Next Article in Journal
Investigation of Ion Release and Antibacterial Properties of TiN-Cu-Nanocoated Nitinol Archwires
Next Article in Special Issue
Correction: Sai, R.; Abumousa, R.A. Impact of Iron Pyrite Nanoparticles Sizes in Photovoltaic Performance. Coatings 2023, 13, 167
Previous Article in Journal
Study on the Effects and Mechanism of Temperature and AO Flux Density in the AO Interaction with Upilex-S Using the ReaxFF Method
Previous Article in Special Issue
Annealing and Doping Effects on Transition Metal Dichalcogenides—Based Devices: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in CoSex and CoTex Anodes for Alkali-ion Batteries

College of Chemistry and Chemical Engineering, Guangzhou University, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(9), 1588; https://doi.org/10.3390/coatings13091588
Submission received: 1 August 2023 / Revised: 6 September 2023 / Accepted: 8 September 2023 / Published: 12 September 2023
(This article belongs to the Special Issue Advanced Materials for Energy Storage and Conversion)

Abstract

:
Transition metal selenides have narrow or zero band-gap characteristics and high theoretical specific capacity. Among them, cobalt selenide and cobalt telluride have some typical problems such as large volume changes, low conductivity, and poor structural stability, but they have become a research hotspot in the field of energy storage and conversion because of their high capacity and high designability. Some of the innovative synthesis, doping, and nanostructure design strategies for CoSex and CoTex, such as CoSe-InCo-InSe bimetallic bi-heterogeneous interfaces, CoTe anchoring MXenes, etc., show great promise. In this paper, the research progress on the multistep transformation mechanisms of CoSex and CoTex is summarized, along with advanced structural design and modification methods such as defect engineering and compositing with MXenes. It is hoped that this review will provide a glimpse into the development of CoSex and CoTex anodes for alkali-ion batteries.

1. Introduction

The demand for clean renewable energy sources such as wind and solar power has increased due to the energy problems and environmental pollution from traditional fossil energy. Lithium-ion batteries (LIBs) were born as high-performance energy storage devices to compensate for the inconvenience of storing clean energy [1,2]. Today, lithium resources are increasingly scarce; thus, sodium-ion batteries (SIBs) and potassium-ion batteries (PIBs) have become two of the hopes to solve the problem of resource shortage [3,4]. The prerequisite for further application of sodium/potassium-ion batteries is to prepare negative electrode materials with appropriate working voltage, excellent cycling performance, and high energy density. Anode materials can be divided into three types: intercalation type (e.g., carbon [5] and TiO2 [6]), alloy type (Si, Ge, Sb [7], etc.), and conversion type (Co3O4 [8], FexS [9], etc.). Although an intercalated anode has a stable structure and good electrochemical cycling stability, it still has the disadvantages of low energy density and limited specific capacity. In addition, the alloy-type anode has a high capacity, but the large volume change caused by the alloy reaction brings the disadvantage of poor stability. Compared with the former two, the conversion negative electrode has a reasonably high specific capacity and a small volume expansion, making it a potential choice for ion batteries’ anodes [10].
In recent years, transition metal chalcogenides (TMCs) have been widely studied as representative conversion-type anode materials due to their stable performance, high theoretical specific capacity, abundant natural resources, diverse types, and low cost. They hold great potential in meeting the demand for high-energy-density anodes in alkali-ion batteries [11]. Compared to transition metal sulfides, transition metal selenides exhibit higher volumetric capacity as anionic battery anodes. Additionally, the weaker M-Se bond compared to M-O and M-S bonds improves the reaction kinetics of transition metal selenides, resulting in higher energy density and electronic conductivity. Moreover, transition metal selenides have larger interlayer spacing, which helps alleviate volume expansion issues during cycling, especially for larger alkali metal ions such as sodium and potassium. Therefore, they possess advantages in electrochemical energy storage applications [12,13,14,15,16,17]. Compared to transition metal oxides and sulfides, transition metal tellurides possess lower electronegativity, the highest conductivity, and the highest volumetric capacity. They are typical layered materials with a large interlayer space, which is beneficial for rapid ion transport in electrodes. This results in excellent electrode wetting and ion diffusion kinetics. So far, many selenides (such as CoSex [18], ZnSex [19], MoSex [20], and FeSex [21]) and tellurides (e.g., NiTex [22], FeTex [23]) have been studied as ion battery anode materials.
Cobalt-based materials have become important materials in various fields (such as hydrodesulfurization and hydrodefluorination) due to their unique catalytic activity, electrochemical performance, and ferromagnetism [24,25,26]. Among them, LiCoO2 is one of the most successful commercial cathode materials for lithium-ion batteries (LIBs) [27]. Based on the success of LiCoO2, research on NaxCoO2 as a cathode material for sodium-ion batteries has also been widely conducted [28]. In terms of anode materials, cobalt has various oxidation states, and cobalt-based chalcogenides have diverse types and high theoretical sodium storage capacity, making them highly favored by researchers. Cobalt-based composite anodes, such as cobalt oxide [29], cobalt phosphide, and cobalt–aluminum alloys, have theoretical capacities of 700~1000 mAh g−1, which is 2~3 times that of graphite anodes. They have diverse crystal structures and low cost, receiving increasing attention in recent years. Among them, CoSex is widely used as an anode material for alkali-ion batteries due to its large capacity, strong conductivity, and weak electronegativity [18,30,31,32,33,34]. CoTex is considered to be a potential choice with excellent electrochemical performance in terms of metallicity, magnetism, electronics, and electrocatalysis [35]. In the past few years, extensive research has shown that cobalt selenides and tellurides can be used as potential anodes with high capacity or improved rate performance. However, their inherent volume changes as conversion anodes, relatively low electronic conductivity, and side reactions can lead to significant reversible capacity loss and low active material utilization. Recently, various energy storage batteries have been working on mitigating these issues and have achieved successes, such as forming hybrid/composite samples to enhance conductivity, optimizing electrode design and electrolyte content, and adjusting voltage windows [31].
At present, some progress has been made in the research of energy storage and conversion of transition metal selenides and tellurides—typically cobalt-based compounds [36]. However, it is necessary to study the application and reaction mechanisms of cobalt selenide and cobalt telluride in the three types of metal-ion battery in a timely and systematic manner. This paper reviews the recent progress in the preparation and application of CoSex/CoTex-based electrodes (Figure 1). The electrode design and reaction mechanism are introduced in detail. Finally, the challenges and opportunities in this field are presented (Figure 2). This review will significantly advance the research on innovative CoSex/CoTex design to improve its application in energy storage devices.

2. Electrochemical Reaction Mechanisms

As anode materials during charging and discharging processes, different metal-based compounds usually have different kinds of reactions, such as intercalation, conversion, and alloying. For example, intercalation–conversion–alloying reactions take place during the charging and discharging of Sn- and Sb-based compounds; Mo-, W-, Fe-, and Ni-based compounds mainly have intercalation and conversion reactions, while Nb-based compounds only have intercalation reactions [49]. The metal-based anode affects its reaction type. TMCs with conversion reactions can accommodate more alkali ions through the multi-electron transfer process and have a higher capacity. In addition, the combination with chalcogenide elements is conducive to the conversion reaction during the charge–discharge process. Due to the weak TM–B (B = Se and Te) bonds, TMBs have a high theoretical capacity, as well as excellent electrical conductivity and convenient ion transport channels. CoxBy has been used as a potential anode in alkali-ion batteries. The storage mechanism between CoxBy in LIBs/SIBs/PIBs is the conversion reaction that eventually generates Co and AxB (A = Li, Na, K). In particular, a few intermediate reactions may occur during the CoxBy discharge processes, as shown in the following equations:
C o x B y + A + + e A C o x B y
C o x B y + A + + e C o x B y 1 + A B
Deciphering the electrochemical reaction mechanism of CoSex and CoTex is significant to enhance their application in batteries. The latest progress on the reaction mechanisms of cobalt selenide and cobalt telluride for the anode material of LIBs, SIBs, and PIBs is summarized below.

2.1. CoSex

Many types of research on the storage mechanisms of CoSex anodes have been conducted. Among them, the storage path of CoSex in PIBs is different from that in SIBs and LIBs. Yu et al. [34] studied the potassium-ion storage mechanism of octahedral CoSe2 nanotubes threaded with N-doped carbon through first-principles calculation (Figure 3a–f). The calculation showed that the formation energies of CoKxSex were all positive, indicating that the insertion of K+ into CoSe2 is difficult to achieve, and it is more inclined to replace Co to generate K2Se.The reason for this is that potassium atoms are too large to penetrate the lattice of CoSe2, whereas the non-stoichiometric selenide CoSe1−x (Co3Se4) has a wider lattice than CoSe2, achieving K+ embedding. The actual reaction process can be explained as follows:
C o S e 2 + 4 3 K + + 4 3 e 1 3 C o 3 K S e 4
C o 3 K S e 4 + K + + e 3 C o S e + K 2 S e
C o S e + 2 K + + 2 e C o + K 2 S e
In contrast to PIBs, Jiang et al. [50] found the appearance of three new peaks in cyclic voltammetry (CV) curves (Figure 3h), indicating that CoSe2 turns into NaxCoSe2, CoSe, and Co. The oxidation peaks at 1.53 V and 1.89 V represent the intermediate product NaxCoSe2 and CoSe2, which is the fully charged product during the charge process, respectively. The locations of these peaks are highly consistent throughout the five cycles of reduction and oxidation, together with the corresponding position of the charging and discharging platform (Figure 3g), testifying to the reversibility of the charge–discharge process. The actual reaction mechanism can be summarized as follows:
C o S e 2 + x N a + + x e N a x C o S e 2
N a x C o S e 2 + ( 2 x ) N a + + ( 2 x ) e C o S e + N a 2 S e
C o S e + 2 N a + + 2 e N a 2 S e + C o
2 N a 2 S e + C o C o S e 2 + 4 N a + + 4 e
In LIBs, Jiang et al. [51] found three prominent peaks of CoSe/G at 1.42, 1.27, and 0.63 V in the first cathodic scan through the first five CV curves in LIBs (Figure 3i). Li+ inserted into the CoSe showed a prominent peak at 1.42 V, corresponding to the formation of LixCoSe. The solid–electrolyte interphase (SEI) formed at 1.27 V. The wide peak at 0.63 V can be ascribed to the conversion reaction from LixCoSe to Li2Se and Co. The anode peaks at 1.28 and 2.1 V signify the oxidation reactions that convert Li2Se and Co into CoSe. The reaction process of CoSe anode in LIBs can be expressed as follows:
C o S e 2 + x L i + + x e L i x C o S e
L i 2 C o S e + 2 x L i + + 2 e C o + L i 2 S e

2.2. CoTex

As an emerging potential anode, the storage mechanism of CoTex is being extensively studied. Xu et al. [38] performed high-resolution transmission electron microscopy (HRTEM), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS) analyses on CoTe2 nanowires with a Te vacancy (o-P-CoTe2/MXene) in different discharge and charge states to study the potassium storage behavior. When the half-cell discharged to 1.0 V, the HRTEM (Figure 3j) showed that the lattice spacing of the (111) plane expanded significantly to 0.296 nm compared with the initial state (Figure 3k, 0.282 nm), which proves the existence of K+ intercalation above 1.0 V. At 0.5 V, peaks of K2Te3 and Co were observed in the ex situ XRD spectra (Figure 3l). This observation confirmed the conversion reaction from CoTe2 to Co and K2Te3. After further discharging to 0.01 V, a peak of K2Te was observed, indicating that K2Te3 was further converted into the final reduction product (K2Te). When the half-cell was charged to 1.4 V, K2Te converted to K2Te3 and Co. When the anode was further charged to 2.6 V, the peak of o-P-CoTe2/MXene reappeared. Ex situ XPS (Figure 3m) spectra further confirmed that the peak of Co 2p3/2 continued to shift from 781.4 eV to 778.4 eV during complete discharge, corresponding to the formation of Co. Therefore, the o-P-CoTe2/MXene anode conversion mechanism in PIBs can be described by the following reactions:
C o T e 2 + x K + + x e K x C o T e 2
3 K x C o T e 2 + 4 3 x K + + 4 3 x e 3 C o + 2 K 2 T e 3
K 2 T e 3 + 4 K + + 4 e 3 K 2 T e
Recently, Li et al. [41] published a more detailed supplement on the storage mechanism of CoTex in PIBs. As shown in the in situ XRD pattern (Figure 4a), after discharging to 1.21 V, the peak of CoTe2 shifted to a lower angle, indicating that K+ was embedded into the CoTe2 lattice and KxCoTe2 formed. Upon further discharging to 0.01 V, the peak of CoTe2 became weaker, corresponding to the conversion of CoTe2 to Co and KxTey. When charged to 3 V, the peak of CoTe2 reappeared. In situ XPS analysis (Figure 4b) showed that the metallic Co phase appeared in the high-resolution Co 2p spectrum when the cell was discharged to 0.8 V. In the Te 3d spectrum, K5Te3 appeared, indicating that KxCoTe2 was converted to K5Te3 at 0.8 V. When discharging to 0.4 V, the cobalt peak strengthened and a new peak of K2Te appeared in the Te 3d spectrum. Upon further discharging to 0.1 V, K5Te3 was almost transformed into K2Te. The electrochemical reaction mechanism was further studied by transmission electron microscope (TEM), energy-dispersive X-ray spectroscopy (EDS), and HRTEM (Figure 4c–h). The TEM and EDS patterns proved that the structure remains stable and the elements are evenly distributed after the cycle. The discharge HRTEM image shows the (100) crystal plane of Co, the (251) crystal face of K5Te3, and the (311) and (200) crystal faces of K2Te (Figure 4d). In addition, upon full charge, the (011) plane of CoTe2 (Figure 4g) appears, which is consistent with the in situ XRD results. In summary, the above results show that KxCoTe2 converted totally to K5Te3, and then K5Te3 converted to K2Te during the discharge process, and the authors clearly express the evolution of polytellurides in Figure 4i. Therefore, the relevant reversible reactions are summarized below:
C o T e 2 + x K + + x e K x C o T e 2
3 K x C o T e 2 + 10 3 x K + + 10 3 x e 3 C o + 2 K 5 T e 3
K 5 T e 3 + K + + e 3 K 2 T e
Zhang et al. [42] clarified the sodium-ion storage mechanism of CoTe2 by selected-area electron diffraction (SAED, Figure 4j–k), HRTEM (Figure 4l–m), and ex situ XRD (Figure 4n). At the end of discharge, the peaks of Co and Na2Te could be clearly observed from the ex situ XRD pattern. Moreover, the signals of Co and Na2Te could be detected in the SAED pattern. The crystal lattice of Na2Te and Co also appeared in HRTEM. After charging to 2.8 V, the CoTe2 phase peak could be collected again. A clear CoTe2 ring was observed in the SAED diagram. At the same time, Zhang analyzed the storage mechanism of Na ions from the CV curve (Figure 4o). A reduction peak formed at about 0.83 V, because CoTe2 formed Co nanocrystals and Na2Te with Na ions. The oxidation peak at about 1.65 V can be ascribed to the reformation of CoTe2. It can be summarized as follows:
C o T e 2 + 4 N a + + 4 e C o + 2 N a 2 T e
C o + 2 N a 2 T e C o T e 2 + 4 N a + + 4 e
Ganesan et al. [52] studied the deep material changes of CoTe2 cycling in LIBs by ex situ XRD (Figure 4p) and extended X-ray absorption fine structure (EXAFS, Co-K-Edge, Figure 4q). In the XRD pattern, when discharging to 0.9 V, it can be seen that the peak of CoTe2 changed to the peak of Li2Te, and CoTe2 also changed to Co in EXAFS, which means that the conversion reaction in the discharge process is that CoTe2 changes to Co and Li2Te. Finally, although no recovery of CoTe2 was observed in the XRD pattern at a full charge of 3.0 V, the main CoTe2 EXAFS peak reappeared, confirming the reversibility of the material. Based on the experimental results, the LIB conversion mechanism for CoTe2 is proposed as follows:
C o T e 2 + x L i + + x e C o + L i 2 T e
It is known that many works have fully studied the reaction mechanisms of cobalt selenide and cobalt telluride in LIBs, SIBs, and PIBs. Many in situ and ex situ analytical characterization methods have emerged. In situ XRD, XPS, and ex situ TEM, HRTEM, etc., can well demonstrate the mechanisms of CoSex and CoTex. However, due to emerging anode design methods such as doping and heterojunction, heteroatoms and conductive carbon materials will bring various synergistic effects to cobalt selenide and cobalt telluride. At present, the mechanistic analysis of cobalt selenide and cobalt telluride anodes mainly focuses on the reaction between alkali ions and CoSex or CoTex. In fact, the analysis of the synergistic mechanism between substances in the anode requires more in-depth research, and at the same time, innovative research methods need to be developed.

