Next Article in Journal
Leading Edge Technologies and Perspectives in Industrial Oilseed Extraction
Next Article in Special Issue
PPy-Coated Mo3S4/CoMo2S4 Nanotube-like Heterostructure for High-Performance Lithium Storage
Previous Article in Journal
The Adsorption Behaviors of CO and H2 to FeO onto CaO Surfaces: A Density Functional Theory Study
Previous Article in Special Issue
Electrospun Sandwich-like Structure of PVDF-HFP/Cellulose/PVDF-HFP Membrane for Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

MoS2/SnS/CoS Heterostructures on Graphene: Lattice-Confinement Synthesis and Boosted Sodium Storage

1
State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, China
2
Quzhou Institute for Innovation in Resource Chemical Engineering, Quzhou 324000, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(16), 5972; https://doi.org/10.3390/molecules28165972
Submission received: 6 July 2023 / Revised: 31 July 2023 / Accepted: 4 August 2023 / Published: 9 August 2023
(This article belongs to the Special Issue Innovative Materials for Energy Storage and Conversion)

Abstract

:
The development of high-efficiency multi-component composite anode nanomaterials for sodium-ion batteries (SIBs) is critical for advancing the further practical application. Numerous multi-component nanomaterials are constructed typically via confinement strategies of surface templating or three-dimensional encapsulation. Herein, a composite of heterostructural multiple sulfides (MoS2/SnS/CoS) well-dispersed on graphene is prepared as an anode nanomaterial for SIBs, via a distinctive lattice confinement effect of a ternary CoMoSn-layered double-hydroxide (CoMoSn-LDH) precursor. Electrochemical testing demonstrates that the composite delivers a high-reversible capacity (627.6 mA h g−1 after 100 cycles at 0.1 A g−1) and high rate capacity of 304.9 mA h g−1 after 1000 cycles at 5.0 A g−1, outperforming those of the counterparts of single-, bi- and mixed sulfides. Furthermore, the enhancement is elucidated experimentally by the dominant capacitive contribution and low charge-transfer resistance. The precursor-based lattice confinement strategy could be effective for constructing uniform composites as anode nanomaterials for electrochemical energy storage.

Graphical Abstract

1. Introduction

Rechargeable sodium-ion batteries (SIBs) were increasingly promising in the field of large-scale electric energy storage for renewable energy and smart grids during the “beyond-lithium-ion-batteries” era, due to the advantageous characteristics of low cost, high abundance, and wide distribution of sodium resources [1,2,3,4]. The larger radius of Na+ (1.02 Å) than that of Li+ (0.76 Å), however, results in the sluggish kinetics of insertion/extraction of Na+ into active materials [5,6,7] and thus hinders the development and application of SIBs. The limitation has triggered a great deal of effort to develop numerous advanced anode nanomaterials for SIBs, such as carbon-based nanomaterials [8,9,10,11], metallic alloys [12,13,14], two-dimensional (2D) metal carbides (MXenes) [15,16,17,18], metal dichalcogenides [19,20,21,22], and metal sulfides. In particular, metal sulfides are considered to be favorable for sodium storage reactions due to their weak metal–sulfur (M–S) bonds, good reversibility and high capacity [23]. Cobalt sulfide (CoS) is an interesting semiconductor material and has broad applications in supercapacitors, lithium-ion batteries, and catalysis [24,25,26], as well as SIBs due to its high-theoretical capacity [27]. However, CoS still suffers from poor cycling performance due to large volume-change-induced electrode pulverization, and also from poor rate performance owing to the sluggish electrode kinetics [28].
To alleviate the problems, effective approaches to improve electrochemical performance mainly focus on hybridizing with carbonaceous materials [29,30,31], designing nano- and sub-nano scale structures [32,33], and/or integrating with other metal sulfides [27,34,35]. Using glucose-assisted hydrothermal reaction and subsequent calcination treatment [36], carbon coated MoS2/CoS nanostructures on carbon cloth (MoS2/CoS@CC) were fabricated, and achieved a reversible capacity of 605 mA h g−1 after 100 cycles at 0.5 A g−1 and rate performance of 366 mA h g−1 at 8.0 A g−1. Using a sol–gel method [37], CoS/Co9S8 quantum dots-embedded N/S-doped carbon nanosheets were prepared, and afforded rate capability (600 and 500 mAh g−1 at 0.2 A g−1 and 10 A g−1, respectively) and a capacity retention of 87% after 200 cycles at 1 A g−1. From cobalt-containing metal–organic frameworks (MOFs) of ZIF-67 precursor [38], CoS/Co9S8 heterostructures embedded within N-doped carbon were synthesized, and provided a capacity of 409.2 mAh g−1 after 100 cycles at 0.1 A g−1 and rate capability of 134.9 mAh g−1 at 10 A g−1. By combining hydrothermal synthesis and annealing treatment [39], binary metal sulfide (CoS@SnS) heterostructure confined within carbon microspheres were acquired, and delivered reversible capacity of 463 mAh g−1 and long durability of 368 mAh g−1 after 1000 cycles at 2 A g−1. Using a solvothermal synthesis and polydopamine (PDA) encapsulation [40], binary CoS/Cu2S nanocages within nitrogen-doped carbon shell (CoS/Cu2S@N-C) were constructed, and displayed a capacity of 435.3 mAh g−1 after 1000 cycles at 2 A g−1 and high-rate capacity of 347.2 mAh g−1 after 2300 cycles at 10 A g−1. All the above anode nanomaterials exhibited highly enhanced electrochemical performances. The enhancements were contributed from the well-designed nanostructures and their homogeneous distributions, as well as hybridization with a high electron conductive carbon/multiple sulfide heterostructure [41], which were achieved mainly via confinement strategies of surface templating and/or three-dimensional encapsulation.
Herein, we report a multi-component MoS2/SnS/CoS heterostructures/graphene (MoS2/SnS/CoS/rGO) composite as a high-rate anode nanomaterial for SIBs. The well-dispersed sulfide composite is derived by pyrolyzing a single-source ternary CoMoSn-layered double-hydroxide (CoMoSn-LDH) precursor via a distinctive lattice confinement effect (Figure 1). LDHs are well-known as one large type of two-dimensional Mg(OH)2-like layered structures that consist of type-and-ratio-tunable metallic cations and anions [42]. The metallic cations are well-orderedly dispersed within the same crystalline layers [42], that is, the synthesis process and crystal growth of one metal cation are confined by, and well-distributed within, the matrix of the other metal cations, which provides a lattice confinement-based strategy to design and prepare well-distributed nanocomposites, such as alloy [43] and mixed metal oxides [44] in others’ studies, as well as sulfides in our recent studies [45,46,47]. As a result, the MoS2/SnS/CoS/rGO composite delivers a greatly boosted reversible specific capacity at 0.1 A g−1 and long-cycled rate capacity at 5.0 A g−1, exceeding those of the single-, bi- and mixed sulfides. The enhancement is further elucidated experimentally by the interfacial interaction between heterostructures, dominant capacitive contribution, and low charge-transfer resistance.

