Jinshan Wang, Feng Li, Si Zhao, Lituo Zheng, Yiyin Huang, and Zhensheng Hong, Uniform nanoplating of metallic magnesium film on titanium dioxide nanotubes as a skeleton for reversible Na metal anode, Int. J. Miner. Metall. Mater., 30(2023), No. 10, pp.1868-1877. https://doi.org/10.1007/s12613-023-2685-7
Cite this article as: Jinshan Wang, Feng Li, Si Zhao, Lituo Zheng, Yiyin Huang, and Zhensheng Hong, Uniform nanoplating of metallic magnesium film on titanium dioxide nanotubes as a skeleton for reversible Na metal anode, Int. J. Miner. Metall. Mater., 30(2023), No. 10, pp.1868-1877. https://doi.org/10.1007/s12613-023-2685-7
Research Article

Uniform nanoplating of metallic magnesium film on titanium dioxide nanotubes as a skeleton for reversible Na metal anode

Author Affilications
  • Corresponding author:

    Lituo Zheng      E-mail: zhenglituo@fjnu.edu.cn

    Zhensheng Hong      E-mail: zshong@fjnu.edu.cn

  • To meet the low-cost concept advocated by the sodium metal anode, this paper reports the use of a pulsed electrodeposition technology with ionic liquids as electrolytes to achieve uniform nanoplating of metallic magnesium films at around 20 nm on spaced titanium dioxide (TiO2) nanotubes (STNA-Mg). First, the sodiophilic magnesium metal coating can effectively reduce the nucleation overpotential of sodium metal. Moreover, three-dimensional STNA can limit the volume expansion during sodium metal plating and stripping to achieve the ultrastable deposition and stripping of sodium metals with a high Coulombic efficiency of up to 99.5% and a small voltage polarization of 5 mV in symmetric Na||Na batteries. In addition, the comparative study of sodium metal deposition behavior of STNA-Mg and STNA-Cu prepared by the same route further confirmed the advantage of magnesium metal to guide sodium metal growth. Finally, the prepared STNA-Mg–Na metal anode and commercial sodium vanadium phosphate cathode were assembled into a full cell, delivering a discharge capacity of 110.2 mAh·g−1 with a retention rate of 95.6% after 110 cycles at 1C rate.
  • The rapid development of energy storage systems for large-scale grids has pushed explosive demand for high-energy-density batteries [15]. Lithium-ion batteries have been well developed for use in portable electronic devices and electric vehicles, but the rising prices of materials have limited their wide application in large energy-storage power stations [68]. Sodium-based batteries are considered a cheap alternative to lithium batteries because sodium has similar electrochemical properties but is more abundant and cost-effective than lithium. With a high energy density of 1166 mAh·g−1 and a low redox potential (−2.71 V versus standard hydrogen electrode) [910], sodium metal is the anode material of choice for various high-energy-density sodium-based batteries, including sodium metal [910], sodium sulfur [1112], and sodium air batteries [1315]. Nevertheless, given the uncontrollable dendrite growth of sodium metal and the chemical reactivity between sodium metal and electrolytes, safety problems and unstable solid electrolyte interphase film are easily generated [1618]. Several strategies have been proposed to solve the problem caused by sodium metal anodes. The structural design of negative electrodes includes a three-dimensional (3D) conductive framework and surface modification of sodiophilic metal particles [1920]. These designs show satisfactory performance but are mostly expensive or supported by complex graphene materials [2122].

    Based on the heterogeneous nucleation reported by Cui and his co-authors and the successful application of magnesium in lithium-metal anodes [23], the modification of magnesium nanoparticles is a feasible method to regulate sodium metal deposition. Magnesium is an abundant and sustainable element in the Earth’s crust [2425]. Moreover, magnesium has strong activity and reducibility, and its surface easily forms a thin oxide film to prevent the further passivation of the metal anode [26]. However, most magnesium reduction still uses electrolysis and thermal reduction, both of which require a high temperature. Therefore, the treatment is relatively complex, and the preparation process is difficult to control.

    In this paper, we show that magnesium metal coating can be uniformly deposited on spaced titanium dioxide (TiO2) nanotubes by pulsed electrodeposition in a low-temperature molten electrolyte, forming spaced titanium dioxide nanotubes (STNA-Mg). STNA-Cu was also prepared via such a green and low-cost nanoplating route. Different metal modifications showed various effects on sodium metal nucleation and deposition. The nucleation barrier of sodium metal on the surface of STNA-Mg was significantly lower than that of STNA-Cu, demonstrating the sodiophilic nature of magnesium metal. Finally, the STNA-Mg framework as a skeleton for sodium metal anode exhibited an outstanding performance with a high Coulombic efficiency, small voltage polarization, and high reversibility during sodium metal plating and stripping. A full cell constructed by STNA-Mg–Na metal anode and sodium vanadium phosphate cathode also demonstrated excellent rate capability and very good cycling stability, which indicates the practicability of such technology in sodium metal batteries.

