Next Article in Journal
Synthesis of 4-(Phenylchalcogenyl)tetrazolo[1,5-a]quinolines by Bicyclization of 2-Azidobenzaldehydes with Phenylchalcogenylacetonitrile
Next Article in Special Issue
Efficient Fe3C-CF Cathode Catalyst Based on the Formation/Decomposition of Li2−xO2 for Li-O2 Batteries
Previous Article in Journal
Investigation of the Relationship between Electronic Structures and Bioactivities of Polypyridyl Ru(II) Complexes
Previous Article in Special Issue
Enhanced Capacity Retention of Li3V2(PO4)3-Cathode-Based Lithium Metal Battery Using SiO2-Scaffold-Confined Ionic Liquid as Hybrid Solid-State Electrolyte
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Stannate-Based Materials as Anodes in Lithium-Ion and Sodium-Ion Batteries: A Review

1
Key Laboratory of Comprehensive and Highly Efficient Utilization of Salt Lake Resources, Qinghai Institute of Salt Lakes, Chinese Academy of Sciences, Xining 810008, China
2
Key Laboratory of Salt Lake Resources Chemistry of Qinghai Province, Xining 810008, China
3
University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(13), 5037; https://doi.org/10.3390/molecules28135037
Submission received: 5 June 2023 / Revised: 19 June 2023 / Accepted: 23 June 2023 / Published: 27 June 2023
(This article belongs to the Special Issue Battery Chemistry: Recent Advances and Future Opportunities)

Abstract

:
Binary metal oxide stannate (M2SnO4; M = Zn, Mn, Co, etc.) structures, with their high theoretical capacity, superior lithium storage mechanism and suitable operating voltage, as well as their dual suitability for lithium-ion batteries (LIBs) and sodium-ion batteries (SIBs), are strong candidates for next-generation anode materials. However, the capacity deterioration caused by the severe volume expansion problem during the insertion/extraction of lithium or sodium ions during cycling of M2SnO4-based anode materials is difficult to avoid, which greatly affects their practical applications. Strategies often employed by researchers to address this problem include nanosizing the material size, designing suitable structures, doping with carbon materials and heteroatoms, metal–organic framework (MOF) derivation and constructing heterostructures. In this paper, the advantages and issues of M2SnO4-based materials are analyzed, and the strategies to solve the issues are discussed in order to promote the theoretical work and practical application of M2SnO4-based anode materials.

Graphical Abstract

1. Introduction

With the rapid development of human society in recent years, the need for energy has grown significantly. Environmental pollution caused by greenhouse gases such as carbon dioxide produced during the consumption of traditional fossil energy has prompted people to accelerate the development and use of green and clean energy [1,2,3]. However, renewable energy sources such as solar and wind are intermittent and diverse, and their large-scale expansion requires efficient and stable energy storage equipment [4,5,6]. Of all energy storage devices, alkali metal ion batteries are considered to be the most promising because of their high energy density, extended cycle life and environmental friendliness [7,8,9]. Since their introduction, lithium-ion batteries (LIBs) have become a mainstream energy storage device for mobile electronic devices, electric vehicles and other equipment [7,10]. However, commercial LIBs currently available are still unable to meet the requirements of large-scale energy storage devices such as electric vehicles and smart grids in terms of size, power output and energy density [11]. In the case of electric vehicles, there is diversity in the choice of cathode materials. BMW uses lithium nickel manganese cobalt oxide (NMC), while Tesla uses lithium nickel cobalt aluminum oxide (NCA), and electric vehicle companies in China often use lithium iron phosphate (LFP). Currently, carbonaceous materials are generally chosen as anode materials [12]. However, carbonaceous materials have a lower theoretical capacity (372 mAh·g−1) and take up about half the volume of the battery. Hence, higher-capacity anode materials need to be developed to replace carbon materials to match the high capacity of the anode materials and improve the energy storage capacity of LIBs. Nevertheless, the continued development and application of LIBs are leading to the rapid depletion of lithium resources, yet sodium resources are quite abundant. Therefore, SIBs, which use sodium resources as the main raw material, can be developed to replace a portion of LIBs, contributing to cost reduction and sustainability [13,14]. SIBs have a similar electrochemical storage mechanism to that of LIBs and can be adapted from LIB technology, allowing them to develop rapidly. Due to their significant advantage of lower cost, SIBs are likely to replace several LIBs as an important classification of energy storage devices. However, due to the relatively large radius of Na+ (0.59 Å for Li+ and 1.02 Å for Na+ in radius), significant volume expansion problems occur during the sodiation/desodiation process, resulting in poor energy density and cycle stability [15,16,17]. Due to this issue, Na+ is difficult to insert between the lattice layers of graphite; therefore, graphite anode material commonly used in LIBs cannot be applied to SIBs. Thus, selecting the right anode material is a key issue for both LIBs and SIBs.
Among the currently studied anode materials, which mainly include graphite, tin, germanium and silicon, a comparison of the capacity and volume variation of the elements in the IVA group is shown (Figure 1). Although the theoretical specific capacity of Sn is not the highest of them all, the volumetric specific capacity of Sn is quite close to that of Si and Ge. In particular, Sn has a much higher electrical conductivity (9.17 × 104 S cm−1) than several other materials. Thus, Sn-based materials such as Sn, SnO2, SnS, M2SnO4 and their composites have been extensively researched as they can have a very high theoretical capacity and a suitable potential platform [18,19,20,21]. Due to conversion and alloying reactions, the theoretical capacity of low-cost Sn dioxide can reach 1494 mAh·g−1, with the conversion reaction contributing 711 mAh·g−1 and the alloying reaction contributing 783 mAh·g−1. This is much higher than that of commonly used graphite materials, making Sn-based oxides a potential candidate for anode electrode materials [22,23,24].
As researchers investigated further, several serious problems with the Sn-based material were identified: (i) During the insertion/extraction of Li+ and Na+, Sn fully converts to Li22Sn5 and Na15Sn4, undergoing severe volume expansion (260% and 420%, respectively). This causes pulverization of the active material, allowing it to break away from the collector and leading to electrode failure [25,26,27,28]. This is detrimental to the cycling stability and reversible capacity of the battery. (ii) The inherent poor electrical and ionic conductivity of the material results in slow reaction kinetics, which reduces high-magnification performance. (iii) The poor reversibility of the conversion reaction between Sn and Li2O or Na2O during the initial cycle results in a large initial irreversible capacity loss as well as a low initial coulombic efficiency (ICE) [29]. In response to these problems, several studies have shown that the doping of Sn-based materials with other transition metal elements to produce metal oxide M2SnO4 can improve the properties [30,31,32]. On the one hand, due to the differences in expansion coefficients between different metal elements, redox reactions are enriched during the ion insertion/extraction process. This forms metal “M” or metal oxide “MOy” with different redox potentials, which can act as buffer substrates between each other to slow down the volume changes during cycling, thus ensuring the structural stability of the electrode material and reducing the capacity decay [33,34,35]. Furthermore, the conversion reaction of M can provide Sn with the required Li2O or Na2O, promoting the reversibility of the Sn conversion reaction. This improves the poor reversibility of the conversion reaction and reduces irreversible capacity loss [36]. On the other hand, combining different nanocrystals allows for the construction of heterogeneous structures at the contact interface. The internal electric field formed at the interface improves surface reaction kinetics, facilitating charge transfer and increasing material conductivity. For example, Mn2SnO4, a combination of Sn and Mn, has “metal-like conductivity” and improves the reaction kinetics [34,37]. In addition to overcoming the problems of tin-based materials, researchers have combined other modification strategies such as nanosizing, carbon compounding, heterostructuring and heteroatom doping.In recent years, stannate-based materials have gained great attention due to their superiority as anodes in LIBs and SIBs. In order to make stannate-based materials a viable alternative to anode materials, researchers have conducted a lot of exploration and proposed various modification strategies to address their problems. Table 1 provides a review of articles from the last decade on the research of different Sn-based materials. Among these excellent reviews, the main materials targeted include metallic Sn, SnO, SnO2, SnS, M2SnO4 and their composites, as shown in Table 1. For example, Huang Bin et al. [38] summarized studies on the mechanisms of action and institutional design of Sn-based oxides, sulfides, alloys and their compound materials for the storage of alkaline earth metal ions, including an analysis of the lithium storage mechanism of Co2SnO4 materials. Dong Sun’s review [39] gives special attention to two tin-based materials, ZnSnO3 and Zn2SnO4, outlining various synthetic strategies, formation mechanisms and morphology and summarizing their various applications. Although some of these excellent previous reviews mention the use of M2SnO4 for LIBs and SIBs, there is a lack of systematic review articles that address the energy storage mechanisms and material design of these materials. However, M2SnO4 cathode materials combining alloying and conversion reactions have shown great potential. It is therefore necessary to fill this gap by reviewing the rapid development of M2SnO4 materials as negative electrodes in the field of rechargeable batteries. In this review, we provide an overview of research on M2SnO4 as an anode in LIBs and SIBs. We systematically discuss the storage mechanisms, design of materials, and electrochemical properties of M2SnO4 anode materials in each section (Figure 2). In addition, we summarize and outline the future directions and prospects of M2SnO4-based anode materials.

2. M2SnO4-Based Anodes for LIBs

2.1. Lithium-Ion Storage

Transition metal stannate M2SnO4 is considered an ideal candidate for LIBs with high theoretical capacity values compared to current commercially available anode electrode materials (graphite) [37,53]. This is mainly since the storage of lithium ions in M2SnO4 materials (M2SnO4; M = Zn, Mn, Co, etc.) is carried out by both a conversion reaction mechanism and an alloying–dealloying mechanism. The lithiation mechanism of M2SnO4 materials is described by a four-stage reaction [23,33,53,54,55,56,57]. As shown in Equation (1), the first step is an irreversible reaction, consisting specifically of the decomposition of M2SnO4 into M and Sn and the formation of LiO2. Next, the second step is the alloying/dealloying reaction as expressed in Equation (2). This reaction allows Sn to bind up to 4.4 Li+, resulting in a high capacity, but also causing severe expansion of the Sn volume, which affects cycling stability. Finally, there is a conversion reaction between Sn and LiO2/SnO2 and M and LiO2/MO2, as indicated in Equations (3) and (4). The reversibility of the conversion reaction is a vital factor in maintaining the high capacity of the M2SnO4-based anode material.
M 2 SnO 4 + 8 Li + + 8 e 2 M + Sn + 4 LiO 2   ( M = Zn ,   Mn ,   Co   et   al )
S n + x Li + + x e L i x S n ( 0 < x < 4.4 )
S n + 2 L i 2 O S n O 2 + 4 Li + + 4 e
M + Li 2 O MO + 2 Li + + 2 e
During this lithiation/delithiation cycle, Reactions (3) and (4) form two metals with different redox potentials (metal–lithium alloys). On the one hand, reactions (3) and (4) can provide additional Li2O to Sn and Mrespectively, which constitutes a “synergistic effect” and promotes the reversibility of both [33]. On the other hand, the two act as a skeleton for each other to buffer the change in the volume of the material, ensuring the structural integrity of the material and reducing capacity degradation [34]. In addition, Figure 3 shows the CV curves of Mn2SnO4 and Co2SnO4, whose responses correspond to the energy storage mechanism described above. It is due to the superiority of this lithium storage mechanism that the development of M2SnO4 materials as anodes for LIBs is of great significance.

