Xingzhong Guo*,
Zichen Wang,
Wenjun Zhu and
Hui Yang*
School of Materials Science and Engineering, Zhejiang University, Xihu District, Hangzhou, 310027, China. E-mail: msewj01@zju.edu.cn; yanghui@zju.edu.cn; Fax: +86-571-8795-3054; Tel: +86-571-87953313 Tel: +86-571-8795-1408
First published on 30th January 2017
Multilayer molybdenum disulfide (MoS2) was facilely prepared by a chelation-assisted sol–gel method with ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O) as the molybdenum source, thioacetamide (CH3CSNH2) as the sulfur source and diethylenetriamine pentaacetic acid (Dtpa) as the chelating agent, subsequently followed by high-temperature calcination. The chelating agent Dtpa ingeniously mediated the chelation reaction of the system and promoted the formation of a monolithic gel. The hexagonal MoS2 crystal (2H-MoS2) with good crystallinity precipitated after calcination at 1000 °C with the Mo and S mass ratio of 1:3. The adjustable MoS2 layers stacked together to form MoS2 flakes, and these flakes aggregated to construct crystalline MoS2 particles. The electrochemical tests showed the possibility of as-prepared MoS2 crystals applied as a negative electrode for lithium ion batteries.
Up to now, a series of MoS2 with different nanostructures such as nanosheets, nanoflowers, microspheres and monolayered MoS2 sheets have been used in lithium ion battery storage due to the high reversible capacity (up to 1290 mA h g−1). Ding et al. successfully prepared MoS2 microspheres with good initial discharge and charge capacities (1160 and 791 mA h g−1) and a capacity of 672 mA h g−1 after 50 cycles.5 Lin et al. produced sulfur-depleted monolayered MoS2 nanocrystals by exfoliating and disintegrating the bulk MoS2 and the products showed excellent catalytic performance on HER.6 Chhowalla et al. demonstrated that metallic 1T phase MoS2 nanosheets prepared by chemical exfoliation can intercalate ions such as H+, Li+ and Na+ efficiently and reach capacitance values from 400 to ∼700 F cm−3.7 Hu et al. synthesized MoS2 nanoflowers with expanded interlayers and used the products as Na-ion battery anode with high discharge capacities and good rate capability.8 Wang et al. prepared single-layer MoS2/graphene composites as the anode electrode of lithium ion battery with better cycle performance and rate capability than pure MoS2 electrode due to the good electron conductivity of graphene and the synergy effect between MoS2 and graphene.38
There have been various synthetic methods to prepare MoS2 and its composites, including hydrothermal synthesis, solid-state process, chemical vapor deposition (CVD), etc.9–11 Wherein, hydrothermal synthesis is a common method for the preparation of well-crystallized MoS2 with various morphologies due to high temperature and pressure conditions.12,13 However those preparation methods almost are expensive, complicated and low productive. Sol–gel method is a new synthetic route for advanced materials based on low-temperature, high production, moderate reacting conditions and molecular-level mixture.14 So sol–gel method could be a cheap and simple way to prepare MoS2 crystal. At present, only Li et al. reported the preparation of porous MoS2 via a sol–gel route using (NH4)2Mo3S13 as precursor.15 However, the dissolving reaction of the precursor (NH4)2Mo3S13 to prepare “Mo3S12” gel is not supposed to be real sol–gel process. In addition, the precursor (NH4)2Mo3S13 is quite rare and very hard to synthesize. Therefore, it is essential to develop an effective and low-cost sol–gel approach to prepare MoS2 materials.
In the present work we demonstrate a novel and facile chelation-assisted sol–gel method to prepare multilayer MoS2. The common and inexpensive ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O), thioacetamide (CH3CSNH2) and diethylenetriamine pentaacetic acid (Dtpa) were employed as molybdenum source, sulfur source and chelating agent, respectively. The mass ratio of Mo and S has an important role on the formation of multilayer MoS2 crystalline phase. The synthesis mechanism, surface morphology, crystal structure and electrochemical performances of the as-prepared MoS2 were also studied.
The XRD patterns of as-prepared samples show the precipitates are hexagonal 2H-MoS2 (JCPDS 37-1492) with different crystallinity (Fig. 2). It is obvious that the MoS2-2 sample displays the sharpest diffraction peaks, indicating the highest crystallinity. It reveals that complete MoS2 crystal tends to be formed when the mass ratio of Mo source and S source is 1:3. The diffraction peak of MoS2-2 sample at 2θ = 14.2° indicates the MoS2 layers stack orderly along (002) direction with a d-spacing of 0.62 nm. The (100) reflection at 2θ = 33.5° and (110) reflection at 2θ = 59.1° can be found in MoS2-3 sample but no clear (002) reflection. It is supposed that some complicated polysulfide molybdenum were synthesized owing to redundant S source in MoS2-3 sample and few MoS2 layers are stacked in the c direction.16 According to the Scherrer formula, D = Kλ/Bcosθ, K = 0.89, λ = 0.154056 nm, θ = 14.2°, B = 0.507 for MoS2-2 and B = 0.591 for MoS2-1. The average dimension of MoS2 in z-axis is roughly estimated at about 15 nm in MoS2-2 sample corresponding to 25 MoS2 layers, which was facilitated by high-temperature calcination. For MoS2-1 sample, the MoS2 grain dimension in z-axis is about 13 nm, corresponding to 21 MoS2 layers approximately. It shows that the mass ratio of Mo and S plays a role on the formation of multilayer MoS2 to some extent.
The three samples were further investigated to analyze the inner structure by Raman spectroscopy (Fig. 3). It has been proved that E12g and A1g peaks of monolayer MoS2 appear at 384.3 and 403 cm−1, respectively.17 With the increase of layer numbers, the in-plane E12g vibration weakens and the out-of-plane A1g vibration strengthens. When the layer number is over 6, the E12g and A1g peaks are observed at 382 and 408 cm−1 stably.18,19 It is seen that the MoS2-2 sample exhibits a strong out-of-plane vibration at 408 cm−1 and a relatively weaker in-plane vibration at 382 cm−1. However, the signal of this Raman spectroscopy is a little weak and red shifts are found in MoS2-1 and MoS2-3 samples. It is supposed that amorphous substances and defects in the resultant sample disorder the uniform MoS2 molecule vibrations, thus causing these phenomena. Due to the selection rules for scattering geometry and limited rejection of the Raleigh scattered radiation, the other two vibration modes E1g and E22g could not be detected.20 It proves that the mass ratio of Mo and S also impacts the inner structure of as-prepared MoS2.
The microstructures and morphologies of as-prepared MoS2-2 sample were observed by SEM and TEM (Fig. 4). It is seen that the MoS2 precipitates exist in the form of particles with the size of <2 μm and some agglomeration (Fig. 4a). Fig. 4b depicts irregular MoS2 flakes aggregate together to form larger particles. As shown in Fig. 4c, the MoS2 nanosheets stack together to form a large flake with different inside thickness. It is clearly observed from Fig. 4d that a nanosheet has some parallel MoS2 crystal fringes with an interlayer distance of 0.62 nm, corresponding to interplanar spacing of 2H-MoS2 (002) plane based on XRD results. The high magnification HR-TEM image of the thin MoS2 layers (Fig. 4e) confirms that the as-prepared MoS2 is typically hexagonal MoS2. The interlayer distance of (100) plane and (110) plane are 0.27 and 0.16 nm, respectively. The selected area electron diffraction (SAED) (Fig. 4f) shows a clear monocrystalline MoS2 diffraction pattern with six inner diffraction spots indicating (100) plane and the outer diffraction spots indicating (110) plane. There is no (002) plane diffraction spots because the electrons incident direction is [001]. Based on above analysis, during the formation of MoS2 particles, large amounts of monocrystalline MoS2 layers in different orientations stacked to form nanosheets, the multilayer nanosheets arranged to form MoS2 flakes, and then the flakes aggregated to construct crystalline MoS2 particles.
We used the sample MoS2-2 as the negative electrode of lithium ion battery because of its superior structure and composition. Fig. 5 depicts the cyclic voltammetry (CV) profile of as-prepared sample MoS2-2. Cyclic voltammetry is a normal method to research the redox reactions of the electrodes. From Fig. 5, two obvious reduction peaks at 0.6 V and 0.4 V are shown in the first cycle. The peak at 0.6 V implies intercalation of lithium ions into MoS2 layers with MoS2 structure transformation from 2H (trigonal prismatic coordination) to the 1T (octahedral coordination).21–23 The other peak at 0.4 V can be attributed to the conversion reaction process of LixMoS2 into Mo and Li2S. In the anodic sweep, the peak in 2.25 V is attributed to the delithiation of Li2S with the reaction process of Li2S − 2e → 2Li+ + S.7,24,25 In the second and third cathodic sweep, three reduction peaks are found at 1.7 V, 1.0 V, and 0.2 V, respectively, which could be due to the following reactions: 2Li+ + S + 2e → Li2S, MoS2 + xLi+ + xe → LixMoS2, and LixMoS2 + (4 − x)Li+ + (4 − x)e → Mo + 2Li2S.26,27
Fig. 6 shows the first three charge–discharge profiles of the sample MoS2-2 electrodes with a cutoff voltage of 0.005–3 V at a current density of 100 mA g−1, which are nearly accord with the CV measurements described above. In the first discharge process, there are two obvious voltage plateaus at 0.5 V and 1.0 V, which indicates the formation of LixMoS2 and the following conversion reaction of LixMoS2 into Mo and Li2S.1,17,28 The slope region below 0.5 V could be attributed to the formation of a solid-electrolyte interphase (SEI) layer.29,30 In the second and third discharge process, three vague potential plateaus could be found at 1.7 V, 1.0 V and 0.25 V, which are in qualitative agreement with the CV results. In the charge process, the MoS2-2 electrode shows a distinct potential plateaus at 2.25 V, which could be due to the reduction of sulfur to polysulfide.24,31,32 Fig. 6 also shows that the MoS2-2 electrode delivers an initial discharge capacity of 1149 mA h g−1 and a reversible charge capacity of 1038 mA h g−1, with a high coulombic efficiency of 90.3%.
Fig. 6 Galvanostatic charge and discharge curves of as-prepared sample MoS2-2 at a current density of 100 mA g−1. |
Fig. 7a depicts the cycling behavior and rate capability of the sample MoS2-2 at a constant current density of 100 mA g−1. Although the first charge and discharge capacities are higher than 900 mA h g−1, the cycling stability of the MoS2-2 electrode is poor with a discharge capacity decrease from 947 to 353 mA h g−1 after 40 cycles. Fig. 7b shows the rate cycling behavior of the sample MoS2-2 electrode. At the current densities of 1.0 A g−1, the capacity rapidly declines below 300 mA h g−1, which cannot compete with the MoS2/graphene composite electrode with good rate performance. Surprisingly, the sample MoS2-2 electrode has a high coulombic efficiency of nearly 100%.
Fig. 7 (a) Cycle performance of as-prepared sample MoS2-2 electrode at a current density of 0.1 A g−1, and (b) their rate capabilities at different current densities. |
Electrochemical impedance spectra (EIS) can be applied to better understand the electrochemical performance of MoS2 electrode. Fig. 8a depicts the Nyquist plots of MoS2-2 electrode after 40th cycle, and Fig. 8b is the equivalent circuit model for the impedance response. The semicircle in high-frequency region is due to the resistance Rf and CPE1 of the solid electrolyte interphase (SEI) film.33,34 The medium-frequency semicircle corresponds to the charge-transfer resistance Rct and CPE2 of the electrode/electrolyte interface.35,36 The inclined line is assigned to the lithium-diffusion within the electrode material.37 According to EIS equivalent circuit in Fig. 8b, Rf and Rct of MoS2-2 electrode are 14.84 and 121.9 Ω, respectively.
This journal is © The Royal Society of Chemistry 2017 |