    Fig. 1 shows the preparation process. For the preparation of interstitial TiO2 nanotube array (STNA), titanium foil (99.99%, 0.1 mm) was first polished with 150-, 500-, and 1000-mesh sandpapers. The polished foil was then cut into 2.5 cm × 1.2 cm pieces, ultrasonicated in acetone, ethanol, and deionized water for 30 min, and dried. The etching electrolyte was prepared with a mass fraction of 0.3% NH4F, 3.7% HF (40wt%), 8% deionized water, and 88% dimethyl sulfoxide (DMSO). For the etching process, a treated titanium sheet was used as the anode and a platinum sheet as the cathode. The etching cell was held at a constant voltage of 40 V at 30°C for 1 h. The etched sample was rinsed in absolute ethanol for 1 h and dried to obtain the intergap TiO2 nanotube array precursor.

    Fig. 1.  Schematic of the preparation process of spaced titanium dioxide nanotube array (STNA) and magnesium-deposited spaced titanium dioxide nanotubes (STNA-Mg).

    The preparation of low-temperature molten salt (ionic liquid electrolyte) was performed in a drying room (air humidity below 30%). Choline chloride and urea with a molar ratio of 0.68:1 were weighed into a 100-mL volumetric flask and dried in a vacuum oven 60°C for 12 h to obtain 50 mL ionic liquid in a molten state.

    For the electrodeposition of magnesium (STNA-Mg) on interstitial TiO2 nanotubes, an appropriate amount of magnesium acetate (Mg(CH3COO)2) was added to 50 mL DMSO, and the mixture was stirred at room temperature for 4 h and then mixed with the prepared ionic liquid at 50°C. The prepared TiO2 nanotubes were used as the cathode, the platinum sheet was used as the anode, and 30 mL ionic liquid was used as the electrolyte. Sample STNA-Mg1 was obtained by direct current (DC) 30 mA power supply (ton) for 100 s. Sample STNA-Mg2 was prepared by energization with a 30 mA current unidirectional pulse for 10 s and cut-off interval (toff) for 20 s with total energization for 100 s. The obtained electrode pieces were washed with alcohol and dried under an argon atmosphere. Other experimental procedures, including testing and characterizaion are described in supplementary information.

    The process of magnesium electrodeposition was explored on the basis of interstitial TiO2 nanotubes. Fig. S1(a) and (b) shows the STNA prepared by the classical anodic oxidation method. STNA was formed by the vertical arrangement of nanotubes with a diameter of 150 nm, a gap of (100 ± 20) nm, and a length of 2 μm. According to the literature survey, no mature method is available for magnesium in deep eutectic solvents (DESs), nor has there been a systematic study on magnesium electrodeposition in DESs. The main problems are the strong self-activity of magnesium and its sensitivity to water and air. The deposited particles easily agglomerate, and the formation of a uniform covering layer at the nanoscale is difficult to achieve. In this work, we show that pulse electrodeposition can refine the grains during cathodic reduction and promote the uniformity of the coating (Fig. 2). Fig. 2(b) and (c) shows that the STNA-Mg1 particles obtained by DC conduction exhibited serious agglomeration and uneven distribution. They accumulated at the mouth of some nanotubes, and the rest showed little to no accumulation. The particle size of sample STNA-Mg2 obtained by pulse electrodeposition was approximately 15–30 nm, with uniform distribution and almost no agglomeration (Fig. 2(e) and (f)). The key steps in the electrodeposition process are the formation of new crystal nuclei and the growth of crystals, and the competition between these two steps directly affects the size of grains formed in the coating [2728]. Three independent parameters can be tuned in pulse electrodeposition, namely pulse current density, pulse conduction time, and pulse turn-off time [29]. When the pulse current is applied to an electrolytic cell, the growth of crystals is hindered due to the existence of the pulse interval, which reduces the epitaxial growth and changes the growth trend, avoiding the formation of large crystals.

    Fig. 2.  (a) Schematic of direct current (DC) power supply, (b, c) scanning electron microscopy (SEM) top view of STNA-Mg1 obtained after 100 s of DC electrification, (d) schematic of unidirectional pulse, (e, f) SEM images of STNA-Mg2 obtained using 30 mA unidirectional pulse for 100 s, (g) X-ray diffraction patterns of STNA, STNA-Mg, and STNA-Cu, (h) transmission electron microscopy (TEM) patterns of STNA-Mg, (i) high-resolution TEM (HRTEM)patterns of STNA-Mg (inset: lattice distance measured in TEM), (j, k) X-ray photoelectron spectroscopy spectra of STNA-Cu, and (l, m) X-ray photoelectron spectroscopy spectra of STNA-Mg.

    At the beginning of the deposition, anions constantly formed new crystal nuclei at different sites, avoiding the continuous growth of anions at the initial nucleation position, such as in DC electrification. Subsequent anions were reduced and reunited step by step at the initial nucleation sites, and the crystal nuclei easily grew in size. We continuously analyzed the microscopic morphology of the section of STNA-Mg2. As shown in Fig. S1(c), a uniform coating layer can be observed on the wall of the interstitial TiO2 nanotubes without evident agglomeration, and the TiO2 nanotubes maintained their original morphology, with the tube gap ranging from 50 to 150 nm. Therefore, we selected the STNA-Mg2 obtained by pulsed electrodeposition as the subsequent electrode for a detailed study and named it STNA-Mg.

