Next Article in Journal
A Multi-Layer Multi-Pass Weld Bead Cross-Section Morphology Extraction Method Based on Row–Column Grayscale Segmentation
Previous Article in Journal
Excellent Electrochromic Properties of Ti4+-Induced Nanowires V2O5 Films
Previous Article in Special Issue
Progress of Metal Chalcogenides as Catalysts for Efficient Electrosynthesis of Hydrogen Peroxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistically Boosting Li Storage Performance of MnWO4 Nanorods Anode via Carbon Coating and Additives

1
Key Laboratory of Functional Materials Physics and Chemistry of Ministry of Education, Jilin Normal University, Changchun 130103, China
2
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(19), 4682; https://doi.org/10.3390/ma17194682
Submission received: 9 July 2024 / Revised: 5 September 2024 / Accepted: 14 September 2024 / Published: 24 September 2024

Abstract

:
Polyanionic structures, (MO4)n−, can be beneficial to the transport of lithium ions by virtue of the open three-dimensional frame structure. However, an unstable interface suppresses the life of the (MO4)n−-based anode. In this study, MnWO4@C nanorods with dense nanocavities have been synthesized through a hydrothermal route, followed by a chemical deposition method. As a result, the MnWO4@C anode exhibits better cycle and rate performance than MnWO4 as a Li-ion battery; the capacity is maintained at 718 mAh g−1 at 1000 mA g−1 after 400 cycles because the transport of lithium ions and the contribution of pseudo-capacitance are increased. Generally, benefiting from the carbon shell and electrolyte additive (e.g., FEC), the cycle performance of the MnWO4@C electrode is also effectively improved for lithium storage.

1. Introduction

In recent years, due to the theoretical capacity of commercial graphite negative electrodes being only 372 mAh g−1, they are no longer able to meet the requirements of the new generation of lithium-ion batteries as high energy density and high power density negative electrodes. Therefore, scientists are exploring a new type of negative electrode material with high specific energy and high energy density to replace graphite negative electrodes [1,2,3,4].
Metal tungstates (MWO4, where M is a 3D divalent transition metal ion with an ionic radius < 1 A°) with the wolframite type of structure can be described as made up of hexagonal close-packed oxygen atoms with certain octahedral sites filled with M2+ and W6+ cations in an ordered way [5,6]. A large amount of research has been conducted on this series of compounds in many application fields, but there is limited literature on the study of metal tungstate salts in the field of electrochemistry [7,8,9,10,11,12,13,14,15]. Fortunately, transition metal oxides have advantages, such as high specific capacity, lower operating voltage, and environmental friendliness, and are considered one of the most promising negative electrode materials. In addition, their preparation cost is low, and they have good economic benefits [16,17,18,19,20,21,22]. The open 3D frameworks containing (MO4)n− polyanions instead of the smaller O2− ions may allow for fast Li+ ion conduction.
In accordance with the aforementioned analysis, MnWO₄@C composites are successfully synthesized in this study via a solvothermal reaction. Surprisingly, the MnWO4@C electrode shows excellent electrochemical performance with a high capacity of 795 mAh g−1 at 200 mA g−1. And the MnWO4@C electrode also shows a superior capacity retention at various current densities from 100 to 5000 mA g−1. While the electrolyte additive can further improve the long-cycle stability of the MnWO4@C electrode, at 1000 mA g−1, the capacity of the MnWO4@C-FEC electrode is maintained at 718 mAh g−1 after 400 cycles.

2. Experimental Section

2.1. Materials Preparation

2.1.1. Materials

Manganese chloride tetrahydrate (MnCl2·4H2O, AR, Macklin, Shanghai, China), Sodium tungstate dihydrate (Na2WO4·2H2O, Xi long hua gong, China), Urea (CH4N2O, Aladdin, China), Dopamine hydrochloride (PDA) (C8H11NO2·HCl, Aladdin, Shanghai, China), and Tris(hydroxymethyl)aminomethane (C4H11NO3, Aladdin, China) were used as the raw materials. All the chemicals and solvents were used as received without any further purification.

2.1.2. Synthesis of MnWO4

As a typical preparing process, 0.3 mmol of MnCl4·4H2O and 0.3 mmol Na2WO4·2H2O were dissolved in 30 mL of deionized water (DIW) to form a solution A. After stirring for 30 min, 9 mmol of urea was added into solution A and stirred for another 30 min. Then, the mixture was transferred into a 50 mL polytetrafluoroethylene reactor, which was sealed and heated at 200 °C for 12 h. After cooling to the room temperature, the brown precipitate was collected and washed several times with DIW and then dried in a vacuum oven at 60 °C for 24 h. MnWO4 was obtained by annealing the intermediates under an air atmosphere at 450 °C for 12 h with a ramping rate of 2 °C min−1.

2.1.3. Synthesis of MnWO4@C

We completely dispersed 100 mg of MnWO4 into 200 mL of Tris-buffer solution. Then, we added 60 mg of PDA to the above mixture and stirred for 12 h. The obtained product was cleaned and centrifuged multiple times using DIW, then dried at 60 °C, and finally baked in a vacuum furnace for 12 h (Figure 1).