2.2. Design of M2SnO4-Based Anodes for LIBs

2.2.1. Nanostructures

Due to the advantages of nanosizing materials such as increasing the specific surface area and shortening the diffusion path of Li+, various structures of electroactive materials are designed and prepared on the nanoscale to improve the rate performance and cycle stability of batteries.
In terms of preparation methods, M2SnO4-based nanomaterials can be synthesized using a variety of techniques, including hydrothermal/solvothermal methods, solid-phase synthesis and sol–gel methods. Early studies by Irvine et al. [56] demonstrated the successful preparation of nanoscale M2SnO4 via solid-state synthesis. By comparing the Li+ storage behavior of these materials, they found that M2SnO4 exhibited better reversibility when M = Mn or Zn and poorer capacity reversibility when M = Mg. More recently, hydrothermal methods have emerged as a popular approach for the preparation of M2SnO4-based nanomaterials due to their advantages in terms of dispersion and ease of control. For instance, pure-phase Mn2SnO4 [58], Zn2SnO4 [59] and Co2SnO4 [23] nanoparticles prepared using this method have been shown to possess excellent electrochemical activity, with initial charge–discharge capacities of 1320, 1900.4 and 1533.1 mAh·g−1, respectively.
Wang et al. [23] recently reported a comparison of the properties of Co2SnO4 nanocrystals prepared using hydrothermal (HT-Co2SnO4) and high-temperature solid-state (SS-Co2SnO4) methods. Their results showed significant differences in the electrochemical properties of these materials (Figure 4c). After 50 cycles, the discharge capacity of SS-Co2SnO4 was 112.8 mAh·g−1 with a capacity retention rate of only 12.2%, which is much lower than that of HT-Co2SnO4 (555.9 mAh·g−1 with a capacity retention of 50.3%). SEM and TEM characterization revealed that HT-Co2SnO4 consisted of uniform spherical nanoparticles with diameters ranging from 80 to 100 nm, while SS-Co2SnO4 had an irregular shape with smaller particles (about 100 nm) coexisting with larger particles (1–2 μm) (Figure 4a,b). The particle size of the nanocrystals may be the main reason for the difference in electrochemical properties, with smaller nanoparticles obtained by the hydrothermal method exhibiting better electrochemical properties compared to those obtained by the high-temperature solid-state method.
In addition, some unique nanometer-size structures have been designed, such as hollow skeletons [60], hollow nanospheres [54], core–shell nanostructures [61], nanowires [62] and nanoplates [62], which are favorable for buffering volume expansion and maintain structural stability. Using a sol–gel method combined with phase separation, Wang et al. [60] synthesized macroporous Co2SnO4 with a hollow skeleton (Figure 5a,b) using polyacrylic acid (PAA) as the phase separation agent. As an anode for LIBs, the macroporous Co2SnO4 exhibited high capacity retention (115.5% at 200 mAh·g−1 after 300 cycles) and provided an ultrahigh specific capacity (921.8 mAh·g−1 at 1 A·g−1). Similarly, Zhang et al. [54] prepared hollow nanospheres (Figure 5c,d) using typical hydrothermal synthesis and heat treatment to serve as an anode for LIBs, which exhibited characteristics such as high multiplicative performance and cycling stability; even at a current density of 1 A·g−1, they still could achieve the reversible specific capacity of 442 mAh·g−1 after 60 cycles. In contrast to the above preparation methods, one-dimensional nanowires are usually prepared using the vapor phase method. Zn2SnO4 nanowires (Figure 5e,f)) were synthesized directly on stainless steel substrates by the vapor transport method. Constant current cycling studies of Zn2SnO4 nanowires at voltages ranging from 0.005 to 3 V and currents of 120 mA·g−1 have shown a reversible capacity of 1000 (±5) mAh·g−1, which is almost constant for the first 10 cycles and decays to 695 mAh·g−1 through 60 cycles thereafter [62]. The design of these special shapes provides additional cushioning space, relieves volume expansion and protects the stability of the structure.
Nanosizing the material and designing it with suitable structures, especially hollow, porous and unique structures, has been shown to effectively improve the bulk expansion of stannate-based materials.

2.2.2. Composited with Carbon Materials

In contrast to nanostructures, which improve material morphology, the incorporation of carbon materials optimizes material composition. By designing a combination of active and carbonaceous materials, the conductivity of the electrode material can be increased and its stability improved by buffering volume expansion. Commonly used carbon sources include glucose [61,63,64], fructose [55], graphene [53,65,66,67,68,69,70,71], carbon nanotubes [72,73] and other organic materials.
Previously, Qi and colleagues [61] synthesized a core–shell nanostructure of Co2SnO4@C using glucose as the carbon source. The uniform carbon layer (5–10 nm) coated on the surface of the Co2SnO4 material (Figure 6a) significantly improved the electrochemical performance of the electrode material. The capacity retention rate of Co2SnO4@C was much higher than that of pure-phase Co2SnO4 in terms of cycling performance (Figure 6b). EIS test results (Figure 6c) showed that the diameter of the semicircle in the mid-frequency region of Co2SnO4@C was smaller than that of Co2SnO4, indicating that it had a smaller charge transfer resistance. Similarly, a core–shell nanorod structure of Zn2SnO4@C was synthesized using a similar method. The nanorod is composed of continuous stripes with the same orientation and has a lattice spacing of about 0.26 nm and a uniform carbon coating layer of 5–10 nm (Figure 6d) [65]. The electrochemical performance of carbon composite materials was compared with pure-phase materials (Figure 5e,f), and it was found that the combination of carbon materials can effectively improve the stability and charge transfer impedance of the electrode. The presence of this carbon coating layer serves as a buffer base to enhance structural stability on one hand and improves electron conductivity on the other hand to accelerate charge transfer on the surface. Therefore, it can exhibit superior electrochemical performance.
Compared with traditional carbon materials, graphene materials possess excellent electrical conductivity, flexible structure and high specific surface area. In particular, they exhibit greatly improved electrochemical properties when composited with metal oxide anode materials [71,74]. Typically, a Zn2SnO4/graphene nanohybrid in flake form was prepared using the in situ hydrothermal method. In this anode system, GNSs act as a buffer to mitigate the volume change and as a separator to inhibit the aggregation of nanoparticles, thus improving the cycling stability. Additionally, the addition of GNSs provides a two-dimensional conductive channel for the Zn2SnO4 nanocrystals, enhancing the rate capability [68]. It is worth mentioning that graphene and M2SnO4 are combined via an electrostatic mechanism, which generally manifests itself as positively charged M2SnO4 nanoparticles being uniformly wrapped in negatively charged graphene under the influence of electrostatic forces [71]. Therefore, some method is needed to positively charge the surface of M2SnO4 to bind to the negatively charged graphene. For example, Co2SnO4 nanoparticles (Co2SnO4 NPs) were dispersed in CoCl2 solution so that they adsorbed Co2+ for positive charge [71]. Similarly, a Co2SnO4 hollow cube (Co2SnO4 HC) was modified with aminopropyltriethoxysilane (APTES) so that its surface was functionalized and positively charged (Figure 7a) [67]. As the uniform wrapping of graphene sheets (Figure 7b,c) effectively alleviates the volume change, the electrodes can maintain excellent conductivity throughout the discharge/charging process, so Co2SnO4 HC@rGO can provide more than 1000 mAh·g−1 capacity at 100 mA·g−1 after 100 cycles, while Co2SnO4 NPs@rGO can achieve the large reversible capacity of 1037.9 mAh·g−1 after 200 cycles. In addition to the above-listed Co2SnO4, the graphene-coated Zn2SnO4 (Figure 7d,e) obtained by electrostatic interaction also exhibits enhanced properties compared to the pure-phase material [69].
In contrast to the direct wrapping of graphene sheets on stannate-based material described above, the work of Rehman et al. [53] designed a bouquet-like nanocomposite structure in which graphene sheets were embedded in nanoparticles containing manganese and tin to achieve a high degree of bonding. The strategy benefits from a unique porous nanostructure in which high electronic pathways provided by graphene sheets are used to enhance electronic conductivity and uniformly distributed nanoparticles accelerate the kinetic reaction with lithium ions. Moreover, the graphene sheet also limits the growth of stannate nanoparticles, reducing the grain size. Due to the unique structure and conductive network of Mn2SnO4@GS, the Mn2SnO4@GS anode material exhibits excellent rate and cycling performance. At a current density of 400 mA·g−1, a specific capacity of about mAh·g−1 can be achieved after 200 cycles, while at a high current density of 2500 mA·g−1, a specific capacity of about 455 mAh·g−1 can still be provided (Figure 8a).
Two-dimensional graphene sheets and one-dimensional carbon nanotubes are the stars of the carbon materials world. Carbon nanotubes (CNTs) are frequently used as carbon materials compounded with metal oxides due to their superior electrical conductivity, high aspect ratio and large specific surface area [75,76,77]. Three-component Mn2SnO4@MWCNTs composites with cubic particles and high porosity anchored on carbon nanotubes were synthesized through a facile hydrothermal method [73]. It is clearly demonstrated by the impedance spectra that due to the synergy between Mn2SnO4 and MWCNTs, the electron transfer resistance in the LIBs is reduced [78], resulting in a much faster charge transfer in Mn2SnO4@MWCNTs than in bare Mn2SnO4 (Figure 8b).
In recent years, metal–organic frameworks (MOFs) have been widely investigated as materials that contain both metal and carbon sources with easily controlled structures [79,80,81,82], offering a new approach to the preparation of carbon-coated materials using the pyrolysis of MOFs [83,84]. Shi et al. [33] designed carbon-coated Mn2SnO4 nanomaterials with a two-dimensional combined micro/nanoscale configuration using a two-step carbonization process based on Mn-based metal–organic frameworks (Mn-MOFs) as a precursor. The Mn-MOFs play a role in providing both the Mn source and the flake-like porous carbon matrix. When comparing the cyclic and rate performance of Mn2SnO4@C with MnO/SnO2@C which has the same elemental composition, it was concluded that the “synergistic effect” in Mn2SnO4 [36] and the unique 2D structure brought by the MOF precursors contribute to the reversibility of the lithium storage reaction (Figure 8c). In contrast to the two-step method described above, Yue et al. [85] directly synthesized a Zn-Sn binary MOF (ZT-MOF) and then used rapid calcination under a reducing atmosphere to obtain Zn2SnO4@C/Sn composites. Interestingly, this work explored the effect of calcination time on the electrochemical properties of the material and found that the material calcined for 1 min had the best performance, mainly since metal Sn grew into large particles as the calcination time increased, which transformed the lithium storage process from surface-controlled to diffusion-controlled.
In summary, incorporating carbon materials through various methods to optimize material composition has proven to be an effective approach for mitigating volume expansion and enhancing the electrical conductivity of M2SnO4 materials. This approach has been widely adopted in the development of M2SnO4-based anode materials.