    In the same molten-salt system, Mg(CH3COO)2 was replaced with copper sulfate as the electrodeposited copper electrolyte, and copper-modified gap TiO2 nanotubes were obtained by the pulsed electrodeposition process and named as STNA-Cu. As shown in Fig. S1(e) and (f), particles with a size of 10–30 nm were uniformly attached to the wall of the nanotube. The cross section shows that the nanoparticles were uniformly attached to the lower two-thirds of the tube mouth. Similarly, Cu modification did not change the geometrical properties of nanotubes substantially. We did not find the peaks of copper metal and magnesium metal in the X-ray diffraction (XRD) pattern of the two samples, and only the anatase phase structure of titanium substrate and TiO2 was observed (Fig. 2(g)). This finding was due to the small sample particles and insufficient strength of the signal to reach the detection limit of XRD. The TEM images in Fig. 2(h) and (i) show the magnesium layer with a thickness of around 20 nm on the surface of the nanotubes; the high-resolution transmission electron microscopy (HRTEM) image (insert in Fig. 2(i)) also reveals the lattice stripes of magnesium metal, whose crystal plane spacing corresponds to the crystal plane of magnesium (101) [30]. In the X-ray photoelectron spectroscopy (XPS) map of the STNA-Cu sample, the binding energy of 952.55 eV in Cu 2p corresponds to that of Cu 2p1/2 at 932.3 eV, which indicates that most of the copper was present as copper metal [29]. The binding energy at 934.2 eV corresponds to Cu–O of Cu 2p3/2, which implies that a small amount of copper was oxidized on the surface. The 530.3 eV binding energy in O 1s corresponds to the typical O–Ti bond of TiO2. In the XPS of STNA-Mg, the binding energy of Mg 2p at 49.6 eV shown in the map indicates the existence of the zero-valence state of magnesium, which also verified the magnesium layer on the surface of the nanotube [31]. Combined with the above characterization analysis, the nanolayer attached by STNA-Mg to TiO2 nanotubes was mainly magnesium metal. The energy-dispersive X-ray spectroscopy (EDX) map in Fig. S2(a) and (b) shows that the nanoparticles of the two samples attached to the tube wall with uniform distribution, and the metal particles were slightly densely distributed in a circular manner. From the color depth of the mapping map of magnesium and copper, the content of magnesium was higher than that of copper, and the peak intensity of the latter was higher than that of the former in the energy spectrum. As presented in Table S1, the atomic content of magnesium was slightly higher than that of Cu.

    According to the findings of lithium-metal anode research, substrate materials that can be alloyed with lithium metal can reduce the nucleation overpotential and induce the uniform nucleation of lithium [23]. The physical and chemical properties of sodium are similar to those of lithium. Similarly, metals that can be alloyed with sodium are considered to play a key role in the nucleation of sodium metal [32]. The nucleation overpotential is a key indicator of sodiophilic property. The current density of 1 mA·cm−2 and deposition capacity of 0.5 mAh·cm−2 were used to test the sodium metal nucleation overpotential of STNA-Mg and STNA-Cu substrates (Fig. 3(a)). As shown in Fig. 3(b) and (c), the nucleation overpotential of the STNA-Mg base was 8 mV, and that of the STNA-Cu base was at least 14 mV. The modification of copper almost had no effect on STNA because the STNA itself had a certain sodium affinity. According to the binary phase diagram shown in Fig. S3, magnesium (<0.1at%, 97.7°C) can form solid solutions with sodium and can be mutually soluble at any ratio above 637°C, which should be an intrinsic advantage of this element. This finding also indicates that magnesium is a sodiophilic material. However, copper cannot form any form of solid solution with sodium at any temperature and ratio.

    Fig. 3.  Nucleation potential diagram of (a) STNA||Na, (b) STNA-Mg||Na, (c) STNA-Cu||Na; the charge–discharge curves of (d) STNA-Mg||Na and (e) STNA-Cu||Na; (f) Coulombic efficiency of Na anode on the three substrates with a capacity of 1 mAh·cm−2.