2.1.4. Material Characterization

A Bruker D8 Focus power type X-ray diffractometer, Cu Kα rays with a wavelength of 0.154056 nm, was used as the radiation source, and X-ray diffraction analysis was performed on the prepared samples at a scanning rate of 5°/min within the 2θ angle range of 10–80° to study their crystalline phases. Detailed observations and analyses of the appearance and microstructure of the synthesized product were conducted using a field emission scanning electron microscope (FE-SEM, model Hitachi S-4800, Hitachi, Tokyo, Japan) under a set acceleration voltage of 10 kV. A transmission electron microscope (TEM, model Tecnai G2, Hillsboro, OR, USA) was used to conduct an in-depth microstructure analysis of the sample under an electric field emission gun with an operating voltage of 200 kV. Then, we evaluated the surface chemical state and elemental composition of Al Kα source samples under X-ray photoelectron spectroscopy (XPS) analysis technology. In addition, the adsorption/desorption isotherm data of nitrogen were obtained using an ASAP 2010 trace adsorption analyzer under a temperature condition of −196 °C, combined with a porosity analysis function. The specific surface area of the sample was accurately calculated using the Brunauer–Emmett–Teller (BET) method.

2.1.5. Electrochemical Measurements

In the glove box, button type batteries were assembled, and the performance of the experimental batteries was tested. In order to prepare the working electrode, the prepared sample, acetylene black, and carboxymethyl cellulose (CMC) were mixed in a mass ratio of 80:10:10 and dissolved in an appropriate amount of deionized water. Sufficient stirring was used to ensure the formation of a uniform and consistent slurry. Then, we evenly coated the prepared slurry on the surface of the copper foil and then placed the copper foil coated with the slurry in a vacuum oven at 60 °C for 12 h of drying treatment. The average mass load of the active substance layer on the copper foil was controlled at 1.0–1.3 mg cm−2. The battery uses lithium foil as the counter electrode/reference electrode and a Celgard 2400 membrane as the separator, and the electrolyte was dissolved in a mixed solvent of ethylene carbonate (EC) and diethyl carbonate (DEC) (volume ratio 1:2) to form a complete half-cell system. For the anode testing for MnWO4@C, we added a mass percentage of 2% (wt) fluoroethylene carbonate (FEC) to the electrolyte to observe its impact on the testing process and results. We conducted a detailed evaluation and testing of the electrochemical performance of the battery samples using a programmable battery tester (LAND CT2001A) at room temperature. Cyclic voltammogram (CV) and electrochemical impedance spectroscopy (EIS) measurements were carried out on a Bio-Logic VMP3 electrochemical workstation.

3. Result and Discussion

3.1. The Characterization of MnWO4 and MnWO4@C

Figure 2a gives the morphology of MnWO4, which shows the shape of the nanorod with a diameter of about 20 nm, and the lengths of most nanorods are less than 200 nm. The TEM image in Figure 2b shows the hierarchically distributed nanorod-like structure more clearly. Figure 2c,d shows that there are plenty of micropores on the surface of nanorods, which can be used as effective channels for electrolyte transformation. The HRTEM image of Figure 2e clearly shows the lattice fringe (002) of the crystal plane of MnWO4, whose lattice spacing is 2.497 Å. The elements Mn, W, and O are evenly distributed, as can be seen from the EDX images in Figure 2f–j [23,24,25,26,27,28].
Figure 3 is TEM and EDX images of MnWO4@C. MnWO4@C also has the nanorod structure similar to MnWO4, as seen in Figure 3a. The nanorods are coated with a ~5.1 nm-thick carbon layer, which is derived from dopamine, as Figure 3b shows. The EDX analysis indicates that the elements Mn, W, O, and N are uniformly distributed in MnWO4@C nanorods. The element N comes entirely from the carbon layer-derived dopamine. Therefore, the range of the element N corresponds to the distribution of the carbon layer. The fuzzy distribution of the element C is caused by the carbon-supporting film in the testing process.
The composition of obtained materials was determined by Powder X-ray diffractometer (XRD), as shown in Figure 4a. From the XRD pattern, it can be seen that the patterns of MnWO4 and MnWO4@C are very similar. And both of them are corresponding to the Standards Card (PDF#13-0434, a = 4.824 Å, b = 5.75 Å, c = 4.99 Å) of MnWO4 very well. The absence of extraneous peaks confirms the high purity of the obtained sample [29,30,31]. Moreover, compared with the pure MnWO4, the inconspicuous protrusions around 21° proves a successful introduction of amorphous carbon into MnWO4@C. X-ray photoelectron spectroscopy (XPS) was used to analyze the chemical valence of elements in the sample. The peaks at 641.1 and 652.0 eV corresponded to the binding energy of Mn 2P3/2 and 2P1/2, respectively (Figure 4b). This shows that Mn is in the +2 oxidation state. The peaks at 33.6 and 36.2 eV are in accordance with the W 4f7/2 and 4f5/2, which indicates W is in the +6 oxidation state (Figure 4c). In the C 1s spectrum (Figure S1b) [32], the signals at 284.5 eV, 285.4, and 288.5 eV correspond to the C=C, C-N (or C-C), and C=O configurations, respectively [25,33,34,35]. N2 adsorption/desorption measurements was used to determine the specific surface area and porous characteristics of MnWO4@C (Figure 4d,e: the latter is nested within Figure 4d). The specific surface area of MnWO4@C composite powder is 60.77 m2 g−1 from the type-III isotherm with a hysteresis loop. The average pore volume and average pore size are 0.18 cm3 g−1, respectively. The specific surface area of MnWO4 composite powder is 35.59 m2 g−1 from the type-III isotherm with a hysteresis loop (Figure S2) [36]. The increase in specific surface area is caused by the carbon shells derived from dopamine.