3. Electrode Design

As conversion-type anodes, CoSex and CoTex can bring enthralling high weight/volume capacity due to multiple electron transfer processes. However, due to the low conductivity of the converted anode materials and the large volume change in the process of intercalation/deintercalation, their practical application has been hindered. Therefore, effective strategies to solve these problems have been explored through carbon recombination [53], core–shell structures [54], defect engineering [55], etc.
Two approaches used to modify cobalt selenides and cobalt tellurides are the introduction of defects (e.g., point defects and heterogeneous interfaces) and structural design. Point defects like vacancies and dopings can interfere with surrounding atoms, resulting in lattice distortion, forming a good adjusted electronic structure with improved conductivity and diffusion coefficients [56]. The inorganic heterostructure of the anode builds an interfacial electric field as a facilitative transport channel, resulting in rapid Li/Na/K ion reaction kinetics. Considering the excellent prospects of interface engineering, Dong et al. [57] realized a “structural function motifs” design in a ZnO/ZnS anode. The ZnO/ZnS heterostructure regulates the state around the Fermi level, narrowing the band gap and, thus, improving the electronic conductivity. The introduction of defect engineering can create synergistic effects that alter the inherent properties of materials, unlike structural design and composite strategies [58,59]. In addition, Co 3d electrons with the low spintronic structure t2g6eg1 are known to produce a Jahn–Teller effect in CoSe2 [60]. Proper doping of metal ions can alleviate Jahn–Teller distortion to a certain extent [61,62] and promote a more stable electrode structure and better cycling stability. Defect engineering is widely used in CoSe2 and has also been reported in recently emerging CoTe2 materials.
In addition to defects, structural designs such as core–shell structures or CoSex and CoTex composite carbon materials [42] can take advantage of carbon’s high conductivity and deformation resistance, not only reducing the risk of electrode crushing but also improving the rate performance. The latest reports of introducing defects or innovative structural designs to CoSex and CoTex are summarized below.

3.1. Introducing Defects

The modification methods of defect engineering include entry-point defects and surface defects. In recent studies, the introduction of point defects has been divided into metallic elements and non-metallic elements. Secondly, surface defects generally introduce heterogeneous structures. Recent advances in the combination of the above engineering with cobalt selenide and cobalt telluride are summarized below.

3.1.1. Doping with Metal Elements

Due to the large volume variation during circulation, the structural solidity and circulation function of metal selenides are weak, which limits the feasibility of their application. To overcome the above problems, some studies have prepared polyselenide metal nanostructures by adding inert elements [63]. The activity of the transition-metal-based materials can further be improved through coupling with other metals and non-metals [64]. The reason for the excellent doping performance of metal atoms is that the introduction of two or more metal atoms into a mixed-metal selenide results in the formation of more multicomponent selenides. Some studies have shown that this result can significantly improve the electrical conductivity of discharge products and further provide a buffer matrix to adapt to the expansion stress [65]. They benefit from strain relief and spectacular pseudocapacitance effects, thus reducing the energy barrier for ionic diffusion and showing considerable improvement in electrochemical performance.
For the voltage window of CoSe2, previous studies have usually shortened it to 0.5~3.0 V to stabilize the reversible capacity of long cycles [66,67]. Ji et al. [50] doped a variety of metal ions such as Ni2+ and Cu2+ into nitrogen-doped carbon-shell-coated CoSe2 particles to solve the problem of capacity loss caused by shortening the window. The doped product CoM–Se2@NC, obtained by carbonizing a bimetallic zeolite imidazole skeleton (ZIFs) precursor, had more outstanding capacity than CoSe2@NC at 0.01~3.0 V. It was shown that the long cycle stability of CoSe2 can be optimized effectively and the inefficient cycle caused by window adjustment can be eliminated under the synergistic action of the participation of another metal ion and the N-doped carbon shell.
Lu et al. [68] introduced Mo into a CoSe2 self-supported nanosheet array (Mo-CoSe2@NC) and investigated the performance of Mo in SIBs. In the XPS diagram (Figure 5a–b), the 3d spectrum of Mo consisted of peaks at 232.3 eV and 235.4 eV, associated with the characteristic spin-orbital bistates of Mo 3d5/2 and 3d3/2, respectively, which implies that Mo6+ ions enter the CoSe2 lattice [69]. The embedded Mo defect in the CoSe2 lattice not only produced more redox sites, but also induced the generation of mixed phases of o-CoSe2 and c-CoSe2, as evidenced by the coexistence of two hybrid phases of CoSe2 in the SAED diagram (Figure 5d).
Wang [70] formed FeCo-Se in situ on a carbon cloth, and this multilayer array was able to provide sufficient active surfaces and shuttle channels for sodium ion exchange, thereby achieving excellent electrochemical kinetics. XRD (Figure 5g) proved that after the Co-Se was doped with Fe, impurities such as cobalt oxide, iron oxide, iron selenide, and FeCo alloy were not produced, implying that Fe ions only enter the Co-Se lattice. Notably, when more than 50% iron salt was added to the Co-Se material, the morphology of Fe-Se changed from a layered nanosheet array into pyramidal particles (Figure 5e–f), with a decline in electrochemical performance.

3.1.2. Doping with Non-Metallic Elements

For cobalt telluride, Hu et al. [71] prepared Cu-doped cobalt telluride hollow carbon nanocells (Cu-Co1−xTe@NC HNBs, Figure 5j–l) by chemically etching CuCo-ZIF nanocells and tellurizing them (Figure 5i), The copper–cobalt tellurium was homogeneously embedded in the nanobox (Figure 5p–r), and Co vacancies were constructed by copper induction (Figure 5n). At the same time, the interaction in the heterogeneous interface between nitrogen-doped carbon and cobalt telluride (Figure 5m) triggered the transfer of the p-band to interface charge transfer, so that the composite exhibited faster ion and electron diffusion kinetics than HNB electrodes, contributing to improved storage performance of lithium-ion batteries (Figure 5h).
Doping with non-metallic elements, especially chalcogenide elements, can not only produce polysulfides with alkali ions to improve the storage capacity, but also induce vacancy formation like metal elements to form a built-in electric field. Yu [34] reported an octahedral CoSe2 sawtooth chain array threaded with N-doped carbon nanotubes (NCNTs) as a high-performance PIB anode. Highly conductive NCNTs form a flexible conductive network that greatly accelerates electron transfer and improves rate performance. Nitrogen-doped NCNTs also provide additional capacity to CS-NCNTs, constituting a versatile composite carbon material.
Wang [33] prepared Co0.85Se1−xSx nanoparticles by using S, which is of the same family as Se, as a point doping tool. The sample was formalized in a metal–organic framework, carbonized in graphene and, finally, cured in situ. Density functional theory (DFT) proved that S doping enhanced the adsorption energy of Co to alkali metal ions (Figure 6a–h). The hollow polyhedral skeleton of the double-carbon shell (Co0.85Se1−xSx @C/G) with a stable structure and S vacancies reduced the volume change, increased the number of electrochemical active sites, and improved the alkali-ion storage performance.
Moreover, given the high conversion capacity of phosphorus, Ye et al. [73] reported in another work that P was successfully doped into CoSe2. It was found through P 2p spectroscopy that additional Co-P and P-Se bonds could be generated by P doping, which could enhance the structural stability of CoSe2. Xu [38] also used P to induce Te vacancies in o-CoTe2 nanowires. Compared with h-CoTe2, the Co 2p and Te 3d peaks of o-CoTe2 showed significantly higher binding energy on the XPS images after P doping. EPR showed that the characteristic Te vacancy signal of o-P-CoTe2/MXene was stronger and wider, indicating that the vacancy content increased significantly. Brunauer–Emmett–Teller (BET) theory showed that P doping increased the specific surface area. These results show that Te vacancy defects and P doping not only provide the active sites but also jointly enhance the structural stability of CoTe2.

3.1.3. Heterojunctions

Heterostructure refers to the geometric structure and connecting interface formed by the combination of two or more materials through physical or chemical methods [74,75]. In the heterojunction, electrons will transition from a high Fermi level to a low Fermi level, resulting in potential difference due to interfacial polarization at the heterojunction [76] and the internal electric field. The existence of the internal electric field effect is helpful to accelerate the diffusion of metal ions and increase the adsorption energy of ions [77,78]. In addition, lattice distortion and charge redistribution taking place at heterogeneous boundaries can not only improve charge-transfer efficiency and conductivity, but also provide additional active sites for reversible redox reactions [79,80]. Therefore, due to the active role of the heterogeneous interface, heterogeneous anodes can obtain a stable nanostructure and excellent storage performance [81]. Therefore, in view of problems such as the low conductivity and large volume expansion of chalcogen-based metal materials, it is considered to be an effective strategy to construct high-quality heterogeneous structural interfaces to improve their storage performance [82].
Previously, Fang [72] reported a bimetallic selenide heterogeneous structure (CoSe2/ZnSe@C) for SIBs. The phase-boundary charge redistribution of Co/Zn was studied by theoretical calculation. The Na+ adsorption energy showed that the ZnSe side had a phase boundary with high electron density, which was more conducive to the adsorption of Na+ ions, accelerating the ion movement and charge conduction, and achieving high reaction kinetics and high electrical conductivity (Figure 6i–k). In addition, the multistep conversion reaction in the Co/Zn heterostructure can effectively relieve the stress of Na+ insertion. CoSe2/ZnSe@C also showed good OER (oxygen evolution reaction) activity, providing additional evidence that the Co/Zn heterogeneous interface accelerated the redox reaction.
Xiao [83] recently also reported a similar application of heteroselenides in SIBs. He developed ZnSe/CoSe2-CN heterostructures with Se vacancies via controllable in situ selenization. HETEM images showed a clear heterogeneous interface between ZnSe and CoSe2, which can create electric fields to promote the movement of ions and electrons. Electron paramagnetic resonance (EPR) spectra showed an obvious signal caused by the selenium vacancy trapping electrons at g = 2.003, confirming the presence of the Se vacancy in the sample. The more induced active sites and the enhanced electronic conductivity introduced by the Se vacancy can improve the electrochemical properties of the materials.
Analogously, Zhang et al. [40] reported a template-assisted strategy to achieve in situ formation of a bimetallic tellurium heterostructure (ZnTe/CoTe2@NC) on N-doped carbon shells. The results showed that both the electron-rich Te sites and the internal electric fields generated by electron transfer from ZnTe to CoTe2 in Te heterojunctions provided rich cation-adsorption sites and promoted interfacial electron transport during the potassic/de-potassic process. Notably, as Zn led to ZIF-8 crystal formation, with the introduction of Zn2+ species into Co-ZIF-67, Zn2+ ions acted like an obstacle to Co-ZIF-67 crystals’ nucleation and growth. Concretely, Zn delays the Co-ZIF-67 nucleation dynamics, thus resulting in relatively small doping particle sizes. The growth of ZIF-67 dominated by Co leads to the production of larger crystals in the reaction–diffusion process [84,85]. The finer ZnTe/CoTe2 nanoparticles prepared by the heterojunction method showed higher structural stability. Ex situ SEM, HRTEM, and EDS were performed on the cycled ZnTe/CoTe2@NC (Figure 6l–s). The ZnTe and CoTe2 nanoparticles kept their original appearance, with no obvious particle aggregation. In addition, the TDOS results presented in Figure 6t–v highlight the discontinuous band gap of Fermi levels and confirm the semiconductor properties. In contrast, ZnTe/CoTe2 has an obvious hybridization zone in the conduction band, resulting in significantly higher TDOS of ZnTe/CoTe2 than that of both ZnTe and CoTe2. This indicates that the electronic conductivity of ZnTe/CoTe2 is significantly improved.
In addition to zinc–cobalt heterojunctions, He et al. [86] chose Ni to produce coral-like NixCo1−xSe2 for the first time using a layered structure. The introduction of Ni formed a hierarchical structure to prevent structural fragmentation, accelerated the electron transmission, and shortened the Na+ diffusion path. Zhu et al. [87] used the carbonization and subsequent selenization process design to encapsulate Ni3Se4@CoSe2@C/CNTs in a core–shell three-dimensional interpenetrating two-carbon frame. Ni3Se4@CoSe2 was evenly dispersed into a three-dimensional carbon skeleton structure/carbon nanotube network, greatly enhancing the conductivity and further achieving fast Na+ diffusion.