    The nucleation overpotential is an indication of the difficulty of sodium metal nucleation on the substrate, and Coulombic efficiency is the most direct variable reflecting the reversibility of metal deposition [3334]. As shown in Fig. 3(f), the STNA-Mg electrode can maintain 500 reversible cycles with a high Coulombic efficiency of up to 99.5%. However, the Coulombic efficiency of the STNA-Cu electrode fluctuated sharply after 200 cycles. This finding indicates that the dissolution capacity of sodium metal on the STNA-Cu substrate gradually deviated from the deposition capacity, and the reason for this phenomenon was the serious side reactions caused by the non-uniform deposition of sodium metal at the interface and the continuous accumulation of dead sodium. As displayed in the charge–discharge curve of STNA-Cu in Fig. 3(d) and (e), with the increased number of cycles, the internal polarization of the half-cell also increased, and the voltage fluctuated violently. The overpotential reached 30 mV in the 280th discharge. However, in the STNA-Mg half-cell, the overpotential of around 15 mV increased slowly with the number of cycles to 500 cycles, showing high, stable, and reversible sodium metal deposition and stripping. In the subsequent cycle with a higher capacity of 2 mAh·cm−2, the STNA-Mg half-cell also showed a high Coulombic efficiency (Fig. S4). The cell did not show an evident abnormal fluctuation of efficiency for 200 cycles, and the average Coulombic efficiency was 99.3% in the first 200 cycles (Fig. S4(b)). This finding indicates that STNA-Mg can maintain a relatively stable sodium deposition and stripping under a large capacity of sodium metal deposition. However, the Coulombic efficiency of STNA-Cu at 2 mAh·cm−2 capacity fluctuated dramatically almost from the beginning, dropping to 40% after less than 90 cycles. This result indicates that in the case of large-capacity sodium metal deposition, more side reactions occurred. The impedance of the half-cell gradually increased, and the charging cut-off voltage was easily reached during the charging state, which led to a significant decline in Coulombic efficiency. The electrochemical impedance spectroscopy (EIS) spectra (Fig. S5(a) and (b)) showed the impedance changes of STNA-Mg and STNA-Cu half cells after 200 cycles in the cycle process. The impedance of the STNA-Cu electrode increased to approximately 600 Ω, which is consistent with the gradual fluctuation of Coulombic efficiency after 200 cycles. This finding indicates that the internal interface impedance of the cell increased. The sodium metal deposition curve also shows that the nucleation resistance of sodium metal deposition and the polarization voltage increased. This result is due to the poor reversibility of sodium metal deposition and stripping, which prevents that part of sodium from being completely peeled off and reacting with electrolyte to form a new solid electrolyte interface (SEI) film and become inactive sodium. The accumulation of inactive sodium easily leads to an increase in impedance. However, the impedance of the STNA-Mg electrode only increased to 250 Ω after 200 cycles. The sodium deposition curve also reveals that its polarization voltage increased slowly with the number of cycles, which indicates that its interface impedance was small and had good stability.

    We predeposited 1 mAh·cm−2 sodium on STNA-Cu and STNA-Mg electrodes to assemble symmetric cells and selected a current density of 0.5 mA·cm−2 to test the interface stability at 50% discharge depth. Fig. 4(a) shows that the STNA-Mg symmetric battery exhibited a very small voltage plateau of 5 mV, which stabilized after 800 h of cycles. In addition, the voltage lag did not increase significantly, which indicates its good interface stability. The STNA-Cu cell also revealed a low overpotential of approximately 28 mV and can maintain a relatively stable voltage platform for more than 400 h. The increase in voltage polarization means that the impedance of the plating/dissolution interface of STNA-Cu sodium metal increased, which was consistent with the previous EIS test results. The Na||Na electrode showed a high voltage platform in the beginning, and the voltage plummeted to 0 after running for 74 h. Thus, bare sodium, which easily forms dendrites and internal short circuits in the process of deposition and stripping, has evident disadvantages as a metal negative electrode.

    Fig. 4.  (a) Time–voltage curves of three symmetrical batteries with a load capacity of 1 mAh·cm−2 and discharge depth of 50% and (b–d) the corresponding enlarged of (a).

    We deposited 0.04 mAh·cm−2 Na on STNA-Mg and STNA-Cu samples to observe the initial nucleation of sodium metal. Fig. 5(a) and (b) shows that sodium was uniformly distributed on the STNA-Mg electrode sheet, which almost completely covered the magnesium coating on the original nanotube. In addition, the tube wall of TiO2 nanotubes also thickened, and the sodium layer tightly adhered to the inner and outer walls of TiO2 nanotubes and the interval between tubes. On the surface of STNA-Cu, most sodium was preferentially deposited on the tube mouth modified by copper nanoparticles and accumulated at the tube mouth after continuous growth (Fig. 5(c) and (d)). This deposition topography easily led to the incomplete peeling of sodium metal during charging, and the side close to the current collector was preferentially peeled. The sodium metal near the separator side dissolved and lost electric contact; thus, the separator side reacted with the electrolyte to form a new SEI film, which resulted in a gradual increase in impedance. Sodium metal tended to remain on the surface due to the influence of polarization. The EDX map of STNA-Mg–Na shows that sodium covered the original magnesium and exhibited a strong signal peak in the EDX spectrum. However, the peak intensity of magnesium was considerably weaker than that before sodium metal deposition due to the weakened signal after magnesium was covered by a large amount of sodium.

    Fig. 5.  Scanning electron microscopy images of nucleation of STNA-Mg (a, b) and STNA-Cu (c, d), electroplating capacity of 0.04 mAh·cm−2 sodium; energy-dispersive X-ray spectroscopy element mapping (e) and energy spectrum (f) of STNA-Mg–Na.