2.2.3. Heterogeneous Structures

The coupling between different nanocrystals builds heterogeneous structures at the contact interface, and the resulting Mott–Schottky heterogeneous junction generates an internal electric field at the interface, which improves the electrochemical dynamics [86,87]. However, the rational design and controlled synthesis of nano-heterostructured anode materials with high performance is still a challenge.
The yolk–shell structure is a classical form of a heterogeneous structure. Ju et al. [88] described the preparation of Mn-Sn-O-C composites with yolk–shell heterostructures by simple spray pyrolysis at various temperatures, where the shell and core portions were MnO-Mn2SnO4-C and Sn-Mn2SnO4-C, respectively. The Mn-Sn-O-C prepared at 900 °C had good structural stability with a discharge capacity of 784 mAh·g−1 over 100 cycles at a current density of 1 A·g−1 (Figure 9a–e). Similar to the yolk–shell structure, Tian et al. [89] used a three-step method combining co-precipitation, hydrothermal treatment and carbonization to design multi-yolk–shell SnO2/Mn2SnO4@C nanoboxes. In this report, it is shown that the SnO2/Mn2SnO4 heterostructure on the one hand produced lattice distortions in the internal material to improve thermodynamic stability, and on the other hand promoted reaction kinetics by hindering the coarsening of Sn, inducing a redistribution of electrons between SnO2 and Mn2SnO4 and accelerating the diffusive adsorption of Li+ through the internal electric field at the heterogeneous interface. Benefiting from the properties brought about by the heterogeneous structure, SnO2/Mn2SnO4@C nanoboxes as anode materials for LIBs exhibit a large reversible capacity (1293 mAh·g−1 at 0.2 A·g−1 after 100 cycles) and a stable long cycle performance (more than 549 cycles at 2 A·g−1).
Zhuang et al. [34] reported a novel heterogeneous composite material with excellent rate performance and cyclability by constructing a sandwich structure of graphene hollow spheres confined to a Mn2SnO4/SnO2 heterostructure (Mn2SnO4/SnO2@SG) as an anode for LIBs. Microscopically, the Mn2SnO4 and SnO2 nanoparticles form a heterogeneous structure next to each other between the graphene spherical shells; macroscopically, the structure appears to be a sandwiched hollow sphere. The Mn2SnO4/SnO2 Mott–Schottky heterojunction generates a strong electric field at the interface, which greatly facilitates the electronic/ionic transport kinetics and therefore shows excellent rate properties (823.8 mAh·g−1 at 5 C) (Figure 9f–h).
The heterogeneous structure is a modification method that involves introducing other materials compositionally to construct special structures. This approach has significantly improved the electrochemical performance of M2SnO4 anode materials. However, the process of constructing these structures can be complex and challenging.

2.2.4. Heteroatom Doping

Doping heteroatoms into the lattice of a material can generate a large number of active sites and improve electronic conductivity. This approach can be applied to M2SnO4 nanomaterials and carbon-based materials to enhance their electrochemical properties. Incorporating heteroatoms into these materials holds great potential for improving their performance in various applications [90].
Wang et al. [91] reported for the first time the fabrication of nanocomposites (Co-ZTO-G-C) consisting of ultrafine (3–5 nm) Co-doped Zn2SnO4 nanoparticles, graphene nanosheets and amorphous carbon layers using a one-step hydrothermal method. Since the radii of Co2+ (0.074 nm) and Zn2+ (0.074 nm) are so close to each other, the substitution of Co2+ for Zn2+ in the lattice of Zn2SnO4 is favorable, and no other byproducts such as cobalt oxide were found by X-ray diffraction analysis, which also proves the successful doping of Co2+. In particular, the crystallinity of the Co-ZTO-G-C nanocomposite is significantly higher than that of the other two nanocomposites, indicating that Co doping enhances the crystallinity of the Zn2SnO4 nanoparticles. The doping of N atoms into carbon-based materials such as N, P, B and S is also a common strategy. Wan et al. [37] doped N atoms into carbon and encapsulated Sn@Mn2SnO4 nanoparticles in N-doped carbon to be used as anode materials for LIBs. The doping of N atoms enhanced the electrical conductivity of the carbon shell and improved the lithium storage capacity.
Whether through the direct introduction of Co+ into nanomaterials to improve their crystallinity or the introduction of N atoms into composite carbon materials to increase active sites and improve electrical conductivity, doping heteroatoms has proven to be a feasible means of modifying M2SnO4 anode materials.

3. M2SnO4-Based Anodes for SIBs

3.1. Sodium Ion Storage

The sodium storage mechanism of the M2SnO4 material in SIBs involves a multi-step electrochemical redox reaction, which can be expressed by the following Equations (5)–(8) [24,57,92,93]. The initial step is an irreversible decomposition reaction as shown in Equation (5), which specifically includes the decomposition of M2SnO4 to M and Sn and the formation of NaO. The second step is the alloying/dealloying reaction expressed in Equation (6), where the formation of NaxSn causes the expansion of the volume. The final Equations (7) and (8) represent the conversion reactions of Sn and M.
M 2 SnO 4 + 8 Na + + 8 e 2 M + Sn + 4 Na 2 O   ( M = Zn ,   Mn ,   Co   et   al . )
Sn + x Na + + x e Na x Sn   ( 0 x 3.75 )
S n + 2 Na 2 O Sn O 2 + 4 Na + + 4 e
M + Na 2 O MO + 2 Na + + 2 e
The introduction of a second metal can enrich redox reactions compared to monometallics [35], and M and MO can be used as “substrates” to attenuate the effect of volume expansion on cycling performance caused by NaxSn generated by alloying reactions during cycling. For example, Huang et al. [94] developed CoMoO4@NC using two metals (Co and Mo) as an SIB anode material, which exhibited a long cycle life and maintained a specific capacity of 190 mAh·g−1 after 3200 cycles even at a high current density of 1 A·g−1.

3.2. Design of M2SnO4-Based Anodes for SIBs

3.2.1. Nanostructures

The use of nanostructures can effectively shorten the diffusion path of Na+, facilitate sodiation/desodiation, and reduce the adverse effects of volume expansion during cycling.
Park et al. [57] used the hydrothermal method to directly synthesize high-purity Zn2SnO4 nanowires and for the first time used them as an anode material for SIBs to investigate their electrochemical properties. In an SIB, Zn2SnO4 has a reversible capacity of 306 mAh·g−1 after 100 cycles at 0.1 C and maintains a high coulombic efficiency of about 99%. Nanowires (NWs) composed of a cubic (Zn2SnO4) phase have a homogeneous morphology, and the NWs are able to form a network to mitigate the volume change, which improves the cycling performance and extends the battery life. In this study, the working mechanism in Zn2SnO4 was also investigated by cyclic voltammetry (CV), I–V combined with ex situ XRD, and it was found that the alloying reaction did not proceed completely enough to reach the NaxSn alloy and lead to capacity reduction, while the reversible conversion reaction helped the capacity recovery. The above findings are consistent with previous studies on Sn and SnO2 [54,95].

3.2.2. Composited with Carbon Materials

Combining Sn-based materials with carbon materials and designing superior structures can effectively improve the material conductivity as well as the structural stability of electrodes and therefore has received a lot of attention.
Graphene is an excellent carbon material. N. Kalaiselvi’s team [92] designed Mn2SnO4 nanoporous cubes with graphene and used them for the first time as an anode material for SIBs. To prepare Mn2SnO4/graphene (MSO/G), the prepared Mn2SnO4 was sonicated with graphene in an ethanol solution, followed by oven drying and calcination under an argon atmosphere to ensure complete wrapping and adhesion of graphene sheets to the Mn2SnO4 nanocubes. In the SIB, the MSO/G composite anode exhibited high electrochemical performance at various current rates in terms of cycling capacity (257 mAh·g−1 after 100 cycles at 100 mA·g−1) and rate capacity (211 mAh·g−1 at 500 mA·g−1) and reached 106 mA h g−1 even after 1000 cycles at a high current density of 1 A·g−1. The presence of graphene networks and MSO nanocube voids greatly mitigates the shock caused by volume changes, ensuring that the composite anode material can maintain structural stability even under long cycling and enhancing the sodium storage capacity.

3.2.3. Heterogeneous Structures

According to the related reports, a heterogeneous structure constructed by combining bimetallic and carbon materials introduces defects, disorders and heterogeneous interfaces that can provide effective channels for the diffusion and adsorption of Na+, thus promoting the kinetics of the reaction [96,97,98].
Tian et al. designed novel nanoboxes with multiple yolk–shell structures (denoted as SnO2/Mn2SnO4@C) containing heterogeneously structured SnO2/Mn2SnO4 nanoparticles as the yolk and phenolic resin-derived carbon as the shell. The procedure for the preparation of SnO2/Mn2SnO4@C is represented in Figure 8a. Specifically, it includes the following steps: firstly, MnCl2, SnCl4 and NaOH are used as raw materials to obtain MnSn(OH)6 nanoboxes using the co-precipitation method; after that, the phenolic resin is wrapped on the surface of MnSn(OH)6 by hydrothermal reaction to obtain PR-MnSn(OH)6 nanoboxes; finally, heat treatment is performed to obtain SnO2/Mn2SnO4@C products. During the heat treatment, MnSn(OH)6 was transformed into heterostructured SnO2/Mn2SnO4 nuclei and PR was carbonized into carbon shells. As an SIB anode, it provides a large reversible capacity of 203 mA h g−1 after 100 cycles. In addition, according to the EIS spectra and fitting results, the SnO2/Mn2SnO4@C anode has a low Rct value (30.55 Ω) and a large DNa coefficient (8.72 × 10−14). The excellent electrochemical performance is mainly attributed to the following: the heterogeneous structure of SnO2/Mn2SnO4 effectively mitigates the volume expansion of Sn→NaxSn and improves the reversibility of conversion and alloying, the lattice distortion and the redistribution of charges at the heterogeneous interface accelerate the migration of Na+, and the hollow structure of SnO2/Mn2SnO4@C nanoboxes and the phenolic resin-derived carbon shell attenuate the crushing and agglomeration of the material.

3.2.4. Heteroatom Doping

The doping of heteroatoms (N, S, B, P, etc.) into carbon materials can provide more reactive sites and thus effectively improve the electrochemical properties of the materials, so this method of modification at the atomic level has gained great attention [99,100].
Using a typical synthesis method of hydrothermal treatment, carbon encapsulation and high-temperature treatment, Wan et al. [37] prepared the composite Sn@Mn2SnO4-NC containing Mn2SnO4 nanoparticles with Sn encapsulated in an N-doped carbon layer. Figure 10a shows synthesis path of Sn@Mn2SnO4-NC. In an SIB, the introduction of heteroatoms improves the electrical conductivity of the material, facilitates electron/ion transfer and accelerates the reaction kinetics, and the hollow structure combined with the carbon material provides the space needed for volume expansion/contraction and protects the material from crushing. The electrochemical reaction of Sn@Mn2SnO4-NC in SIBs is mainly controlled by pseudocapacitance according to the CV curves (Figure 10b,c). In addition, the capacitance contribution of SIBs is higher than that of LIBs at the same scan rate, which is mainly attributed to the fact that the radius of Na+ is larger than that of Li+, which hinders the insertion of ions.
Similarly, Kim et al. [99] reported the synthesis of Zn2SnO4 nanoparticles with uniformly encapsulated nitrogen-doped carbon layers using dopamine as a single carbon and nitrogen source, and the thickness of the carbon layer was adjusted by tuning the degree of polymerization of dopamine. The presence of the N-doped carbon cladding layer enhances the ionic/electronic conductivity, alleviates the volume expansion during electrode cycling, and prevents the aggregation of nanoparticles and the adverse effects caused by contact with the electrolyte.