3.2. Structural Design

Nanomaterials, especially spherical and small nanocrystals, have their isotopic physical properties, large active surface, and high packing density [88,89,90]. More active sites can be exposed by delicate structures (e.g., sea urchin CoSe2 [88], CoSe microspheres [91]). By using carbon coatings, CoSex and CoTex can be well maintained in nanoform and improved in terms of cycling stability by using constraint effects (such as CoSe-C@C [92] and o/h-CoTe2 [93]). Combination with structural carbon can produce interfacial interactions, which can effectively improve diffusion kinetics, electrical conductivity, and structural compatibility, thus achieving excellent electrochemical performance [10,94]. For example, three-dimensional conductive carbon network structures with large channels can provide diffusion paths, improve the migration rate of ions in the battery material, and reduce the limitation of poor conductivity of CoSex and CoTex [66]. Metal–organic frameworks (MOFs) possess exceptionally high surface area and porosity, allowing for the adsorption and storage of gases, liquids, and even small molecules. This property makes them promising for applications such as energy storage, separation, and catalysis. Moreover, by optimizing the composite structure, the electrochemical properties of the composites can be further improved, and the volume strain of the active nanomaterials in the carbon matrix can be reduced. Therefore, carbon materials such as reduced graphene oxide (rGO) and MXenes are not only electrically conductive but also have good ductility, which can greatly enhance the structural compatibility of CoSex and CoTex anode materials.
However, MOFs, MXenes, and other carbon-based materials have their limitations. For example, MOFs are prone to acid–base interference and structural collapse. MXene etching processes can be complex and hazardous. Additionally, low energy density is a common drawback of these materials. Therefore, recently, researchers have been consolidating the carbon-based structure of transition-metal-based MOFs through annealing and synthesizing high-performance ion battery anodes using vapor-phase methods [95]. High-capacity ion battery anode materials can be obtained through green etching of MXenes using only grinding and annealing via molten salt methods [96]. These structural design approaches are frequently employed in the design of cobalt selenide and cobalt telluride anodes. The latest progress in combining the abovementioned methods with cobalt selenide and cobalt telluride for alkali-metal-ion battery cathodes is summarized below.

3.2.1. Freestanding Electrodes

Common electrodes are made of an active material composited with a binder and conductive carbon on the current collector, and the presence of the binder and collector fluid especially affects the energy density. Therefore, the development of freestanding electrode structures to increase the proportion of active materials and eliminate the negative effects of the binder is a promising direction [97].
Electrospinning is used to produce individual electrode materials, allowing the active material to be embedded in a network of carbon fibers. The carbonized material can be cut so as to be used as an electrode, avoiding mechanical losses caused by mixing adhesives with conductive materials and coatings [98,99]. Self-supported electrodes can alleviate electrode pulverization and construct interwoven electron/ion transport networks, thus improving electrochemical energy storage performance.
Zhan et al. [47] innovatively prepared a self-supported cobalt telluride electrode for SIBs. They synthesized CoTe2@NMCNFs via electrospinning and subsequent in situ tellurization (Figure 7a–f). Material characterizations revealed NMCNFs with N-doped active sites and porous carbon networks with high specific surface area. These positive effects promote electrolyte penetration and mitigate the aggregation effect in the CoTe2 nanoparticles by reducing the volume changes during the cycle (Figure 7g–j). The long cycle of CoTe2@NMCNFs shows a high reversible Na+ storage capacity, long-term cycle stability, and high rate capacity. In general, the overall conductivity of the electrode is greatly improved, and Na+ diffusion is promoted, thanks to the strong three-dimensional crosslinked composite fiber and the absence of voluminous inactive components.
Meanwhile, Park et al. [100] prepared cobalt selenite (p-CoSeO3-CNF) embedded in carbon composite nanofibers by electrospinning. Among them, the heat-treated nanofibers containing cobalt and selenium formed CoSe2 nanocluster intermediates. Then, through the final oxidation step, CoSe2 was converted to amorphous CoSeO3. Part of the carbon fiber was changed into CO2 gas during the oxidation process, leaving pores and forming mesopores in the nanofibers. The carbon and pores limited the volume change of CoSeO3 during repeated sodification/desodification processes, allowing the electrolyte to permeate easily and reducing the diffusion path of sodium ions. The large network carbon structure accelerated electron transport, and the high performance of p-CoSeO3-CNF was demonstrated. Electrospinning has also been reported for SIB and LIB anode materials other than CoSe2 [32,101].

3.2.2. Carbon-Cloth-Based Electrodes

Commercially available carbon cloth (CC) material, consisting of hundreds of fibers woven tightly together, ensures that the fibers are interconnected and can be rolled into different shapes without compromising their integrity [102]. The stable cycling performance of CC-based electrodes makes them a competitive option for industrial applications of metal-ion batteries [103].
According to the different raw materials, carbon fiber is mainly divided into three substrates: asphalt-based carbon fiber, PAN-based carbon fiber, and viscose-based carbon fiber. PAN-based carbon fiber is rich in raw materials and superior to the other two types. Therefore, PAN-based carbon fiber is the most widely used, accounting for more than 90% [104]. The difference between the growth of active substances in PAN-based carbon cloth and the embedding of active substances in electrospinning is that carbon cloth generally penetrates active substances into existing carbon substrates via infiltration methods, hydrothermal methods, and other methods. However, electrospinning uses PAN, solvent, and metal salt to dissolve directly into the active material. Both have their advantages, such as the toughness of carbon cloth and the advantage of electrospinning for in situ doping treatments, which can enhance the synergistic effects of carbon nanowires and active substances [105].
To date, carbon cloth combined with active substances has been used in fuel cells [106], catalysis [107], lithium–sulfur batteries [108,109], supercapacitors [102], etc. There have also been some studies on carbon cloth combined with metallic chalcogenide compounds as negative electrodes for LIBs [110]. For cobalt selenide, Wang et al. [70] grew FeCoSe on carbon cloth, and Lu et al. [48] synthesized Mo-doped CoSe2 nanosheets, proving that the carbon cloth and cobalt selenide have a good synergistic effect and are more conducive to the formation of bimetallic synergies (Figure 8). However, cobalt telluride combined with carbon cloth as a negative electrode for LIBs, SIBs, and PIBs has rarely been reported.

3.2.3. MOF-Based Electrodes

MOFs with a high specific surface area and adjustable pore structure are widely used in the fields of calcification and energy conversion due to their stable structure, large specific surface area, and diverse functions [95,111,112]. Under appropriate conditions, MOFs can be used as precursors and/or templates to derive metal chalcogenides with regular topography and excellent electrochemical properties (Figure 9) [113,114,115]. Notably, MOFs have proven to be ideal templates for the preparation of transition metal selenides and tellurides. In recent reports, it has been shown that transition metal selenide and telluride nanocrystals can be embedded 1–3D porous carbon substrates to form efficient conductive networks [116]. In addition, the relatively high surface area and large pores of the high-dimensional conductive substrate enhance the wettability of the electrolyte to the material and promote ion diffusion, thereby improving the storage performance [117]. Therefore, MOFs as a template for the preparation of nanoscale electroactive materials and porous carbon matrix hybrid structures have broad application prospects.
More recently, Xiao [39] prepared hollow In2Se3/CoSe2 nanorods by growing Co-ZIF-67 on the surface of In-MIL-68 and via in situ selenization. During selenization, selenium atoms penetrate from the surface of the nanorod to the inside, while indium atoms diffuse outward in opposite directions and at different speeds, creating the Hilken–Dahl effect and forming hollow structures [118]. At the same time, the mass ratio of ZIF-67 to MIL-68 grown on the surface was studied by changing the concentration of cobalt nitrate. The results showed that the ZIF-67 nanoparticles became smaller and could be uniformly deposited on the surface of MIL-68 with the decrease in the concentration of cobalt nitrate in solution (Figure 10).
Wang et al. [119] prepared cobalt selenide (CoSe2-CNS) anchored on carbon nanosheets by selenizing cobalt embedded in 1D-MOF carbon nanosheets (Co-CNS). CoSe2 grown in situ on carbon nanosheets showed close contact between CoSe2 and carbon. At the same time, CoSe2-CNS also has typical two-dimensional structural characteristics, which are conducive to electrolyte penetration and ion/electron transport. In addition, the carbon matrix can alleviate the volume expansion of CoSe2 during the desodification process. As the result of these advantages, CoSe2-CNS provides a stable cycle capacity of 468 mA h g−1, together with a high rate capability of 352 mA h g−1 at 10,000 mA g−1.

3.2.4. MXene-Based Electrodes

MXenes, which include transition metal carbides or carbon nitride, have proven to be novel materials with great potential for energy storage [120,121]. MXenes have high electrical conductivity, large interlayer spacing, and excellent mechanical properties, enabling them to act as a fluid collector and buffer deformation at the same time [122]. Their electrical and electronic properties can be customized through modifications to the MXenes’ functional groups, solid solution structures, or stoichiometry to meet the needs of different scenarios [123]. As a result, titanium carbide has demonstrated outstanding catalytic activity and selectivity as a co-catalyst in various catalytic reactions, finding wide applications in energy conversion, environmental protection, and organic synthesis [124].
Xu [38] used P-induced o-CoTe2 nanowires to anchor MXene nanosheets (o-P-CoTe2/MXene). The elastic MXene formed a tight interfacial interaction with P-CoTe2 (Figure 11). DFT calculations further showed that the new superstructure had higher electronic conductivity, enhanced the adsorption capacity of K+ ions, and reduced the energy barrier to ion diffusion. Meanwhile, the entire battery was cycled through with negligible capacity loss even after bending. This proves that the combination with MXenes can greatly improve the stability of the material. Hong [91] found a strong Co-O-Ti covalent interaction in CoSe2@MXene. It assembled the inner hollow CoSe2 microsphere and the outer MXene via electrostatic self-assembly. This covalent bond facilitates electron/ion transport and enhances the structural durability of the CoSe2@MXene hybrid, thus improving the rate performance and cyclic stability. It has been reported that the polar surface of MXenes can inhibit the polysulfides’ shuttle effect [125,126]. Wang [127] was therefore inspired to grow CoSe2 nanorods in situ on the surface of an MXene via a simple one-step hydrothermal method. Importantly, the stability of the CoSe2/MXene cycle in this study was better than that of the pure CoSe2 cycle, demonstrating that Ti3C2Tx can inhibit the undervoltage failure caused by NaxSe instability.

3.3. Summary of Electrode Design

Cobalt selenide, as a high-capacity conversion anode, has been widely studied in alkali-metal-ion batteries. The new electrode design methods, such as doping with composite conductive materials (e.g., MXenes), can show excellent high capacity and high stability. However, the wide application and innovation of cobalt selenide in potassium-ion batteries still need to be improved. Cobalt telluride, as a compound of the same series, is still in its infancy in the design of advanced anode materials. Some novel designs of cobalt telluride show excellent electrochemical performance, which indicates the great potential of cobalt telluride. However, innovative designs of cobalt selenide are still to be developed; for example, co-doping with multiple non-metallic elements, carbon cloth growth, and electrospinning in situ growth methods have not been tried. At the same time, the electrochemical reaction mechanism of various synergistic effects is still unclear, and products with ultra-long periods of stability are still to be developed.

4. Electrode Synthesis

In addition to the elemental composition of the material, the morphology, crystal phase, and structure of the material also have a great influence on its properties. The microstructure of materials, such as their size and morphology, will affect their electron energy density, as well as influencing their optical and electrical properties [90,107,108,128]. Therefore, choosing appropriate preparation methods to synthesize materials with special structures is the key to improving the properties of the materials. Conventional preparation methods of cobalt-based anode materials include, but are not limited to, hydrothermal/solvothermal methods [32,101,113,116], heat treatment methods [52,93,129], sol–gel methods [130], vapor deposition methods [34], and electrospinning [100]. At the same time, the usual method for preparing selenides and tellurides is the gas-phase method [34].
Hydrothermal reaction is a typical synthesis route. Under certain temperature and high-pressure conditions, cobalt-based materials with high phase purity and crystallinity can be synthesized. For example, Lei et al. [131] synthesized uniform CoTe2 nanorods by the hydrothermal method, without any surfactant or mineralizer for 20 h, using ascorbic acid as a reducing agent. The study found that simply adjusting the concentration of sodium hydroxide can produce different forms, including short rod-shaped bodies with many spines and flower-like hierarchies (Figure 12a–f). Hydrothermal methods are widely used in the synthesis of various materials, due to their advantages of fewer experimental steps, changeable morphology, and simple operation. However, hydrothermal methods also have some shortcomings, such as too many uncontrollable factors, easily leading to an uneven appearance, and a generally long hydrothermal reaction time. As a high-temperature treatment, the vapor-phase method can be used alone or in combination with a template method for intermediate processes such as selenides and tellurides. Cobalt-based electrode materials with an abundant pore structure and high electrochemical performance can be obtained by heat treatment. Heat treatment is almost a necessary experimental method for MOF-series materials, but this method has high energy consumption, and for MOF materials the sintering temperature is too high, which can easily lead to the collapse of the structure. The above methods are conducive to the formation of high-functional structures for CoSex/CoTex, but high-temperature pyrolysis is always necessary in the process of in situ carbonization. To achieve energy savings and high yields, it is urgent to explore more effective modification or combination methods such as magnetron sputtering [132], microwave [133], and molten salt etching [96,134] methods. The next section summarizes the current CoSex/CoTex anode synthesis methods for LIBs, SIBs, and PIBs.