    When the deposition capacity of sodium metal reached 0.2 mAh·cm−2, sodium metal was completely deposited on the surface of the polar sheet (Fig. 6(a) and (b)), showing a porous network on the surface of STNA-Cu. Sodium was preferentially deposited on the surface of Cu particles, and a hollow shape appeared in the middle with the increase in sodium deposition. The number of sodium metals deposited at the nozzle gradually increased and connected to each other. In addition, from the cross-section images in Fig. 6(c) and (d), some space still existed between the original nanotubes, and dendrites grew on some voids. This deposition morphology was unfavorable to the reversible deposition and stripping of sodium metal. The terminal sodium metal was easily separated from the substrate and dissolved in the electrolyte after the root was preferentially stripped. However, the pole plate of STNA-Mg showed excellent sodium deposition regulation performance. Fig. 6(e) and (f) shows that with the increase in sodium metal deposition, the nanotubes were gradually filled, and a large number of gaps between the tubes were continuously filled. Only the 20 nm pipe diameter was left between some pipes, and the diameter decreased to around 40 nm. In addition, from the cross-sections of Fig. 6(g) and (h), the gap between the nanotubes was filled to the brim. The gap between the tubes was filled from 50–100 nm to 10–30 nm, which was in sharp contrast to that in the STNA-Cu sample. Most sodium metals were deposited into and between the tubes. Therefore, no sodium metal accumulation occurred on the surface of the pole sheet. Given the mechanical stability of TiO2 nanotubes, the skeleton can contain these sodium metals and become an excellent carrier space without destroying the original structure morphology. We further studied the condition of 0.5 mAh·cm−2 sodium deposited on STNA-Mg. Fig. S6(a) and (b) shows that sodium metal has been deposited on the surface of the sheet, and the original magnesium-coated TiO2 nanotubes have been filled. Sodium metal has been deposited on the surface of the polar plate as a smooth thin layer. The cross-sections in Fig. S6(c) and (d) also reveal that the space between the tubes was filled, and the shape of the nanotubes can be vaguely observed. The results indicate that the interstitial TiO2 nanotube skeleton modified by magnesium coating exhibited a good regulatory function on the deposition of sodium metal and limited the deposition of sodium metal in the lower limit domain of the nanometer scale.

    Fig. 6.  Scanning electron microscopy (SEM) images of STNA-Cu deposited with 0.2 mAh·cm−2 sodium from the top (a, b) and cross section (c, d) views. SEM images of STNA-Mg deposited with 0.2 mAh·cm−2 sodium from the top (e, f) and cross section (g, h) views.

    After observing the deposition behavior of sodium metal, we studied its stripping to determine the reversibility of sodium metal deposition behavior. Fig. 7(a) and (b) shows the residual sodium on the surface, which should be the residual sodium after the deposited sodium failed to be completely stripped. Moreover, Fig. 7(c) and (d) shows a large amount of sodium residue on the surface of the STNA-Cu half-cell after 250 cycles, which piled up into blocks of various shapes and almost clogged the tube mouth. This finding shows that in the process of sodium metal cycling, some sodium failed to peel off successfully and remained in the negative electrode. This portion of sodium then underwent side reactions with the electrolyte to generate a new SEI film on the surface. This event led to constant sodium buildup. However, the Coulombic efficiency of the STNA-Cu half-cell fluctuated sharply after 250 cycles, occasionally rising to 130%, which indicates that the complete stripping of sodium metal has become more difficult, and the deposition and stripping of sodium metal on the STNA-Cu electrode sheet was irreversible. Fig. 7(e) and (f) shows the STNA-Mg electrode after 50 cycles of charging and discharging. The electrode sheet was still an independent interstitial TiO2 nanotube. However, the tube wall slightly thickened, which means that a relatively stable SEI film has been formed. After 500 cycles (Fig. 7(g)), the morphology of the nanotube array and magnesium nanocoating remained stable, and the degree of blurring on the surface increased, which can be ascribed to the thickening of the SEI film. As shown in Fig. 7(h) and (i), after 500 cycles, the gap between the STNA-Mg electrode tube and the tube was still clear, and the tube gap changed from 10 nm in the discharge state to 50–100 nm in the initial morphology, which more intuitively indicates that sodium metal has excellent reversibility of deposition and peeling on the STNA-Mg electrode. The structural design of magnesium-coated interstitial nanotubes can effectively regulate the nucleation deposition and stripping of sodium metal, which also explains why the STNA-Mg half-cell can maintain a high Coulombic efficiency after 500 cycles.

    Fig. 7.  Top view of SEM image of the sodium stripping state of STNA-Cu after 50 (a, b) and after 250 (c, d) cycles of Coulombic efficiency test; top view of SEM image of the sodium stripping state of STNA-Mg after 50 (e, f) and 500 (g) cycles of Coulombic efficiency test; (h, i) cross section diagram of STNA-Mg after 500 cycles.

    Finally, to prove the feasibility of the design in practical application, we assembled a full cell with an NVP positive electrode and STNA-Mg and STNA-Cu with sodium deposition and studied its electrochemical performance (Fig. 8). Fig. 8(a) shows the rate performance of the three full batteries. STNA-Mg–Na showed the best rate performance, with a discharge capacity of 115 mAh·g−1 at 1C and 108 mAh·g−1 at 10C. Fig. 8(c) shows STNA-Mg–Na||NVP in 1, 50, and 100 cycles of charge and discharge curves. STNA-Mg–Na||NVP presented a constantly smaller voltage polarization at the beginning of the cycle and after approximately 20 mV. In addition, the STNA-Mg–Na||NVP (Fig. 8(e)) at 5C showed good cycle stability with a first-time Coulombic efficiency of 89.8% and a capacity retention rate of 88% afterward (100.2 mAh·g−1 after 500 cycles), which indicates its good applicability.

    Fig. 8.  Electrochemical performance of STNA-Mg–Na||NVP full cells (NVP: Na3V2(PO4)3): (a) rate performance of 0.5C–10C, (b) charge and discharge curve of 0.5C–10C, (c) charge and discharge curves at 1st, 50th, and 100th cycle, (d) cycle performance at 1C, and (e) cycle performance at 5C.

