman spectra of the MXene, MoS2 and MoS2/MXene composites. As for the Raman spectrum of MXene (Figure S5), the region between 230 and 470 cm−1 indicates the in-plane (Eg) vibrations of the surface groups attached to the titanium atoms; the region between 580 and 730 cm−1 can be assigned mainly to carbon vibrations (Eg and A1g), which further confirms the successful etching of MXene. As for MoS2 and MoS2/MXene composites (Figure 2b), two typical peaks at around 400 cm−1 belong to the E12g and A1g vibrational modes of MoS2 [36]. The distinct D band and G band can be observed around 1350 cm−1 and 1550 cm−1, which belong to the defects of the carbon-based material and the vibrations of the sp2 hybridized carbon atoms, respectively. Clear shifts can be detected of the D band and G band of the MoS2/MXene composites towards the higher energy regions, demonstrating the recovery of the conjugate structure after the nitrogen atoms incorporated into the lattice of MXene [37]. In addition, the intensity ratios (ID/IG) of MoS2 and MoS2/MXene are 0.86 and 0.92, respectively, suggesting the existence of a great number of defects and vacancies in the MoS2/MXene material, which facilitates the lithium-ion diffusion and also provides sufficient space for the storage of lithium ions. The FTIR spectra of MXene/MoS2 are presented in Figure S6, the peaks near 580 cm−1 and 1100 cm−1 could be assigned to the stretching vibrations of Ti–C and C–O bonds, respectively, caused by the characteristic structure of Ti3C2. In addition, the peaks at 700–900 cm−1 are attributed to the Ti–O–Mo bond. Additionally, the peak near 615 cm−1 corresponds to the vibrational absorption peak of the S–S bond of MoS2. It can be concluded that MoS2 behaves in a strong interaction with MXene and realizes stable constructions of the as-prepared MXene/MoS2 composite. Furthermore, the chemical valence state of the MoS2/MXene material surface was investigated by means of XPS. From the full spectrum (Figure 2c), the elements of O, Ti, Mo, C and S elements can be detected, and trace amounts of N elements are also observed. Additionally, as shown in Figure S7a, F, O, Ti, C and Cl are clearly detected in the XPS spectrum of MXene. The presence of Cl may come from the molten salt process. Furthermore, in the high-resolution Mo 3d spectrum of MXene/MoS2 composite (Figure 2d), two distinct peaks in pure MoS2 at 229.1 eV (Mo 3d5/2) and 232.2 eV (Mo 3d3/2) can be ascribed to Mo4+. In addition, the peak near 235.5 eV is attributed to Mo6+ (oxidation of the MoS2 surface). The peak at 226.3 eV is attributed to the Mo–S bond. Correspondingly, in the MoS2/MXene composite, in addition to the four peaks mentioned above, the peaks at 230.4 and 233.3 eV are ascribed to Mo 3d5/2 and Mo 3d3/2, respectively, suggesting that Mo–S–C and Mo–O–C bonds are formed. These bonds enhance the interaction of MoS2 and MXene and enhance the structural stability of the material during cycling. The high-resolution spectrum of Ti 2p in Figure 2e contains peaks associated with Ti–O (sp1), Ti2+ (2p1/2), Ti–C, Ti–O (sp3), Ti–N and Ti–S. And it is noteworthy that the two peaks at 456.7 and 455.6 eV correspond to Ti–N and Ti–S bonds, respectively [38,39], indicating that MoS2 is firmly anchored to the surface of MXene through the Ti-S bond, which strengthens the structural stability of the material. Additionally, the presence of Ti–N bonds indicates the successful doping of N elements. The peaks at 533.6, 532.1 and 530.4 eV in the O 1s high-resolution spectrum (Figure 2f) are attributed to C–Ti-(OH)x, C–Ti/Mo–Ox and Ti–O bonds, respectively. The N 1s high-resolution spectrum in Figure 2g has four peaks. The peaks at 397.2 and 399.8 eV are attributed to pyridine nitrogen and pyrrole nitrogen, respectively, and the strong peak at 395.1 eV is attributed to Mo 3p3/2. The peak at 396.1 eV corresponds to the Ti–N bond, which coincides with the Ti 2p fine spectrum, and nitrogen doping can improve the electrochemical performance of the material [40]. In addition, the atomic percentages of the individual elemental content of the materials analyzed by XPS data are shown in Table S2. The masses of Mo, S and Ti elements in the synthesized materials were also quantified by ICP-MS (Table S3). Meanwhile, the MoS2/MXene displays a specific surface area of 41.67 m2 g−1 and pore size mainly distributed between 2 and 5 nm (Figure S8). The large specific surface area combined with rich mesopores facilitates the ion transportation and electrolyte diffusion.
