Home Journals Progress in Chemistry
Progress in Chemistry

Abbreviation (ISO4): Prog Chem      Editor in chief: Jincai ZHAO

About  /  Aim & scope  /  Editorial board  /  Indexed  /  Contact  / 
Review

Metal-Organic Frameworks and Their Derivative Nano Anode Materials

  • Haotian Ma 1, 2 ,
  • Rujin Tian , 1, * ,
  • Zhongsheng Wen 2
Expand
  • 1 College of Materials Science and Engineering,Dalian Jiaotong University,Dalian 116028, China
  • 2 Department of Materials,Dalian Maritime University,Dalian 116026, China
*Corresponding author e-mail:

Received date: 2023-05-04

  Revised date: 2023-09-19

  Online published: 2023-12-18

Abstract

Anode is one of the important components for lithium ion battery. Many technical bottlenecks (such as lower ionic-electronic conductivity, huge volume effect and easy pulverization resulted from long-term charge/discharge process) prevent the development and large scale application of traditional anode materials. As a novel kind of advanced multi-functional materials, Metal-organic frameworks (MOFs) and their derivative materials behave enough pore structures promoting rapid migration of Li+ and electron, and high specific surface areas providing abundant active sites for electrochemical reaction. Importantly, tunable structure and chemical composition of the MOFs and their derivative materials can be further optimized by changing parameters of synthesis process, thereby markedly increases specific capacity and cycle stability of lithium ion batteries. Herein, the recent progress in the MOFs and their derivative materials used as anode for lithium ion batteries are reviewed systematically, and the relationships between their preparation methods, microstructures, morphologies and corresponding electrochemical properties are summarized detailly. The urgent problems and challenges of this class of anode materials for lithium ion batteries are also analyzed. On the basis of resonable choosing organic ligands and metal centers, some effective measures for improving performances of lithium storage are proposed by combining with the variability and particularity of structure of the MOFs and their derivative materials, and the feasible strategies for commercialization application are suggested. Finally, the perspective and future development in design and fabrication of the new types of nano porous anodes with high energy efficiencies in relation with the next generation lithium ion battery are further discussed.

Contents

1 Introduction

1.1 Conversion mechanism

1.2 Insertion/extraction mechanism

1.3 Absorption/desorption mechanism

2 Pristine MOFs

2.1 Co-MOFs

2.2 Zn-MOFs

2.3 Mn-MOFs

2.4 Fe-MOFs

2.5 Ni-MOFs

2.6 Cu-MOFs

2.7 Sn-MOFs

2.8 Other metal-based MOFs

3 MOFs-derived metal compounds

3.1 Monometal oxides

3.2 Bimetal oxides

3.3 Other metal compounds

4 MOFs-derived porous carbon

5 MOFs-derived composites

5.1 MOFs/metal compounds

5.2 MOFs/carbon-based materials

5.3 metal oxide/Metal oxide

5.4 Metal oxide/carbon-based materials

5.5 Metal sulfide/carbon-based materials

5.6 Other metal compound/carbon-based materials

5.7 metal/metal oxide/carbon-based materials

6 Conclusion and outlook

Cite this article

Haotian Ma , Rujin Tian , Zhongsheng Wen . Metal-Organic Frameworks and Their Derivative Nano Anode Materials[J]. Progress in Chemistry, 2023 , 35(12) : 1807 -1846 . DOI: 10.7536/PC230502

1 Introduction

Energy crisis, environmental pollution and climate change have seriously hindered economic development and social progress. The research and development of new and efficient clean energy and energy storage equipment has become a key issue to be solved urgently. Therefore, the focus of attention and research in the field of energy is mainly on various electrochemical energy conversion technologies such as secondary batteries, fuel cells, solar cells and supercapacitors, especially lithium-ion batteries with the advantages of fast charging speed, high specific energy and long cycle life[1]. Lithium-ion batteries are composed of electrodes, electrolytes and separators, and energy storage is mainly achieved by reversible electrochemical reactions of electrode materials[2]. The theoretical capacity (372 mAh/G) of the traditional graphite anode is limited, and the rate capability is low[3,4]; On the contrary, the theoretical capacity of silicon anode (4200 mAh/G) is higher, and there are many disadvantages in the charge-discharge process, such as the decrease of structural stability due to volume change, the continuous growth of SEI (Solid Electrolyte Interphase) layer and the continuous consumption of Solid Electrolyte Interphase[5,6]. Therefore, the design and preparation of anode materials must meet the following requirements: on the one hand, to maintain good structural stability during long-term cycling, to alleviate volume and stress changes to a certain extent, and to avoid serious pulverization and structural collapse; On the other hand, good wettability and conductivity ensure sufficient active sites to participate in the electrochemical reaction. In other words, only the rational design, preparation and selection of high-performance anode materials can meet the requirements of lithium-ion batteries in terms of energy density, cycle life and safety.
MOFs are porous crystal materials with periodic structure formed by coordination chemistry, using metal ions/clusters as nodes and organic ligands as linkers. Due to the characteristics of both inorganic and organic materials, MOFs have obvious advantages in specific surface area, porosity, structural controllability and the number of active sites. For example, porosity provides more Li+ diffusion paths and intercalation, which is beneficial to the migration of Li+ and electrons. The special structure ensures the long-term stable cycle of the lithium-ion battery. For the first time, Li et al. Used MOF-177 directly as the anode of lithium-ion batteries, and the electrochemical performance was significantly degraded due to structural collapse and irreversible decomposition[7]. However, it is undeniable that this innovative work has pointed out the research direction and laid the research foundation for the subsequent exploration of MOFs and their derivative anode materials with high energy density and long cycle life. Transition metal compounds, porous carbons and their composites prepared by different methods, such as pyrolysis and annealing, using MOFs as precursors or templates, all belong to MOFs-derived porous nanomaterials, which retain the characteristics of MOFs bulk materials, such as controllable chemical composition, adjustable pore structure and large specific surface area. The lithium storage performance of MOFs and their derived anode materials mainly depends on their specific surface area, pore size, particle size and particle size distribution, which directly affect the electrode wettability, ion and electron transport and diffusion capabilities to a large extent, and play a decisive role in improving the electrochemical performance and stability of lithium-ion battery anode. The most effective and direct way to comprehensively improve the comprehensive performance of MOFs and their derived anode materials is to start with the selection of organic ligands, the design of spatial structure and micro-morphology, the optimization of process and the regulation of function.
In this paper, the lithium storage mechanism of MOFs is briefly described, and the research progress of MOFs bulk materials and their derivatives (metal oxides and metal sulfides, porous carbon and composites) as anode materials for lithium ion batteries in recent years is reviewed.The relationship between preparation methods, process parameters, nanoporous structure and electrochemical performance was analyzed, and the development direction and application prospect of MOFs and its derived nano-anode materials were prospected.
According to the electrochemical reactions occurring during charge-discharge cycles, the lithium storage mechanisms of MOFs materials are generally divided into conversion (Li+ replaces the central metal ion to form a metal element or metal oxide), intercalation/deintercalation (the structure of MOFs remains stable, and Li+ is intercalated/deintercalated in the transport pathway), and adsorption/desorption mechanisms.

1.1 Conversion mechanism

In the structure of MOFs, both redox metal ions and organic ligands can be used as active sites. The conversion mechanism is to use the valence change of the central metal to achieve effective lithium storage. According to whether the metal element is produced during the charge-discharge cycle, it can be divided into reversible and irreversible conversion mechanisms. In the process of reversible transformation, the central metal only changes its valence and does not produce simple metal, so the reversibility is obviously improved. In the process of irreversible conversion reaction, the formation of metal elements makes the structure of MOFs irreversibly destroyed, and even the electrode structure will be pulverized after only one charge-discharge cycle, which affects the capacity and coulombic efficiency of MOFs to a certain extent. Due to the decomposition and structural collapse of MOFs, the reversible capacity of most MOFs anode materials based on conversion mechanism (such as MOF-177 and Fe-MOF (MIL-88B)) after multiple cycles is only provided by the metal or metal oxide generated by the decomposition of MOFs, and the cycle stability is poor[7][8]. Therefore, the irreversible conversion reaction is the key factor for the capacity fading of MOFs anode, and the reversible conversion reaction of metal center is the source of electrode capacity stabilization.

1.2 De-intercalation mechanism

In the stable porous structure of some MOFs, organic ligands contain unsaturated bonds, and functional groups such as amino, carboxyl and benzene rings have the ability to store and transfer charges. In the electrochemical reaction process, energy is reversibly stored mainly by opening and reconnecting unsaturated chemical bonds.Different types of unsaturated chemical bonds are used to increase the number of active sites involved in electrochemical reactions, promote the rapid transmission and reversible intercalation/deintercalation of Li+, and improve the lithium storage capacity. Such as MIL-101 (Fe), Fe-MOF, and 2D/2D Ti3C2Tx/NiCo MOF heterostructures achieve effective lithium storage through this mechanism[9~11]. Therefore, it is necessary to rationally select organic ligands and optimize the structural design to obtain MOFs anode materials with high specific capacity and cycle stability.

1.3 Adsorption/desorption mechanism

The adsorption/desorption mechanism relies on the electrostatic interaction of the electrode surface to store energy, rather than chemical bonding, and the capacity is usually very low. Although there are few examples of reversible lithium storage using adsorption/desorption mechanism (such as Na+storage in hard carbon), it should be paid enough attention to porous MOFs electrode materials with high specific surface area[12].

2 Bulk materials of MOFs

The basic components of MOFs materials are metal ions (or metal ion clusters) and organic ligands, which mainly rely on the large number of active sites to promote redox reaction kinetics, and use the unique porosity to maintain structural stability and promote the rapid migration and diffusion of Li+ to obtain good comprehensive performance. MOFs as electrode materials only need to meet one of two conditions: one is that they do not contain Li element, and need to use their own metal ions or the active functional groups of organic ligands in the framework structure to carry out redox reactions to achieve the insertion/extraction of Li+; The other is that it contains Li element itself, and Li+ has synergistic effect with other metal ions in the process of redox reaction. MOF-177(Zn4O(BTB)2) is the first MOFs bulk material directly used as the anode of lithium-ion batteries, which has low electrochemical performance and can not meet the needs of practical applications[7]. The initial specific capacity of some MOFs synthesized on this basis is very low (about 100 ~ 350 mAh/G) at low current density (20 ~ 50 mA/G), and the specific capacity decays rapidly after a certain cycle[9,13~18]. Fig. 1 shows the important time nodes in the development of MOFs and their derived anode materials.
图1 锂离子电池MOFs及其衍生负极材料的发展时间轴

Fig. 1 Brief summary of the development history of MOFs and their derivative anode materials for Li-ion batteries

MOFs are usually hollow, spherical and polyhedral, and contain transition metal ions such as Zn and Ni or metal ion clusters, which can be used as redox active substances in electrochemical reactions. The morphology and structure of MOFs can be controlled and optimized by changing the reaction temperature, adjusting the ratio of solvent to reactant, and adding surfactant.

2.1 Co-MOFs

Co-MOFs are one of the more widely studied anode materials for lithium-ion batteries. Li et al. Prepared optimized CoBTC-EtOH hollow microspheres by hydrothermal method[19]. The reversible capacity of CoBTC-EtOH anode is 856 mAh/G after 100 cycles at 100 mA/G. After 500 cycles at a high current density of 2 A/G, it is still 473 mAh/G. Although the Li+ can be intercalated into the organic part of Co-BTC, the Co2+ always remains ionic and does not directly participate in the charge-discharge cycle. Similarly, after 200 cycles at 100 mA/G, the capacity of the conchoidal Co2(OH)2BDC(S-Co-MOF) is as high as 1021 mAh/G; The capacity is still 435 mAh/G after 1000 cycles at 1 A/G[10]. This good cycling stability and outstanding lithium storage capacity depend on the organic moiety-dominated intercalation/deintercalation mechanism to ensure the efficient transport of Li+. In addition, the contribution of pseudocapacitive characteristics to the Li+ intercalation/deintercalation ability is also important, that is, the pores of S-Co-MOF absorb an appropriate amount of Li+, resulting in a higher capacity than the theoretical value at low rates. Compared with Co-MOFs, the micro-flower-shaped H-Co-MOF anode composed of nanosheets has high pore volume and specific surface area, which can provide sufficient lithium storage sites, and can also further improve the performance with the help of the ion buffer zone in the interface layer[20]. Obviously, the hierarchical structure is an important factor affecting the performance of most Co-MOFs anodes, which rely on a single metal or ligand to participate in electrochemical reactions[21]. Ning et al. Synthesized morphology-controlled Co-MOFs (u-CoTDA) ultrathin nanosheets by ultrasonic method, and the capacity, rate capability and long-term cycling stability of this anode material not only depend on the characteristics of more active sites for Li+ intercalation/deintercalation, less diffusion barriers and lower mechanical strain,Ultrathin nanosheets can also increase the proportion of coordinatively unsaturated metal active sites exposed on the surface and improve the redox activity of metal centers, that is, the ligands of Co (II) and u-CoTDA participate in the redox process at the same time[22]. Therefore, it is feasible to rationally design nanosized MOFs to shorten the ion/electron diffusion distance and improve the diffusion kinetics.
In addition, because the coordination compound is easy to decompose during charge-discharge cycles, the construction of one-dimensional chain or multi-dimensional framework structure can effectively improve its solid-state stability and capacity reversibility. The metal ions and the bridging ligands are not connected by π-π stacking interaction or covalent bonds, but are coordinated by weak hydrogen bonds to form a one-dimensional chain structure, which does not affect the size of the Li+ diffusion channel of the electrode material and the dissolution of the active organic in the electrolyte, and the capacity is improved mainly through the combination of two active elements, and the coulombic efficiency is up to 98.3%[23]. It can be seen that the participation of aromatic chelating ligands and metal centers is a new mechanism for lithium storage in coordination compounds. Co2(DOBDC)(Co-MOFs) anode was prepared by a modified hydrothermal method, and most of the capacity was derived from the pseudocapacitive characteristics including surface or near-surface redox reactions and a large number of ion intercalation, rather than the diffusion-controlled faradaic effect, which may be related to the microporous structure characteristics and high specific surface area[24]. The reversible capacity at 2 A/G does not decrease significantly and remains at 408.2 mAh/G. Due to the synergistic effect of metal centers and organic ligands and the uniform nanowire morphology during cycling, the carboxylic acid-based one-dimensional coordination polymer (Co-BDCN) anode not only has a high capacity of 1132 mAh/G after 100 cycles at 100 mA/G, but also has excellent cycle stability[25]. Depending on the metal-ligand coordination mechanism and intermolecular interactions, metal ions (or metal ion clusters) and organic bridging ligands can also assemble into two-dimensional or three-dimensional frameworks. In the serrated worm-like nano-Co-BTC, the chain structure effectively promotes the reversible charge transfer, and the conjugated carboxylate has a strong π-π interaction, which makes the capacity, cycle stability and coulombic efficiency outstanding, but the internal structure transformation and electrochemical mechanism still need to be further explored[26]. CoC8H4O4(CoTPA) is a kind of lamellar MOFs containing carbonyl groups, and both Co2+ and organic moieties containing carbonyl and aromatic structures possess electrochemical active sites, while multi-electron transfer reactions provide high capacity and good capacity retention[27]. This is inconsistent with the lithium storage mechanism of common MOFs anode. The three-dimensional Co2(OH)2BDC with high purity and crystallinity prepared by Gou et al. By solvothermal method is superior to other known BDC-based anode materials in terms of specific capacity, cycle stability and rate capability, especially the effect of solvent ratio on the purity of the product is obvious[28]. Coordination polymers belong to weak framework structures, which absorb relatively less solvent and are more prone to reversible transformation reactions. Fei et al. Used a simple template-free method to synthesize CoC6H2O5(H2O)2 hollow microspheres[29]. Under the combined action of the crystal structure, the reversible conversion reaction and the oxygen atom in the aromatic ring, the cycle stability is obviously higher than that of the non-hollow microsphere.

2.2 Zn-MOFs

In terms of structure, size, compositional functionality, permanent porosity and specific surface area, MOFs are promising electrode materials, but their capacity and cycle stability are not ideal. For example, the initial discharge capacity of MOF-177(Zn4O(1,3,5-benzenetribenzoate)2) is 400 mAh/G, and after the first discharge, metal zinc and lithium salts are irreversibly generated, and the charge-discharge cycle is realized through the reversible alloying of zinc and lithium, but the capacity decays rapidly due to the structural collapse in the subsequent cycle process, so it is not suitable as an anode material for lithium-ion batteries[7]. The formate-bridged rhombic structure MOFs(Zn3(HCOOH)6) reacts reversibly with Li during cycling by relying on a turnover mechanism[30]. Compared with Li2O, the thermodynamic feasibility of lithium formate to transition metal formate is better, and the matrix involved in the cycling process is lithium formate rather than the typical Li2O. The reversible formation or regeneration of FOR1 MOF is the key to obtain good lithium storage ability, and its electrochemical performance is significantly better than that of MOF-177. On the one hand, the post-modification of hydrophobic functional framework can be used to construct stable MOFs. On the other hand, internal pores with polar functions can chemically capture and strongly interact with guest atoms or molecules. Lin et al. Designed and synthesized BMOF(Zn(IM)1.5(abIM)0.5) with outstanding chemical and thermal stability, which can store lithium reversibly mainly through adsorption mechanism and functionalized pores.At the same time, the host-guest interaction between Li and amine functional groups/N atoms in imidazole can not be ignored, that is, the lithium storage capacity can be further improved by increasing the number of active N-rich functional sites and the ratio of surface area/pore volume[31]. In addition, the effects of the molar ratio of the initial compounds, reaction temperature and time on the final product were also systematically studied. Therefore, the application of Zn-MOFs anode is largely limited by its low conductivity and capacity, as well as poor cycling stability.

2.3 Mn-MOFs

The characteristics of low cost, low toxicity, safety, and environmental friendliness make Mn-MOFs have great potential for energy storage applications. In the [Mn(tfbdc)(4,4'-bpy)(H2O)2](Mn-LCP) of the two-dimensional microporous structure, the Mn (II) is weakly antiferromagnetically coupled[32]. Although the ideal capacity is not obtained due to the conversion mechanism, the electrochemical performance may be improved by self-assembling variable valence metal ions and low molecular weight organic ligands, which provides a new idea for the preparation of MOF anode materials. Using surfactants to control the shape, growth, and initial nucleation stability of building units, Fei et al. Prepared Mn2(C6H2O4S)·2H2O microspheres using a soft template method[33]. The results show that the reversible capacity is 645.7 mAh/G after 250 cycles at 400 mA/G. Coulombic efficiency approaches 100% even after 650 cycles at 500 mA/G. Compared with MOF materials with irregular shapes, the new microsphere anode has better cycle stability and is very potential for lithium-ion batteries. Conjugated carboxylates are weak electron-attracting ligands with strong π-π interactions, which act as redox centers to coordinate with lithium and can maintain structural integrity and crystallinity in lithium-ion batteries. Maiti et al. Synthesized Mn-BTC MOF with coexistence of structural pores and intergranular pores by solvothermal method[34]. The specific capacity of the anode material is 694 mAh/G at 100 mA/G; It also shows good rate capability when the potential is lower than 2. 0 V. Li+ intercalates into the organic part of carboxylic acid group and benzene ring in Mn-BTC MOF, and the electron-donating effect is the main driving force for lithium storage, and its redox reaction does not belong to the conversion mechanism. Xiong et al. Synthesized uniform single crystal Mn-PBA(Mn[Fe(CN)6]0.6667·nH2O) cubes at room temperature[35]. In addition to the inherent three-dimensional framework structure, which is conducive to Li+ transport, the larger free space can accommodate transition metal ions and some small molecules, and also maintain the charge balance and redox reaction through the change of iron valence. Therefore, the capacity, rate capability, cycle stability and coulombic efficiency are all outstanding. Mn-1,4-BDC @ 200 was prepared from Mn-1,4-BDC MOF by heat treatment (200 ℃ for 24 H). Both of them have the same characteristic bands and framework structure, but the morphology is changed, that is, the loose and uniform lamellar structure is split into small pieces[36]. Reinsch and Stock and Li et al. Considered that the crystallinity and reversible intercalation/deintercalation of Li+ of MOFs are closely related to the coordination solvent in the pores[37][19]. Therefore, the reversible capacity of Mn-1,4-BDC @ 200 anode is as high as 974 mAh/G after 100 cycles at 100 mA/G, further demonstrating the positive effect of removing the coordinating solvent molecules on improving the lithium storage performance. Mn-MOF(Mn2(OH)2BDC) is composed of randomly arranged close-packed ultrathin nanosheets, which has the advantages of low metal ion radius and high specific surface area[38]. During the charge-discharge cycle, the connection between some BDC2- and surface metal ions is interrupted, resulting in the formation of coordinatively unsaturated metal sites on the exposed surface, and the metal redox activity is improved, and the central metal ions and aromatic ligands BDC2- participate in lithium storage. Under the combined effect of these factors, the reversible capacity, rate capability and Li+ diffusion coefficient of the anode material are high.

2.4 Fe-MOFs

F F Érey et al. First used MIL-53 (Fe) as an anode material for lithium-ion batteries, with a capacity of only 75 mAh/G and unsatisfactory cycle stability[39]. Benzoquinone can absorb electrochemically active molecules and act as a redox mediator to accelerate electron transfer and increase the initial capacity of MIL-53 (Fe), but the exchange of benzoquinone and electrolyte molecules will cause capacity fading in subsequent cycles[40]. In addition, the reversibility of electrochemical cycling of MIL-68 (Fe) is not as good as that of MIL-53 (Fe) because the pore shape is not conducive to Li+ diffusion[41]. Under the condition of no precursor and pH adjustment, Hu et al. Reasonably selected iron source and directly prepared nano-sized Fe-BTC MOF with carboxylate as ligand[42]. The electrochemical performance of the Fe-BTC MOF anode, especially the rate capability and coulombic efficiency, which are closely related to practical applications, is quite outstanding due to the heterogeneous catalytic oxidation and stable structure during cycling. In addition, the flexible structure MIL-88A synthesized by solvothermal method belongs to the MIL series of Fe-MOFs materials[43]. Restricted by its poor conductivity and the formation of SEI layer, it is not conducive to Li+ intercalation/deintercalation, and the initial discharge specific capacity is 140. 5 mAh/G. Under the action of external conditions such as guest molecules, the flexible framework structure is easy to cause the expansion and contraction of pores, and some electrolytes and electrode materials are continuously consumed to produce irreversible by-products (lithium hydroxide, lithium oxide and lithium carbonate), which lead to rapid capacity decay in the subsequent cycle process. There is no doubt that improving the reversibility of Li+ intercalation/deintercalation process while improving the conductivity of nanostructured materials is an effective way to further improve the performance of electrode materials. Shen et al. Designed and synthesized Fe-MOF (Fe-MIL-88B) nanorods with polyhedral structure, and measured the half-cell and full-cell performance of their composition[44]. The results show that the higher capacity and cycling stability mainly depend on the transition metal clusters and organic ligands in Fe-MIL-88B anode, which can help to better understand the lithiation/delithiation behavior of small organic molecules in advanced electrode materials. Shin et al. Suggested that the capacity fading of MIL-101 (Fe) anode materials is not because of the collapse of the framework structure, but depends on the time-dependent irreversible oxidation (Fe2+→Fe3+), that is, the Fe2+→Fe3+ relaxation promotes the irreversible accumulation of Li in MIL-101 (Fe)[9]. Obviously, precise control of the electronic structure can effectively improve the reversibility of MOFs electrode materials. The morphology of Fe (Zn) -BDC @ 300 was improved, the specific surface area increased and the charge transfer resistance decreased under the action of the structure and morphology directing agent Zn(NO3)2, but the effect of Zn content was negligible. Therefore, the capacity and cycling stability of this anode material are higher[45].

2.5 Ni-MOFs

Li et al. Synthesized waffle-like Ni-MOF(Ni2(OH)2BDC) by ultrasonic method[38]. However, the large atomic radius and steric effect of Ni, the low specific surface area, the lack of unsaturated redox active sites and the small control of surface capacitance are not conducive to the migration and diffusion of electrons/ions, which makes it difficult to maintain the original nanostructure, and the reversible capacity and rate performance are not ideal. On the basis of choosing low-cost terephthalic acid as the ligand, the Ni-MOF synthesized by Zhang et al. By one-step hydrothermal method has a reversible specific capacity of 620 mAh/G after 100 cycles at 100 mAh/G, and has a high rate performance, but the mechanism of lithium storage and its influencing factors have not been studied in detail[46]. Compared with carboxylate and/or pyridine ligands, MOFs bearing azo groups are chemically and thermally stable[47]. The H2Me4bpz ligand linked to Ni2+ provides a strong support for the structural flexibility and stability of Ni-MOF(Ni-H2Me4bpz), and the volume change in the process of Li+ deintercalation is alleviated to some extent, and the original crystal structure is retained[14]. Therefore, the specific capacity and cycling performance of the 2D ripple hierarchical structure Ni-H2Me4bpz porous anode are very outstanding. It can be seen that although Ni-MOFs materials can effectively alleviate the volume effect and improve the ion transport kinetics, they have obvious shortcomings such as insufficient thermal stability and low charge/discharge potential, and there is still a certain gap in meeting the practical application.

2.6 Cu-MOFs

Cu-MOFs contain abundant active components for redox reactions, and the capacity reduction first occurs at the cluster (Cu/Cu2+), which may be related to the incomplete redox and insufficient structural stability of Cu2+ located at the metal cluster site. The high specific surface area Cu-BDC MOF([Cu2(C8H4O4)4]n) porous anode was prepared based on the optimized in situ electrochemical synthesis conditions, and the first cycle discharge capacity was 227 mAh/G, which was about 95% of the theoretical capacity. Compared with Zn4O(1,3,5-tribenzoate)2, the rate capability of MOFs anode with terephthalic acid ligand is higher[48]. However, under the influence of a large number of pores, amorphization reaction and side reaction between Li+ and electrolyte, the initial charge/discharge capacity has a large difference[7,30]. Among the alkali metal ions, Li+ binds to MOFs with the highest affinity, but the number of binding sites provided and the binding mode mainly depend on the coordination tendency with the central metal ion and the Li+ content. In Cu3(BTC)2, coordinatively unsaturated metal ions can bind to small molecules. The Cu3(BTC)2 anode prepared by solvothermal method for the first time has a reversible capacity of 740 mAh/G at 96 mA/G. The capacity retention was nearly 100% after 50 cycles at 383 mA/G[49]. The results show that the charge storage mechanism may be related to the redox reaction of carboxylate ligands, which can not be explained by the conventional conversion mechanism. Hu Xiaoshi prepared the [Cu2(cit)(H2O)2]n anode material based on the modified hydrothermal method, and the reversible capacity (950 mAh/G) and cycling stability were higher than those of the previously reported Cu-MOFs[50]. After the first discharge, the Cu2+ of the metal center is almost completely converted to elemental Cu, and the reversibility is gradually enhanced during the charge process, and Cu is oxidized to CuxO again, resulting in an abnormal increase in capacity.