2. Results

The CoMoSn-LDH precursor/GO was prepared using a traditional co-precipitation method. The resulting X-ray diffraction (XRD) pattern of the precursor shows the characteristic basal (001) diffraction peaks and non-basal (012) diffraction peak (Figure S1a), reflecting the formation of ternary CoMoSn-LDH precursor. The scanning electron microscope(SEM) image further confirms that the precursor was grown on the supporting GO (Figure S1b). Subsequently, the CoMoSn-LDH precursor/GO was subjected to the calcination and sulfurization at 300 °C for 2 h in Ar. The resulting product (Figure 2a) shows the diffraction peaks at 14.4° and 32.7°, which are assigned to the (002) crystal plane of hexagonal 2H-MoS2 (PDF #37-1492); and the diffraction peaks at 22.0°, 26.0°, 27.5°, 31.5°, 39.0°, 44.7° and 56.7°, which are attributed to the crystal planes of (110), (120), (021), (111), (131), (141) and (042) for SnS (PDF #39-0354), respectively. The other diffraction peaks at 30.6°, 35.3°, 46.9° and 54.4° correspond to the (100), (101), (102) and (110) crystal planes of CoS (PDF#39-0354), respectively. All the aforementioned XRD results demonstrate the co-existence of the MoS2, SnS, and CoS phases in the composite. Moreover, the SEM image (Figure S2) shows that numerous nanosheets are dispersed on the rGO support, and the energy dispersive spectrometer (EDS) mapping images for SEM shows the uniform distribution of the elements of Co, Mo, Sn, S and C in the MoS2/SnS/CoS@rGO composite (Figure S3).
Furthermore, the high-resolution transmission electron microscopy (TEM) reveals the MoS2 nanosheets are confined on rGO support (as shown by a checkmark in Figure 2b). Distinct interfaces are also identified between MoS2, SnS, and CoS (as shown by two dotted boxes in Figure 2c), which strongly suggests the lattice confinement of the CoMoSn-LDH precursor. Three different lattice fringes are quantitatively determined: a lattice spacing of 0.62 nm corresponding to the (002) plane of MoS2 (Figure 2c), 0.19 nm to the (102) plane of CoS, and 0.28 nm to the (111) plane of SnS (Figure 2d).
The chemical composition and element valence of the composite were characterized by X-ray photoelectron spectra (XPS). Figure 3a shows a XPS full spectrum that demonstrates the presence of Co, Mo, Sn, C, S, and O elements. The high-resolution Co 2p spectrum was divided into two spin-orbit doublets and two oscillatory satellite peaks (Figure 3b). The former spin-orbit doublets with binding energies at 779.2/780.7 eV and 794.2/795.8 eV correspond to Co 2p3/2 and Co 2p1/2, respectively, which are assigned to the Co2+/Co3+ redox pair; while the latter satellite peaks at 786.5 and 802.1 eV are related to Co 2p3/2 and Co 2p1/2, respectively. The Sn 3d spectrum shows two strong peaks at 487.4 eV and 495.8 eV, which are assigned to Sn 3d5/2 and Sn 3d3/2, respectively, and thus proves that SnS mainly exists in a state of Sn2+ (Figure 3c). The Mo 3d spectrum was fitted by two peaks at 229.3 eV and 232.4 eV, which are assigned to Mo4+ 3d5/2 and Mo4+ 3d3/2, respectively, and also demonstrates that Mo exists in a tetravalent form. Another peak at 235.5 eV was also visible, indicating that the partial Mo4+ at the surface was oxidized to Mo6+ due to the exposure to air (Figure 3d). The S 2p spectrum was de-convoluted into two peaks at 162.2 eV and 163.3 eV (Figure 3e), which are attributed to S 2p3/2 and S 2p1/2, respectively. Note that one peak at 164.5 eV was observed, which indicates the existence of the C–S bond in the composite; this result indicates that the rGO carbon substrate is tightly bound to the three metal sulfide materials, which is more favorable for electronic migrate and maintaining structural stability. Moreover, the high-resolution C 1s spectrum was fitted by those four peaks at 284.8 eV (corresponding to C–C), 285.8 eV (C–S), 287.2 eV (C–O), and 289.5 eV (O–C=O) (as shown in Figure 3f).
The carbonaceous graphitization degree was estimated using the Raman spectroscopy (Figure S4a). Two strong peaks are visible at 221.0 cm−1 and 310.8 cm−1, which reflect the presence of SnS and MoS2, respectively. In particular, two pronounced peaks appear at 1340.6 cm−1 and 1596.7 cm−1, which are attributed to the carbonaceous D band (carbon atom lattice defect) and G band (carbon atom sp2 hybridization), respectively. The value of intensity ratio (ID/IG) was thus calculated to be 1.21, which is indicative of a relatively low degree of graphitization. Furthermore, the exact carbon content in the composite was examined using Thermal gravimetric (TG) analysis. Two main stages of weight loss were clearly visible (Figure S4b): the first stage occurred before 450 °C, which involves both the removal of water and the oxidation of metal sulfide MoS2/SnS/CoS; the second stage ranged between 400 °C and 670 °C, which corresponds to the combustion of carbon. As a result, the carbonaceous content in the MoS2/SnS/CoS@rGO composite was calculated to be 18.6 wt%. In addition, the specific surface area and pore size distribution of MoS2/SnS/CoS@rGO were quantitatively determined by measuring N2 adsorption and desorption isotherms. The specific surface area of the composite was calculated to be 30.3 m2 g−1 according to the Brunauer–Emmett–Teller (BET) method (Figure S4c). The pore size distributions ranged from 3 and 10 nm (Figure S4d), clearly reflecting the mesoporous feature. The mesoporous size distribution, together with the relatively low specific surface area (30.3 m2 g−1), is well-recognized to hold the electrolyte and facilitate the Na+ diffusion, and thus to improve rate performance [48,49]. As a result, by combining the above-obtained characterization results, the MoS2/SnS/CoS@rGO composite was successfully obtained, and had the advantageous characteristics capable of boosting the sodium-storage performance: well-dispersed multi-component MoS2/SnS/CoS with the interfaces, highly electron conductive rGO support (18.6 wt%), appropriate specific surface area (30.3 m2 g−1) and mesoporous size distribution.
The electrochemical performance of MoS2/SnS/CoS@rGO was evaluated by measuring cyclic voltammetry (CV) curves (Figure S5) at a scan rate of 0.1 mV s−1 between 0.01 and 3.0 V. During the first reduction scan, the first reduction peak at 0.78 V is attributed to both the decomposition of the electrolyte and the formation of irreversible solid-electrolyte interface (SEI). Three peaks were observed at 0.21 V, 0.58 V, and 0.91 V, corresponding to the alloying reaction of Sn and Na and indicating that the alloying reaction is a multi-step process. Another reduction peak at 1.19 V is attributed to the process of converting from MoS2 to NaxMoS2. During the oxidation sweep, two peaks at 0.18 V and 0.70 V were visible, which are assigned to the multi-step Na-Sn dealloying reaction. The subsequent two broad peaks at 1.00 V and 1.37 V appeared, due to the conversion reaction of Sn and Na2S to SnS. Another oxidation peak at 1.82 V appeared, owing to the reversible transformation reaction between metallic Co, Mo and Na2S matrix. Consequently, the mechanism reactions for sodium storage are presented in the following equations with Gibbs (ΔG, kJ/mol).