4. Conclusions and Outlook

In conclusion, we have presented a comprehensive analysis of the development of M2SnO4-based nanomaterials as anodes for LIBs and SIBs. By incorporating other metallic elements into Sn-based materials to form M2SnO4 compounds (where M = Zn, Mn, Co, etc.), researchers have enriched the redox reactions involved in energy storage and created metal “M” or metal oxide “MO” buffer substrates. This effectively mitigates the issue of volume expansion during charge/discharge cycles. Table 2 and Table 3 summarizes a range of M2SnO4-based anode materials, categorized by synthesis techniques, morphological structures and electrochemical performance in LIBs and SIBs.
To improve the electrochemical performance of M2SnO4 as a battery anode material, researchers have explored various approaches, starting with pure-phase M2SnO4-based compounds. A primary focus has been on reducing the size of M2SnO4 materials to the nanoscale and designing various nanostructures (e.g., nanorods, nanowires, nanosheets, hollow nanospheres) to increase specific surface area and shorten Li+ and Na+ diffusion paths. Nanosized M2SnO4 materials consistently exhibit superior electrochemical properties compared to their larger counterparts. Three synthesis methods were discussed in the previous section: high-temperature solid phase, hydrothermal/solvothermal and vapor phase. Of these, hydrothermal synthesis is widely used due to its ability to produce high-purity materials with small particle sizes and controllable morphologies.
Subsequently, researchers have combined M2SnO4 with various carbon carriers to form carbon composites. In these composites, carbon carriers are encapsulated on the M2SnO4 nanostructure in the form of lamellae or shells. Carbon materials have been compounded with M2SnO4 using electrostatic forces to combine graphene with M2SnO4, high-temperature carbonization to encapsulate organic carbon sources (e.g., dopamine, fructose) on the surface of M2SnO4, and pyrolysis of MOF materials. The incorporation of carbon has a dual effect: it buffers the volume expansion of the active material during charge/discharge cycles and improves the electrical conductivity of the composite. As a result, hybrid M2SnO4-based materials with carbon exhibit enhanced electrochemical performance as anode materials.
Other researchers have coupled M2SnO4 with different nanomaterials to create heterogeneous structures with contact interfaces between different nanocrystals. These structures include yolk–shell, sandwich and multi-yolk–shell nanobox structures. The materials chosen for coupling are typically corresponding metal oxides and carbon carriers. The lattice distortion and charge redistribution at the heterogeneous interface between the two nanomaterials facilitate rapid Li+ and Na+ migration, thus improving reaction kinetics. Additionally, the unique structures created by this strategy can mitigate the crushing and agglomeration of active materials and enhance battery cycling stability.
Although nanosizing enhances the mass-specific capacity of materials, it concurrently diminishes their volumetric loading. Additionally, excessive carbon content may result in inadequate capacity. Modification strategies, such as the formation of heterostructures or doping with foreign elements, often necessitate intricate synthesis procedures. Consequently, selecting suitable modification techniques for distinct materials is of paramount importance. Table 2 and Table 3 provide detailed information on recent developments in M2SnO4-based anode materials for LIBs and SIBs. These tables summarize the composition, synthesis methods, material morphology and electrochemical properties of these materials.
As performance requirements for M2SnO4-based anode materials continue to increase, researchers are focusing on reducing particle size and studying the physicochemical properties of these materials in greater depth. However, many conventional experimental tools and characterization methods have significant limitations. Simulation calculations have emerged as an important tool in materials research and are widely used in the design of electrode materials, as well as in the calculation, prediction, validation, optimization and simulation of properties. For material design and preparation, simulation can provide an effective way to avoid unnecessary experiments. For mechanistic studies, it can elucidate the factors influencing electrochemical reactions and reveal microscopic mechanisms at the atomic level. Thus, simulation calculations will be a powerful aid in the study of M2SnO4-based anode materials.
Volume expansion and its associated problems remain major challenges in the development of M2SnO4-based materials. To address these issues, researchers are exploring the compounding of stannate materials with various other materials (including carbon carriers, metals and their oxides) and creating unique and refined morphological structures to improve and maintain the high capacity of stannate-based anode materials. Reducing grain size to less than 10 nm could be an effective strategy in terms of material size. In terms of material composition, the heterogeneous interface between different materials and the strength of chemical bonding should be considered. In terms of material morphology, a reasonable layout between the conductive network formed by the carbon material, active material and extra buffer space provided by hollow porous structures is crucial. In terms of heteroatom doping, introducing multiple heteroatoms into a suitable lattice structure could be a direction for future development. For example, while striving for the high weight capacity that nanosizing brings, it is important to also consider the volume capacity required for practical applications.
At the industrial application level, we suggest that using simple preparation methods to obtain M2SnO4@C composites by combining lower-cost carbon materials with M2SnO4 is the most promising direction. Although these low-cost carbon materials do not have the exceptional properties of graphene or carbon nanotubes, they can effectively mitigate volume expansion, prevent shedding and aggregation and improve the cycling stability of stannate-based anode materials. Additionally, more attention should be paid to comprehensive performance rather than solely pursuing excellence in one aspect.
M2SnO4-based materials are already widely used as anode materials in LIBs and SIBs due to their high theoretical capacity. Similarly, potassium-ion, magnesium-ion and calcium-ion batteries have similar reaction mechanisms, and M2SnO4-based materials could be potential candidates for their anode materials. In addition to these applications, M2SnO4-based materials may also have potential in emerging energy storage technologies. For example, lithium-ion hybrid capacitors (LICs) are a new type of energy storage device that combines a pre-lithiated anode from an LIB with an electric double-layer capacitor (EDLC)-type cathode. This device combines the high energy density of LIBs with the high power density and long cycle life of EDLCs and is considered to be a highly promising energy storage device. Stannate-based anode materials with high capacity are potential candidates for LIC anodes.
Despite the many challenges in the development of M2SnO4-based materials, we believe that they have great potential for use in the next generation of high-performance LIBs and SIBs. We hope that this review will be helpful in the development of M2SnO4-based anode materials.