2.7 Other metal-based MOFs

In addition to the above transition metal MOFs, researchers have also made useful attempts on other metal-based MOFs anode materials. Wu et al. First synthesized Sn-MOF(C24H16Li2O14Sn) polyhedra composed of Sn2+ and phthalic acid anions by one-step reflux method[51]. Due to the effective mitigation of structural collapse caused by volume expansion/contraction during cycling and the maintenance of structural stability, this Sn-MOF anode exhibits a high capacity (610 mAh/G) and cycling stability, with a capacity retention of more than 92% even after 200 cycles. On this basis, Wu et al. Further compared the effects of preparation methods on the morphology and lithium storage performance of Sn-MOF anode[52]. Different from the hydrothermal method, the Sn-MOF prepared by the reflux method has more uniform morphology, smaller particle size, larger specific surface area and lower charge exchange resistance, which can provide higher lithium storage activity and faster Li+ diffusion kinetics, and the reversible capacity and coulombic efficiency are correspondingly higher. The agaric-like hierarchical Al-FumA MOFs anode is composed of intertwined ultrathin nanosheets, with a reversible capacity of 392 mAh/G at 37.5 mA/G and no capacity degradation after 100 cycles. The capacity is 258 mAh/G at 37.5 A/G[53]. Low impedance, fast and reversible deintercalation/intercalation of Li+, and structural stability are important factors that determine the good lithium storage performance of Al-FumA MOFs. In addition, whether the framework structure is beneficial to Li+ transport and the organic part plays a decisive role in the lithium storage mechanism, or the high thermal stability and specific surface area are beneficial to the reversible migration and storage of Li+, the three-dimensional porous Pb-MOF or Cd-MOF anode materials synthesized by solvothermal method show good cycle stability and rate performance[16,18]. The new Ti-MOF anode has an initial discharge capacity of 1590.24 mAh/G at 100 mA/G, and especially maintains good reversibility after 8000 cycles[54]. Density functional theory calculations prove that the carboxyl functional group acts as an electrochemical active site to reversibly bind to Li+, and the Ti octahedron acts as a stabilizing framework. This is attributed to the stable porous structure and sufficient redox sites of Ti-MOF, but the whole electrochemical process is mainly controlled by the diffusion of Li+, which leads to the decrease of rate capability, and it is necessary to increase the pore size and select small molecular electrolyte solvents to improve the capacity and pseudocapacitive behavior.
A large number of studies have shown that monometallic MOFs are generally difficult to meet the requirements of cyclicity and structural stability simultaneously. Bimetallic MOFs or metal-based MOFs mixtures can exert the synergistic effect between different metal ions to obtain more excellent comprehensive properties. For example, Han et al. Prepared Li/Ni-NTC anode materials by rheological phase reaction, which combined the respective advantages of Ni-NTC and Li-NTC in conductivity and structural stability, and its electrochemical capacity and cycle stability were significantly improved, providing an important basis for the development of a new generation of MOFs anode materials by designing and optimizing the molecular structure of aromatic derivatives[55]. Saravanan et al. Synthesized Zn1.5Co1.5(HCO2)6 with Zn2+ instead of Co2+, which significantly reduced the amount of expensive and toxic Co2+, improved the capacity and cycle stability of electrode materials under the combined action of conversion and alloying/dealloying reactions and formate, and provided a simple and friendly synthesis method for the preparation of MOFs electrode materials with long life and high thermal stability[30].
Some literature reported bulk anode materials of MOFs and their electrochemical performances are summarized in Table 1. Although some progress has been made in the research of MOFs bulk materials directly used as negative electrodes for lithium-ion batteries, there are still some problems to be solved. On the one hand, metal-based MOFs themselves are not conductive enough. On the other hand, the structural controllability of metal-based MOFs is poor due to the constraints of preparation methods. In most cases, only a single metal or organic ligand participates in the electrochemical reaction, which can easily lead to the structural collapse of metal-based MOFs. These factors directly or indirectly affect the reversible capacity, cycle stability and rate capability. Therefore, in the preparation of MOFs anode materials, not only the main structure should be kept basically unchanged, but also ligands with active functional groups (such as carboxyl, amino, benzene ring, etc.) And variable valence metals should be selected to increase the number of active sites for lithium storage and theoretical specific capacity, and the porous framework structure should be fully utilized to accelerate Li+ transport and reversible intercalation/deintercalation.
表1 MOFs本体负极材料及其电化学性能

Table 1 Pristine MOFs as anode materials and their electrochemical performances

Materials Voltage range (V) Current density (mA/g) Cycle number Overall capacity (mAh/g) Initial discharge/
charge capacity
(mAh/g)
Initial coulomb
efficiency
(%)
Specific surface area
(m2/g)
ref
CoBTC-EtOH 0.01-3.0 100/2000 100/500 856/473 1790.3/879 49.15 17.7 19
S-Co-MOF 0.01-3.0 100/500/1000 200/700/1000 1021/601/435 1964/1564 80.4 10.4 10
H-Co-MOF 100/2000 100/700 1345/828 2147/1432 66.7 49.9 20
u-CoTDA 0.01-3.0 100/1000/2000 100/300/400 946/790/548 1631/- 75.2 52.6 22
[Co1.5L(H2O)4]n 0.01-3.0 50 50 431 1978/869 23
Co2(DOBDC) 0.01-3.0 100/500 100/200 878.5/526.1 1409/785 56 24
Co-BDCN 0.01-3.0 100 100 1132 1439/1015 70.54 24.5 25
Co-BTC 0.01-3.0 100 200 750 1739/622 36 18.5 26
CoTPA 0.005-2.8 60 100 700 1938/1004 51.8 27
Co2(OH)2BDC 0.02-3.0 50 100 650 1385/1005 72.8 28
CoC6H2O5(H2O)2 0.05-3.0 100/1250 95/499 549.8/513.4 -/- 29
Zn3(HCOOH)6 0.005-3.0 60 60 560 1344/693 30
BMOF 100 200 190 -/- 821 31
Mn-LCP 0.01-2.5 50 50 390 1807/- 32
CMPS-1 0.05-3.0 400/500 250/650 645.7/588.3 1631.8/- 33
Mn-BTC 0.01-2.0 103/1030/2060 100/100/100 694/400/250 1717/694 40.4 23.8 34
Mn-PBA 0.01-3.0 200 100 295.7 1123.7/544.5 48.5 499.8 35
Mn-1,4-BDC@200 0.01-3.0 100 100 974 1746/706.4 40.5 6.135 36
Mn-UMOFNs 0.01-3.0 100/1000 100/300 1187/818 -/- 57 32.65 38
Fe-BTC 0.01-3.0 100 100 1021.5 1765.5/683.2 38.7 1125 42
MIL-88A 0.01-3.0 10 4 40.5 140.5/5.3 4 43
Fe-MIL-88B 0.005-3.0 60 400 744.5 1507/949.9 63 44
Fe-BDC@300 0.01-3.0 100 120 324.1 1330.6/- 45
Ni-UMOFNs 0.01-3.0 100 100 546 1833/1226 67 15.04 38
Ni-MOF 0.01-3.0 100 100 620 1984/1369 46
Ni-Me4bpz 0.01-3.0 50 100 120 320/- 67 14
[Cu2(C8H4O4)4]n 0.01-2.5 48 50 161 1492/194 747 48
Cu3(BTC)2 0.05-3.0 96/191/383 50/50/50 740/644/474 1497/641 489.4 49
[Cu2(cit)(H2O)2]n 0.01-3.0 100/2000 500/500 608.5/321.5 -/- 50
Sn-MOF 0-3.0 20 200 610 1017/450 67.437 51
Sn-MOF 0.01-3.0 50 100 250 -/- 16.96 52
Al-FumA MOFs 0.01-3.0 37.5/37500 100/100 392/258 1509/899 45.5 260.1 53
Pb-MOF 0.01-3.0 100/500 500/500 489/380 1522/678 38 725 16
Cd-MOF 0.1-3.0 100 100 302 710/435 821 18
Ti-MOF 0.01-3.0 200/400 200/500 296/175.34 1590.24/- 621 54
Li/Ni-NTC 0.01-3.0 100 80 482 1084/601 55
Zn1.5Co1.5(HCO2)6 0.005-3.0 60 60 510 1344/693 30
Conductive MOF materials have high conductivity, specific surface area and abundant active sites. Compared with other traditional carbon materials and metal oxide materials, the pore structure has more obvious advantages. Cu-CAT nanowires synthesized by solvothermal method show high Li+ diffusivity and conductivity[56]. The specific capacity of the anode material is about 631 mAh/G at a current density of 0. 2 A/G, even close to 381 mAh/G at a high rate of 2 A/G, and the capacity retention rate is about 81% after 500 cycles at 5 A/G. In particular, the energy density of the Cu-CAT nanowire-based full cell is as high as 275 Wh/kg. Similarly, the diffusion coefficient of Li+ in the deintercalation process of one-dimensional Ni-CAT nanorods synthesized by hydrothermal method reaches 10-9~10-10cm2/s, and the discharge capacity is about 889 mAh/G at the current density of 0.1 A/G, and its rate capability and cycle stability are also quite outstanding, as shown in Fig. 2[57]. A two-dimensional conducting MOF (Cu-HHTQ) facilitates the Li+ and electron transport and ensures the resilience of the electrode, and has a discharge capacity of about 657. 6 mAh/G at 0. 6 a/G[58]. Half-cells and full-cells constructed with Co-CAT anode materials also exhibit excellent lithium storage performance, which is mainly attributed to the conductive nature of the MOF structure contributing to charge migration, electrolyte penetration, and regulation of volume change[59]. These studies show that conductive MOF is a very potential anode material, and it is necessary to further select low-cost organic ligands and improve product stability to accelerate the application and development of this new advanced energy storage material.
图2 (a)CV曲线(0.1mV/s), (b) 潜在储锂位点 (Ⅰ 苯环; Ⅱ孔隙; Ⅲ层间空间), (c) 初始3次充放电曲线(0.2 A/g), (d) 倍率性能, (e)与其他负极在不同电流密度时的容量,(f) 0.2和0.5A/g时Ni-CAT NRs的循环性能[57]

Fig. 2 (a) CV curves (0.1 mV/s), (b) potential lithium-storage sites (Ⅰ, benzene rings; Ⅱ, pores; Ⅲ, interlaminar space), (c) initial three charge and discharge plots (0.2 A/g), (d) rate behaviour, (e) comparison of capacities with other anodes at various current densities, and (f) cycling properties at 0.2 and 0.5 A/g of the Ni-CAT NRs[57]

Considering the limited lithium storage capacity of MOFs bulk anode materials, MOFs derivatives have become a hot research topic in recent years. These derivatives are new functional materials prepared by different synthetic methods using MOFs as templates or precursors, mainly including metal oxides, porous carbon, other metal compounds and nanocomposites synthesized on the basis of them. Due to the high controllability of synthesis conditions and chemical reaction system, the product composition is more uniform, and the structure and morphology are more diverse, these MOFs derivatives and their composites can play a more advantageous role and obtain the expected performance when used as anode materials for lithium-ion batteries.

3 MOFs derived metal compound

Most of the traditional transition metal oxides (Cu, Mn, Fe, Co, Ni, etc.) Have the outstanding advantages of high theoretical specific capacity, chemical stability, low cost, abundant resources and environmental friendliness.However, it is difficult to replace graphite as a high-performance anode material because of its obvious volume change, serious ion agglomeration and insufficient internal conductivity during long-term charge-discharge cycles, which lead to serious pulverization, capacity fading, poor cycle stability and short cycle life of electrode materials. As mentioned above, MOFs materials have complex porous structures and optional central metal ions, but their low conductivity seriously restricts their direct use as anode materials. It is a feasible method to use MOFs as templates or precursors to convert them into new metal oxides by controlling the synthesis conditions. Metal ions immobilized in MOFs are converted into the corresponding oxides during the synthesis process, and other elements (such as carbon, nitrogen, and hydrogen) are oxidized into gases to escape, forming interconnected pores and porous shells within the micro/nano structure. According to the difference of reaction mechanism, these preparation methods include self-pyrolysis in inert atmosphere, chemical reaction with solution or gas (steam), and chemical etching. Novel porous metal oxides with controlled structure and composition can be obtained without additional templates or complicated synthesis procedures. Therefore, the key to improve the electrochemical performance lies in the uniform distribution of metal ions/clusters through the structure and morphology control of transition metal oxides, which mainly depends on the synthesis process, the type of metal ions/clusters, the morphology and size of functional groups, and the structural characteristics of MOFs. It can be seen that functional porous MOFs-derived metal oxides with hierarchical or hollow micro/nanostructures have great potential in improving the specific capacity and prolonging the cycle life compared with other metal oxides with non-porous structures or low surface areas[60].
Common MOFs-derived metal oxides (TMOs) include monometallic oxides and bimetallic oxides, and the lithium storage capacity mainly depends on factors such as the type of metal ion/cluster and the inherent framework structure, morphology, and pore size.

3.1 Monometallic oxide

It is well known that the advantage of metal oxide electrode materials lies in the chemical transformation during charge and discharge, that is, one or more electrons exchange and replace the intercalated Li+. The research on monometallic oxides derived from MOFs as precursors or templates started earlier, and the synthesis methods are relatively mature, and the process routes are simple and easy to control. In particular, low-density MOFs are often accompanied by significant volume shrinkage during transformation, which is more suitable for the synthesis of hollow nano-functional materials. Common MOFs-derived mono-oxide metal oxide anode materials include NiO, Co3O4, and CuO, etc., and the effect of metal type on their electrochemical performance is particularly obvious.
The reversible capacity of CuO nanoparticles prepared with copper phenylalanine as the precursor is 505. 3 mAh/G after 200 cycles at 100 mA/G, but the capacity is only 111. 6 mAh/G at 1000 mA/G, which shows that the rate capability is low[61]. Wu et al. Synthesized porous CuO hollow octahedra using Cu-MOF (Cu-btc) as a template[62]. In addition to the contribution of the porous hollow structure, wide pore size distribution and appropriate specific surface area (49.6 m2/g) to the cycling stability and rate capability, the { 111 } facet is more favorable for Li+ transport than other facets, that is, the facet structure also affects the electrode performance[63]. Although the reversible capacity (538 mA · H/G) of nanostructured spherical CuO synthesized based on pyrolytic Cu-MOF (MOF-199) is significantly better than the reported results, further regulation and optimization of the process are needed to meet the requirements of anode materials[64]. Obviously, the morphology and porosity of these copper oxides change with the subtle changes of MOF template and pyrolysis process, and the power density or cycle life is very limited. In addition, due to the retention of the morphology and structural characteristics of Cu-BTC, the stability of the main structure is not destroyed during cycling, and the charge storage process controlled by the surface redox pseudocapacitive behavior makes the CuO octahedral anode show ultra-high capacity and rate performance.This pseudocapacitive property is related to the ultrathin fine two-dimensional nanosheet subunit, while the hierarchical porous structure is a prerequisite for the extrinsic pseudocapacitance in the nanosheet subunit[65]. More importantly, this room temperature solid-state transformation method significantly reduces energy consumption and avoids the release of toxic gases, which is a green synthesis strategy to accelerate the development of micro/nano-structured metal oxide anodes.
The electrode potential of Mn2O3 is low and the energy density is high, but the low electrochemical activity and high volume change in the process of Li+ intercalation/deintercalation directly affect its use as an anode material. Hierarchical structure Mn2O3 porous octahedron is composed of nano-sized units, which can reduce the interfacial contact resistance and avoid agglomeration during cycling, and the higher specific surface area and internal volume in the mesoporous structure provide more active sites and shorten the Li + intercalation/deintercalation diffusion distance[66]. Affected by the formation of SEI layer and electrolyte decomposition, the Mn-LCP template-based Mn2O3 porous anode has a low initial coulombic efficiency (74%), but it has an advantage over micron-sized Mn2O3 in charge exchange kinetics, which indicates that the appropriate pore structure can significantly improve the cycle stability and rate capability[67]. Considering that bulk or common large hollow structural materials are often not conducive to alleviating volume expansion due to lack of internal space or too large internal space, Cao et al. Designed and synthesized a mini-hollow Mn2O3 polyhedron.In the internal small pores of the hierarchical nanostructure, the nanoparticles rearranged due to the nanosize effect are uniformly dispersed, which induces more active sites, improves the reaction kinetics and forms more oxidation products[68]. Therefore, it is very important to reasonably control the size of the hollow structure to improve the electrochemical performance of the conversion mechanism anode materials. The hollow Mn2O3 microspheres with hierarchical porous shell structure were generated from the non-uniform shrinkage caused by non-equilibrium heat treatment, and the electrochemical performance changed with the heating rate, although it was improved[69]. Hu et al. Prepared mesoporous MnOx micro-cuboids using Mn-MOF-74 as a precursor[70]. On the one hand, the special 3D hierarchical structure promotes ion migration and reduces the exchange charge resistance. On the other hand, high specific surface area, electrochemical activity and structural stability lay the foundation for good lithium storage performance. Obviously, high-performance metal oxide electrode materials with complex 3D micro/nano structures can be prepared by morphology and dimensionality control, and some breakthroughs have been made in low-cost and pollution-free synthesis technology.
Co3O4 with high theoretical capacity is the most concerned Co-based oxide anode material, but it is also plagued by problems such as volume expansion during cycling. Pyrolytic MOFs materials can effectively reduce the surface area, increase the pore size, improve the conductivity and keep the diffusion path unobstructed. Although the specific surface area of agglomerated Co3O4 nanoparticles synthesized by Co-MOF([Co3(HCOO)6]·(DMF)4) for the first time is low, the capacity, rate capability and cycle life are obviously improved, and the electrode performance can be further reasonably optimized by agglomerate structure control with variable primary and secondary particle sizes[71]. Li et al. Employed a template-free self-assembly strategy to fabricate one-dimensional hierarchically porous Co3O4 nanorods composed of interconnected nanoparticles[72]. The modification of nanoparticles and the hierarchical porous structure can shorten the diffusion distance of ions and charges and alleviate the volume change. At the same time, one-dimensional nanorods provide more exposed active sites and larger electrode/electrolyte contact area. Therefore, the capacity of the anode material is 628 mAh/G after 350 cycles at 1000 mA/G; After more than 600 cycles at 5000 mA/G, the capacity fade was only 0.068% per cycle. Compared with bulk materials, porous Co3O4 wrinkled nanosheets have higher aspect ratio and surface area, and more available active sites, all of which are beneficial to their use as anode materials with high specific capacity and rate capability[73]. In addition, the porous Co3O4 cuboid anode synthesized by direct pyrolysis of Co-BTC precursor is also improved in lithium storage and rate capability[74]. Han et al believed that the parallelepiped of porous hollow Co3O4 composed of uniform and continuous nanoparticles could significantly improve the reversible capacity, rate performance and cycle stability.The importance of three-dimensional nanostructured electrochemically active materials in the field of energy storage is emphasized, and this simple solid phase conversion method can also be used to prepare other porous hollow metal oxides with fine morphology[75]. In order to overcome the shortcomings of traditional Co3O4 electrodes, Hu et al. Fabricated Co3O4 nanocages composed of porous shells and hollow structures loaded with a large number of nanoparticles using the non-equilibrium interdiffusion (Kirkendall effect) of Prussian blue analogue (PBA,Co3[Co(CN)6]2)[76]. Compared with the dense hollow structure, this unique structure is more conducive to the intercalation/deintercalation of Li+ in the active material and accelerates the diffusion rate, and the capacity is stable at 1465 mAh/G after 50 cycles at a high current density (300 mA/G). Mesoporous nanostructured Co3O4 was synthesized from MOF-71 template by low temperature (300 ℃) pyrolysis and long time (12 H) reaction, and its electrochemical performance was significantly improved by reasonable pore volume, high specific surface area and low particle size[77]. Gou et al. Further proved that only appropriate temperature and time can remove the ligand and obtain unique morphology and structure during the pyrolysis of Co-MOFs precursor[78]. With the combined effect of mesopores and nanoparticles, the process-optimized Co3O4 anode has a reversible capacity of 1058.9 mAh/G after 100 cycles at 200 mA/G; The capacity is maintained at 348 mAh/G at 2000 mA/G. In addition, under the same pyrolysis process, porous MOFs derivatives often have different properties due to the retention of precursor morphology. For example, the electrochemical performance of the twin-hemispherical Co3O4 anode is more outstanding than that of the flower-shaped Co3O4 with a relatively large surface area, and the reversible capacity is also higher than that of the commercial graphite anode (372 mAh/G), due to the smaller particle size of the uniform nanoparticles, the appropriate pore size, and the denser hierarchical structure[79]. There is also a view that the final morphology of metal oxides derived from pyrolytic MOFs is related to factors such as organic chain type, metal ions and organic amines, and does not depend on their initial morphology[80]. In contrast, different kinds of Co3O4 hollow dodecahedra were derived due to the change of synthesis process even if ZIF-67 precursor was selected, and the formation mechanism was similar to that of core-shell microspheres synthesized by pyrolysis of solid precursor[81][82]. Due to the advantages of the size, specific surface area and structure of the primary nanoparticles, the utilization of the active material is significantly improved, and the reversible capacity of the hollow dodecahedral mid-sphere Co3O4 anode is as high as 1550 mAh/G, and the cycling stability is ideal (1335 mAh/G after 100 cycles). This similar phenomenon has been observed in Co3O4 hollow microspheres with complex internal structure[83]. Hexagonal Co3O4 nanorings are pyrolytic derivatives of Co-MOFs assisted by organic amines, and the final morphology is determined by the Co2+ release rate of the reaction system and the steric hindrance of the organic chains[80]. Compared with commercial Co3O4 particles, the unique morphology and single crystal characteristics make the reversible capacity and stability of hexagonal Co3O4 nanorings anode higher.
Iron oxide has been widely studied as an anode due to its stable structure and high theoretical capacity (1000 mAh/G), but its low conductivity and cycle stability restrict its development. Spindle-shaped α-Fe2O3 anode materials can be prepared by using Fe-MOFs as templates or precursors, but different preparation methods will cause changes in the electrode structure[84,85]. For example, the product of the one-step method is a spherical α-Fe2O3 with nanometer size; The products of the two-step method are mesoporous materials composed of small-sized α-Fe2O3 nanoparticles, which provide larger specific surface area and more electrochemical active sites, and greatly shorten the Li+ diffusion distance, thus significantly improving the lithium storage performance. With the help of microwave irradiation for homogeneous nucleation and the inherent framework structure and porosity of MOFs (MIL-53-Fe) template, Guo et al. Synthesized mesoporous nanostructured Fe2O3 with different morphologies by simply adjusting the irradiation time[86]. Among them, the porous core-shell structure of Fe2O3 octahedron performs well in terms of reversible capacity, cycle life and rate capability due to its special structure, which is significantly better than the porous and hollow anode materials with single structure. Metal-based nanomaterials with hollow structures can effectively alleviate the accompanying volume change during the conversion of low-density MOFs into metal oxides. Zhang et al. Prepared Fe2O3 microboxes with different shell structures by using the characteristics of Prussian blue cube oxidative decomposition and synchronous growth of iron oxide shell[87]. Due to the hierarchical hollow porous structure with high crystallinity and structural stability, which helps to maintain the initial nanostructure during cycling, the capacity and cycling stability of this Fe2O3 microbox anode are high. Compared with the widely used liquid phase method, this solid phase method is easier to synthesize uniform anisotropic hollow structures on a large scale, but the morphology and structural characteristics of the final products are often different due to the change of calcination temperature. From this point of view, further exploring the coating of the internal space and the construction of core-shell or other hollow structure anode materials are conducive to the realization of structural efficiency and value.
Soundharrajan et al. Found that the reversible capacity of NiO nanoparticle anode was 748 and 410 mAh/G after 100 cycles at 500 and 1000 mA/G, respectively, which mainly depends on its particle size, surface area and structural stability, and its comprehensive performance is comparable to that of electrode materials with other supporting materials (such as graphene or graphene oxide)[88]. Although the capacity of the nano/microstructured NiO porous anode is high (about 400 mAh/G), attention should be paid to the effects of calcination temperature and time on its structure and morphology[89]. In addition, restricted by the poor stability of crystal structure, small contact area with electrolyte and low reactivity with Li+, the NiO anode derived by two-step calcination of Ni-MOF(Ni-H2Me4bpz) leads to capacity fading and reduced cycle performance due to pulverization, while the porous hollow NiO nanospheres can accelerate the volume change in the process of Li+ intercalation/deintercalation kinetics and coordination conversion[14]. The preparation methods of inorganic hollow nano-structured materials generally include Kirkendall effect, electrochemical replacement, chemical etching and template methods. Among them, the template method is an important means to prepare hollow metal oxides with different morphologies by using MOFs. For example, NiO quasi-nanospheres derived from Ni-MOF are typical porous hollow structures, which are composed of a large number of self-aggregation particles with an average diameter of about 40 nm[90]. As an anode material, it has a reversible capacity of 760 mAh/G after 100 cycles at 200 mA/G, which is even higher than the theoretical capacity of NiO electrode (718 mAh/G) based on the conversion mechanism. These additional capacities are due to the Faradaic effect of electrolyte side reactions and/or the pseudocapacitive behavior of reversible formation of polymer-like films. More importantly, the capacity is still 392 mAh/G at a high current density (3200 mA/G).
Compared with other traditional metal oxides, TiO2 electrodes are prone to structural collapse and do not form SEI layers during charge-discharge cycling, especially the low rate performance caused by low conductivity and volume effect, which directly affects their industrial application[91~94]. Within the MOFs crystal, the metal and oxygen atoms are arranged in periodic atomic levels, which can be completely transformed into metal oxides without long-range atomic migration and retain permanent pores to some extent. At the same time, the typical hydrophobic pores generated by calcination are conducive to electrolyte penetration and Li+ diffusion, which are the basic conditions for the preparation of new porous metal oxide anode materials by MOFs template method, and the technical route is completely different from the traditional inorganic materials synthesis method. For example, Wang et al. First used MIL-125 (Ti) as a precursor to prepare anatase TiO2 porous anode, and the reversible capacity and rate capability were significantly improved, which provided a new idea for exploring new MOFs-derived TiO2 materials with different structures and Ti contents[95]. In addition, the 2D GeO2 nanosheets also showed an ultrahigh lithium storage capacity during long-term cycling (1393 mAh/G reversible capacity after 350 cycles at 100 mA/G)[96].