3.2. The Behavior of MnWO4 and MnWO4@C on Li+ Storage

In order to evaluate the lithium storage capacity of MnWO4 and MnWO4@C, a lithium metal anode was employed as a symmetrical electrode to assemble a button battery. The MnWO4@C electrodes deliver an excellent capacity of 795 mAh g−1 at 200 mA g−1 after initial 100 cycles. This value is much higher than 340 mAh g−1 of bare MnWO4, which showed a spanning decay and 60% capacity retention after only 100 cycles, as shown in Figure 5a. This outstanding performance is also seen in rate performances (Figure 5b). MnWO4@C electrodes show superior capacity retention at various current densities from 100 to 5000 mA g−1. The average capacities of 875, 815, 687, 587, 447, and 287 mAh g−1 are obtained for MnWO4@C at 100, 200 500, 1000, 2000, and 5000 mA g−1, respectively; by contrast, 697, 408, 195, 102, 43, and 11 mAh g−1 were obtained for MnWO4. Even at high current densities of 2000 and 5000 mA g−1, its capacity retention can still be maintained at 51% and 32.8%. As the current densities return from 5000 to 100 mA g−1, the specific capacity comes back to 287, 500, 640, 735, 827, and 915 mAh g−1, indicating a good stability of MnWO4@C. Figure 5c shows the corresponding galvanostatic discharge–charge curve at different current densities of MnWO4@C. The cyclic voltammetry also certificates the stable cycling performance of MnWO4@C in a half-cell of LIBs at a scan rate of 0.1 mV s−1 in the voltage range from 0.01 to 3.0 V, as shown in Figure 5d. There are obviously differences between the first discharge curve and the subsequent cycle. In the first cycle, the peak at 1.2 V originates from the reduction of Mn2+ to Mn, and the reduced peak locked at 0.35 V corresponds to W6+ to W, as shown in Equation (1). In the subsequent anodic scan, there are two distinct peaks at 1.35 and 2.0 V, which corresponded to the transformation of Mn and W into MnO and WO3, respectively [37,38], as shown in Equation (2). The reaction process is as follows [39,40]:
Discharge:
MnWO4 + 8Li+ +8e → Mn + W + 4Li2O
Charge:
Mn + W + 4Li2O → MnO + WO3 +8Li+ + 8e
In the subsequent cycles, a pair of reduction peaks at 1.3 and 0.5 V in the cathodic scan and a pair of oxidation peaks at 1.3 and 2.0 V in the anodic scan were found. In subsequent loops, the oxidation peak at 2.0 V (peak I) gradually disappeared with the increase in the number of cycles, which proved the irreversibility of the W to WO3 response. In addition, the peak locked at 1.3 V does not change significantly in subsequent cycles, which proves that the electrochemical reaction is highly reversible. In the cathodic scan, the intensity of the peak II decreases gradually with the number of cycles increasing, which proved that the reversibility of the conversion reaction declined. This phenomenon is consistent with the change trend of the oxidation peak I. However, the intensity of the peak III becomes more and more obvious as the number of cycles increases, which proves that the reversibility of the reaction is getting much higher. This phenomenon corresponds to the change trend of the oxidation peak of 1.3 V in the anodic scan. When the current density is 1000 mA g−1, the capacity is maintained at 600 mAh g−1 after 200 cycles, which further proves the excellent electrochemical performance of MnWO4@C.