Author Contributions

Conceptualization, Z.-H.F. and A.-J.J.; methodology, A.-J.J. and Y.-K.D.; software, T.S. and Z.-H.Z.; validation, A.-J.J., Z.-W.L. and S.-C.Z.; formal analysis, A.-J.J.; investigation, Y.-K.D. and A.-J.J.; resources, Z.-H.F.; data curation, Y.-K.D.; writing—original draft preparation, Y.-K.D.; writing—review and editing, Y.-K.D.; visualization, A.-J.J. and Y.-K.D.; supervision, Z.-H.F. and A.-J.J.; project administration, T.S. and Z.-H.Z.; funding acquisition, Z.-H.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Youth Innovation Promotion Association CAS (2018466) and the CAS “Light of West China” Program (Y810041019).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are thankful for the financial support of the Youth Innovation Promotion Association CAS (2018466) and the CAS “Light of West China” Program (Y810041019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, M.; Wang, H.; Luo, W.; Sherrell, P.C.; Chen, J.; Yang, J. Heterogeneous Single-Atom Catalysts for Electrochemical CO2 Reduction Reaction. Adv. Mater. 2020, 32, 2001848. [Google Scholar] [CrossRef] [PubMed]
  2. Ouyang, Y.; Ye, H.; Xia, X.; Jiao, X.; Li, G.; Mutahir, S.; Wang, L.; Mandler, D.; Lei, W.; Hao, Q. Hierarchical Electrodes of NiCo2S4 Nanosheets-Anchored Sulfur-Doped Co3O4 Nanoneedles with Advanced Performance for Battery-Supercapacitor Hybrid Devices. J. Mater. Chem. A 2019, 7, 3228–3237. [Google Scholar] [CrossRef]
  3. Feng, K.; Wang, F.; Yang, X.; Zhang, H.; Li, X.; Zhang, H. LiCr(MoO4)2: A New High Specific Capacity Cathode Material for Lithium Ion Batteries. J. Mater. Chem. A 2019, 7, 567–573. [Google Scholar] [CrossRef]
  4. Whittingham, M.S. Ultimate Limits to Intercalation Reactions for Lithium Batteries. Chem. Rev. 2014, 114, 11414–11443. [Google Scholar] [CrossRef]
  5. Chu, S.; Cui, Y.; Liu, N. The Path towards Sustainable Energy. Nat. Mater. 2017, 16, 16–22. [Google Scholar] [CrossRef]
  6. Wu, C.; Zhu, G.; Wang, Q.; Wu, M.; Zhang, H. Sn-Based Nanomaterials: From Composition and Structural Design to Their Electrochemical Performances for Li- and Na-Ion Batteries. Energy Storage Mater. 2021, 43, 430–462. [Google Scholar] [CrossRef]
  7. Li, W.-H.; Liang, H.-J.; Hou, X.-K.; Gu, Z.-Y.; Zhao, X.-X.; Guo, J.-Z.; Yang, X.; Wu, X.-L. Feasible Engineering of Cathode Electrolyte Interphase Enables the Profoundly Improved Electrochemical Properties in Dual-Ion Battery. J. Energy Chem. 2020, 50, 416–423. [Google Scholar] [CrossRef]
  8. Zhao, B.; Mattelaer, F.; Kint, J.; Werbrouck, A.; Henderick, L.; Minjauw, M.; Dendooven, J.; Detavernier, C. Atomic Layer Deposition of ZnO–SnO2 Composite Thin Film: The Influence of Structure, Composition and Crystallinity on Lithium-Ion Battery Performance. Electrochim. Acta 2019, 320, 134604. [Google Scholar] [CrossRef]
  9. Ying, H.; Han, W.-Q. Metallic Sn-Based Anode Materials: Application in High-Performance Lithium-Ion and Sodium-Ion Batteries. Adv. Sci. 2017, 4, 1700298. [Google Scholar] [CrossRef]
  10. Liu, P.; Zhu, K.; Xu, Y.; Bian, K.; Wang, J.; Tai, G.; Gao, Y.; Luo, H.; Lu, L.; Liu, J. Hierarchical Porous Intercalation-Type V2O3 as High-Performance Anode Materials for Li-Ion Batteries. Chem. Eur. J. 2017, 23, 7538–7544. [Google Scholar] [CrossRef]
  11. Schmuch, R.; Wagner, R.; Hörpel, G.; Placke, T.; Winter, M. Performance and Cost of Materials for Lithium-Based Rechargeable Automotive Batteries. Nat. Energy 2018, 3, 267–278. [Google Scholar] [CrossRef]
  12. Xin, F.; Whittingham, M.S. Challenges and Development of Tin-Based Anode with High Volumetric Capacity for Li-Ion Batteries. Electrochem. Energy Rev. 2020, 3, 643–655. [Google Scholar] [CrossRef]
  13. Rojo, T.; Hu, Y.-S.; Forsyth, M.; Li, X. Sodium-Ion Batteries. Adv. Energy Mater. 2018, 8, 1800880. [Google Scholar] [CrossRef] [Green Version]
  14. Vaalma, C.; Buchholz, D.; Weil, M.; Passerini, S. A Cost and Resource Analysis of Sodium-Ion Batteries. Nat. Rev. Mater. 2018, 3, 18013. [Google Scholar] [CrossRef]
  15. Kim, T.-H.; Shin, J.; Lee, K.-S.; Cho, E. Origin of the Superior Electrochemical Performance of Amorphous-Phase Conversion-Reaction-Based Electrode Materials for Na-Ion Batteries: Formation of a Bicontinuous Metal Network. ACS Appl. Mater. Interfaces 2020, 12, 22721–22729. [Google Scholar] [CrossRef]
  16. Li, Z.; Zhao, H. Recent Developments of Phosphorus-Based Anodes for Sodium Ion Batteries. J. Mater. Chem. A 2018, 6, 24013–24030. [Google Scholar] [CrossRef]
  17. Hwang, J.-Y.; Myung, S.-T.; Sun, Y.-K. Sodium-Ion Batteries: Present and Future. Chem. Soc. Rev. 2017, 46, 3529–3614. [Google Scholar] [CrossRef] [Green Version]
  18. Abe, J.; Takahashi, K.; Kawase, K.; Kobayashi, Y.; Shiratori, S. Self-Standing Carbon Nanofiber and SnO2 Nanorod Composite as a High-Capacity and High-Rate-Capability Anode for Lithium-Ion Batteries. ACS Appl. Nano Mater. 2018, 1, 2982–2989. [Google Scholar] [CrossRef]
  19. Kim, H.; Kim, H.; Muhammad, S.; Um, J.H.; Sher Shah, M.S.A.; Yoo, P.J.; Yoon, W.-S. Catalytic Effect of Reduced Graphene Oxide on Facilitating Reversible Conversion Reaction in SnO2 for Next-Generation Li Rechargeable Batteries. J. Power Sources 2020, 446, 227321. [Google Scholar] [CrossRef]
  20. Gao, C.; Jiang, Z.; Wang, P.; Jensen, L.R.; Zhang, Y.; Yue, Y. Optimized Assembling of MOF/SnO2/Graphene Leads to Superior Anode for Lithium Ion Batteries. Nano Energy 2020, 74, 104868. [Google Scholar] [CrossRef]
  21. Bhattacharya, P.; Lee, J.H.; Kar, K.K.; Park, H.S. Carambola-Shaped SnO2 Wrapped in Carbon Nanotube Network for High Volumetric Capacity and Improved Rate and Cycle Stability of Lithium Ion Battery. Chem. Eng. J. 2019, 369, 422–431. [Google Scholar] [CrossRef]
  22. Zhang, R.; Upreti, S.; Stanley Whittingham, M. Tin-Iron Based Nano-Materials as Anodes for Li-Ion Batteries. J. Electrochem. Soc. 2011, 158, A1498. [Google Scholar] [CrossRef]
  23. Wang, G.; Gao, X.P.; Shen, P.W. Hydrothermal Synthesis of Co2SnO4 Nanocrystals as Anode Materials for Li-Ion Batteries. J. Power Sources 2009, 192, 719–723. [Google Scholar] [CrossRef]
  24. Kebede, M.A. Tin Oxide–Based Anodes for Both Lithium-Ion and Sodium-Ion Batteries. Curr. Opin. Electrochem. 2020, 21, 182–187. [Google Scholar] [CrossRef]
  25. Zhang, F.; Wang, J.; Liu, S.; Du, Y. Effects of the Volume Changes and Elastic-Strain Energies on the Phase Transition in the Li-Sn Battery. J. Power Sources 2016, 330, 111–119. [Google Scholar] [CrossRef]
  26. Zhang, M.; Wang, T.; Cao, G. Promises and Challenges of Tin-Based Compounds as Anode Materials for Lithium-Ion Batteries. Int. Mater. Rev. 2015, 60, 330–352. [Google Scholar] [CrossRef]
  27. Muñoz-Márquez, M.Á.; Saurel, D.; Gómez-Cámer, J.L.; Casas-Cabanas, M.; Castillo-Martínez, E.; Rojo, T. Na-Ion Batteries for Large Scale Applications: A Review on Anode Materials and Solid Electrolyte Interphase Formation. Adv. Energy Mater. 2017, 7, 1700463. [Google Scholar] [CrossRef]
  28. Che, H.; Chen, S.; Xie, Y.; Wang, H.; Amine, K.; Liao, X.-Z.; Ma, Z.-F. Electrolyte Design Strategies and Research Progress for Room-Temperature Sodium-Ion Batteries. Energy Environ. Sci. 2017, 10, 1075–1101. [Google Scholar] [CrossRef]
  29. Hu, R.; Zhang, H.; Lu, Z.; Liu, J.; Zeng, M.; Yang, L.; Yuan, B.; Zhu, M. Unveiling Critical Size of Coarsened Sn Nanograins for Achieving High Round-Trip Efficiency of Reversible Conversion Reaction in Lithiated SnO2 Nanocrystals. Nano Energy 2018, 45, 255–265. [Google Scholar] [CrossRef]
  30. Mueller, F.; Bresser, D.; Chakravadhanula, V.S.K.; Passerini, S. Fe-Doped SnO2 Nanoparticles as New High Capacity Anode Material for Secondary Lithium-Ion Batteries. J. Power Sources 2015, 299, 398–402. [Google Scholar] [CrossRef] [Green Version]
  31. Guo, J.; Zhu, H.; Sun, Y.; Tang, L.; Zhang, X. Flexible Foams of Graphene Entrapped SnO2–Co3O4 Nanocubes with Remarkably Large and Fast Lithium Storage. J. Mater. Chem. A 2016, 4, 16101–16107. [Google Scholar] [CrossRef]
  32. Reddy, M.V.; Subba Rao, G.V.; Chowdari, B.V.R. Metal Oxides and Oxysalts as Anode Materials for Li Ion Batteries. Chem. Rev. 2013, 113, 5364–5457. [Google Scholar] [CrossRef] [PubMed]
  33. Shi, X.; Lin, X.; Liu, S.; Li, A.; Chen, X.; Zhou, J.; Ma, Z.; Song, H. Flake-like Carbon Coated Mn2SnO4 Nanoparticles as Anode Material for Lithium-Ion Batteries. Chem. Eng. J. 2019, 372, 269–276. [Google Scholar] [CrossRef]
  34. Zhuang, H.; Han, M.; Ma, W.; Ou, Y.; Jiang, Y.; Li, W.; Liu, X.; Zhao, B.; Zhang, J. Sandwich-Structured Graphene Hollow Spheres Limited Mn2SnO4/SnO2 Heterostructures as Anode Materials for High-Performance Lithium-Ion Batteries. J. Colloid Interface Sci. 2021, 586, 1–10. [Google Scholar] [CrossRef] [PubMed]
  35. Muralidharan, N.; Nallathamby, K. CuCo2S4: Versatile Anode for High Capacity and High Rate for Lithium and Sodium Ion Battery Application. J. Phys. Chem. Solids 2021, 151, 109902. [Google Scholar] [CrossRef]
  36. Chen, J.S.; Li, C.M.; Zhou, W.W.; Yan, Q.Y.; Archer, L.A.; Lou, X.W. One-Pot Formation of SnO2 Hollow Nanospheres and α-Fe2O3@SnO2 Nanorattles with Large Void Space and Their Lithium Storage Properties. Nanoscale 2009, 1, 280–285. [Google Scholar] [CrossRef]
  37. Wan, S.; Liu, Q.; Cheng, M.; Chen, Y.; Chen, H. Binary-Metal Mn2SnO4 Nanoparticles and Sn Confined in a Cubic Frame with N-Doped Carbon for Enhanced Lithium and Sodium Storage. ACS Appl. Mater. Interfaces 2021, 13, 38278–38288. [Google Scholar] [CrossRef]
  38. Huang, B.; Pan, Z.; Su, X.; An, L. Tin-Based Materials as Versatile Anodes for Alkali (Earth)-Ion Batteries. J. Power Sources 2018, 395, 41–59. [Google Scholar] [CrossRef]
  39. Sun, S.; Liang, S. Morphological Zinc Stannate: Synthesis, Fundamental Properties and Applications. J. Mater. Chem. A 2017, 5, 20534–20560. [Google Scholar] [CrossRef]
  40. Chen, J.S.; Lou, X.W.D. SnO2-Based Nanomaterials: Synthesis and Application in Lithium-Ion Batteries. Small 2013, 9, 1877–1893. [Google Scholar] [CrossRef]
  41. Zhao, Q.; Ma, L.; Zhang, Q.; Wang, C.; Xu, X. SnO2-Based Nanomaterials: Synthesis and Application in Lithium-Ion Batteries and Supercapacitors. J. Nanomater. 2015, 2015, 850147. [Google Scholar] [CrossRef] [Green Version]