3.2 Bimetallic oxide

Generally speaking, most of the traditional bimetallic oxide anodes belong to the spinel structure, which mainly improves the performance by the synergistic effect of transition metal atoms with different activities.However, the problems of poor conductivity, low activation energy of electron transfer and ion diffusion coefficient, few vacancies to store Li+ and high polarization potential during cycling have seriously restricted its large-scale application. MOFs-derived bimetallic oxides containing different metal ions are expected to solve these problems to some extent. It should be emphasized that these bimetallic oxides are not a mixture of two metal oxides in the physical sense, and there may be a constraint effect between the metal elements. Therefore, in addition to the special porous structure and abundant redox active sites, bimetallic oxides are more conductive than monometallic oxides, with significantly shorter electron transport pathways, faster Li+ diffusion and migration, higher reversible capacity, and stable operation even at high current densities[97]. For example, bimetallic oxides such as ZnFe2O4, CoFe2O4, and CuFe2O4 are only partially reduced during cycling, followed by reoxidation to improve the reversible capacity[98][99][100].
Benefiting from the unique porous structure and chemical composition, as well as the rapid migration of a large number of Li+ and electrons, the electrochemical and rate performances of MOFs-based derived MnCo2O4 nanowires are superior to those of traditional spinel-structured MnCo2O4 in the long-term charge-discharge process, and the preparation method is also significantly improved[101]. The specific surface area of two-dimensional nanostructured materials is larger, which has more advantages in alleviating volume change, exposing active surface and shortening the diffusion distance of Li+ and electrons. Wang et al. Partially replaced Co in Co3O4 with Zn and successfully synthesized ZnxCo3-xO4 nanosheets to improve the lithium storage performance by adjusting the metal element ratio and rationally designing the porous structure, combined with two-dimensional morphology and nano-sized building units[97]. Similarly, the porous ZnCo2O4 nanosheet anode has a reversible capacity of 1640.8 mAh/G after 50 cycles at 100 mA/G; The discharge capacity is still 581.3 mAh/G after 190 cycles at 1500 mA/G[102]. This structural design method can not only greatly reduce the cost and toxicity of electrode materials, but also be used to prepare two-dimensional porous bimetallic oxide anodes derived from other MOFs, which is expected to achieve commercial applications. With the help of interconnected porous structure, structural stability enhanced by a large number of mesopores, high specific surface area formed by gas evolution, and retained morphological characteristics of MOFs, the capacity of Mn1.8Fe1.2O4 nanocubes was 827 mAh/G after 60 cycles at 200 mA/G[103]. The mixed valence state (Fe2+↔Fe3+ and Mn2+↔Mn3+) of the multimetallic center provides the expected conductivity for effective energy storage, probably due to the defect effect mechanism. This obvious difference just proves the significance of lowering the synthesis temperature for the bimetallic oxide anode. Hollow structural materials generally contain controllable internal pores and functional shells, but most of the single-shell structures are often too simple and relatively single in composition, which makes it difficult to maintain structural integrity in the process of repeated insertion/extraction of Li+, which inevitably limits the improvement of cycle performance of MOF-derived hollow metal oxides. Therefore, reasonable structural design of MOF-derived materials is the development direction of high-performance anode. Wu et al. Used bimetallic MOF (Ni-Co-BTC) as a precursor to prepare spinel-type NixCo3-xO4 multi-shelled hollow spheres composed of nanosized subunits in the shell and large pore spaces between adjacent shells by a self-sacrificial template method[104]. The reversible capacity is 1109.8 mAh/G after 100 cycles at 100 mA/G, and the rate capability is outstanding. In addition, adjusting the chemical composition can also improve the electrochemical performance of the NixCo3-xO4 anode. Although the synthesis technology of hollow spherical bimetallic oxides has been relatively mature, the design and preparation of non-spherical hollow complex bimetallic oxides still face challenges. Wu et al. Prepared porous spinel-type ZnxCo3-xO4 hollow polyhedron anode materials by pyrolysis-induced transformation of heterogeneous bimetallic Zn-Co-ZIF[105]. On the one hand, the high-symmetry hollow structure composed of nano-sized subunits in the shell can provide more degrees of freedom and form atomic steps, rather than asymmetric structures. The high-density atomic steps on these atomic planes are beneficial to the reaction between Li and ZnxCo3-xO4, so the cycle stability and rate capability are good. On the other hand, the hollow structure effectively suppresses the initial nanoparticle aggregation and dissolution in the electrolyte. Therefore, this low-cost and simple method is expected to be used to prepare other hetero-hollow structured bimetallic metal oxides and construct advanced energy conversion and storage electrodes.
As mentioned above, transition metal oxides often face the problems of electrode pulverization caused by volume effect, rapid capacity decay and reduced cycle stability. In bimetallic oxides, metal ions react electrochemically in the matrix of different material systems, and the combination of two metals provides abundant redox reaction sites and improves conductivity. Under such conditions, the matrix, either active or inactive for lithium, undergoes a volume change in a progressive manner rather than in a fixed pattern, relying on the unreacted component to coordinate the strain produced by the reacted component. As a precursor or template for the preparation of new nano-functional materials, nano-MOFs have unique hollow structure and high porosity and specific surface area, which make it easier to select and control different metal ions and organic ligands. Especially in the conversion process, the pores and long-range ordered structure of nano-MOFs provide a convenient way for the rapid transport of small molecules and ions. Porous hollow CoFe2O4 nanocubes and NiFe2O4 nanocages were synthesized by using Co[Fe(CN)6]0.667 and Ni2Fe(CN)6 as precursors or templates, and the element hybridization characteristics made the two anode materials change in volume periodically during cycling, showing excellent performance in both rate capability and cycling stability without being affected by the increase of current density[106,107]. Similarly, Co-Ni-O bimetallic oxide is a derivative obtained by further heat treatment based on the microwave-assisted synthesis of Co-Ni bimetallic MOF nanorods[108]. For example, the reversible capacity of mesoporous Ni0.3Co2.7O4 nanorods reached 1410 mAh/G after 200 cycles at a low current density (100 mA/G) and maintained a high reversible capacity at 5000 mA/G. The excellent electrochemical performance is mainly due to the synergistic effect between the interconnected mesoporous nanostructure and the two active metal oxide components (Co3O4 and NiO). Obviously, most of these bimetallic oxides are derived from bimetallic MOFs, in which the bimetallic elements include Co, Ni, Zn, Mn, Fe, etc. Different from this, Chen et al. Used MIL-25 (Ti-MOF) and LiNO3 as titanium source and lithium source, respectively, to prepare porous Li4Ti5O12 anode by solid state sintering method, which not only had good cycle stability and rate performance, but also provided a new idea for the preparation of other electrode materials[109]. The precursor/template and main electrochemical performance of some MOFs-derived metal oxide anode materials are shown in Table 2.
表2 MOFs衍生金属氧化物负极材料及其电化学性能

Table 2 MOFs-derived metal oxides as anode materials and their electrochemical performances

Materials Template/precursor Voltage range (V) Current density (mA/g) Cycle number Overall capacity (mAh/g) Initial discharge/charge capacity (mAh/g) Initial coulomb efficiency(%) Specific surface area
(m2/g)
ref
CuO Cu(L-Phe)2 0.1~3.0 100/1000 200/500 505.3/116.7 -/- 62
CuO Cu-BTC 0.05~3.0 100 100 470 1208/- 40 49.6 63
CuO MOF-119 0.005~3.0 2000 40 210 1208/538 64
CuO Cu-BTC 0.01~3.0 500/1000/2000 200/400/400 1062/615/423 1334.7/836.1 49.75 65
Mn2O3 Mn-MIL-100 0.1~3.0 200 100 755 1668/1003 40.45 66
Mn2O3 Mn-LCP 0.01~3.0 1000 250 705 1158/852 74 15.34 67
Mn2O3 Mn-MOF 0~3.0 400/1000 450/1200 1370/819.8 -/- 68
Mn2O3 Mn-BTC 0.01~3.0 100 60 582 3404/1559 46 38.5 69
Mn3O4 Mn-MOF-74 0.01~3.0 200/2000 400/400 890.7/437.1 1078.9/625.1 80.5 70
Co3O4 [Co3(HCOO)6](DMF)4 0.01~3.0 50/100 50/100 965/730 1118/- 5.3 71
Co3O4 Co-MOF 0.01~3.0 1000/5000 350/600 628/412 1402/879 62.7 42 72
Co3O4 CoBDC MOF 0.01~3.0 100/1000 160/200 1477/775 1392/961 69.09 133.74 73
Co3O4 Co-BTC 0.00~3.0 100 60 886 2082/1061 51 10.44 74
Co3O4 Co-MOF 0.01~3.0 100 50 1115 1608/1080 43 75
Co3O4 PBA 300 50 1465 1557/- 66.5 76
Co3O4 MOF-71 0.001~3.0 200 60 913 1286.1/879.5 68 59 77
Co3O4 Co2(NDC)2DMF2 0.01~3.0 200/2000 100/100 1058.9/348 1504.2/976.7 40 78
Co3O4-a Co-MOF 0.01~3.0 100 90 470.3 1325.5/1003.5 75.7 22.6 79
Co3O4 ZIF-67 0.01~3.0 100/100 100/60 1335/1265 1735/1083 45 81
Co3O4 Co-MOF 0.01~3.0 100 100 1370 1324/1034 20.1 80
α-Fe2O3 MIL-88 0.01~3.0 200 50 911 1372/940 69 75 84
α-Fe2O3 Fe-MOF 0.005~3.0 100 40 921.6 1487/1024 85
Fe2O3-2 MIL-53 0.005~3.0 100/1000 200/500 1176/744 1456/1048 93.1 86
Fe2O3 PB 0.01~3.0 200 30 945 -/- 25.4 87
NiO MOF-C 0.005~2.5 500/1000 100/100 748/410 2134/1303 61 36 88
NiO Ni-MOF 0.01~3.0 15 100 380 900/480 24 89
NiO Ni-MOF 0.0~3 200 100 760 1149/850 28.6 90
TiO2 MIL-125 1.0~3.0 168/840/1680 500/500/500 166/106.5/71 168/- 220 95
GeO2 Ge-MOF 0.005~3.0 100 350 1393 2079/1315 63.2 12.9 96
MnCo2O4 Mn-Co-MOF 0.01~3.0 100 100 929 1496/963 64 31.69 101
Zn-NPs ZIF-L 0.01~3.0 100 100 143 1245.9/692.2 55.6 47.6 97
ZnCo2O4 ZnCo-8-hydroxyquinoline 0.01~3.0 100/1500 50/25 1640.8/348.1 1710.2/1273.5 74.5 118 102
Mn1.8Fe1.2O4 Mn3[Fe(CN)6]2·nH2O 0.01~3.0 200 60 827 2312/1337 57.8 124 103
NixCo3-xO4 Ni-Co-BTC 0.005~3.0 100/1000/2000 100/300/300 1109.8/832/673 1619.2/1139.3 70 96.7 104
ZnxCo3-xO4 Zn-Co-ZIF 0.01~3.0 100 50 990 1272/969 76.2 65.58 105
CoFe2O4 Co[Fe(CN)6]0.667 0.01~3.0 1352/1190 85.3 102.692 106
NiFe2O4 Ni2Fe(CN)6 0.01~3.0 914 200 1071 1245/1152 260.9 107
Ni0.3Co2.7O4 Co/Ni-MOF-74 0.005~3.0 100/2000/5000 200/500/500 1410/812/656 1737/1189 28.5 108
Li4Ti5O12 MIL-125 1.0~3.0 500 700 120.3 184.9/149.1 80.6 109

3.3 Other metal compounds

A large number of studies have shown that metal oxides derived from MOFs are superior to traditional transition metal oxides in terms of specific capacity, cycle stability and rate capability. Inspired by this, many researchers have turned their attention to other metal compounds derived from MOFs, aiming to make full use of the advantages of MOFs and use simple preparation methods to construct nano-anodes with different structural characteristics to meet the needs of lithium-ion batteries in lithium storage capacity. Yu et al. Used a two-step diffusion-controlled method to synthesize hollow prisms composed of interconnected bubble-like nanosized CoS2 subunits, and the lithium storage performance was significantly better than that of the previously reported Co sulfide-based anode[110][111,112]. The ultra-thin-walled nanobubbles significantly shorten the diffusion distance of the Li+ and improve the electrochemical kinetics. More importantly, the multi-level hollow interior of the hierarchical prism can effectively coordinate the structural stress during the repeated charging and discharging process. Metal phosphides have lower redox potentials and higher theoretical capacities (e.g., 926 and 894 mAh/G for FeP and CoP, respectively). The bimetallic NiCoP nanocage retains the quasi-polyhedral characteristics of the ZIF-67 template, and the inner and outer shells are composed of many wrinkled nanosheets with a large number of micropores on the surface and connected with each other. This multi-channel structure effectively shortens the diffusion distance of ions and electrons and relieves the diffusion-induced stress. At the same time, the free space between adjacent nanosheets can coordinate the stress change caused by the volume change[113]. Therefore, the NiCoP anode material has a complete structure without agglomeration, shrinkage and collapse during cycling, and has good reversible capacity, rate capability and cycle stability.

4 MOFs derived porous carbon materials

Traditional carbon materials have abundant reserves, high conductivity and chemical stability, and limited energy and power density and lithium storage capacity. Carbon nanotubes, carbon nanofibers, carbon films, graphene and other new anode materials have more advantages in specific surface area, morphology control, structural stability and electrochemical performance, but it is difficult to get rid of the harsh synthesis conditions and high cost, which is not conducive to industrial application. It is well known that MOFs materials can realize the coexistence of metal elements, organic molecules or other guests, and the high specific surface area and developed pore structure provide sufficient space for Li+ intercalation/deintercalation. As mentioned above, most MOFs materials have more or less deficiencies in conductivity, stability and reversibility, and their electrochemical performance decreases rapidly after several charge-discharge cycles. Admittedly, MOFs materials are also ideal sacrificial templates or precursors for derivatives such as porous carbon, graphene (568 mAh/G), and carbon nanotubes (625 mAh/G)[114][115]. In addition to the central metal ion and pore size, further studies have found that poor thermal stability is the main obstacle for MOFs to be derived from carbon materials by heat treatment. Structural collapse occurs even with low temperature heat treatment, hindering intercalation/deintercalation of Li+ and reducing capacity[116].
In the structure of MOFs, the organic ligand can provide sufficient carbon source and abundant active sites, and the derived carbon materials well retain the characteristics of high specific surface area and uniform pore size distribution of the precursor. On the one hand, high-temperature calcination in a specific protective atmosphere (Ar or N2) can effectively avoid oxidation reaction, the synthesis process is easy to control, and the operation is relatively simple and convenient; On the other hand, the special framework structure has high stability, which can improve the migration rate and charge exchange ability of Li+ and alleviate the volume change during charging and discharging[117,118]. During the pyrolysis of MOFs, the temperature directly affects the final composition, morphology, specific surface area and porosity of the derived carbon materials. Generally speaking, higher pyrolysis temperature is beneficial to the formation of ordered conductive graphite. On the contrary, excessive heating increases the thermal brittleness of MOFs, and even causes partial or complete collapse of the structure. Zeolitic imidazole frameworks (ZIFs) have higher porosity and smaller pore size than other MOFs when the volume is the same, which is beneficial to more ion and electron transport under the same rate. The 3D porous carbon anode was prepared by direct calcination of MOF-5(Zn4O(BDC)3), with a specific surface area up to 1880 m2/g and initial charge-discharge capacities of 2983 and 1084 mAh/G at 100 mA/G, respectively[119]. Takamura et al. Attributed the higher irreversible capacity and lower coulombic efficiency to SEI layer formation, irreversible intercalation of Li+ related to carboxyl functional groups, and surface physisorbed water[120]. In fact, the SEI layer not only prevents excessive organic solvent from co-intercalation and acts as a good conductor of Li+ to promote Li cycling, but also avoids the direct contact between the strong reducing agent LiCx and the electrolyte, inhibiting harmful side reactions. After 100 cycles at 100 mA/G, the 3D porous carbon anode has a capacity of 1015 mAh/G, which is significantly higher than that of commercial graphite (372 mAh/G). Considering the structural design and assembly, Li et al. Synthesized porous carbon with dense pores and multifractal structure by vacuum pyrolysis of Zn-MOF precursor.Macroscopic pores and mesopores provide short migration and diffusion paths for Li+ by optimizing electrolyte penetration conditions, while subnanometer pores with fractal structure provide a large number of active sites for lithium storage and form a transition zone rich in lithium atoms[121]. According to the theory of fractal surface electrochemical reaction, the surface or interface with higher fractal dimension can provide larger electrochemical active surface area and lower internal resistance[122,123]. By virtue of the pore space distribution in the carbon matrix, the rate performance and cycle stability of the porous carbon anode are significantly improved, that is, the lithium storage performance of the porous carbon anode is significantly improved by the combined effect of bulk, surface and defect lithium storage modes. Luo et al. Used a new Co-MOF template to prepare porous carbon with hierarchical irregular structure by high temperature carbonization and etching[124]. Similarly, Peng et al. Synthesized porous carbon by calcination and etching treatment of a Cd-MOF precursor[125]. The discharge and charge capacities of the anode material are as high as 2486 and 1683 mAh/G at 300 mA/G, respectively, and the initial coulombic efficiency is 68%. The high capacity loss caused by the formation of SEI layer and electrolyte decomposition is also a common problem of most anode materials. However, the reversible capacity was maintained at 1285 mAh/G after 300 cycles at 300 mA/G. In addition, the flexible binderless porous carbon nanofibers are prepared by pyrolysis and etching of PAN/Zn-MOFs at high temperature (800 ℃) in an inert atmosphere, and this relatively simple process provides a basis for the preparation of one-dimensional high-performance anode materials[126]. Obviously, in order to obtain ideal porous carbon materials, the precursor or template of MOFs containing high melting point metals (Fe, Co, Cu and Cd, etc.) must be removed by subsequent etching treatment after pyrolysis. It is worth noting that porous carbon materials can be obtained directly from MOFs containing low melting point metals such as Zn without etching treatment when calcined above 900 ℃. Therefore, for MOFs-derived carbon materials, whether etching treatment is needed in the preparation process depends on the type of metal ions in the MOFs structure and the pyrolysis temperature.
Due to the surface superhydrophobicity and insufficient active sites, the lithium storage performance of nanoporous carbon anode is not ideal. In order to preserve the morphological characteristics of MOFs as much as possible, element doping is an effective way to improve the structural stability. Doping heterogeneous atoms can effectively improve the capacity, surface wettability, electrical conductivity and mechanical properties of nanoporous carbon electrodes. By selecting appropriate organic ligands, heterogeneous atoms such as N, S, P, and B can be doped into the structure of MOFs. The electronegativity of N (3.5) is higher than that of C (3.0), and the atomic radius of N is also smaller. The ligand doped with N has a strong adsorption effect on the Li+, and the N atoms combined in the graphite network are beneficial to the strong interaction between the carbon structure doped with N and the Li+, which improves the charge density and thermal stability of the structure, and promotes the electron and ion exchange. Therefore, heterogeneous N-doped carbon materials are often synthesized by calcination of MOFs containing N-rich organic ligands, which have high specific capacity and cycle stability. The N-modified porous carbon material was prepared by direct carbonization and etching of Cu-MONFs nanofibers, and the N content was 8. 62 wt%[127]. This highly disordered nanostructured material has a large number of defects and more favorable sites for Li+ binding, while the binding of N atoms can form n-type conductive materials and improve electrochemical reactivity and conductivity. The discharge capacity of the anode material is 853.1 mAh/G after 800 cycles at 500 mA/G, which is higher than that of ordered graphite. Due to the existence of a large number of edge-doped N atoms on the inner surface, which can provide abundant active sites for Li+ adsorption and affect the structure and electron distribution of NCNFs anode materials, the electrochemical performance is significantly improved, as shown in Fig. 3[128]. Similarly, hierarchically structured N-P-C porous carbon microspheres are Ni-ZIF synthesized in situ by carbonization-etching[129]. During the carbonization process, the Ni2+ has a strong catalytic effect on the pyrolysis of N-containing organic ligands, and the high content of pyrrole and pyridine N increases the defect concentration and forms more abundant active lithium storage sites. The reversible capacity of this anode material is 725 mAh/G at 100 mA/G, and it still maintains 570 mAh/G after 200 cycles. In particular, N-doping can improve the surface polarity, conductivity, and electron-donor propensity of nanoporous carbon. N-doped nanoporous carbon (N-NPC) was derived from the direct carbonization of Al-MOF template with specific surface area and total pore volume up to 1244 m2/g and 0.96 cm3/g,N content of 10%, respectively[130]. At 100 mA/G, the initial specific discharge capacity and reversible capacity of N-NPC anode are 820 and 720 mAh/G, respectively, especially the reversible capacity is about twice of the theoretical capacity of graphite (372 mAh/G). The capacity fading at different stages during cycling mainly comes from the formation of SEI layer on the surface and the structural change of electrode materials at high current density. The porous carbon electrode derived from MOFs can retain the morphological characteristics of the bulk material of MOFs, and its structure and composition can be controlled by changing the pyrolysis temperature. Surface N doping induces pseudocapacitance between electrolyte ions and porous carbon electrode. In-situ doping with N-containing precursors can increase the N content of the final product and avoid the formation of oxygen-containing functional groups, which is more advantageous than post-treatment of carbon sources with N-containing reagents. Polydentate TCNQ is a redox-active organic ligand with a high N content of 27.5 wt%. Therefore, the solid-state physical properties of N-rich TCNQ-based MOFs are distinctive. Tong et al. Prepared N-doped carbon materials with uniform distribution using TCNQ-based Sr-MOFs as sacrificial templates under N2 atmosphere[131]. The initial discharge capacity of N-C-550 (9.98 wt% N) anode is as high as 1043 mAh/G at 100 mA/G, which is affected by the formation of SEI layer and the decomposition of electrolyte, and the initial coulombic efficiency is only 64.7%. The discharge capacity is 736.8 mAh/G after 50 cycles at 100 mA/G. The discharge capacity is 460.7 mAh/G at 800 mA/G, which is due to the high N content, specific surface area and porosity. The results show that the carbonization parameters directly affect the content and type of doped N (especially pyridine-N). Pyridine-N and pyrrole-N can introduce a large number of surface defects to form disordered carbon structures, and these N atoms and oxygen functional groups can act as electrochemical reaction active sites.
图3 NCNFs在Li+半电池中的电化学性能: (a) NCNFs-800在0.1 mV/s时的CV曲线, (b) NCNFs-800在0.1 A/g时的充放电曲线, (c) NCNFs在电流密度0.1~10 A/g时的倍率性能, (d) NCNFs-800在2 A/g时的循环性能[128]

Fig. 3 The electrochemical performances of NCNFs in Li+ half cells. (a) CV curves of NCNFs-800 at 0.1 mV/s. (b)Discharge/charge profiles of NCNFs-800 at 0.1 A/g. (c) Cycling performances of NCNFs at 100 mA/g. (d) Rate capability of NCNFs at a current density from 0.1 to 10 A/g.(e) Cycling performance of NCNFs-800 at 2 A/g[128]

Heteroatom doping can change the host electronic structure of carbon materials, thus improving the specific capacity and rate capability of porous carbon anode. In most cases, the doped heterogeneous atoms are localized only within the lattice. Due to the lack of sufficient space in the graphite structure, only doping with large-sized functional groups such as NH2 and NO2 at the edge can not destroy the crystal structure. Among the common doping groups, the nitro group has the highest adsorption energy for Li, even higher than pyridine N and pyrrole N. Without selecting organic matter as the N source, Yang et al. Synthesized nitro-edge-modified porous carbon (N-C) by high-temperature pyrolysis of Cu-MOF precursor in N2 and subsequent etching treatment[132]. Although the N-C anode exhibited irreversible capacity fading at low current density (100 mA/G) and the initial coulombic efficiency was not satisfactory, the coulombic efficiency increased to about 99% after 1500 cycles at high current density (1000 mA/G). In addition to the stronger adsorption ability of nitro groups and more convenient intercalation of Li+, the role of large-area edge nitro doping in improving the conductivity of porous carbon electrodes and increasing the capacity as a reservoir is also important. Theoretical and experimental studies have shown that the lithium storage ability of N-doped graphene anode depends on the N-doping level. When the amount of N doping is more than 10 wt%, the large number of N atoms in the two-dimensional honeycomb lattice causes the structural stability of most N-doped carbon materials to decrease[133,134]. In order to avoid the formation of oxygen-containing functional groups, Zheng et al. Prepared N-doped microporous carbon polyhedra (17.72 wt% N) by direct carbonization of ZIF-8 (34 wt% N), which can be regarded as a large number of small graphene-like particle aggregates[135]. Relying on sufficient surface active sites and edge-trapped heterogeneous N atoms (especially pyridine or pyrrole N), the specific capacity of the anode material is as high as 2132 mAh/G after 50 cycles at 100 mA/G; Still 785 mAh/G after 1000 cycles at 5 A/G. This outstanding performance is attributed to the unique N-doped structure, which also provides a new idea for high level N-doping of the host lattice and the edge. The main electrochemical performances of some typical MOFs-derived porous carbon anode materials are listed in Table 3.
表3 MOFs衍生多孔碳负极材料及其电化学性能

Table 3 MOFs-derived porous carbon as anode materials and their electrochemical performances

Materials Template/precursor Voltage range (V) Current density (mA/g) Cycle number Overall capacity (mAh/g) Initial discharge/charge capacity (mAh/g) Initial coulomb efficiency(%) Specific
surface
area
(m2/g)
ref
3D porous carbon Zn4O(BDC)3 0.01~3.0 100 100 1015 2983/1084 1880 114
porous carbon Zn-MOF 0.01~3.0 74 50 2016 -/2458 2587 121
Porous carbon Co-MOF 0.01~3.0 100 49 549 3066/946 688 124
Porous carbon Cd-MOF 0.01~3.0 300 300 1285 2486/1683 68 1796 125
Porous CNFs ZIF-8 0.01~3.0 100 200 520 570/390 68 126
Nitrogen-modified carbon Cu-MONFs 0.01~3.0 500/5000 800/1000 853.1/440 1584.4/942.1 59.5 7.275 127
N-C Cu-MOF 0.01~3.0 100/1000 100/1500 890/588 2037/1039 51 128
N-P-C Ni-ZIF 0.01~3.0 100 200 570 1497/725 48.5 320.8 129
N-NPC Al-MOF 0.01~3.0 1000 400 352 820/720 87.8 1244 130
N-C-550 Sr-MOF 0.1~3.0 100 50 736.8 1043/675 64.71 131
N-C-800 ZIF-8 0.01~3.0 100/5000 50/1000 1147/785 3487/2037 58.4 134

5 Composite material

In addition to the bulk materials of MOFs and their derived porous carbon and metal oxides, which are directly used as anode materials for lithium-ion batteries, the organic combination of MOFs and their derivatives with other materials or their surface modification can also improve the lithium storage performance. Composite material is a new type of functional material which is synthesized by optimizing and combining two or more materials with different properties by advanced preparation technology. Compared with single component materials, composite materials not only maintain the good properties of each component material, but also obtain ideal comprehensive properties through the complementarity and correlation of each component. It is feasible to combine MOFs or MOFs derivatives with other materials with high conductivity and theoretical capacity to prepare corresponding composites, and to further improve the electrochemical performance of MOFs-based composite anode materials by using special nanostructures to avoid agglomeration and alleviate the volume effect.