The electrochemical performance was evaluated by a half-cell system. Figure 3a illustrates the CV curves of MoS2/MXene in the first three cycles at 0.1 mV s−1 scanning rate and in a voltage range of 0.01 to 3.0 V (vs. Li+/Li) for the study of the lithium ion lithiation/de-lithiation behavior. Within the initial cathodic scan curve, the peaks at 0.5–1.0 V are attributed to the formation of LixMoS2 in the Li+ intercalated into the MoS2 and Li+ intercalated into the MXene interlayer to form LiyTi3C2Tx, along with the generation of solid electrolyte interphase (SEI) film, and the peak at around 1.9 V is probably attributed to the reduction of LixMoS2 to Mo and Li2S. The reduction peak at 0.01 V is attributed to the electrochemical decomposition of MoS2, resulting in the intercalation of Mo metal into Li2S. Within the initial anodic scan curve, the peak around 1.6 V is attributed to the oxidation of Mo and the de-intercalating of Li+ between MXene layers, while the peak at ~2.3 V relates to the de-lithiating of Li2S [41]. In the two subsequent cycles, the CV curves nearly overlap, suggesting a stable structure of the electrode. The relevant electrochemical reaction equations are shown in equations as follows:
T i 3 C 2 T x + y L i + + y e L i y T i 3 C 2 T x
M o S 2 + x L i + + x e L i x M o S 2 #
L i x M o S 2 + ( 4 x ) L i + + ( 4 x ) e 2 L i 2 S + M o
The corresponding charge/discharge profiles for different cycle times at 100 mA g−1 are illustrated in Figure 3b. The plateaus of 1.37 and 1.8 V for the discharge process relate to the formation of LixMoS2 and Li2S, while the plateaus of 1.6 and 2.3 V for the charging process are equivalent to the oxidation of Mo or the de-lithiation of LiyTi3C2Tx and Li2S, which is in line with the results of the CV curves. The initial Coulombic efficiency (ICE) of the material was calculated to be 81.60%, which was far superior to pure MoS2 and MXene, followed by a Coulombic efficiency of 96.24% in the second cycle, and it was able to maintain a high Coulombic efficiency in the subsequent cycles (98.99% at 50 cycles and 99.36% at 100 cycles). In addition, the composite has a low capacity decay rate and good cycling performance with charge/discharge capacities of 719/731 mA h g−1, respectively, after 100 cycles. The lower Coulombic efficiency in the first cycle may be due to irreversible capacity decay and the decomposition of electrolytes for SEI. The cycling performance of the MoS2/MXene electrode and the counterparts with different MoS2 and MXene ratios at 100 mA g−1 was investigated (Figure 3c). It can be found that the MXene with unique structure delivers a very stable cycling performance (maintaining 94% capacity retention compared to the third cycle after 100 cycles). Pure MoS2 electrode behaves with a first discharge capacity of 1261 mA h g−1, but the discharge capacity drops to 588 mA h g−1 after 100 cycles. This could be attributed to the structural collapse caused by the volume expansion during the charging and discharging process. When incorporating with MXene, the volume expansion problem can be alleviated, and thus the cycling stability can be enhanced. All electrodes show capacity decay in the first 20 cycles, which could be due to the formation of SEI film. Among different compositions, the MoS2/MXene electrode behaves the best cycling stability, retaining a discharge capacity of 731 mA h g−1 after 100 cycles and having a Coulombic efficiency of approximately 99%. In comparison, the MoS2/MXene-2:1, MoS2/MXene-3:1 and MoS2/MXene-5:1 electrodes retain reversible capacities of 534 mA h g−1, 704 mA h g−1 and 691 mA h g−1, respectively. For the MoS2/MXene-2:1 and MoS2/MXene-3:1 electrodes, the cycling stability of the materials was improved compared to the pristine MoS2, but the capacities were lower due to the large proportion of MXene content, while the capacity decay was more serious for MoS2/MXene-5:1, probably due to the high MoS2 content, meaning the active material would accumulate and agglomerate on the MXene surface. By the above comparison, the electrode MoS2/MXene with the ratio of 4:1 behaves with the most excellent electrochemical properties. Moreover, the electrochemical properties of the MXene was studied (Figure S9), and it delivers a first discharge capacity of 355 mA h g−1 with initial Coulombic efficiency of only 40%. Additionally, the rate performance of these electrodes was also investigated at different current densities (Figure 3d). The capacity of MoS2/MXene at current densities of 100, 200, 500, 1000 and 2000 mA g−1 were 834 mA h g−1 (3rd cycle), 778 mA h g−1 (13th cycle), 713 mA h g−1 (23rd cycle), 655 mA h g−1 (33rd cycle) and 526 mA h g−1 (43rd cycle), respectively, and when the current density backed to 100 mA g−1 again, the capacity could be retained at 852 mA h g−1 (83rd cycle). Clearly, the capacities of the MoS2/MXene-2:1 and MoS2/MXene-3:1 electrodes at different current densities were higher than those of pure MoS2 electrode under the same conditions. For MoS2/MXene-5:1, although the reversible capacity increases, the rate performance decreased due to the aggregation and destruction of the composite structure during cycling, which is also consistent with the cycling performance results. The charge/discharge profiles of MoS2/MXene electrode at the corresponding number of cycles are shown in Figure 3e. As can be seen, although the overpotential (ΔV) increases with the current density, it still retains a clear charge/discharge plateau, indicating an excellent rate performance. However, at high current densities, ion diffusion and electron transfer become slower due to the slow redox reaction resulting from the generation of the concentration polarization [42,43,44]. The capacity values of different materials at the 3rd, 13th, 23rd, 43rd, 53rd, 63rd, 73rd and 83rd cycles are listed in Table S4. In addition, the electrochemical performances of previously reported MoS2-based electrodes are summarized in comparison with the MoS2/MXene electrode in this work (Figure 3f) [29,45,46,47,48,49,50]. It can be seen that at different current densities the capacity retention of this work is at a relatively high level.