    In conclusion, we demonstrated the low-temperature and green nanoplating of metal films on a 3D substrate via the application of pulsed electrodeposition technology in ionic liquids. The metallic magnesium coating can effectively reduce the nucleation overpotential of sodium metal, and the 3D gap in spaced TiO2 nanotubes can provide a spatially controlled deposition. As a result, ultrastable sodium metal plating and stripping can achieve small polarization and high Coulombic efficiency. This study systematically compared the deposition and stripping of sodium metal on the surfaces of two different metal modification materials and concluded that magnesium coating is completely superior to copper in the deposition and stripping of sodium metal. As ionic liquids are fully in line with the concept of green chemistry and magnesium is a cheap and readily available metal, this research not only opens up a new and green method of metal nanoplating but also has a good application prospect in the field of sodium metal batteries.

    This work was financially supported by the National Natural Science Foundation of China (No. 51874099) and the National Science Foundation of Fujian Province’s Key Project, China (No. 2021J02031). Z.S. Hong thanks the support from the open fund from the Academy of Carbon Neutrality of Fujian Normal University, China (No. CZH2022-06).

    Zhensheng Hong is a youth editorial board member for this journal and was not involved in the editorial review or the decision to publish this article. All authors confirm that they have no competing interests or financial ties that could influence the outcomes or interpretation of this research.

    The online version contains supplementary material available at https://doi.org/10.1007/s12613-023-2685-7