3.3. Kinetics Comparison of MnWO4 and MnWO4@C

The synthesis of N-doped carbon results in defects and pores being uniformly distributed throughout the material, creating a vast array of active sites and a homogeneous mesoporous structure, which enhances the electron conductivity and lithium-ion diffusion rate. Compared to MnWO4, MnWO4@C exhibits superior kinetic and exemplary electrochemical properties. Figure S3 shows the Nyquist plots of MnWO4 and MnWO4@C [29]. The initial impedance semicircle observed in the high-frequency region predominantly represents the charge transfer resistance (Rct) at the electrode/electrolyte boundary. Subsequently, the second capacitive impedance arc emerges in the low-frequency region, attributable to the capacitive reactance (Rc) of the adsorption layer generated on the electrode surface. The Rct value of MnWO4@C is 48.8 Ω, which is much lower than 64.7 Ω of MnWO4. In addition, the capacitive reactance of MnWO4@C (Rc = 16.7 Ω) is also lower than that of bare MnWO4 (31.1 Ω), which implies MnWO4@C has a higher lithium-ion diffusion rate and better electrochemical performance. This proves that N-doped carbon not only relieves the stress of the lithium insertion process to a certain extent, maintaining the stability of the material structure but also forms a more stable SEI. The galvanostatic intermittent titration technique (GITT) test is also applied, which shows that MnWO4@C has better dynamic performance than MnWO4, as shown in Figure 6 and Figure S4 [32] and Table S1 (tested at 0.1 A g−1, pulse time of 20 min, and relaxation time of 30 min). This is primarily attributable to the enhanced electrical conductivity of the active material, which is enabled by the N-doped carbon shells on the surface of MnWO4@C and effectively mitigates the volume effect during the cycling process, thereby facilitating the formation of a stable SEI film. Furthermore, surface imperfections (such as holes and cracks) can facilitate the formation of transport channels, which are conducive to the diffusion and transport of lithium ions. This, in turn, enhances the electrochemical cycling performance of the composites. The equation as follows [41]:
D = 4 L 2 π   τ ( Δ E s Δ E t ) 2
in which L is the distance of Li+ diffusion, (equal to the electrode thickness, cm), τ stands for the relaxation time (s), ΔEs is the potential change via the current pulse, and ΔEt is the voltage change caused by constant current charging (discharging). The values of each parameter are shown in Figure 6b. According to the above equation, the lithium-ion diffusion rates under different delithiation and lithiation states are calculated, as shown in Tables S1 and S2. In addition, the cycle performance is comparable to previously reported results regarding Mn-based materials for LIBs (Table S3).
As shown in Figure 7a, in order to delve deeper into the details of the charge storage process, we conducted cyclic voltammetry (CV) measurements and analyzed the correlation between the peak current (i) and the scanning rate (v) based on the power law i = avb. This analysis aims to reveal how the charge storage mechanism changes at different scanning rates, thereby providing a deeper understanding of the performance of a battery or electrode materials. Particularly, when the value of b is equal to 0.5, it indicates a semi-infinite linear diffusion control process, which means that the charge transfer is mainly limited by the diffusion rate of ions inside the electrode material; when the value of b reaches 1, it indicates that the process is surface controlled or has capacitive characteristics—that is, the storage and release of charges mainly occur on the surface or near the surface area of the electrode material, and the speed is very fast, almost not limited by diffusion. By plotting log i versus log v, the b values of the three redox peaks (peak 1, peak 2, and peak 3) are determined to be 0.74, 0.7, 1, and 0.79 (Figure 7b), indicating a dominant capacitive behavior of MnWO4@C during the electrochemical reaction. The equation i = k1v + k2v1/2 can be used to accurately quantify the proportion of capacitive contribution (k1v) and diffusive contribution (k2v1/2) in the total current. To determine the constant values of k1 and k2 at a specific potential, a simulation and fitting can be performed by plotting the relationship between iv−1/2 and v1/2. This method helps to separate and quantify these two different charge transfer mechanisms. The Figure 7c,d summarizes detailed contribution of capacitive behavior at various scan rates. The contributions are 62.98%, 74.67%, 79.12%, 81.82%, 83.62%, 84.33%, 86.83%, and 90.31% at the scan rates of 0.1, 0.3, 0.5, 0.7, 0.9, 1.0, 1.5, and 3.0 mV s−1, respectively. The experimental results show that, among the total capacity of MnWO4@C, capacitive charge storage accounts for a significant proportion. The enhanced charge transfer performance and fast capacitive charge storage capacity of the carbon shell in MnWO4@C are mainly attributed to its unique characteristics, including the rich amorphous structure, porous structure, and large specific surface area. In contrast, the contributions of MnWO4 are 29.6%, 42.1%, 48.4%, 52.6%, 55.8%, 57.1%, 62.0%, and 69.7% at the scan rates of 0.1, 0.3, 0.5, 0.7, 0.9, 1.0, 1.5, and 3.0 mV s−1, respectively, as Figure S5 shows [42].

3.4. The Influence of an Electrolyte Additive

Fluoroethylene carbonate (FEC) is employed as a film-forming additive in electrolytes, whereby it forms stable complexes with lithium ions. This complexation constrains the migration of lithium ions, which in turn inhibits the occurrence of the reduction reaction of the boundary electrolyte, thus improving the cycle life of the battery. Concurrently, the reaction between FEC and lithium ions at the interface generates a dense film, which restricts the diffusion of solvent molecules and lithium ions in the electrolyte to the electrode surface, thereby reducing the self-discharge phenomenon of the battery [43,44,45,46]. Therefore, 2 wt% FEC was added in the electrolyte as an additive on the MnWO4@C electrode. The outstanding performance is represented in rate performances (Figure 8a). MnWO4@C-FEC electrodes shows high-capacity retention at various current densities from 100 to 5000 mA g−1. The average capacities of 878, 831, 758, 704, 611, and 458 mAh g−1 are obtained for MnWO4@-FEC at the current densities of 100, 200 500, 1000, 2000, and 5000 mA g−1, respectively, which are higher than MnWO4@C. As the current densities return from 5000 to 100 mA g−1, the specific capacity comes back to 678, 798, 824, 938, and 1017 mAh g−1, indicating more stability of MnWO4@C-FEC. Figure 8b shows the corresponding galvanostatic discharge–charge curve at different current densities of MnWO4@C-FEC, which is similar to the MnWO4@C anode. The capacity is maintained at 718 mAh g−1 at 1000 mA g−1 after 400 cycles, which further proves the superior electrochemical performance after adding FEC. However, the capacity of the MnWO4@C anode began to decrease after 200 cycles.

4. Conclusions

Through an efficient hydrothermal synthesis method, we have successfully prepared a compact nanocavity structure MnWO4@C nanorods. The formation mechanism of the nanocavity structure inside these nanorods is similar to the process of inside-out Ostwald ripening. The prepared MnWO4@C nanorods with dense nanocavities have been a substantial improvement in rate performance and specific capacity in LIB compared with MnWO4. On the one hand, the good electrochemical performance is mainly attributed to its unique nanocavity geometry and optimized conversion mechanism during the cycling process. These two factors jointly promote efficient electrode–electrolyte interface contact, establish a short-range electron transfer path, and provide sufficient lithium-ion storage sites during the insertion and extraction process of lithium ions. On the other hand, the standout electrochemical performance could be attributed to the carbon shell derived from dopamine. Moreover, both the transmission rate of lithium ions and the contribution of pseudo-capacitance increased because of the electroconductive and poriferous carbon shell. And the negative effects of volume expansion are also attenuated.
The carbon coating layer greatly improves the electrochemical properties of MnWO4 nanorods. However, with the cycle number increasing, the volume expansion of the electrode becomes more and more serious, and then the structure of the material and electrode are totally destroyed. Therefore, the electrolyte additive can form a stable SEI film and effectively restrain the capacity loss caused by the damage to the electrode structure in the long cycle process.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17194682/s1, Figure S1: XPS spectra of MnWO4@C, (a) O 1s, (c) C 1s; Figure S2: N2 adsorption and desorption isotherms of MnWO4; Figure S3: Nyquist plot of MnWO4 and MnWo4@C; Figure S4: GITT curves of the MnWO4 electrode (discharge/charge state). (b) E vs t profile for one GITT test. (c) DLi+ of MnWO4 during the charge and discharge processes; Figure S5: Kinetic analysis of MnWO4 as LIB anodes. (a) CV curves of MnWO4 at different scan rates, (b) log i versus log v plots at each redox peak, contribution ratio of the capacitive and diffusion-controlled charge at (c) 0.5 mV s−1 and (d) different scan rates; Table S1: Lithiation states; Table S2: Delithiation states; Table S3: Comparison of electrochemical performance in LIBs with previous works.