5.1 MOFs/carbon-based materials

Only when the electrode reaction is completely reversible during the charge-discharge process, the bulk materials of MOFs directly used as the negative electrode can achieve ideal cycle stability. However, the poor conductivity of MOFs will inevitably affect their performance to a certain extent. Cao Na et al. Used liquid phase deposition method to make MOFs grow in situ on carbon cloth, and prepared Co-MOFs/CF composites without binder[136]. The initial discharge specific capacity of the anode was 1621.3 mAh/G at 50 mA/G, and the specific capacity remained at 445.1 mAh/G even after 100 charge-discharge cycles, which was related to the synergistic effect between the hierarchical structure of MOFs and the highly conductive carbon fibers, which was beneficial to ion migration and storage. Li et al. Synthesized MIL-101 (Cr)/graphene oxide composites by a simple solvothermal method, that is, MIL-101 (Cr) octahedra were uniformly distributed on the surface of graphene oxide with hierarchical structure, but the combination of graphene oxide did not affect the formation of MIL-102 (Cr)[137]. The initial discharge capacity of MIL-101 (Cr)/GO anode is almost 2 times higher than that of MIL-101 (Cr) due to the increase of contact area between electrode and electrolyte caused by high specific surface area. The porous structure and the conductive graphene significantly shorten the diffusion time and accelerate the phase change reaction, the agglomeration of the electrode material and the pulverization of the active material caused by volume expansion/contraction in the charge-discharge process are also effectively inhibited, and the rate performance and cycle stability are significantly improved. On the basis of Cu-MOF prepared by coprecipitation method, Gao Guoliang et al. Synthesized Cu-MOF/rGO composite in situ, using carbonyl as binding site to realize reversible intercalation/deintercalation of Li+[138]. The specific discharge capacity of the anode material is 520 mAh/G after 50 charge-discharge cycles at 50 mA/G. In addition to its own good conductivity, rGO also avoids the direct contact between the electrolyte and Cu-MOF to some extent, inhibiting the SEI layer thickening. In addition, the flexibility of rGO also improves the volume effect and capacity fading caused by the lack of structural stability of MOF, and achieves good results in improving the performance of electrode materials. Considering that F contributes to the effective nucleation of MOFs and improves the reversibility of lithiation/delithiation, Wei et al. Realized the in situ self-assembly of hollow urchin-like F-Co-MOF with rGO via a solvothermal method[139][140]. The reversible capacity of F-Co-MOF/rGO anode is 1202.0 mAh/G at 100 mA/G and the coulombic efficiency is 99.95% after 50 cycles due to the synergistic effect and the promotion of Li+ intercalation/deintercalation. Especially, the full cell composed of nanostructured F-Co-MOF/rGO anode and LiFePO4 also exhibited good cycling stability. Jin et al. Prepared Fe-MOF/RGO composites composed of two-dimensional RGO nanosheets and their coated Fe-MOF octahedral particles by solvothermal method, and focused on the effect of RGO content on their lithium storage capacity[141]. The results show that only the Fe-MOF/RGO (5%) anode exhibits the best reversible capacity (2259.3 mAh/G at 500 mA/G), rate capability, and cycle stability. The electrochemical performance of MIL-53 (Fe) anode material mainly depends on the redox reaction of Fe3+, but the terephthalic acid ligand does not exert its electrochemical activity due to poor conductivity and thick surface SEI layer. Benefiting from the chemical passivity and surface hydrophilicity of RGO, which can reduce the contact area between the electrolyte and MIL-53 (Fe), avoid the surface reaction of the electrolyte, activate the carboxyl functional groups and improve the conductivity, the three-dimensional porous MIL-53 (Fe) @ RGO anode has better reversible specific capacity and rate capability than MIL-53[142]. However, due to the limited metal chelating ability of RGO and the absence of any carboxyl functional group on the basal plane, the rate performance improvement of Fe-MOF/RGO anode is not obvious. Co2(OH)2BDC is a typical MOF-based electrode material for aromatic chelating ligand controlled Li+ intercalation/deintercalation reaction[10]. Considering that abundant carboxyl functional groups can significantly improve the metal chelation strength of MOF and graphene and form a suitable three-dimensional structure, Li et al. Synthesized covalently reinforced Co2(OH)2BDC/ carboxylated graphene (CoCGr) composites by surfactant-free solvothermal method[143]. The outstanding electrochemical performance of CoCGr-5 anode is mainly attributed to the Li+ intercalation/deintercalation reaction controlled by the organic part and the bicontinuous electron/ion transfer path between Co-BDC and CGr. Meanwhile, the CGr loading not only affects the specific surface area and pore volume of CoCGr anode, but also affects the rate capability and cycle stability. In addition, the surface lithium storage reaction is Li+ fast intercalation/deintercalation process, which can also provide additional capacity and improve the rate performance[144]. Using H2BDC non-covalently functionalized graphene oxide as nucleation sites and structure-directed templates to promote the growth of MOFs, He et al. Successfully fabricated MOF/RGOn composites[145]. The reduction of transition metal ions (Mn2+) and the subsequent lithiation of organic ligands are the direct reasons for the higher reversible capacity of the MOF/RGOn anode, and the long-range conducting RGO network can also ensure efficient electron transport and provide mechanical flexibility to the MOF, avoiding drastic volume changes and self-aggregation. Therefore, the electrochemical performance of MOFs-based electrode materials can be improved by relying on other electron-withdrawing groups to bridge organic ligands and metal ions[146]. The electrochemical performances of MOFs/porous carbon composite anode materials are summarized in Table 4.
表4 MOFs/多孔碳复合负极及其电化学性能

Table 4 MOFs/ metal compound composites as anode materials and their electrochemical performances

Materials Voltage range (V) Current density (mA/g) Cycle number Overall capacity(mAh/g) Initial discharge/
charge capacity
(mAh/g)
Initial coulomb efficiency(%) Specific surface area (m2/g) ref
Co-MOFs/CF 0.01~3.0 50 100 445.1 1621.3/976.7 60.2 163.4 136
MIL-101(Cr)/GO 0.01~3.0 200 40 40.3 445.7/- 3081 137
Cu-MOF/RGO 0.01~3.0 50 50 520 872.7/- 45.8 138
F-Co-MOF/RGO 0.01~3.0 100/2000 50/550 1202/771.5 2464.2/1- 73.45 140
Fe-MOF/RGO(5) 0.01~3.0 500 200 1010.3 2055.9/891.1 43.3 141
MIL-53(Fe)@RGO 0.01~3.0 100 100 550 -/- 42.3 240.9 142
Co-BDC/CGr 0.01~3.0 100/1000 100/400 1368/818 2566/- 75.42 58.151 143
MOF/RGOn 0.01~3.0 100/1000 100/500 715/348 1677.5/732.6 43.7 145

5.2 MOFs/metal compound

Core-shell structure is a typical core-shell structure and partial hollow structure in which one component material is coated with another component material and assembled in a hierarchical and orderly manner. In a broad sense, it is a general term for a typical core-shell structure with no gap between the core and the shell[147]. The advantages of core-shell MOFs-based composites are: (1) high conductivity of the core material; (2) the porous core-shell structure effectively slows down the volume expansion in the charge-discharge cycle process and provides a certain buffer space; (3) Avoid direct contact of the core material with the electrolyte. Therefore, the reasonable selection of MOFs shell and core materials and the precise control of the size, morphology, thickness and distribution of the shell are undoubtedly very beneficial to improve the structural stability and electrochemical performance of MOFs-based nano-anode materials. The Fe3O4@MOF composite synthesized by the hierarchical assembly method consists of a Fe3O4 microsphere core with a porous MOFs (HKUST-1) shell[148]. After 100 charge-discharge cycles at 100 mA/G, the reversible specific capacity of the composite anode is about 1001. 5 mAh/G, which is significantly better than that of pure Fe3O4(696 mAh/g). Among the known MOFs, HKUST-1 with π-conjugated benzene ring and carboxylate has high conductivity and capacity retention, and can also provide additional capacity[34]. HKUST-1 is beneficial to electron and Li+ migration, and the excellent rate performance of the Fe3O4@MOF composite anode must be related to the improved reaction kinetics. The inner core can be CNTs, metal/nonmetal oxides, organics, and MOFs because it is not limited by composition, dimensionality, size (tens of nanometers to several micrometers), and shape (one-dimensional nanowires to three-dimensional hollow spheres). This opens up the possibility of hybrid hybridization of the ZIF-8 shell with different core materials and obtaining the expected performance. The Zn2SnO4/ZIF-8 nanocomposite anode consists of a Zn2SnO4 nanoparticle core and a ZIF-8 shell on its surface[149]. In the process of Li-Sn alloying/dealloying, the nanoporous properties release the stress caused by severe volume expansion, while the conductivity and thickness of ZIF-8 shell directly affect its cycle stability. In addition, MOFs can also serve as growth templates for internal active nanoparticles in core-shell structures. For example, Wang et al. Utilized the porosity of MIL-101 (Cr) to embed SnO2 nanoparticles inside the pores[150]. Due to the absence of solvent in the pores, the heat-treated multi-core/shell structured SnO2@MIL-101(Cr) nanocomposite is a concave octahedron with a smooth surface. The MIL-101 (Cr) protective shell alleviated the volume change of SnO2 and inhibited the aggregation of powdered nanoparticles, and the lithium storage performance of the SnO2@MIL-101(Cr) composite anode was significantly improved. It can be seen that although the abundant pores and the redox reaction of metal ions are beneficial to Li+ intercalation/deintercalation, it is still necessary to further simplify the synthesis method and improve the conductivity of MOFs/metal oxide composite anode. Affected by the poor overlap of p and d orbitals of metal ions and insulating organic ligands, most MOFs without any structural modification are typically poor conductors. Controlling the composition and morphology and optimizing the rigid-flexible framework structure can improve the conductivity of MOFs. Wang et al. Prepared CuS @ Cu-BTC polyhedra by a simple sulfidation reaction, in which rod-like CuS was distributed on the surface of Cu-BTC[151]. The active material CuS can improve the conductivity, while Cu-BTC can provide porosity and chemical stability, which can effectively inhibit the conversion of Li2S into polysulfides by improving the reversibility of the conversion reaction. The capacity of CuS (70 wt%) @ Cu-BTC anode reached 1609 mAh/G (higher than the theoretical capacity) after 200 cycles at 100 mA/G. ~ 490 mAh/G at 1000 mA/G. It can be seen that the content of CuS is not only related to the sulfidation time, but also affects the porosity and morphology of CuS @ Cu-BTC composite anode, and then affects its lithium storage performance.

5.3 Metal oxide/metal oxide

The conductivity of transition metal oxides is relatively low, which is easy to produce obvious volume effect and serious agglomeration in the cycle process, and even cause electrode pulverization and rapid capacity decay. Only when the lithium storage sites are abundant, the Li+ at the electrode/electrolyte interface and the adsorption/desorption table/interface migrate rapidly, and the diffusion distance of Li+/ electrons is shortened, the electrode materials can obtain good lithium storage performance. Metal oxide/metal oxide composites are generally composed of dual or multiple metal oxides, which can play the role of different metal oxides at the same time, and can also enhance the mechanical stability, electrochemical reactivity and ionic conductivity of each component by means of special structure and synergistic effect.This requires that the design and synthesis of metal oxide/metal oxide composites should not only consider the properties and functions of each component material, but also provide a conductive network for efficient electron and ion transport by delicately constructing and assembling porous structures to uniformly distribute them on the nanoscale. For example, hollow core-shell metal oxide/metal oxide nanocomposites with different morphologies can maintain the integrity of the electrode to the greatest extent, prevent capacity fading and prolong the cycle life. Therefore, metal oxide/metal oxide nanocomposites have become one of the research hotspots in the field of anode materials for lithium-ion batteries.
Using bimetallic Co-Ce-MOF as a precursor, Kang et al. Designed and synthesized a heterostructured Co3O4/CeO2 composite[152]. The Co3O4/CeO2 (molar ratio of Co/Ce = 5 ∶ 1) anode has a high reversible capacity (1131.2 mAh/G after 100 cycles at 100 mA/G) and cycle stability (835.3 mAh/G at 2000 mA/G), and the good lithium storage ability is mainly due to the conversion reaction and interfacial capacitance. However, the introduction of semiconductor CeO2 does not affect the diffusion of Li+ in the Co3O4, which provides a basis for the modification of electrode materials by rare earth components. The electrochemical performance of hollow MOFs derived anode is related to the interfacial interaction between the core and the shell and the stress distribution, but the effect of the expansion/contraction of the core on the structural stability can not be ignored. Zhang et al. Synthesized Fe2O3/SnO2, Fe2O3/GeO2, and Fe2O3/Al2O3 multi-component microboxes with complex hollow structures using a controlled template reaction method[153]. Since the beneficial effect of the hollow hierarchical structure is maximized, the volume expansion/contraction is effectively suppressed, the diffusion distance of the Li+ is shortened, and the capacity retention rate of these multi-shell composite anodes is high. In addition to the known advantages of the hollow structure, the synergistic effect of the tightly combined Fe2O3 and SnO2 can also provide additional capacity for the multi-component microbox anode, and the specific lithium storage mechanism needs to be further studied. Li et al. Prepared ZnO/NiO nanospheres composed of nanorod shell and microsphere core by controlling the calcination temperature. The unique core-shell structure retained from bimetallic Zn-Ni MOFs is beneficial to electrolyte diffusion and ion migration. The reasonably hybridized nanosized ZnO and NiO improve the electrochemical kinetics and reactivity with Li+, effectively suppressing the volume effect[154]. Therefore, the high specific surface area ZnO/NiO composite anode shows excellent specific capacity (1008.6 mAh/G after 200 cycles at 0.1 A/G), rate capability (437.1 mAh/G at 2 A/G), and long-term cycling stability (592.4 mAh/G after 1000 cycles at 0.5 A/G). Xu et al. Prepared hollow Co3O4/TiO2 polyhedra using cation exchange and subsequent heat treatment[155]. During the synthesis, ZIF-67 served both as a host for the exchange of Ti+ and as a template for the formation of Co3O4/TiO2 composites. The structurally stable TiO2 continuously provides high capacity for Co3O4 and also provides a conductive path for fast electron transport. After 200 cycles at 500 mA/G, the reversible capacity of the Co3O4/TiO2 anode is 662 mAh/G, which is about 96.9% of the initial capacity, and is significantly better than that of the pure Co3O4 anode with the same structure. In particular, this novel synthesis method can also be used to construct more complex nanostructured composites. Hu et al. Fabricated hollow octahedral CuO/Cu2O composite by pyrolytic [Cu3(btc)2]n template, and analyzed the formation process of porous hollow structure in combination with non-equilibrium interdiffusion, volume decrease, and gas release[156]. According to the theoretical capacity of CuO (670 mAh/G) and Cu2O(670 mAh/g), the conversion mechanism and the proportion of CuO, the theoretical capacity of CuO/Cu2O composite anode was deduced to be 520 mAh/G.
Although the cyclicity and lithium storage performance of CuO/Cu2O anode materials have been improved due to the hollow porous shell and different components of metal oxides, the problems of low inherent conductivity, poor transport kinetics and electrode pulverization have not been fundamentally solved. In order to solve the volume effect of metal oxide electrodes, Guo et al. Designed and synthesized CuO @ NiO hollow multilayer spheres derived from Cu-Ni-BTC, which are composed of a three-layer sphere-in-sphere structure and interconnected nanoparticles, and the elements Ni and Cu are distributed in a gradient from core to shell[157]. After 200 charge-discharge cycles, the reversible capacity of CuO @ NiO composite anode is 1061 mAh/G, which is much higher than the theoretical capacity of CuO and NiO anode, which is mainly related to the stepwise intercalation reaction initiated by the commensurate composition from core to shell, the smooth Li+/ electron diffusion and the coordinated volume change. Metal-oxygen clusters in MOFs can be directly transformed into porous transition metal oxides without long-range atomic migration during pyrolysis. The Cr2O3@TiO2 nanocomposite is a concave octahedron synthesized by calcination of the MIL-101(Cr)@TiO2 template in argon and air sequentially, which is a typical multi-core core-shell structure[158]. Due to the protective effect of the chemically stable TiO2 shell, the pulverization of internal particles is effectively inhibited, and the structural stability of the electrode material is improved. The reversible capacity of the Cr2O3@TiO2 composite anode is 510 mAh/G after 500 cycles at 0.5 mA/G, which is higher than that of nano Cr2O3(200 mAh/g). More importantly, its capacity recovery rate is also outstanding. The synthesis of mesoporous CuO@TiO2 octahedra involves two stages: uniform growth of TiO2 shell on Cu-MOF (HKSUT-1) template and controllable temperature pyrolysis[159]. This core-shell structure design fully combines the advantages of different components and solves the pulverization problem of most metal oxide-based electrodes caused by severe volume change. After 200 cycles at 100 mA/G, the reversible specific capacity of the CuO@TiO2 composite anode is 692 mAh/G, which is significantly better than that of pure CuO and the previously reported CuO-based nanomaterials. At high current densities of 1600 and 3200 mA/G, the capacities were still 441 and 387 mAh/G, respectively. Prussian blue and Prussian blue materials with FCC structure are common templates or precursors for the synthesis of metal oxides.Its derivatives generally have the characteristics of high specific surface area, connected internal pores and uniform element distribution from inside to outside, and low-cost composite materials with controllable composition and special morphology can be obtained by adjusting the preparation process and precursor composition. Porous Fe2O3-CuO cubes are composites prepared by pyrolysis of a bimetallic MOF synthesized based on Prussian blue (Fe4[Fe(CN)6]3)[160]. After 120 cycles at 500 mA/G, the specific capacity of the Fe2O3-CuO composite anode is 744 mAh/G; The discharge capacity at high current density (2000 mA/G) is still higher than 500 mAh/G. In particular, the cycle stability and the rate capability are obviously superior to those of other known hybrid transition metal oxides. Fig. 4 is a schematic diagram of the synthesis of the Fe2O3 nanotube @Co3O4 composite, and it can be seen that the Co3O4 host matrix is uniformly coated by the Fe2O3 nanotubes[161].
图4 Fe2O3纳米管@Co3O4复合材料的合成示意图: (Ⅰ) MIL-88B纳米棒、Co2+和 2-methylimidazole (2-MIM)自组装成MIL-88B@ZIF-67复合材料; (Ⅱ) 经空气中热处理转变为Fe2O3纳米管@Co3O4复合材料[161]

Fig. 4 Schematic illustration of the formation process of the Fe2O3 nanotubes@Co3O4 composite. (Ⅰ) Self-assembly of MIL-88B nanorods, Co2+ ions, and 2-methylimidazole (2-MIM) to a MIL-88B@ZIF-67 composite. (Ⅱ) Transformation to Fe2O3 nanotubes@Co3O4 composite through thermal treatment in air[161]

The hollow structure electrode can effectively alleviate the stress-induced structural changes during the long-term electrochemical reaction. Compared with the single-shell hollow structure, the multi-shell hollow structure has the advantages of high mass fraction of active material, complex internal pores, large specific surface area and controllable shell composition. The multi-shell Co3O4@Co3V2O8 nanobox is pyrolyzed from ZIF-67 @ amorphous -Co3V2O8 in air, and the number and composition of its shells can be precisely controlled by simply adjusting the amount of VOT in the reaction system[162]. Similarly, NiFe2O4/Fe2O3 hierarchical nanotubes and Fe2O3@NiCo2O4 porous nanocages were derived from the pyrolysis of core-shell structured MOF precursors (Fe2Ni MIL-88/Fe MIL-88 and Co3[Fe(CN)6]2@Ni3[Co(CN)6]2, respectively)[163,164]. Although the two composite anode materials do not retain the structural characteristics of the precursor, their electrochemical performance is significantly higher than that of the single component metal oxide derived from MOFs.
Zhong et al. Prepared nanostructured Co3O4-CoFe2O4 composites by pyrolyzing MOF-74-FeCo, whose morphology and composition mainly depend on the Fe3+/Co2+ molar ratio in the bimetallic MOF[165]. Compared with pure Co3O4 and other Co3O4-CoFe2O4 composite anodes, only Co3O4-CoFe2O4-12 exhibited the highest initial discharge capacity (1328 mAh/G), reversible capacity (940 mAh/G after 80 cycles at 100 mA/G), and rate capability. This is because Fe3+ promotes the transformation of MOF-74-Co from an irregular morphology to a relatively regular spherical structure, resulting in the formation of fine and dispersed Co3O4-CoFe2O4 nanoparticles, whose phase and structural stability are not changed by repeated charge-discharge cycles, and the specific capacity hardly decays after 80 cycles. Xu et al. Successfully prepared ZnO/ZnCo2O4 nanosheets composed of both primary ZnO and ZnCo2O4 nanoparticles with uniform size and interconnection by one-step pyrolysis of Zn-Co-MOF precursor[166]. A reversible capacity of 1016 mAh/G and a coulombic efficiency of almost 99% were obtained for the three-dimensional hierarchical ZnO/ZnCo2O4 porous anode after 250 cycles at 2 a/G. The capacity is still 630 mAh/G at 10 A/G. On the one hand, the size and morphology of MOFs core/shell composites are directly controlled by selecting the appropriate template. On the other hand, special functional groups are used to modify the surface of the core component to control the binding and coating of the MOFs core, and to inhibit the growth of MOFs on the precursor solution and the core surface. Benefiting from the multi-component functional complementation and structural characteristics, the specific capacity, cycle life, and rate capability of the Ni(OH)2@ZIF-8(Zn,Co) nanowire-derived core-shell NiO@ZnCo2O4 composite anode are significantly improved, while this synthesis method provides the possibility for the MOFs shell to grow on the surface of core materials with arbitrary chemical composition, structure, dimensionality, and size[167]. Hierarchical ZnO/ZnFe2O4 mesoporous submicron cubes are composed of uniformly dispersed ZnO and highly electrochemically active ZnFe2O4 nanophase[168]. Compared with ZnFe2O4, the excellent electrochemical performance of the ZnO/ZnFe2O4 composite anode mainly depends on the unique structure and the synergistic effect of two-component ZnO and ZnFe2O4 active phase. The discharge capacity of the porous ZnO/ZnFe2O4 composite anode composed of nanoparticles with a particle size of about 5 nm was 804 mAh/G after 500 cycles at 1000 mA/G. 496 mAh/G even after 1000 cycles at 2000 mA/G[169]. In addition to the porous nanostructure providing additional space to coordinate the volume change, the ZnO phase as a buffer not only spatially separates the coexisting ZnFe2O4 nanophase, but also further prevents its self-aggregation during cycling[170]. The electrochemical properties of the metal oxide/metal oxide composite anode materials are summarized in detail in Table 5.
表5 金属氧化物/金属氧化物复合负极材料及其电化学性能

Table 5 Metal oxide/metallic oxide composites as anode materials and their electrochemical performances

Materials Template/precursor Voltage range (V) Current density (mA/g) Cycle
number
Overall capacity (mAh/g) Initial discharge/
charge capacity
(mAh/g)
Initial
coulomb
efficiency
(%)
Specific surface area (m2/g) ref
5Co3O4/CeO2 Co-Ce-MOF 0.01~3.0 100/500 100/300 1131.2/901.4 1090.1/873.6 77.6 19.265 152
Fe2O3/SnO2 Fe-MOF 0.05~3.0 200 100 500 1751/904 43 153
ZnO/NiO Zn-Ni-MOF 0.005~3.0 100/500 200/1000 1008.6/592.4 1221.7/769.2 62.9 21.34 154
Co3O4/TiO2 ZIF-67 0.01~3.0 500 200 642 662/535 155
CuO/Cu2O [Cu3(btc)2]n 0.01~3.0 100 250 740 727/513 9 156
CuO@NiO Cu-Ni-BTC 0.05~3.0 100 200 1061 1218/856 16.3 157
Cr2O3@TiO2 MIL-101(Cr)
@TiO2
0.05~3.0 0.5C 500 510 1138/- 146 158
CuO@TiO2 HKUST-1/TiO2 0.01~3.0 100 200 692 1092/780 71.4 88.9 159
Fe2O3-CuO PB 0.01~3.0 500 120 744 1070/795 74 160
NiFe2O4@Fe2O3 Fe2Ni MIL-88/
FeMIL-88
0.01~3.0 100 100 936.9 1400.9/989.1 39.2 163
Fe2O3@NiCo2O4 Co3[Fe(CN)6]2
@Ni3[Co(CN)6]2
0.01~3.0 100 100 1079.6 1311.4/902.7 12.72 164
Co3O4-CoFe2O4-12 MOF-74-FeCo-xy 0.01~3.0 100/500 80/80 940/598 1328/918 165
ZnO/ZnCo2O4 ZnO@ZIF-8 NRAs 0.01~3.0 1000/2000 200/250 870/1016 1299/987 76 20 166
NiO/ZnCo2O4 Ni(OH)2@ZIF-8 0.01~3.0 2000 100 1002 -/- 44 167
ZnO/ZnFe2O4 Zn3[Fe(CN)6]2 0.01~3.0 1000/2000 200/200 837/701 1892/1371 70 54.3 168
ZnO/ZnFe2O4 Prussian Blue 0.01~3.0 1000/2000 500/100 804/497 1293/826 63.8 39.0 169
ZnO/ZnFe2O4 ZnFe PBA 0~3.0 200 200 704 998.4/704.9 70.6 170

5.4 Metal oxide/carbon-based material

It is well known that the performance of electrode materials mainly depends on their basic properties, morphology and structure. Low intrinsic conductivity and obvious volume effect seriously restrict the application of transition metal oxides (such as Co, Ni and Fe) with high theoretical capacity in the field of anode materials. On the basis of fully retaining the activity of MOFs derivatives, combined with the advantages of carbon-based materials such as carbon nanotubes, graphene, carbon fibers and amorphous carbon in terms of stability and conductivity, different metal oxide/carbon-based materials can be directly synthesized by using MOFs containing metal ion centers and organic ligands as templates or precursors. Although the control process is complex and there are many influencing factors, the comprehensive performance of anode materials can be improved by optimizing the heat treatment process and accurately controlling the morphology and structure.
Compared with solid materials with the same size, low-density hollow structural materials have clear internal voids and higher specific surface area, especially the characteristics of difficult pulverization, coordinated volume change and micro/nano structure, which make them suitable for energy storage applications. The core-shell hierarchical porous Co3O4/C dodecahedral composite anode derived from ZIF-67 by two-step heat treatment exhibited a specific capacity of 1100 mAh/G after 120 cycles at 200 mA/G, which is much higher than the theoretical capacity of Co3O4 (890 mAh/G)[171]. Kinetic analysis shows that most of the stored charge comes from the capacitive process, and the rate-determining step is surface confinement rather than solid-state diffusion[172]. In addition, the improvement of electrode conductivity by carbon matrix cannot be ignored. In the solid-solid conversion process, the porosity and long-range order of the MOFs template facilitate the rapid entry and exit of small molecules and ions, and the temperature and atmosphere directly affect the structure and composition of the derivatives. MOFs-derived core-shell nanomaterials usually have high specific surface area and stable hollow structure, which will not affect the product morphology due to template removal. By fully combining the advantages of the retained core-shell structure of MOFs and the matrix of different material systems, the multi-component electrochemical reaction occurs on the hybrid matrix of different material systems, which effectively solves the serious pulverization of electrodes and the resulting rapid capacity decay. In this case, the defined matrix causes the volume change in a progressive manner rather than in a fixed pattern, i.e., the strain produced by the reacted phase is coordinated by the unreacted components. At the same time, the combination of carbon-based materials can also provide abundant redox reaction sites for transition metal oxides and improve conductivity[173]. Similarly, both core-shell ZnO @ C and Co3O4@C polyhedra are derived from MOFs[174]. With the synergistic effect of the three-dimensional porous structure and the surface conductive carbon layer, the reversible capacity of the ZnO @ C and Co3O4@C composite anode is 526 and 721 mAh/G after 500 cycles at 250 mA/G, respectively; More than 1000 charge-discharge cycles at 2.0 A/G. The kinetic analysis indicates that the surface-controlled pseudocapacitive behavior directly affects the lithium storage ability of the ZnO @ C and Co3O4@C composite anode. It is worth noting that the binding force of MOFs-derived metal oxide/carbon-based nanocomposites is stronger than that of metal oxide directly coated with carbon layer. The mesoporous spindle-shaped CuO/C composite prepared by one-step annealing well retains the Cu-MOF template morphology synthesized by microwave-assisted method[175]. Although they are all typical hollow structures, the electrochemical performance of porous CuO/C microcube anode is far inferior to that of mesoporous spindle-shaped CuO/C composite anode due to the different composition of Cu-MOF template and pyrolysis temperature[176]. Peng et al. Prepared mesoporous Mn3O4/C composite by one-step calcination of 3D Mn-PBA template and studied the effect of calcination temperature on its surface morphology[177]. The results show that the reversible capacity and rate capability of the Mn3O4/C microsphere anode are significantly better than those of the sponge network Mn3O4/C anode[178]. After 100 cycles at 50 and 500 mA/G, the specific capacity of the 3D rhombic MnO/C composite anode is 835 and 586 mAh/G, respectively, which is much higher than that of the porous carbon and Mn2O3 derived from the same Mn-MOFs[179]. This is because of the beneficial effects of the uniformly distributed MnO nanoparticles on the surface and the conductive carbon matrix containing a large number of pores in increasing the active sites for lithium storage, shortening the ion and electron transport distance, avoiding agglomeration, and alleviating the volume effect. In addition, because the morphology of Mn (PTA) -MOFs changes with the molar ratio of Mn2+ and benzoic acid, Sun et al. Synthesized MnO/C composites with different morphologies on the basis of self-sacrifice and thermal conversion[180]. Only the reversible capacity and rate capability of the spindle-shaped MnO/C microrods were higher, and the capacity did not decay significantly even after 200 cycles at a high current density of 1000 mA/G. Therefore, not only the reasonable selection of organic ligands and their dosage is very necessary, but also the difference of the connection mode between the same metal ion and different organic ligands affects the final morphology and spatial structure of metal oxide/carbon composites, and the electrochemical performance is also different.
Although the controlled heat treatment process can simultaneously synthesize ordered nanostructured transition metal-based nanoparticles and carbon frameworks, there are still many challenges in other methods to uniformly embed nanoparticles into porous carbon frameworks. The hierarchical ZnO/C hollow spheres were derived from the precursor Zn-BTC by one-step annealing treatment, and the ultra-small ZnO nanodots (average particle diameter < 5 nm) were uniformly dispersed in the carbon nanosheet matrix on the surface of the porous hollow structure[181]. The reversible capacity of ZnO/C composite anode is 919 mAh/G after 100 cycles at 100 mA/G. The reversible capacity is maintained at 741 mAh/G after 120 cycles at 500 mA/G. It should be noted that the irreversible capacity fading at the first cycle is due to the irreversible reduction of Li+ during the formation of surface SEI layer, which is a common problem for most anode materials. The ZnOQDs @ C composite is obtained by controlled pyrolysis and structural reorganization of IRMOF-1, which is composed of ZnO QDs without inherent defects or agglomeration and porous carbon matrix[182]. The reversible capacity of the hierarchical structure anode exceeds the theoretical prediction, and the cycle performance is also better than the reported value[183]. In the highly long-range ordered structure of the MOF precursor, each ZnO quantum dot is separated by amorphous carbon that lacks long-range order, and interfacial charge storage at the surface of the ZnO quantum dot and weak binding of lithium, reaction with the electrolyte, and charge separation at the SEI layer may occur and provide additional capacity[184]. In contrast, the known carbon-coated ZnO nanoparticles are more or less agglomerated during charge-discharge cycles, which seems to ignore other possible ways of lithium storage[185]. Compared with the commercial Fe3O4 anode and the spindle-shaped Fe2O3 anode prepared by direct carbonization of Fe-MOFs in air, the porous Fe3O4/C octahedral anode has a higher capacity, that is, a specific capacity of 861 mAh/G after 100 cycles at 100 mA/G[186]. The cycling stability and rate performance of the Fe3O4/C composite anode are also outstanding due to the porous nanostructure enhancing the diffusion speed of Li+ and the highly conductive randomly doped amorphous carbon framework, which effectively prevent the collapse of the electrode structure and inhibit the undesirable growth of SEI layer. The Fe3O4@C nanocomposite was prepared from Fe-MOFs with small organic ligands (formic acid) via high-temperature carbonization, and the Fe3O4 nanoparticles were coated by carbon layers (< 1 nm)[187]. This is obviously completely different from other core-shell metal oxide @ C composites with metal oxide as the core. The reversible capacity of the Fe3O4@C nanoanode is 1041 mAh/G after 50 cycles at 100 mA/G; The rate capability is lower than that of the Fe3O4/C anode synthesized by in situ pyrolysis[188]. Wang et al. Successfully prepared SnO2@C nanocomposites by taking advantage of the strong adsorbability of porous MOFs, that is, monodisperse SnO2 nanoparticles distributed within the three-dimensional porous carbon network[189]. Especially, the porous carbon structure not only improves the wettability of the electrode/electrolyte and ensures the ion mobility of the electrode reaction, but also prevents the SnO2 nanoparticles from falling off from the electrode. Therefore, the SnO2@C nanocomposite anode exhibits good reversible specific capacity, cycling stability, and durability. During the pyrolysis of In-MOF In N2 atmosphere, the reduction potential of In ions to form In2O3 nanoparticles is lower than − 0.27 V. Jin et al. First treated In-MOF (MIL-68 (In)) by one-step calcination to homogeneously embed In2O3 nanoparticles into carbon matrix[190]. The initial discharge capacity of the porous In2O3/C nanocomposite anode is about 1410 mAh/G; The reversible discharge capacity is 720 mAh/G after 150 cycles at 100 mA/G. However, the effects of different precursors on the composition, morphology and properties of In2O3/C composites need to be investigated. The MOF-derived In2O3/HPNC is an ultrastable composite anode material, and the morphological characteristics are shown in Fig. 5[191]. After 2000 cycles at a current density of 1000 mA/G, the specific capacity of the In2O3/HPNC composite anode was still as high as 623 mA H/G, and the capacity fading was only 0.017% per cycle. This is because the N-doped hierarchical porous structure can not only ensure the mechanical strength and coordinate the volume expansion of In2O3 nanocrystals, but also provide sufficient electron and Li+ interpenetration paths and fast migration during charge-discharge process. Due to the inherent chemical passivity and high hydrophobicity, it is difficult to directly deposit other materials on the surface of multi-walled carbon nanotubes (MWCNTs). The hierarchically porous MWCNTs/Co3O4 composite anode containing nano-sized units was prepared by embedding MWCNTs into Co3O4 polyhedrons through heat treatment of MWCNTs/ZIF-67 precursor[192]. Similarly, nano-MWCNTs/ZnO composite anode was synthesized using ZIF-8/MWCNTs as precursor[193]. Under the joint action of Co3O4 or ZnO with high theoretical capacity, MWCNTs with high conductivity and a polyhedral structure, the porous MWCNTs/Co3O4 or MWCNTs/ZnO nanocomposite anode is superior to the corresponding metal oxide in specific capacity, cycle stability and rate capability, and shows super lithium storage capacity and wide application prospect. For the NiO/CNT composite anode, CNTs are not only distributed on the surface of NiO microspheres, but also embedded in the interior, thus connecting with each other and forming a 3D conductive network[194]. Especially, the maximum reversible capacity of NiO/CNTs-10 anode is 812 mAh/G after 100 cycles at 100 mA/G. It is still stable at 502 mAh/G after 300 cycles at a high current density of 2000 mA/G. Obviously, the incorporation of one-dimensional CNTs into metal oxides is an effective way to improve the transport speed of Li+ and the lithium storage performance. Although the hierarchical tubular structure CNT/Co3O4 composite exhibits excellent rate capability and electrochemical reactivity with a specific capacity of 782 and 577 mAh/G at 1000 and 4000 mA/G, respectively, the two-step synthesis process is significantly more complex[195]. In addition, one-dimensional carbon fiber is one of the ideal templates for the synthesis of electrode materials because of its fast electron transfer rate, high strength and strong buffer volume effect. For example, the porous CFs@Co3O4 composite is composed of carbon fibers and Co3O4 polyhedra on its surface, which well retains the structural morphology and size uniformity of the CFs @ ZIF-67 precursor[196]. Benefiting from the buffering effect of the porous core-shell structure and the synergistic effect between Co3O4 with higher theoretical capacity and carbon fiber, the CFs@Co3O4 composite anode shows good capacity, cycle stability and rate performance, which is very potential to replace traditional metal oxide and carbon fiber anode materials. Flexible graphene can also be used as a conductive matrix for MOFs to prepare metal oxide/graphene composites due to its special physical, chemical and mechanical properties. Ji et al. Synthesized a free-standing porous 3DGN/CuO composite anode by in-situ self-assembly growth of Mn-MOF on the 3DGN matrix by impregnation method, and then by subsequent heat treatment, that is, CuO octahedral nanoparticles were uniformly distributed in the three-dimensional graphene framework[197].
图5 In2O3/HPNC复合材料的形貌特征:(a)与(b) SEM图像;(c)与(d) TEM图像;(e) HRTEM图像(内部为SAED谱)(f~k) TEM暗场图像和相应元素分布[191]

Fig. 5 Morphological features of In2O3/HPNC composite: (a) and (b) SEM images, (c) and (d) TEM images, (e) HRTEM images, and the inset is the SAED pattern, and (f~k)Dark-field TEM image and corresponding elemental mapping[191]

Similarly, Shao et al. Synthesized NiO/GF composites by growing Ni-MOF on the surface of graphene foam by solvothermal method and then heat treating at high temperature[198]. Flexible conductive graphene not only supports active materials, but also prevents NiO particles from diffusing into the electrolyte. It can be seen that precise control of active material loading, rational design of porous hierarchical nanostructure and utilization of interfacial interaction are the key to obtain the best lithium storage performance. Although the electrochemical properties of 3DGN/CuO and NiO/GF binderless anodes are better than those of the corresponding metal oxide and graphene electrodes, the limited active sites and easy pulverization still limit their application to some extent. Reduced graphene oxide (rGO) has no oxygen functional groups in its structure, and has high conductivity, chemical stability, flexibility, specific surface area and porosity. Considering that the organic combination of multifunctional MOFs and graphene fiber structure can obtain more novel functions and physicochemical properties, Zhang et al. Prepared one-dimensional porous Fe2O3/rGO composite by two-step annealing treatment of MIL-88-Fe/GO precursor[199]. Compared with the composite aerogel prepared by common freeze-drying and annealing processes, the Fe2O3/rGO composite anode has similar electrochemical active surface area and higher density, and has more excellent specific capacity, cycling stability and rate capability. The initial Coulombic efficiency of the nanoporous rGO/Co3O4 composite anode is about 70%, and the rate capability and long-term cycling stability are outstanding due to the SEI layer formation, interfacial lithium storage, partial undecomposed Li2O, and irreversible decomposition of the electrolyte. Obviously, the positive effect of rGO on improving the conductivity and reducing the charge exchange resistance at the electrode/electrolyte interface cannot be ignored[200]. In addition, GO/Ni-MOFs were synthesized by the strong electrostatic adsorption of GO on the surface of MOFs, and then the RGO/NiO composite anode was obtained by subsequent heat treatment[201]. The high specific surface area and medium-size porous structure ensure the effective charge transfer efficiency during charge-discharge process, and the active RGO outer layer protects the electrode structure from destruction.
N is close to the C atom diameter, but the electronegativity is slightly higher. The inessential defects produced by N doping can improve the reactivity and conductivity of carbon[130]. In addition, the doping of N atoms can also change the electronic characteristics, provide more active sites for the adsorption of Li+, enhance the interaction between carbon structure and Li+, and improve the diffusion and migration kinetics of Li+, which are beneficial to improve the electrochemical performance of electrode materials. Although the pyrolytically synthesized MnO/C-N composite well retains the morphological characteristics of the rutile Mn-MOFs (Mn-PBI) precursor, the size is slightly reduced due to pyrolysis and volume shrinkage[202]. Temperature is an important parameter affecting the amount of N-doping when MOFs are used as sacrificial templates to prepare high-performance electrode materials. The capacity (higher than the theoretical capacity of MnO, 765 mAh/G), cycling stability and rate capability of the MnO/C-N nanoanode are higher. In particular, the higher N doping content (5. 10 wt%) makes the MnO/C-N-500 anode structure electron-poor, which is easier to combine with Li atoms strongly and accept more charges from the Li+[203,128]. In order to make all active sites of electrode materials participate in the electrochemical reaction, Chu et al. Made NiO nanocrystals grow on N-doped porous carbon substrate by optimizing the MOFs (Ni-NTA) derivation strategy, which solved the problem of NiO nanoparticles falling off from the electrode surface due to mechanical stress during repeated charge-discharge[204][205]. Therefore, the rod-like NiO @ N-C composite anode with rough surface has a specific capacity of 1373 mAh/G after 200 cycles at 100 mA/G. After 1000 cycles at 1000 mA/G, the capacity is still as high as 877 mAh/G. These results indicate that the conductivity and electrochemical kinetics of the NiO @ N-C composite anode are good, which is more conducive to the diffusion, transport and storage of Li+. In the process of direct pyrolysis of MOFs template, the high thermal stability is helpful for the in situ transformation of organic chains into N-doped carbon materials, avoiding the volatilization of residual organic matter. Therefore, as sacrificial templates and auxiliary carbon and nitrogen precursors, nitrogen-containing MOFs with high thermal stability can be directly calcined in air to synthesize metal oxide/N-doped porous carbon composites. Upon direct calcination of the Fe-ZIF precursor, the catalytic effect of Fe atoms accelerates the decomposition and carbonization of organic ligands, resulting in uniform intercalation of ultrafine Fe2O3 particles into the N-doped carbon spherical shell[206,207]. Compared with the core-shell structure, the utilization rate of atoms in this unique internal hollow structure is higher. The hollow interior and N-doped carbon matrix of the Fe2O3@N-C composite anode can be regarded not only as a nanoelectrochemical reactor, but also as a storage for Li+ and electrolyte. Therefore, the reversible capacity is still stable at 1142 mAh/G after 100 cycles at 1000 mA/G. The three-dimensional hierarchical NCW@Fe3O4/NC composite anode is a nanomaterial synthesized by self-assembling Fe-MOFs on N-doped carbon nanowires and then carbonizing, and the ultra-small Fe3O4 nanodots are uniformly embedded in the N-doped carbon nanowires[208]. After 600 cycles at 1 A/G, the reversible capacity of the NCW@Fe3O4/NC composite anode is 1741 mAh/G; The capacity is still 723 mAh/G at a high current density of 10 A/G. Obviously, this self-assembly process is not only suitable for other types of MOFs to uniformly coat one-dimensional carbon nanotubes, two-dimensional graphene and three-dimensional N-doped carbon nanomesh, but also can realize the controllable growth of other functional materials (such as metal oxides, sulfides and selenides) on different carbon nets. Co3O4/N-C polyhedron, fish-scale structure Co3O4/N-C, Co3O4/N-PC dodecahedron, and starfish-like Co3O4@N-C are porous nanomaterials synthesized by using Co-TATB, nitrogen-rich Co-MOFs, ZIF-67, and Co-MOFs as sacrificial templates or precursors, respectively, with highly uniform distribution of Co3O4 nanoparticles on the N-doped carbon matrix[209][210][211][212]. Without exception, the capacity, rate performance and cycle stability of these composite anodes are significantly improved, but there are still some differences due to the preparation process, morphology and structure, and the amount of N doping. The hollow structure of transition metal oxides is easy to construct, but the size distribution, morphology uniformity, pore volume, shell definition and functional groups are difficult to control. Ding et al. Prepared N-doped Co3O4@PC microspheres with multi-shelled hollow structure by thermal oxidation of Co-MOFs precursor at different temperatures[213]. The coexisting pyrrole and pyridinic N can also provide additional electrochemical active sites, although the amount of N doping is low. The reversible capacity of N-doped Co3O4@PC anode is 1701 mAh/G after 60 cycles at 100 mA/G, especially 427 mAh/G at 5000 mA/G, due to the combined effects of the porous shell structure composed of nanocrystals, the gap between the shells and the porous carbon. The results show that there are two ways to combine CNTs with nano-sized metal oxides: one is to form metal oxides on the surface of CNTs, which retains the electron and ion migration path, but the volume expansion can not be effectively alleviated. The other is to intercalate metal oxides into the pores of CNTs, which can alleviate the volume expansion, but hinder the electron and ion migration to some extent. Obviously, the in-situ growth of CNTs and metal oxides is a feasible method to solve the above contradiction. As bifunctional metal-organic precursors and self-sacrificial templates, MOFs enable the construction of heterogeneous atom-doped carbon-based architectures with specific surface properties through controlled pyrolysis or post-treatment. Actually, the interlayer spacing of 2D MOFs is wide and the interlayer interaction is weak, which is favorable for Li+ intercalation/deintercalation. Compared with the typical three-dimensional structure, the thin layer can provide shorter ion migration distance and expose more active centers. The heterostructured CoO-NCNTs composite anode is derived from layered Co-rich 2D-MOFs, the active CoO coated on the top of NCNTs provides good ionic conductivity and alleviates the volume change, and the tightly entwined NCNTs constitute an effective conductive network for ion migration[214]. At 500 mA/G, the cycle life of the CoO-NCNTs composite anode is more than 2000 cycles, and the capacity fading per cycle is only 0.0063%. Making full use of the advantages of MOF and graphene-based aerogel, a large number of Co3O4 particles of ZIF-67 @ NGA were embedded into the N-doped graphene network after heat treatment, which solved the problem that the metal oxide particles could only be deposited on the graphene sheets but could not enter the interlayer[215]. The reversible capacity of this porous hierarchical structure Co3O4@NGN composite anode is not only higher than that of graphite and bulk Co3O4 anode, but also higher than that of MWCNTs/Co3O4 anode[192]. In addition, the ZnO @ C composite anode is composed of spherical particles with uniform submicron size, and each core-shell composite particle contains a ZnO core, a conformal carbon layer with precisely controlled thickness, and a boundary region with uniform intermingling inside. The good charge-discharge reversibility, cycle life, and rate capability mainly depend on the synergistic effect of the N-doped conductive carbon layer, higher structural/interfacial stability, and the LiF-rich SEI layer to improve the Li+ diffusivity[216].
In addition, doping metal elements is also the main means to improve the lithium storage performance of metal oxide/carbon matrix composites. For example, Co can simultaneously improve the conductivity and carbon graphitization degree of Co-doped ZnO @ C (CZO @ C) composite anode[217]. After 50 cycles at 100 mA/G, the reversible capacity of the CZO @ C composite anode is 725 mAh/G, which is significantly higher than that of ZnO @ C (725 mAh/G). Similar to ZnO, the reaction of CZO @ C composite anode with Li includes not only the reversible transformation of metal oxide into nano-sized metal and lithium oxide matrix, but also the alloying/dealloying process of lithium and metal. Li et al. Obtained Ti-doped mesoporous CoO @ C octahedra by pyrolyzing bimetallic Co-Ti-MOF to embed ultra-small CoO nanoparticles into the surface and interior of the carbon framework[218]. As the heterogeneous Ti doping introduces more defects and reduces the nanoparticle size, which is beneficial to the maximization of reactive active sites and the improvement of pseudocapacitive behavior, the CoO @ C nanocomposite anode has a reversible capacity of up to 1180 mAh/G after 150 cycles at 200 mA/G. Although the addition of new metal elements improves the structural stability and reversible capacity of two-component metal oxides, the conductivity is still reduced by agglomeration during cycling. Metal ions enter the structure of MOFs through pores and gaps, and bind to and integrate with the active sites. Therefore, on the basis of the interconnection of metal ions and organic chains, the introduction of heterogeneous metal ions can form a new type of MOFs precursor with controllable structure. Zhang et al. Synthesized C/NiCo2O4 composites by heat treatment of bimetallic Ni-Co-MOF precursor[219]. Due to the cross-connection of NiCo2O4 and amorphous carbon to effectively alleviate the volume expansion and improve the conductivity, the specific capacity, rate capability, and cycle stability of the C/NiCo2O4 anode material are good. Similar to other metal oxide anodes, the obvious capacity fading and increasing trends of C/NiCo2O4 anodes come from the formation of SEI layer and the activity of materials, respectively. The theoretical capacity of hierarchical perovskite structure CoTiO3 is about 520 mAh/G. Although the structure remains stable during cycling, the low conductivity seriously restricts the application of CoTiO3 electrode. The amorphous carbon formed after the carbonization of Co-MOF was uniformly coated on the surface of CoTiO3 nanocrystals under the catalysis of high temperature and metal ions[220]. In addition to making up for the lack of conductivity of CoTiO3, the amorphous carbon layer can also avoid the agglomeration of magnetic CoTiO3 particles and improve the stability of the structure. The reversible capacity of the C/CoTiO3 composite anode was 610 mAh/G after 1400 cycles at 2000 mA/G. The NiFe2O4/N-C composite anode was derived from bimetallic NiFe-MOF calcined in N2[221]. The three-dimensional mesoporous nanostructure shortens the electron/ion migration path and improves the diffusion coefficient of the Li+, and the NiFe2O4 nanoparticles are uniformly distributed and fully utilized during repeated charge-discharge cycles. At the same time, N doping improves the conductivity of the carbon matrix and accelerates the internal electron and ion transport. After 50 cycles at 100 mA/G, the reversible capacity of the NiFe2O4/N-C composite anode is 760 mAh/G, and the rate capability is also better than that of NiFe2O4/CNTs and NiFe2O4/MWCNTs anodes[222][223]. Due to the complementary effect of the one-dimensional hierarchical porous structure and multi-components, the lithium storage performance of the core-shell CNTs@ZnCo2O4 nanowire anode is significantly improved[167]. On the one hand, the abundant redox sites in metal oxides promote the reversibility of electrochemical reactions and inhibit active material agglomeration. On the other hand, CNTs, as an effective conductive and buffering matrix, improve the conductivity and strain coordination ability. Importantly, the MOFs shell can be coated with surfactants of arbitrary chemical composition, structure, dimensionality, and size to modify the core material, not to mention the dimensionality and size of MOFs. The CoFe2O4/GNS composite anode is synthesized by a one-step solvothermal method, and the capacity and the rate capability are obviously improved[224]. This is not only inseparable from the beneficial effect of higher theoretical capacity CoFe2O4, but also fully combines the advantages of graphene nanosheets in conductivity, specific surface area and structural flexibility. Yang et al. First synthesized CoFe2O4/GNS nanocomposite using bimetallic MOFs and graphene oxide via one-step chemical precipitation and subsequent pyrolysis[225]. Compared with pure CoFe2O4 nanoparticles, the reversible capacity, capacity retention, and rate capability of the CoFe2O4/GNS nanocomposite anode are significantly improved, which is not only related to the synergistic effect of Co and Fe oxides, but also the contribution of graphene nanosheets to coordinate the volume change and enlarge the electrolyte/electrode contact area cannot be neglected. The ZnxMnO@C composite retains the hollow hexagonal nanodisc structure of the MOF precursor, and the ultrafine ZnxMnO nanoparticles in each subunit are uniformly embedded into the continuous carbon matrix[226]. Benefiting from the protective effect of the carbon matrix, the hierarchical hollow structure composed of two-dimensional subunits stacked in parallel, and the fast reaction kinetics derived from the pseudocapacitive characteristics, the ZnxMnO@C anode has a specific capacity of 1050 mAh/G after 200 cycles at 100 mA/G; The rate capability is 713 and 330 mAh/G at 1 and 10 A/G, respectively. Especially, the stable cycle reaches 120 times when it is combined with the positive electrode of LiMn2O4 to form a full battery.
It is difficult to directly derive mixed metal oxide/carbon composites from mixed MOFs, and their development and application are obviously limited. With the uniform porous hierarchical nanostructure, the in situ formation of carbon layer and the strong electron interaction between Fe3O4 and MnO, the MnO-doped Fe3O4@C nanosphere anode based on Mn-doped MIL-53 (Fe) precursor has a capacity of 1297.5 mAh/G after 200 cycles at 200 mA/G[227]. The large amount of Co2+ in the ZIF-67 structure has a catalytic effect on the formation of conductive sp2- graphitic carbon when heated in N2. The porous carbon layer of the core-shell structured MoO2@C composite consists of a large amount of N-doped carbon and CoO nanoparticles[228]. Pyridine N and low C content (about 3.24 wt%) increase the conductivity and active material ratio, respectively, and MoO2 and CoO nanoparticles accelerate the Li+ diffusion kinetics and provide additional capacity, respectively, which are beneficial to the improvement of the electrochemical performance of the MoO2@C composite anode. The multilayer ball-in-ball hollow nanostructured Co3O4/NiO/C composite was prepared by sequential carbonization and oxidation of CoNi-MOF, in which the shell wall of each ball is composed of uniform nanorods[229]. After assembling the battery and cycling for 200 times, not only the morphology of the Co3O4/NiO/C anode did not change, but also the Co3O4 and NiO did not agglomerate, and the electrode structure remained intact. The Fe-Mn-O/C microspheres were derived from the direct calcination of Fe/Mn-MOF-74 in inert atmosphere, and the uniformly dispersed Fe2O3 and Mn3O4 nanoparticles were coated with thin carbon layers[230]. The synergetic effect of the hollow spherical structure, hierarchical pores and two-component metal oxides (Fe2O3 and Mn3O4) makes the Fe-Mn-O/C composite anode have a capacity of 1294 mAh/G after 200 cycles at 100 mA/G; The capacity is stable at 315 mAh/G at 5 A/G. The unique composition, atomic-level structure, and suitable chemical stability facilitate the immobilization of pyrolytically derived transition metal oxides of MOFs in porous carbon frameworks with controllable topology. Wu et al. Modified Co3O4@CuO microspheres with carboxyl-functionalized graphene quantum dots (GQDs) to synthesize a Co3O4@CuO@GQDs composite anode[231]. In addition to the core-shell structure, high porosity, and multi-step lithium storage energy affecting their lithium storage performance, the role of GQDs in improving the specific surface area, conductivity, and the number of active sites, as well as enhancing the affinity of Li+ is also important. Using FeIII-MOF-5 as precursor and self-sacrificial template, Zou et al. Successfully constructed uniform nanostructured ZnO/ZnFe2O4/C octahedra[232]. The rate capability, reversible capacity, and cycling performance of the ZnO/ZnFe2O4/C composite anode are quite outstanding due to the effects of the hollow internal structure, ultrafine constituent units, conductive elastic buffer framework, and high porosity. This low-cost, convenient and controllable synthesis process is expected to be mass-produced, and can also be used to prepare other porous carbon-stabilized electrode materials with high power and energy density. The core-shell ZnO/Ni3ZnC0.7/C composite anode exhibited a stable reversible capacity of 1002 mAh/G after 750 cycles at 500 mA/G[233]. The coulombic efficiency, cyclicity, and rate capability of the ZnO/ZnFe2O4@C composite anode are quite outstanding under the combined effect of factors such as the core-shell hollow ZnO/ZnFe2O4 mesoporous nanospheres, the uniformly coated ultrathin N-doped carbon layer, the nanosized heterogeneously connected mixed ZnO-ZnFe2O4, and the interconnected carbon matrix throughout the internal ZnO/ZnFe2O4[234]. Niu et al. Fabricated N-doped C@ZnO/ZnCo2O4/CuCo2O4 nanocomposites using Co-Cu-ZIF-8 as a sacrificial template[235]. Multicomponent metal oxides provide them with higher conductivity and richer redox kinetics due to their lower activation energy for electron transfer. A large amount of N doping makes the porous hollow structure electron-poor and easier to combine with the Li+. The reversible capacity of N-doped C@ZnO/ZnCo2O4/CuCo2O4 nanoanode is 1742 mAh/G after 500 cycles at 300 mA/G; The capacity is still 1009 and 667 mAh/G at high current densities of 3 and 10 A/G, respectively. Obviously, not all active materials react with Li+ at the same time during the repeated charge-discharge process, and other component active materials that do not participate in the reaction shoulder the role of relieving stress, coordinating volume change and inhibiting electrode pulverization. Therefore, the good lithium storage performance of MOFs-based derived hybrid multi-component composite electrodes mainly depends on the mutual synergism and inherent structural advantages of each component. The precursors/templates and main electrochemical performances of some metal oxide/carbon-based composite anode materials are systematically summarized in Table 6.
表6 金属氧化物/碳基复合负极材料及其电化学性能

Table 6 Metal oxide/carbon-based composites as anode materials and their electrochemical performances

Materials Template/precursor Voltage range (V) Current density (mA/g) Cycle
number
Overall capacity (mAh/g) Initial discharge
capacity/charge
capacity
(mAh/g)
Initial coulomb efficiency(%) Specific surface area (m2/g) ref
Co3O4/C ZIF-67 0.01~3.0 200 120 1100 1209/864 71.0 179.4 171
ZnO@C PPy@ZIF-8 0.01~3.0 250/1000/2000 500/500/1000 526/397/275 1106.2/665.8 60.2 174
Co3O4/C PPy@ZIF-67 0.01~3.0 250/1000/2000 500/500/1000 721/372/272 1112/645 58 174
CuO/C Cu-MOF 0.01~3.0 100 200 789 1259/- 76 131.7 175
CuO/C [Cu3(btc)2]n 0.01~3.0 100 200 510.5 1150.9/450.4 46.2 16 176
Mn3O4/C Mn-PBA 0.01~3.0 200 500 1032 1500/1205 80.3 137 177
Mn3O4/C MOF 0.01~3.0 200/500/700 100/120/120 770/651/592 1186/722 60.8 8.0 178
MnO/C Mn-MOF 0.01~3.0 50/500 150/500 884/648 1321.6/779.2 59 313 179
MnO/C Mn(PTA)-MOFs 0~3.0 600/1000 100/200 804/800 -/- 309 180
ZnO/C Zn-BTC 0.01~3.0 500 120 741 1205/715 59 198 181
Fe3O4/C Fe-MOFs 0.01~3.0 100 100 861 1044.2/- 82.4 27 186
Fe3O4@C Fe-MOF 0.01~3.0 100 80 776.8 1714/1333 78 4.57 187
SnO2@C HKUST-1 0.001~3.0 100 200 880 2134/1208 474 189
In2O3/C MIL-68(In) 0.01~3.0 100 150 720 1410/- 43 152 190
MWCNTs/Co3O4 MWCNTs/ZIF-67 0.01~3.0 100 100 813 1171/812 62.9 192
MWCNTs/ZnO ZIF-8/MWCNTs 0.01~3.0 200 100 419.8 1477/854 94.13 193
NiO/CNTs-10 Ni-MOF/CNTs 0.005~3.0 100/2000 100/300 812/502 1100/- 134.68 194
CNT/Co3O4 ZIF-67 0~3.0 1000/4000 200/200 782/577 1840/1281 93.9 195
CFs@Co3O4 CFs@ZIF-67 0.01~3.0 100 150 420 630/369.9 63 532.4 196
3DGN/CuO Cu-BTC 0.01~3.0 100 50 405 569/422 74 197
NiO/GF Ni-MOF/GF 0.01~3.0 100 50 640 903/612 67.8 119 198
Fe2O3/rGO MIL-88-Fe/GO 0.01~3.0 500/5000 200/500 846.9/610.3 1478/971 199
RGO@Co3O4 GO@ZIF 0.01~3.0 100 100 974 1451/- 70 198.54 200
RGO/NiO GO/Ni-MOFs 0~3.0 100 200 440 681/678 99.49 201
MnO/C-N-500 Mn-PBI 0.01~3.0 300 100 1085 1507/1143 75.8 146.4 203
NiO@N-C Ni-NTA 0~3.0 50/4000 300/1200 921/450 1220/1009 82.3 204
Fe2O3@N-C Fe-ZIF 0.01~3.0 100/1000 50/100 1573/1142 1696/1368 80.7 27.1 206
NCW@Fe3O4/NC NCW@Fe-ZIFs 0.01~3.0 100/1000 170/600 1963/1741 2867/1585 55.3 52.7 208
Co3O4/N-C Co-TATB MOFs 0.0~3.0 1000 200 620 1062/- 75.0 209
Co3O4/N-C N-rich Co-MOF 0.01~3.0 1000 500 612 1210/613 51 21.5 210
Co3O4/N-PC ZIF-67 0.05~3.0 100 100 892 1730/1321 76.4 97 211
Co3O4@N-C Co-MOF 0.1~3.0 500 300 795 1385/1055 76 30.16 212
MS-Co3O4@PC Co-MOFs 0.01~3.0 100/1000 60/500 1701/601 1470/1188 80.8 22.1 213
CoO-NCNTs 2D-MOFs 0.01~3.0 500 2000 583 1156/945 81.7 86.26 214
Co3O4@NGN ZIF-67@NGA 0.01~3.0 200/1000 100/400 955/676 976/865 52.3 50 215
ZnO@C(30) ZnO@ZIF-8 0.02~3.0 100/1000 100/300 539/498 1065/664 62 216
Co-doped ZnO@C Co-MOF-5s 0.01~3.0 100 50 725 903/- 217
Ti-doped-CoO@C Co-Ti-MOF 0.01~3.0 200 150 1180 1749/830.7 241.4 218
C/CoTiO3 Co-MOF 0.005~3.0 100/2000 100/1400 630/610 -/- 75.8 62.2 220
NiFe2O4/N-C NiFe-MOF 0.005~3.0 100/500 50/50 760/610 1193/773 64.8 146.2 221
NiFe2O4/CNTs Fe2Ni MIL-88 0.01~3.0 100/2000 100/100 624.6/250 1348/1030 76.4 70.73 222
CNTs@ZnCo2O4 ZIF-8 0.01~3.0 100 100 750 682.7/435.6 78 167
CoFe2O4/GNS Co-Fe-BTC 0.01~3.0 100 100 1061.7 1413/1058 24.5 225
Zn0.5MnO@C Zn-Mn-BTC 100/5000 200 /500 1050/408 1565.9/954.6 30.8 226
MnO-doped Fe3O4@C Mn-doped MIL-53
(Fe)
0.01~3.0 200 200 1297.5 1281.4/938.6 73.3 63.3 227
Co3O4/NiO/C CoNi-MOFs 0.01~3.0 1000 1000 776 1522/907 136 229
Fe-Mn-O/C Fe/Mn-MOF-74 0.0~3.0 100 200 1294 1333/837 158.2 230
Co3O4@CuO
@GQDs
Co-Cu-BTC 0.005~3.0 100 200 1054 1352/816 60 36 231
ZnO/ZnFe2O4/C MOF-5 0.005~3.0 500/2000 100/100 1390/988 1385/1047 75.6 140 232
ZnO/Ni3ZnC0.7/C Zn-MOF/Ni 0.01~3.0 500 750 1002 1743/1015 58.2 112 233
ZnO/ZnFe2O4@C ZFC 0.01~3.0 1000 500 718 1392/1059 76.1 80 234
C@ZnO/ZnCo2O4/CuCo2O4 Co-Cu-ZIF-8 0.01~3.0 300/3000/10000 500/500/500 1742/1009/664 2430/1967 80.1 549.7 235

5.5 Metal sulfide/carbon-based material

Metal sulfides generally have lower cost and higher theoretical capacity, higher conductivity than the corresponding metal oxides, and relatively small volume expansion during Li+ intercalation/deintercalation. Therefore, metal sulfides have obvious advantages in structural stability and reaction kinetics[236]. For example, the Co9S8 anode with a theoretical specific capacity of 544 mAh/G mainly realizes Li+ intercalation/deintercalation through chemical reaction Co9S8+16Li++16e-↔9Co+8Li2S[237]. However, metal sulfides are also difficult to avoid electrode pulverization and structural instability caused by volume change, and even more prone to dissolution of polysulfides in electrolyte, resulting in capacity fading, rate capability and cycle stability reduction. Zhang et al. Synthesized H-Co9S8@C hexagonal prism using Co-MOF-74 as precursor[238]. The highly conductive porous carbon adsorbs and captures intermediate reaction products (such as polysulfides) to reduce the interfacial resistance between metal sulfide particles, while the ultrafine and uniform hollow structure Co9S8 nanoparticles provide the maximum space to buffer stress changes and avoid agglomeration. After 250 cycles at 100 mA/G, the capacity of the H-Co9S8@C composite anode still showed an upward trend and maintained at 900. 5 mAh/G, especially the rate capability and cycle stability were also significantly better than those of S-Co9S8@C. The Co9S8@NMCN composite nanomaterials retained the size uniformity and morphological structure of the original ZIF-67, and the Co9S8 nanoparticles were completely coated by the N-doped carbon layer[239]. Zeng et al. Synthesized Co9S8/N-C nanocomposite by in situ pyrolysis and sulfidation process to make Co9S8 uniformly dispersed in N-rich carbon hollow spherical shell[240]. The discharge capacity of Co9S8/N-C anode is 784 mAh/G at 1 C, which is higher than the theoretical capacity of Co9S8. The discharge capacity at 4 C is still stable at 518 mAh/G. These results suggest that the N atom changes the charge distribution on the surface of the carbon matrix, and the N-doped carbon matrix provides more active sites for lithium storage reaction, which promotes the Li+ and electron migration in the electrode by improving the conductivity and electrochemical activity. In addition, the reversible formation/dissolution of the SEI layer and the polymer-like/gel layer formed in the first cycle also provides additional capacity[241]. The strong reducing organic ligand in ZIF-67 can react with divalent transition metal ions to form the corresponding metal. Sulfur-containing organic compounds and sulfur powder or H2S gas are generally selected as sulfur sources in traditional sulfurization reactions. Although transition metal sulfides (such as CdS and ZnS) have high sulfur content and are rarely used as sulfur sources, their decomposition products are easy to control and recycle, which provides the possibility of green synthesis of some multifunctional sulfides with special structures and components. Wang et al. Synthesized Co9S8/S-NC nanomaterials by solid-phase ion-exchange and diffusion methods, in which CdS replaced common sulfur sources such as sulfur powder, and ZIF-67 was used as a sacrificial template, reducing agent, and C, N, Co source[242]. The results show that the capacity of the Co9S8/S-NC anode synthesized at the annealing temperature of 800 ℃ is 500 mAh/G after 300 cycles at 1000 mA/G, and the cycle stability is good. Obviously, the preparation of MOFs-derived sulfide/carbon-based composites by sulfurization reaction with additional sulfur source is mostly complex. Chen et al. Directly used the Co-MOF containing sulfonic acid groups as the precursor to prepare the N/S Co-doped 3D composite (Co9S8/NSC) for the first time by virtue of the in situ sulfidation reaction of sulfonic acid functional groups and Co2+ to generate Co sulfides during pyrolysis, that is, Co9S8 nanoparticles were uniformly distributed on the N/S Co-doped porous carbon matrix[243]. The Co9S8/NSC anode material exhibited a reversible capacity of 1179 mAh/G at 100 mA/G and a specific capacity of 789 mAh/G after 1000 cycles at 2000 mA/G. This is because N doping enhances the chemisorption of S atoms by the carbon matrix and improves the cycling stability[244]. Wu et al. Obtained three-dimensional CoS @ PCP/CNTs polyhedral composites by simultaneous pyrolysis and vulcanization of ZIF-67 template using a one-step method[245]. During the reaction, Co2+ from the pyrolysis of ZIF-67 reacts with sulfur powder to form Co sulfide, which subsequently acts as a catalyst to promote the in situ growth of carbon nanotubes on the surface of carbon polyhedra. The strong coupling of porous hollow interior structure, ultra-small building unit, flexible conductive carbon matrix, and sulfide/carbon interface enables the CoS @ PCP/CNTs anode to exhibit ultrahigh reversible capacity (1668 mAh/G) and rate capability. Wang et al. Prepared N-doped porous carbon/cobalt sulfide (NC/CoS2) composite by a simple low-temperature sulfurization process using nanocrystalline ZIF-67 as precursor, and studied the effects of CoS2 and carbon matrix particle size on battery performance[246]. With the combination of ultrafine CoS2 particles and N-rich porous carbon layer, the specific capacity of NC/CoS2 anode is 560 mAh/G after 50 cycles at 100 mA/G; The reversible capacity is maintained at 410 mAh/G at a high current density of 2500 mA/G. The binderless CoS2-N-C/3DGN composite anode is prepared based on the organic combination of solution immersion reaction and carbonization-sulfurization treatment, and the excellent electrochemical performance mainly comes from the synergistic effect of the CoS2-N-C with high electrochemical activity, the low-density three-dimensional graphene framework with high conductivity and the macroscopic porous structure[247]. In addition to providing space for CoS2 volume expansion and effectively preventing pulverization, host structures such as graphene and porous carbon can also absorb and trap polysulfide intermediates. The reasonably designed hollow structure and the local pseudocapacitance characteristic, which is beneficial to the rapid insertion/extraction of Li+, provide a strong guarantee for the excellent rate performance of the CoS2/NSCNHF anode, with a coulombic efficiency of about 95.0% (except for the first and second cycles) after 100 cycles of charge and discharge, still maintaining a fairly high capacity (845.0 mAh/G)[248].
MOFs are difficult to combine with carbon-based materials during pyrolysis, so hollow metal sulfide/carbon nanocomposites are rarely synthesized by hydrothermal method. Tian et al. Prepared nanosized Co3S4/MNCNT composite from solvothermally synthesized ZIF-67/MWCNT precursor by sulfurization and subsequent heat treatment, in which multi-walled carbon nanotubes penetrated the hollow Co3S4 quasi-polyhedron and formed a conductive network[249]. After 500 cycles at 2000 mA/G, the specific capacity of the Co3S4/MNCNT composite anode is 976.5 mAh/G, and the coulombic efficiency is close to 100%, which is better than that of most known cobalt sulfide nanocomposite anodes. Yang et al. First used micro/nanostructured Co-BTC instead of nanocrystalline ZIF-67 polyhedra as the precursor to synthesize the rostonium-like Co1-xS/C composite through low temperature carbonization and subsequent sulfidation reaction, and the pseudocapacitive characteristics and redox reaction diffusion kinetics provided more contributions to improve the electrochemical performance[250]. In addition, this simple self-templating method is also suitable for the preparation of other MOFs-derived three-dimensional hierarchical micro/nano structure electrode materials. Heteroatom O can induce a large number of defects, provide additional reaction sites and change the growth kinetics of carbon layer, while S can sometimes promote the dissolution of active material Fe and prevent rapid passivation of the electrode. Therefore, with the synergistic effect of O-doped carbon and special nanosheet morphology, the capacity of FeS/C composite anode is 830 mAh/G after 150 cycles at 100 mA/G. The capacity is 460 mAh/G at 5 A/G[251]. Combined with chemical etching of self-sacrificial template Fe-MOF (MIL-53) and sulfidation reaction, Yin et al. Fabricated porous FeS2@POC concave octahedron[252]. After 100 cycles at 100 mA/G, the reversible specific capacity of the FeS2@POC nanocomposite anode is 1074 mAh/G, which is about three times that of commercial graphite; The capacity is still 607 mAh/G after 200 cycles at 2000 mA/G. It should be noted that the intact carbon matrix not only provides a stable surface for the formation of the SEI layer, but also inhibits its continuous cracking and re-formation. The core-shell C@Fe7S8 nanorods prepared by solid-state chemical sulfidation process contain about 13% carbon and 87%Fe7S8, and the specific surface area is up to 277 m2/g[253]. After 170 cycles at 500 mA/G, the lithium storage capacity of this conversion-type composite anode is 1148 mAh/G, especially the rate capability at 2000 mA/G is more outstanding (657 mAh/G), which is significantly higher than that of Fe3O4/C anode. Obviously, S is very important for the preparation of C@Fe7S8 core-shell nanorods based hierarchical porous composites, and the additional capacity mainly comes from the continuous formation of electroactive polymer colloid film on the surface of the hierarchical porous structure and the interfacial lithium storage provided by the interfacial pseudo-capacitive lithium storage mechanism. In addition to the high specific surface area, porous structure, and conductive carbon matrix, the intimate combination of metal sulfide nanoparticles and carbon framework contributes to the enhanced conductivity, reaction kinetics, and structural stability of ZnS/PC composite anode[254]. On the basis of N-doping and carbon coating to improve the conductivity and stabilize the SEI layer, the new ZnS @ NC composite anode has a capacity of 853 mAh/G after 500 cycles at 500 mA/G by making full use of the strong interaction between N-doped carbon materials and Li+[255]. Although the lithium storage properties of ZnS/carbon composites with two-dimensional or three-dimensional structures have been improved to some extent, the inhomogeneity of ZnS/carbon composites synthesized by solid phase and coprecipitation methods needs to be further improved. For example, the three-dimensional hollow ZnS NR @ HCP composite anode synthesized by simultaneous carbonization and sulfurization of ZIF-8 template not only realizes the controllable growth of crystal morphology, but also ensures the compatibility of different subunits in nano-size[256].
The sheet-like ZnS/C composite anode was synthesized in situ by pyrolysis of MOFs precursor composed of Zn2+ and sulfur-containing organic ligands[257]. The results show that the pyrolysis temperature is closely related to the pore collapse, carbon content, morphology, polymorphic transformation and crystallinity, which ultimately affect the lithium storage performance of ZnS/C composite anode. Although α-MnS anode material has a high theoretical capacity (616 mAh/G) and a low redox potential, its lithium storage performance is not ideal due to the volume effect. The intercalation of nano-sized transition metal sulfides into porous carbon matrices (or layers) such as carbon nanotubes, amorphous carbon, or graphene is expected to solve these problems, but due to the complexity or high cost of the process, this method has rarely been used to prepare α-MnS based composites, and the mechanism of lithium storage has not yet been clarified. Heteroatoms S and N can improve the wettability of the electrode-electrolyte and provide additional lithium storage sites. In the α-MnS/SCMFs microrod composite, ultrafine α-MnS nanoparticles are homogeneously embedded within the S-doped carbon mesoporous framework[258]. The specific capacity of the composite anode is 1383 mAh/G after 300 cycles at 200 mA/G, and the kinetic analysis shows that the pseudocapacitive behavior is beneficial to rapid lithium storage. More importantly, the rate capability, long-term stability, and Coulombic efficiency of the α-MnS/SCMFs/Cu//NCM full cell are also quite outstanding.
In addition to the preparation of metal sulfide/C matrix composites by sulfurization reaction with different sulfur sources, the organic combination of MOFs-derived carbon materials and metal sulfides can also effectively compensate for the insufficient conductivity and cycle stability of metal sulfides. The N-doped carbon material derived from ZIF-8 (N content up to 34wt%) has the characteristics of better conductivity, faster Li+ diffusion rate, low working voltage and high theoretical capacity, and the core-shell structure MoS2@NC anode is a composite material synthesized by in-situ growth of the N-doped carbon layer derived from ZIF-8 on the surface of a MoS2 microsphere[259]. In contrast, the uniform growth of two-dimensional ultrathin MoS2 nanosheets on the active surface of N-doped porous carbon dodecahedra derived from direct pyrolysis of ZIF-8 can also be achieved without any treatment and surfactant assistance,The MNCD core not only acts as a Li+ reservoir and an electron transport medium, but also facilitates the ultrafast charge/discharge of the outer MoS2 shell and improves the structure, thereby improving the cycling stability of the porous MNCD@MoS2 composite anode[260]. During the carbonization process, the simple confined reaction of MOF contributes to the uniform distribution of MoS2 nanocrystals within the porous carbon pores, as shown in Fig. 6[261]. This strong confinement effect can avoid the growth and agglomeration of MoS2, thus forming abundant active centers and edges. Influenced by the fast migration rate of Li+ (~10-9cm2/s) and the obvious improvement of electrical conductivity, the MoS2⊂C hybrid shows ultrahigh rate capability and cycling stability. The core-shell NiCo2S4@D-NC composite anode prepared with thiourea as the sulfur source is composed of double N-doped carbon layers and their coated NiCo2S4 nanoparticles[262]. The rate performance and cycle stability of the NiCo2S4@D-NC composite anode are better than those of the NiCo2S4@NC anode, because the outer carbon layer effectively inhibits the dissolution of intermediate products (polysulfides), enhances the structural integrity of the electrode and improves the conductivity. In order to give full play to the advantages of bimetallic sulfides, carbon materials, and hollow structures, Aslam et al. Fabricated porous core-shell ZnCoS@Co9S8/NC composite anode, with special emphasis on the importance of in situ kinetic control and ion migration control through intermetallic synergism on the structure-related active sites after synthesis[263]. After 500 cycles at 500 mA/G, the specific capacity of the ZnCoS@Co9S8/NC composite anode is 1813 mAh/G; 1095 mAh/G after 400 cycles at 2000 mA/G. The main electrochemical properties of some typical metal sulfide/carbon-based composite anode materials are listed in detail in Table 7.
图6 (a)合成过程示意图, (b) SEM图像, (c) 高分辨TEM图像,(d) 合成的MoS2⊂C的STEM-EDS分布 ((c)的内部为SAED谱)[261]

Fig. 6 (a) Schematic illustration of the fabrication process, (b) SEM image, (c) high-resolution TEM image and (d) STEM-EDS mapping of the as-synthesized MoS2⊂C hybrids (inset of (c) showing the corresponding SAED pattern)[261]

表7 金属硫化物/碳基复合负极材料及其电化学性能

Table 7 Metal sulfide/carbon-based composites as anode materials and their electrochemical performances

Materials Template/precursor Voltage range (V) Current density (mA/g) Cycle
number
Overall Capacity (mAh/g) Initial discharge capacity/charge capacity (mAh/g) Initial
coulomb
efficiency
(%)
Specific surface area (m2/g) ref
H-Co9S8@C Co-MOF-74 0.01~3.0 100/500 250/50 900.5/655 1119.5/867.3 77.4 127.1 238
Co9S8@NMCN ZIF-67 0.01~3.0 100 80 988 1705/1125 66 76.9 239
Co9S8/N-C ZIF-67 0.01~3.0 544 400 784 1260/900 71.48 125.9 240
Co9S8/S-NC ZIF-67 0~3.0 1000 300 500 879.7/529.3 60.17 150.5 242
Co9S8/NSC Sulfonate-based Co-MOF 0.01~3.0 100/2000 200/1000 1179/789 1816.9/862 47.5 228.1 243
CoS@PCP/
CNTs-600
ZIF-67 0.01~3.0 200 100 1668 2083/1246 101.5 245
NC/CoS2-650 ZIF-67 0.1~3.0 100/2500 50/50 560/410 1100/ 246
CoS2-N-C/3DGN Co-MOF 0.01~3.0 100 100 409.5 833.5/666.3 79.9 247
CoS2/NSCNHF ZIF8@ZIF67 0.01~3.0 100/1000 100/200 845/549.9 1155.6/739.6 64 234.25 248
Co3S4/MNCNT ZIF-67/MWCNT 0.01~3.0 200/2000 50/500 1281.2/976.5 1644.2/1055 64.17 112 249
Co1-xS/C Co-BTC 0.01~3.0 200/1000 100/700 791/667 1290/932 72 124 250
FeS/C Fe-MOFs 0.01~3.0 100 150 830 1702/972 57 0.015 251
FeS2@POC Fe-MOFs 0.01~3.0 100/2000 100/200 1074/607 1394/1115 80 44.2 252
C@Fe7S8 MIL-88 0.01~3.0 500 170 1148 1072/761 71 277 253
ZnS/PC MOF-5 0.01-2.5 100 300 438 1220/- 296.8 254
ZnS@NC ZIF-8 0.01~3.0 200/500 150/450 521.8/853 1284.9/840.6 34.6 191.45 255
α-MnS/SCMFs/Cu Mn-MOF 0.01~3.0 200/1500 300/1000 1383/601 1115/- 69 109 258
MoS2@NC-2 ZIF-8 0.005~3.0 100 50 715 1864/- 50 259
ZnCoS@Co9S8/NC ZIF-67@ZIF-8/
ZIF-67
0.01~3.0 500/2000 500/400 1814/1095 2895/2182 270.46 263

5.6 Other metal compounds/carbon-based materials

Compared with oxides and sulfides, metal selenides have higher energy density and faster electrochemical reaction kinetics. Due to the relatively weak bonding strength between metal and selenium, chemical bond breaking and conversion reactions are prone to occur during charge-discharge cycles, and the electrochemical performance is often reduced due to volume change. Only by optimizing the chemical composition and microstructure to improve the specific surface area and porosity can the capacity and cycle performance requirements be met. Metal selenide/carbon-based composites are usually selenized during the synthesis of precursors or after calcination, which is similar to the synthesis of metal oxide (or sulfide)/carbon-based composites. The CoSe @ C composite anode is synthesized in situ by carbonization of ZIF-67 precursor and subsequent selenization reaction, and the CoSe nanoparticles are uniformly distributed in the porous carbon polyhedron[264]. Among them, CoSe has higher reactivity with lithium and faster conversion reaction kinetics, and PCP provides conductive support for CoSe and alleviates stress-induced structural changes. Using 3DG/MOF aerogel as a template, Jiang et al. Designed and synthesized hierarchically structured 3DG/Fe7Se8@C composites, that is, nanoparticles composed of a Fe7Se8 rich inner core and a carbon rich outer shell distributed within a three-dimensional graphene framework[265]. After 120 cycles at 100 mA/G, the reversible capacity of the flexible 3DG/Fe7Se8@C composite anode is 884.1 mAh/G, which is significantly higher than the theoretical capacity of Fe7Se8 (418 mAh/G); It is still 815.2 mAh/G after 2500 cycles at 1 A/G.
With the synergistic lithium storage of the two metal selenides, the stable SEI layer and the pseudocapacitive characteristics, the reversible capacity of the mesoporous bundle-like Ni-Co-Se/C composite anode is 2061 mAh/G after 300 cycles at 500 m A/G, which is much higher than the theoretical capacity[266]. The hierarchical rod-like Bi2Se3@C composite anode prepared by the in-situ selenization method still maintains the initial morphology characteristics after 1000 cycles, and the structure is not obviously damaged.The outstanding specific capacity and rate capability are mainly determined by the stable micro/nano structured porous carbon matrix and pseudocapacitive behavior, and the electrochemical lithium storage reaction mechanism indicates that some Bi2Se3 transform from rhombohedral to orthorhombic structure after repeated charge-discharge cycles[267]. As mentioned above, the introduction of heterogeneous N atoms into the carbon matrix produces a large number of lattice defects and forms a disordered carbon structure, which improves the conductivity and accelerates the electron migration. Liu et al. controlled the morphology and size of the ZnSe/NC composite anode by adjusting the particle size of the ZIF-8 precursor, for example, the ZnSe/NC-300 anode had a reversible discharge capacity of 724.4 mAh/G after 500 cycles at 1000 mA/G[268]. Similarly, Tao et al. First used MOF-Ni as a template to synthesize Ni2P/NC porous spheres via in situ low-temperature phosphating reaction, and the resulting assembled LiNi1/3Co1/3Mn1/3O2‖Ni2P/NC full cell exhibited good rate performance and cycle life[269]. In addition, the structural characteristics of transition metal phosphide/carbon-based composite anodes are diverse, not limited to the common hollow and spherical ones. For example, the specific capacity, cycle stability and rate performance of the hierarchical porous CoP/C nanobox composite anode are good, and the electrochemical reaction mechanism needs to be further clarified[270]. Xia et al. Employed a simple MOF-derived self-template method and precisely regulated the pyrolysis temperature to successfully prepare a composition-controlled CoxP-NC composite, that is, CoxP nanoparticles embedded in an N-doped polyhedral porous carbon matrix generated in situ by a ZIF-67 template[271]. The CoxP-NC-800 composite anode is mainly composed of Co2P and CoP, and has a reversible capacity of 1224 mA/G after 100 cycles at 100 mA/G; The cycle life is up to 1800 cycles at a high current density of 1000 mA/G. Therefore, it is necessary to study the evolution of composition and structure in the process of electrochemical reaction with the help of advanced technologies such as in situ spectroscopy and in situ electron microscopy. During the carbonization process, the carbon layer generated in situ Mo3(BTC)2 of the precursor restricts the growth and serious agglomeration of MoC nanocrystals, and ensures good high-temperature stability. The optimized MoC @ C-700 anode has a reversible capacity of 647.3 mAh/G at 200 mA/G; The capacity is still 509.8 mAh/G after 500 cycles at 2 A/G[272]. This is not only related to the special pomegranate-like structure to reduce electrolyte consumption and form SEI layer, but also the carbonization temperature, MoC nanocrystal size and carbon layer content affect the electron migration and ion diffusion kinetics of MoC @ C composite anode. Yan Haoran et al. Obtained Ni2B@C composite materials by carbonizing and boronizing Ni-MOF in turn, and analyzed the relationship between temperature and its morphology and electrochemical performance[273]. The initial discharge specific capacity of the Ni2B@C-500 anode is 1033.88 mAh/G, and the capacity is 467.17 mAh/G after 500 cycles at 500 mA/G. The good cycling stability is due to the integrity of the uniform octahedral structure and the difficulty of crushing. Table 8 shows the related electrochemical properties of other metal compound/carbon-based composite anode materials reported in some literatures. Compared with metal oxide/C and sulfide/C composites, there are relatively few studies on MOF-derived metal phosphide/C and metal carbide/C anode materials, especially the mechanism of electrochemical lithium storage needs to be studied in detail.
表8 其他金属化合物/碳基复合负极材料及其电化学性能

Table 8 Other metal compounds/carbon-based composites as anode materials and their electrochemical performances

Materials Template/precursor Voltage range (V) Current density (mA/g) Cycle
number
Overall Capacity (mAh/g) Initial discharge
capacity/charge
capacity
(mAh/g)
Initial
coulomb
efficiency
(%)
Specific surface area (m2/g) ref
CoSe@PCP ZIF-67 0.005~3.0 200/1000 100/500 675/708.2 902/603 73.5 76.94 264
3DG/Fe7Se8@C 3DG/MOF 0.01~3.0 200/1000 120/250 884.1/815.2 265
Ni-Co-Se/C-600 Ni-Co-BTC 0.01~3.0 1000/3000 500/900 1514/852 821/634 77 127 266
Bi2Se3@C Bi-MOF 0.01~3.0 200/1000 200/5000 637/543 -/642 68 76 267
ZnSe/NC-300 ZIF-8 0.01~3.0 100 500 724.4 906.66/547.48 60.3 93.926 268
Ni2P/NC MOF-Ni 0.01~3.0 500 800 450.4 1240.5/649.5 34.5 269
CoP/C ZIF-67 0.01~3.0 500 1000 523 1522/1110 72.9 67.2 270
CoxP-NC-800 ZIF-67 0.01~3.0 100/1000 100/1800 1224/400 2450/1469 326.5 271
MoC@C-700 Mo3(BTC)2 0.01~3.0 1000 500 509.8 990.8/674.3 65.3 187 272

5.7 Metal/metal oxide/carbon-based material

In MOFs materials, the central metal ions interconnected by organic ligands are in a dispersed and isolated state. The highly carbonized organic ligand can avoid metal agglomeration during heat treatment, which provides the possibility of preparing nanocomposite anode by using porous carbon-coated metal particles and metal oxides, and is expected to significantly improve its cycle stability and life.
Co3O4/Co/ carbon nanocages were prepared from Co-MOFs (ZIF-67) precursors by sequential carbonization and oxidation treatments, and Co3O4 and Co nanoparticles were dispersively embedded in the N-doped carbon nanocage matrix[274]. The hollow dodecahedron structure with uniformly distributed pores not only solves the problem of particle agglomeration and nanostructure cracking caused by charge-discharge cycles, but also provides a continuous flexible conductive carbon framework for rapid transmission of ions and electrons. With the advantages of reversible specific capacity, coulombic efficiency, rate capability and cycle stability, Co3O4/Co/ carbon nanocages have great potential as anode materials. Zhong et al. Synthesized a series of core-shell structured Co/Co3O4@N-C nanocomposites using Co-MOF as precursor at different carbonization temperatures[275]. Since the carbonization temperature not only affects the N and C content, the degree of carbon graphitization and electrical conductivity, but also promotes the gradual aggregation of Co/Co3O4 nanoparticles. Therefore, the optimized Co/Co3O4@N-C-700 composite anode has the highest initial discharge capacity (1535 mAh/G), reversible capacity (903 mAh/G after 100 cycles at 100 mA/G), and best rate capability (774 mAh/G after 100 cycles at 1000 mA/G). Chen et al. Prepared the core-shell structure Sn/C-ZnO composite by a two-step method, in which the metal Sn originated from the reduction of ZIF-8 coated SnO2 particles, and the flow of liquid Sn in the pores of MOFs during high temperature pyrolysis was used to limit the shrinkage of pore size and surface area, effectively retaining the porous structure characteristics of MOFs, as shown in Fig. 7[276]. The results show that the thicker the C-ZnO shell, the higher and more stable the discharge capacity of Sn/C-ZnO composite anode during charge-discharge cycles. He et al. First prepared SnO2/Co@C nanocubes using a two-step solvent-free thermal solid state reaction[277]. On the one hand, Co distributed in the SnO2 promotes electron conduction and inhibits the migration and subsequent agglomeration of Sn during cycling, which improves the structural stability of the electrode and maintains the activity of Sn. Co, on the other hand, consumes excess Li2O and provides extra capacity, releasing unusable Sn encapsulated by passive Li2O and promoting its reversible transformation. Therefore, the reversible discharge capacity of the SnO2/Co@C anode is 800 mAh/G after 100 cycles at 200 mA/G; The capacity is 400 mAh/G after 1800 cycles at 5 A/G. The NiO/Ni/graphene composite material is prepared by continuous carbonization and oxidation treatment of a Ni-MOF precursor, is composed of conformal graphene and ultrafine NiO/Ni nanoparticles coated with the graphene, and belongs to a porous nano hollow sphere-in-sphere structure[278]. The NiO/Ni/graphene composite anode exhibits outstanding reversible specific capacity (1144 mAh/G), cyclability (almost no capacity fading after 1000 cycles at 2000 mA/G), and rate capability (805 mAh/G at 2000 mA/G) due to the graphene buffer layer improving the conductivity, promoting the formation of a stable SEI layer, and ensuring the structural stability. The capacity retention of the reasonably designed Ni @ ZnO/CNF composite anode is 88% after 100 cycles at 100 mA/G. The specific capacity is about 497 mAh/G after 10 cycles at 1000 mA/G[279]. The results show that Ni reacts with Li2O and Zn to form Li-Zn alloy, and C nanocages and heterogeneous N atoms form a conductive framework and act as a buffer framework for the volume change of ZnO. In addition, the carbon framework retains the shortened rhombohedral morphology and the original pores, which is more conducive to the rapid intercalation/deintercalation of Li+. Especially, this simple and low-cost synthesis method is also suitable for the preparation of other high-performance unsupported anode materials. The relevant electrochemical performances of some metal/metal oxide/carbon-based composite anode materials are summarized in Table 9.
图7 (A) SnZCw前驱体和(B)不同放大倍数的SnZCw的SEM图像;SnZCw的(C) TEM图像、(D) HRTEM图像和(E) 选区电子衍射; (F) 不同放大倍数的SnZCd TEM图像[276]

Fig. 7 SEM images of (A) the precursor of SnZCw, and (B) SnZCw at different magnifications; (C) TEM image, (D) HRTEM image, and (E) the SAED pattern of SnZCw; (F) TEM images at different magnifications of SnZCd[276]

表9 金属/金属氧化物/碳基复合负极材料及其电化学性能

Table 9 Metal/metal oxide/carbon-based composites materials as anode materials and their electrochemical performances

Materials Template/precursor Voltage range (V) Current density (mA/g) Cycle
number
Overall Capacity (mAh/g) Initial discharge
capacity/charge
capacity
(mAh/g)
Initial
coulomb
efficiency
(%)
Specific surface area (m2/g) ref
Co3O4/Co/C ZIF-67 0.01~3.0 100/2000 60/600 801/505 1158/867 75 183.9 274
Co/Co3O4@N-C-700 NUM-6 0.01~3.0 100/1000 100/100 903/774 1535/830 250.4 275
Sn/C-ZnO ZIF-8 0.01~3.0 100 50 515.6 1118.7/- 64.2 32.9 276
SnO2/Co@C CoSnO3 @MOF 0.01-2.5 200/5000 100/1800 800/400 1300/857 66 84 277
NiO/Ni/Graphene Ni-MOFs 0.005~3.0 2000 250 1180 1759/1144 104 278
Ni@ZnO/CNF Ni@ZIF-8 0.01~3.0 100 100 1051 1547/1100 71 38.6 279
The above analysis shows that although MOFs and their derived anode materials have made significant progress in the selection of central metal ions and organic ligands, structural design and optimization, and performance improvement, it is undeniable that these materials still have some gaps in meeting practical applications. The advantages and disadvantages of MOFs and their derived anode materials are summarized in Table 10.
表10 MOFs及其衍生负极材料的优点和缺点

Table 10 Advantages and disadvantages of MOFs and their derivative anode materials

Materials Advantages Disadvantages
MOFs abundant active sites
large specific surface area
high porosity
adjustable composition, morphology and structure
low cost
low conductivity
poor structure stability
fast capacity decay
Porous carbon large specific surface area
high porosity
simple synthesis method and mild synthesis condition
high thermal stability
no subsequent complex physical or chemical activation
insufficient capacity
uncontrolled structure evolution
inherent structure properties determined from micropore
Single metal oxide simple synthesis method
controllable synthesis path
Large specific surface area
high porosity
controlled structure and composition
limited species
lower conductivity
easy volume expansion and pulverization
lower rate capability and cycle performance
Double metal oxides stronger synergistic effect between different metal elements
abundant redox sites
faster reaction kinetics and activity
lower conductivity
easy volume expansion and pulverization
lower rate capability and cycle performance
Other metal compounds wide variety
good mechanical and thermodynamic stability
higher theoretical capacity
complicated synthesis process
higher cost
lower reaction kinetics under high load
lower utilization
MOFs/ Metal oxide abundant active sites
high porosity
alleviated volume expansion and pulverization
lower conductivity
MOFs/C high porosity
better structure stability
higher conductivity
limited capacity
Metal oxide/Metal oxide stronger synergistic effect between different metal elements
improved electrode integrity
alleviated volume effect
lower conductivity
easy volume expansion and pulverization
limited capacity and cycle stability
Metal oxide/C higher conductivity
better structure stability
faster electron transportation
stronger synergistic effect
complicated synthesis method
uncontrolled synthesis condition
Metal sulfide/C higher conductivity
better structure stability
stronger synergistic effect
slower reaction kinetics
faster capacity decay
Other metal compounds/C abundant active sites
higher utilization rate of active substances
faster electron transportation
stronger synergistic effect
insufficient performance under high load
ambiguous reaction mechanism
Metal/Metal oxide/C higher conductivity
better structure stability
faster electron transportation
stronger synergistic effect
complicated synthesis process
ambiguous reaction mechanism

6 Conclusion and prospect

MOFs anode materials have both the characteristics and pore structure of inorganic and organic materials, which can realize reversible lithium storage during charge-discharge process, but also have the defects of insufficient conductivity, limited lithium storage sites, low reversible capacity and poor structural stability.Various porous nanostructured derivatives (such as transition metal oxides, sulfides, phosphides, selenides, and carbon-based materials) and their composites prepared by using MOFs as templates or precursors are expected to solve these problems to varying degrees. In this paper, the research progress of MOFs and their derivative nano-anode materials in recent years is systematically summarized, and the relationship between preparation methods, pore structure, chemical composition and lithium storage mechanism, electrochemical performance and cycle life is emphatically expounded according to the respective characteristics of MOFs derivatives and their composites. Obviously, on the basis of fully combining process optimization and chemical composition, morphology and structure control, the specific surface area, conductivity, ion and electron migration rate and volume effect of these porous anode materials have been effectively improved.The reversible capacity, rate performance and cycle stability have been significantly improved, but most of them still face the problems of relatively complex synthesis process, harsh conditions, high cost and environmentally unfriendly factors. More importantly, most of these studies are limited to half-cell performance testing, rarely involving the comprehensive evaluation of the whole cell, and even lack of in-depth discussion and analysis of key issues such as battery capacity fading, which to some extent restricts the development and commercial application of lithium ion battery anode materials.
The design and synthesis of MOFs and their derived anode materials should start with the improvement of Li+ and electron transfer, as well as the contact area between the electrolyte and the electrode, focusing on the electrode structure, surface morphology, chemical composition and porosity, so as to improve the overall transport efficiency of the battery. The key problems to be solved in future research work mainly focus on the following four aspects: First, the precise control of chemical composition, structure and specific surface area can be achieved by reasonable selection of metal ion centers, variable organic ligands, optimization of process parameters and recrystallization.That is to say, relatively stable metal ions are selected as the connection points of organic ligands to provide more active sites for lithium storage and maintain structural stability by means of the interaction of uniformly distributed metal ions, open framework structure, chemical modification or functional heterogeneous atom doping and multi-component assembly. The second is to design complex porous nano-space structures (such as hollow or core-shell structures) or adopt conductive matrix composite methods, balance and utilize the advantages of different lithium storage mechanisms, and maximize the synergistic effect of metal centers, organic ligands and pore structures to meet the requirements of lithium storage performance. The third is to comprehensively and thoroughly study the surface or near-surface ion and electron migration, as well as the possible accompanying electrochemical adsorption and pseudocapacitive behavior, so as to further improve and clarify the relationship between complex morphology and lithium storage performance. To sum up, improving the existing synthesis methods or combining the advantages of different methods will become the development trend of the preparation of some special and complex nanostructured anode materials, but there are still some challenges to achieve precise control, and cost-effective and controllable synthesis is the prerequisite for the practical application of MOFs and their derivative nanoanode materials. At the same time, the combination of theoretical calculation and experiment can further understand the relationship between lithium storage mechanism, structure and performance of the electrode, and provide theoretical basis and support for the structural design and performance optimization of the electrode. With the continuous development and improvement of electrode preparation and material characterization techniques, the electrode reaction kinetics of MOFs derivatives and their composite anode materials during charge-discharge cycles will be thoroughly clarified. The fourth is to reduce the dependence on expensive or rare raw materials, simplify the complexity of the synthesis process, establish an efficient green synthesis and evaluation system, implement theoretical mining and experimental prediction into production practice, and effectively promote the industrialization process of high-performance MOFs-derived nano-anode materials. As a potential advanced functional intensive electrode material, the application prospect of MOFs derivatives and their composites will be broader, which is of great significance for promoting the development of lithium-ion batteries.
[1]
Zhong M, Kong L J, Li N, Liu Y Y, Zhu J, Bu X H. Coord. Chem. Rev., 2019, 388: 172.

[2]
Jiang J, Li Y Y, Liu J P, Huang X T, Yuan C Z, David Lou X W. Adv. Mater., 2012, 24(38): 5166.

[3]
Lux L, Williams K, Ma S Q. CrystEngComm, 2015, 17(1): 10.

[4]
Jiang Y, Yue J L, Guo Q B, Xia Q Y, Zhou C, Feng T, Xu J, Xia H. Small, 2018, 14(19): 1704296.

[5]
Li Z J, Du Z, Zhang J, Chen J W, Wang G, Wang R L. Prog. Chem., 2019, 31(1): 201.

(李振杰, 钟杜, 张洁, 陈金伟, 王刚, 王瑞林. 化学进展, 2019, 31: 201.).

[6]
Zhao Y, Kang Y, Jin Y, Wang L, Tian G, He X. Prog. Chem., 2019, 31: 613.

(赵云, 亢玉琼, 金玉红, 王莉, 田光宇, 何向明. 化学进展, 2019, 31: 613.).

[7]
Li X X, Cheng F Y, Zhang S N, Chen J. J. Power Sources, 2006, 160(1): 542.

[8]
C, Lin Y, Zhao Y, Wang JinD, Chen L, Shen C. Mater. Sci. Technol., 2017, 33(8): 768.

[9]
Shin J, Kim M, Cirera J, Chen S, Halder G J, Yersak T A, Paesani F, Cohen S M, Meng Y S. J. Mater. Chem. A, 2015, 3(8): 4738.

[10]
Li C, Hu X S, Lou X B, Zhang L J, Wang Y, Amoureux J P, Shen M, Chen Q, Hu B W. J. Mater. Chem. A, 2016, 4(41): 16245.

[11]
Xiao T, Jin J, Zhang Y F, Xi W, Wang R, Gong Y S, He B B, Wang H W. Electrochim. Acta, 2022, 427: 140851.

[12]
Matei Ghimbeu C, GÓrka J, Simone V, Simonin L, Martinet S, Vix-Guterl C. Nano Energy, 2018, 44: 327.

[13]
Nie P, Shen L F, Luo H F, Ding B, Xu G Y, Wang J, Zhang X G. J. Mater. Chem. A, 2014, 2(16): 5852.

[14]
An T C, Wang Y H, Tang J, Wang Y, Zhang L J, Zheng G F. J. Colloid Interface Sci., 2015, 445: 320.

[15]
Zhao C C, Shen C, Han W Q. RSC Adv., 2015, 5(26): 20386.

[16]
Lin X M, Niu J L, Lin J, Wei L M, Hu L, Zhang G, Cai Y P. Inorg. Chem., 2016, 55(17): 8244.

[17]
Lin X M, Niu J L, ChenD N, Lu Y N, Zhang G, Cai Y P. CrystEngComm, 2016, 18(36): 6841.

[18]
Hu L, Lin X M, Mo J T, Lin J, Gan H L, Yang X L, Cai Y P. Inorg. Chem., 2017, 56(8): 4289.

[19]
Li C, Lou X B, Shen M, Hu X S, Guo Z, Wang Y, Hu B W, Chen Q. ACS Appl. Mater. Interfaces, 2016, 8(24): 15352.

[20]
Chen L, Yang W J, Wang J B, Chen C R, Wei MD. Chem. A Eur. J., 2018, 24(50): 13362.

[21]
Xing J J, Shi F N, XueD F. Chem. Res., 2020, 31(2): 95.

(邢锦娟, 史发年, 薛冬峰. 化学研究, 2020, 31(2): 95.).

[22]
Ning Y Q, Lou X B, Li C, Hu X S, Hu B W. Chem. A Eur. J., 2017, 23(63): 15984.

[23]
Li G H, Yang H, Li F C, Cheng F Y, Shi W, Chen J, Cheng P. Inorg. Chem., 2016, 55(10): 4935.

[24]
Liao Y X, Li C, Lou X B, Wang P, Yang Q, Shen M, Hu B W. J. Colloid Interface Sci., 2017, 506: 365.

[25]
Wang P, Lou X B, Li C, Hu X S, Yang Q, Hu B W. Nano Micro Lett., 2018, 10(2): 19.

[26]
GeD H, Peng J, Qu G L, Geng H B, Deng Y Y, Wu J J, Cao X Q, Zheng J W, Gu H W. New J. Chem., 2016, 40(11): 9238.

[27]
Wang L P, Zhao M J, Qiu J L, Gao P, Xue J, Li J Z. Energy Technol., 2017, 5(4): 637.

[28]
Gou L, Hao L M, Shi Y X, Ma S L, Fan X Y, Xu L, LiD L, Wang K. J. Solid State Chem., 2014, 210(1): 121.

[29]
Fei H L, Liu X, Li Z W. Chem. Eng. J., 2015, 281: 453.

[30]
Saravanan K, Nagarathinam M, Balaya P, Vittal J J. J. Mater. Chem., 2010, 20(38): 8329.

[31]
Lin Y C, Zhang Q J, Zhao C C, Li H L, Kong C L, Shen C, Chen L. Chem. Commun., 2015, 51(4): 697.

[32]
Liu Q, Yu L L, Wang Y, Ji Y Z, Horvat J, Cheng M L, Jia X Y, Wang G X. Inorg. Chem., 2013, 52(6): 2817.

[33]
Fei H L, Liu X, Li Z W, Feng W J. Electrochim. Acta, 2015, 174: 1088.

[34]
Maiti S, Pramanik A, Manju U, Mahanty S. ACS Appl. Mater. Interfaces, 2015, 7(30): 16357.

[35]
Xiong P X, Zeng G J, Zeng L X, Wei MD. Dalton Trans., 2015, 44(38): 16746.

[36]
Hu H P, Lou X B, Li C, Hu X S, Li T, Chen Q, Shen M, Hu B W. New J. Chem., 2016, 40(11): 9746.

[37]
Reinsch H, Stock N. CrystEngComm, 2013, 15(3): 544.

[38]
Li C, Hu X S, Tong W, Yan W S, Lou X B, Shen M, Hu B W. ACS Appl. Mater. Interfaces, 2017, 9(35): 29829.

[39]
FÉrey G, Millange F, Morcrette M, Serre C, Doublet M L, Grenèche J M, Tarascon J M. Angew. Chem. Int. Ed., 2007, 46(18): 3259.

[40]
de Combarieu G, Morcrette M, Millange F, Guillou N, Cabana J, Grey C P, Margiolaki I, FÉrey G, Tarascon J M. Chem. Mater., 2009, 21(8): 1602.

[41]
Fateeva A, Horcajada P, Devic T, Serre C, Marrot J, Grenèche J M, Morcrette M, Tarascon J M, Maurin G, FÉrey G. Eur. J. Inorg. Chem., 2010, 2010(24): 3789.

[42]
Hu X S, Lou X B, Li C, Ning Y Q, Liao Y X, Chen Q, Mananga E S, Shen M, Hu B W. RSC Adv., 2016, 6(115): 114483.

[43]
Du J. MasteralDissertation of Xi’an University of Science and Technology, 2017.

(杜婕. 西安科技大学硕士论文, 2017.).

[44]
Shen L S, Song H W, Wang C X. Electrochim. Acta, 2017, 235: 595.

[45]
Lou X B, Hu H P, Li C, Hu X S, Li T, Shen M, Chen Q, Hu B W. RSC Adv., 2016, 6(89): 86126.

[46]
Zhang Y, Niu Y B, Liu T, Li Y T, Wang M Q, Hou J K, Xu M W. Mater. Lett., 2015, 161: 712.

[47]
Park K S, Ni Z, CôtÉ A P, Choi J Y, Huang RD, Uribe-Romo F J, Chae H K, O’Keeffe M, Yaghi O M. Proc. Natl. Acad. Sci. U. S. A., 2006, 103(27): 10186.

[48]
Senthil Kumar R, Nithya C, Gopukumar S, Anbu Kulandainathan M. Energy Technol., 2014, 2(11): 921.

[49]
Maiti S, Pramanik A, Manju U, Mahanty S. Microporous Mesoporous Mater., 2016, 226: 353.

[50]
Hu X S. DoctoralDissertation of East China Normal University, 2018.

(胡小诗. 华东师范大学博士论文, 2018.).

[51]
Wu N, Wang W, Kou L Q, Zhang X, Shi Y R, Li T H, Li F, Zhou J M, Wei Y. Chem. A Eur. J., 2018, 24(24): 6330.

[52]
Wu N, Jia T, Shi Y R, Yang Y J, Li T H, Li F, Wang Z. Ionics, 2020, 26(3): 1547.

[53]
Wang Y, Qu Q T, Liu G, Battaglia V S, Zheng H H. Nano Energy, 2017, 39: 200.

[54]
Xia S B, Yu S W, Yao L F, Li F S, Li X, Cheng F X, Shen X, Sun C K, Guo H, Liu J J. Electrochim. Acta, 2019, 296: 746.

[55]
Han X Y, Yi F, Sun T L, Sun J T. Electrochem. Commun., 2012, 25: 136.

[56]
Guo L Z, Sun J F, Zhang W H, Hou L R, Liang L W, Liu Y, Yuan C Z. ChemSusChem, 2019, 12(22): 5051.

[57]
Guo L Z, Sun J F, Sun X, Zhang J Y, Hou L R, Yuan C Z. Nanoscale Adv., 2019, 1(12): 4688.

[58]
Yan J, Cui Y T, Xie M, Yang G Z, BinD S, LiD. Angew. Chem. Int. Ed., 2021, 60(46): 24467.

[59]
Mao P C, Fan H L, Liu C, Lan G X, Huang W, Li Z P, Mahmoud H, Zheng R G, Wang Z Y, Sun H Y, Liu Y G. Sustain. Energy Fuels, 2022, 6(17): 4075.

[60]
Sun M H, Huang S Z, Chen L H, Li Y, Yang X Y, Yuan Z Y, Su B L. Chem. Soc. Rev., 2016, 45(12): 3479.

[61]
Yao Y, Hou H Y, Liu X X, Tian C, Meng K, Lan J, Xu J L, Feng M M. J. Synth. Cryst., 2020, 49(7): 1242.

(姚远, 侯宏英, 刘显茜, 田川, 孟堃, 兰建, 徐加雷, 冯蒙蒙. 人工晶体学报, 2020, 49(7): 1242.).

[62]
Wu R B, Qian X K, Yu F, Liu H, Zhou K, Wei J, Huang Y Z. J. Mater. Chem. A, 2013, 1(37): 11126.

[63]
Xiao X L, Liu X F, Zhao H, ChenD F, Liu F Z, Xiang J H, Hu Z B, Li YD. Adv. Mater., 2012, 24(42): 5762.

[64]
Banerjee A, Singh U, Aravindan V, Srinivasan M, Ogale S. Nano Energy, 2013, 2(6): 1158.

[65]
Hu X S, Li C, Lou X B, Yang Q, Hu B W. J. Mater. Chem. A, 2017, 5(25): 12828.

[66]
Zhang B W, Hao S J, XiaoD R, Wu J S, Huang Y Z. Mater.Des., 2016, 98: 319.

[67]
Bai Z C, Zhang Y H, Zhang Y W, Guo C L, Tang B, SunD. J. Mater. Chem. A, 2015, 3(10): 5266.

[68]
Cao K Z, Jiao L F, Xu H, Liu H Q, Kang H Y, Zhao Y, Liu Y C, Wang Y J, Yuan H T. Adv. Sci., 2016, 3(3): 1500185.

[69]
Zheng F C, Xu S H, Yin Z C, Zhang Y G, Lu L. RSC Adv., 2016, 6(96): 93532.

[70]
Hu X S, Lou X B, Li C, Yang Q, Chen Q, Hu B W. ACS Appl. Mater. Interfaces, 2018, 10(17): 14684.

[71]
Liu B, Zhang X B, Shioyama H, Mukai T, Sakai T, Xu Q. J. Power Sources, 2010, 195(3): 857.

[72]
Li X, Tian XD, Yang T, Song Y, Liu Y M, Guo Q G, Liu Z J. J. Alloys Compd., 2018, 735: 2446.

[73]
Li A, Zhong M, Shuang W, Wang C P, Liu J, Chang Z, Bu X H. Inorg. Chem. Front., 2018, 5(7): 1602.

[74]
Zheng F C, Yin Z C, Xia H Y, Zhang Y G. Mater. Lett., 2017, 197: 188.

[75]
Han Y, Zhao M L, Dong L, Feng J M, Wang Y J, LiD J, Li X F. J. Mater. Chem. A, 2015, 3(45): 22542.

[76]
Hu L, Yan N, Chen Q W, Zhang P, Zhong H, Zheng X R, Li Y, Hu X Y. Chem. A Eur. J., 2012, 18(29): 8971.

[77]
Li C, Chen T Q, Xu W J, Lou X B, Pan L K, Chen Q, Hu B W. J. Mater. Chem. A, 2015, 3(10): 5585.

[78]
Gou L, Ma L, Zhao M J, Liu P G, Wang XD, Fan X Y, LiD L. J. Mater. Sci., 2019, 54(2): 1529.

[79]
Zhang L M, Yan B, Zhang J H, Liu Y J, Yuan A H, Yang G. Ceram. Int., 2016, 42(4): 5160.

[80]
Su P P, Liao S C, Rong F, Wang F Q, Chen J, Li C, Yang Q H. J. Mater. Chem. A, 2014, 2(41): 17408.

[81]
Shao J, Wan Z M, Liu H M, Zheng H Y, Gao T, Shen M, Qu Q T, Zheng H H. J. Mater. Chem. A, 2014, 2(31): 12194.

[82]
Li J F, Wang J Z, Liang X, Zhang Z J, Liu H K, Qian Y T, Xiong S L. ACS Appl. Mater. Interfaces, 2014, 6(1): 24.

[83]
Wang J Y, Yang N L, Tang H J, Dong Z H, Jin Q, Yang M, Kisailus D, Zhao H J, Tang Z Y, Wang D. Angew. Chem., 2013, 125(25): 6545.

[84]
Xu XD, Cao R G, Jeong S, Cho J. Nano Lett., 2012, 12(9): 4988.

[85]
Banerjee A, Aravindan V, Bhatnagar S, MhamaneD, Madhavi S, Ogale S. Nano Energy, 2013, 2(5): 890.

[86]
Guo W X, Sun W W, Lv L P, Kong S F, Wang Y. ACS Nano, 2017, 11(4): 4198.

[87]
Zhang L, Wu H B, Madhavi S, Hng H H, David Lou X W. J. Am. Chem. Soc., 2012, 134(42): 17388.

[88]
Soundharrajan V, Sambandam B, Song J J, Kim S, Jo J, Duong P T, Kim S, Mathew V, Kim J. J. Energy Chem., 2018, 27(1): 300.

[89]
Zhang F, JiangD G, Zhang X G. Nano Struct. Nano Objects, 2016, 5: 1.

[90]
Xu J M, Tang H B, Xu T T, WuD, Shi Z F, Tian Y T, Li X J. Ionics, 2017, 23(12): 3273.

[91]
Bi Z H, Paranthaman M P, Guo B K, Unocic R R, Meyer H M, Bridges C A, Sun X G, Dai S. J. Mater. Chem. A, 2014, 2(6): 1818.

[92]
Mai Y Y, Zhang F, Feng X L. Nanoscale, 2014, 6(1): 106.

[93]
Lee S, Ha J, Choi J, Song T, Lee J W, Paik U. ACS Appl. Mater. Interfaces, 2013, 5(22): 11525.

[94]
Hao B, Yan Y, Wang X B, Chen G. ACS Appl. Mater. Interfaces, 2013, 5(13): 6285.

[95]
Wang Z Q, Li X, Xu H, Yang Y, Cui Y J, Pan H G, Wang Z Y, Chen B L, Qian GD. J. Mater. Chem. A, 2014, 2(31): 12571.

[96]
Zhang W B, Pang H C, Sun W W, Lv L P, Wang Y. Electrochem. Commun., 2017, 84: 80.

[97]
Wang X X, Xue H J, Na Z L, YinD M, Li Q, Wang C L, Wang L M, Huang G. J. Power Sources, 2018, 396: 659.

[98]
Wang CD, Li Y, Ruan Y J, Jiang J J, Wu Q H. Mater. Today Energy, 2017, 3: 1.

[99]
Xiong Q Q, Tu J P, Shi S J, Liu X Y, Wang X L, Gu CD. J. Power Sources, 2014, 256: 153.

[100]
Xing Z, Ju Z C, Yang J, Xu H Y, Qian Y T. Electrochim. Acta, 2013, 102: 51.

[101]
Yang L G, Wang X, Zheng F C. J. Mater. Sci. Mater. Electron., 2019, 30(17): 16687.

[102]
Du J C, Tang Y H, Wang Y, Shi P H, Fan J C, Xu Q J, Min Y L. Dalton Trans., 2018, 47(22): 7571.

[103]
Zheng F C, ZhuD Q, Shi X H, Chen Q W. J. Mater. Chem. A, 2015, 3(6): 2815.

[104]
Wu L L, Wang Z, Long Y, Li J, Liu Y, Wang Q S, Wang X, Song S Y, Liu X G, Zhang H J. Small, 2017, 13(17): 1604270.

[105]
Wu R B, Qian X K, Zhou K, Wei J, Lou J, Ajayan P M. ACS Nano, 2014, 8(6): 6297.

[106]
Guo H, Li T T, Chen W W, Liu L X, Yang X J, Wang Y P, Guo Y C. Nanoscale, 2014, 6(24): 15168.

[107]
Guo H, Li T T, Chen W W, Liu L X, Qiao J L, Zhang J J. Sci. Rep., 2015, 5: 13310.

[108]
Li H, Liang M, Sun W W, Wang Y. Adv. Funct. Mater., 2016, 26(7): 982.

[109]
Chen C, Qian S, Ding Y, Yao T H, Guo J H, Wang H K. J. Funct. Mater., 2020, 51(10): 10116.

(陈川, 钱森, 丁一, 姚天浩, 郭经红, 王红康. 功能材料, 2020, 51(10): 10116.).

[110]
Yu L, Yang J F, David Lou X W. Angew. Chem. Int. Ed., 2016, 128(43): 13620.

[111]
Zhou Y L, YanD, Xu H Y, Feng J K, Jiang X L, Yue J, Yang J, Qian Y T. Nano Energy, 2015, 12: 528.

[112]
Sennu P, Christy M, Aravindan V, Lee Y G, Nahm K S, Lee Y S. Chem. Mater., 2015, 27(16): 5726.

[113]
Duan J L, Zou Y L, Li Z Y, Long B. Powder Technol., 2019, 354: 834.

[114]
Liu F, Song S Y, XueD F, Zhang H J. Adv. Mater., 2012, 24(8): 1089.

[115]
Oktaviano H S, Yamada K, Waki K. J. Mater. Chem., 2012, 22(48): 25167.

[116]
Zhang H B, Nai J W, Yu L, David Lou X W. Joule, 2017, 1(1): 77.

[117]
Zhang Y, Sha M, Fu Q, Zhao H, Lei Y. Mater. Today Sustain., 2022, 18: 100156.

[118]
Chen Y Y, Du W Q, Dou B X, Chen J H, Hu L, Zeb A, Lin X M. CrystEngComm, 2022, 24(15): 2729.

[119]
Zuo L, Chen S H, Wu J F, Wang L, Hou H Q, Song Y H. RSC Adv., 2014, 4(106): 61604.

[120]
Takamura T, Awano H, Ura T, Sumiya K. J. Power Sources, 1997, 68(1): 114.

[121]
Li A, Tong Y, Cao B, Song H H, Li Z H, Chen X H, Zhou J S, Chen G, Luo H M. Sci. Rep., 2017, 7: 40574.

[122]
Turon Teixidor G, Park B Y, Mukherjee P P, Kang Q, Madou M J. Electrochim. Acta, 2009, 54(24): 5928.

[123]
Park B Y, Zaouk R, Wang C L, Madou M J. J. Electrochem. Soc., 2007, 154(2): P1.

[124]
Luo Y M, Sun L, Xu F, Wang Z Q. Key Eng. Mater., 2017, 727: 705.

[125]
Peng H J, Hao G X, Chu Z H, Lin Y W, Lin X M, Cai Y P. RSC Adv., 2017, 7(54): 34104.

[126]
Zheng G X, Chen M H, Zhang H R, Zhang J W, Liang X Q, Qi M L, Yin J H. Surf. Coat. Technol., 2019, 359: 384.

[127]
Shen C, Zhao C C, Xin F X, Cao C, Han W Q. Electrochim. Acta, 2015, 180: 852.

[128]
Chu K N, Hu M L, Song B, Chen S L, Li J Y, Zheng F C, Li Z Q, Li R, Zhou J Y. RSC Adv., 2023, 13(9): 5634.

[129]
He X L, Cai Y Q, Zhao W, Zhuang Q C, Ju Z C. J. Phys. Chem. Solids, 2020, 147: 109639.

[130]
Shi X Z, Gong J, Kierzek K, Michalkiewicz B, Zhang S, Chu P K, Chen X C, Tang T, Mijowska E. New J. Chem., 2019, 43(26): 10405.

[131]
Tong Y L, Ji D, Wang P, Zhou H, Akhtar K, Shen X P, Zhang J H, Yuan A H. RSC Adv., 2017, 7(40): 25182.

[132]
Yang Y, Zheng F C, Xia G L, Lun Z Y, Chen Q W. J. Mater. Chem. A, 2015, 3(36): 18657.

[133]
Mao Y, Duan H, Xu B, Zhang L, Hu Y S, Zhao C C, Wang Z X, Chen L Q, Yang Y S. Energy Environ. Sci., 2012, 5(7): 7950.

[134]
Yu Y X. Phys. Chem. Chem. Phys., 2013, 15(39): 16819.

[135]
Zheng F C, Yang Y, Chen Q W. Nat. Commun., 2014, 5: 5261.

[136]
Cao N, Du H L, Wang J L, Ma W X, Ma W L, Tian C. J. Chin. Ceram. Soc., 2018, 46(12): 1748.

(曹娜, 杜慧玲, 王金磊, 马武祥, 马万里, 田超. 硅酸盐学报, 2018, 46(12): 1748.).

[137]
Li G H, Li F C, Yang H, Cheng F Y, Xu N, Shi W, Cheng P. Inorg. Chem. Commun., 2016, 64: 63.

[138]
Gao G L, WangD Y, Zeng Q, Shen C. J. South China Norm. Univ. Nat. Sci. Ed., 2018, 50(2): 30.

(高国梁, 王德宇, 曾群, 沈彩. 华南师范大学学报(自然科学版), 2018, 50(2): 30.).

[139]
HongD Y, Hwang Y K, Serre C, FÉrey G, Chang J S. Adv. Funct. Mater., 2009, 19(10): 1537.

[140]
Wei R P, Dong Y T, Zhang Y Y, Zhang R, Al-Tahan M A, Zhang J M. J. Colloid Interface Sci., 2021, 582: 236.

[141]
Jin Y, Zhao C C, Sun Z X, Lin Y C, Chen L, WangD Y, Shen C. RSC Adv., 2016, 6(36): 30763.

[142]
Zhang C H, Hu W Q, Jiang H, Chang J K, Zheng M S, Wu Q H, Dong Q F. Electrochim. Acta, 2017, 246: 528.

[143]
Li C, Lou X B, Yang Q, Zou Y M, Hu B W. Chem. Eng. J., 2017, 326: 1000.

[144]
Wang J, Polleux J, Lim J, Dunn B. J. Phys. Chem. C, 2007, 111(40): 14925.

[145]
He S H, Li Z P, Ma L M, Wang J Q, Yang S R. New J. Chem., 2017, 41(23): 14209.

[146]
Vermoortele F, Vandichel M, van de Voorde B, Ameloot R, Waroquier M, van Speybroeck V, de VosD E. Angew. Chem. Int. Ed., 2012, 51(20): 4887.

[147]
Zhu W, Chen Z, Pan Y, Dai R Y, Wu Y, Zhuang Z B, WangD S, Peng Q, Chen C, Li YD. Adv. Mater., 2019, 31(38): 1800426.

[148]
Sun X M, Gao G, YanD W, Feng C Q. Appl. Surf. Sci., 2017, 405: 52.

[149]
Zheng X Z, Li Y F, Xu Y X, Hong Z S, Wei MD. CrystEngComm, 2012, 14(6): 2112.

[150]
Wang B X, Wang Z Q, Cui Y J, Yang Y, Wang Z Y, Qian GD. RSC Adv., 2015, 5(103): 84662.

[151]
Wang P, Shen M Q, Zhou H, Meng C F, Yuan A H. Small, 2019, 15(47): 1903522.

[152]
Kang Y, Zhang Y H, Shi Q, Shi H W, XueD F, Shi F N. J. Colloid Interface Sci., 2021, 585: 705.

[153]
Zhang L, Wu H B, David Lou X W. J. Am. Chem. Soc., 2013, 135(29): 10664.

[154]
Li J B, YanD, Hou S J, Lu T, Yao Y F, ChuaD H C, Pan L K. Chem. Eng. J., 2018, 335: 579.

[155]
Xu W W, Cui XD, Xie Z Q, Dietrich G, Wang Y. Electrochim. Acta, 2016, 222: 1021.

[156]
Hu L, Huang Y M, Zhang F P, Chen Q W. Nanoscale, 2013, 5(10): 4186.

[157]
Guo W X, Sun W W, Wang Y. ACS Nano, 2015, 9(11): 11462.

[158]
Wang B X, Wang Z Q, Cui Y J, Yang Y, Wang Z Y, Chen B L, Qian GD. Microporous Mesoporous Mater., 2015, 203: 86.

[159]
WangD P, Fu M S, Ha Y, Wang H, Wu R B. J. Colloid Interface Sci., 2018, 529: 265.

[160]
Yang X, Tang Y B, Huang X, Xue H T, Kang W P, Li W Y, Ng T W, Lee C S. J. Power Sources, 2015, 284: 109.

[161]
Zhang S L, Guan B Y, Wu H B, David Lou X W. Nano Micro Lett., 2018, 10(3): 44.

[162]
Lu Y, Yu L, Wu M, Wang Y, David Lou X W. Adv. Mater., 2018, 30(1): 1702875.

[163]
Huang G, Zhang F F, Zhang L L, Du X C, Wang J W, Wang L M. J. Mater. Chem. A, 2014, 2(21): 8048.

[164]
Huang G, Zhang L L, Zhang F F, Wang L M. Nanoscale, 2014, 6(10): 5509.

[165]
Zhong M, YangD H, Kong L J, Shuang W, Zhang Y H, Bu X H. Dalton Trans., 2017, 46(45): 15947.

[166]
Xu X H, Cao K Z, Wang Y J, Jiao L F. J. Mater. Chem. A, 2016, 4(16): 6042.

[167]
Huang G, YinD M, Wang L M. J. Mater. Chem. A, 2016, 4(39): 15106.

[168]
Hou L R, Lian L, Zhang L H, Pang G, Yuan C Z, Zhang X G. Adv. Funct. Mater., 2015, 25(2): 238.

[169]
Yang X, Xue H T, Yang QD, Yuan R, Kang W P, Lee C S. Chem. Eng. J., 2017, 308: 340.

[170]
CaiD P, Zhan H B, Wang T H. Mater. Lett., 2017, 197: 241.

[171]
Wu Y Z, Meng J S, Li Q, Niu C J, Wang X P, Yang W, Li W, Mai L Q. Nano Res., 2017, 10(7): 2364.

[172]
Zhao K N, Liu F N, Niu C J, Xu W W, Dong Y F, Zhang L, Xie S M, Yan M Y, Wei Q L, ZhaoD Y, Mai L Q. Adv. Sci., 2015, 2(12): 1500154.

[173]
Liu L X, Guo H, Liu J J, Qian F, Zhang C H, Li T T, Chen W W, Yang X J, Guo Y C. Chem. Commun., 2014, 50(67): 9485.

[174]
Zhang J, Chu R X, Chen Y L, Jiang H, Zeng Y B, Chen X, Zhang Y, Huang N M, Guo H. J. Alloys Compd., 2019, 797: 83.

[175]
Peng H J, Hao G X, Chu Z H, He C L, Lin X M, Cai Y P. J. Alloys Compd., 2017, 727: 1020.

[176]
Yin H, Yu X X, Li Q W, Cao M L, Zhang W, Zhao H, Zhu M Q. J. Alloys Compd., 2017, 706: 97.

[177]
Peng H J, Hao G X, Chu Z H, Lin J, Lin X M, Cai Y P. Cryst. GrowthDes., 2017, 17(11): 5881.

[178]
Sambandam B, Soundharrajan V, Song J J, Kim S, Jo J, TungD P, Kim S, Mathew V, Kim J. Inorg. Chem. Front., 2016, 3(12): 1609.

[179]
Tang B. MasteralDissertation of Beijing University of Chemical Technology, 2017.

(唐波. 北京化工大学硕士论文, 2017.).

[180]
SunD, Tang Y G, YeD L, Yan J, Zhou H S, Wang H Y. ACS Appl. Mater. Interfaces, 2017, 9(6): 5254.

[181]
Yang T, Liu Y G, Huang Z H, Liu J W, Bian P J, Ling CD, Liu H, Wang G X, Zheng R K. J. Alloys Compd., 2018, 735: 1079.

[182]
Yang S J, Nam S, Kim T, Im J H, Jung H, Kang J H, Wi S, Park B, Park C R. J. Am. Chem. Soc., 2013, 135(20): 7394.

[183]
Wang H B, Pan Q M, Cheng Y X, Zhao J W, Yin G P. Electrochim. Acta, 2009, 54(10): 2851.

[184]
Jamnik J, Maier J. Phys. Chem. Chem. Phys., 2003, 5(23): 5215.

[185]
Farrusseng D, Aguado S, Pinel C. Angew. Chem. Int. Ed., 2009, 48(41): 7502.

[186]
Chen Y Q, Zheng L, Fu Y Y, Zhou R H, Song Y H, Chen S H. RSC Adv., 2016, 6(89): 85917.

[187]
Yang L, Tian Y, Ge P, Zhao G G, Pu T C, Yang Y C, Zou G Q, Hou H S, Huang L P, Ji X B. ChemElectroChem, 2018, 5(22): 3426.

[188]
Zhao L, Liu W, Liu S, Wang J F, Wang H L, Chen J X. J. Mater. Chem. A, 2015, 3(27): 14210.

[189]
Wang M H, Yang H, Zhou X L, Shi W, Zhou Z, Cheng P. Chem. Commun., 2016, 52(4): 717.

[190]
Jin L N, Zhao X S, Qian X Y, Wang S W, Shen X Q, Dong MD. Mater. Lett., 2017, 199: 176.

[191]
Xu H J, Wang L, Zhong J, Wang T, Cao J H, Wang Y Y, Li X Q, Fei H L, Zhu J, Duan XD. Energy Environmental Mater., 2020, 3(2): 177.

[192]
Huang G, Zhang F F, Du X C, Qin Y L, YinD M, Wang L M. ACS Nano, 2015, 9(2): 1592.

[193]
Zou Y L, Qi Z G, Ma Z S, Jiang W J, Hu R W, Duan J L. J. Electroanal. Chem., 2017, 788: 184.

[194]
Xu Y Q, Hou S J, Yang G, Lu T, Pan L K. J. Solid State Electrochem., 2018, 22(3): 785.

[195]
Chen Y, Yu L, David Lou X. Angew. Chem. Int. Edit., 2016, 55(20): 5990.

[196]
Wang F X, Han Q G, Yi Z, GengD, Li X, Wang Z, Wang L M. J. Electroanal. Chem., 2017, 807: 196.

[197]
JiD, Zhou H, Tong Y L, Wang J P, Zhu M Z, Chen T H, Yuan A H. Chem. Eng. J., 2017, 313: 1623.

[198]
Shao J X, Zhou H, Feng J H, Zhu M Z, Yuan A H. J. Alloys Compd., 2019, 784: 869.

[199]
Zhang L, Liu W X, Shi W H, Xu X L, Mao J, Li P, Ye C Z, Yin R L, Ye S F, Liu X Y, Cao X H, Gao C. Chem. A Eur. J., 2018, 24(52): 13689.

[200]
YinD M, Huang G, Sun Q J, Li Q, Wang X X, YuanD X, Wang C L, Wang L M. Electrochim. Acta, 2016, 215: 410.

[201]
Tian S Y, Zheng G X, Liu Q, Ren M Y, Yin J H. Int. J. Electrochem. Sci., 2019, 14(10): 9459.

[202]
Niu J L, Hao G X, Lin J, He X B, Sathishkumar P, Lin X M, Cai Y P. Inorg. Chem., 2017, 56(16): 9966.

[203]
Wang Z H, Xiong X Q, Qie L, Huang Y H. Electrochim. Acta, 2013, 106: 320.

[204]
Chu K N, Li Z Q, Xu S K, Yao G, Xu Y, Niu P, Zheng F C. J. Alloys Compd., 2021, 854: 157264.

[205]
Zhang Q Y, Liu F J, Gao P A, Zhao P, Guo H X, Wang L, Wan Z L. Mater. Lett., 2020, 268: 127366.

[206]
Zheng F C. DoctoralDissertation of University of Science and Technology of China, 2015.

(郑方才. 中国科学技术大学博士论文, 2015.).

[207]
Zheng F C, He M N, Yang Y, Chen Q W. Nanoscale, 2015, 7(8): 3410.

[208]
Wang Y, Gao Y J, Shao J, Holze R, Chen Z, Yun Y X, Qu Q T, Zheng H H. J. Mater. Chem. A, 2018, 6(8): 3659.

[209]
Kang W P, Zhang Y, Fan L L, Zhang L L, Dai F N, Wang R M, SunD F. ACS Appl. Mater. Interfaces, 2017, 9(12): 10602.

[210]
Han X, Chen W M, Han X G, Tan Y Z, SunD. J. Mater. Chem. A, 2016, 4(34): 13040.

[211]
Hou Y, Li J Y, Wen Z H, Cui S M, Yuan C, Chen J H. Nano Energy, 2015, 12: 1.

[212]
Sun Y, Huang F Z, Li S K, Shen Y H, Xie A J. Nano Res., 2017, 10(10): 3457.

[213]
Ding Y C, Hu L H, HeD C, Peng Y Q, Niu Y J, Li Z Q, Zhang X X, Chen S H. Chem. Eng. J., 2020, 380: 122489.

[214]
Pang Y C, Chen S, Xiao C H, Ma SD, Ding S J. J. Mater. Chem. A, 2019, 7(8): 4126.

[215]
Sui Z Y, Zhang P Y, Xu M Y, Liu Y W, Wei Z X, Han B H. ACS Appl. Mater. Interfaces, 2017, 9(49): 43171.

[216]
Park J, Ju J B, Choi W, Kim S O. J. Alloys Compd., 2019, 773: 960.

[217]
Yue H Y, Shi Z P, Wang Q X, Cao Z X, Dong H Y, Qiao Y, Yin Y H, Yang S T. ACS Appl. Mater. Interfaces, 2014, 6(19): 17067.

[218]
Li J K, WangD, Zhou J S, Hou L, Gao F M. ChemElectroChem, 2019, 6(3): 917.

[219]
Zhang X, Cao W J, Zou W W, ZhaoD Y, Zhao H B, Fang J H. Ferroelectrics, 2019, 547(1): 59.

[220]
Li J K, WangD, Zhou J S, Hou L, Gao F M. J. Alloys Compd., 2019, 793: 247.

[221]
Zhang J L, Chen Z H. Front. Mater., 2020, 7: 178.

[222]
Zou Y L, Li Z Y, Liu Y L, Duan J L, Long B. J. Alloys Compd., 2020, 820: 153085.

[223]
Mujahid M, Ullah Khan R, Mumtaz M, Mubasher, Soomro S A, Ullah S. Ceram. Int., 2019, 45(7): 8486.

[224]
Cai M C, Cai S R, Zheng M S, Dong Q F. J. Electrochem., 2014, 20(2): 101.

(蔡默超, 蔡森荣, 郑明森, 董全峰. 电化学, 2014, 20(2): 101.).

[225]
Yang H, Zhang K, Wang Y, Yan C, Lin S. J. Phys. Chem. Solids, 2017, 115: 317.

[226]
WangD, Zhou W W, Zhang R, Huang X X, Zeng J J, Mao Y F, Ding C Y, Zhang J, Liu J P, Wen G W. J. Mater. Chem. A, 2018, 6(7): 2974.

[227]
He Z S, Wang K, Zhu S S, Huang L A, Chen M M, Guo J F, Pei S E, Shao H B, Wang J M. ACS Appl. Mater. Interfaces, 2018, 10(13): 10974.

[228]
Zhang W, Wang B, Luo H, Jin F, Ruan T T, WangD L. J. Alloys Compd., 2019, 803: 664.

[229]
Wang Y Z, Kong M G, Liu Z W, Lin C C, Zeng Y. J. Mater. Chem. A, 2017, 5(46): 24269.

[230]
Sun W W, Chen S, Wang Y. Dalton Trans., 2019, 48(6): 2019.

[231]
Wu M H, Chen H Q, Lv L P, Wang Y. Chem. Eng. J., 2019, 373: 985.

[232]
Zou F, Hu X L, Li Z, Qie L, Hu C C, Zeng R, Jiang Y, Huang Y H. Adv. Mater., 2014, 26(38): 6622.

[233]
Zhao Y C, Li X, Liu JD, Wang C G, Zhao Y Y, Yue G H. ACS Appl. Mater. Interfaces, 2016, 8(10): 6472.

[234]
Yuan C Z, Cao H, Zhu S Q, Hua H, Hou L R. J. Mater. Chem. A, 2015, 3(40): 20389.

[235]
Niu J L, Zeng C H, Peng H J, Lin X M, Sathishkumar P, Cai Y P. Small, 2017, 13(47): 170215.

[236]
Li J F, Han L, Li Y Q, Li J L, Zhu G, Zhang X J, Lu T, Pan L K. Chem. Eng. J., 2020, 380: 122590.

[237]
Lu M J, Liao C, Jiang C, Du Y, Zhang Z, Wu S P. Electrochim. Acta, 2017, 250: 196.

[238]
Zhang L G, Li H, Xie H T, Chen T X, Yang C, Wang JD. J. Mater. Res., 2018, 33(10): 1496.

[239]
Mujtaba J, Sun H Y, Huang G Y, Zhao Y Y, Arandiyan H, Sun G X, Xu S M, Zhu J. RSC Adv., 2016, 6(38): 31775.

[240]
Zeng P Y, Li J W, Ye M, Zhuo K F, Fang Z. Chem. A Eur. J., 2017, 23(40): 9517.

[241]
Grugeon S, Laruelle S, Dupont L, Tarascon J M. Solid State Sci., 2003, 5(6): 895.

[242]
Wang F, Li K, Wang X, Li J Q, Pan J, Feng J, Liu K, Song S Y, Zhang H J. ACS Appl. Energy Mater., 2018, 1(11): 6242.

[243]
Chen L, Yang W J, Li X Y, Han L J, Wei MD. J. Mater. Chem. A, 2019, 7(17): 10331.

[244]
Paraknowitsch J P, Thomas A. Energy Environ. Sci., 2013, 6(10): 2839.

[245]
Wu R B, Wang D P, Rui X H, Liu B, Zhou K, Law A W K, Yan Q Y, Wei J, Chen Z. Adv. Mater., 2015, 27(19): 3038.

[246]
Wang Q F, Zou R Q, Xia W, Ma J, Qiu B, Mahmood A, Zhao R, Yang Y, XiaD G, Xu Q. Small, 2015, 11(21): 2511.

[247]
Song J B, Zhang C Y, Zhang J H, Zhou H, Chen L, Bian L L, Yuan A H. J. Nanopart. Res., 2019, 21(5): 90.

[248]
Wang J L, Wang J W, Han L F, Liao C, Cai W, Kan Y C, Hu Y. Nanoscale, 2019, 11(43): 20996.

[249]
Tian R, Zhou Y, Duan H N, Guo Y P, Li H, Chen K F, XueD F, Liu H Z. ACS Appl. Energy Mater., 2018, 1(2): 402.

[250]
Yang T, YangD X, Liu Y G, Liu J, Chen Y F, Bao L, Lu X X, Xiong Q Q, Qin H Y, Ji Z G, Ling CD, Zheng R K. Electrochimica Acta, 2018, 290: 193.

[251]
Zhao J G, Hu Z, SunD Z, Jia H, Liu X M. Nanomaterials, 2019, 9(4): 492.

[252]
Yin W H, Li W Y, Wang K, Chai W W, Ye W K, Rui Y C, Tang B. Electrochim. Acta, 2019, 318: 673.

[253]
Huang W, Li S, Cao X Y, Hou C Y, Zhang Z, Feng J K, Ci L J, Si P C, Chi Q J. ACS Sustain. Chem. Eng., 2017, 5(6): 5039.

[254]
Fu Y, Zhang Z A, Yang X, Gan Y Q, Chen W. RSC Adv., 2015, 5(106): 86941.

[255]
Ding H, Huang H C, Zhang X K, Xie L, Fan J Q, Jiang T, ShiD A, Ma N, Tsai F C. ChemElectroChem, 2019, 6(22): 5617.

[256]
Chen Z L, Wu R B, Wang H, Jiang Y K, Jin L, Guo Y H, Song Y, Fang F, SunD L. Chem. Eng. J., 2017, 326: 680.

[257]
Wu HD, Li G, Li Y, Geng Z X, Ren T Q, Cai T F, Yang Z X. Cryst. Res. Technol., 2019, 54(6): 1800281.

[258]
Ma Y, Ma Y J, Kim G T, Diemant T, Behm R J, GeigerD, Kaiser U, Varzi A, Passerini S. Adv. Energy Mater., 2019, 9(43): 1902077.

[259]
Xue H L, Yue S, Wang J, Zhao Y, Li Q, Yin M M, Wang S S, Feng C H, Wu Q, Li H S, ShiD X, Jiao Q Z. J. Electroanal. Chem., 2019, 840: 230.

[260]
Shao J, Gao T, Qu Q T, Shi Q, Zuo Z C, Zheng H H. J. Power Sources, 2016, 324: 1.

[261]
Hu C, Ma K, Hu Y J, Chen A P, Saha P, Jiang H, Li C Z. Green Energy Environ., 2021, 6(1): 75.

[262]
Yuan D X, Huang G, YinD M, Wang X X, Wang C L, Wang L M. ACS Appl. Mater. Interfaces, 2017, 9(21): 18178.

[263]
Aslam M K, Ahmad Shah S S, Li S, Chen C G. J. Mater. Chem. A, 2018, 6(29): 14083.

[264]
Li J B, YanD, Lu T, Yao Y F, Pan L K. Chem. Eng. J., 2017, 325: 14.

[265]
Jiang T C, Bu F X, Liu B L, Hao G L, Xu Y X. New J. Chem., 2017, 41(12): 5121.

[266]
Yang T, Liu Y G, YangD X, Deng B B, Huang Z H, Ling CD, Liu H, Wang G X, Guo Z P, Zheng R K. Energy Storage Mater., 2019, 17: 374.

[267]
Yang T, Liu J W, Yang D X, Mao Q N, Zhong J S, Yuan Y J, Li X Y, Zheng X, Ji Z G, Liu H, Wang G X, Zheng R K. ACS Appl. Energy Mater., 2020, 3(11): 11073.

[268]
Liu H, Li Z, Zhang L, Ruan H, Hu R. Nanoscale Res. Lett., 2019, 14: 237.

[269]
Tao S, Cui P X, Cong S, Chen S M, WuD J, Qian B, Song L, Marcelli A. Sci. China Mater., 2020, 63(9): 1672.

[270]
Wang X X, Na Z L, YinD M, Wang C L, Wu Y M, Huang G, Wang L M. ACS Nano, 2018, 12(12): 12238.

[271]
Xia G L, Su J W, Li M S, Jiang P, Yang Y, Chen Q W. J. Mater. Chem. A, 2017, 5(21): 10321.

[272]
Qi J, Shi Z P, Li X, Gao B X, Wang H, Yang L C, Tang Y, Zhu M. J. Alloys Compd., 2019, 786: 284.

[273]
Yan H R, Qiu F, Wang K L, ZhangD P, Chen J H, Niu F E. Plast. Sci. Technol., 2020, 48(11): 7.

(闫浩然, 邱帆, 汪楷丽, 张大鹏, 陈君华, 牛斐洱. 塑料科技, 2020, 48(11): 7.).

[274]
Zhou K Q, Lai L F, Zhen Y C, Hong Z S, Guo J H, Huang Z G. Chem. Eng. J., 2017, 316: 137.

[275]
Zhong M, He W W, Shuang W, Liu Y Y, Hu T L, Bu X H. Inorg. Chem., 2018, 57(8): 4620.

[276]
Chen S H, Zhou R H, Chen Y Q, Li P, Song Y H, Wang L. Int. J. Electrochem. Sci., 2016, 11(12): 10522.

[277]
He Q, Liu J S, Li Z H, Li Q, Xu L, Zhang B X, Meng J S, Wu Y Z, Mai L Q. Small, 2017, 13(37): 1701504.

[278]
Zou F, Chen Y M, Liu K W, Yu Z T, Liang W F, Bhaway S M, Gao M, Zhu Y. ACS Nano, 2016, 10(1): 377.

[279]
Joshi B, Samuel E, Il Kim Y, Kim M W, Jo H S, Swihart M T, Yoon W Y, Yoon S S. Chem. Eng. J., 2018, 351: 127.

Outlines

/