NaxMoS2 + (4 − x) Na+ + (4 − x) e ⇌ Mo + 2Na2S  ∙∙ (ΔG = 58.672)
NaxMoS2 + (4 − x) Na+ + (4 − x) e ⇌ Mo + 2Na2S  ∙∙ (ΔG = 58.672)
CoS + 2Na+ + 2e ⇌ Co + Na2S  ∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙ (ΔG = 53.461)
4Sn + 15Na+ + 15e ⇌ Na15Sn4  ∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙∙ (Alloying reaction)
The first three galvanostatic charge/discharge curves were recorded for the MoS2/SnS/CoS@rGO composite electrode at 0.1 A g−1. Figure 4a shows that the MoS2/SnS/CoS@rGO composite achieves the first discharge/charge capacities of 875.0 mA h g−1 and 742.3 mA h g−1, generating an initial Coulombic efficiency (ICE) of 84.8%. Figure 4b shows the comparison of cycling performance between MoS2/SnS/CoS@rGO, CoS@rGO, MoS2/CoS@rGO and MoS2/SnS/CoS@rGO-M, at 0.1 A g−1. The MoS2/SnS/CoS@rGO composite delivers a high-reversible capacity of 627.6 mA h g−1 after 100 cycles at 0.1 A g−1, which is superior to those of CoS@rGO (325.2 mA h g−1) and MoS2/CoS@rGO (363.4 mAh g−1). By contrast, the specific capacity of the MoS2/SnS/CoS@rGO-M rapidly decayed to less than 600 mA h g−1 after 80 cycles at 0.1 A g−1. Note that an overlap occurs between the second and third discharge/charge profiles. Further comparison of reversible capacity shows that the capacity retentions of the 5th, 10th, 11th, 100th cycles were determined to be 93.0%, 89.2%, 89.2% and 84.3%, respectively, reflecting a decrease-then-quasi-stable cycling behavior after the completion of 100 cycles at 0.1 A g−1 (as shown in Figure 4b). The trend is attributed to the following possibilities. (i) The formation of the SEI film and the irreversible Na+ intercalation/extraction, (ii) the initial pulverization and the subsequently stable MoS2/SnS/CoS heterostructures on the underlying rGO support. Such a type of decrease-then-stable trend was also reported early for single-atom cobalt-doped MoS2/carbon [50], free-standing mesoporous Co(OH)x/Co2P [51], and hollow double-shell MnS@C nanospheres [52]. The comparison strongly suggests the satisfactory cycling stability of the MoS2/SnS/CoS@rGO electrode, which is predominantly attributed to the synergism between the multiple sulfides.
Rate capability was further evaluated for the electrodes at densities from 0.1 to 5.0 A g−1. Figure 4c shows that the MoS2/SnS/CoS@rGO composites afforded specific capacities of 632.2, 579.1, 537.4, 513.2, 496.3, 439.8 and 371.0 mA h g−1 at 0.1, 0.2, 0.5, 0.8, 1.0, 2.0 and 5.0 A g−1, respectively; and was almost capable of returning to the initial state at 0.1 A g−1. However, the three electrodes of CoS@rGO, MoS2/CoS@rGO, and MoS2/SnS/CoS@rGO -M all exhibited much lower specific capacities than those of the MoS2/SnS/CoS@rGO composite at the corresponding current densities. In particular, at 5.0 A g−1, the MoS2/SnS/CoS@rGO composite afforded a high-reversible-specific capacity of 304.9 mA h g−1 after 1000 cycles (Figure 4d), with Coulombic efficiencies of ca. 100% (as shown by the golden line in Figure 4d). However, the three electrodes of CoS@rGO, MoS2/CoS@rGO, and MoS2/SnS/CoS@rGO-M exhibited rapid degradation to 150 mA h g−1 after 400 cycles, 180.1 mA h g−1 after 500 cycles, and 137.2 mA h g−1 after 600 cycles, respectively. As a result, the MoS2/SnS/CoS@rGO possesses the greatly boosted specific capacity and long-cycle stability compared with those of the CoS@rGO, MoS2/CoS@rGO, and MoS2/SnS/CoS@rGO-M. Moreover, further comparison of cycling performance is summarized in Table S1 [27,36,37,38,39,40,53,54,55,56,57], which reflects that the MoS2/SnS/CoS@rGO composite exhibits a comparable cycling stability to those of the sulfide anode nanomaterials reported previously. In particular, the rate performance was compared between MoS2/SnS/CoS@rGO and those reported sulfide anode nanomaterials for SIBS [38,39,40,55,56,57,58,59], demonstrating the excellent rate capability of the MoS2/SnS/CoS@rGO (Figure S6).
The above-absolved electrochemical results demonstrate that the LDH-derived MoS2/SnS/CoS@rGO composite exhibits greatly enhanced electrochemical performances for SIBs compared with those of the three electrodes of the individual CoS@rGO, binary MoS2/CoS@rGO, and the mixed MoS2/SnS/CoS@rGO-M. The higher enhancement is mainly attributed to the following possibilities. The multiple-component MoS2/SnS/CoS presents the heterostructural nanodomains and thereby plays a synergistic role in boosting the electrochemical performances. It is well-known that the interface between different components could effectively reduce the kinetic barrier of Na+ diffusion, and thereby provide active sites for achieving efficient secondary ion storage [60,61,62]. To support the aspect, the Nyquist plots were measured by recording electrochemical impedance spectroscopy (EIS) for the electrodes. Obviously, the MoS2/SnS/CoS@rGO composite exhibits a much smaller semicircle and a larger slope than those of the other electrodes (Figure S7), strongly suggesting that the former electrode has a lower charge-transfer resistance and lower Na+ diffusion resistance compared with those of the latter electrodes.
Furthermore, comparison of XPS spectra was conducted between the LDH-derived MoS2/SnS/CoS@rGO and the MoS2/SnS/CoS@rGO-M. The main Sn 3d of MoS2/SnS/CoS@rGO shift to a lower binding energy compared with those of MoS2/SnS/CoS@rGO-M, reflecting that there is more charge density on the Sn atoms; whereas the main Mo 3d peaks shift to a higher bonding energy compared with those of MoS2/SnS/CoS@rGO-M, indicating that there is less charge density on the Mo atoms. The electron transfers between SnS and MoS2, resulting from the interface between MoS2 and SnS, are due to the lattice-confinement of the LDH precursor, indeed facilitating the electrochemical performance during the charging/discharging processes.
Moreover, the Na+ storage kinetics of MoS2/SnS/CoS@rGO was examined by measuring CV curves at different scan rates from 0.2 to 1.2 mV s−1. Figure 5a shows that both the cathodic and anodic peaks appeared to shift to increasingly low and high potentials, respectively. As a result, the slopes of the linear relationship between the logarithmic currents (i) and scan rates (v) at the redox peak potentials were calculated to be 0.87, 0.77, 0.82, and 0.76, respectively (Figure 5b), all of which are closely approximate to 1.0 and thus indicating that the whole electrochemical process of MoS2/SnS/CoS@rGO is mainly controlled by capacitive contribution. Furthermore, the capacitive contribution was quantitatively determined in accordance with the formula: i = k1v + k2v1/2, in which where k1v means the fraction of capacitive contribution. Consequently, the capacitive contribution at 1.2 mV s−1 was 76.3% (Figure 5c). At the varied scan rates from 0.2, 0.4, 0.6, 0.8, 1.0, to 1.2 mV s−1, the capacitive contribution was increased from 51.0%, 56.3%, 62.3%, 67.7%, 72.2%, to 76.3% (Figure 5d). The results demonstrate that the capacitive contributions contribute dominantly to the sodium-ion storage especially at high-current densities.