Author Contributions

Writing-Original, D.W.; Draft; Supervision, Z.W. and Y.L.; Visualisation, C.W.; Data Curation, D.Y.; Conceptualisation, L.W.; Funding Acquisition, Y.C. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 52171210), the Program for the Development of Science and Technology of Jilin Province (Item No. 20220201130GX), and Suppored by Key Laboratory of Functional Materials Physics and Chemistry (Jilin Normal University), Ministry of Education, China (No.202207).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article [and its Supplementary Materials].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Armand, M.; Tarascon, J.-M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef] [PubMed]
  2. Dunn, B.; Kamath, H.; Tarascon, J.M. Electrical energy storage for the grid: A battery of choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef]
  3. Liu, C.; Qiu, Y.; Liu, Y. Novel 3D grid porous Li4Ti5O12 thick electrodes fabricated by 3D printing for high performance lithium-ion batteries. J. Adv. Ceram. 2022, 11, 295–307. [Google Scholar] [CrossRef]
  4. Chen, C.; Lee, C.S.; Tang, Y.B. Fundamental Understanding and Optimization Strategies for Dual-Ion Batteries: A Review. Nano-Micro Lett. 2023, 15, 121. [Google Scholar] [CrossRef]
  5. Hou, J.B.; Zhang, K.; Xiao, J.H.; Xu, Z.Q.; Gao, W.J.; Gao, X.Y.; Zhou, S.K.; Jiao, Z.Z.; Yi, M.R.; Yin, Y.H.; et al. Modified tungsten oxide as a binder-free anode in lithium-ion battery for improving electrochemical stability. Tungsten 2022, 4, 356–369. [Google Scholar] [CrossRef]
  6. Montemayor, S.M.; Fuentes, A.F. Electrochemical characteristics of lithium insertion in several 3D metal tungstates (MWO4, M = Mn, Co, Ni and Cu) prepared by aqueous reactions. Ceram. Int. 2004, 30, 393–400. [Google Scholar] [CrossRef]
  7. Tiwari, A.; Singh, V.; Nagaiah, T.C. Tuning the MnWO4 morphology and its electrocatalytic activity towards oxygen reduction reaction. J. Mater. Chem. A 2018, 6, 2681–2692. [Google Scholar] [CrossRef]
  8. Amina, M.; Amna, T.; Hassan, M.S.; Al Musayeib, N.M.; Al-Deyab, S.S.S.; Khil, M.-S. Low temperature synthesis of Manganese tungstate nanoflowers with antibacterial potential: Future material for water purification. Korean J. Chem. Eng. 2016, 33, 3169–3174. [Google Scholar] [CrossRef]
  9. Jiao, L.; Sun, G.R.; Wang, Y.; Zhang, Z.X.; Wang, Z.; Wang, H.R.; Li, H.B.; Feng, M. Defective TiO2 hollow nanospheres as photo-electrocatalysts for photo-assisted Li-O2 batteries. Chin. Chem. Lett. 2022, 33, 4008–4012. [Google Scholar] [CrossRef]
  10. Muthamizh, S.; Suresh, R.; Giribabu, K.; Manigandan, R.; Kumar, S.P.; Munusamy, S.; Narayanan, V. MnWO4 nanocapsules: Synthesis, characterization and its electrochemical sensing property. J. Alloys Compd. 2015, 619, 601–609. [Google Scholar] [CrossRef]
  11. Liu, H.; Di, J.; Wang, P.; Gao, R.; Tian, H.; Ren, P.F.; Yuan, Q.X.; Huang, W.X.; Liu, R.P.; Liu, Q.; et al. A novel design of 3D carbon host for stable lithium metal anode. Carbon Energy 2022, 4, 654–664. [Google Scholar] [CrossRef]
  12. Raj, B.G.S.; Acharya, J.; Seo, M.-K.; Khil, M.-S.; Kim, H.-Y.; Kim, B.-S. One-pot sonochemical synthesis of hierarchical MnWO4 microflowers as effective electrodes in neutral electrolyte for high performance asymmetric supercapacitors. Int. J. Hydrog. Energy 2019, 44, 10838–10851. [Google Scholar] [CrossRef]
  13. Rani, B.J.; Ravi, G.; Ravichandran, S.; Ganesh, V.; Ameen, F.; Al-Sabri, A.; Yuvakkumar, R. Electrochemically active XWO4 (X = Co, Cu, Mn, Zn) nanostructure for water splitting applications. Appl. Nanosci. 2018, 8, 1241–1258. [Google Scholar] [CrossRef]
  14. Xiang, H.; Xiang, Y.; Dai, F.Z. High-entropy ceramics: Present status, challenges, and a look forward. J. Adv. Ceram. 2021, 10, 385–441. [Google Scholar] [CrossRef]
  15. Li, Y.H.; Zhang, L.; Yen, H.Y.; Zhou, Y.C.; Jang, G.; Yuan, S.L.; Wang, J.H.; Xiong, P.X.; Liu, M.L.; Park, H.S.; et al. Single-Phase Ternary Compounds with a Disordered Lattice and Liquid Metal Phase for High-Performance Li-Ion Battery Anodes. Nano-Micro Lett. 2023, 15, 63. [Google Scholar] [CrossRef]