In order to comprehensively understand the synergistic effect between MoS2 and MXene, we drew a schematic diagram comparing the lithiation and de-lithiation processes of the MoS2, MXene and MoS2/MXene anode materials. As shown in Figure 4a, the pure MXene lamellar material undergoes different degrees of re-stacking during the lithiation process, which reduces the layer spacing and negatively affect the ion transport in the lamellar material. While the pure MoS2 lamellar material undergoes irreversible volume expansion during the lithiation process, which also degrades the electrochemical performance of the material, the above two materials have undergone different degrees of volume changes during the lithiation process, leading to irreversible phenomenon during the de-lithiation process. The MoS2/MXene composite obtained by one-step hydrothermal method solves the above problems well. On the one hand, the existence of MXene slows down the volume expansion of MoS2 during cycling, and on the other hand, MoS2 can alleviate the restacking of MXene. The MoS2 and MXene complement each other and make up for each other’s shortcomings. We also studied the EIS and the GITT of the electrodes to better elucidate the transfer kinetics of the electrodes. Figure 4b displays the Nyquist plot obtained from the EIS test. The plot consists of two parts, a semicircle in high-frequency region and a diagonal line in low-frequency region, belonging to the charge transfer resistance Rct and the diffusion-related Warburg impedance (Zw), respectively, which indicates that the electrochemical behavior is controlled by charge transfer and ion diffusion [51]. We can find that MoS2 displays the largest Rct and MXene shows the smallest Rct. There is no significant difference in the Rct (after one cycle) of the MoS2/MXene-2:1, MoS2/MXene-3:1, MoS2/MXene and MoS2/MXene-5:1 electrodes, but the slope in the low frequency region shows that the MoS2/MXene electrode has the smallest Zw. The EIS analysis shows that the introduction of MXene reduces the impedance of electron diffusion and improves the electrical conductivity of the material. Moreover, it is found that the sample MoS2/MXene has the lowest Rct and Zw, which also corresponds to the previous electrochemical performance. The Li+ diffusion coefficient (DLi+) was calculated based on the results of the constant GITT and the detailed calculation process is presented in the Supporting Information. The test was carried out by applying a current pulse of 50 mA g−1 for 30 min with a chirality time of 30 min. The results of the test are given in Figure 4c–d, and the DGITT of MoS2/MXene is higher in the first discharge platform than in the second one because the first discharge platform belongs to the embedding of protons and the second one links with the embedding of lithium ions, a phenomenon attributed to the diffusion coefficient of lithium ions itself is lower than that of protons, and the charging platform also exhibits a similar pattern [52,53,54]. After comparison, it was found that the time required to perform a charge/discharge cycle for MoS2/MXene was shorter than that of the MoS2, indicating a more rapid embedding/disembedding process of lithium ions. By comparing the calculated DLi+, the results represent the energy barrier for Li+ diffusion. Apparently, the average DLi+ of MoS2/MXene is larger than MoS2, and it can be seen that MoS2 is nearly half an order of magnitude lower than MoS2/MXene; meanwhile, both have similar variation patterns, further indicating their excellent diffusion kinetic behavior. By the above EIS and GITT analysis, the introduction of MXene lamellar can provide sufficient surface electron mobility and efficient ion transport to promote electrochemical kinetic behavior.

4. Conclusions

In summary, the MoS2/MXene-x:1 (x = 2, 3, 4, 5) composites were prepared via a facile hydrothermal strategy and applied to the corresponding electrode for lithium-ion batteries. It was demonstrated by Raman, XRD and XPS that the synthesized MoS2 formed a strong combination with the MXene through the Ti–S and Ti–O–Mo bonds; meanwhile, the N element derived from the decomposition of thioacetamide was incorporated into the composite to enhance the electrochemical properties. As expected, the corresponding MoS2/MXene electrode delivers initial charge/discharge capacities of 887/1087 mA h g−1, respectively, at 100 mA g−1 with an ICE of 81.6%, and a discharge capacity of 731 mA h g−1 can be retained after 100 cycles. Furthermore, the EIS and GITT investigations demonstrate that the introduction of MXene is beneficial to improve electronic conductivity, boost reaction kinetics and enhance stability. Additionally, the nitrogen-doped Ti3C2Tx can effectively suppress the volume change of MoS2 and further enhance the electrical conductivity of the entire electrode, whereas MoS2 can reduce the self-stacking of Ti3C2Tx. These excellent results demonstrate the promising potential of the MoS2/MXene electrode for energy storage systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/batteries8100156/s1, Figure S1: One-Step Hydrothermal Reaction induced Nitrogen-Doped MoS2/MXene Composites with Superior Lithium-Ion Storage; Table S1: One-Step Hydrothermal Reaction induced Nitrogen-Doped MoS2/MXene Composites with Superior Lithium-Ion Storage.

Author Contributions

Conceptualization, H.G. and H.L.; methodology, C.G. and M.L.; software, G.Z. and Y.X.; formal analysis, J.L.; investigation, M.L.; resources, J.X.; data curation, J.C.; writing—original draft preparation, C.G. and M.L.; writing—review and editing, all authors; supervision, H.G. and H.L.; project administration, H.L.; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Australian Research Council (ARC, DP180102297, FT180100705).