  42. Zhao, Y.; Li, X.; Yan, B.; Li, D.; Lawes, S.; Sun, X. Significant Impact of 2D Graphene Nanosheets on Large Volume Change Tin-Based Anodes in Lithium-Ion Batteries: A Review. J. Power Sources 2015, 274, 869–884. [Google Scholar] [CrossRef]
  43. Liu, L.; Xie, F.; Lyu, J.; Zhao, T.; Li, T.; Choi, B.G. Tin-Based Anode Materials with Well-Designed Architectures for next-Generation Lithium-Ion Batteries. J. Power Sources 2016, 321, 11–35. [Google Scholar] [CrossRef]
  44. Zhao, M.; Zhao, Q.; Qiu, J.; Xue, H.; Pang, H. Tin-Based Nanomaterials for Electrochemical Energy Storage. RSC Adv. 2016, 6, 95449–95468. [Google Scholar] [CrossRef]
  45. Zhao, S.; Li, S.; Guo, T.; Zhang, S.; Wang, J.; Wu, Y.; Chen, Y. Advances in Sn-Based Catalysts for Electrochemical CO2 Reduction. Nano Micro Lett. 2019, 11, 62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Liang, J.-M.; Zhang, L.-J.; XiLi, D.-G.; Kang, J. Research Progress on Tin-Based Anode Materials for Sodium Ion Batteries. Rare Met. 2020, 39, 1005–1018. [Google Scholar] [CrossRef]
  47. Mou, H.; Xiao, W.; Miao, C.; Li, R.; Yu, L. Tin and Tin Compound Materials as Anodes in Lithium-Ion and Sodium-Ion Batteries: A Review. Front. Chem. 2020, 8, 141. [Google Scholar] [CrossRef]
  48. Dalapati, G.K.; Sharma, H.; Guchhait, A.; Chakrabarty, N.; Bamola, P.; Liu, Q.; Saianand, G.; Sai Krishna, A.M.; Mukhopadhyay, S.; Dey, A.; et al. Tin Oxide for Optoelectronic, Photovoltaic and Energy Storage Devices: A Review. J. Mater. Chem. A 2021, 9, 16621–16684. [Google Scholar] [CrossRef]
  49. Shinde, P.; Rout, C.S. Advances in Synthesis, Properties and Emerging Applications of Tin Sulfides and Its Heterostructures. Mater. Chem. Front. 2021, 5, 516–556. [Google Scholar] [CrossRef]
  50. Van Daele, K.; De Mot, B.; Pupo, M.; Daems, N.; Pant, D.; Kortlever, R.; Breugelmans, T. Sn-Based Electrocatalyst Stability: A Crucial Piece to the Puzzle for the Electrochemical CO2 Reduction toward Formic Acid. ACS Energy Lett. 2021, 6, 4317–4327. [Google Scholar] [CrossRef]
  51. Ansari, M.Z.; Ansari, S.A.; Kim, S.-H. Fundamentals and Recent Progress of Sn-Based Electrode Materials for Supercapacitors: A Comprehensive Review. J. Energy Storage 2022, 53, 105187. [Google Scholar] [CrossRef]
  52. Harish, S.; Sathyakam, P.U. A Review of Tin Selenide-Based Electrodes for Rechargeable Batteries and Supercapacitors. J. Energy Storage 2022, 52, 104966. [Google Scholar] [CrossRef]
  53. Ur Rehman, W.; Xu, Y.; Sun, X.; Ullah, I.; Zhang, Y.; Li, L. Bouquet-Like Mn2SnO4 Nanocomposite Engineered with Graphene Sheets as an Advanced Lithium-Ion Battery Anode. ACS Appl. Mater. Interfaces 2018, 10, 17963–17972. [Google Scholar] [CrossRef]
  54. Zhang, R.; He, Y.; Xu, L. Controllable Synthesis of Hierarchical ZnSn(OH)6 and Zn2SnO4 Hollow Nanospheres and Their Applications as Anodes for Lithium Ion Batteries. J. Mater. Chem. A 2014, 2, 17979–17985. [Google Scholar] [CrossRef]
  55. Liang, K.; Cheang, T.; Wen, T.; Xie, X.; Zhou, X.; Zhao, Z.; Shen, C.; Jiang, N.; Xu, A. Facile Preparation of Porous Mn2SnO4/Sn/C Composite Cubes as High Performance Anode Material for Lithium-Ion Batteries. J. Phys. Chem. C 2016, 120, 3669–3676. [Google Scholar] [CrossRef]
  56. Zhong, K.; Xia, X.; Zhang, B.; Li, H.; Wang, Z.; Chen, L. MnO Powder as Anode Active Materials for Lithium Ion Batteries. J. Power Sources 2010, 195, 3300–3308. [Google Scholar] [CrossRef]
  57. Lim, Y.R.; Jung, C.S.; Im, H.S.; Park, K.; Park, J.; Cho, W.I.; Cha, E.H. Zn2GeO4 and Zn2SnO4 Nanowires for High-Capacity Lithium- and Sodium-Ion Batteries. J. Mater. Chem. A 2016, 4, 10691–10699. [Google Scholar] [CrossRef]
  58. Lei, S.; Tang, K.; Chen, C.; Jin, Y.; Zhou, L. Preparation of Mn2SnO4 Nanoparticles as the Anode Material for Lithium Secondary Battery. Mater. Res. Bull. 2009, 44, 393–397. [Google Scholar] [CrossRef]
  59. Kim, K.; Annamalai, A.; Park, S.H.; Kwon, T.H.; Pyeon, M.W.; Lee, M.-J. Preparation and Electrochemical Properties of Surface-Charge-Modified Zn2SnO4 Nanoparticles as Anodes for Lithium-Ion Batteries. Electrochim. Acta 2012, 76, 192–200. [Google Scholar] [CrossRef]
  60. Wang, J.; Wang, J.; Guo, X.; Yang, H. Construction of Macroporous Co2SnO4 with Hollow Skeletons as Anodes for Lithium-Ion Batteries. Gels 2022, 8, 257. [Google Scholar] [CrossRef]
  61. Qi, Y.; Du, N.; Zhang, H.; Wu, P.; Yang, D. Synthesis of Co2SnO4@C Core-Shell Nanostructures with Reversible Lithium Storage. J. Power Sources 2011, 196, 10234–10239. [Google Scholar] [CrossRef]
  62. Cherian, C.T.; Zheng, M.; Reddy, M.V.; Chowdari, B.V.R.; Sow, C.H. Zn2SnO4 Nanowires versus Nanoplates: Electrochemical Performance and Morphological Evolution during Li-Cycling. ACS Appl. Mater. Interfaces 2013, 5, 6054–6060. [Google Scholar] [CrossRef] [PubMed]
  63. Jiang, T.; Tian, X.; Gu, H.; Zhu, H.; Zhou, Y. Zn2SnO4@C Core–Shell Nanorods with Enhanced Anodic Performance for Lithium-Ion Batteries. J. Alloys Compd. 2015, 639, 239–243. [Google Scholar] [CrossRef]
  64. Yuvaraj, S.; Amaresh, S.; Lee, Y.S.; Selvan, R.K. Effect of Carbon Coating on the Electrochemical Properties of Co2SnO4 for Negative Electrodes in Li-Ion Batteries. RSC Adv. 2014, 4, 6407–6416. [Google Scholar] [CrossRef]
  65. Chen, C.; Ru, Q.; Hu, S.; An, B.; Song, X.; Hou, X. Co2SnO4 Nanocrystals Anchored on Graphene Sheets as High-Performance Electrodes for Lithium-Ion Batteries. Electrochim. Acta 2015, 151, 203–213. [Google Scholar] [CrossRef]
  66. Wang, B.-Y.; Wang, H.-Y.; Ma, Y.-L.; Zhao, X.-H.; Qi, W.; Jiang, Q.-C. Facile Synthesis of Fine Zn2SnO4 Nanoparticles/Graphene Composites with Superior Lithium Storage Performance. J. Power Sources 2015, 281, 341–349. [Google Scholar] [CrossRef]
  67. Wang, B.; Zhang, J.; Xia, Z.; Shu, K. Facile Synthesis of Graphene Encapsulated Co2SnO4 Nanoparticles as Enhanced Anode Materials for Lithium-Ion Batteries. J. Nanosci. Nanotechnol. 2018, 18, 5679–5685. [Google Scholar] [CrossRef]
  68. Song, W.; Xie, J.; Hu, W.; Liu, S.; Cao, G.; Zhu, T.; Zhao, X. Facile Synthesis of Layered Zn2SnO4/Graphene Nanohybrid by a One-Pot Route and Its Application as High-Performance Anode for Li-Ion Batteries. J. Power Sources 2013, 229, 6–11. [Google Scholar] [CrossRef]
  69. Zhao, Y.; Huang, Y.; Sun, X.; Huang, H.; Wang, K.; Zong, M.; Wang, Q. Hollow Zn2SnO4 Boxes Wrapped with Flexible Graphene as Anode Materials for Lithium Batteries. Electrochim. Acta 2014, 120, 128–132. [Google Scholar] [CrossRef]
  70. Cui, L.; Li, X.; Yin, C.; Wang, J.; Li, S.; Zhang, Q.; Kang, S. Reduced Graphene Oxide Wrap Buffering Volume Expansion of Mn2SnO4 Anodes for Enhanced Stability in Lithium-Ion Batteries. Dalton Trans. 2019, 48, 504–511. [Google Scholar] [CrossRef]
  71. Zhang, J.; Liang, J.; Zhu, Y.; Wei, D.; Fan, L.; Qian, Y. Synthesis of Co2SnO4 Hollow Cubes Encapsulated in Graphene as High Capacity Anode Materials for Lithium-Ion Batteries. J. Mater. Chem. A 2014, 2, 2728–2734. [Google Scholar] [CrossRef]
  72. Wang, G.; Liu, Z.Y.; Liu, P. Co2SnO4-Multiwalled Carbon Nanotubes Composite as a Highly Reversible Anode Material for Lithium-Ion Batteries. Electrochim. Acta 2011, 56, 9515–9519. [Google Scholar] [CrossRef]
  73. Ali, G.; Mehboob, S.; Ahmad, M.; Akbar, M.; Kim, S.-O.; Ha, H.Y.; Chung, K.Y. High-Rate Lithium Storage and Kinetic Investigations of a Cubic Mn2SnO4@Carbon Nanotube Composite Anode. J. Alloys Compd. 2020, 823, 153789. [Google Scholar] [CrossRef]
  74. Huang, X.; Qi, X.; Boey, F.; Zhang, H. Graphene-Based Composites. Chem. Soc. Rev. 2012, 41, 666–686. [Google Scholar] [CrossRef]
  75. Schnorr, J.M.; Swager, T.M. Emerging Applications of Carbon Nanotubes. Chem. Mater. 2011, 23, 646–657. [Google Scholar] [CrossRef] [Green Version]
  76. Mishra, A.K.; Ramaprabhu, S. Magnetite Decorated Multiwalled Carbon Nanotube Based Supercapacitor for Arsenic Removal and Desalination of Seawater. J. Phys. Chem. C 2010, 114, 2583–2590. [Google Scholar] [CrossRef]
  77. Fan, Z.; Yan, J.; Zhi, L.; Zhang, Q.; Wei, T.; Feng, J.; Zhang, M.; Qian, W.; Wei, F. A Three-Dimensional Carbon Nanotube/Graphene Sandwich and Its Application as Electrode in Supercapacitors. Adv. Mater. 2010, 22, 3723–3728. [Google Scholar] [CrossRef]
  78. Hays, K.A.; Banek, N.A.; Wagner, M.J. High Performance Arsenic: Multiwall Carbon Nanotube Composite Anodes for Li-Ion Batteries. J. Electrochem. Soc. 2017, 164, A1635–A1643. [Google Scholar] [CrossRef]
  79. Xi, K.; Cao, S.; Peng, X.; Ducati, C.; Vasant Kumar, R.; Cheetham, A.K. Carbon with Hierarchical Pores from Carbonized Metal–Organic Frameworks for Lithium Sulphur Batteries. Chem. Commun. 2013, 49, 2192. [Google Scholar] [CrossRef]
  80. Wang, M.; Yang, H.; Zhou, X.; Shi, W.; Zhou, Z.; Cheng, P. Rational Design of SnO2@C Nanocomposites for Lithium Ion Batteries by Utilizing Adsorption Properties of MOFs. Chem. Commun. 2016, 52, 717–720. [Google Scholar] [CrossRef]
  81. Guan, B.Y.; Yu, X.Y.; Wu, H.B.; Lou, X.W.D. Complex Nanostructures from Materials Based on Metal-Organic Frameworks for Electrochemical Energy Storage and Conversion. Adv. Mater. 2017, 29, 1703614. [Google Scholar] [CrossRef] [PubMed]
  82. Zhu, B.; Wen, D.; Liang, Z.; Zou, R. Conductive Metal-Organic Frameworks for Electrochemical Energy Conversion and Storage. Coord. Chem. Rev. 2021, 446, 214119. [Google Scholar] [CrossRef]
  83. Shi, X.; Liu, S.; Tang, B.; Lin, X.; Li, A.; Chen, X.; Zhou, J.; Ma, Z.; Song, H. SnO2/TiO2 Nanocomposites Embedded in Porous Carbon as a Superior Anode Material for Lithium-Ion Batteries. Chem. Eng. J. 2017, 330, 453–461. [Google Scholar] [CrossRef]
  84. Li, A.; Song, H.; Bian, Z.; Shi, L.; Chen, X.; Zhou, J. ZnO Nanosheet/Squeezebox-like Porous Carbon Composites Synthesized by in Situ Pyrolysis of a Mixed-Ligand Metal–Organic Framework. J. Mater. Chem. A 2017, 5, 5934–5942. [Google Scholar] [CrossRef]
  85. Yue, H.; Shi, Z.; Wang, L.; Li, X.; Dong, H.; Yin, Y.; Yang, S. Rapid Calcination Synthesis of Zn2SnO4@C/Sn Composites for High-Performance Lithium Ion Battery Anodes. J. Alloys Compd. 2017, 723, 1018–1025. [Google Scholar] [CrossRef]