  16. Dai, Y.M.; Chen, Q.J.; Hu, C.C.; Huang, Y.Y.; Wu, W.Y.; Yu, M.L.; Sun, D.; Luo, W. Copper fluoride as a low-cost sodium-ion battery cathode with high capacity. Chin. Chem. Lett. 2022, 33, 1435–1438. [Google Scholar] [CrossRef]
  17. Park, G.D.; Lee, J.-K.; Kang, Y.C. Synthesis of Uniquely Structured SnO2 Hollow Nanoplates and Their Electrochemical Properties for Li-Ion Storage. Adv. Funct. Mater. 2017, 27, 1603399–1603408. [Google Scholar] [CrossRef]
  18. Kong, D.; Wang, Y.; Liu, B.; Huang, Z.; Yang, H.Y. Seed-assisted growth of α-Fe2O3 nanorod arrays on reduced graphene oxide: A superior anode for high-performance Li-ion and Na-ion batteries. J. Mater. Chem. A 2016, 4, 11800–11811. [Google Scholar] [CrossRef]
  19. Zhang, W.; Zhang, J.; He, T.; Hu, L.; Wang, R.; Mai, L.; Mu, S. Hierarchical three-dimensional MnO nanorods/carbon anodes for ultralong-life lithium-ion batteries. J. Mater. Chem. A 2016, 4, 16936–16945. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Gao, X.; Wang, B.; Liu, H.; Wu, H.; Liu, H.; Dou, S. Nitrogen-Doped Graphene Ribbon Assembled Core–Sheath MnO@Graphene Scrolls as Hierarchically Ordered 3D Porous Electrodes for Fast and Durable Lithium Storage. Adv. Funct. Mater. 2016, 26, 7754–7765. [Google Scholar] [CrossRef]
  21. Wang, D.; Zhao, Z.Y.; Wang, P.; Wang, S.M.; Feng, M. Synthesis of MOF-derived nitrogen-doped carbon microtubules via template self-consumption. Rare Met. 2022, 41, 2582–2587. [Google Scholar] [CrossRef]
  22. Zhang, K.; Li, P.; Ma, M.; Park, J.H. Core-Shelled Low-Oxidation State Oxides@Reduced Graphene Oxides Cubes via Pressurized Reduction for Highly Stable Lithium Ion Storage. Adv. Funct. Mater. 2016, 26, 2959–2965. [Google Scholar] [CrossRef]
  23. Zhou, S.; Huang, J.; Zhang, T.; Ouyang, H.; Li, A.; Zhang, Z. Effect of variation Mn/W molar ratios on phase composition, morphology and optical properties of MnWO4. Ceram. Int. 2013, 39, 5159–5163. [Google Scholar] [CrossRef]
  24. Wang, Y.; Yang, L.; Wang, Y.; Wang, X.; Han, G. Morphology-controlled synthesis and characterization of MnWO4 nanocrystals via a facile, additive-free hydrothermal process. J. Alloys Compd. 2016, 654, 246–250. [Google Scholar] [CrossRef]
  25. Wang, W.; Wu, N.; Zhou, J.M.; Li, F.; Wei, Y.; Li, T.H.; Wu, X.L. MnWO4 nanoparticles as advanced anodes for lithium-ion batteries: F-doped enhanced lithiation/delithiation reversibility and Li-storage properties. Nanoscale 2018, 10, 6832–6836. [Google Scholar] [CrossRef]
  26. Yu, S.H.; Liu, B.; Mo, M.S.; Huang, J.H.; Liu, X.M.; Qian, Y.T. General Synthesis of Single-Crystal Tungstate Nanorods/Nanowires: A Facile, Low-Temperature Solution Approach. Adv. Funct. Mater. 2003, 13, 639–647. [Google Scholar] [CrossRef]
  27. Chen, S.-J.; Chen, X.-T.; Xue, Z.; Zhou, J.-H.; Li, J.; Hong, J.-M.; You, X.-Z. Morphology control of MnWO4 nanocrystals by a solvothermal route. J. Mater. Chem. 2003, 13, 1132–1135. [Google Scholar] [CrossRef]
  28. Yang, L.; Wang, Y.; Wang, Y.; Wang, X.; Wang, L.; Han, G. Shape-controlled synthesis of MnWO4 nanocrystals via a simple hydrothermal method. J. Alloys Compd. 2013, 578, 215–219. [Google Scholar] [CrossRef]
  29. Gao, G.; Dang, W.; Wu, H.; Zhang, G.; Feng, C. Synthesis of MnWO4@C as novel anode material for lithium ion battery. J. Mater. Sci. Mater. Electron. 2018, 29, 12804–12812. [Google Scholar] [CrossRef]
  30. Ramesan, M.T.; Abdu Raheem, V.P.; Jayakrishnan, P.; Pradyumnan, P.P. Acrylonitrile butadiene rubber (NBR)/manganous tungstate (MnWO4) nanocomposites: Characterization, mechanical and electrical properties. AIP Conf. Proc. 2014, 1620, 3–9. [Google Scholar]
  31. En Zhang, Zheng Xing, Ji Wang, Zhicheng Ju, Yitai Qian, Enhanced energy storage and rate performance induced by dense nanocavities inside MnWO4 nanobars. RSC Adv. 2012, 2, 6748–6751. [CrossRef]
  32. Peng, J.M.; Chen, Z.Q.; Li, Y.; Hu, S.J.; Pan, Q.C.; Zheng, F.H.; Wang, H.Q.; Li, Q.Y. Conducting network interface modulated rate performance in LiFePO4/C cathode materials. Rare Met. 2022, 41, 951–959. [Google Scholar] [CrossRef]