  • [1]
    T.F. Liu, X.K. Yang, J.W. Nai, et al., Recent development of Na metal anodes: Interphase engineering chemistries determine the electrochemical performance, Chem. Eng. J., 409(2021), art. No. 127943. DOI: 10.1016/j.cej.2020.127943
    [2]
    B. Sun, P. Xiong, U. Maitra, et al., Design strategies to enable the efficient use of sodium metal anodes in high-energy batteries, Adv. Mater., 32(2020), No. 18, art. No. 1903891. DOI: 10.1002/adma.201903891
    [3]
    C. Delmas, Sodium and sodium-ion batteries: 50 years of research, Adv. Energy Mater., 8(2018), No. 17, art. No. 1703137. DOI: 10.1002/aenm.201703137
    [4]
    Z.Y. Feng, W.J. Peng, Z.X. Wang, et al., Review of silicon-based alloys for lithium-ion battery anodes, Int. J. Miner. Metall. Mater., 28(2021), No. 10, p. 1549. DOI: 10.1007/s12613-021-2335-x
    [5]
    L.F. Wang, J.Y. Wang, L.Y. Wang, M.J. Zhang, R. Wang, and C. Zhan, A critical review on nickel-based cathodes in rechargeable batteries, Int. J. Miner. Metall. Mater., 29(2022), No. 5, p. 925. DOI: 10.1007/s12613-022-2446-z
    [6]
    B. Lee, E. Paek, D. Mitlin, and S.W. Lee, Sodium metal anodes: Emerging solutions to dendrite growth, Chem. Rev., 119(2019), No. 8, p. 5416. DOI: 10.1021/acs.chemrev.8b00642
    [7]
    Y. Zhao, K.R. Adair, and X.L. Sun, Recent developments and insights into the understanding of Na metal anodes for Na-metal batteries, Energy Environ. Sci., 11(2018), No. 10, p. 2673. DOI: 10.1039/C8EE01373J
    [8]
    L.L. Fan and X.F. Li, Recent advances in effective protection of sodium metal anode, Nano Energy, 53(2018), p. 630. DOI: 10.1016/j.nanoen.2018.09.017
    [9]
    Z.P. Li, K.J. Zhu, P. Liu, and L.F. Jiao, 3D confinement strategy for dendrite-free sodium metal batteries, Adv. Energy Mater., 12(2022), No. 4, art. No. 2100359. DOI: 10.1002/aenm.202100359
    [10]
    W. Liu, P.C. Liu, and D. Mitlin, Review of emerging concepts in SEI analysis and artificial SEI membranes for lithium, sodium, and potassium metal battery anodes, Adv. Energy Mater., 10(2020), No. 43, art. No. 2002297. DOI: 10.1002/aenm.202002297
    [11]
    F. Jin, B. Wang, J.L. Wang, et al., Boosting electrochemical kinetics of S cathodes for room temperature Na/S batteries, Matter, 4(2021), No. 6, p. 1768. DOI: 10.1016/j.matt.2021.03.004
    [12]
    H. Kim, M.K. Sadan, C. Kim, et al., Enhanced reversible capacity of sulfurized polyacrylonitrile cathode for room-temperature Na/S batteries by electrochemical activation, Chem. Eng. J., 426(2021), art. No. 130787. DOI: 10.1016/j.cej.2021.130787
    [13]
    X.T. Lin, Y.P. Sun, Q. Sun, et al., Reviving anode protection layer in Na–O2 batteries: Failure mechanism and resolving strategy, Adv. Energy Mater., 11(2021), No. 11, art. No. 2003789. DOI: 10.1002/aenm.202003789
    [14]
    J.F. Xie, Z. Zhou, and Y.B. Wang, Metal–CO2 batteries at the crossroad to practical energy storage and CO2 recycle, Adv. Funct. Mater., 30(2020), No. 9, art. No. 1908285. DOI: 10.1002/adfm.201908285
    [15]
    Q.Y. Guo and Z.J. Zheng, Rational design of binders for stable Li–S and Na–S batteries, Adv. Funct. Mater., 30(2020), No. 6, art. No. 1907931. DOI: 10.1002/adfm.201907931
    [16]
    X.J. Lai, Z.M. Xu, X.F. Yang, et al., Long cycle life and high-rate sodium metal batteries enabled by regulating 3D frameworks with artificial solid-state interphases, Adv. Energy Mater., 12(2022), No. 10, art. No. 2103540. DOI: 10.1002/aenm.202103540
    [17]
    H. Wang, C.L. Wang, E. Matios, and W.Y. Li, Critical role of ultrathin graphene films with tunable thickness in enabling highly stable sodium metal anodes, Nano Lett., 17(2017), No. 11, p. 6808. DOI: 10.1021/acs.nanolett.7b03071
    [18]
    X.Y. Zheng, C. Bommier, W. Luo, L.H. Jiang, Y.N. Hao, and Y.H. Huang, Sodium metal anodes for room-temperature sodium-ion batteries: Applications, challenges and solutions, Energy Storage Mater., 16(2019), p. 6. DOI: 10.1016/j.ensm.2018.04.014
    [19]
    Z.X. Wang, Z.X. Huang, H. Wang, et al., 3D-printed sodiophilic V2CTx/rGO-CNT MXene microgrid aerogel for stable Na metal anode with high areal capacity, ACS Nano, 16(2022), No. 6, p. 9105. DOI: 10.1021/acsnano.2c01186
    [20]
    X.M. Xia, X. Lv, Y. Yao, et al., A sodiophilic VN interlayer stabilizing a Na metal anode, Nanoscale Horiz., 7(2022), No. 8, p. 899. DOI: 10.1039/D2NH00152G
    [21]
    J.L. Liang, W.W. Wu, L. Xu, and X.H. Wu, Highly stable Na metal anode enabled by a multifunctional hard carbon skeleton, Carbon, 176(2021), p. 219. DOI: 10.1016/j.carbon.2021.01.144
    [22]
    Z.W. Sun, Y.D. Ye, J.W. Zhu, et al., Regulating sodium deposition through gradiently-graphitized framework for dendrite-free Na metal anode, Small, 18(2022), No. 18, art. No. 2107199. DOI: 10.1002/smll.202107199
    [23]
    K. Yan, Z.D. Lu, H.W. Lee, et al., Selective deposition and stable encapsulation of lithium through heterogeneous seeded growth, Nat. Energy, 1(2016), art. No. 16010. DOI: 10.1038/nenergy.2016.10
    [24]
    S.A. Ferdousi, L.A. O’Dell, J. Sun, Y. Hora, M. Forsyth, and P.C. Howlett, High-performance cycling of Na metal anodes in phosphonium and pyrrolidinium fluoro(sulfonyl)imide based ionic liquid electrolytes, ACS Appl. Mater. Interfaces, 14(2022), No. 13, p. 15784. DOI: 10.1021/acsami.1c24812
    [25]
    C.L. Wei, L.W. Tan, Y.C. Zhang, Z.R. Wang, J.K. Feng, and Y.T. Qian, Towards better Mg metal anodes in rechargeable Mg batteries: Challenges, strategies, and perspectives, Energy Storage Mater., 52(2022), p. 299. DOI: 10.1016/j.ensm.2022.08.014
    [26]
    L.F. Zhao, Z. Hu, Z.Y. Huang, et al., In situ plating of Mg sodiophilic seeds and evolving sodium fluoride protective layers for superior sodium metal anodes, Adv. Energy Mater., 12(2022), No. 32, art. No. 2200990. DOI: 10.1002/aenm.202200990
    [27]
    N. Shahverdi, A. Montazeri, A. Khavandi, H.R. Rezaei, and F. Saeedi, Fabrication of nanohydroxyapatite-chitosan coatings by pulse electrodeposition method, J. Inorg. Organomet. Polym. Mater., 32(2022), No. 12, p. 4649. DOI: 10.1007/s10904-022-02468-w
    [28]
    B.S. Pan, Y.J. Yao, L. Peng, Q.X. Zhang, and Y. Yang, Ultrasound-assisted pulse electrodeposition of cobalt films, Mater. Chem. Phys., 241(2020), art. No. 122395. DOI: 10.1016/j.matchemphys.2019.122395
    [29]
    T.A. Green, X. Su, and S. Roy, Pulse electrodeposition of copper in the presence of a corrosion reaction, J. Electrochem. Soc., 168(2021), No. 6, art. No. 062515. DOI: 10.1149/1945-7111/ac0a21
    [30]
    B.Q. Cheng, X.J. Zhao, Y. Zhang, H.W. Chen, I. Polmear, and J.F. Nie, Co-segregation of Mg and Zn atoms at the planar η1-precipitate/Al matrix interface in an aged Al–Zn–Mg alloy, Scripta. Mater., 185(2020), p. 51. DOI: 10.1016/j.scriptamat.2020.04.004
    [31]
    R. Davidson, A. Verma, D. Santos, et al., Mapping mechanisms and growth regimes of magnesium electrodeposition at high current densities, Mater. Horiz., 7(2020), No. 3, p. 843. DOI: 10.1039/C9MH01367A
    [32]
    S. Tang, Y.Y. Zhang, X.G. Zhang, et al., Stable Na plating and stripping electrochemistry promoted by in situ construction of an alloy-based sodiophilic interphase, Adv. Mater., 31(2019), No. 16, art. No. 1807495. DOI: 10.1002/adma.201807495
    [33]
    J.D. Wan, R. Wang, Z.X. Liu, et al., A double-functional additive containing nucleophilic groups for high-performance Zn-ion batteries, ACS Nano, 17(2023), No. 2, p. 1610. DOI: 10.1021/acsnano.2c11357
    [34]
    Y. Li, M.H. Chen, B. Liu, Y. Zhang, X.Q. Liang, and X.H. Xia, Heteroatom doping: An effective way to boost sodium ion storage, Adv. Energy Mater., 10(2020), No. 27, art. No. 2000927. DOI: 10.1002/aenm.202000927
  • Related Articles