  86. Kennedy, T.; Bezuidenhout, M.; Palaniappan, K.; Stokes, K.; Brandon, M.; Ryan, K.M. Nanowire Heterostructures Comprising Germanium Stems and Silicon Branches as High-Capacity Li-Ion Anodes with Tunable Rate Capability. ACS Nano 2015, 9, 7456–7465. [Google Scholar] [CrossRef]
  87. Zheng, Y.; Zhou, T.; Zhang, C.; Mao, J.; Liu, H.; Guo, Z. Boosted Charge Transfer in SnS/SnO2 Heterostructures: Toward High Rate Capability for Sodium-Ion Batteries. Angew. Chem. 2016, 128, 3469–3474. [Google Scholar] [CrossRef]
  88. Ju, H.; Hong, Y.; Cho, J.; Kang, Y. Strategy for Yolk-Shell Structured Metal Oxide-Carbon Composite Powders and Their Electrochemical Properties for Lithium-Ion Batteries. Carbon 2016, 100, 137–144. [Google Scholar] [CrossRef]
  89. Tian, J.; Yang, L.; Zha, L.; Wang, R.; Huang, S.; Xu, G.; Wei, T.; Li, H.; Cao, J.; Wei, X. Heterostructured Multi-Yolk-Shell SnO2/Mn2SnO4@C Nanoboxes for Stable and Highly Efficient Li/Na Storage. J. Power Sources 2021, 506, 230243. [Google Scholar] [CrossRef]
  90. Yuan, Y.; Chen, Z.; Yu, H.; Zhang, X.; Liu, T.; Xia, M.; Zheng, R.; Shui, M.; Shu, J. Heteroatom-Doped Carbon-Based Materials for Lithium and Sodium Ion Batteries. Energy Storage Mater. 2020, 32, 65–90. [Google Scholar] [CrossRef]
  91. Wang, H.-Y.; Wang, B.-Y.; Meng, J.-K.; Wang, J.-G.; Jiang, Q.-C. One-Step Synthesis of Co-Doped Zn2SnO4–Graphene–Carbon Nanocomposites with Improved Lithium Storage Performances. J. Mater. Chem. A 2015, 3, 1023–1030. [Google Scholar] [CrossRef]
  92. Suresh babu, G.N.; Kalaiselvi, N. Graphene Encapsulated Mn2SnO4/G Nanocubes as High-Capacity and Cycle-Stable Anode Material for Sodium Ion Batteries. J. Alloys Compd. 2021, 889, 161679. [Google Scholar] [CrossRef]
  93. Zou, G.; Hou, H.; Zhao, G.; Ge, P.; Yin, D.; Ji, X. N-Rich Carbon Coated CoSnO3 Derived from in Situ Construction of a Co-MOF with Enhanced Sodium Storage Performance. J. Mater. Chem. A 2018, 6, 4839–4847. [Google Scholar] [CrossRef]
  94. Huang, X.; Zhang, W.; Zhou, C.; Yang, L.; Wang, H.; Gao, Q.; Zhu, M. N-Doped Carbon Encapsulated CoMoO4 Nanorods as Long-Cycle Life Anode for Sodium-Ion Batteries. J. Colloid Interface Sci. 2020, 576, 176–185. [Google Scholar] [CrossRef]
  95. Li, Z.; Ding, J.; Mitlin, D. Tin and Tin Compounds for Sodium Ion Battery Anodes: Phase Transformations and Performance. Acc. Chem. Res. 2015, 48, 1657–1665. [Google Scholar] [CrossRef]
  96. Ou, X.; Cao, L.; Liang, X.; Zheng, F.; Zheng, H.-S.; Yang, X.; Wang, J.-H.; Yang, C.; Liu, M. Fabrication of SnS2/Mn2SnS4/Carbon Heterostructures for Sodium-Ion Batteries with High Initial Coulombic Efficiency and Cycling Stability. ACS Nano 2019, 13, 3666–3676. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, Z.; Fei, S.; Wu, C.; Xin, P.; Huang, S.; Selegård, L.; Uvdal, K.; Hu, Z. Integrated Design of Hierarchical CoSnO3@NC@MnO@NC Nanobox as Anode Material for Enhanced Lithium Storage Performance. ACS Appl. Mater. Interfaces 2020, 12, 19768–19777. [Google Scholar] [CrossRef]
  98. He, T.; Feng, J.; Ru, J.; Feng, Y.; Lian, R.; Yang, J. Constructing Heterointerface of Metal Atomic Layer and Amorphous Anode Material for High-Capacity and Fast Lithium Storage. ACS Nano 2019, 13, 830–838. [Google Scholar] [CrossRef] [PubMed]
  99. Denis, P.A. Heteroatom Codoped Graphene: The Importance of Nitrogen. ACS Omega 2022, 7, 45935–45961. [Google Scholar] [CrossRef]
  100. Li, Z.; Lin, J.; Li, B.; Yu, C.; Wang, H.; Li, Q. Construction of Heteroatom-Doped and Three-Dimensional Graphene Materials for the Applications in Supercapacitors: A Review. J. Energy Storage 2021, 44, 103437. [Google Scholar] [CrossRef]
  101. An, B.; Ru, Q.; Hu, S.; Song, X.; Li, J. Facile Synthesis and Electrochemical Performance of Co2SnO4/Co3O4 Nanocomposite for Lithium-Ion Batteries. Mater. Res. Bull. 2014, 60, 640–647. [Google Scholar] [CrossRef]
  102. An, B.; Ru, Q.; Hu, S.; Song, X.; Chen, C. Enhanced Electrochemical Performance of Nanomilling Co2SnO4/C Materials for Lithium Ion Batteries. Ionics 2015, 21, 2485–2493. [Google Scholar] [CrossRef]
  103. Wu, B.; Zhang, S.; Yao, F.; Zhang, F.; Xu, S. Synergistic Lithium Storage of a Multi-Component Co2SnO4/Co3O4/Al2O3/C Composite from a Single-Source Precursor. RSC Adv. 2015, 5, 69932–69938. [Google Scholar] [CrossRef]
  104. Miao, X.; Ge, X.; Wang, P.; Zhao, D.; Yin, L. Size-Tunable SnO2/Co2SnO4 Nanoparticles Loaded 3D Reduced Graphene Oxide Aerogel Architecture as Anodes for High Performance Lithium Ion Batteries. Electrochim. Acta 2020, 356, 136769. [Google Scholar] [CrossRef]
  105. Mani, V.; Babu, G.; Kalaiselvi, N. Li2MnSnO4/C Anode for High Capacity and High Rate Lithium Battery Applications. Electrochim. Acta 2014, 133, 347–353. [Google Scholar] [CrossRef]
  106. Wang, J.; Wang, J.; Liu, W.; Guo, X.; Yang, H. Sol–Gel Synthesis of Dictyophora-Shaped Hierarchically Porous Mn2SnO4/C Materials as Anodes for Li-Ion Batteries. New J. Chem. 2021, 45, 9538–9549. [Google Scholar] [CrossRef]
  107. Yuan, W.S.; Tian, Y.W.; Liu, G.Q. Synthesis and Electrochemical Properties of Pure Phase Zn2SnO4 and Composite Zn2SnO4/C. J. Alloys Compd. 2010, 506, 683–687. [Google Scholar] [CrossRef]
  108. Zhang, R.; He, Y.; Li, A.; Xu, L. Facile Synthesis of One-Dimensional Mn3O4/Zn2SnO4 Hybrid Composites and Their High Performance as Anodes for LIBs. Nanoscale 2014, 6, 14221–14226. [Google Scholar] [CrossRef]
  109. Gao, Z.; Zhang, X.; Hu, H.; Guo, D.; Zhao, H.; Yu, H. Influencing Factors of Low- and High-Temperature Behavior of Co-Doped Zn2SnO4–Graphene–Carbon Nanocomposite as Anode Material for Lithium-Ion Batteries. J. Electroanal. Chem. 2017, 791, 56–63. [Google Scholar] [CrossRef] [Green Version]
  110. Chen, H.; Zhou, J.; Guo, F.; Wang, Y.; Chen, Y.; Liang, Y.; Xu, Y.; Zhang, H. High-Performance Zn2SnO4 Anodes Enabled by MOF-Derived MnO Decoration and Carbon Confinement for Lithium-Ion Batteries. CrystEngComm 2021, 23, 2590–2598. [Google Scholar] [CrossRef]
  111. Kim, N.; Shim, J.-H.; Jae, W.; Song, J.; Kim, J. Zn2SnO4 Particles Coated with N-Doped Carbon as an Anode Material for Lithium and Sodium-Ion Batteries. J. Alloys Compd. 2019, 786, 346–355. [Google Scholar] [CrossRef]
  112. Song, J.; Lu, X.; Tian, Q.; Cui, L.; Chen, J.; Sui, Z. Dual-Stable Engineering Enables High-Performance Zn2SnO4-Based Lithium-Ion Battery Anode. J. Alloys Compd. 2022, 910, 164924. [Google Scholar] [CrossRef]
Figure 1. Capacity and volume change comparison of IVA group elements [9]. Copyright 2017, Wiley.
Figure 1. Capacity and volume change comparison of IVA group elements [9]. Copyright 2017, Wiley.
Molecules 28 05037 g001
Figure 2. Key elements presented in this overview.
Figure 2. Key elements presented in this overview.
Molecules 28 05037 g002
Figure 3. (a) First few cyclic voltammetry curves of Mn2SnO4@GS [53]. Copyright 2018, American Chemical Society. (b) First few cyclic voltammetry curves of Co2SnO4. Copyright 2014, Elsevier.
Figure 3. (a) First few cyclic voltammetry curves of Mn2SnO4@GS [53]. Copyright 2018, American Chemical Society. (b) First few cyclic voltammetry curves of Co2SnO4. Copyright 2014, Elsevier.
Molecules 28 05037 g003
Figure 4. SEM images of the products synthesized via hydrothermal reaction (a) and high-temperature solid-state reaction (b). SEM images at high magnification are inserted. (c) Cycle performance of HT-Co2SnO4 and SS-Co2SnO4 at 30 mA·g−1 [23]. Copyright 2009, Elsevier.
Figure 4. SEM images of the products synthesized via hydrothermal reaction (a) and high-temperature solid-state reaction (b). SEM images at high magnification are inserted. (c) Cycle performance of HT-Co2SnO4 and SS-Co2SnO4 at 30 mA·g−1 [23]. Copyright 2009, Elsevier.
Molecules 28 05037 g004
Figure 5. SEM images of (a,b) Co2SnO4 hollow skeletons [60]. Copyright 2022, MDPI. TEM images of (c,d) Zn2SnO4 hollow nanospheres [54]. Copyright 2014, Royal Society of Chemistry. TEM images of (e,f) Zn2SnO4 nanowires [62]. Copyright 2013, American Chemical Society.
Figure 5. SEM images of (a,b) Co2SnO4 hollow skeletons [60]. Copyright 2022, MDPI. TEM images of (c,d) Zn2SnO4 hollow nanospheres [54]. Copyright 2014, Royal Society of Chemistry. TEM images of (e,f) Zn2SnO4 nanowires [62]. Copyright 2013, American Chemical Society.
Molecules 28 05037 g005
Figure 6. (a) TEM images of Co2SnO4@C-thick nanostructures, cycle performance at a current density of 100 mA·g−1 and Nyquist plots [61]. Copyright 2011, Elsevier. (b) TEM images of Zn2SnO4@C core–shell nanorods, cycle performance at a current density of 100 mA·g−1 and Nyquist plots [63]. Copyright 2015, Elsevier.
Figure 6. (a) TEM images of Co2SnO4@C-thick nanostructures, cycle performance at a current density of 100 mA·g−1 and Nyquist plots [61]. Copyright 2011, Elsevier. (b) TEM images of Zn2SnO4@C core–shell nanorods, cycle performance at a current density of 100 mA·g−1 and Nyquist plots [63]. Copyright 2015, Elsevier.
Molecules 28 05037 g006
Figure 7. (a) Schematic illustration of the formation of Co2SnO4 HC@rGO; TEM images of (b) Co2SnO4 HC and (c) Co2SnO4 HC@rGO [71]. Copyright 2014, Royal Society of Chemistry. (d) SEM and (e) TEM images of GN-wrapped Zn2SnO4 boxes [69]. Copyright 2014, Elsevier.
Figure 7. (a) Schematic illustration of the formation of Co2SnO4 HC@rGO; TEM images of (b) Co2SnO4 HC and (c) Co2SnO4 HC@rGO [71]. Copyright 2014, Royal Society of Chemistry. (d) SEM and (e) TEM images of GN-wrapped Zn2SnO4 boxes [69]. Copyright 2014, Elsevier.
Molecules 28 05037 g007
Figure 8. (a) SEM and TEM images of Mn2SnO4@GS, rate capability at various current densities and cycle performance at a current density of 500 mA·g−1 [53]. Copyright 2018, American Chemical Society. (b) SEM and TEM images of Mn2SnO4@MWCNTs, Nyquist plots and cycle performance at a current density of 100 mA·g−1 [73]. Copyright 2020, Elsevier. (c) SEM and TEM images of Mn2SnO4@C, rate capability at various current densities and cycle performance at a current density of 100 mA·g−1 [33]. Copyright 2019, Elsevier.