  33. Wang, Y.T.; Chen, X. Single crystal growth and electrochemical studies of garnet-type fast Li-ion conductors. Tungsten 2022, 4, 263–268. [Google Scholar] [CrossRef]
  34. Zhang, B.; Wang, L.; Zhang, H.; Xu, H.; He, X. Revelation of the transition-metal doping mechanism in lithium manganese phosphate for high performance of lithium-ion batteries. Battery Energy 2022, 1, 20220020. [Google Scholar] [CrossRef]
  35. Yao, S.; Xing, L.; Dong, Y.; Wu, X. Hierarchical WO3@MnWO4 core-shell structure for asymmetric supercapacitor with ultrahigh cycling performance at low temperature. J. Colloid Interface Sci. 2018, 531, 216–224. [Google Scholar] [CrossRef]
  36. Zhang, X.; Jiang, Y.; Liu, B.; Yang, W.; Li, J.; Niu, P.; Jiang, X. High-performance phototransistor based on individual high electron mobility MnWO4 nanoplate. J. Alloys Compd. 2018, 762, 933–940. [Google Scholar] [CrossRef]
  37. Petnikota, S.; Srikanth, V.V.S.S.; Nithyadharseni, P.; Reddy, M.V.; Adams, S.; Chowdari, B.V.R. Sustainable Graphenothermal Reduction Chemistry to Obtain MnO Nanonetwork Supported Exfoliated Graphene Oxide Composite and its Electrochemical Characteristics. ACS Sustain. Chem. Eng. 2015, 3, 3205–3213. [Google Scholar] [CrossRef]
  38. Choi, M.J.; Baek, J.H.; Kim, J.Y.; Jang, H.W. Highly textured and crystalline materials for rechargeable Li-ion batteries. Battery Energy 2023, 2, 20230010. [Google Scholar] [CrossRef]
  39. Gu, X.X.; Yang, Z.G.; Qiao, S.; Shao, C.B.; Ren, X.L.; Yang, J.J. Exploiting methylated amino resin as a multifunctional binder for high-performance lithium-sulfur batteries. Rare Met. 2021, 40, 529–536. [Google Scholar] [CrossRef]
  40. Hyun-Woo Shim, Ah-Hyeon Lim, Jae-Chan Kim, Gwang-Hee Lee, Dong-Wan Kim, Hydrothermal realization of a hierarchical, flowerlike MnWO4@MWCNTs nanocomposite with enhanced reversible Li storage as a new anode material. Chem. Asian J. 2013, 8, 2851–2858. [CrossRef]
  41. Wei, J.; Ma, J.; Wang, W.; Li, T.; Wu, N.; Zhang, D. Study of the effect of F-doping on lithium electrochemical behavior in MnWO4 anode nanomaterials. J. Inorg. Organomet. Polym. Mater. 2021, 31, 3175–3182. [Google Scholar] [CrossRef]
  42. Cheng, Y.; Wang, S.; Zhou, L.; Chang, L.; Liu, W.; Yin, D.; Yi, Z.; Wang, L. SnO2 quantum dots: Rational design to achieve highly reversible conversion reaction and stable capacities for lithium and sodium storage. Small 2020, 16, 2000681. [Google Scholar] [CrossRef] [PubMed]
  43. Choi, N.-S.; Yew, K.H.; Lee, K.Y.; Sung, M.; Kim, H.; Kim, S.-S. Effect of fluoroethylene carbonate additive on interfacial properties of silicon thin-film electrode. J. Power Sources 2006, 161, 1254–1259. [Google Scholar] [CrossRef]
  44. Komaba, S.; Ishikawa, T.; Yabuuchi, N.; Murata, W.; Ito, A.; Ohsawa, Y. Fluorinated Ethylene Carbonate as Electrolyte Additive for Rechargeable Na Batteries. ACS Appl. Mater. Interfaces 2011, 3, 4165–4168. [Google Scholar] [CrossRef]
  45. Song, S.P.; Yang, C.; Jiang, C.Z.; Wu, Y.M.; Guo, R.; Sun, H.; Yang, J.L.; Xiang, Y.; Zhang, X.K. Increasing ionic conductivity in Li0.33La0.56TiO3 thin-films via optimization of processing atmosphere and temperature. Rare Met. 2022, 41, 179–188. [Google Scholar] [CrossRef]
  46. Zhang, S.S. A review on electrolyte additives for lithium-ion batteries. J. Power Sources 2006, 162, 1379–1394. [Google Scholar] [CrossRef]
Figure 1. Scheme of the preparation process.
Figure 1. Scheme of the preparation process.
Materials 17 04682 g001
Figure 2. (a) SEM of MnWO4; (be) TEM and HRTEM of MnWO4; (fj) EDX of MnWO4.
Figure 2. (a) SEM of MnWO4; (be) TEM and HRTEM of MnWO4; (fj) EDX of MnWO4.
Materials 17 04682 g002
Figure 3. (a) TEM of MnWO4@C; (b) HRTEM of MnWO4@C; (ci) EDX of MnWO4@C.
Figure 3. (a) TEM of MnWO4@C; (b) HRTEM of MnWO4@C; (ci) EDX of MnWO4@C.
Materials 17 04682 g003
Figure 4. Physical characterization of MnWO4@C. (a) XRD patterns of MnWO4 and MnWO4@C. XPS spectra of the (b) Mn2p, (c) W4f, (d) N2 adsorption and desorption isotherms, and (e) pore size distribution of MnWO4@C.