5. Performance Summary

This section focuses on the latest advances in the synthesis of various cobalt-based selenides and tellurides for LIBs, SIBs, and PIBs, and it summarizes the key properties and synthesis methods of various heterostructured anodes, as shown in the following Table 1 and Table 2.
In general, cobalt selenide has been shown to be a high-energy anode for LIBs, SIBs, and PIBs, which greatly improves their electrochemical performance through independent electrodes without binders and with innovative heterojunction design. The reaction mechanism has also been studied by theoretical calculation and in situ techniques. As a potential electrode with high area capacity, cobalt telluride has many surprising applications in PIBs, and some electrode design methods that have proven excellent for cobalt selenide, such as heterojunctions and composite MXenes, have also been efficiently practiced for CoTe. However, compared with the metal oxides, sulfides, or selenides, the existing modification strategies are still insufficient. In addition, electrochemical performance can be enhanced by electrode synthesis and preparation. For example, studies on the compatibility of cobalt telluride with electrolytes and binders are still lacking. For practical applications, in future, more investigations and full battery evaluation should be emphasized.

6. Summary and Expectations

Cobalt selenide and cobalt telluride show promising applications in LIBs, SIBs, and PIBs due to their unique crystal structures, diverse structural designs, and high theoretical capacities. In this review, the progress of CoSe and CoTe was comprehensively summarized, not only focusing on the running mechanism and electrochemical performance, but also putting forward challenges to the synthesis strategies. The mechanism of Co-based chalcogenides in energy storage reactions is complex because it may have mechanisms for multilevel transformation. Often, conversion reactions are indispensable, which can lead to high weight capacities. However, disadvantages remain, such as large volume expansion and sluggish kinetics. In response to this series of problems, effective improvement strategies have been reported, including the design of nanocomposite carbon materials and the introduction of defects. Firstly, 0–3D conductive carbon nanocomposites can improve conductivity, create abundant active sites, and enhance electrolyte penetration. In addition, the introduced vacancies and heterogeneous interfaces can accelerate the conduction of ions and electrons. Table 1 and Table 2 summarize the latest modified properties and synthesis methods, which can provide a reference point for innovative strategies for cobalt selenide and cobalt telluride materials. Nevertheless, the application of cobalt selenide and cobalt telluride anode materials in alkali-metal-ion batteries still requires extensive efforts. Therefore, we propose some prospective directions to promote further research to enable cobalt selenide and cobalt telluride to achieve higher energy storage efficiency.

6.1. Replenishing Existing Strategies

Heteroatom doping is an effective strategy to increase the numbers of ion storage locations and ion mobility channels, thereby facilitating faster electron transport. However, it is often reported that only one metallic element or non-metallic element (such as the common elements S and P) is selected to be mixed into cobalt selenide or cobalt telluride. At the same time, to improve the conductivity and active sites of CoSex and CoTex, the selection of composite carbon materials has been limited to traditional carbon materials, such as rGO, dopamine, super-C, or metal–organic frameworks. In addition, many reports testify that MXenes have the advantages of large interlayer spacing, good conductivity, and heteroatom doping. In summary, the new strategy of multi-element doping and synthesis and the novel MXene composite method can improve the performance of cobalt telluride and cobalt selenide as anodes for alkali-ion batteries.

6.2. Expanded CBs’ Anode Material Synthesis Method

Compared to CoOx and CoSx, the available modification strategies are still limited. Currently, constructed multilayer, porous, or three-dimensional mesh structures can provide short transmission channels and large specific surface areas with active sites. This helps to improve material toughness and electrolyte contact, thus improving long-cycle performance. Relevant synthesis methods have been reported for CoSex, but less often for CoTex. Based on this, we should further explore the latter’s electrochemical performance in LIBs/SIBs/PIBs. For example, three-dimensional nanonetworks of CoTex based on hydrogels may be a good synthesis method. Furthermore, binder-free electrodes for CoTex are rarely reported. Independent electrodes can be prepared via carbon cloth chemical deposition and electrospinning to avoid the obstruction of poor-conducting binders and low-capacity conductive agents, thus saving resources, reducing impurities, and improving the energy density.