Figure 8. (a) SEM and TEM images of Mn2SnO4@GS, rate capability at various current densities and cycle performance at a current density of 500 mA·g−1 [53]. Copyright 2018, American Chemical Society. (b) SEM and TEM images of Mn2SnO4@MWCNTs, Nyquist plots and cycle performance at a current density of 100 mA·g−1 [73]. Copyright 2020, Elsevier. (c) SEM and TEM images of Mn2SnO4@C, rate capability at various current densities and cycle performance at a current density of 100 mA·g−1 [33]. Copyright 2019, Elsevier.
Molecules 28 05037 g008
Figure 9. (a) Schematic illustration of the fabrication process of multi-yolk–shell SnO2/Mn2SnO4@C; (b) cycling performance of powders prepared at various temperatures, initial charge/discharge curves of powders prepared at various temperatures (300 °C/700 °C/900 °C); (ce) TEM image of powders prepared at various temperatures [88]. Copyright 2016, Elsevier. (f) Schematic illustration of the fabrication process of Mn2SnO4/SnO2@GO; (g) SEM and TEM images; (h) cycling performance and rate performance at various current densities of Mn2SnO4/SnO2@SG [34]. Copyright 2021, Elsevier.
Figure 9. (a) Schematic illustration of the fabrication process of multi-yolk–shell SnO2/Mn2SnO4@C; (b) cycling performance of powders prepared at various temperatures, initial charge/discharge curves of powders prepared at various temperatures (300 °C/700 °C/900 °C); (ce) TEM image of powders prepared at various temperatures [88]. Copyright 2016, Elsevier. (f) Schematic illustration of the fabrication process of Mn2SnO4/SnO2@GO; (g) SEM and TEM images; (h) cycling performance and rate performance at various current densities of Mn2SnO4/SnO2@SG [34]. Copyright 2021, Elsevier.
Molecules 28 05037 g009
Figure 10. (a) Schematic diagram of the synthesis route of Sn@Mn2SnO4-NC. (b) CV curves at various scan rates (I, II, III: specific redox peaks) (c) contribution ratios of the pseudocapacitive- and diffusion-controlled capacity at different scan rates for SIBs. Copyright 2021, American Chemical Society.
Figure 10. (a) Schematic diagram of the synthesis route of Sn@Mn2SnO4-NC. (b) CV curves at various scan rates (I, II, III: specific redox peaks) (c) contribution ratios of the pseudocapacitive- and diffusion-controlled capacity at different scan rates for SIBs. Copyright 2021, American Chemical Society.
Molecules 28 05037 g010
Table 1. Existing review literature on Sn-based materials.
Table 1. Existing review literature on Sn-based materials.
Sr
No.
Title of PaperReviewed MaterialYearCovers StannateRef.
1SnO2-Based Nanomaterials: Synthesis and Application in Lithium-Ion BatteriesSnO2-based nanomaterials2013No[40]
2SnO2-Based Nanomaterials: Synthesis and Application in Lithium-Ion Batteries and SupercapacitorsSnO2-based nanomaterials2015No[41]
3Significant impact of 2D graphene nanosheets on large volume change tin-based anodes in lithium-ion batteries: A reviewSn, SnO2, SnS2, MxSnOy (stannates)2015Section[42]
4Tin-based anode materials with well-designed architectures for next generation lithium-ion batteriesSn-based multi-component intermetallics, SnO2, SnS22016No[43]
5Tin-based nanomaterials for electrochemical energy storageSn; Sn–M (M-Co, Ni, Cd, Zn, Fe), SnO2, SnS2, Sn4P3, SnF22016No[44]
6Morphological zinc stannate: synthesis, fundamental properties and applicationsZnSnO3 and Zn2SnO42017Only
Zn-based
[39]
7Metallic Sn-Based Anode Materials: Application in High-Performance Lithium-Ion and Sodium-Ion BatteriesSn; Sn in carbon matrix, Sn alloy2017No[9]
8Tin-based materials as versatile anodes for alkali (earth)-ion batteriesSnO2, SnS2, Sn Alloy, MxSnOy2018Section[38]
9Advances in Sn-Based Catalysts for Electrochemical CO2 ReductionSnO2, SnS2, tin alloy and its composites2019No[45]
10Tin oxide–based anodes for both lithium-ion and sodium-ion batteriesSnO2 and its composites2020No[24]
11Research progress on tin-based anode materials for sodium ion batteriesSn, SnO2, SnSe, SnS
and its composites
2020No[46]
12Tin and Tin Compound Materials as Anodes in Lithium-Ion and Sodium-Ion Batteries: A ReviewTin alloy, SnO2, SnS2 and its composites2020No[47]
13Challenges and Development of Tin-Based Anode with High Volumetric Capacity for Li-Ion BatteriesTin alloy and its composites2020No[12]
14Tin oxide for optoelectronic, photovoltaic and energy storage devices: a reviewSnO2 and its composites2021No[48]
15Advances in Synthesis, Properties and Emerging Applications of Tin Sulfides and its HeterostructuresSnxSy and its composites2021No[49]
16Sn-Based Electrocatalyst Stability: A Crucial Piece to the Puzzle for the Electrochemical CO2 Reduction toward Formic AcidSn, SnO2 and its composites2021No[50]
17Sn-based nanomaterials: From composition and structural design to their electrochemical performances for Li- and Na-ion batteriesSn, SnO2, SnS, SnS2 and its composites2021No[6]
18Fundamentals and recent progress of Sn-based electrode materials for supercapacitors: A comprehensive reviewSnO2, SnSx and its composites2022No[51]
19A review of tin selenide-based electrodes for rechargeable batteries and supercapacitorsSnSe, SnSe2 and its composites2022No[52]
Table 2. Selected M2SnO4-based composites and their electrochemical performances for LIBs (ICE: initial coulombic efficiency (%); cycling stability: specific capacity (mAh·g−1)/cycle/current density (A·g−1)).
Table 2. Selected M2SnO4-based composites and their electrochemical performances for LIBs (ICE: initial coulombic efficiency (%); cycling stability: specific capacity (mAh·g−1)/cycle/current density (A·g−1)).
MaterialSynthesis MethodMorphologyICECycling StabilityRef.
Co2SnO4hydrothermal reactionnanoparticles71556/50/0.03[23]
Co2SnO4@Chydrothermal and heat treatmentcore–shell nanostructures-474/75/0.1[61]
Co2SnO4@MWCNTshydrothermal reaction3D network89898/50/0.05[72]
Co2SnO4/Co3O4co-precipitation methodspherical and polyhedral74.4702/50/0.1[101]
Co2SnO4@Csonochemical and hydrothermalcubic phase59742/30/0.04[64]
Co2SnO4 HC@rGOhydrothermal and heat treatmenthollow cubes691016/100/0.1[71]
Co2SnO4@Cco-precipitation process and high-energy ball millingspherical particles58573.8/100/0.2C[102]
Co2SnO4/Ghydrothermalnanoparticles691061/100/0.1[65]
Co2SnO4/Co3O4/Al2O3/Cco-precipitationparticle-like morphology-1170/100/0.1[103]
Co2SnO4 NPs@rGOhydrothermalnanoparticles63.41037/200/0.2[67]
SnO2/Co2SnO4@rGOAsol-gel and heat treatmentcubic phase70588/1500/1[104]
Co2SnO4/Csol–gel method combined with phase separationhollow skeletons79.5582/500/1[60]
Mn2SnO4hydrothermal and thermal decompositionnanoparticles42-[58]
Li2MnSnO4/Coxalyl dihydrazide-assisted combustion methodspherical morphology69610/100/0.1[105]
MnO/Mn2SnO4/C@ Sn/Mn2SnO4/Cspray pyrolysisyolk–shell66784/100/1[88]
Mn2SnO4/Sn/Chydrothermal and heat treatmentporous cubes59.6908/100/0.5[88]
Mn2SnO4@GShydrothermal and heat treatmentbouquet-like611070/200/0.4[53]
Mn2SnO4@rGOhydrothermal and heat treatmentnanoparticles-542/100/0.1[70]
Mn2SnO4@Chydrothermal and heat treatmentflake-like69.2986/100/0.1[33]
Mn2SnO4@MWCNTshydrothermal and heat treatmentcubic and nanotube72611/100/0.1C[73]
SnO2/Mn2SnO4@Chydrothermalmulti-yolk–shell nanoboxes50.551293/100/0.2[89]
Sn@ Mn2SnO4-NChydrothermal and heat treatmentcubic frame73.88823/600/1[37]
Mn2SnO4/Csol–gel and heat treatmentDictyophora-shaped hierarchically porous-784/500/1[106]
Mn2SnO4/SnO2@SGheat treatmenthollow spheres71.41180/100/0.1C[34]
Zn2SnO4/Chydrothermal and carbothermic reductionnanoparticles61563/40/-[107]
Zn2SnO4vapor transport/hydrothermalnanowires/nanoplates41470/50/-[62]
Zn2SnO4/Gin situ hydrothermallayered54688/50/0.2[68]
Mn3O4/Zn2SnO4hydrothermalnanorod/nanoneedle59.6529/50/0.5[108]
Zn2SnO4hydrothermalhollow nanospheres66.2602.5/60/0.1[54]
Zn2SnO4/Gco-precipitation and alkali etching methodhollow boxes62678.2/45/0.3[69]
Zn2SnO4@Chydrothermal and carbonization approachcore–shell nanorods53.6495/100/0.1[63]
Zn2SnO4/Ghydrothermalnanoparticles57.4492/500/0.5[66]
Zn2SnO4–graphene–carbonhydrothermalnanoparticles62.3461/200/0.2[91]
Zn2SnO4hydrothermalnanowires52.5983/100/0.1[57]
Zn2SnO4@C/Sncalcinationlarge spheres92.41140/100/0.1[85]
Co–ZTO–G–Chydrothermal and heat treatment-62.3695/50/0.1C[109]
Zn2SnO4/N-doped carbon compositehydrothermal and heat treatmentspherical shaped particles71.2992.4/100/0.6[110]
LC@Zn2SnO4@MnO/C(MOF)solvothermal method and high-temperature annealing treatmentporous micro/nanostructures65.91185.6/150/0.2[111]
Zn2SnO4@V@PCcarbonizationyolk–shell57.2438/600/1[112]
Table 3. Selected M2SnO4-based composites and their electrochemical performances for SIBs (ICE: initial coulombic efficiency (%); cycling stability: specific capacity (mAh·g−1)/cycle/current density (A·g−1)).
Table 3. Selected M2SnO4-based composites and their electrochemical performances for SIBs (ICE: initial coulombic efficiency (%); cycling stability: specific capacity (mAh·g−1)/cycle/current density (A·g−1)).
MaterialSynthesis MethodMorphologyICECycling StabilityRef.
Zn2SnO4hydrothermalnanowires52.5306/100/0.1[57]
Zn2SnO4/NChydrothermal and heat treatmentspherical shaped particles-324.4/100[110]
Mn2SnO4/Ghydrothermal and heat treatmentnanocubes45.6106/1000/1[92]
SnO2/Mn2SnO4@Chydrothermal and heat treatmentnanoboxes65.9203/100/0.2[89]
Sn@ Mn2SnO4-NChydrothermal and heat treatmentcubic frame64185.8/7000/2[37]
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

Duan, Y.-K.; Li, Z.-W.; Zhang, S.-C.; Su, T.; Zhang, Z.-H.; Jiao, A.-J.; Fu, Z.-H. Stannate-Based Materials as Anodes in Lithium-Ion and Sodium-Ion Batteries: A Review. Molecules 2023, 28, 5037. https://doi.org/10.3390/molecules28135037

AMA Style

Duan Y-K, Li Z-W, Zhang S-C, Su T, Zhang Z-H, Jiao A-J, Fu Z-H. Stannate-Based Materials as Anodes in Lithium-Ion and Sodium-Ion Batteries: A Review. Molecules. 2023; 28(13):5037. https://doi.org/10.3390/molecules28135037

Chicago/Turabian Style

Duan, You-Kang, Zhi-Wei Li, Shi-Chun Zhang, Tong Su, Zhi-Hong Zhang, Ai-Jun Jiao, and Zhen-Hai Fu. 2023. "Stannate-Based Materials as Anodes in Lithium-Ion and Sodium-Ion Batteries: A Review" Molecules 28, no. 13: 5037. https://doi.org/10.3390/molecules28135037

Article Metrics

Back to TopTop