    [1]Qiong Long, Yun-bo Zhong, Huai Wang, Tian-xiang Zheng, Jun-feng Zhou, Zhong-ming Ren. Effects of magnetic fields on Fe-Si composite electrodeposition [J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(12): 1175-1186. DOI: 10.1007/s12613-014-1025-3
    [2]Lei Li, Zhi Wang, Ming-yong Wang, Yi Zhang. Modulation of active Cr(III) complexes by bath preparation to adjust Cr(III) electrodeposition [J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(9): 902-908. DOI: 10.1007/s12613-013-0813-5
    [3]A. Jegan, R. Venkatesan. Characterization and optimization of pulse electrodeposition of Ni/nano-Al2O3 composite coatings [J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(5): 479-485. DOI: 10.1007/s12613-013-0754-z
    [4]Ting Liu, Zhan-cheng Guo, Zhi Wang, Ming-yong Wang. Effects of gravity on the electrodeposition and characterization of nickel foils [J]. International Journal of Minerals, Metallurgy and Materials, 2011, 18(1): 59-65. DOI: 10.1007/s12613-011-0400-6
    [5]Xiao-long Li, Ke-feng Cai, Hui Li, Ling Wang, Chi-wei Zhou. Electrodeposition and characterization of thermoelectric Bi2Se3 thin films [J]. International Journal of Minerals, Metallurgy and Materials, 2010, 17(1): 104-107. DOI: 10.1007/s12613-010-0118-x
    [6]Shu-huei Hsieh, Wen-jauh Chen, Pei-i Wei, Jiing-herng Lee. Synthesis of the CuInSe2 thin film for solar cells using the electrodeposition technique and Taguchi method [J]. International Journal of Minerals, Metallurgy and Materials, 2009, 16(1): 101-107. DOI: 10.1016/S1674-4799(09)60017-0
    [7]Qiaoying Zhou, Hongliang Ge, Guoying Wei, Qiong Wu. Influence of bath composition on the electrodeposition of cobalt-molybdenum amorphous alloy thin films [J]. International Journal of Minerals, Metallurgy and Materials, 2008, 15(5): 611-617. DOI: 10.1016/S1005-8850(08)60114-0
    [8]Qinghua Huang, Wei Wang, Falong Jia, Zhirong Zhang. Preparation of Bi2-xSbxTe3 thermoelectric films by electrodeposition [J]. International Journal of Minerals, Metallurgy and Materials, 2006, 13(3): 277-280. DOI: 10.1016/S1005-8850(06)60058-3
    [9]Huajun Guo, Xinhai Li, Luping Liu, Xinming Zhang, Zhixing Wang, Wenjie Peng, Bingquan Zhu. Effects of temperature and concentration of sulfuric acid on the electrodeposition of grainy electrolytic manganese dioxide [J]. International Journal of Minerals, Metallurgy and Materials, 2005, 12(6): 553-557.
    [10]T. Kikuchi, M. Sakairi, H. Takahashi. Fabrication of independent nickel microstructures with anodizing of aluminum, laser irradiation, and electrodeposition [J]. International Journal of Minerals, Metallurgy and Materials, 2003, 10(3): 39-42.
  • Cited by

    Periodical cited type(3)

    1. Chen Chen, Hongyu Xue, Qilin Hu, et al. Construction of 3D porous Cu1.81S/nitrogen-doped carbon frameworks for ultrafast and long-cycle life sodium-ion storage. International Journal of Minerals, Metallurgy and Materials, 2025, 32(1): 191. DOI:10.1007/s12613-024-2890-z
    2. Zehua Chen, Jielei Huang, Ting Wang, et al. Self-templated formation of hierarchically porous TiO2 hollow microspheres with fast lithium-uptake performance. Applied Surface Science, 2024, 654: 159462. DOI:10.1016/j.apsusc.2024.159462
    3. Qipeng Lu, Zhihong Du, Jie Wang, et al. Editorial for special issue on renewable energy conversion, utilization and storage. International Journal of Minerals, Metallurgy and Materials, 2023, 30(10): 1855. DOI:10.1007/s12613-023-2746-y

    Other cited types(0)

Catalog

    Figures(8)

    Share Article

    Article Metrics

    Article views (564) PDF downloads (29) Cited by(3)

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return