3. Materials and Methods

3.1. Preparation of CoMoSn-LDH/GO Precursor

The CoMoSn-LDH/GO precursor was synthesized using a traditional co-precipitation method (as shown in Scheme S1). A brown graphene sol (GO), with a concentration 6.73 mg mL−1, was synthesized using the modified Hummers method [63]. A 50 mg of GO sol was well dispersed into 100 mL of deionized water, and then ultrasonicated for 2 h to acquire a uniform 0.5 mg mL−1 sol. A mixed salt solution was prepared by dissolving Na2MoO4∙2H2O (2 mmol), SnCl2 (3 mmol) and Co(NO3)2·6H2O (6 mmol) into 50 mL of deionized water, respectively, followed by 2 h of ultrasonic to obtain a clear solution. Note that concentrated HCl was added drop wisely to prevent MoO42− from oxidizing Sn2+ to Sn4+ under the acidic condition. Subsequently, a mixed alkali solution was obtained by dissolving NaOH (40 mmol) and Na2CO3 (12 mmol) into 100 mL of deionized water under ultrasonication. The GO suspension was poured into a 500 mL four-necked flask, followed by dropping the mixed salt solution under mechanical stirring. After the completion of adding the mixed salt solution, the mixed alkali solution was also added drop wisely into the four-necked flask until the pH = 10.0 under subsequent stirring for 30 min. The as-prepared slurry was transferred quickly into a 100 mL Teflon-sealed autoclave, and kept for 24 h of aging at 120 °C. After the temperature was lowered naturally to room temperature, the CoMoSn-LDH/GO precursor was acquired by centrifugation and rinsing with the alternative deionized water/ethanol until pH = 7.0, followed by 12 h of drying in a vacuum oven at 60 °C.

3.2. Preparation of MoS2/SnS/CoS@rGO

As shown in Scheme S2, both 0.2 g of CoMoSn-LDH/GO precursor and 0.2 g thioacetamide (TAA, CH3CSNH2) were dispersed into 60 mL of deionized water, followed by 30 min of magnetic stirring. The mixed dispersion was then poured into a 100 mL autoclave and maintained at 200 ° C for 24 h. After the temperature cooled to room temperature, the precipitate was collected by centrifugation and rinsed alternatively with the deionized water and ethanol, and kept for 12 h of drying at 60 °C. The MoS2/SnS/CoS@rGO product was obtained by subjecting the above-prepared precipitate to 120 min of calcination at 300 °C in a corundum porcelain boat in Ar with a heating rate of 5 °C min−1.
For comparison, MoS2/CoS@rGO and CoS@rGO were prepared as counterparts under the same experimental conditions without adding SnCl2, or Na2MoO4∙2H2O/SnCl2, respectively. A MoS2/SnS/CoS@rGO-M composite was directly prepared under the same experimental conditions of Na2MoO4, Na2[Sn(OH)6], Co(NO3)2∙6H2O and CH3CSNH2, without involving the formation of the LDH precursor.

3.3. Materials Characterization

The crystal structure of the sample was tested using XRD (Bruker Discover, Bruker, Berlin, Germany), with Cu Kα as the emission source (λ = 0.15418 nm). The test conditions are 10–80° and the scanning speed is 10° min−1. The SEM (Zeiss Supreme 55, Zeiss, Oberkochen, Germany) was used to observe the microscopic topography and size of the sample. The microstructure of the sample as well as lattice fringes were analyzed using HRTEM (JEM-2100F, JEOL, Peabody, MA, USA). The elements on the sample surface were analyzed using the XPS (Axis Supra, Shimadzu, Kyoto, Japan) radiation source as the Al target with a vacuum degree of 2 × 10−9 Pa; Nitrogen isothermal adsorption and desorption test (BET) uses a fully automatic adsorption instrument (ASAP-2020, Mack, Greensboro, NC, USA) to analyze the specific surface area and pore size distribution of the sample. Carbon material defects in samples were analyzed using Raman spectroscopy (RM 20001, Renishaw, West Dundee, IL, USA). The excitation wavelength is 532 nm, and the scanning range is 1000–2000 cm−1.

3.4. Electrochemical Testing

All the half cells were subjected to electrochemical testing using the CR2032 coin cells that were assembled in an Ar-filled glove box (with H2O, O2 both below 0.1 ppm). The working electrodes were acquired by mixing the active materials (MoS2/SnS/CoS@rGO and its counterparts), Super P and poly(vinyl difluoride) (PVDF) at a mass ratio of 7:2:1, followed by casting homogenously onto copper foils. The loading of the active material was ca. 1.5 mg cm−2 on each electrode foil. The counter electrode and the separator involved were sodium foil and Whatman glass fiber, respectively. The electrolyte used was a mixture consisting of NaPF6 (1.0 M) and ethylene glycol/dimethyl ether (DG/DME), the latter of which were prepared at a 1:1 mass ratio. Both cyclic voltammogram (CV) curves and electrochemical impedance spectra (EIS) were acquired on a commercial 760E electrochemistry workstation (Shanghai, China). The CV curves were recorded at a scan rate of 0.1 mV s−1 within the voltage range from 0.01 to 3.0 V (vs. Na+/Na), while the EIS data were obtained within the frequency range between 100 kHz and 0.1 Hz at an open circuit potential. All galvanostatic discharge/charge profiles were recorded on a LANDCT-2001A battery tester.