Author Contributions

Writing—review and editing, Y.Z.; supervision, Z.S. and D.Q.; resources, D.H. and L.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (22204028, 22104021, 22204159), Young Talent Support Project of Guangzhou Association for Science and Technology (QT-2023-003), Guangdong Basic and Applied Basic Research Fund Project (2022A1515110451), Guangzhou University Graduate Student Innovation Ability Cultivation Funding Program (2022GDJC-M06), and Science and Technology Projects in Guangzhou (202201010245, 2023A03J0029).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grey, C.P.; Hall, D.S. Prospects for Lithium-Ion Batteries and beyond—A 2030 Vision. Nat. Commun. 2020, 11, 6279. [Google Scholar] [CrossRef] [PubMed]
  2. Huy, V.P.H.; So, S.; Kim, I.T.; Hur, J. Self-Healing Gallium Phosphide Embedded in a Hybrid Matrix for High-Performance Li-Ion Batteries. Energy Storage Mater. 2021, 34, 669–681. [Google Scholar] [CrossRef]
  3. Zhang, X.; He, Q.; Xu, X.; Xiong, T.; Xiao, Z.; Meng, J.; Wang, X.; Wu, L.; Chen, J.; Mai, L. Insights into the Storage Mechanism of Layered VS2 Cathode in Alkali Metal-Ion Batteries. Adv. Energy Mater. 2020, 10, 1904118. [Google Scholar] [CrossRef]
  4. Sun, Y.; Wang, H.; Wei, W.; Zheng, Y.; Tao, L.; Wang, Y.; Huang, M.; Shi, J.; Shi, Z.C.; Mitlin, D. Sulfur-Rich Graphene Nanoboxes with Ultra-High Potassiation Capacity at Fast Charge: Storage Mechanisms and Device Performance. ACS Nano 2021, 15, 1652–1665. [Google Scholar] [CrossRef]
  5. Hao, Q.; Jia, G.; Wei, W.; Vinu, A.; Wang, Y.; Arandiyan, H.; Ni, B.J. Graphitic Carbon Nitride with Different Dimensionalities for Energy and Environmental Applications. Nano Res. 2020, 13, 18–37. [Google Scholar] [CrossRef]
  6. Kim, H.; Kim, M.J.; Yoon, Y.H.; Nguyen, Q.H.; Kim, I.T.; Hur, J.; Lee, S.G. Sb2Te3–TiC–C Nanocomposites for the High-Performance Anode in Lithium-Ion Batteries. Electrochim. Acta 2019, 293, 8–18. [Google Scholar] [CrossRef]
  7. Xie, H.; Kalisvaart, W.P.; Olsen, B.C.; Luber, E.J.; Mitlin, D.; Buriak, J.M. Sn–Bi–Sb Alloys as Anode Materials for Sodium Ion Batteries. J. Mater. Chem. A Mater. 2017, 5, 9661–9670. [Google Scholar] [CrossRef]
  8. Zhou, Y.; Sun, W.; Rui, X.; Zhou, Y.; Ng, W.J.; Yan, Q.; Fong, E. Biochemistry-Derived Porous Carbon-Encapsulated Metal Oxide Nanocrystals for Enhanced Sodium Storage. Nano Energy 2016, 21, 71–79. [Google Scholar] [CrossRef]
  9. Xiao, Y.; Hwang, J.Y.; Belharouak, I.; Sun, Y.K. Na Storage Capability Investigation of a Carbon Nanotube-Encapsulated Fe1–xS Composite. ACS Energy Lett. 2017, 2, 364–372. [Google Scholar] [CrossRef]
  10. Zhao, Y.; Wang, L.P.; Sougrati, M.T.; Feng, Z.; Leconte, Y.; Fisher, A.; Srinivasan, M.; Xu, Z. A Review on Design Strategies for Carbon Based Metal Oxides and Sulfides Nanocomposites for High Performance Li and Na Ion Battery Anodes. Adv. Energy Mater. 2017, 7, 1601424. [Google Scholar] [CrossRef]
  11. Hartmann, F.; Etter, M.; Cibin, G.; Liers, L.; Terraschke, H.; Bensch, W. Superior Sodium Storage Properties in the Anode Material NiCr2S4 for Sodium-Ion Batteries: An X-ray Diffraction, Pair Distribution Function, and X-ray Absorption Study Reveals a Conversion Mechanism via Nickel Extrusion. Adv. Mater. 2021, 33, 2101576. [Google Scholar] [CrossRef] [PubMed]
  12. Park, C.; Samuel, E.; Joshi, B.; Kim, T.; Aldalbahi, A.; El-Newehy, M.; Yoon, W.Y.; Yoon, S.S. Supersonically Sprayed Fe2O3/C/CNT Composites for Highly Stable Li-Ion Battery Anodes. Chem. Eng. J. 2020, 395, 125018. [Google Scholar] [CrossRef]
  13. Ni, J.; Sun, M.; Li, L. Highly Efficient Sodium Storage in Iron Oxide Nanotube Arrays Enabled by Built-In Electric Field. Adv. Mater. 2019, 31, e1902603. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, X.; Liu, Y.; Han, H.; Zhao, Y.; Ma, W.; Sun, H. Polyaniline Coated Fe3O4 Hollow Nanospheres as Anode Materials for Lithium Ion Batteries. Sustain. Energy Fuels 2017, 1, 915–922. [Google Scholar] [CrossRef]
  15. Xiao, Y.; Miao, Y.; Hu, S.; Gong, F.; Yu, Q.; Zhou, L.; Chen, S. Structural Stability Boosted in 3D Carbon-Free Iron Selenide through Engineering Heterointerfaces with Se-P Bonds for Appealing Na+-Storage. Adv. Funct. Mater. 2022, 33, 2210042. [Google Scholar] [CrossRef]
  16. Sun, Z.; Shi, W.; Chen, J.; Gu, Z.; Lai, S.; Gan, S.; Xie, J.; Li, Q.; Qu, D.; Wu, X.L.; et al. Engineering Honeycomb-like Carbon Nanosheets Encapsulated Iron Chalcogenides: Superior Cyclability and Rate Capability for Sodium Ion Half/Full Batteries. Electrochim. Acta 2022, 431, 141084. [Google Scholar] [CrossRef]
  17. Sun, Z.; Wu, X.; Gu, Z.; Han, P.; Zhao, B.; Qu, D.; Gao, L.; Liu, Z.; Han, D.; Niu, L. Rationally Designed Nitrogen-Doped Yolk-Shell Fe7Se8/Carbon Nanoboxes with Enhanced Sodium Storage in Half/Full Cells. Carbon 2020, 166, 175–182. [Google Scholar] [CrossRef]
  18. Liu, Y.; Deng, Q.; Li, Y.; Li, Y.; Zhong, W.; Hu, J.; Ji, X.; Yang, C.; Lin, Z.; Huang, K. CoSe@N-Doped Carbon Nanotubes as a Potassium-Ion Battery Anode with High Initial Coulombic Efficiency and Superior Capacity Retention. ACS Nano 2021, 15, 1121–1132. [Google Scholar] [CrossRef]
  19. Li, X.; Han, Z.; Yang, W.; Li, Q.; Li, H.; Xu, J.; Li, H.; Liu, B.; Zhao, H.; Li, S.; et al. 3D Ordered Porous Hybrid of ZnSe/N-Doped Carbon with Anomalously High Na+ Mobility and Ultrathin Solid Electrolyte Interphase for Sodium-Ion Batteries. Adv. Funct. Mater. 2021, 31, 2106194. [Google Scholar] [CrossRef]
  20. Ge, J.M.; Fan, L.; Wang, J.; Zhang, Q.; Liu, Z.; Zhang, E.; Liu, Q.; Yu, X.; Lu, B. MoSe2/N-Doped Carbon as Anodes for Potassium-Ion Batteries. Adv. Energy Mater. 2018, 8, 1801477. [Google Scholar] [CrossRef]
  21. Liu, Y.; Yang, C.; Li, Y.; Zheng, F.; Li, Y.; Deng, Q.; Zhong, W.; Wang, G.; Liu, T. FeSe2/Nitrogen-Doped Carbon as Anode Material for Potassium-Ion Batteries. Chem. Eng. J. 2020, 393, 124590. [Google Scholar] [CrossRef]
  22. Sun, D.; Liu, S.; Zhang, G.; Zhou, J. NiTe2/N-Doped Graphitic Carbon Nanosheets Derived from Ni-Hexamine Coordination Frameworks for Na-Ion Storage. Chem. Eng. J. 2019, 359, 1659–1667. [Google Scholar] [CrossRef]
  23. Park, G.D.; Kang, Y.C. Conversion Reaction Mechanism for Yolk-Shell-Structured Iron Telluride-C Nanospheres and Exploration of Their Electrochemical Performance as an Anode Material for Potassium-Ion Batteries. Small Methods 2020, 4, 2000556. [Google Scholar] [CrossRef]
  24. Bao, S.J.; Li, Y.; Li, C.M.; Bao, Q.; Lu, Q.; Guo, J. Shape Evolution and Magnetic Properties of Cobalt Sulfide. Cryst. Growth Des. 2008, 8, 3745–3749. [Google Scholar] [CrossRef]
  25. Wang, L.H.; Teng, X.L.; Qin, Y.F.; Li, Q. High Electrochemical Performance and Structural Stability of CoO Nanosheets/CoO Film as Self-Supported Anodes for Lithium-Ion Batteries. Ceram. Int. 2021, 47, 5739–5746. [Google Scholar] [CrossRef]
  26. Ren, L.L.; Wang, L.H.; Qin, Y.F.; Li, Q. One-Pot Synthesized Amorphous Cobalt Sulfide with Enhanced Electrochemical Performance as Anodes for Lithium-Ion Batteries. Front. Chem. 2022, 9, 818255. [Google Scholar] [CrossRef] [PubMed]
  27. Lyu, Y.; Wu, X.; Wang, K.; Feng, Z.; Cheng, T.; Liu, Y.; Wang, M.; Chen, R.; Xu, L.; Zhou, J.; et al. An Overview on the Advances of LiCoO2 Cathodes for Lithium-Ion Batteries. Adv. Energy Mater. 2021, 11, 2000982. [Google Scholar] [CrossRef]
  28. Xiang, X.; Zhang, K.; Chen, J. Recent Advances and Prospects of Cathode Materials for Sodium-Ion Batteries. Adv. Mater. 2015, 27, 5343–5364. [Google Scholar] [CrossRef]
  29. Kim, W.; Shin, D.; Seo, B.; Chae, S.; Jo, E.; Choi, W. Precisely Tunable Synthesis of Binder-Free Cobalt Oxide-Based Li-Ion Battery Anode Using Scalable Electrothermal Waves. ACS Nano 2022, 16, 17313–17325. [Google Scholar] [CrossRef]
  30. Xiao, Q.; Song, Q.; Zheng, K.; Zheng, L.; Zhu, Y.; Chen, Z. CoSe2 Nanodots Confined in Multidimensional Porous Nanoarchitecture towards Efficient Sodium Ion Storage. Nano Energy 2022, 98, 107326. [Google Scholar] [CrossRef]
  31. Zhou, Y.; Han, Y.; Zhang, H.; Sui, D.; Sun, Z.; Xiao, P.; Wang, X.; Ma, Y.; Chen, Y. A Carbon Cloth-Based Lithium Composite Anode for High-Performance Lithium Metal Batteries. Energy Storage Mater. 2018, 14, 222–229. [Google Scholar] [CrossRef]
  32. Liu, J.; Liang, J.; Wang, C.; Ma, J. Electrospun CoSe@N-Doped Carbon Nanofibers with Highly Capacitive Li Storage. J. Energy Chem. 2019, 33, 160–166. [Google Scholar] [CrossRef]
  33. Wang, C.; Zhang, B.; Xia, H.; Cao, L.; Luo, B.; Fan, X.; Zhang, J.; Ou, X. Composition and Architecture Design of Double-Shelled Co0.85Se1−xSx@Carbon/Graphene Hollow Polyhedron with Superior Alkali (Li, Na, K)-Ion Storage. Small 2020, 16, e1905853. [Google Scholar] [CrossRef] [PubMed]
  34. Yu, Q.; Jiang, B.; Hu, J.; Lao, C.Y.; Gao, Y.; Li, P.; Liu, Z.; Suo, G.; He, D.; Wang, W.; et al. Metallic Octahedral CoSe2 Threaded by N-Doped Carbon Nanotubes: A Flexible Framework for High-Performance Potassium-Ion Batteries. Adv. Sci. 2018, 5, 1800782. [Google Scholar] [CrossRef]
  35. Yang, E.; Ji, H.; Jung, Y. Two-Dimensional Transition Metal Dichalcogenide Monolayers as Promising Sodium Ion Battery Anodes. J. Phys. Chem. C 2015, 119, 26374–26380. [Google Scholar] [CrossRef]
  36. Ni, W.; Li, X.; Shi, L.Y.; Ma, J. Research Progress on ZnSe and ZnTe Anodes for Rechargeable Batteries. Nanoscale 2022, 14, 9609–9635. [Google Scholar] [CrossRef]
  37. Yang, S.H.; Park, S.K.; Park, G.D.; Lee, J.H.; Kang, Y.C. Conversion Reaction Mechanism of Ultrafine Bimetallic Co–Fe Selenides Embedded in Hollow Mesoporous Carbon Nanospheres and Their Excellent K-Ion Storage Performance. Small 2020, 16, e2002345. [Google Scholar] [CrossRef]
  38. Xu, X.; Zhang, Y.; Sun, H.; Zhou, J.; Liu, Z.; Qiu, Z.; Wang, D.; Yang, C.; Zeng, Q.; Peng, Z.; et al. Orthorhombic Cobalt Ditelluride with Te Vacancy Defects Anchoring on Elastic MXene Enables Efficient Potassium-Ion Storage. Adv. Mater. 2021, 33, 2100272. [Google Scholar] [CrossRef]
  39. Xiao, S.; Li, X.; Zhang, W.; Xiang, Y.; Li, T.; Niu, X.; Chen, J.S.; Yan, Q. Bilateral Interfaces in In2Se3–CoIn2–CoSe2 Heterostructures for High-Rate Reversible Sodium Storage. ACS Nano 2021, 15, 13307–13318. [Google Scholar] [CrossRef]
  40. Zhang, C.; Li, H.; Zeng, X.; Xi, S.; Wang, R.; Zhang, L.; Liang, G.; Davey, K.; Liu, Y.; Zhang, L.; et al. Accelerated Diffusion Kinetics in ZnTe/CoTe2 Heterojunctions for High Rate Potassium Storage. Adv. Energy Mater. 2022, 12, 2202577. [Google Scholar] [CrossRef]
  41. Li, Q.; Peng, J.; Zhang, W.; Wang, L.; Liang, Z.; Wang, G.; Wu, J.; Fan, W.; Li, H.; Wang, J.; et al. Manipulating the Polytellurides of Metallic Telluride for Ultra-Stable Potassium-Ion Storage: A Case Study of Carbon-Confined CoTe2 Nanofibers. Adv. Energy Mater. 2023, 13, 2300150. [Google Scholar] [CrossRef]
  42. Zhang, G.; Liu, K.; Zhou, J. Cobalt Telluride/Graphene Composite Nanosheets for Excellent Gravimetric and Volumetric Na-Ion Storage. J. Mater. Chem. A Mater. 2018, 6, 6335–6343. [Google Scholar] [CrossRef]
  43. Cho, J.S.; Won, J.M.; Lee, J.K.; Kang, Y.C. Design and Synthesis of Multiroom-Structured Metal Compounds-Carbon Hybrid Microspheres as Anode Materials for Rechargeable Batteries. Nano Energy 2016, 26, 466–478. [Google Scholar] [CrossRef]
  44. Li, J.; Yan, D.; Lu, T.; Yao, Y.; Pan, L. An Advanced CoSe Embedded within Porous Carbon Polyhedra Hybrid for High Performance Lithium-Ion and Sodium-Ion Batteries. Chem. Eng. J. 2017, 325, 14–24. [Google Scholar] [CrossRef]
  45. Yang, X.; Wang, S.; Yu, D.Y.W.; Rogach, A.L. Direct Conversion of Metal-Organic Frameworks into Selenium/Selenide/Carbon Composites with High Sodium Storage Capacity. Nano Energy 2019, 58, 392–398. [Google Scholar] [CrossRef]
  46. Sun, Z.; Gu, Z.; Shi, W.; Sun, Z.; Gan, S.; Xu, L.; Liang, H.; Ma, Y.; Qu, D.; Zhong, L.; et al. Mesoporous N-Doped Carbon-Coated CoSe Nanocrystals Encapsulated in S-Doped Carbon Nanosheets as Advanced Anode with Ultrathin Solid Electrolyte Interphase for High-Performance Sodium-Ion Half/Full Batteries. J. Mater. Chem. A Mater. 2022, 10, 2113–2121. [Google Scholar] [CrossRef]