Figure 4. Physical characterization of MnWO4@C. (a) XRD patterns of MnWO4 and MnWO4@C. XPS spectra of the (b) Mn2p, (c) W4f, (d) N2 adsorption and desorption isotherms, and (e) pore size distribution of MnWO4@C.
Materials 17 04682 g004
Figure 5. Electrochemical characteristics of MnWO4 and MnWO4@C. (a) Comparative cycle performances of MnWO4 and MnWO4@C at a current density of 200 mA g−1. (b) Rate capacities of MnWO4 and MnWO4@C at current densities of 100, 200, 500, 1000, 2000, and 5000 mA g−1. (c) galvanostatic discharge–charge curves of MnWO4@C electrodes at 0.1–5 A g−1. (d) CV curves of MnWO4@C electrodes at a scan rate of 0.1 mV s−1. (e) the galvanostatic discharge–charge cycling performances and the coulombic efficiency of MnWO4 and MnWO4@C electrodes at a current density of 1000 mA g−1.
Figure 5. Electrochemical characteristics of MnWO4 and MnWO4@C. (a) Comparative cycle performances of MnWO4 and MnWO4@C at a current density of 200 mA g−1. (b) Rate capacities of MnWO4 and MnWO4@C at current densities of 100, 200, 500, 1000, 2000, and 5000 mA g−1. (c) galvanostatic discharge–charge curves of MnWO4@C electrodes at 0.1–5 A g−1. (d) CV curves of MnWO4@C electrodes at a scan rate of 0.1 mV s−1. (e) the galvanostatic discharge–charge cycling performances and the coulombic efficiency of MnWO4 and MnWO4@C electrodes at a current density of 1000 mA g−1.
Materials 17 04682 g005
Figure 6. (a) GITT curves of the MnWO4@C electrode (discharge/charge state). (b) E vs. t profile for one GITT test. (c) DLi+ of MnWO4@C during the charge and discharge processes.
Figure 6. (a) GITT curves of the MnWO4@C electrode (discharge/charge state). (b) E vs. t profile for one GITT test. (c) DLi+ of MnWO4@C during the charge and discharge processes.
Materials 17 04682 g006
Figure 7. Kinetic analysis of MnWO4@C as LIB anodes. (a) CV curves of MnWO4@C at different scan rates, (b) log i versus log v plots at each redox peak, the contribution ratio of the capacitive and diffusion-controlled charge at (c) 0.5 mV s−1, and (d) different scan rates.
Figure 7. Kinetic analysis of MnWO4@C as LIB anodes. (a) CV curves of MnWO4@C at different scan rates, (b) log i versus log v plots at each redox peak, the contribution ratio of the capacitive and diffusion-controlled charge at (c) 0.5 mV s−1, and (d) different scan rates.
Materials 17 04682 g007
Figure 8. Electrochemical characteristics of MnWO4@C and MnWO4@C-FEC. (a) Comparative rate capacity of MnWO4@C and MnWO4@C-FEC at current densities of 100, 200, 500, 1000, 2000, and 5000 mA g−1. (b) galvanostatic discharge–charge curves of MnWO4@C-FEC electrodes at 0.1–5 A g−1. (c) the long cycle performance and coulombic efficiency of MnWO4@C and MnWO4@C-FEC electrodes at a current density of 1000 mA g−1.
Figure 8. Electrochemical characteristics of MnWO4@C and MnWO4@C-FEC. (a) Comparative rate capacity of MnWO4@C and MnWO4@C-FEC at current densities of 100, 200, 500, 1000, 2000, and 5000 mA g−1. (b) galvanostatic discharge–charge curves of MnWO4@C-FEC electrodes at 0.1–5 A g−1. (c) the long cycle performance and coulombic efficiency of MnWO4@C and MnWO4@C-FEC electrodes at a current density of 1000 mA g−1.
Materials 17 04682 g008
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

Wang, D.; Wang, Z.; Wang, C.; Yin, D.; Liang, Y.; Wang, L.; Cheng, Y.; Feng, M. Synergistically Boosting Li Storage Performance of MnWO4 Nanorods Anode via Carbon Coating and Additives. Materials 2024, 17, 4682. https://doi.org/10.3390/ma17194682

AMA Style

Wang D, Wang Z, Wang C, Yin D, Liang Y, Wang L, Cheng Y, Feng M. Synergistically Boosting Li Storage Performance of MnWO4 Nanorods Anode via Carbon Coating and Additives. Materials. 2024; 17(19):4682. https://doi.org/10.3390/ma17194682

Chicago/Turabian Style

Wang, Duo, Zhaomin Wang, Chunli Wang, Dongming Yin, Yao Liang, Limin Wang, Yong Cheng, and Ming Feng. 2024. "Synergistically Boosting Li Storage Performance of MnWO4 Nanorods Anode via Carbon Coating and Additives" Materials 17, no. 19: 4682. https://doi.org/10.3390/ma17194682

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