4. Conclusions

We have demonstrated the MoS2/SnS/CoS@rGO composite as a high-rate anode nanomaterial for SIBs. (1) The composite was prepared via a distinctive lattice confinement of the ternary CoMoSn-LDH precursor, and endowed with the advantageous characteristics of boosting electrochemical performances: (i) well-distributed multi-component CoS/MoS2/SnS heterostructures with the interface interaction between Mo and Sn cations; (ii) highly electron conductive rGO support (18.6 wt.%), (iii) appropriate specific surface area (30.3 m2 g−1) and mesoporous size distribution. (2) Consequently, the composite delivered satisfactory sodium-storage performances: a high-reversible capacity of 620.3 mA h g−1 after 100 cycles at 0.1 A g−1, and excellent rate capacity of 371.0 mA h g−1 after 1000 cycles at 5.0 A g−1; both of which surpass those of the three counterparts of sing- and bi-sulfides, as well as the non-LDH-precursor-derived multiple sulfides. (3) Furthermore, the substantial enhancement was experimentally elaborated by the low charge-transfer resistance and diffusion resistance demonstrated using EIS testing, and the dominant capacitive contribution. The results show that the LDH precursor-based synthesis route could promise an effective approach to design and prepare well-dispersed multiple-component transition-metal compounds as anode nanomaterials for SIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28165972/s1, Figure S1: XRD pattern and SEM image of CoMoSn-LDH/GO; Figure S2: SEM image of MoS2/SnS/CoS@rGO; Figure S3: SEM/EDS image of each element of MoS2/SnS/CoS@rGO; Figure S4: Raman, TGA curve, N2 adsorption/desorption curves and pore size distribution of MoS2/SnS/CoS@rGO; Figure S5: The initial three CV curves of MoS2/SnS/CoS@rGO at 0.1 mV s−1; Figure S6: Comparison of the rate capability of MoS2/SnS/CoS@rGO with other reported works; Figure S7: Nyquist plots of MoS2/SnS/CoS@rGO, MoS2/CoS@rGO, MoS2/SnS/CoS@rGO-M; Scheme S1: Preparation illustration of CoMoSn-LDH/GO precursor; Scheme S2: Preparation illustration of MoS2/SnS/CoS@rGO; Table S1: Comparison of cycle performances.