  47. Zhang, W.; Wang, X.; Wong, K.W.; Zhang, W.; Chen, T.; Zhao, W.; Huang, S. Rational Design of Embedded CoTe2 nanoparticles in Freestanding N-Doped Multichannel Carbon Fibers for Sodium-Ion Batteries with Ultralong Cycle Lifespan. ACS Appl. Mater. Interfaces 2021, 13, 34134–34144. [Google Scholar] [CrossRef]
  48. Chen, Z.; Chen, M.; Yan, X.; Jia, H.; Fei, B.; Ha, Y.; Qing, H.; Yang, H.; Liu, M.; Wu, R. Vacancy Occupation-Driven Polymorphic Transformation in Cobalt Ditelluride for Boosted Oxygen Evolution Reaction. ACS Nano 2020, 14, 6968–6979. [Google Scholar] [CrossRef]
  49. Fan, H.; Mao, P.; Sun, H.; Wang, Y.; Mofarah, S.S.; Koshy, P.; Arandiyan, H.; Wang, Z.; Liu, Y.; Shao, Z. Recent Advances of Metal Telluride Anodes for High-Performance Lithium/Sodium-Ion Batteries. Mater. Horiz. 2022, 9, 524–546. [Google Scholar] [CrossRef]
  50. Jiang, Y.; Zou, G.; Hou, H.; Li, J.; Liu, C.; Qiu, X.; Ji, X. Composition Engineering Boosts Voltage Windows for Advanced Sodium-Ion Batteries. ACS Nano 2019, 13, 10787–10797. [Google Scholar] [CrossRef]
  51. Jiang, Y.; Xie, M.; Wu, F.; Ye, Z.; Zhang, Y.; Wang, Z.; Zhou, Y.; Li, L.; Chen, R. Cobalt Selenide Hollow Polyhedron Encapsulated in Graphene for High-Performance Lithium/Sodium Storage. Small 2021, 17, 2102893. [Google Scholar] [CrossRef] [PubMed]
  52. Ganesan, V.; Nam, K.H.; Park, C.M. Robust Polyhedral CoTe2–C Nanocomposites as High-Performance Li-A Nd Na-Ion Battery Anodes. ACS Appl. Energy Mater. 2020, 3, 4877–4887. [Google Scholar] [CrossRef]
  53. Park, S.; Park, S.; Mathew, V.; Sambandam, B.; Hwang, J.Y.; Kim, J. A New Tellurium-Based Ni3TeO6-Carbon Nanotubes Composite Anode for Na-Ion Battery. Int. J. Energy Res. 2022, 46, 16041–16049. [Google Scholar] [CrossRef]
  54. Ge, P.; Li, S.; Xu, L.; Zou, K.; Gao, X.; Cao, X.; Zou, G.; Hou, H.; Ji, X. Hierarchical Hollow-Microsphere Metal–Selenide@Carbon Composites with Rational Surface Engineering for Advanced Sodium Storage. Adv. Energy Mater. 2019, 9, 1803035. [Google Scholar] [CrossRef]
  55. Miao, Y.; Xiao, Y.; Hu, S.; Chen, S. Chalcogenides Metal-Based Heterostructure Anode Materials toward Na+-Storage Application. Nano Res. 2022, 16, 2347–2365. [Google Scholar] [CrossRef]
  56. Banhart, F.; Kotakoski, J.; Krasheninnikov, A.V. Structural Defects in Graphene. ACS Nano 2011, 5, 26–41. [Google Scholar] [CrossRef]
  57. Dong, C.; Zhang, X.; Dong, W.; Lin, X.; Cheng, Y.; Tang, Y.; Zhao, S.; Li, G.; Huang, F. ZnO/ZnS Heterostructure with Enhanced Interfacial Lithium Absorption for Robust and Large-Capacity Energy Storage. Energy Environ. Sci. 2022, 15, 4738–4747. [Google Scholar] [CrossRef]
  58. Arandiyan, H.; Mofarah, S.S.; Sorrell, C.C.; Doustkhah, E.; Sajjadi, B.; Hao, D.; Wang, Y.; Sun, H.; Ni, B.J.; Rezaei, M.; et al. Defect Engineering of Oxide Perovskites for Catalysis and Energy Storage: Synthesis of Chemistry and Materials Science. Chem. Soc. Rev. 2021, 50, 10116–10211. [Google Scholar] [CrossRef]
  59. Wang, Y.; Arandiyan, H.; Chen, X.; Zhao, T.; Bo, X.; Su, Z.; Zhao, C. Microwave-Induced Plasma Synthesis of Defect-Rich, Highly Ordered Porous Phosphorus-Doped Cobalt Oxides for Overall Water Electrolysis. J. Phys. Chem. C 2020, 124, 9971–9978. [Google Scholar] [CrossRef]
  60. Wang, X.; Li, F.; Li, W.; Gao, W.; Tang, Y.; Li, R. Hollow Bimetallic Cobalt-Based Selenide Polyhedrons Derived from Metal-Organic Framework: An Efficient Bifunctional Electrocatalyst for Overall Water Splitting. J. Mater. Chem. A Mater. 2017, 5, 17982–17989. [Google Scholar] [CrossRef]
  61. Fang, C.; Huang, Y.; Zhang, W.; Han, J.; Deng, Z.; Cao, Y.; Yang, H. Routes to High Energy Cathodes of Sodium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1501727. [Google Scholar] [CrossRef]
  62. Billaud, J.; Singh, G.; Armstrong, A.R.; Gonzalo, E.; Roddatis, V.; Armand, M.; Rojo, T.; Bruce, P.G. Na0.67Mn1−XMgxO2 (0 ≤ x ≤ 0.2): A High Capacity Cathode for Sodium-Ion Batteries. Energy Environ. Sci. 2014, 7, 1387–1391. [Google Scholar] [CrossRef]
  63. Narsimulu, D.; Kakarla, A.K.; Vamsi Krishna, B.N.; Shanthappa, R.; Yu, J.S. Nitrogen-Doped Reduced Graphene Oxide Incorporated Ni2O3–Co3O4@MoS2 Hollow Nanocubes for High-Performance Energy Storage Devices. J. Alloys Compd. 2022, 922, 166131. [Google Scholar] [CrossRef]
  64. Jana, J.; Sharma, T.S.K.; Chung, J.S.; Choi, W.M.; Hur, S.H. The Role of Surface Carbide/Oxide Heterojunction in Electrocatalytic Behavior of 3D-Nanonest Supported FeiMoj Composites. J. Alloys Compd. 2023, 946, 169395. [Google Scholar] [CrossRef]
  65. Pellow, M.A.; Emmott, C.J.M.; Barnhart, C.J.; Benson, S.M. Hydrogen or Batteries for Grid Storage? A Net Energy Analysis. Energy Environ. Sci. 2015, 8, 1938–1952. [Google Scholar] [CrossRef]
  66. Ge, P.; Hou, H.; Li, S.; Huang, L.; Ji, X. Three-Dimensional Hierarchical Framework Assembled by Cobblestone-Like CoSe2@C Nanospheres for Ultrastable Sodium-Ion Storage. ACS Appl. Mater. Interfaces 2018, 10, 14716–14726. [Google Scholar] [CrossRef]
  67. Ge, P.; Hou, H.; Li, S.; Yang, L.; Ji, X. Tailoring Rod-Like FeSe2 Coated with Nitrogen-Doped Carbon for High-Performance Sodium Storage. Adv. Funct. Mater. 2018, 28, 1801765. [Google Scholar] [CrossRef]
  68. Lu, M.; Liu, C.; Li, X.; Jiang, S.; Yao, Z.; Liu, T.; Yang, Y. Mixed Phase Mo-Doped CoSe2 Nanosheets Encapsulated in N-Doped Carbon Shell with Boosted Sodium Storage Performance. J. Alloys Compd. 2022, 922, 166265. [Google Scholar] [CrossRef]
  69. Wang, M.Y.; Wang, X.L.; Yao, Z.J.; Xie, D.; Xia, X.H.; Gu, C.D.; Tu, J.P. Molybdenum-Doped Tin Oxide Nanoflake Arrays Anchored on Carbon Foam as Flexible Anodes for Sodium-Ion Batteries. J. Colloid Interface Sci. 2020, 560, 169–176. [Google Scholar] [CrossRef]
  70. Wang, P.; Huang, J.; Zhang, J.; Wang, L.; Sun, P.; Yang, Y.; Yao, Z. Coupling Hierarchical Iron Cobalt Selenide Arrays with N-Doped Carbon as Advanced Anodes for Sodium Ion Storage. J. Mater. Chem. A Mater. 2021, 9, 7248–7256. [Google Scholar] [CrossRef]
  71. Hu, L.; Li, L.; Zhang, Y.; Tan, X.; Yang, H.; Lin, X.; Tong, Y. Construction of Cobalt Vacancies in Cobalt Telluride to Induce Fast Ionic/Electronic Diffusion Kinetics for Lithium-Ion Half/Full Batteries. J. Mater. Sci. Technol. 2022, 127, 124–132. [Google Scholar] [CrossRef]
  72. Fang, G.; Wang, Q.; Zhou, J.; Lei, Y.; Chen, Z.; Wang, Z.; Pan, A.; Liang, S. Metal Organic Framework-Templated Synthesis of Bimetallic Selenides with Rich Phase Boundaries for Sodium-Ion Storage and Oxygen Evolution Reaction. ACS Nano 2019, 13, 5635–5645. [Google Scholar] [CrossRef] [PubMed]
  73. Ye, J.; Li, X.; Xia, G.; Gong, G.; Zheng, Z.; Chen, C.; Hu, C. P-Doped CoSe2 Nanoparticles Embedded in 3D Honeycomb-like Carbon Network for Long Cycle-Life Na-Ion Batteries. J. Mater. Sci. Technol. 2021, 77, 100–107. [Google Scholar] [CrossRef]
  74. Shin, I.; Cho, W.J.; An, E.S.; Park, S.; Jeong, H.W.; Jang, S.; Baek, W.J.; Park, S.Y.; Yang, D.H.; Seo, J.H.; et al. Spin–Orbit Torque Switching in an All-Van Der Waals Heterostructure. Adv. Mater. 2022, 34, e2101730. [Google Scholar] [CrossRef]
  75. Pan, L.; Grutter, A.; Zhang, P.; Che, X.; Nozaki, T.; Stern, A.; Street, M.; Zhang, B.; Casas, B.; He, Q.L.; et al. Observation of Quantum Anomalous Hall Effect and Exchange Interaction in Topological Insulator/Antiferromagnet Heterostructure. Adv. Mater. 2020, 32, 2001460. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, H.; Niu, R.; Liu, J.; Guo, S.; Yang, Y.; Liu, Z.; Li, J. Electrostatic Self-Assembly of 2D/2D CoWO4/g–C3N4 p—n Heterojunction for Improved Photocatalytic Hydrogen Evolution: Built-in Electric Field Modulated Charge Separation and Mechanism Unveiling. Nano Res. 2022, 15, 6987–6998. [Google Scholar] [CrossRef]
  77. Huang, S.; Wang, Z.; von Lim, Y.; Wang, Y.; Li, Y.; Zhang, D.; Yang, H.Y. Recent Advances in Heterostructure Engineering for Lithium–Sulfur Batteries. Adv. Energy Mater. 2021, 11, 2003689. [Google Scholar] [CrossRef]
  78. Li, Y.; Zhang, J.; Chen, Q.; Xia, X.; Chen, M. Emerging of Heterostructure Materials in Energy Storage: A Review. Adv. Mater. 2021, 33, 2100855. [Google Scholar] [CrossRef]
  79. Liang, L.; Gu, W.; Wu, Y.; Zhang, B.; Wang, G.; Yang, Y.; Ji, G. Heterointerface Engineering in Electromagnetic Absorbers: New Insights and Opportunities. Adv. Mater. 2022, 34, 2106195. [Google Scholar] [CrossRef]
  80. Wang, S.; Yang, Y.; Quan, W.; Hong, Y.; Zhang, Z.; Tang, Z.; Li, J. Ti3+-Free Three-Phase Li4Ti5O12/TiO2 for High-Rate Lithium Ion Batteries: Capacity and Conductivity Enhancement by Phase Boundaries. Nano Energy 2017, 32, 294–301. [Google Scholar] [CrossRef]
  81. Luo, D.; Ma, C.; Hou, J.; Zhang, Z.; Feng, R.; Yang, L.; Zhang, X.; Lu, H.; Liu, J.; Li, Y.; et al. Integrating Nanoreactor with O–Nb–C Heterointerface Design and Defects Engineering Toward High-Efficiency and Longevous Sodium Ion Battery. Adv. Energy Mater. 2022, 12, 2103716. [Google Scholar] [CrossRef]
  82. Yang, C.; Liang, X.; Ou, X.; Zhang, Q.; Zheng, H.S.; Zheng, F.; Wang, J.H.; Huang, K.; Liu, M. Heterostructured Nanocube-Shaped Binary Sulfide (SnCo)S2 Interlaced with S-Doped Graphene as a High-Performance Anode for Advanced Na+ Batteries. Adv. Funct. Mater. 2019, 29, 1807971. [Google Scholar] [CrossRef]
  83. Xiao, Y.; Miao, Y.; Wan, S.; Sun, Y.K.; Chen, S. Synergistic Engineering of Se Vacancies and Heterointerfaces in Zinc-Cobalt Selenide Anode for Highly Efficient Na-Ion Batteries. Small 2022, 18, 2202582. [Google Scholar] [CrossRef] [PubMed]
  84. Saliba, D.; Ammar, M.; Rammal, M.; Al-Ghoul, M.; Hmadeh, M. Crystal Growth of ZIF-8, ZIF-67, and Their Mixed-Metal Derivatives. J. Am. Chem. Soc. 2018, 140, 1812–1823. [Google Scholar] [CrossRef] [PubMed]
  85. Venna, S.R.; Jasinski, J.B.; Carreon, M.A. Structural Evolution of Zeolitic Imidazolate Framework-8. J. Am. Chem. Soc. 2010, 132, 18030–18033. [Google Scholar] [CrossRef]
  86. He, Y.; Luo, M.; Dong, C.; Ding, X.; Yin, C.; Nie, A.; Chen, Y.; Qian, Y.; Xu, L. Coral-like NixCo1−xSe2 for Na-Ion Battery with Ultralong Cycle Life and Ultrahigh Rate Capability. J. Mater. Chem. A Mater. 2019, 7, 3933–3940. [Google Scholar] [CrossRef]
  87. Zhu, H.; Li, Z.; Xu, F.; Qin, Z.; Sun, R.; Wang, C.; Lu, S.; Zhang, Y.; Fan, H. Ni3Se4@CoSe2 Hetero-Nanocrystals Encapsulated into CNT-Porous Carbon Interpenetrating Frameworks for High-Performance Sodium Ion Battery. J. Colloid Interface Sci. 2022, 611, 718–725. [Google Scholar] [CrossRef]
  88. Zhang, K.; Park, M.; Zhou, L.; Lee, G.H.; Li, W.; Kang, Y.M.; Chen, J. Urchin-Like CoSe2 as a High-Performance Anode Material for Sodium-Ion Batteries. Adv. Funct. Mater. 2016, 26, 6728–6735. [Google Scholar] [CrossRef]
  89. Ko, Y.N.; Choi, S.H.; Kang, Y.C. Hollow Cobalt Selenide Microspheres: Synthesis and Application as Anode Materials for Na-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 6449–6456. [Google Scholar] [CrossRef]
  90. Hu, J.-Z.; Liu, W.-J.; Zheng, J.-H.; Li, G.-C.; Bu, Y.-F.; Qiao, F.; Lian, J.-B.; Zhao, Y. Coral-like Cobalt Selenide/Carbon Nanosheet Arrays Attached on Carbon Nanofibers for High-Rate Sodium-Ion Storage. Rare Met. 2023, 42, 916–928. [Google Scholar] [CrossRef]
  91. Hong, L.; Ju, S.; Yang, Y.; Zheng, J.; Xia, G.; Huang, Z.; Liu, X.; Yu, X. Hollow-Shell Structured Porous CoSe2 Microspheres Encapsulated by MXene Nanosheets for Advanced Lithium Storage. Sustain. Energy Fuels 2020, 4, 2352–2362. [Google Scholar] [CrossRef]
  92. Gu, X.; Zhang, L.; Zhang, W.; Liu, S.; Wen, S.; Mao, X.; Dai, P.; Li, L.; Liu, D.; Zhao, X.; et al. A CoSe-C@C Core-Shell Structure with Stable Potassium Storage Performance Realized by an Effective Solid Electrolyte Interphase Layer. J. Mater. Chem. A Mater. 2021, 9, 11397–11404. [Google Scholar] [CrossRef]
  93. Fan, H.; Liu, C.; Lan, G.; Mao, P.; Zheng, R.; Wang, Z.; Liu, Y.; Sun, H. Uniform Carbon Coating Mediated Multiphase Interface in Submicron Sized Rodlike Cobalt Ditelluride Anodes for High-Capacity and Fast Lithium Storage. Electrochim. Acta 2023, 439, 141614. [Google Scholar] [CrossRef]