Data Availability Statement

Not applicable.

Acknowledgments

C.G. and M.L. contributed equally to this work. All authors thank the support from “Joint International Laboratory on Environmental and Energy Frontier Materials” and “Innovation Research Team of High-Level Local Universities in Shanghai”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, T.; Long, M.; Xiao, J.; Liu, H.; Wang, G. Recent research on emerging organic electrode materials for energy storage. Energy Mater. 2021, 1, 09. [Google Scholar] [CrossRef]
  2. Jun, X.; Xiao, L.; Kaikai, T.; Mengqi, L.; Jun, C.; Dandan, W.; Hong, G.; Hao, L. Enhanced electrochemical performance of Li-rich cathode material for lithium-ion batteries. Surf. Innov. 2022, 10, 119–127. [Google Scholar]
  3. Lu, J.; Chen, Z.W.; Pan, F.; Cui, Y.; Amine, K. High-Performance Anode Materials for Rechargeable Lithium-Ion Batteries. Electrochem. Energy Rev. 2018, 1, 35–53. [Google Scholar] [CrossRef]
  4. Nitta, N.; Wu, F.X.; Lee, J.T.; Yushin, G. Li-ion battery materials: Present and future. Mater. Today 2015, 18, 252–264. [Google Scholar] [CrossRef]
  5. Pramanik, A.; Chattopadhyay, S.; De, G.; Mahanty, S. Efficient energy storage in mustard husk derived porous spherical carbon nanostructures. Mater. Adv. 2021, 2, 7463–7472. [Google Scholar] [CrossRef]
  6. Pramanik, A.; Maiti, S.; Sreemany, M.; Mahanty, S. Rock-Salt-Templated Mn3O4 Nanoparticles Encapsulated in a Mesoporous 2D Carbon Matrix: A High Rate 2 V Anode for Lithium-Ion Batteries with Extraordinary Cycling Stability. Chemistryselect 2017, 2, 7854–7864. [Google Scholar] [CrossRef]
  7. Anasori, B.; Lukatskaya, M.R.; Gogotsi, Y. 2D metal carbides and nitrides (MXenes) for energy storage. Nat. Rev. Mater. 2017, 2, 98. [Google Scholar] [CrossRef]
  8. Enyashin, A.N.; Ivanoyskii, A.L. Structural and Electronic Properties and Stability of MXenes Ti2C and Ti3C2 Functionalized by Methoxy Groups. J. Phys. Chem. C 2013, 117, 13637–13643. [Google Scholar] [CrossRef]
  9. Han, M.K.; Yin, X.W.; Wu, H.; Hou, Z.X.; Song, C.Q.; Li, X.L.; Zhang, L.T.; Cheng, L.F. Ti3C2 MXenes with Modified Surface for High-Performance Electromagnetic Absorption and Shielding in the X-Band. ACS Appl. Mater. Interfaces 2016, 8, 21011–21019. [Google Scholar] [CrossRef]
  10. Karlsson, L.H.; Birch, J.; Halim, J.; Barsoum, M.W.; Persson, P.O.A. Atomically Resolved Structural and Chemical Investigation of Single MXene Sheets. Nano Lett. 2015, 15, 4955–4960. [Google Scholar] [CrossRef] [Green Version]
  11. Lipatov, A.; Alhabeb, M.; Lukatskaya, M.R.; Boson, A.; Gogotsi, Y.; Sinitskii, A. Effect of Synthesis on Quality, Electronic Properties and Environmental Stability of Individual Monolayer Ti3C2 MXene Flakes. Adv. Electron. Mater. 2016, 2, 00255. [Google Scholar]
  12. Guo, X.; Gao, H.; Wang, S.; Yang, G.; Zhang, X.; Zhang, J.; Liu, H.; Wang, G. MXene-Based Aerogel Anchored with Antimony Single Atoms and Quantum Dots for High-Performance Potassium-Ion Batteries. Nano Lett. 2022, 22, 1225–1232. [Google Scholar] [CrossRef]
  13. Gao, H.; Zhou, T.F.; Zheng, Y.; Zhang, Q.; Liu, Y.Q.; Chen, J.; Liu, H.K.; Guo, Z.P. CoS Quantum Dot Nanoclusters for High-Energy Potassium-Ion Batteries. Adv. Funct. Mater. 2017, 27, 02634. [Google Scholar] [CrossRef]
  14. Wang, Y.K.; Liu, M.C.; Cao, J.Y.; Zhang, H.J.; Kong, L.B.; Trudgeon, D.P.; Li, X.H.; Walsh, F.C. 3D Hierarchically Structured CoS Nanosheets: Li+ Storage Mechanism and Application of the High-Performance Lithium-Ion Capacitors. ACS Appl. Mater. Interfaces 2020, 12, 3709–3718. [Google Scholar] [CrossRef]
  15. Zhang, S.L.; Ying, H.J.; Huang, P.F.; Wang, J.L.; Zhang, Z.; Yang, T.T.; Han, W.Q. Rational Design of Pillared SnS/Ti3C2Tx MXene for Superior Lithium-Ion Storage. ACS Nano 2020, 14, 17665–17674. [Google Scholar] [CrossRef]
  16. Zhang, X.Y.; Shi, H.T.; Liu, L.S.; Min, C.Y.; Liang, S.T.; Xu, Z.W.; Xue, Y.L.; Hong, C.X.; Cai, Z.J. Construction of MoS2/Mxene heterostructure on stress-modulated kapok fiber for high-rate sodium-ion batteries. J. Colloid Interface Sci. 2022, 605, 472–482. [Google Scholar] [CrossRef] [PubMed]
  17. Xiao, Y.; Lee, S.H.; Sun, Y.K. The Application of Metal Sulfides in Sodium Ion Batteries. Adv. Energy Mater. 2017, 7, 01329. [Google Scholar] [CrossRef]
  18. Liu, C.; Wang, Y.; Sun, J.; Chen, A. A Review on Applications of Layered Phosphorus in Energy Storage. Trans. Tianjin Univ. 2020, 26, 104–126. [Google Scholar] [CrossRef] [Green Version]
  19. Su, F.; Hou, X.C.; Qin, J.Q.; Wu, Z.S. Recent Advances and Challenges of Two-Dimensional Materials for High-Energy and High-Power Lithium-Ion Capacitors. Batteries Supercaps 2020, 3, 10–29. [Google Scholar] [CrossRef] [Green Version]
  20. Wang, L.L.; Zhang, Q.F.; Zhu, J.Y.; Duan, X.D.; Xu, Z.; Liu, Y.T.; Yang, H.G.; Lu, B.A. Nature of extra capacity in MoS2 electrodes: Molybdenum atoms accommodate with lithium. Energy Storage Mater. 2019, 16, 37–45. [Google Scholar] [CrossRef]
  21. Ding, S.J.; Zhang, D.Y.; Chen, J.S.; Lou, X.W. Facile synthesis of hierarchical MoS2 microspheres composed of few-layered nanosheets and their lithium storage properties. Nanoscale 2012, 4, 95–98. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, X.H.; Wang, Z.Y.; Yu, M.Z.; Xiu, L.Y.; Qiu, J.S. Stabilizing the MXenes by Carbon Nanoplating for Developing Hierarchical Nanohybrids with Efficient Lithium Storage and Hydrogen Evolution Capability. Adv. Mater. 2017, 29, 07017. [Google Scholar] [CrossRef] [PubMed]
  23. Yuan, Z.Y.; Wang, L.L.; Li, D.D.; Cao, J.M.; Han, W. Carbon-Reinforced Nb2CTx MXene/MoS2 Nanosheets as a Superior Rate and High-Capacity Anode for Sodium-Ion Batteries. ACS Nano 2021, 15, 7439–7450. [Google Scholar] [CrossRef]
  24. Zhang, Y.L.; Mu, Z.J.; Yang, C.; Xu, Z.K.; Zhang, S.; Zhang, X.Y.; Li, Y.J.; Lai, J.P.; Sun, Z.H.; Yang, Y.; et al. Rational Design of MXene/1T-2H MoS2-C Nanohybrids for High-Performance Lithium-Sulfur Batteries. Adv. Funct. Mater. 2018, 28, 07578. [Google Scholar] [CrossRef]
  25. Chen, Y.L.; Su, F.Y.; Xie, H.Q.; Wang, R.P.; Ding, C.H.; Huang, J.D.; Xu, Y.X.; Ye, L.Q. One-step construction of S-scheme heterojunctions of N-doped MoS2 and S-doped g-C3N4 for enhanced photocatalytic hydrogen evolution. Chem. Eng. J. 2021, 404, 126498. [Google Scholar] [CrossRef]
  26. Han, J.H.; Zhang, S.C.; Song, Q.G.; Yan, H.Y.; Kang, J.H.; Guo, Y.R.; Liu, Z.F. The synergistic effect with S-vacancies and built-in electric field on a TiO2/MoS2 photoanode for enhanced photoelectrochemical performance. Sustain. Energ. Fuels 2021, 5, 509–517. [Google Scholar] [CrossRef]
  27. Nguyen, D.C.; Tran, D.T.; Doan, T.L.L.; Kim, D.H.; Kim, N.H.; Lee, J.H. Rational Design of Core@shell Structured CoSx@Cu2MoS4 Hybridized MoS2/N,S-Codoped Graphene as Advanced Electrocatalyst for Water Splitting and Zn-Air Battery. Adv. Energy Mater. 2020, 10, 03289. [Google Scholar] [CrossRef]
  28. Xie, D.; Xia, X.H.; Wang, Y.D.; Wang, D.H.; Zhong, Y.; Tang, W.J.; Wang, X.L.; Tu, J.P. Nitrogen-Doped Carbon Embedded MoS2 Microspheres as Advanced Anodes for Lithium- and Sodium-Ion Batteries. Chem.-Eur. J. 2016, 22, 11617–11623. [Google Scholar] [CrossRef]
  29. Chen, C.; Xie, X.; Anasori, B.; Sarycheva, A.; Makaryan, T.; Zhao, M.; Urbankowski, P.; Miao, L.; Jiang, J.; Gogotsi, Y. MoS2-on-MXene Heterostructures as Highly Reversible Anode Materials for Lithium-Ion Batteries. Angew. Chem.-Int. Edit. 2018, 57, 1846–1850. [Google Scholar] [CrossRef]
  30. Hu, Z.; Kuai, X.; Chen, J.; Sun, P.; Zhang, Q.; Wu, H.-H.; Zhang, L. Strongly Coupled MoS2 Nanocrystal/Ti3C2 Nanosheet Hybrids Enable High-Capacity Lithium-Ion Storage. ChemSusChem 2020, 13, 1485–1490. [Google Scholar] [CrossRef]
  31. Xu, X.; Zhao, R.S.; Ai, W.; Chen, B.; Du, H.F.; Wu, L.S.; Zhang, H.; Huang, W.; Yu, T. Controllable Design of MoS2 Nanosheets Anchored on Nitrogen-Doped Graphene: Toward Fast Sodium Storage by Tunable Pseudocapacitance. Adv. Mater. 2018, 30, 1800658. [Google Scholar] [CrossRef]
  32. Li, J.B.; Tang, S.C.; Li, Z.Q.; Wang, C.Y.; Li, J.L.; Li, X.D.; Ding, Z.B.; Pan, L.K. Crosslinking Nanoarchitectonics of Nitrogen-doped Carbon/MoS2 Nanosheets/Ti3C2Tx MXene Hybrids for Highly Reversible Sodium Storage. ChemSusChem 2021, 14, 5293–5303. [Google Scholar] [CrossRef] [PubMed]
  33. Li, M.; Lu, J.; Luo, K.; Li, Y.B.; Chang, K.K.; Chen, K.; Zhou, J.; Rosen, J.; Hultman, L.; Eklund, P.; et al. Element Replacement Approach by Reaction with Lewis Acidic Molten Salts to Synthesize Nanolaminated MAX Phases and MXenes. J. Am. Chem. Soc. 2019, 141, 4730–4737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Lin, W.J.; Huang, Y.H.; He, G.Q. Unique CoS hierarchitectures for high-performance lithium ion batteries. Crystengcomm 2018, 20, 6727–6732. [Google Scholar] [CrossRef]
  35. Zhou, Y.P.; Li, J.; Yang, Y.; Luo, B.; Zhang, X.; Fong, E.L.E.; Chu, W.; Huang, K.M. Unique 3D flower-on-sheet nanostructure of NiCo LDHs: Controllable microwave-assisted synthesis and its application for advanced supercapacitors. J. Alloy. Compd. 2019, 788, 1029–1036. [Google Scholar] [CrossRef]
  36. Li, C.R.; Song, H.; Mao, C.M.; Peng, H.R.; Li, G.C. A novel MoS2 nanosheets-decorated Sb@Sb2S3@C tubular composites as anode material for high performance lithium ion battery. J. Alloy. Compd. 2019, 786, 169–176. [Google Scholar] [CrossRef]
  37. Yu, J.L.; Zeng, M.L.; Zhou, J.; Chen, H.D.; Cong, G.T.; Liu, H.C.; Ji, M.W.; Zhu, C.Z.; Xu, J. A one-pot synthesis of nitrogen doped porous MXene/TiO2 heterogeneous film for high-performance flexible energy storage. Chem. Eng. J. 2021, 426, 130765. [Google Scholar] [CrossRef]
  38. Luo, J.M.; Zheng, J.H.; Nai, J.W.; Jin, C.B.; Yuan, H.D.; Sheng, O.W.; Liu, Y.J.; Fang, R.Y.; Zhang, W.K.; Huang, H.; et al. Atomic Sulfur Covalently Engineered Interlayers of Ti3C2 MXene for Ultra-Fast Sodium-Ion Storage by Enhanced Pseudocapacitance. Adv. Funct. Mater. 2019, 29, 08107. [Google Scholar] [CrossRef]
  39. Wang, Z.; Li, X.Y.; Xuan, C.X.; Li, J.J.; Jiang, Y.L.; Xiao, J.R. Photo-synergetic nitrogen-doped MXene/reduced graphene oxide sandwich-like architecture for high-performance lithium-sulfur batteries. Int. J. Energy Res. 2021, 45, 2728–2738. [Google Scholar] [CrossRef]
  40. Ye, Z.Q.; Jiang, Y.; Li, L.; Wu, F.; Chen, R.J. Self-Assembly of 0D-2D Heterostructure Electrocatalyst from MOF and MXene for Boosted Lithium Polysulfide Conversion Reaction. Adv. Mater. 2021, 33, 2101204. [Google Scholar] [CrossRef]
  41. He, F.Y.; Tang, C.; Liva, Y.D.; Li, H.T.; Du, A.J.; Zhang, H.J. Carbon-coated MoS2 nanosheets@CNTs-Ti3C2 MXene quaternary composite with the superior rate performance for sodium-ion batteries. J. Mater. Sci. Technol. 2022, 100, 101–109. [Google Scholar] [CrossRef]
  42. Bao, W.Z.; Shuck, C.E.; Zhang, W.X.; Guo, X.; Gogotsi, Y.; Wang, G.X. Boosting Performance of Na-S Batteries Using Sulfur-Doped Ti3C2Tx MXene Nanosheets with a Strong Affinity to Sodium Polysulfides. ACS Nano 2019, 13, 11500–11509. [Google Scholar] [CrossRef] [PubMed]
  43. Dong, Y.F.; Zheng, S.H.; Qin, J.Q.; Zhao, X.J.; Shi, H.D.; Wang, X.H.; Chen, J.; Wu, Z.S. All-MXene-Based Integrated Electrode Constructed by Ti3C2 Nanoribbon Framework Host and Nanosheet Interlayer for High-Energy-Density Li-S Batteries. ACS Nano 2018, 12, 2381–2388. [Google Scholar] [CrossRef] [PubMed]
  44. Mou, J.R.; Li, Y.J.; Liu, T.; Zhang, W.J.; Li, M.; Xu, Y.T.; Zhong, L.; Pan, W.H.; Yang, C.H.; Huang, J.L.; et al. Metal-Organic Frameworks-Derived Nitrogen-Doped Porous Carbon Nanocubes with Embedded Co Nanoparticles as Efficient Sulfur Immobilizers for Room Temperature Sodium-Sulfur Batteries. Small Methods 2021, 5, 00455. [Google Scholar] [CrossRef] [PubMed]
  45. Bai, J.; Zhao, B.C.; Zhou, J.F.; Si, J.G.; Fang, Z.T.; Li, K.Z.; Ma, H.Y.; Dai, J.M.; Zhu, X.B.; Sun, Y.P. Glucose-Induced Synthesis of 1T-MoS2/C Hybrid for High-Rate Lithium-Ion Batteries. Small 2019, 15, 05420. [Google Scholar] [CrossRef]
  46. Huang, M.J.; Chen, H.H.; He, J.; An, B.H.; Sun, L.N.; Li, Y.L.; Ren, X.Z.; Deng, L.B.; Zhang, P.X. Ultra small few layer MoS2 embedded into three-dimensional macro-micro-mesoporous carbon as a high performance lithium ion batteries anode with superior lithium storage capacity. Electrochim. Acta 2019, 317, 638–647. [Google Scholar] [CrossRef]
  47. Lin, J.; Shi, Y.H.; Li, Y.F.; Wu, X.L.; Zhang, J.P.; Xie, H.M.; Sun, H.Z. Confined MoS2 growth in a unique composite matrix for ultra-stable and high-rate lithium/sodium-ion anodes. Chem. Eng. J. 2022, 428, 131103. [Google Scholar] [CrossRef]
  48. Wang, X.L.; Li, G.; Seo, M.H.; Hassan, F.M.; Hoque, M.A.; Chen, Z.W. Sulfur Atoms Bridging Few-Layered MoS2 with S-Doped Graphene Enable Highly Robust Anode for Lithium-Ion Batteries. Adv. Energy Mater. 2015, 5, 1501106. [Google Scholar] [CrossRef]
  49. Wu, H.Y.; Zhang, X.E.; Wu, Q.H.; Han, Y.; Wu, X.Y.; Ji, P.L.; Zhou, M.; Diao, G.W.; Chen, M. Confined growth of 2D MoS2 nanosheets in N-doped pearl necklace-like structured carbon nanofibers with boosted lithium and sodium storage performance. Chem. Commun. 2020, 56, 141–144. [Google Scholar] [CrossRef]
  50. Zhang, S.; Zeng, Y.W.; Wang, Z.T.; Zhao, J.; Dong, G.D. Glycerol-controlled synthesis of MoS2 hierarchical architectures with well-tailored subunits and enhanced electrochemical performance for lithium ion batteries. Chem. Eng. J. 2018, 334, 487–496. [Google Scholar] [CrossRef]
  51. Cao, Y.P.; Chen, H.; Shen, Y.P.; Chen, M.; Zhang, Y.L.; Zhang, L.Y.; Wang, Q.; Guo, S.J.; Yang, H. SnS2 Nanosheets Anchored on Nitrogen and Sulfur Co-Doped MXene Sheets for High-Performance Potassium-Ion Batteries. ACS Appl. Mater. Interfaces 2021, 13, 17668–17676. [Google Scholar] [CrossRef] [PubMed]
  52. Cao, J.M.; Sun, Z.Q.; Li, J.Z.; Zhu, Y.K.; Yuan, Z.Y.; Zhang, Y.M.; Li, D.D.; Wang, L.L.; Han, W. Microbe-Assisted Assembly of Ti3C2Tx MXene on Fungi-Derived Nanoribbon Heterostructures for Ultrastable Sodium and Potassium Ion Storage. ACS Nano 2021, 15, 3423–3433. [Google Scholar] [CrossRef] [PubMed]
  53. Shi, M.J.; Wang, B.; Chen, C.; Lang, J.W.; Yan, C.; Yan, X.B. 3D high-density MXene@MnO2 microflowers for advanced aqueous zinc-ion batteries. J. Mater. Chem. A 2020, 8, 24635–24644. [Google Scholar] [CrossRef]
  54. Zhu, X.D.; Cao, Z.Y.; Wang, W.J.; Li, H.J.; Dong, J.C.; Gao, S.P.; Xu, D.X.; Li, L.; Shen, J.F.; Ye, M.X. Superior-Performance Aqueous Zinc-Ion Batteries Based on the In Situ Growth of MnO2 Nanosheets on V2CTx MXene. ACS Nano 2021, 15, 2971–2983. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Schematic diagram of the synthesis process of MoS2/MXene. (b) SEM images of MoS2; (c) HRTEM images of MXene; (d) SEM images of MoS2/MXene; (e–f) TEM images of MoS2/MXene, where inset (f) is SEAD pattern of MoS2/MXene; (g) HRTEM images of MXene in the MoS2/MXene composite, where inset is its HRTEM image at higher magnification; (hm) elemental mapping of MoS2/MXene.
Figure 1. (a) Schematic diagram of the synthesis process of MoS2/MXene. (b) SEM images of MoS2; (c) HRTEM images of MXene; (d) SEM images of MoS2/MXene; (e–f) TEM images of MoS2/MXene, where inset (f) is SEAD pattern of MoS2/MXene; (g) HRTEM images of MXene in the MoS2/MXene composite, where inset is its HRTEM image at higher magnification; (hm) elemental mapping of MoS2/MXene.
Batteries 08 00156 g001
Figure 2. (a) XRD patterns of MoS2/MXene−x:1 (x = 2, 3, 4, 5) and MXene; (b) Raman spectra of MoS2/MXene composites and MoS2; (c) XPS full spectra of MoS2/MXene, MoS2 and MXene; high-resolution (d) Mo 3d XPS spectra of MoS2 and MoS2/MXene; high-resolution (e) Ti 2p (f) O 1s and (g) N 1s XPS spectra of MoS2/MXene.
Figure 2. (a) XRD patterns of MoS2/MXene−x:1 (x = 2, 3, 4, 5) and MXene; (b) Raman spectra of MoS2/MXene composites and MoS2; (c) XPS full spectra of MoS2/MXene, MoS2 and MXene; high-resolution (d) Mo 3d XPS spectra of MoS2 and MoS2/MXene; high-resolution (e) Ti 2p (f) O 1s and (g) N 1s XPS spectra of MoS2/MXene.
Batteries 08 00156 g002
Figure 3. (a) CV curves of MoS2/MXene at the scan rate of 0.1 mV s−1; (b) charge/discharge profiles of MoS2/MXene at 100 mA g−1; (c) cycling performance of MoS2, MXene and MoS2/MXene−x:1 (x = 2, 3, 4, 5) at 100 mA g−1; (d) rate performance of MoS2, MXene and MoS2/MXene−x:1 (x = 2, 3, 4, 5) at different current densities (100, 200, 500, 1000 and 2000 mA g−1); (e) charge/discharge profiles of MoS2/MXene electrode at different current densities; (f) comparison of the capacity retention of as-prepared MoS2/MXene with other related materials.
Figure 3. (a) CV curves of MoS2/MXene at the scan rate of 0.1 mV s−1; (b) charge/discharge profiles of MoS2/MXene at 100 mA g−1; (c) cycling performance of MoS2, MXene and MoS2/MXene−x:1 (x = 2, 3, 4, 5) at 100 mA g−1; (d) rate performance of MoS2, MXene and MoS2/MXene−x:1 (x = 2, 3, 4, 5) at different current densities (100, 200, 500, 1000 and 2000 mA g−1); (e) charge/discharge profiles of MoS2/MXene electrode at different current densities; (f) comparison of the capacity retention of as-prepared MoS2/MXene with other related materials.
Batteries 08 00156 g003
Figure 4. (a) Schematic diagram of the processes of lithiation and de-lithiation in MXene, MoS2 and MXene/MoS2; (b) Nyquist plots of MoS2, MXene, MoS2/MXene−2:1, MoS2/MXene−3:1, MoS2/MXene and MoS2/MXene−5:1 after the first charge/discharge cycle; GITT results of (c) MoS2 and (d) MoS2/MXene.
Figure 4. (a) Schematic diagram of the processes of lithiation and de-lithiation in MXene, MoS2 and MXene/MoS2; (b) Nyquist plots of MoS2, MXene, MoS2/MXene−2:1, MoS2/MXene−3:1, MoS2/MXene and MoS2/MXene−5:1 after the first charge/discharge cycle; GITT results of (c) MoS2 and (d) MoS2/MXene.
Batteries 08 00156 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gong, C.; Long, M.; Xiao, J.; Li, J.; Chen, J.; Xiao, Y.; Zhang, G.; Gao, H.; Liu, H. One-Step Hydrothermal Reaction Induced Nitrogen-Doped MoS2/MXene Composites with Superior Lithium-Ion Storage. Batteries 2022, 8, 156. https://doi.org/10.3390/batteries8100156

AMA Style

Gong C, Long M, Xiao J, Li J, Chen J, Xiao Y, Zhang G, Gao H, Liu H. One-Step Hydrothermal Reaction Induced Nitrogen-Doped MoS2/MXene Composites with Superior Lithium-Ion Storage. Batteries. 2022; 8(10):156. https://doi.org/10.3390/batteries8100156

Chicago/Turabian Style

Gong, Cheng, Mengqi Long, Jun Xiao, Jiayi Li, Jun Chen, Yang Xiao, Guilai Zhang, Hong Gao, and Hao Liu. 2022. "One-Step Hydrothermal Reaction Induced Nitrogen-Doped MoS2/MXene Composites with Superior Lithium-Ion Storage" Batteries 8, no. 10: 156. https://doi.org/10.3390/batteries8100156

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