Author Contributions

R.Z.: methodology, writing—original draft, writing—review and editing. Y.D.: writing—original draft. Y.S.: investigation, methodology. W.Z.: writing—review and editing. S.X.: conceptualization, funding acquisition, supervision, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Basic Science Center Project of the National Natural Science Foundation of China (Grant No. 22288102), the Beijing Natural Science Foundation (Grant No. 2222020), and the National Natural Science Foundation of China (Grant No. 22279006).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Most of the data used during the preparation of the manuscript are included in the Section 2 and Section 3.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Gervillié-Mouravieff, C.; Boussard-Plédel, C.; Huang, J.; Leau, C.; Blanquer, L.A.; Ben Yahia, M.; Doublet, M.-L.; Boles, S.T.; Zhang, X.H.; Adam, J.L.; et al. Unlocking cell chemistry evolution with operando fibre optic infrared spectroscopy in commercial Na(Li)-ion batteries. Nat. Energy 2022, 7, 1157–1169. [Google Scholar] [CrossRef]
  2. Su, R.; Zhu, W.; Liang, K.; Wei, P.; Li, J.; Liu, W.; Ren, Y. Mnx+ Substitution to Improve Na3V2(PO4)2F3-Based Electrodes for Sodium-Ion Battery Cathode. Molecules 2023, 28, 1409. [Google Scholar] [CrossRef] [PubMed]
  3. Tan, S.; Yang, H.; Zhang, Z.; Xu, X.; Xu, Y.; Zhou, J.; Zhou, X.; Pan, Z.; Rao, X.; Gu, Y.; et al. The Progress of Hard Carbon as an Anode Material in Sodium-Ion Batteries. Molecules 2023, 28, 3134. [Google Scholar] [CrossRef] [PubMed]
  4. Zhu, L.; Yin, B.; Zhang, Y.; Wu, Q.; Xu, H.; Duan, H.; Shi, M.; He, H. A Multifunctional Coating on Sulfur-Containing Carbon-Based Anode for High-Performance Sodium-Ion Batteries. Molecules 2023, 28, 3335. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, X.; Wu, X.-Y.; Chang, B.; Wang, K.-X. Recent progress on germanium-based anodes for lithium ion batteries: Efficient lithiation strategies and mechanisms. Energy Storage Mater. 2020, 30, 146–169. [Google Scholar] [CrossRef]
  6. Song, N.-J.; Guo, N.; Ma, C.; Zhao, Y.; Li, W.; Li, B. Modulating the Graphitic Domains and Pore Structure of Corncob-Derived Hard Carbons by Pyrolysis to Improve Sodium Storage. Molecules 2023, 28, 3595. [Google Scholar] [CrossRef]
  7. Sun, X.; Luo, F. Sodium Storage Properties of Carbonaceous Flowers. Molecules 2023, 28, 4753. [Google Scholar] [CrossRef]
  8. Bommier, C.; Surta, T.W.; Dolgos, M.; Ji, X. New Mechanistic Insights on Na-Ion Storage in Nongraphitizable Carbon. Nano Lett. 2015, 15, 5888. [Google Scholar] [CrossRef]
  9. Xiao, B.; Soto, F.A.; Gu, M.; Han, K.S.; Song, J.; Wang, H.; Engelhard, M.H.; Murugesan, V.; Mueller, K.T.; Reed, D.; et al. Lithium-Pretreated Hard Carbon as High-Performance Sodium-Ion Battery Anodes. Adv. Energy Mater. 2018, 8, 1801441. [Google Scholar] [CrossRef]
  10. Yao, X.; Ke, Y.; Ren, W.; Wang, X.; Xiong, F.; Yang, W.; Qin, M.; Li, Q.; Mai, L. Defect-Rich Soft Carbon Porous Nanosheets for Fast and High-Capacity Sodium-Ion Storage. Adv. Energy Mater. 2018, 9, 1803260. [Google Scholar] [CrossRef]
  11. Zhong, X.; Wu, Y.; Zeng, S.; Yu, Y. Carbon and Carbon Hybrid Materials as Anodes for Sodium-Ion Batteries. Chem. Asian J. 2018, 13, 1248–1265. [Google Scholar] [CrossRef]
  12. Choi, Y.S.; Byeon, Y.W.; Park, J.H.; Seo, J.H.; Ahn, J.P.; Lee, J.C. Ultrafast Sodiation of Single-Crystalline Sn Anodes. ACS Appl. Mater. Interfaces 2018, 10, 560–568. [Google Scholar] [CrossRef]
  13. Tan, H.; Chen, D.; Rui, X.; Yu, Y. Peering into Alloy Anodes for Sodium-Ion Batteries: Current Trends, Challenges, and Opportunities. Adv. Funct. Mater. 2019, 29, 1808745. [Google Scholar] [CrossRef]
  14. Yang, M.; Liu, L.; Yan, H.; Zhang, W.; Su, D.; Wen, J.; Liu, W.; Yuan, Y.; Liu, J.; Wang, X. Porous nitrogen-doped Sn/C film as free-standing anodes for lithium ion batteries. Appl. Surf. Sci. 2021, 551, 149246. [Google Scholar] [CrossRef]
  15. Chai, W.; Yang, F.; Yin, W.; You, S.; Wang, K.; Ye, W.; Rui, Y.; Tang, B. Bi2S3/C nanorods as efficient anode materials for lithium-ion batteries. Dalton Trans. 2019, 48, 1906–1914. [Google Scholar] [CrossRef]
  16. Chen, Q.; Sun, S.; Zhai, T.; Yang, M.; Zhao, X.; Xia, H. Yolk-Shell NiS2 Nanoparticle-Embedded Carbon Fibers for Flexible Fiber-Shaped Sodium Battery. Adv. Energy Mater. 2018, 8, 1800054. [Google Scholar] [CrossRef]
  17. Liu, X.; Li, S.; Zai, J.; Jin, Y.; Zhan, P.; Huang, Y.; Tie, X.; Qi, R.; Qian, X. Flower-like SnS2 composite with 3D pyrolyzed bacterial cellulose as the anode for lithium-ion batteries with ultralong cycle life and superior rate capability. Dalton Trans. 2019, 48, 833–838. [Google Scholar] [CrossRef]
  18. Luo, F.; Ma, D.; Li, Y.; Mi, H.; Zhang, P.; Luo, S. Hollow Co3S4/C anchored on nitrogen-doped carbon nanofibers as a free-standing anode for high-performance Li-ion batteries. Electrochim. Acta 2019, 299, 173–181. [Google Scholar] [CrossRef]
  19. Yang, M.; Chang, X.; Wang, L.; Wang, X.; Gu, M.; Huang, H.; Tang, L.; Zhong, Y.; Xia, H. Interface Modulation of Metal Sulfide Anodes for Long-Cycle-Life Sodium-Ion Batteries. Adv. Mater. 2023, 35, 2208705. [Google Scholar] [CrossRef]
  20. Lim, Y.V.; Li, X.L.; Yang, H.Y. Recent Tactics and Advances in the Application of Metal Sulfides as High-Performance Anode Materials for Rechargeable Sodium-Ion Batteries. Adv. Funct. Mater. 2020, 31, 2006761. [Google Scholar] [CrossRef]
  21. Xiao, F.; Yang, X.; Wang, H.; Yu, D.Y.W.; Rogach, A.L. Hierarchical CoS2/N-Doped Carbon@MoS2 Nanosheets with Enhanced Sodium Storage Performance. ACS Appl. Mater. Interfaces 2020, 12, 54644–54652. [Google Scholar] [CrossRef] [PubMed]
  22. Shen, J.; Wu, N.; Xie, W.; Li, Q.; Guo, D.; Li, J.; Liu, G.; Liu, X.; Mi, H. Realizing Ultrafast and Robust Sodium-Ion Storage of Iron Sulfide Enabled by Heteroatomic Doping and Regulable Interface Engineering. Molecules 2023, 28, 3757. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, Y.; Yang, C.; Zhang, Q.; Liu, M. Recent progress in the design of metal sulfides as anode materials for sodium ion batteries. Energy Storage Mater. 2019, 22, 66–95. [Google Scholar] [CrossRef]
  24. Sennua, P.; Christyb, M.; Aravindanc, V.; Lee, Y.-G.; Nahmb, K.S.; Lee, Y.-S. Two Dimensional Mesoporous Cobalt Sulfide Nanosheets as a Superior Anode for a Li-Ion Battery and a Bifunctional Electrocatalyst for the Li-O2 System. Chem. Mater. 2015, 27, 5726–5735. [Google Scholar] [CrossRef]
  25. Wang, Q.; Zou, R.; Xia, W.; Ma, J.; Qiu, B.; Mahmood, A.; Zhao, R.; Yang, Y.; Xia, D.; Xu, Q. Facile Synthesis of Ultrasmall CoS2 Nanoparticles within Thin N-Doped Porous Carbon Shell for High Performance Lithium-Ion Batteries. Small 2015, 11, 2511–2517. [Google Scholar] [CrossRef]
  26. Wu, R.; Wang, D.P.; Rui, X.; Liu, B.; Zhou, K.; Law, A.W.; Yan, Q.; Wei, J.; Chen, Z. In-situ formation of hollow hybrids composed of cobalt sulfides embedded within porous carbon polyhedra/carbon nanotubes for high-performance lithium-ion batteries. Adv. Mater. 2015, 27, 3038–3044. [Google Scholar] [CrossRef] [PubMed]
  27. Fu, Y.; Li, J.; Wang, H.; Zhang, J.; Ma, Z.; Yi, Q.; Liu, J.; Wang, S. Synthesis of FeS2/CoS heterostructure microspheres as anodes for high-performance Li-ion batteries. J. Solid. State Electrochem. 2023. [Google Scholar] [CrossRef]
  28. Zhou, Q.; Liu, L.; Huang, Z.; Yi, L.; Wang, X.; Cao, G. Co3S4@polyaniline nanotubes as high-performance anode materials for sodium ion batteries. J. Mater. Chem. A 2016, 4, 5505–5516. [Google Scholar] [CrossRef]
  29. He, Q.; Xiao, F.; Chen, R.; Yao, T.; Wu, Y.; Wang, H. Covalent sulfur/CoS2-decorated honeycomb-like carbon hierarchies for high-performance sodium storage with ultra-long high-rate cycling stability. J. Alloys Compd. 2023, 960, 170866. [Google Scholar] [CrossRef]
  30. Mu, M.; He, H.; Gan, Y.; Yu, J.; Mou, J.; Yuan, J.; Li, F.; Wang, X.; Li, X.; Zhang, X.; et al. Three-dimensional porous C/CoS nanocomposites for a long-life and high-rate potassium storage. Mater. Res. Bull. 2023, 165, 112337. [Google Scholar] [CrossRef]
  31. Wang, S.; Zhou, Z.; Wen, B.; Zhang, Z.; Yang, G.; Yan, W. Controlled growth of Co9S8 nanoparticle-embedded carbon nanosheets/carbon nanofibers toward high-performance sodium storage. J. Colloid Interface Sci. 2023, 648, 644–653. [Google Scholar] [CrossRef] [PubMed]
  32. Li, X.; Zhang, W.; Kang, H.; Liu, H.; Yang, B.; Li, Z.; Li, Z. Self-assembled CoS2/NiCo2S4/RGO nanohybrids as advanced electrode for hybrid supercapacitor with enhanced energy density and ultra-long durability. J. Energy Storage 2023, 67, 107528. [Google Scholar] [CrossRef]
  33. Wang, Y.; Sun, C.; Zhao, H.; Cui, J.; Zhang, Y.; He, W. A cycle-durable hollow nanoneedle structure with a nanosheet as a conductive substrate CoS1.097/Ni9S8@RGO to enhance supercapacitor performance. Dalton Trans. 2023, 52, 7731–7738. [Google Scholar] [CrossRef] [PubMed]
  34. Cao, S.; Liu, Q.; Chen, H.; Li, G.; Zhu, H. Quadra layer core-shell-core-shell structured Co9S8@MnS-NC nanorod composites for enhanced sodium-ion batteries. Electrochim. Acta 2023, 459, 142569. [Google Scholar] [CrossRef]
  35. Li, Y.; Zan, L.; Chen, J. Improving the Reaction Kinetics by Annealing MoS2/PVP Nanoflowers for Sodium-Ion Storage. Molecules 2023, 28, 2948. [Google Scholar] [CrossRef] [PubMed]
  36. Xue, F.; Fan, F.; Zhu, Z.; Zhang, Z.; Gu, Y.; Li, Q. MoS2/CoS heterostructures grown on carbon cloth as free-standing anodes for high-performance sodium-ion batteries. Nanoscale 2023, 15, 6822–6829. [Google Scholar] [CrossRef]
  37. Guo, Q.; Ma, Y.; Chen, T.; Xia, Q.; Yang, M.; Xia, H.; Yu, Y. Cobalt Sulfide Quantum Dot Embedded N/S-Doped Carbon Nanosheets with Superior Reversibility and Rate Capability for Sodium-Ion Batteries. ACS Nano 2017, 11, 12658–12667. [Google Scholar] [CrossRef]
  38. Qi, Y.; Zhang, T.; Wu, N.; Ding, H.; Yu, G.; Lian, J.; Xu, L.; Qiu, J.; Li, S. Rational Design of the CoS/Co9S8@NC Composite Enabling High-Rate Sodium-Ion Storage. ACS Appl. Energy Mater. 2021, 4, 5574–5582. [Google Scholar] [CrossRef]
  39. Wan, S.; Cheng, M.; Chen, H.; Zhu, H.; Liu, Q. Nanoconfined bimetallic sulfides (CoSn)S heterostructure in carbon microsphere as a high-performance anode for half/full sodium-ion batteries. J. Colloid Interface Sci. 2022, 609, 403–413. [Google Scholar] [CrossRef]
  40. Huang, X.; Tao, K.; Han, T.; Li, J.; Zhang, H.; Hu, C.; Niu, J.; Liu, J. Long-Cycling-Life Sodium-Ion Battery Using Binary Metal Sulfide Hybrid Nanocages as Anode. Small 2023, 19, 2302706. [Google Scholar] [CrossRef]
  41. Duan, Y.-K.; Li, Z.-W.; Zhang, S.-C.; Su, T.; Zhang, Z.-H.; Jiao, A.-J.; Fu, Z.-H. Stannate-Based Materials as Anodes in Lithium-Ion and Sodium-Ion Batteries: A Review. Molecules 2023, 28, 5037. [Google Scholar] [CrossRef]
  42. Sideris, P.J.; Nielsen, U.G.; Gan, Z.; Grey, C.P. Mg/Al Ordering in Layered Double Hydroxides Revealed by Multinuclear NMR Spectroscopy. Science 2008, 321, 113. [Google Scholar] [CrossRef] [Green Version]
  43. Zhang, X.; Cui, G.; Feng, H.; Chen, L.; Wang, H.; Wang, B.; Zhang, X.; Zheng, L.; Hong, S.; Wei, M. Platinum-copper single atom alloy catalysts with high performance towards glycerol hydrogenolysis. Nat. Commun. 2019, 10, 5812. [Google Scholar] [CrossRef] [Green Version]
  44. Zhao, X.; Zhang, F.; Xu, S.; Evans, D.G.; Duan, X. From Layered Double Hydroxides to ZnO-based Mixed Metal Oxides by Thermal Decomposition: Transformation Mechanism and UV-Blocking Properties of the Product. Chem. Mater. 2010, 22, 3933–3942. [Google Scholar] [CrossRef]
  45. Liu, S.; Dong, Y.; Deng, C.; Chen, F.; Su, Y.; Li, S.Y.; Xu, S. Low-content SnO2 nanodots on N-doped graphene: Lattice-confinement preparation and high-performance lithium/sodium storage. Dalton Trans. 2023, 52, 1642–1649. [Google Scholar] [CrossRef]
  46. Mohili, R.; Hemanth, N.R.; Jin, H.; Lee, K.; Chaudhari, N. Emerging high entropy metal sulphides and phosphides for electrochemical water splitting. J. Mater. Chem. A 2023, 11, 10463–10472. [Google Scholar] [CrossRef]
  47. Wu, C.; Zhang, M.; Chen, F.; Kang, H.; Xu, S.; Xu, S. IrCo alloy nanoparticles supported on N-doped carbon for hydrogen evolution electrocatalysis in acidic and alkaline electrolytes. Dalton Trans. 2020, 49, 13339–13344. [Google Scholar] [CrossRef]
  48. Lou, S.J.; Xu, D.Q.; Xu, Z.Y. Mild and versatile nitrate-promoted C-H bond fluorination. Angew. Chem. Int. Ed. Engl. 2014, 53, 10330–10335. [Google Scholar] [CrossRef]
  49. Sun, N.; Qiu, J.; Xu, B. Understanding of Sodium Storage Mechanism in Hard Carbons: Ongoing Development under Debate. Adv. Energy Mater. 2022, 27, 12. [Google Scholar] [CrossRef]
  50. Wen, X.; Feng, W.; Li, X.; Yang, J.; Du, R.; Wang, P.; Li, H.; Song, L.; Wang, Y.; Cheng, M.; et al. Diatomite-Templated Synthesis of Single-Atom Cobalt-Doped MoS2/Carbon Composites to Boost Sodium Storage. Adv. Mater. 2023, 2211690. [Google Scholar] [CrossRef]
  51. Wu, M.; Wang, J.; Liu, Z.; Liu, X.; Duan, J.; Yang, T.; Lan, J.; Tan, Y.; Wang, C.; Chen, M.; et al. Engineering Co—P Alloy Foil to a Well-Designed Integrated Electrode Toward High-Performance Electrochemical Energy Storage. Adv. Mater. 2023, 35, 2209924. [Google Scholar] [CrossRef]
  52. Ma, Z.; Song, A.; Liu, Z.; Guo, Y.; Yang, X.; Li, Q.; Fan, Y.; Dai, L.; Tian, H.; Qin, X.; et al. Nanoconfined Expansion Behavior of Hollow MnS@Carbon Anode with Extended Lithiation Cyclic Stability. Adv. Energy Mater. 2023, 28, 33. [Google Scholar] [CrossRef]
  53. Deng, Q.; Liu, X.; Li, Z.; Fan, H.; Zhang, Y.; Yang, H.Y. Cobalt-nickel bimetallic sulfide (NiS2/CoS2) based dual-carbon framework for super sodium ion storage. J. Colloid Interf. Sci. 2023, 633, 480–488. [Google Scholar] [CrossRef]
  54. Zhang, Y.; Lv, C.; Wang, X.; Chen, S.; Li, D.; Peng, Z.; Yang, D. Boosting Sodium-Ion Storage by Encapsulating NiS (CoS) Hollow Nanoparticles into Carbonaceous Fibers. ACS Appl. Mater. Interfaces 2018, 10, 40531–40539. [Google Scholar] [CrossRef]
  55. Fang, Y.; Guan, B.Y.; Luan, D.; Lou, X.W.D. Synthesis of CuS@CoS2 Double-Shelled Nanoboxes with Enhanced Sodium Storage Properties. Angew. Chem. Int. Ed. 2019, 58, 7739–7743. [Google Scholar] [CrossRef] [PubMed]
  56. Li, X.; Liang, H.; Qin, B.; Wang, M.; Zhang, Y.; Fan, H. Rational design of heterostructured bimetallic sulfides (CoS2/NC@VS4) with VS4 nanodots decorated on CoS2 dodecahedron for high-performance sodium and potassium ion batteries. J. Colloid Interf. Sci. 2022, 625, 41–49. [Google Scholar] [CrossRef]
  57. Liu, C.; Lu, Q.; Gorbunov, M.V.; Omar, A.; Martinez, I.G.G.; Zhao, P.; Hantusch, M.; Permana, A.D.C.; He, H.; Gaponik, N.; et al. Ultrasmall CoS nanoparticles embedded in heteroatom-doped carbon for sodium-ion batteries and mechanism explorations via synchrotron X-ray techniques. J. Energy Chem. 2023, 79, 373–381. [Google Scholar] [CrossRef]
  58. Liu, J.; Xu, Y.G.; Kong, L.B. Synthesis of polyvalent ion reaction of MoS2/CoS2-RGO anode materials for high-performance sodium-ion batteries and sodium-ion capacitors. J. Colloid Interf. Sci. 2020, 575, 42–53. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Zhan, R.; Xu, Q.; Liu, H.; Tao, M.; Luo, Y.; Bao, S.; Li, C.; Xu, M. Circuit board-like CoS/MXene composite with superior performance for sodium storage. Chem. Eng. J. 2019, 357, 220–225. [Google Scholar] [CrossRef]
  60. Chen, C.; Hu, X.; Zhang, B.; Miaob, L.; Huang, Y. Architectural design and phase engineering of N/Bcodoped TiO2(B)/anatase nanotube assemblies for high-rate and long-life lithium storage. J. Mater. Chem. A 2015, 3, 22591. [Google Scholar] [CrossRef]
  61. Bai, D.; Wang, F.; Lv, J.; Zhang, F.; Xu, S. Triple-Confined Well-Dispersed Biactive NiCo2S4/Ni0.96S on Graphene Aerogel for High-Efficiency Lithium Storage. ACS Appl. Mater. Interfaces 2016, 8, 32853–32861. [Google Scholar] [CrossRef]
  62. Wang, R.; Xue, X.; Lu, W.; Liu, H.; Lai, C.; Xi, K.; Che, Y.; Liu, J.; Guo, S.; Yang, D. Tuning and understanding the phase interface of TiO2 nanoparticles for more efficient lithium ion storage. Nanoscale 2015, 7, 12833–12838. [Google Scholar] [CrossRef]
  63. Zhang, H.; Zhang, W.; Huang, F. Graphene inducing graphitization: Towards a hard carbon anode with ultrahigh initial coulombic efficiency for sodium storage. Chem. Eng. J. 2022, 434, 134503. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the preparation of MoS2/SnS/CoS@rGO composite derived from a single-source ternary CoMoSn-LDH/GO precursor. The checkmark shows the concept of cationic dispersion, which is based on the Mg/Al ordering in MgAl-LDHs [42].
Figure 1. Schematic diagram of the preparation of MoS2/SnS/CoS@rGO composite derived from a single-source ternary CoMoSn-LDH/GO precursor. The checkmark shows the concept of cationic dispersion, which is based on the Mg/Al ordering in MgAl-LDHs [42].
Molecules 28 05972 g001
Figure 2. The MoS2/SnS/CoS@rGO composite: (a) XRD patterns, (b,c) TEM images, and (d) HRTEM images.
Figure 2. The MoS2/SnS/CoS@rGO composite: (a) XRD patterns, (b,c) TEM images, and (d) HRTEM images.
Molecules 28 05972 g002
Figure 3. XPS spectra for the MoS2/SnS/CoS@rGO composite: (a) a full survey spectrum, (b) Co 2p, (c) Mo 3d, (d) Sn 3d, (e) S 2p, and (f) C 1s.
Figure 3. XPS spectra for the MoS2/SnS/CoS@rGO composite: (a) a full survey spectrum, (b) Co 2p, (c) Mo 3d, (d) Sn 3d, (e) S 2p, and (f) C 1s.
Molecules 28 05972 g003
Figure 4. Comparison of electrochemical performance for SIBs. (a) The first three Galvanostatic discharge/charge profiles of MoS2/SnS/CoS@rGO at 0.1 A g−1, (b) cycling performance at 0.1 A g−1, (c) rate performances from 0.1 to 5.0 A g−1, and (d) long-life cycling performance at 5.0 A g−1.
Figure 4. Comparison of electrochemical performance for SIBs. (a) The first three Galvanostatic discharge/charge profiles of MoS2/SnS/CoS@rGO at 0.1 A g−1, (b) cycling performance at 0.1 A g−1, (c) rate performances from 0.1 to 5.0 A g−1, and (d) long-life cycling performance at 5.0 A g−1.
Molecules 28 05972 g004
Figure 5. The Na+ kinetics analysis of MoS2/SnS/CoS@rGO electrode. (a) CV curves at different scan rates between 0.2 and 1.2 mV s−1, (b) the linear relationship between log(i) and log(ν), (c) capacitive contribution at 1.2 mV s−1, and (d) capacitive contributions at various scan rates from 0.2 to 1.2 mV s−1.
Figure 5. The Na+ kinetics analysis of MoS2/SnS/CoS@rGO electrode. (a) CV curves at different scan rates between 0.2 and 1.2 mV s−1, (b) the linear relationship between log(i) and log(ν), (c) capacitive contribution at 1.2 mV s−1, and (d) capacitive contributions at various scan rates from 0.2 to 1.2 mV s−1.
Molecules 28 05972 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, R.; Dong, Y.; Su, Y.; Zhai, W.; Xu, S. MoS2/SnS/CoS Heterostructures on Graphene: Lattice-Confinement Synthesis and Boosted Sodium Storage. Molecules 2023, 28, 5972. https://doi.org/10.3390/molecules28165972

AMA Style

Zhang R, Dong Y, Su Y, Zhai W, Xu S. MoS2/SnS/CoS Heterostructures on Graphene: Lattice-Confinement Synthesis and Boosted Sodium Storage. Molecules. 2023; 28(16):5972. https://doi.org/10.3390/molecules28165972

Chicago/Turabian Style

Zhang, Ruyao, Yan Dong, Yu Su, Wenkai Zhai, and Sailong Xu. 2023. "MoS2/SnS/CoS Heterostructures on Graphene: Lattice-Confinement Synthesis and Boosted Sodium Storage" Molecules 28, no. 16: 5972. https://doi.org/10.3390/molecules28165972

Article Metrics

Back to TopTop