  94. Nie, P.; Le, Z.; Chen, G.; Liu, D.; Liu, X.; Wu, H.B.; Xu, P.; Li, X.; Liu, F.; Chang, L.; et al. Graphene Caging Silicon Particles for High-Performance Lithium-Ion Batteries. Small 2018, 14, 1800635. [Google Scholar] [CrossRef]
  95. Xu, H.; Cao, J.; Shan, C.; Wang, B.; Xi, P.; Liu, W.; Tang, Y. MOF-Derived Hollow CoS Decorated with CeOx Nanoparticles for Boosting Oxygen Evolution Reaction Electrocatalysis. Angew. Chem. 2018, 130, 8790–8794. [Google Scholar] [CrossRef]
  96. Huang, P.; Ying, H.; Zhang, S.; Zhang, Z.; Han, W.Q. In Situ Fabrication of MXene/CuS Hybrids with Interfacial Covalent Bonding via Lewis Acidic Etching Route for Efficient Sodium Storage. J. Mater. Chem. A Mater. 2022, 10, 22135–22144. [Google Scholar] [CrossRef]
  97. Jin, T.; Han, Q.; Wang, Y.; Jiao, L. 1D Nanomaterials: Design, Synthesis, and Applications in Sodium–Ion Batteries. Small 2018, 14, 1703086. [Google Scholar] [CrossRef]
  98. Nie, S.; Liu, L.; Liu, J.; Xie, J.; Zhang, Y.; Xia, J.; Yan, H.; Yuan, Y.; Wang, X. Nitrogen-Doped TiO2–C Composite Nanofibers with High-Capacity and Long-Cycle Life as Anode Materials for Sodium-Ion Batteries. Nanomicro Lett. 2018, 10, 71. [Google Scholar] [CrossRef]
  99. La Monaca, A.; Paolella, A.; Guerfi, A.; Rosei, F.; Zaghib, K. Electrospun Ceramic Nanofibers as 1D Solid Electrolytes for Lithium Batteries. Electrochem. Commun. 2019, 104, 106483. [Google Scholar] [CrossRef]
  100. Park, J.S.; Park, G.D.; Kang, Y.C. Exploration of Cobalt Selenite–Carbon Composite Porous Nanofibers as Anode for Sodium-Ion Batteries and Unveiling Their Conversion Reaction Mechanism. J. Mater. Sci. Technol. 2021, 89, 24–35. [Google Scholar] [CrossRef]
  101. Li, F.; Li, L.; Yao, T.; Liu, T.; Zhu, L.; Li, Y.; Lu, H.; Qian, R.; Liu, Y.; Wang, H. Electrospinning Synthesis of Porous Carbon Nanofiber Supported CoSe2 Nanoparticles towards Enhanced Sodium Ion Storage. Mater. Chem. Phys. 2021, 262, 124314. [Google Scholar] [CrossRef]
  102. Lin, J.; Zhong, Z.; Wang, H.; Zheng, X.; Wang, Y.; Qi, J.; Cao, J.; Fei, W.; Huang, Y.; Feng, J. Rational Constructing Free-Standing Se Doped Nickel-Cobalt Sulfides Nanotubes as Battery-Type Electrode for High-Performance Supercapattery. J. Power Sources 2018, 407, 6–13. [Google Scholar] [CrossRef]
  103. Xia, Z.; Sun, H.; He, X.; Sun, Z.; Lu, C.; Li, J.; Peng, Y.; Dou, S.; Sun, J.; Liu, Z. In Situ Construction of CoSe2@vertical-Oriented Graphene Arrays as Self-Supporting Electrodes for Sodium-Ion Capacitors and Electrocatalytic Oxygen Evolution. Nano Energy 2019, 60, 385–393. [Google Scholar] [CrossRef]
  104. Li, Z.; Zhang, J.; Lou, X.W.D. Hollow Carbon Nanofibers Filled with MnO2 Nanosheets as Efficient Sulfur Hosts for Lithium-Sulfur Batteries. Angew. Chem. 2015, 127, 13078–13082. [Google Scholar] [CrossRef]
  105. Xin, D.; He, S.; Han, X.; Zhang, X.; Cheng, Z.; Xia, M. Co-Doped 1T-MoS2 Nanosheets Anchored on Carbon Cloth as Self-Supporting Anode for High-Performance Lithium Storage. J. Alloys Compd. 2022, 921, 166099. [Google Scholar] [CrossRef]
  106. Wang, G.; Chen, J.; Cai, P.; Jia, J.; Wen, Z. A Self-Supported Ni–Co Perselenide Nanorod Array as a High-Activity Bifunctional Electrode for a Hydrogen-Producing Hydrazine Fuel Cell. J. Mater. Chem. A Mater. 2018, 6, 17763–17770. [Google Scholar] [CrossRef]
  107. Liu, B.; Li, H.; Cao, B.; Jiang, J.; Gao, R.; Zhang, J. Few Layered N, P Dual-Doped Carbon-Encapsulated Ultrafine MoP Nanocrystal/MoP Cluster Hybrids on Carbon Cloth: An Ultrahigh Active and Durable 3D Self-Supported Integrated Electrode for Hydrogen Evolution Reaction in a Wide PH Range. Adv. Funct. Mater. 2018, 28, 1801527. [Google Scholar] [CrossRef]
  108. Sun, T.; Huang, C.; Shu, H.; Luo, L.; Liang, Q.; Chen, M.; Su, J.; Wang, X. Porous NiCo2S4 Nanoneedle Arrays with Highly Efficient Electrocatalysis Anchored on Carbon Cloths as Self-Supported Hosts for High-Loading Li–S Batteries. ACS Appl. Mater. Interfaces 2020, 12, 57975–57986. [Google Scholar] [CrossRef]
  109. Zeng, M.; Wang, M.; Zheng, L.; Gao, W.; Liu, R.; Pan, J.; Zhang, H.; Yang, Z.; Li, X. In Situ Enhance Lithium Polysulfides Redox Kinetics by Carbon Cloth/MoO3 Self-Standing Electrode for Lithium–Sulfur Battery. J. Mater. Sci. 2022, 57, 10003–10016. [Google Scholar] [CrossRef]
  110. Kim, I.; Park, S.W.; Kim, D.W. Carbon-Coated Tungsten Diselenide Nanosheets Uniformly Assembled on Porous Carbon Cloth as Flexible Binder-Free Anodes for Sodium-Ion Batteries with Improved Electrochemical Performance. J. Alloys Compd. 2020, 827, 154348. [Google Scholar] [CrossRef]
  111. Li, T.; Bai, Y.; Wang, Y.; Xu, H.; Jin, H. Advances in Transition-Metal (Zn, Mn, Cu)-Based MOFs and Their Derivatives for Anode of Lithium-Ion Batteries. Coord Chem. Rev. 2020, 410, 213221. [Google Scholar] [CrossRef]
  112. Wang, H.; Zhu, Q.L.; Zou, R.; Xu, Q. Metal-Organic Frameworks for Energy Applications. Chem 2017, 2, 52–80. [Google Scholar] [CrossRef]
  113. Miao, W.; Zhang, Y.; Li, H.; Zhang, Z.; Li, L.; Yu, Z.; Zhang, W. ZIF-8/ZIF-67-Derived 3D Amorphous Carbon-Encapsulated CoS/NCNTs Supported on CoS-Coated Carbon Nanofibers as an Advanced Potassium-Ion Battery Anode. J. Mater. Chem. A Mater. 2019, 7, 5504–5512. [Google Scholar] [CrossRef]
  114. Cheng, Z.; Du, T.; Gong, H.; Zhou, L. In-Situ Synthesis of Fe7S8 on Metal Sites of MOFs as High-Capacity and Fast-Kinetics Anodes for Sodium Ion Batteries. J. Alloys Compd. 2023, 940, 168854. [Google Scholar] [CrossRef]
  115. Guo, W.; Sun, W.; Lv, L.P.; Kong, S.; Wang, Y. Microwave-Assisted Morphology Evolution of Fe-Based Metal-Organic Frameworks and Their Derived Fe2O3 Nanostructures for Li-Ion Storage. ACS Nano 2017, 11, 4198–4205. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, Y.; Pan, A.; Ding, L.; Zhou, Z.; Wang, Y.; Niu, S.; Liang, S.; Cao, G. Nitrogen-Doped Yolk-Shell-Structured CoSe/C Dodecahedra for High-Performance Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 3624–3633. [Google Scholar] [CrossRef]
  117. Feng, J.; Luo, S.H.; Yan, S.X.; Zhan, Y.; Wang, Q.; Zhang, Y.H.; Liu, X.; Chang, L.J. Rational Design of Yolk–Shell Zn-Co-Se@N-Doped Dual Carbon Architectures as Long-Life and High-Rate Anodes for Half/Full Na-Ion Batteries. Small 2021, 17, e2101887. [Google Scholar] [CrossRef]
  118. Yin, Y.; Rioux, R.M.; Erdonmez, C.K.; Hughes, S.; Somorjai, G.A.; Alivisatos, A.P. Formation of Hollow Nanocrystals through the Nanoscale Kirkendall Effect. Science 2004, 304, 711–714. [Google Scholar] [CrossRef]
  119. Wang, B.; Miao, X.; Dong, H.; Ma, X.; Wu, J.; Cheng, Y.; Geng, H.; Li, C.C. In Situ construction of Active Interfaces towards Improved High-Rate Performance of CoSe2. J. Mater. Chem. A Mater. 2021, 9, 14582–14592. [Google Scholar] [CrossRef]
  120. Pang, J.; Mendes, R.G.; Bachmatiuk, A.; Zhao, L.; Ta, H.Q.; Gemming, T.; Liu, H.; Liu, Z.; Rummeli, M.H. Applications of 2D MXenes in Energy Conversion and Storage Systems. Chem. Soc. Rev. 2019, 48, 72–133. [Google Scholar] [CrossRef]
  121. Yu, H.; Wang, Y.; Jing, Y.; Ma, J.; Du, C.F.; Yan, Q. Surface Modified MXene-Based Nanocomposites for Electrochemical Energy Conversion and Storage. Small 2019, 15, e1901503. [Google Scholar] [CrossRef] [PubMed]
  122. Sun, H.; Zhang, Y.; Xu, X.; Zhou, J.; Yang, F.; Li, H.; Chen, H.; Chen, Y.; Liu, Z.; Qiu, Z.; et al. Strongly Coupled Te-SnS2/MXene Superstructure with Self-Autoadjustable Function for Fast and Stable Potassium Ion Storage. J. Energy Chem. 2021, 61, 416–424. [Google Scholar] [CrossRef]
  123. Kannan, K.; Sadasivuni, K.K.; Abdullah, A.M.; Kumar, B. Current Trends in MXene-Based Nanomaterials for Energy Storage and Conversion System: A Mini Review. Catalysts 2020, 10, 495. [Google Scholar] [CrossRef]
  124. Nguyen, V.H.; Nguyen, B.S.; Hu, C.; Nguyen, C.C.; Nguyen, D.L.T.; Dinh, M.T.N.; Vo, D.V.N.; Trinh, Q.T.; Shokouhimehr, M.; Hasani, A.; et al. Novel Architecture Titanium Carbide (Ti3C2Tx) Mxene Cocatalysts toward Photocatalytic Hydrogen Production: A Mini-Review. Nanomaterials 2020, 10, 602. [Google Scholar] [CrossRef] [PubMed]
  125. Ran, J.; Gao, G.; Li, F.T.; Ma, T.Y.; Du, A.; Qiao, S.Z. Ti3C2 MXene Co-Catalyst on Metal Sulfide Photo-Absorbers for Enhanced Visible-Light Photocatalytic Hydrogen Production. Nat. Commun. 2017, 8, 13907. [Google Scholar] [CrossRef] [PubMed]
  126. Peng, Q.; Guo, J.; Zhang, Q.; Xiang, J.; Liu, B.; Zhou, A.; Liu, R.; Tian, Y. Unique Lead Adsorption Behavior of Activated Hydroxyl Group in Two-Dimensional Titanium Carbide. J. Am. Chem. Soc. 2014, 136, 4113–4116. [Google Scholar] [CrossRef]
  127. Wang, M.; Peng, A.; Zeng, M.; Chen, L.; Li, X.; Yang, Z.; Chen, J.; Guo, B.; Ma, Z.; Li, X. Facilitating the Redox Conversion of CoSe2 nanorods by Ti3C2Tx to Improve the Electrode Durability as Anodes for Sodium-Ion Batteries. Sustain. Energy Fuels 2021, 5, 6381–6391. [Google Scholar] [CrossRef]
  128. Acharya, J.; Ojha, G.P.; Pant, B.; Park, M. Construction of Self-Supported Bimetallic MOF-Mediated Hollow and Porous Tri-Metallic Selenide Nanosheet Arrays as Battery-Type Electrodes for High-Performance Asymmetric Supercapacitors. J. Mater. Chem. A Mater. 2021, 9, 23977–23993. [Google Scholar] [CrossRef]
  129. Sarkar, D.; Das, D.; Nagarajan, S.; Mitlin, D. Thermally Fabricated Cobalt Telluride in Nitrogen-Rich Carbon Dodecahedra as High-Rate Potassium and Sodium Ion Battery Anodes. Sustain. Energy Fuels 2022, 6, 3582–3590. [Google Scholar] [CrossRef]
  130. Zhao, H.; Qi, Y.; Liang, K.; Li, J.; Zhou, L.; Chen, J.; Huang, X.; Ren, Y. Interface-Driven Pseudocapacitance Endowing Sandwiched CoSe2/N-Doped Carbon/TiO2 Microcubes with Ultra-Stable Sodium Storage and Long-Term Cycling Stability. ACS Appl. Mater. Interfaces 2021, 13, 61555–61564. [Google Scholar] [CrossRef]
  131. Lei, Y.X.; Miao, N.X.; Zhou, J.P.; Hassan, Q.U.; Wang, J.Z. Novel Magnetic Properties of CoTe Nanorods and Diversified CoTe2 Nanostructures Obtained at Different NaOH Concentrations. Sci. Technol. Adv. Mater. 2017, 18, 325–333. [Google Scholar] [CrossRef] [PubMed]
  132. Keles, O.; Karahan, B.D.; Eryilmaz, L.; Amine, R.; Abouimrane, A.; Chen, Z.; Zuo, X.; Zhu, Z.; Al-Hallaj, S.; Amine, K. Superlattice-Structured Films by Magnetron Sputtering as New Era Electrodes for Advanced Lithium-Ion Batteries. Nano Energy 2020, 76, 105094. [Google Scholar] [CrossRef]
  133. Sun, C.; Gao, L.; Yang, Y.; Yan, Z.; Zhang, D.; Bian, X. Ultrafast Microwave-Induced Synthesis of Lithiophilic Oxides Modified 3D Porous Mesh Skeleton for High-Stability Li-Metal Anode. Chem. Eng. J. 2023, 452, 139407. [Google Scholar] [CrossRef]
  134. Huang, P.; Ying, H.; Zhang, S.; Zhang, Z.; Han, W.Q. Molten Salts Etching Route Driven Universal Construction of MXene/Transition Metal Sulfides Heterostructures with Interfacial Electronic Coupling for Superior Sodium Storage. Adv. Energy Mater. 2022, 12, 2202052. [Google Scholar] [CrossRef]
  135. Hu, H.; Zhang, J.; Guan, B.; Lou, X.W.D. Unusual Formation of CoSe@carbon Nanoboxes, Which Have an Inhomogeneous Shell, for Efficient Lithium Storage. Angew. Chem. 2016, 128, 9666–9670. [Google Scholar] [CrossRef]
  136. Sui, Y.; Guo, J.; Chen, X.; Guan, J.; Chen, X.; Wei, H.; Liu, Q.; Wei, B.; Geng, H. Highly Dispersive CoSe2 Nanoparticles Encapsulated in Carbon Nanotube-Grafted Multichannel Carbon Fibers as Advanced Anodes for Sodium-Ion Half/Full Batteries. Inorg. Chem. Front. 2022, 9, 5217–5225. [Google Scholar] [CrossRef]
  137. He, Y.; Li, H.; Yu, J.; Wang, Q.; Dai, Y.; Shan, R.; Zhou, G. Hierarchical Mesoporous Binary Nickel Cobalt Selenides Nanospheres for Enhanced Sodium-Ion Storage. J. Alloys Compd. 2023, 938, 168608. [Google Scholar] [CrossRef]
  138. Zhao, J.; Wu, H.; Li, L.; Lu, S.; Mao, H.; Ding, S. A CoSe2-Based 3D Conductive Network for High-Performance Potassium Storage: Enhancing Charge Transportation by Encapsulation and Restriction Strategy. Mater. Chem. Front. 2021, 5, 5351–5360. [Google Scholar] [CrossRef]
  139. Jiang, Y.; Wu, F.; Ye, Z.; Zhou, Y.; Chen, Y.; Zhang, Y.; Lv, Z.; Li, L.; Xie, M.; Chen, R. Superimposed Effect of Hollow Carbon Polyhedron and Interconnected Graphene Network to Achieve CoTe2 Anode for Fast and Ultralong Sodium Storage. J. Power Sources 2023, 554, 232174. [Google Scholar] [CrossRef]
  140. Zhao, W.; Zhang, W.; Lei, Y.; Wang, L.; Wang, G.; Wu, J.; Fan, W.; Huang, S. Dual-Type Carbon Confinement Strategy: Improving the Stability of CoTe2 Nanocrystals for Sodium-Ion Batteries with a Long Lifespan. ACS Appl. Mater. Interfaces 2022, 14, 6801–6809. [Google Scholar] [CrossRef]
  141. Ding, Y.; Wang, W.; Bi, M.; Guo, J.; Fang, Z. CoTe Nanorods/RGO Composites as a Potential Anode Material for Sodium-Ion Storage. Electrochim. Acta 2019, 313, 331–340. [Google Scholar] [CrossRef]
  142. Yang, S.; Park, G.D.; Kang, Y.C. Conversion Reaction Mechanism of Cobalt Telluride-Carbon Composite Microspheres Synthesized by Spray Pyrolysis Process for K-Ion Storage. Appl. Surf. Sci. 2020, 529, 147140. [Google Scholar] [CrossRef]
Figure 1. Typical recent advances in the synthesis of CoSex/CoTex–based electrodes, Refs. [37,38,39,40,41,42,43,44,45].
Figure 1. Typical recent advances in the synthesis of CoSex/CoTex–based electrodes, Refs. [37,38,39,40,41,42,43,44,45].
Coatings 13 01588 g001
Figure 2. This paper summarizes from four perspectives, including electrochemical mechanisms, electrode design, synthesis, and performance; Refs. [38,42,46,47,48].
Figure 2. This paper summarizes from four perspectives, including electrochemical mechanisms, electrode design, synthesis, and performance; Refs. [38,42,46,47,48].
Coatings 13 01588 g002
Figure 3. (a) The formation energy of conversion products. (bf) Schematic molecular structures using in this work; Ref. [34]. (g) The cycling performance and (h) CV curve; Ref. [50]. (i) CoSe/G CV curve; Ref. [51]. (j) HRTEM for o–P–CoTe2/MXene and (k) after discharging to 1.0 V. (l) Ex situ XRD and (m) ex situ XPS Co 2p; Ref. [38].
Figure 3. (a) The formation energy of conversion products. (bf) Schematic molecular structures using in this work; Ref. [34]. (g) The cycling performance and (h) CV curve; Ref. [50]. (i) CoSe/G CV curve; Ref. [51]. (j) HRTEM for o–P–CoTe2/MXene and (k) after discharging to 1.0 V. (l) Ex situ XRD and (m) ex situ XPS Co 2p; Ref. [38].
Coatings 13 01588 g003
Figure 4. (a) In situ XRD patterns. (b) XPS of Co 2p and Te 3d at discharged state of (b1) 0.8 V, (b2) 0.4 V, and (b3) 0.1 V. (c,f) TEM, (d,g) HRTEM, and (e,h) EDS images at 0.1 V and 3 V, respectively. (i) Schematic diagram of the PIB storage mechanism of the CoTe2@NPCNFs@NC; Ref. [41]. CoTe2/G (j,k) SAED patterns, (l,m) HRTEM, and (n) ex situ XRD at 0.01 V and 2.8 V in the first cycle. (o) CV curve; Ref. [42]. (p) Ex situ XRD. (q) EXAFS after Li–embedded/Li–detached; Ref. [52].
Figure 4. (a) In situ XRD patterns. (b) XPS of Co 2p and Te 3d at discharged state of (b1) 0.8 V, (b2) 0.4 V, and (b3) 0.1 V. (c,f) TEM, (d,g) HRTEM, and (e,h) EDS images at 0.1 V and 3 V, respectively. (i) Schematic diagram of the PIB storage mechanism of the CoTe2@NPCNFs@NC; Ref. [41]. CoTe2/G (j,k) SAED patterns, (l,m) HRTEM, and (n) ex situ XRD at 0.01 V and 2.8 V in the first cycle. (o) CV curve; Ref. [42]. (p) Ex situ XRD. (q) EXAFS after Li–embedded/Li–detached; Ref. [52].
Coatings 13 01588 g004
Figure 5. XPS of Mo–CoSe2@NC with (a) survey and (b) Mo 3d. (c) HRTEM and (d) SAED; Ref. [68]. (e,f) Typical low– and high–magnification SEM (scanning electron microscope) images and (g) FeCo–Se XRD; Ref. [70]. (h) Li–ion batteries’ performance. (i) Schematic representation of the synthesis of Cu–Co1−xTe@NC HNBs with (jl) SEM and TEM images, (m,n) HRTEM images, and (o) SAED pattern and (pr) elemental mappings, Ref. [71].
Figure 5. XPS of Mo–CoSe2@NC with (a) survey and (b) Mo 3d. (c) HRTEM and (d) SAED; Ref. [68]. (e,f) Typical low– and high–magnification SEM (scanning electron microscope) images and (g) FeCo–Se XRD; Ref. [70]. (h) Li–ion batteries’ performance. (i) Schematic representation of the synthesis of Cu–Co1−xTe@NC HNBs with (jl) SEM and TEM images, (m,n) HRTEM images, and (o) SAED pattern and (pr) elemental mappings, Ref. [71].
Coatings 13 01588 g005
Figure 6. (ah) DFT in crystal structures and alkali–ion adsorption sites for (a,c,e,g) Co0.85Se and (b,d,f,h) S–doped Co0.85Se; Ref. [33]. (i) The differential charge density between CoSe2 and ZnSe in the phase boundary is calculated. (j) Averaged electrostatic potential. (k) The Na atom adsorption energy of ZnSe, CoSe2, and CoZn–Se; Ref. [72]. (lo) SEM, TEM, HRTEM, and EDS images, respectively, for a ZnTe/CoTe2@NC anode at 0.01 V and (ps) 3.0 V. (tv) Total density of states (TDOS) for CoTe2, ZnTe, and ZnTe/CoTe2, respectively; Ref. [40].
Figure 6. (ah) DFT in crystal structures and alkali–ion adsorption sites for (a,c,e,g) Co0.85Se and (b,d,f,h) S–doped Co0.85Se; Ref. [33]. (i) The differential charge density between CoSe2 and ZnSe in the phase boundary is calculated. (j) Averaged electrostatic potential. (k) The Na atom adsorption energy of ZnSe, CoSe2, and CoZn–Se; Ref. [72]. (lo) SEM, TEM, HRTEM, and EDS images, respectively, for a ZnTe/CoTe2@NC anode at 0.01 V and (ps) 3.0 V. (tv) Total density of states (TDOS) for CoTe2, ZnTe, and ZnTe/CoTe2, respectively; Ref. [40].
Coatings 13 01588 g006
Figure 7. (ac) SEM and (df) TEM images of PAN/PS/Co(ac)2·4H2O. (g,h) SEM and (i,j) TEM images of CoTe2@NMCNFs after 300 cycles at 200 mA g−1. Ref. [47].
Figure 7. (ac) SEM and (df) TEM images of PAN/PS/Co(ac)2·4H2O. (g,h) SEM and (i,j) TEM images of CoTe2@NMCNFs after 300 cycles at 200 mA g−1. Ref. [47].
Coatings 13 01588 g007
Figure 8. (A) Synthesis of FeCo–Se@NC and (BD) SEM images of the material in each step; Ref. [68].
Figure 8. (A) Synthesis of FeCo–Se@NC and (BD) SEM images of the material in each step; Ref. [68].
Coatings 13 01588 g008
Figure 9. Transition metals have various valence states and can form various complex structures, such as Fe–based MIL–53 with yolk–shell octahedron morphologies: (a) process of MIL–53(Fe) growth and (b) the formation process of Fe2O3–2 and Fe2O3–6; Ref. [115].
Figure 9. Transition metals have various valence states and can form various complex structures, such as Fe–based MIL–53 with yolk–shell octahedron morphologies: (a) process of MIL–53(Fe) growth and (b) the formation process of Fe2O3–2 and Fe2O3–6; Ref. [115].
Coatings 13 01588 g009
Figure 10. SEM images of In-MOF: Co (NO3)2 = (a) 1:8, (b) 1:6, (c) 1:4, and (d) 1:3; Ref. [39].
Figure 10. SEM images of In-MOF: Co (NO3)2 = (a) 1:8, (b) 1:6, (c) 1:4, and (d) 1:3; Ref. [39].
Coatings 13 01588 g010
Figure 11. (a) Preparation process. SEM images of (b) MXene, (c) h–CoTe2/MXene, and (d) o–P–CoTe2/MXene. (e,g) TEM and HRTEM images of o–P–CoTe2/MXene; (f) HRTEM of h–CoTe2/MXene. (h) EDS spectrum of o–P–CoTe2/MXene; Ref. [38].
Figure 11. (a) Preparation process. SEM images of (b) MXene, (c) h–CoTe2/MXene, and (d) o–P–CoTe2/MXene. (e,g) TEM and HRTEM images of o–P–CoTe2/MXene; (f) HRTEM of h–CoTe2/MXene. (h) EDS spectrum of o–P–CoTe2/MXene; Ref. [38].
Coatings 13 01588 g011
Figure 12. (af) SEM for CoTe2 synthesized with different NaOH concentrations: 0.0 M, 0.2 M, 0.5 M, 1.2 M, 1.5 M, and 2.0 M. Ref. [131].
Figure 12. (af) SEM for CoTe2 synthesized with different NaOH concentrations: 0.0 M, 0.2 M, 0.5 M, 1.2 M, 1.5 M, and 2.0 M. Ref. [131].
Coatings 13 01588 g012
Table 1. Review of the properties of CoSex electrode materials.
Table 1. Review of the properties of CoSex electrode materials.
Sample NameSynthesis MethodStructureCycle Performance (mAh g−1, A g−1)Rate Performance
(mAh g−1, A g−1)
Initial Coulombic Efficiency
(%, A g−1)
Battery TypeRef./Year
CoSe2/carbon nanoboxesAnnealingNanoboxes860/0.2/100 th
660/1/100 th
686/278.3%/0.2LIB[135]/2016
CoSe@NCNTsAnnealingNanowires326/0.5/100 th278/395%/0.1PIB[18]/2021
CoSe2-CNSAnnealingNanosheets~250/10/2000 th352/10 SIB[103]/2019
CoSe2@NC/CNTsAnnealingNanowires480/1/200 th
369/10/800 th
360/10 SIB[136]/2022
Mo–CoSe2@NCElectrodeposition
+ selenization
Self-supporting nanosheet672/0.5/200 th313/547.9%/0.5SIB[68]/2022
In2Se3/CoSe2–450AnnealingMIL-68@ZIF-67
core–shell
445.0/0.5/200 th
297.5/5/2000 th
205.5/10/2000 th
371.6/2061.2%/0.1SIB[39]/2021
TNC–CoSe2Precipitation + sol–gel + annealingMicrocubes511.2/0.2/200 th464/6.488.26%/0.2SIB[130]/2021
ZnSe/CoSe2–CNAnnealingSpherical
particles
547.1/0.5/300 th
422.6/3/300 th
362.1/2097.7%/0.1SIB[83]/2022
CoSe2@MXeneAnnealingPorous hollow shell
encapsulated
by MXene nanosheets
910/0.2/100 th
1279/1/1000 th
465/571.9%/0.2LIB[91]/2020
Ni0.47Co0.53Se2SolvothermalCoral-like321/2/2000 th277/1598.5%/0.1SIB[86]/2019
p–CoSeO3–CNFElectrospun
+ annealing
Composite porous nanofibers288/0.2/200 th207/556.7%/0.5SIB[100]/2021
CoSe2/CoSeAnnealingHierarchical hollow raspberry-like superstructure1361/1/1000 th
579/2/2000 th
315/5/1000 th
406/5 LIB[54]/2019
Ni0.33Co0.67Se2Solvothermal
selenization
Hierarchical mesoporous nanospheres301.9/1/1300 th314.5/885.4%/0.5SIB[137]/2023
CoSe2/ZnSeAnnealingBimetallic heterostructure~250/8/4000 th
~200/10/4000 th
263/1072.3%/0.1SIB[72]/2019
CoSSe@C/GSolvothermal
+ annealing
Double-carbon shells636.2/2/1400 th
353/1/200 th
208.1/0.2/100 th
254.5/10
266.6/10
195.7/2
68.09%/0.1
71.2%/0.1
48.3%/0.1
LIB
SIB
PIB
[33]/2020
NCNF@CSChemical vapor deposition + solvothermalOctahedral threaded N-doped carbon nanotubes253/0.2/100 th
173/2/600 th
196/269.3%/0.2PIB[34]/2018
3DG/CoSe2@CNWsHydrogelMultidimensional porous nanoarchitecture543/0.1/100
302/2/500 th
~320/263.5%/0.1SIB[30]/2022
CoSe2@NC/rGO-5Two-step
co-precipitation+ pyrolysis
Sandwich-like527.5/0.1/100 th
226/0.5/400 th
206/5
157/10
72.6%/0.1PIB[138]/2021
Table 2. Review of the properties of CoTex electrode materials.
Table 2. Review of the properties of CoTex electrode materials.
SampleSynthesis MethodStructureCycle Performance xRate Performance
(mAh g−1, A g−1)
Initial Coulombic Efficiency
(%, A g−1)
Battery TypeRef./Year
CoTe2–CAnnealingPolyhedron500/0.1/200 th
480/1/200 th
386/370.52%/0.1
56.07%/0.1
LIB
SIB
[52]/2020
CoTe2@3DGAnnealingPolyhedron + rGO191/0.05/350 th
103/1/4500 th
169/266.7%/0.05SIB[139]/2023
CoTe2@NMCNFsElectrospinning
+ annealing
Carbon fiber network261.2/0.2/300 th
143.5/2/4000 th
152.4/1057.1%/0.2SIB[47]/2021
CoTe2@3DPNCAnnealingDual-type carbon216.5/0.2/200 th164.2/587.6%/0.2SIB[140]/2022
o–P–CoTe2
/MXene
Solvothermal
annealing
MXene373.7/0.2/200 th
232.3/2/2000 th
168.2/2072.9%/0.05PIB[38]/2021
CTNRs/rGOOne-pot solvothermalNanorods/rGO306/0.05/100 th176/256%/0.05SIB[141]/2019
CoTe2–CSpray pyrolysisMicrosphere295.8/0.2/100 th163.7/269.2%/0.2PIB[142]/2020
Cu–Co1xSe2
@NC
Chemical etching
tellurization
Hollow nanoboxes796/1/800 th~400/5~80%/1LIB[71]/2022
CoTe2/GOne-pot solvothermalNanosheets356/0.05/100 th246/578.1%/0.05SIB[42]/2018
ZnTe/CoTe2
@NC
AnnealingBimetallic heterostructure317.5/0.5/1000 th
254.5/2/2000 th
165.2/5/5000 th
190.2/558.3%/0.1PIB[40]/2022
o/h–CoTe2One-step hydrothermalSubmicron-sized rods807/0.12/200 th
438/0.6/400 th
305.8/375.2%/0.12LIB[93]/2023
CoTe@NCDSimultaneous pyrolysis–tellurium melt impregnationCarbon dodecahedra300/0.1/100 th
200/0.1/100 th
207/2
58/2
57%/0.1
50%/0.05
SIB
PIB
[129]/2022
CoTe2@NPCNFs@NCElectrospinningN-doped porous carbon nanofibers409.1/0.05/50 th
198/0.5/600 th
120/2/1000 th
148.9/257.2%/0.05PIB[41]/2023
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Y.; Sun, Z.; Qu, D.; Han, D.; Niu, L. Recent Advances in CoSex and CoTex Anodes for Alkali-ion Batteries. Coatings 2023, 13, 1588. https://doi.org/10.3390/coatings13091588

AMA Style

Zhang Y, Sun Z, Qu D, Han D, Niu L. Recent Advances in CoSex and CoTex Anodes for Alkali-ion Batteries. Coatings. 2023; 13(9):1588. https://doi.org/10.3390/coatings13091588

Chicago/Turabian Style

Zhang, Yuqi, Zhonghui Sun, Dongyang Qu, Dongxue Han, and Li Niu. 2023. "Recent Advances in CoSex and CoTex Anodes for Alkali-ion Batteries" Coatings 13, no. 9: 1588. https://doi.org/10.3390/coatings13091588

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop