Recent advances in porous multimetallic alloy-based anodes for rechargeable alkali metal-ion batteries
Abstract
The rechargeable alkali metal-ion batteries (RAMIBs) are highly promising candidates for next-generation efficient energy storage devices, owing to their outstanding theoretical specific capacities and extremely low electrochemical potentials. However, RAMIBs possess unsuitable lifespans, low mechanical durability and inevitable side reactions attributable to their inherent severe volumetric/structure alteration during the charge-discharge cycles. These hitches could be solved using porous multimetallic alloy-based anodes, due to their impressive specific capacities, low working potential, low cost, and earth-abundance, which can meet sustainability and practical application needs. Meanwhile, great surface area, electrical conductivity, structural stability, and ability to accommodate the generated alkali metal ions can yield satisfactory coulomb efficiency and long durability. Immense efforts are dedicated to rationally designing porous multimetallic alloy-based anodes for RAMIBs, so it is essential to provide timely updates on this research area. Herein, we reviewed recent advances in porous multimetallic alloy-based anodes (i.e., Sn, Mn, Mo, Co, V, and Fe) for RAMIBs (i.e., lithium-ion batteries, sodium-ion batteries, and potassium-ion batteries. This is rooted in the engineering approaches (i.e., template-based, hydrothermal/solvothermal, chemical reduction, electrochemical deposition, sol-gel, and electrospinning) to fundamental insights (i.e., mechanisms, key parameters, and calculations) and precise evaluation for structural changes, and mechanisms by various experimental, theoretical, and in-situ analysis to optimizing their performance. Also, advances in RAMIBs recycling and circular economy were discussed. Eventually, we highlighted the current drawbacks and provided proposed perspectives to solve these issues and enable practical utilization of such anodes for large-scale applications.
Keywords
INTRODUCTION
There are global concerns regarding environmental threats (i.e., global warming) and nonrenewable fossil fuel exhaustion, which raise the need for gas conversion reactions[1,2] and developing green energy sources[3-8]. Lithium-ion batteries (LIBs) with their trustworthy power source are the most common and efficient energy storage devices widely used in various portable devices and electric cars, but the high cost and earth-rarity of Li are crucial issues in the sustainable future[9-12]. The efficiency and performance of LIBs are shaped by the anode (i.e., shape and composition) and electrolytes[11-13]. Rechargeable alkali metal-ion batteries (RAMIBs) emerged as a promising alternative to traditional Li-batteries, owing to their tolerable cost, reliable theoretical specific capacities (Li = 3,860 mAh/g; Na = 1,166 mAh/g; K = 685 mAh/g), and low electrochemical redox voltage [i.e., Li = 3.04 V, Na = 2.71 V, and K = 2.93 V vs. Standard Hydrogen Electrode (SHE)][14]. However, the alkali metal-ion batteries are still not up to the level of commercial scale, owing to their viability to dendrite growth, repetitive breakage or regeneration of the solid electrolyte interphase, causing subsequent unreliable safety, low energy efficiency, and low durability. Immense efforts devoted to solving these barriers lie in the development of controlled size, composition, and morphology of the anodes, new electrolytes (i.e., ether-based electrolytes, solid-state electrolytes, and multifunctional electrolyte additives), and the design of new hierarchical supports[10,15-18]. These solutions led to decreasing the local current density, controlling alkali-metal deposition, enhancing mechanical properties, increasing the ionic conductivity, ceasing dendrite growth, and stabilizing alkali metal anodes from corrosion. However, the intrinsically great activity of alkali metals and the significant volume alteration during plating or stripping are critical barriers for practical applications of RAMIBs[19,20].
Another promising solution is the rational formation of porous multimetallic alloy-based anodes (PMMAs), especially porous transition metals, which are the most promising to solve the current hitches in RAMIBs, owing to their impressive specific capacities and low working voltage[16,21,22]. This is in addition to the low cost and abundance of transition metals, which can meet the requirements of both sustainability and practical application[23,24]. Moreover, PMMA anodes act as a protective layer, which enhances the cyclic durability, rate capability, and plating/stripping Coulombic efficiency (CE) of RAMIBs[25]. Besides, PMMAs are endowed with enhanced specific capacity (SC), alkaline metal (Li+/Na+/K+) ion diffusion/kinetics, stable architectural integrity, negligible volume expansion and pulverization in RAMIBs, compared to their monometallic analogous[26,27]. Notably, using at least one alkali metal with PMMAs decreases the high reactivity of alkali metal, provides abundant channels for quick ion transfer, and acts as mechanical supports, hence preferring quick and orderly metal plating and stripping[23]. Additionally, direct use of PMMAs with their interphases can plausibly serve as an artificially protective layer to tailor the ion flux, increase rapid ion diffusion, and inhibit the side reactions between the electrolyte and electrode. This has led to an exponential increase in publications (i.e., 100 articles, cited ~1,500 times) [Figure 1A], with much attention on the rational design of PMMAs for alkali metal-ion batteries; hence, it is important to provide timely updates on this research area. Although there are various recent excellent reviews on alkali metal batteries and metal anodes [Supplementary Table 1], the PMMA anodes for LIBs, sodium-ion batteries (SIBs), and potassium-ion batteries (PIBs) have not yet been reviewed to the best of our knowledge[14,27,28].
Figure 1. (A) Database literature survey using keywords "multimetallic alloy anodes for batteries" according to the "Web of Science" and (B) review focus.
This review summarized the progress in the fabrication methods (i.e., template-based,
FUNDAMENTAL AND MECHANISMS OF RECHARGEABLE ALKALI-ION BATTERIES
The RAMIBs, including LIBs, SIBs and PIBs, comprise four basic parts: anodes, cathodes, electrolytes and separators. The electrodes (anodes and cathodes) are good electrical conductors, while electrolyte functions as an ionic conductor, as briefly explained below.
• Anode: This is a negative electrode (metal or alloy) where negatively charged ions migrate.
• Cathode: This is a positive electrode (layered structure of metallic oxide, sulfides, and/or carbon materials) where positively charged ions migrate.
• Electrolyte: This is a solution containing dissociated salts that allows the movement of ions between the anode and cathode.
• Separator: This is a permeable or porous membrane that keeps the anode and cathode apart to eschew electrical short circuits but allows the movement of ionic charge carriers required for the passage of current in the battery.
The anode and cathode are often made up of different conducting materials, kept apart by the separator and immersed in the electrolyte solution. A typical RAMIB is shown in Figure 2 with similar mechanisms. For example, in LIBs, charging and discharging ensues via the movement of A+-ions between the cathode and anode, coupled with exchange of electrons (e-s). Particularly, the charging Li/Na/K is delithiated/desodiated/depotassiated from the cathode (i.e., Li-based compound) and the interlayer anode [i.e., metal alloys (MAs) or carbon] are lithiated/sodiated/potassiated with Li/Na/K, and vice-versa during the discharge process, as given in Eqs. 1 and 2, where A represents Li+ or Na+ or K+ ions and MAs are metal alloys.
Parameters of indicator and calculations
In addition to the materials’ cost, toxicity and safety of the components of RAMIBs, their performance can be assessed by various parameters of indicators, including cell voltage, C-rate, SC, capacity retention (CR), ED, power density, CE, voltage efficiency and energy efficiency, which aid in the rational design and fabrication of suitable intercalation electrodes for high ED batteries[29-31]. These are briefly discussed below:
• Cell voltage: This is the driving force for the movement of alkali metal ions and electrons from one electrode to the other. Thermodynamically, it is the electrochemical potential difference, which is proportional to Gibb’s free energy (∆G), as given in
Here, n represents the total number of electrons, Faraday’s constant F = 96,485.3 (C/mol), and V refers to the cell voltage.
Discharge rate or C-rate: This is the rate at which a battery is discharged/charged in relation to its capacity. For instance, 1C-rate indicates that a battery is fully discharged in 60 min (1 h), and 10 Ah capacity of a battery gives a 1C-rate equivalent to 10 A.
• Specific capacity (SC): This is the amount of electrons intercalated/deintercalated per unit mass of the electrode during cycling. The SC is reported with mass or volume of an active electrode, called gravimetric SC (GSC, mAh/g) or volumetric SC (VSC, mAh/cm3), respectively, and calculated using
Here, the number of available sites for active Mn+ ions intercalated into the host matrix of the battery’s capacity. Also, the storage capacity can be measured with respect to unit area, called areal capacity (mAh/cm2). The VSC is particularly crucial when designing batteries for high energy applications, such as electric vehicles and grid storage.
• Capacity retention (CR, cyclability/stability): This is the number of times a battery can be discharged/charged at both fixed and various current densities. For practical application in industries, a good battery should retain at least 80% of its initial SC after multiple cycles (charged/discharged).
• Energy density (ED): This is the overall energy stored in a battery, which is calculated by multiplying the working voltage (V) and reversible capacity (C), as given in
Here, i indicates the current, and t represents time. Electrodes having higher potential (voltage), which is a function of the amount of alkali metal ion intercalated, and layered structure with various compositions facilitate higher ED. The ED can be expressed as gravimetric (Wh/kg) or volumetric (Wh/L).
• Power density: This is the optimal power delivered by a battery with regard to its electrode mass or volume. It shows how fast work is done with the availed energy in the battery, as given in
Here, i, V, and Rint denote the current, voltage, and internal resistance, respectively. The Rint is a function of ionic conductivity of the alkali metal ions, electrical conductivity of electrodes and reaction kinetics during cycling. The power density can be reported per unit mass [i.e., gravimetric power density (W/kg)] or unit volume [i.e., volumetric power density (W/L)]. Surface reactivity, conductivity, surface morphology, electrochemical, chemical and structural stability of electrodes are the contributory factors for optimized power density.
• Coulombic efficiency (CE): This is the ratio of energy withdrawn from a battery during discharge and energy restored during charging, as given in
A good battery should have close to 100% CE.
• Voltage/voltaic efficiency: This is the ratio of mean discharging voltage to charging voltage. High active surface area and electrical conductivity of electrodes enable increased voltage efficiency.
• Energy efficiency: This is the ratio of energy densities during discharge and charge process.
The cyclability, ED, power density, and rate capability are dependent on the ionic and electrical conductivities of the electrodes. Hence, a battery must maintain low internal resistance and higher conductivities to obtain excellent performance. Meanwhile, low molar mass electrodes enable higher capacity and rate capability.
ENGINEERING METHODS FOR POROUS MULTIMETALLIC ALLOY-BASED ANODES
The PMMA anodes are controllably synthesized by various methods, including template-based, hydrothermal/solvothermal, chemical reduction, electrodeposition, sol-gel, co-precipitations and electrospinning.
Template-based methods
The PMMAs are mostly and best prepared via the template-based methods, i.e., hard-templates or soft-templates[3,27]. The hard-template is appropriate to tailor the preparation of well-ordered PMMAs, initiated by the infiltration of metal precursors into the template by co-deposition or in-situ deposition or
Hydrothermal/solvothermal method
The synthesis of PMMAs using hydrothermal methods involves the mixing of aqueous metal precursors (i.e., acetate, chloride, nitrate and hydroxide dissolution in water), where a pH modular/additive (i.e., KI, KBr or KOH) serves as a catalyst when magnetically stirred or ultrasonicated, and surfactants (i.e., polyvinylpyrrolidone (PVP), pluronic F127) as structure-directing agents[33,34]. The aqueous solution is sealed in an autoclave and heated at elevated temperature for some time (2-24 h), and then the anticipated product is washed by centrifugation for several cycles and dried in a vacuum oven, prior to calcination [Figure 4A]. Solvothermal method is similar to hydrothermal method, but it uses non-aqueous solvents, such as dimethylformamide, ethylene glycol, isopropanol, ethanol, and butanol, which may act as a structure-directing agent. The fabrication methods of the PMMAs via hydrothermal/solvothermal method entails initial nucleation and subsequently direct attachment growth; meanwhile, the surfactant is adsorbed or interacts with the anion additives (Br-, I-, OH-) and sometimes in-situ etching-induced porous architectures[35-37]. The hydrothermal/solvothermal method is easy, effective and appropriate for the preparation of PMMAs, their oxides, chalcogenides, spinels, perovskites, high entropy alloys, and composites, with or without supports [i.e., carbon, nickel foam (NF)] in various architectures (nanospheres, nanocubes, nanosponges, nanocages, flower-like architectures, etc.).
Chemical reduction method
The chemical reduction method uses suitable reducing agents, such as NaBH4, ascorbic acid, PVP, etc.[38-40]. The formation mechanism of the PMMAs follows a burst nucleation of metal with high standard reduction potential, which eventually serve as seeds for other metal(s) atoms to form multiple crystallites with large surface-free energy; then, these crystallites aggregate and undergo continuous direct-attachment/diffusion-limited growth across the interface through Brownian motion [Figure 4B][41]. At rapid reduction rate, the as-formed nuclei grow along the single-crystalline lines to yield polyhedron shapes (i.e., cubes, tetrahedrons, and octahedrons) with Oh and Td symmetry. In contrast, slow reduction rate affords a five-fold twinned nuclei growth into polyhedrons with Ih symmetry (dodecahedrons and icosahedrons). Also, the structures of the PMMAs are a function of the adsorption model of the structure-directing agents[42], including ionic polymers [i.e., cetyltrimethylammonium bromide (CTAB) and cetyltrimethylammonium chloride (CTAC)] and non-ionic polymers (i.e., Pluronic F127,
Electrochemical deposition method
The preparation of PMMAs via electrochemical deposition is low-cost, rapid, single-step and low energy-intensive (i.e., low potential and room temperature) in the form of thin films, coatings and various shapes/compositions. The fabrication process involves reduction/oxidation of metal ions in an electrolyte (acid, neutral or alkaline), which are deposited on the conductive materials when certain potential is applied, utilizing a 3-electrode configuration, comprising the counter electrode (i.e., Pt wire and graphite), reference electrode (i.e., Ag/AgCl, and SCE) and working electrode (i.e., glassy carbon, carbon paper, carbon cloth sheet) [Figure 5A]. The modulation of the shapes and compositions of the PMMAs are functions of the electrolyte types, applied potential, cell set-up and deposition time. This method is easy and could be combined with other approaches for the synthesis of thin film; however, it requires a lot of precaution and skills for controlled morphologies and compositions.
Sol-gel method
Controlled texture, size and surface features of PMMAs in the form of aerogels, xerogels or cryogels are prepared via sol-gel method, which is cost-effective, easy to implement, and provides high quality and large yield[43,44]. The fabrication process of PMMAs through this method follows several steps: hydrolysis, condensation, aging & drying and crystallization [Figure 5B][45]. Hydrolysis involves the mixing of metal alkoxide precursors and solvent, where the nucleophilic oxygen of water attaches to the metal alkoxide to remove it from the formation of the alcohol functional group (M-OH). This is affected by several parameters [nature of R-group, alkoxy functional group steric hindrance, temperature, pH, aqueous or non-aqueous solvents, ratio of solvent to metal alkoxide, and the amounts of catalysts (acid or base)]. These parameters can sometimes make it difficult to control the synthesis and the structure of the obtained gel depending on the catalyst, owing to the rate of hydrolysis. The hydrolysis rate is often faster in alkaline media than in acidic media[46]. The condensation is the formation of a polymer network, composed of metal oxide linkages, due to the removal of solvent (water or alcohol) to form a sol. This step raises the solution’s viscosity during the polymer network growth and forms a porous structure within the gel-phase. Aging is the continuous polycondensation within the solution and precipitation of the gel network, which lowers porosity, but increases thickness of the colloidal nanoparticles. Drying is the evaporation during gelation, whose rate significantly affects the gel structure. The network structure of the gel may be affected by liquids removal: if the structure is retained, an aerogel (high pore volume and surface area) is formed; if it collapses, a xerogel (low pore volume and surface area) is formed; but it is removed at low temperature, a cryogel is formed. The drying of gel requires heat treatment to remove surface OH groups and compress to form crystalline PMMAs[47].
Electrospinning method
The electrospinning is a useful and worthwhile technology for the preparation of 1D nanofibrous materials for nanotechnological applications in biomedicine, fuel cells, energy storage, electronics, and sensors[48,49]. This method involves four main components: a high-voltage power supply, a syringe pump, a spinneret (i.e., a hypodermic needle with a blunt tip) and a conductor collector[50]. It utilizes an electrodynamic process to electrify liquid droplets of metal precursors to produce a jet that is then stretched and elongated to give PMMA nanofibers [Figure 5C]. The fabrication mechanism involves the extrusion of metal precursor liquid from a spinneret to afford a suspended droplet, due to the surface tension; then, the electrostatic repulsion of the surface charges of the same sign deforms the droplet into a Taylor cone, where a charged jet is ejected. The jet is extended initially in a straight line, then subjected to vigorous whipping motions, owing to bending instabilities, stretches to finer diameters, and solidifies rapidly, resulting in deposition of solid porous PMMA nanofibers on the conductor substrate[51,52].
ALKALI METAL-ION BATTERIES
Lithium-ion batteries
In the last decade, numerous studies have explored the use of MA anodes for rechargeable Lithium-ion batteries (LIBs). Nevertheless, research on PMMA anodes has been relatively limited. PMMA anodes are considered preferable because of their superior specific capacities relative to carbon anodes. However, during cycling, these anodes are conspired with large capacity irreversibility at the 1st cycle and meager cycling performance, owing to volume variation. These hiccups are being surmounted by tailoring the nanostructures of the alloys and/or supporting them with porous conducting materials [i.e., carbon, NF, and copper foam (CF)]. Hence, the PMMAs for LIB anodes are classified according to the main active atoms/elements in the alloy compositions: (i) tin (Sn)-based alloy anodes; (ii) cobalt (Co)-based alloy anodes; (iii) manganese (Mn)-based alloy anodes; and (iv) other alloy anodes-containing active elements/atoms (i.e., Fe, V, Ni, Zn, Ti, Si).
Porous multimetallic tin (Sn)-based alloy anodes
The high theoretical capacity, cheap nature and suitable working conditions of Sn have positioned it as a capable anode for imminent LIBs. The practical application of Sn anodes is limited by the huge volume change during Li+-ions lithiation and delithiation, which results in massive irreversible capacity loss and poor life cycle. Alloying of Sn to other metals/composites with carbon having tailored morphologies and augmented synergism is desirable for boosted electrochemical properties of energy storage systems[53,54], particularly improving the anodic performance of porous multimetallic Sn-based alloy anodes in LIBs[55]. For example, the construction of Sn on a porous Cu foam substrate (Sn-Cu6Sn5/Cu foam) was achieved by electrodeposition. This configuration exhibited superior performance as a LIB anode at higher initial SC (821.5 mAh/g@0.1 A/g) compared to Sn/Cu foil (655.5 mAh/g); besides, CR = 58% after 50 cycles, due to the formation of porous Cu6Sn5 interface that eschewed volume expansion and pulverization/
Sb-doped SnO2 nanoparticles grown on N-doped graphene-carbon nanotubes hydrogels (ATO/N-GCH) were prepared by a hydrothermal method and then freeze-drying to obtain the aerogel counterpart
Figure 6. (A) Hydrothermal synthesis, (B) SEM, (C) rate capability and (D and E) cyclic performance of ATO/N-GCA, ATO/G and SnO2/N-GCA. This figure is quoted with permission from Ref.[58] Copyright 2017 Elsevier. (F) TEM and (G) galvanostatic charge/discharge curves of CoSn2, Co + Sn and Sn. (H) TEM, (I and J) galvanostatic charge/discharge and (K) cycling stability of
Porous multimetallic cobalt (Co)-based alloy anodes
In recent times, Co-based materials (i.e., oxides, phosphides, nitrides, sulfides and alloys) are receiving huge attention as impressive anodes and electrocatalysts for LIBs and other electrochemical applications, owing to their excellent electrochemistry[61]. Despite this, the anode is still affected with volume variations triggered by the lithiation/delithiation that lead to the pulverization and detachment, thereby causing capacity fading[62]. This issue could be circumvented by rational nanoarchitecture design and alloying with or without conductive materials in facilitating the rapid Li+ ions diffusion and enduring strain induced by the volume enlargement, thereby augmenting the electrochemical activity and cyclability[63,64]. For instance, NiCo2O4 anchored reduced graphene oxide (rGO/NiCo2O4) was systematically synthesized by citrate-assisted hydrothermal/annealing techniques, which enabled interconnected nanosheets on hierarchical porous wrinkled film to give high initial SC (1,697.9 mAh/g@0.2 A/g), notable CR and rate capability[65]. The high capacity was traced to contributory effects of the hierarchical porous support that optimized electrode/electrolyte interface for rapid Li+-ions flux/diffusion. Appropriate hierarchical nanocomposites are excellent materials for outstanding reversible Li+-ion storage. Hard-template annealing was utilized to construct CuO/CoO core/shell arrays on Cu foams [Figure 7A], which gave very good LIB anode performance at initial SC (1,216 mAh/g@0.1 A/g). Then, this capacity increased to 1,364 mAh/g after
Figure 7. (A) Hard-template fabrication method, (B) charge/discharge curves, (C) cyclic performance at 0.1 A/g and (D) SEM of tubular CuO/CoO core/shell arrays on Cu foam. This figure is quoted with permission from Ref.[66] Copyright 2015 Elsevier. (E) Field emission scanning electron microscopy (FESEM), (F) charge/discharge curves for 1st, 3rd, 50th and 100th cycles, and (G) cycling performance of NiCo2O4/NiO-HD at 0.2 A/g. This figure is quoted with permission from Ref.[32] Copyright 2015 the Royal Society of Chemistry. (H) charge/discharge curves of 1st, 2nd and 100th cycles at 1.0 A/g, and (I) FESEM of NiCo2O4. Reproduced with permission from Ref.[68] Copyright 2016 Elsevier.
Various self-supported hierarchical porous multimetallic oxides (ZnCo2O4, NiCo2O4, CuxCo3-xO4,
Mesoporous FeCo2O4 was synthesized via sol-gel/calcination, which favorably utilized its multicomponent effect, octahedral architectural stability and meso-porosity for superb LIB anode performance at increased initial SC (2,436 mAh/g@0.2 A/g) and CR (95.2%)[69]. The 3D Zn-Ni-Co oxide enclosed graphene nanosheets (ZNCO/GNS) were prepared by chemical reduction/hard-template annealing [Figure 8A][70]. This was based on the wrapping of ZNCO microspheres within the GNS to form core/shell architectures, evidenced by SEM [Figure 8B], that prevented agglomeration of active ZNCO, volume variations, and boosted electrons/Li+-ions mobility, which resulted in good Li+-ion storage at initial SC
Figure 8. (A) Chemical reduction/annealing synthesis, (B) SEM, and (C) charge/discharge curves at 0.1 A/g for 1st-3rd and 50th cycles of ZNCO/GNS. (D) TEM, (E) Charge-discharge, cycling life and Columbic efficiency, and (F) rate capability of Te@ZnCo2O4 nanofibers. This figure is quoted with permission from Refs.[70,72] Copyright 2015 Wiley-VCH. (G) TEM, (H) 1st-3rd charge/discharge profiles and (I) Cyclic performance at 0.5 A/g of CoMoO4-CoO/S@rGO and analogs. This figure is quoted with permission from Ref.[75] Copyright 2022 the Royal Society of Chemistry.
Hierarchical porous CoMoO4-CoO/S@rGO, prepared from MOF-template via
Porous multimetallic manganese (Mn)-based alloy anodes
Manganese oxide (MnxOy)-based materials possess inimitable outer valence electrons, physical and chemical features that makes it amazing LIB anodes[77,78]. Besides their low cost, environmental benignity, ease of preparation and high theoretical capacity (1,232 mAh/g). Nevertheless, large volume change during cycling and low electrical conductivity of MnxOy anodes are the drawbacks that lead to low rate capability and cycling durability[79,80]. These drawbacks are solved by deliberately fabricating porous nanostructured MnxOy alloy and/or carbon composites with controlled sizes and morphologies[81-83]. For instance, solvothermal/annealing methods were utilized to fabricate porous quasi-mesocrystal ZnMn2O4, following oriented attachment and Ostwald ripening mechanisms for the formation of distinct twin-microsphere structures, 3D hierarchical porosity and Li alloying reaction, beneficially utilized for good LIB anodes at initial SC (1,100 mAh/g@0.5 A/g), CR (77.8%), rate capability and long-standing cycling durability[84]. However, porous ZnMnO3 spherulite, prepared by co-precipitation/annealing, had inner atomic synergism of Zn-O and Mn-O that decreased its LIB anode performance at lower initial SC (1,294 mAh/g@0.5 A/g), but improved CR (67.9%) than ZnO-MnO2 equimolar power mixture (1,333 mAh/g; 60.4%) and nano-sized MnO2 (1,553 mAh/g)[85]. This study showed that the synergistic effect of inner atoms of Zn-O and MnO2 could lead to significant initial capacity fading of ZnMnO3, but improved rate capability and cyclability. The construction of 3D porous CoMn2O4 on NF (CoMn2O4/NF) was achieved by hard-template/annealing, with greater initial discharge SC (1,833 mAh/g@0.4 A/g) and CR (74.5%), compared to un-annealed
Controlled synthesis of multi-shelled nanostructured Ni-Co-Mn oxides by self-templated hydrothermal/annealing method [Figure 9A] that enabled various architectures, including quadruple-shelled hollow microspheres (QS-HS), triple-shelled hollow microspheres (TS-HS) and double-shell hollow microspheres (DS-HS), proved by TEM [Figure 9B-D][88]. The QS-HS afforded abundant interfacial electrode-electrolyte interaction, numerous active sites, and favorable Li+-ions transfer and volume accommodation to deliver high LIB anodes at initial SC (1,761.8 mAh/g@0.2 A/g), rate capability (1,073.6 mAh/g@5 A/g), gravimetric energy densities (1,261.7 Wh/kg) and good cyclability, compared with TS-HS and DS-HS
Figure 9. (A) Controlled hydrothermal/annealing synthesis, (B-D) TEM, (E) specific capacity and coulombic efficiency, (F) rate performance, and (G) gravimetric energy densities of Ni-Co-Mn oxides QS-HS, TS-HS and DS-HS. This figure is quoted with permission from Ref.[88] Copyright 2017 the American Chemical Society. (H) capacity retention at 0.1 A/g after 200 cycles and coulombic efficiency, (I) rate capacity at various current densities, and (J) SEM of np-Ni@NiO/MnO/NF. (K-M) charge/discharge profiles of
A series of NiMn hydroxides modified with [p-phthalic acid (PTA)] and/or activated carbon (AC) [i.e., Ni(OH)2/Mn(OH)2 (MN-1), NiO/Mn3O4 (MN-2), MN-1/P, Ni(OH)2/Mn(OH)2/AC) (MN-1/C), Ni(OH)2/Mn(OH)2/PTA/AC (MN-1/PC), NiO/Mn3O4/PTA (MN-2/P), NiO/Mn3O4/AC (MN-2/C), and NiO/Mn3O4/PTA/AC (MN-2/PC)] were synthesized by in-situ electro-conversion/calcination and studied as LIB anodes[91]. The MN-1/P showed higher initial discharge SC (1,554 mAh/g) than MN-1/PC (1,518 mAh/g), MN-1 (1,386 mAh/g), Mn-2/P (1,128 mAh/g) and MN-2 (1,068 mAh/g) [Figure 9K-M], besides its CR (100%), owing to the successful wrapping of NiMn nanosheets by the PTA, dissolved by alkaline medium
Other porous multimetallic-based alloy anodes
Other PMMAs with main active atoms/elements, including Fe, Mo, Ni, Si, Ti, V, and Zn, and their composites have been demonstrated as inspiring anodes for LIBs because of their excellent features (i.e., chemical stability, high ionic/electrical conductivity and lithiophilicity) that facilitate facile Li+-ions mobility and inhibit Li dendrite growth and volume expansion[93-95]. For example, the fabrication of TiO2 on Fe3O4/CNTs (TFCs) was achieved by hydrothermal/calcination [Figure 10A], based on the bottom-up assembly mechanism with TiO2, carbon nanotubes (CNTs) and FexOy as buffer material, conducting network/effective anchorage and active nanoteeth, respectively, proved by HRTEM [Figure 10B][96]. The TFCs exhibited higher initial SC (1,589 mAh/g@0.5 A/g) than FexOy/CNTs (FCs, 771 mAh/g) and CNTs
Figure 10. (A) Hydrothermal/calcination, (B) HRTEM and (C-E) galvanostatic charge/discharge curves of TFCs, FCs and CNTs at
The 3D ordered hierarchical porous amorphous vanadium and molybdenum oxides
Inverse spinel-type NiFe2O4 anchored porous carbon fibers (NFO@C), synthesized by electrospinning/calcination, delivered superior LIB anodes to NFO, which was attributable to exclusive mesoporous fibrous structures, high surface area, fast Li+-ions diffusion, and improved electrical conductivity that eventually eschewed pulverization and agglomeration[100]. The physicochemical merits of mesoporous fibrous supports could significantly boost the electrochemical storage capacity of PMMA oxides. Facile evaporation-solidification/annealing was adopted to prepare vacancy-enriched Ni3ZnC0.7 nanohybrids, with much focus on stoichiometric and non-stoichiometric ratios, where the latter gave higher Li+-ion storage capacity
Sodium-ion batteries
The abundance and low cost of sodium (Na) have spurred the Sodium-ion batteries (SIBs) among the most auspicious energy storage technology[107,108]. The practicability of the SIBs lies largely on robust fabrication of electrodes with high SC and long cyclability. The development of suitable anodes is more thought-provoking; in this regard, the PMMA anodes have great potential for SIBs with high energy because of their high gravimetric/volumetric SCs and easy Na+-ion insertion/ejection[109-111]. The utilization of PMMA anodes for viable SIBs is conspired with enormous capacity fading, irreversible capacity, rate capability, low coulombic efficiency and poor cycling durability that lead to volume expansion[112,113]. This hitch is solved by incorporating conductive materials with their unique physicochemical merits that could accelerate the reaction kinetics and alleviate capacity fading[114-116]. Hence, the modulated physicochemical features of the PMMA anodes result in improved Na+-ion storage and stability. For example, a hydrothermal method was exploited to prepare Sb-doped SnO2 nanoparticles grown on N-doped graphene-carbon nanotubes aerogels (ATO/N-GCA) with spherical Sb-SnO2 on sheet-like nanotubes, high porosity and electrical conductivity that delivered increased initial SC (409 mAh/g@0.1 C), CR (74%) and lower charge transfer resistance
Figure 11. (A) Electrospinning/annealing synthesis, (B) SEM, (C) TEM, (D) rate performance, and (E) ultra-long cyclability of SnSb/N-PCNWs. This figure is quoted with permission from Ref.[117] Copyright 2019 Elsevier. (F) SEM, (G) charge/discharge curves at 0.2 A/g for 1st-5th cycles, (H) rate capability, and (I) long-term cyclability and coulombic efficiency of ZMS@FCs. This figure is quoted with permission from Ref.[120] Copyright 2022 Wiley-VCH.
The construction of bimetallic metals core-shells (SnSe2/CoSe2@C), achieved by selenization/carbonization and tested as SIB anodes, showed higher initial SC (455 mAh/g@0.1 A/g), rate capability and CR (77.6%) than SnSe2/CoSe2 (300.31 mAh/g) and mechanically-mixed SnSe2-CoSe2 (29 mAh/g)[118]. This study proved that the addition of carbon to SnSe2/CoSe2 improved the conductivity, retained the architectural design, and prevented capacity fading. Also, poor chemical bonding in multiple metal selenides results in capacity fading and structural damage. A self-adaptive Sn-Bi@C composite, prepared by hydrothermal/annealing, had Bi-core with Sn shell that exhibited uniform stress circulation, alleviated structural strain, improved electrical conductivity and Na+-ions diffusion kinetics, advantageous to deliver high SC
Porous carbon-supported Zn-Mn selenides (ZMS@FC) were constructed by solvothermal/calcination, and examined as SIB anodes[120]. The ZMS@FC was endowed with flower-like metal selenides on flakes support, proved by SEM [Figure 11F] that enabled outstanding initial SC (494.8 mAh/g@0.2 A/g), great rate capability, and CR (74.7%) [Figure 11G-I], compared to MnSe@C (230.9 mAh/g) and ZnSe@C
Also, ultrathin Cu4Mo6Se8 grown on carbon skeleton (CMSe/C), achieved by chelation/annealing and tested as SIB anode, exhibited exceptional rate capability, initial discharge SC (514 mAh/g@0.5 A/g) and CR (107.7%), relative to MoSe2 (203 mAh/g), ascribable to its copious void space, homogeneous morphology, ultrathin nanosheet arrays and improved conductivity[121]. Moreover, a series of rod-shaped
Figure 12. (A) In-situ XRD during 1st and 2nd charge-discharge cycles, (B) atomic structure model, (C) density of states, (D) Na+ ions adsorption energies, and (E) diffusion energy barriers of Na+ ions of NiCoSe2 and CoSe. (F) Top/side view of most stable adsorption sites, (G) charge density differences of each structure and (H) simulated density of states of VS4/SnS2/MXene, VS4/MXene and SnS2/MXene. This figure is adapted from Ref.[123,124] Copyright permission 2024 Elsevier.
VS4/SnS2 confined MXene (VS4/SnS2@MXene), prepared by hydrothermal/etching/self-assembly, was demonstrated as an anode for Na+-ion storage, proven experimentally and theoretically[124]. The
Potassium-ion batteries
The Potassium-ion batteries (PIBs) are recently drawing so much interest as an alternative to LIBs and SIBs because of their chemical and economic merits[126]. Their chemical quality is traced to low potential of K+/K (-2.88 V vs. SHE) in carbonate ester electrolytes[127], which infers a high ED with K+-ions as the charge carrier and a low risk of potassium (K) plating[128]. Also, K+-ions boost the ionic conductivity and rapid diffusion rate. The economic advantage is due to the essentially limitless and worldwide distribution of K reserve at low cost [i.e., potassium carbonate (K2CO3, $1,281/tonne) compared to lithium carbonate
Figure 13. (A) Melt-spinning/chemical dealloying synthesis, (B) SEM, (C) cycling and (D) rate performance for 50 cycles of NPCuBi and NPBi. This figure is quoted with permission from Refs.[131] Copyright 2020 Wiley-VCH. (E) SEM, (F) 1st-3rd charge/discharge curves and (G) cycling performance at 0.1 A/g of FeCo@PAZ-C. (H) galvanostatic charge/discharge curves for 1st-3rd, 50th and 100th at 0.1 A/g, (I) rate capability, and (J) long-term cyclability of ZCS@NC@C@rGO, ZCS@NC@C and ZCS@NC. This figure is quoted with permission from Ref.[132,133] Copyright 2024, 2023 Elsevier.
Flower-like porous carbon-embedded Zn-Mn selenides (ZMS@FC), made via solvothermal/calcination, demonstrated exceptional performance as PIB anodes. They exhibited a great initial SC
Germanium-based alloy anodes (GeV4S8) were rationally fabricated by physical mixing/annealing to incorporate Ge atoms into V-S framework with loosely packed and metallic media that facilitated rapid
Figure 14. (A) Density of state (DOS), (B) K+-ions adsorption energies, (C) K+-ions diffusion energy barriers of GeV4S8, V5S8, GeS2 and Ge, and (D-F) charge density difference maps with side views of K/GeV4S8, K/V5S8, and K/GeS2. (G and H) Structural models and corresponding charge density differences for the adsorption of K+ ions on {110} and {111} facets, (I and J) energy and diffusion barrier of BiFeO3. This figure is reproduced with permission from Ref.[134,137] Copyright 2024 Wiley-VCH.
Solvothermal/calcination methods were explored for the construction of FeS2/MoS2 on N-doped carbon (FMS@NC) with spherical flower-like morphology, synergistic effect to alleviate volume expansion and conductivity to give good PIB anodic performance at SC (585.0 mAh/g@0.05 A/g), CR (93.9%) and cycling capability[138]. Systematic construction of heterostructure alloys with conductive materials is viable for high-performance PIB anodes and stability. Although PMMAs are promising anodes for PIBs, only very few studies have been reported as far as we found. The in-situ construction of PMMAs with Cu substrate revealed the best K+-ion storage capacity and long-term durability. Thus, there is a need for this area of research to be extensively exploited and optimized.
RAMIBs RECYCLING AND CIRCULAR ECONOMY
The recycling of RAMIBs is pivotal for developing a sustainable circular economy, mitigating environmental hazards, and promoting a sustainable future, but it remains a significant challenge[139]. This is attributed to the complex chemistry and the need for specific processes to recover rare metals (i.e., Li, Ni, Al, and Co)[140,141], which offer an alternative way to decrease the consumption of these metals besides reducing wastes. Life cycle assessments, including all steps from the extraction of anode materials to end-of-life treatment, are needed to estimate the environmental footprint of RAMIBs recycling processes to determine areas that require prompt improvements to enhance sustainability[142]. For example, unveiling the trade-offs between the environmental benefits of material recycling and the encumbrances of collection and treatment can direct the process towards more eco-friendly practices[142]. Also, the rational design of advanced recycling technologies and life cycle thinking in managing RAMIBs is vital to attain a circular economy[143-145].
Considering the widespread applications of the RAMIB devices for portable electronics, electric vehicles and renewable energy systems, their environmental impact assessment, including (i) raw materials mining and resources extraction; (ii) production stage; (iii) usage phase; and (iv) end of life and recycling, is a crucial area of concern. The extraction of the anode materials (Li, Na, K) from mineral ores and brine deposits could result in water depletion, contamination and environmental disorders. For instance, the mining of cobalt, manganese and nickel results in total emission of 8.1 GtCO2 equivalent in 2050[146], which causes considerable environmental degradation, such as soil and water pollution, deforestation and harmful consequences on the immediate localities. Also, the manufacturing of RAMIBs (LIBs, SIBs and PIBs) requires high energy consumption, which contributes to greenhouse gas emissions and carbon footprint, which are highly concentrated in three countries: China (45%), Indonesia (13%) and Australia (9%)[146]. Besides, the use of various chemicals during the production of RAMIBs could pose the risks of hazardous chemical spillage and pollution if not optimally managed. However, during their use for electric vehicles and renewable energy systems, the RAMIBs are highly effective in reducing the overall greenhouse gas emissions, as no pollutants are emitted, making them cleaner than fossil fuel-based systems[147]. After complete usage (i.e., end of life stage), the RAMIBs are either disposed of or recycled: inappropriate disposal of the RAMIBs could result in water and soil pollution with toxic materials such as heavy metals and pose fire hazards with environmental and safety concerns; meanwhile recycling of valuable materials in spent RAMIBs is complex and not economically viable[148]. Hence, the current recycling rate of RAMIBs is extremely low. Significant enhancement of the technology and infrastructures is a necessity for sustainability. Although there are environmental concerns associated with the production of RAMIBs, the devices still significantly reduce the emissions of greenhouse gas by 38% in 2050[146], particularly for transportation and renewable energy systems. Their sustainability efforts are directed toward improving the RAMIB technology with much focus on reducing total reliance on essential raw materials, advancing recyclability and devising more mining sustainable and manufacturing processes, implementing stricter policy regulations, and promoting sustainable practices in the RAMIB lifecycle.
The policymakers who can promote a circular economy for RAMIBs recycling need proper and clear policy measures to drive sustainable practices and enhance resource efficiency through various processes[149]. This includes material management (i.e., supply chain traceability, material transport/storage/recovery), awareness (i.e., stakeholder engagement and educational campaigns), incentivizing circular business models (i.e., financial incentives, and market signals), and worldwide cooperation[149].
The market size and growth of RAMIBs recycling are noteworthy, indicating the growing importance of sustainability in the battery industry. Mainly, the global market size of battery recycling is nearly 1.83 billion USD in 2023 and is expected to grow by about 37.6% from 2024 to 2030[150]. The market share of RAMIBs is about 75% of the global battery market share[150]. Some successful examples of battery recycling occurred in China, which leads the global battery recycling with over 500,000 metric tons, compared with the USA and Europe, both of which had a recycling capacity of nearly 200,000 metric tons[151,152]. The leading technologies for recycling batteries in China include recycling electric-vehicle batteries and utilization of spent vehicle batteries, which could reach a market size of (3.59 billion Euros) by 2025, and reusing electric-vehicle batteries operate below 70%-80% capacity after 4-6 years in slower light electric vehicles and stationary energy storage application[153,154]. In Europe, the recycling market for batteries will reach 130GWh by 2030, which is about 700 kilotons of recycling capacity required, and will increase by three times by 2040[152].
CONCLUSION AND OUTLOOK
In brief, this article reviews the rational design of PMMA (i.e., Sn, Mn, Mo, Co, V, and Fe) anodes using various approaches (i.e., template-based, hydrothermal/solvothermal, chemical reduction, electrochemical deposition, sol-gel, dealloying and electrospinning) for the RAMIBs (i.e., LIBs, SIBs, and PIBs). This is besides the fundamental aspects (i.e., mechanisms and calculations), along with the deep discussions on the key descriptive factors of RAMIBs (i.e., lifespans, ED, structural changes, and mechanisms) corroborated with the summary tables, schemes, and figures. The results discussed above warranted that the performance of PMMAs in LIBs is significantly superior to SIBs and PIBs, respectively, as shown in the higher SC, CR and cycling durability [Tables 1-3]. Meanwhile, the most active anodes in LIBs were FeCo2O4 octahedral
Figure 15. Summary of the outlook highlighting the recent advances, challenges and solutions proffered for future research.
Comparative porous transition metal alloy anodes for LIBs
Anodes | Synthesis method | Morphology | Electrolyte | Capacity/mAh/g@current density (A/g) | Cycles | Capacity retention/% | Ref. |
Sn/Cu foam | Electrodeposition | Grape-like foam | 1 M LiPF6 in EC/DMC (1:1) | 821.5@0.1 | 50 | 58.0 | [56] |
Sn-Sb/Cu | Template-like electrodeposition/annealing | Sponge-like foam | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 651.9@0.5 | 30 | 82.9 | [57] |
ATO/N-GCA | Hydrothermal | Spherical metals on sheet-like nanotube | 1 M LiPF6 in EC/DMC (1:1) | 942@1.0 | 1,000 | 72.7 | [58] |
CoSn2 | Chemical reduction | Tetragonal | 1 M LiPF6 in EC/DMC (1:1)/FEC(3%) | 650@2.0 | 5,000 | - | [59] |
CoSn2Ox | Chemical reduction/ball milling | Spherical | 1 M LiPF6 in EC/DMC (1:1) | 525@2.0 | 1,500 | 92.0 | [60] |
rGO/NiCo2O4 | Hydrothermal/annealing | Nanosheets on porous wrinkled films | 1 M LiPF6 in EC/DMC (1:1) | 1,697.9@0.2 | 50 | 61.0 | [65] |
CuO/CoO core/shell array on Cu foam | Hard-template and annealing | Nanosheets on nanotubes | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 1,216@0.1 | 1,000 | 93.8 | [66] |
ZnCo2O4/NF | Hard template/calcination | Flower-like nanosheets | 1 M LiPF6 in EC/DMC/EMC (1:1:1) | 1,544@0.1 | 100 | 82.0 | [67] |
NiCo2O4/NiO-HD | Solvothermal/calcination | Hollow dodecahedron | 1 M LiPF6 in EC/DMC (1:1) | 1,622@0.2 | 100 | 63.5 | [32] |
NiCo2O4 | Solvothermal/annealing | Rose flower-like nanosheets | 1 M LiPF6 in EC/DEC (1:1) | 1,282@1.0 | 100 | 94.0 | [68] |
FeCo2O4 | Sol-gel/annealing | Octahedral | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 2,436@0.2 | 200 | 95.2 | [69] |
ZNCO/GNS | Hard-template/calcination/chemicalreduction | Microspheres core/nanosheet shell | 1 M LiPF6 in EC/DMC/EMC (1:1:1) | 1,429@0.1 | 50 | 95.0 | [70] |
ZnCo2O4 film/NF | Hard-template/annealing | 3D porous networks | 1 M LiPF6 in EC/DEC (1:1) | 1,726@0.4 | 100 | 63.0 | [71] |
Te@ZnCo2O4 | Templated/annealing | Core-shell nanofibers | 1 M LiPF6 in EC/DEC (1:1) | 1,364@0.1 | 100 | ~100 | [72] |
Zn-Co-S@NS-CP | Template pyrolysis | Hollow polyhedron | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 1,298.1@0.57 | 400 | - | [73] |
CoO-ZnO@NC-450 | Hydrothermal/annealing | Spherical nanoparticles | 1 M LiPF6 in EC/DMC (1:1) | 975.2@0.2 | 100 | 81.6 | [74] |
CoMoO4-CoO/S@rGO | Templated hydrothermal/Annealing | Spherical metals on polyhedron sheets | 1 M LiPF6 in EC/DEC (1:1)/FEC (5 wt.%) | 1,672@0.5 | 150 | 62.1 | [75] |
CoO/Cu | Hard-template/pulsed electrodeposition | Nanowire core/spherical shell | 1 M LiPF6 in EC/DEC (1:1) | 1,215@0.5 | 250 | 68.1 | [76] |
ZnMn2O4 | solvothermal/annealing | Twin-microspheres | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 1,106@0.5 | 130 | 77.8 | [84] |
ZnMnO3 | Co-precipitation/annealing | Spherulites | 1 M LiPF6 | 1,294@0.5 | 150 | 67.9 | [85] |
CoMn2O4/NF | Hard-template/annealing | 3D reticular-like | 1 M LiPF6 in EC/DMC (1:1) | 1,833@0.4 | 100 | 74.5 | [86] |
ZnMn2O4/NF | Hard-template/annealing | 3D reticular-like | 1 M LiPF6 in EC/DMC (1:1) | 2,211@0.4 | 100 | 94 | [87] |
Ni-Co-Mn oxide QS-HS | Self-templated hydrothermal/annealing | Hollow microspheres | 1 M LiPF6 in EC/DMC (1:1) | 1,761.8@0.2 | 250 | 79.5 | [88] |
np-Ni@NiO/MnO/NF | Induction melting/dealloying/annealing | Coarsened grains | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 960@0.1 | 200 | 105 | [89] |
ZMO-G500 | Solvothermal/calcination | Spherical metal on nanosheets | 1 M LiPF6 in EC/DMC (1:1) | 1,100@0.1 | 200 | 99.95% | [90] |
MN-1/P | In-situ electroconversion/calcination | Flower-like nanosheets | 1 M LiPF6 in EC/DMC (1:1) | 1,554@0.1 | 1,000 | ~100 | [91] |
Cu/Mn3O4@SC | Template annealing | Microspheres | 1 M LiPF6 in EC/DEC (1:1) | 766.2@0.2 | 1,000 | 66.3 | [92] |
TFCs | Hydrothermal/calcination | Spherical metals on nanotubes | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 1,589@0.5 | 450 | 58.0 | [96] |
3D-OHP-a-VOx/MoOy | Freeze-dried/annealing | (3D) ordered hierarchical porous sheets | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 1,705@0.1 | 50 | 82.1 | [97] |
Zn3V2O8 | Solvothermal/annealing | Hollow nanocages | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 1,906@0.1 | 200 | 70.7 | [98] |
CoFe2O4 | Melt spinning/dealloying | Nanoplates | 1 M LiPF6 in EC/DMC (1:1) | 2,134.7@0.1 | 200 | 99.1 | [99] |
NFO@C | Electrospinning/annealing | Spherical metals on fibers | 1 M LiPF6 in EC/DMC/DEC (1:1:1) | 1,030@0.1 | 100 | 72.3 | [100] |
Ni3ZnC0.7 | Evaporation-solidification/annealing | Irregular spheres | 1 M LiPF6 in EC/DEC (1:1) | 1,100@0.05 | 1,000 | 100 | [101] |
NiSix/Si/C | Ball milling/annealing | Irregular spheres | 1 M LiPF6 in EC/EMC (1:1) | 1,060@0.1 | 200 | 72 | [102] |
NiFe-NiFe2O4/rGO-8 | Chemical reduction/annealing | Spherical metals on wrinkled sheets | 1 M LiPF6 in EC/DMC (1:1) | 1,362@0.1 | 130 | 79.4 | [103] |
NiO/Co3O4/Fe3O4 | Precipitation/calcination | Hollow nanocages | 1 M LiPF6 in EC/DMC/EMC (1:1:1) | 1,052@0.5 | 400 | 85.7 | [104] |
NiCo2O4@Fe2O3 | Hydrothermal/calcination | Flower-like porous channels | 1 M LiPF6 in EC/DEC (1:1) | 1,640@1.0 | 400 | 57.6 | [105] |
CoNiSe2/C-700 | Template annealing | Nanododecahedral | 1 M LiPF6 | 2,125.5@0.1 | 1,000 | 98 | [106] |
Comparative porous transition metal alloy anodes for SIBs
Anodes | Synthesis method | Morphology | Electrolytes | Capacity/mAh/g@current density (A/g) | Cycles | Capacity retentio/% | Ref. |
ATO/N-GCA | Hydrothermal | Spherical metals on sheet-like nanotubes | 1 M NaClO4 in EC/FEC (1:1)/FEC(5%) | 409@0.1 | 500 | 74.0 | [58] |
SnSb/N- PCNWs | Electrospinning/annealing | Spherical metals in nanofibers | 1 M NaClO4 in EC/DMC (1:1)/ | 400@0.05 | 10,000 | 100 | [117] |
SnSe2/CoSe2@C | Selenization/carbonization | Nanobox core/shell | 1 M NaPF6 in DIGLYME/DOL (1:1) | 455@0.1 | 500 | 77.6 | [118] |
Sn-Bi@C | Hydrothermal/annealing | nanosphere core-shell on layer support | 1 M NaPF6 in DME | 462@0.1 | 2,000 | 86.9 | [119] |
ZMS@FC | Solvothermal/calcination | Flower-like metals on porous flakes | 1 M SHCF in DEGDME | 494.8@0.2 | 1,000 | 74.7 | [120] |
Cu4Mo6Se8/C | Chelation/annealing | Ultrathin nanosheets polyhedral C skeleton | 1 M NaPF6 in DME | 514@0.5 | 2,400 | 107.7 | [121] |
CoSb2Se4@Sb2Se3 | Hydrothermal | Rod-shaped | 1 M NaCF3SO3 in DEGDME | 437.9@0.1 | 160 | 81.4 | [122] |
NiCoSe2@NSC | Chemical reduction/annealing | Spherical nanoparticles on porous nanosheets | 1 M NaPF6 in DME | 572.5@0.2 | 1,000 | 85.1 | [123] |
VS4/SnS2@MXene | Hydrothermal/etching/self-assembly | Nanosheets on layered MXene | 1 M NaPF6 in EC/DEC (1:1)/FEC(5%) | 2,508.4@0.1 | 200 | 82.8 | [124] |
Mn-NiCoS-10 | Solvothermal/annealing | Nanosheets | 1 M NaPF6 in DME | 662.6@0.3 | 1,000 | 85.4 | [125] |
Comparative porous multimetallic alloy anodes for PIBs
Anodes | Synthesis method | Morphology | Electrolytes | Capacity/mAh/g@current density (A/g) | Cycles | Capacity retention/% | Ref. |
NPCuBi | Melt-spinning/chemical dealloying | Nanospheres with bicontinuous open channel | 1 M KFSI in DME | 1,045@0.05 | 50 | > 93.0 | [131] |
ZMS@FC | Solvothermal/calcination | Flower-like metals on porous flakes | 1 M KPF6 in EC/DEC (1:1) | 494.8@0.2 | 1,000 | 74.7 | [120] |
FeCo@PAZ-C | Template/annealing | Spherical nanoparticle core on nanosphere shell | 1 M KPF6 | 1,232.0@0.1 | 500 | 95 | [132] |
ZnSe/Co0.85Se@NC@C@rGO | Co-precipitation/selenization/annealing | Microcubic core-shell | 1 M KFSI in DME | 422.2@0.1 | 2,000 | 55.5 | [133] |
GeV4S8 | Physical mixing/annealing | Nanocubes | 3 M KFSI in DME | 400.0@0.5 | 1,000 | 80.0 | [134] |
CoS2/SnS2@SC | Vulcanization annealing | Core-shell | 1.5 M KFSI in EC/DEC (1:1) | 1,026.0@0.1 | 1,500 | 76.0 | [135] |
(CoFe)Sex-rGO-CNT | Spray pyrolysis | Microspheres on hierarchical supports | 1 M KFSI in EC/DEC | 600.0@0.5 | 200 | ˃ 90.0 | [136] |
BiFeO3-MF | Hydrothermal | Flower-like spheres | 3 M KFSI in DME | 606.0@0.5 | 5,000 | 98.2 | [137] |
FMS@NC | Solvothermal/calcination | Spherical flower-like | 1.5 M KFSI in EC/DEC (1:1) | 585.0@0.05 | 3,000 | 93.9 | [138] |
• Many studies have addressed storage mechanisms, but PMMA anode failure mechanisms in RAMIBs are rarely reported, necessitating further investigation alongside ambiguous dendrite suppression and related chemical reactions[27]. This goal can be achieved by initially examining electrochemical merits of custom full cells, using experimental, theoretical, and in-situ studies to address problems in future electrode design[21].
• The preference for synergetic or strain effects in PMMAs, particularly with active alkali metals, remains unclear. Some reports indicate that synergism can enhance electrode durability and capacity[27]. Systematic studies, proper metal selection, and in-situ characterization techniques [TEM, XRD, X-ray photoelectron spectroscopy (XPS)] are essential to determine the preferred effect and monitor structural changes during the electrochemical reactions.
• Current fabrication methods face barriers such as hazardous chemicals, low yield, and complexity, necessitating new green and simple synthesis methods for atomic-level metal mixing and improved surface features and structure stability[155,156].
• To optimize the performance of PMMA anodes, factors such as electrolyte type, additives, pH, and composition must be considered. Exploring cost-effective and durable electrolytes is crucial, with ionic liquid electrolytes and solid electrolytes being the most promising options[157,158]. Electrochemical impedance spectroscopy (EIS) analysis can estimate electrolyte resistance and provide insights on interfacial interactions[159].
• Interfacial engineering enhances kinetics and chemical durability in alkali metal ions intercalation/deintercalation[25]. Hierarchical porous morphologies with a high surface area-to-volume ratio and support materials, including carbon, improve metal dispersion and stability against aggregation[15,160,161]. Support can be optimized via in-situ fabrication of PMMAs with strong metal-support interaction.
• Encapsulation of low-temperature liquid MAs inside porous MA under high pressure offers a self-healing anode or substrate, preventing dendritic growth of alkali metals. However, new assembly approaches and electrolytes are required[162].
DECLARATIONS
Authors’ contributions
Data collection and writing: Ipadeola AK
Conceptualization, writing, and supervision: Eid K
Review, supervision, and funding: Abdullah AM
Availability of data and materials
The materials are available upon request.
Financial support and sponsorship
This work was supported by (i) the Qatar National Research Fund (QNRF, a Division of the QRDI Council) through the Academic Research Grant (ARG) program, Grant# ARG01-0524-230330, and (ii) Qatar University through an International Research Collaboration Co-Fund program, Grant# IRCC-2023-157.
Conflicts of interest
All authors declared that there are no conflicts of interest.
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Copyright
© The Author(s) 2024.
Supplementary Materials
REFERENCES
1. Eid K, Gamal A, Abdullah AM. Graphitic carbon nitride-based nanostructures as emergent catalysts for carbon monoxide (CO) oxidation. Green Chem 2023;25:1276-310.
2. Yu Y, Lv Z, Liu Z, et al. Activation of Ga liquid catalyst with continuously exposed active sites for electrocatalytic C-N coupling. Angew Chem Int Ed 2024;136:e202402236.
3. Lu Q, Zhao X, Luque R, Eid K. Structure-activity relationship of tri-metallic Pt-based nanocatalysts for methanol oxidation reaction. Coordin Chem Rev 2023;493:215280.
4. Abdelgawad A, Salah B, Lu Q, et al. Template-free synthesis of M/g-C3N4 (M = Cu, Mn, and Fe) porous one-dimensional nanostructures for green hydrogen production. J Electroanal Chem 2023;938:117426.
5. Gu Y, Nie N, Liu J, et al. Enriching H2O through boron nitride as a support to promote hydrogen evolution from non-filtered seawater. EcoEnergy 2023;1:405-13.
6. Liu H, Li J, Arbiol J, Yang B, Tang P. Catalytic reactivity descriptors of metal-nitrogen-doped carbon catalysts for electrocatalysis. EcoEnergy 2023;1:154-85.
7. Yu W, Huang H, Qin Y, et al. The synergistic effect of pyrrolic-N and pyridinic-N with Pt under strong metal-support interaction to achieve high-performance alkaline hydrogen evolution. Adv Energy Mater 2022;12:2200110.
8. Lu Q, Li J, Eid K, et al. Facile one-step aqueous-phase synthesis of porous PtBi nanosponges for efficient electrochemical methanol oxidation with a high CO tolerance. J Electroanal Chem 2022;916:116361.
9. Ganesan V, Kim DH, Park CM. Robust CoP2-C hollow nanoboxes: superior anodes for Li- and Na-ion batteries. J Energy Stor 2024;79:110197.
10. Nam KH, Ganesan V, Kim DH, Jeong S, Jeon KJ, Park CM. SiSe2 for superior sulfide solid electrolytes and Li-ion batteries. ACS Appl Mater Interfaces 2024;16:643-54.
11. Tian H, Song A, Tian H, et al. Single-atom catalysts for high-energy rechargeable batteries. Chem Sci 2021;12:7656-76.
12. Peng Q, Rehman J, Eid K, et al. Vanadium Carbide (V4C3) MXene as an efficient anode for Li-ion and Na-ion batteries. Nanomaterials 2022;12:2825.
13. Ma F, Liu Y, Huang T, Du X, Lu Q, Kid K. Facile in situ polymerization synthesis of poly(ionic liquid)-based polymer electrolyte for high-performance solid-state batteries. Energy Convers Manag X 2024;22:100570.
14. Wu J, Chen X, Fan W, Li X, Mai Y, Chen Y. Rationally designed alloy phases for highly reversible alkali metal batteries. Energy Stor Mater 2022;48:223-43.
15. Ganesan V, Lee Y, Jung H, Park C. Porous polyhedral carbon matrix for high-performance Li/Na/K-ion battery anodes. Carbon Lett 2023;33:2189-98.
16. Kim Y, Kim K, Seo H, Lee S, Park C, Kim J. Surfactant-derived porous Sn2Nb2O7-graphene oxide composite as Li- and Na-ion storage materials. J Alloys Compd 2022;910:164943.
17. Kim T, Jeon K, Park C. Black P@MO (M = Mg, Al, or Ti) composites as superior Li-ion battery anodes. Chem Eng J 2021;424:130366.
18. Glushenkov A. Recent commentaries on the expected performance, advantages and applications of sodium-ion batteries. Energy Mater 2023;3:300010.
19. Tian H, Tian H, Yang W, et al. Stable hollow-structured silicon suboxide-based anodes toward high-performance lithium-ion batteries. Adv Funct Mater 2021;31:2101796.
20. Ma Z, Song A, Liu Z, et al. Nanoconfined expansion behavior of hollow MnS@Carbon anode with extended lithiation cyclic stability. Adv Funct Mater 2023;33:2301112.
21. Qi S, Deng J, Zhang W, Feng Y, Ma J. Recent advances in alloy-based anode materials for potassium ion batteries. Rare Met 2020;39:970-88.
22. Wang J, Wang Y, Zhang P, Zhang D, Ren X. Preparation and electrochemical properties of binary SixSb immiscible alloy for lithium ion batteries. J Alloy Compd 2014;610:308-14.
23. Li G, Guo S, Xiang B, et al. Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries. Energy Mater 2022;2:200020.
24. Shao Y, Jin Z, Li J, Meng Y, Huang X. Evaluation of the electrochemical and expansion performances of the Sn-Si/graphite composite electrode for the industrial use. Energy Mater 2022;2:200004.
25. Pathak AD, Chanda UK, Samanta K, Mandal A, Sahu KK, Pati S. Selective leaching of Al from hypereutectic Al-Si alloy to produce nano-porous silicon (NPS) anodes for lithium ion batteries. Electrochim Acta 2019;317:654-62.
26. Zhang M, Mou X, Zhou X, Wang J, Li H, Wang C. Metal compound-based heterostructures in anodes promote high capacity and fast reaction kinetic for lithium/sodium-ion storage: a review. ChemElectroChem 2024;11:e202300573.
27. Wang X, Tang S, Guo W, Fu Y, Manthiram A. Advances in multimetallic alloy-based anodes for alkali-ion and alkali-metal batteries. Materi Today 2021;50:259-75.
28. Pirayesh P, Tantratian K, Amirmaleki M, et al. From nanoalloy to nano-laminated interfaces for highly stable alkali-metal anodes. Adv Mater 2023;35:e2301414.
29. Liu J, Wang J, Xu C, et al. Advanced energy storage devices: basic principles, analytical methods, and rational materials design. Adv Sci 2018;5:1700322.
30. Ipadeola AK, Eid K, Abdullah AM. Porous transition metal-based nanostructures as efficient cathodes for aluminium-air batteries. Curr Opin Electroche 2023;37:101198.
31. Ipadeola AK, Haruna AB, Gaolatlhe L, et al. Efforts at enhancing bifunctional electrocatalysis and related events for rechargeable zinc-air batteries. ChemElectroChem 2021;8:3998-4018.
32. Sun C, Yang J, Rui X, et al. MOF-directed templating synthesis of a porous multicomponent dodecahedron with hollow interiors for enhanced lithium-ion battery anodes. J Mater Chem A 2015;3:8483-8.
33. Su Z, Chen T. Porous Noble metal electrocatalysts: synthesis, performance, and development. Small 2021;17:e2005354.
34. Tang J, Xu J, Ye Z, et al. Synthesis of flower-like cobalt, nickel phosphates grown on the surface of porous high entropy alloy for efficient oxygen evolution. J Alloy Compd 2021;885:160995.
35. Adegoke KA, Maxakato NW. Porous metal oxide electrocatalytic nanomaterials for energy conversion: oxygen defects and selection techniques. Coord Chem Rev 2022;457:214389.
36. Sahoo DP, Das KK, Mansingh S, Sultana S, Parida K. Recent progress in first row transition metal Layered double hydroxide (LDH) based electrocatalysts towards water splitting: a review with insights on synthesis. Coord Chem Rev 2022;469:214666.
37. Mofokeng TP, Ipadeola AK, Tetana ZN, Ozoemena KI. Defect-engineered nanostructured Ni/MOF-derived carbons for an efficient aqueous battery-type energy storage device. ACS Omega 2020;5:20461-72.
38. Salah B, Ipadeola AK, Khan A, et al. Unveiling the electrochemical CO oxidation activity on support-free porous PdM (M = Fe, Co, Ni) foam-like nanocrystals over a wide pH range. Energy Convers Manag 2023;20:100449.
39. Ipadeola AK, Abdelgawad A, Salah B, et al. Self-standing foam-like Pd-based alloys nanostructures for efficient electrocatalytic ethanol oxidation. Int J Hydrogen Energy 2023;48:30354-64.
40. Salah B, Ipadeola AK, Abdullah AM, Ghanem A, Eid K. Self-standing Pd-based nanostructures for electrocatalytic CO oxidation: do nanocatalyst shape and electrolyte pH matter? Int J Mol Sci 2023;24:11832.
41. Lu S, Eid K, Lin M, Wang L, Wang H, Gu H. Hydrogen gas-assisted synthesis of worm-like PtMo wavy nanowires as efficient catalysts for the methanol oxidation reaction. J Mater Chem A 2016;4:10508-13.
42. Airo MA, Otieno F, Mxakaza L, et al. Probing the stoichiometry dependent catalytic activity of nickel selenide counter electrodes in the redox reaction of iodide/triiodide electrolyte in dye sensitized solar cells. RSC Adv 2020;10:39509-20.
43. Navas D, Fuentes S, Castro-Alvarez A, Chavez-Angel E. Review on Sol-Gel synthesis of perovskite and oxide nanomaterials. Gels 2021;7:275.
44. He X, Liao J, Wang S, et al. From nanomelting to nanobeads: nanostructured SbxBi1-x alloys anchored in three-dimensional carbon frameworks as a high-performance anode for potassium-ion batteries. J Mater Chem A 2019;7:27041-7.
45. Parashar M, Shukla VK, Singh R. Metal oxides nanoparticles via sol-gel method: a review on synthesis, characterization and applications. J Mater Sci Mater Electron 2020;31:3729-49.
46. Issa AA, El-Azazy M, Luyt AS. Kinetics of alkoxysilanes hydrolysis: an empirical approach. Sci Rep 2019;9:17624.
47. Danks AE, Hall SR, Schnepp Z. The evolution of ‘sol-gel’ chemistry as a technique for materials synthesis. Mater Horiz 2016;3:91-112.
48. Subbiah T, Bhat GS, Tock RW, Parameswaran S, Ramkumar SS. Electrospinning of nanofibers. J Appl Polym Sci 2005;96:557-69.
50. Huang C, Thomas N. Fabricating porous poly(lactic acid) fibres via electrospinning. Eur Polym J 2018;99:464-76.
51. Sun B, Long Y, Zhang H, et al. Advances in three-dimensional nanofibrous macrostructures via electrospinning. Prog Polym Sci 2014;39:862-90.
52. Liao Y, Loh C, Tian M, Wang R, Fane AG. Progress in electrospun polymeric nanofibrous membranes for water treatment: Fabrication, modification and applications. Prog Polym Sci 2018;77:69-94.
53. Ni J, Zhu X, Yuan Y, et al. Rooting binder-free tin nanoarrays into copper substrate via tin-copper alloying for robust energy storage. Nat Commun 2020;11:1212.
54. Huang B, Pan Z, Su X, An L. Tin-based materials as versatile anodes for alkali (earth)-ion batteries. J Power Sources 2018;395:41-59.
55. Mou H, Xiao W, Miao C, Li R, Yu L. Tin and Tin compound materials as anodes in lithium-ion and sodium-ion batteries: a review. Front Chem 2020;8:141.
56. Nam DH, Kim RH, Han DW, Kwon HS. Electrochemical performances of Sn anode electrodeposited on porous Cu foam for Li-ion batteries. Electrochimica Acta 2012;66:126-32.
57. Zhao H, Zhang G, Jiang C, He X. An electrochemical and structural investigation of porous composite anode materials for LIB. Ionics 2012;18:11-8.
58. Cui J, Yao S, Huang J, et al. Sb-doped SnO2/graphene-CNT aerogels for high performance Li-ion and Na-ion battery anodes. Energy Stor Mater 2017;9:85-95.
59. Wang S, He M, Walter M, Krumeich F, Kravchyk KV, Kovalenko MV. Monodisperse CoSn2 and FeSn2 nanocrystals as high-performance anode materials for lithium-ion batteries. Nanoscale 2018;10:6827-31.
60. Walter M, Doswald S, Krumeich F, et al. Oxidized Co-Sn nanoparticles as long-lasting anode materials for lithium-ion batteries. Nanoscale 2018;10:3777-83.
61. Li Z, Xue H, Wang J, Tang Y, Lee C, Yang S. Reduced graphene oxide/marcasite-type cobalt selenide nanocrystals as an anode for lithium-ion batteries with excellent cyclic performance. ChemElectroChem 2015;2:1682-6.
62. Han Z, Wang B, Liu X, Wang G, Wang H, Bai J. Peapod-like one-dimensional (1D) CoP hollow nanorods embedded into graphene networks as an anode material for lithium-ion batteries. J Mater Sci 2018;53:8445-59.
63. Cui C, Wei Z, Zhou G, et al. Quasi-reversible conversion reaction of CoSe2/nitrogen-doped carbon nanofibers towards long-lifetime anode materials for sodium-ion batteries. J Mater Chem A 2018;6:7088-98.
64. Dong W, Shen D, Yang S, et al. First-principles Study of Mechanical and Electronic Properties of Co-Sn Intermetallics for Lithium Ion Battery Anode. Chem Res Chin Univ 2018;34:235-40.
65. Gao G, Wu HB, Lou XW. Citrate-assisted growth of NiCo2O4 nanosheets on reduced graphene oxide for highly reversible lithium storage. Adv Energy Mater 2014;4:1400422.
66. Wang J, Zhang Q, Li X, Zhang B, Mai L, Zhang K. Smart construction of three-dimensional hierarchical tubular transition metal oxide core/shell heterostructures with high-capacity and long-cycle-life lithium storage. Nano Energy 2015;12:437-46.
67. Zhang Q, Wang J, Dong J, et al. Facile general strategy toward hierarchical mesoporous transition metal oxides arrays on three-dimensional macroporous foam with superior lithium storage properties. Nano Energy 2015;13:77-91.
68. Xu J, Su D, Bao W, Zhao Y, Xie X, Wang G. Rose flower-like NiCo2O4 with hierarchically porous structures for highly reversible lithium storage. J Alloys Compd 2016;684:691-8.
69. Zhu H, Sun Y, Zhang X, Tang L, Guo J. Evaporation-induced self-assembly synthesis of mesoporous FeCo2O4 octahedra with large and fast lithium storage properties. Mater Lett 2016;166:1-4.
70. Zhang Q, Chen H, Han X, et al. Graphene-encapsulated nanosheet-assembled zinc-nickel-cobalt oxide microspheres for enhanced lithium storage. ChemSusChem 2016;9:186-96.
71. Yuan J, Chen C, Hao Y, et al. A facile synthetic strategy to three-dimensional porous ZnCo2O4 thin films on Ni foams for high-performance lithium-ion battery anodes. J Electroanal Chem 2017;787:158-62.
72. Huang G, Li Q, Yin D, Wang L. Hierarchical porous Te@ZnCo2O4 nanofibers derived from Te@Metal-organic frameworks for superior lithium storage capability. Adv Funct Mater 2017;27:1604941.
73. Tong H, Gong Z, Huang Y, et al. Bimetallic zinc-cobalt sulfides embedded within N, S-codoped hollow carbon polyhedra for superior lithium-ion batteries. Appl Surf Sci 2024;652:159233.
74. Liu Y, Wang L, Chen G, et al. Nitrogen-containing bimetal oxides derived from cobalt-zinc MOF as anode materials for lithium-ion batteries. J Phys Chem Solids 2024;185:111728.
75. Chen J, Zhu K, Liang P, et al. Ultrahigh reversible lithium storage of hierarchical porous Co-Mo oxides via graphene encapsulation and hydrothermal S-doping. J Mater Chem A 2022;10:5373-80.
76. Chen M, Zhou W, Qi M, Zhang J, Yin J, Chen Q. Reconstruction of copper shell on metal oxides as enhanced nanoarrays electrodes for lithium ion batteries. Mater Res Bull 2017;86:308-12.
77. Kamran U, Park S. Hybrid biochar supported transition metal doped MnO2 composites: efficient contenders for lithium adsorption and recovery from aqueous solutions. Desalination 2022;522:115387.
78. Gu X, Yue J, Li L, Xue H, Yang J, Zhao X. General Synthesis of MnOx (MnO2, Mn2O3, Mn3O4, MnO) hierarchical microspheres as lithium-ion battery anodes. Electrochim Acta 2015;184:250-6.
79. Feng L, Xuan Z, Zhao H, et al. MnO2 prepared by hydrothermal method and electrochemical performance as anode for lithium-ion battery. Nanoscale Res Lett 2014;9:290.
80. Ashokkumar K, Dhanapandian S, Suthakaran S, Krishnakumar N, Anandan M. Synthesis of MnO2 nanoparticles and its effective utilization as high-performance of supercapacitor. Mater Today Proc 2022;49:2675-8.
81. Jiang H, Hu Y, Guo S, Yan C, Lee PS, Li C. Rational design of MnO/carbon nanopeapods with internal void space for high-rate and long-life Li-ion batteries. ACS Nano 2014;8:6038-46.
82. Yue J, Gu X, Chen L, et al. General synthesis of hollow MnO2, Mn3O4 and MnO nanospheres as superior anode materials for lithium ion batteries. J Mater Chem A 2014;2:17421-6.
83. Gu X, Yue J, Chen L, et al. Coaxial MnO/N-doped carbon nanorods for advanced lithium-ion battery anodes. J Mater Chem A 2015;3:1037-41.
84. Liu Y, Bai J, Ma X, Li J, Xiong S. Formation of quasi-mesocrystal ZnMn2O4 twin microspheres via an oriented attachment for lithium-ion batteries. J Mater Chem A 2014;2:14236-44.
85. Liu X, Zhao C, Zhang H, Shen Q. Facile synthesis of porous ZnMnO3 spherulites with a high lithium storage capability. Electrochim Acta 2015;151:56-62.
86. Yuan J, Chen C, Hao Y, et al. Three-dimensionally porous CoMn2O4 thin films grown on Ni foams for high-performance lithium-ion battery anodes. J Mater Sci 2017;52:5751-8.
87. Yuan J, Chen C, Hao Y, et al. Fabrication of three-dimensional porous ZnMn2O4 thin films on Ni foams through electrostatic spray deposition for high-performance lithium-ion battery anodes. J Alloy Compd 2017;696:1174-9.
88. Luo D, Deng YP, Wang X, et al. Tuning shell numbers of transition metal oxide hollow microspheres toward durable and superior lithium storage. ACS Nano 2017;11:11521-30.
89. Zhang S, Zhang Z, Kang J, et al. Double-shelled nanoporous NiO nanocrystal doped MnO/Ni network for high performance lithium-ion battery. Electrochim Acta 2019;320:134542.
90. Lee J, Seo S, Kim D. Hierarchical Zn1.67Mn1.33O4/graphene nanoaggregates as new anode material for lithium-ion batteries. Int J Energy Res 2019;43:1735-46.
91. Lin S, Zhang T. Electrochemical in-situ generation of Ni-Mn MOF nanomaterials as anode materials for lithium-ion batteries. J Alloy Compd 2023;942:168926.
92. Miao X, Wang L, Shi Y, et al. Cu- and S-doped multielement composite Cu/Mn3O4@SC microspheres derived from bimetallic CuMn-MOF as anode materials for lithium-ion batteries. J Mater Sci 2024;59:3930-46.
93. Zheng M, Tang H, Li L, et al. Hierarchically nanostructured transition metal oxides for lithium-ion batteries. Adv Sci 2018;5:1700592.
94. Lin D, Liu Y, Liang Z, et al. Layered reduced graphene oxide with nanoscale interlayer gaps as a stable host for lithium metal anodes. Nat Nanotechnol 2016;11:626-32.
95. Mwonga PV, Ipadeola AK, Naidoo SR, Quandt A, Ozoemena KI. Annealing boosts the supercapacitive properties of molybdenum disulfide powder. Electroanalysis 2020;32:2642-9.
96. Liu J, Qian D, Feng H, et al. Designed synthesis of TiO2-modified iron oxides on/among carbon nanotubes as a superior lithium-ion storage material. J Mater Chem A 2014;2:11372.
97. Zhao D, Qin J, Zheng L, Cao M. Amorphous vanadium oxide/molybdenum oxide hybrid with three-dimensional ordered hierarchically porous structure as a high-performance Li-ion battery anode. Chem Mater 2016;28:4180-90.
98. Yin Z, Qin J, Wang W, Cao M. Rationally designed hollow precursor-derived Zn3V2O8 nanocages as a high-performance anode material for lithium-ion batteries. Nano Energy 2017;31:367-76.
99. Wang Z, Fei P, Xiong H, Qin C, Zhao W, Liu X. CoFe2O4 nanoplates synthesized by dealloying method as high performance Li-ion battery anodes. Electrochim Acta 2017;252:295-305.
100. Dong T, Wang G, Yang P. Electrospun NiFe2O4@C fibers as high-performance anode for lithium-ion batteries. Diam Relat Mater 2017;73:210-7.
101. Song H, Su J, Wang C. Vacancies revitalized Ni3ZnC0.7 bimetallic carbide hybrid electrodes with multiplied charge-storage capability for high-capacity and stable-cyclability lithium-ion storage. ACS Appl Energy Mater 2018;1:5008-15.
102. Ruttert M, Siozios V, Winter M, Placke T. Synthesis and comparative investigation of silicon transition metal silicide composite anodes for lithium ion batteries. Z Anorg Allg Chem 2019;645:248-56.
103. Chen H, Shen X, Xu K, et al. NiFe-NiFe2O4/rGO composites: controlled preparation and superior lithium storage properties. J Am Ceram Soc 2021;104:6696-708.
104. Zheng G, Deng Y, Yu X, Song M. Fe-based frameworks in situ derived 3D Ni-Co-Fe nanocage TMO anode for LIB batteries. Ionics 2022;28:5489-98.
105. Zheng G, Yu X, Huang X, Deng Y, Yuan Z, Song M. Construction of three-dimensional crumpled Ni-Co TMOs for electrochemical energy storage. Electroanalysis 2023;35:e202200480.
106. Han Q, Zhang W, Zhu L, et al. MOF-derived bimetallic selenide CoNiSe2 nanododecahedrons encapsulated in porous carbon matrix as advanced anodes for lithium-ion batteries. ACS Appl Mater Interfaces 2024;16:6033-47.
107. Yu XY, Lou XW. Mixed metal sulfides for electrochemical energy storage and conversion. Adv Energy Mater 2018;8:1701592.
108. Wang S, Fang Y, Wang X, Lou XWD. Hierarchical microboxes constructed by SnS nanoplates coated with nitrogen-doped carbon for efficient sodium storage. Angew Chem Int Ed 2019;58:760-3.
109. Li W, Chou SL, Wang JZ, Kim JH, Liu HK, Dou SX. Sn4+xP3 @ amorphous Sn-P composites as anodes for sodium-ion batteries with low cost, high capacity, long life, and superior rate capability. Adv Mater 2014;26:4037-42.
110. Ramireddy T, Xing T, Rahman MM, et al. Phosphorus-carbon nanocomposite anodes for lithium-ion and sodium-ion batteries. J Mater Chem A 2015;3:5572-84.
111. Chen C, Fu K, Lu Y, et al. Use of a tin antimony alloy-filled porous carbon nanofiber composite as an anode in sodium-ion batteries. RSC Adv 2015;5:30793-800.
112. Luo W, Shen F, Bommier C, Zhu H, Ji X, Hu L. Na-ion battery anodes: materials and electrochemistry. ACC Chem Res 2016;49:231-40.
114. Wang X, Fan L, Gong D, Zhu J, Zhang Q, Lu B. Core-shell Ge@Graphene@TiO2 nanofibers as a high-capacity and cycle-stable anode for lithium and sodium ion battery. Adv Funct Mater 2016;26:1104-11.
115. Yue C, Yu Y, Sun S, et al. High performance 3D Si/Ge nanorods array anode buffered by TiN/Ti interlayer for sodium-ion batteries. Adv Funct Mater 2015;25:1386-92.
116. Fang Y, Yu XY, Lou XWD. Formation of Hierarchical Cu-Doped CoSe2 microboxes via sequential ion exchange for high-performance sodium-ion batteries. Adv Mater 2018;30:e1706668.
117. Gu H, Yang L, Zhang Y, et al. Highly reversible alloying/dealloying behavior of SnSb nanoparticles incorporated into N-rich porous carbon nanowires for ultra-stable Na storage. Energy Stor Mater 2019;21:203-9.
118. Zhang Y, Zhong W, Tan P, Niu Y, Zhang X, Xu M. Heterogeneous interface design of bimetallic selenide nanoboxes enables stable sodium storage. Inorg Chem Front 2021;8:4796-805.
119. Chen J, Zhang G, Xiao J, et al. A stress self-adaptive bimetallic stellar nanosphere for high-energy sodium-ion batteries. Adv Funct Mater 2024;34:2307959.
120. Zhou P, Zhang M, Wang L, et al. MOFs-derived flower-like hierarchically porous Zn-Mn-Se/C composite for extraordinary rate performance and durable anode of sodium-ion and potassium-ion batteries. Small 2022;18:e2203964.
121. Wang S, Zou R, Liu Q, Chen H. Bimetallic selenide Cu4Mo6Se8 nanosheet arrays grown on a carbon skeleton via MOF-derived with enhanced electrochemical kinetics for high-performance sodium-ion batteries. J Mater Chem A 2023;11:8710-8.
122. Guo Y, Li X, Li Z, et al. Rational construction of Sb2Se3 wrapped bimetallic selenide as anode material toward efficient sodium storage performance. Ceram Int 2024;50:14959-67.
123. Zhang L, Xie P, Zhang X, Zhu B, Liu T, Yu J. Facile synthesis of NiCoSe2@carbon anode for high-performance sodium-ion batteries. J Colloid Interface Sci 2024;662:1075-85.
124. Hou YL, Chen JZ, Zhang BH, Wang HY, Wen WX, Zhao DL. Fast ion/electron transport enabled by MXene confined bimetallic sulfides with heterostructure toward highly effective lithium/sodium storage. Chem Eng J 2024;479:147914.
125. He S, Wang Z, Qiu W, Zhao H, Lei Y. Effect of partial cation replacement on anode performance of sodium-ion batteries. Batteries 2024;10:44.
126. Xu Y, Titirici M, Chen J, et al. 2023 roadmap for potassium-ion batteries. J Phys Energy 2023;5:021502.
127. Marcus Y. Thermodynamic functions of transfer of single ions from water to nonaqueous and mixed solvents: part 3 - standard potentials of selected electrodes. Pure Appl Chem 1985;57:1129-32.
128. Matsuda Y, Nakashima H, Morita M, Takasu Y. Behavior of some ions in mixed organic electrolytes of high energy density batteries. J Electrochem Soc 1981;128:2552-6.
130. Kim H, Kim JC, Bianchini M, Seo D, Rodriguez-Garcia J, Ceder G. Recent progress and perspective in electrode materials for K-ion batteries. Adv Energy Mater 2018;8:1702384.
131. Wu X, Zhang W, Wu N, Pang S, He G, Ding Y. Exploration of nanoporous CuBi binary alloy for potassium storage. Adv Funct Mater 2020;30:2003838.
132. Xie Y, Wang X, Zhang H, et al. Bimetallic alloy nanoparticles embedded in N-doped carbon-based as an anode for potassium-ion storage material. J Electroanal Chem 2024;959:118178.
133. Zhang R, Luo Q, Gong J, et al. Multilevel spatial confinement of transition metal selenides porous microcubes for efficient and stable potassium storage. J Colloid Interface Sci 2023;644:10-8.
134. Chen W, Hu K, Zheng H, et al. GeV4S8: a novel bimetallic sulfide for robust and fast potassium storage. Small 2024;20:e2311638.
135. Ma Y, Ouyang Y, Liang H, et al. Heterostructured CoS2/SnS2 encapsulated in sulfur-doped carbon exhibiting high potassium ion storage capacity. J Coll Interface Sci 2024;661:671-80.
136. Kim S, Kim TH, Park HK, Kang YC, Cho JS, Park GD. Aerosol-assisted synthesis of 3D hybridized reduced graphene oxide-carbon nanotube composite microsphere with cobalt-iron selenide nanocrystal as anode materials for potassium-ion batteries. J Energy Stor 2024;83:110683.
137. Guo J, Wang L, Hu A, Zhang J, Xiao Z. 3D micro-flower structured BiFeO3 constructing high energy efficiency/stability potassium ion batteries over wide temperature range. Adv Funct Mater 2024;34:2313300.
138. Liang H, Wang X, Shi J, et al. Design of heterostructured hydrangea-like FeS2/MoS2 encapsulated in nitrogen-doped carbon as high-performance anode for potassium-ion capacitors. J Colloid Interf Sci 2024;664:96-106.
139. Toro L, Moscardini E, Baldassari L, et al. A systematic review of battery recycling technologies: advances, challenges, and future prospects. Energies 2023;16:6571.
140. Yu X, Li W, Gupta V, et al. Current challenges in efficient lithium-ion batteries’ recycling: a perspective. Glob Chall 2022;6:2200099.
141. González YC, Alcaraz L, Alguacil FJ, González J, Barbosa L, López FA. Study of the carbochlorination process with CaCl2 and water leaching for the extraction of lithium from spent lithium-ion batteries. Batteries 2023;9:12.
142. Porzio J, Scown CD. Life-cycle assessment considerations for batteries and battery materials. Adv Energy Mater 2021;11:2100771.
143. Lorero I, Campo M, Del Rosario G, López FA, Prolongo SG. New manufacturing process of composites reinforced with ZnO nanoparticles recycled from alkaline batteries. Polymers 2020;12:1619.
144. Lorero I, Campo M, Arribas C, Prolongo MG, López FA, Prolongo SG. Epoxy composites reinforced with ZnO from waste alkaline batteries. Materials 2022;15:2842.
145. Alguacil FJ, Lopez FA. Separation iron(III)-manganese(II) via supported liquid membrane technology in the treatment of spent alkaline batteries. Membranes 2021;11:991.
146. Llamas-Orozco JA, Meng F, Walker GS, et al. Estimating the environmental impacts of global lithium-ion battery supply chain: a temporal, geographical, and technological perspective. PNAS Nexus 2023;2:pgad361.
147. Balali Y, Stegen S. Review of energy storage systems for vehicles based on technology, environmental impacts, and costs. Renew Sustain Energy Rev 2021;135:110185.
148. Larcher D, Tarascon JM. Towards greener and more sustainable batteries for electrical energy storage. Nat Chem 2015;7:19-29.
149. Helbig C, Hillenbrand M. Principles of a circular economy for batteries. In: Passerini S, Barelli L, Baumann M, Peters J, Weil M, editors. Emerging battery technologies to boost the clean energy transition. The materials research society series. Cham: Springe; 2024.
150. Automated guided vehicle market size & share report 2030. Automated guided vehicle market trends. Available from: https://www.grandviewresearch.com/industry-analysis/automated-guided-vehicle-agv-market [Last accessed on 2 Aug 2024].
151. Tankou A, Bieker G, Hall D. Scaling up reuse and recycling of electric vehicle batteries: assessing challenges and policy approaches. In: Proc ICCT; 2023. pp. 1-138. Available from: https://theicct.org/wp-content/uploads/2023/02/recycling-electric-vehicle-batteries-feb-23.pdf [Last accessed on 2 Aug 2024].
152. Navarro RP, Seidel P, Lenz L, Kolk M, Krug A. European battery recycling: an emerging cross-industry convergence. In: Arthur D. Litle; 2022. Available from:https://www.adlittle.mx/sites/default/files/viewpoints/ADL_European_battery_recycling.pdf [Last accessed on 2 Aug 2024].
153. Yu W, Guo Y, Shang Z, Zhang Y, Xu S. A review on comprehensive recycling of spent power lithium-ion battery in China. eTransportation 2022;11:100155.
154. Yoo E, Lee U, Kelly JC, Wang M. Life-cycle analysis of battery metal recycling with lithium recovery from a spent lithium-ion battery. Resour Conserv Recy 2023;196:107040.
155. Wang X, Guo W, Fu Y. High-entropy alloys: emerging materials for advanced functional applications. J Mater Chem A 2021;9:663-701.
156. Tian H, Tian H, Wang S, et al. High-power lithium-selenium batteries enabled by atomic cobalt electrocatalyst in hollow carbon cathode. Nat Commun 2020;11:5025.
157. Wang R, Wang H, Zhao H, et al. Highly fluorinated co-solvent enabling ether electrolyte for high-voltage lithium ion batteries with graphite anode. Energy Mater 2023;3:300040.
158. Lei K, Wang J, Chen C, et al. Recent progresses on alloy-based anodes for potassium-ion batteries. Rare Met 2020;39:989-1004.
159. Jin C, Liu T, Sheng O, et al. Rejuvenating dead lithium supply in lithium metal anodes by iodine redox. Nat Energy 2021;6:378-87.
160. Sun J, Liu C, Zheng P, et al. Exploring the potential of one-dimensional van der Waals material V2PS10/carbon composite: an anode design paradigm for lithium-ion batteries. Next Mater 2023;1:100053.
161. Tian H, Liang J, Liu J. Nanoengineering carbon spheres as nanoreactors for sustainable energy applications. Adv Mater 2019;31:e1903886.
Cite This Article
How to Cite
Ipadeola, A. K.; Abdullah, A. M.; Eid, K. Recent advances in porous multimetallic alloy-based anodes for rechargeable alkali metal-ion batteries. Energy Mater. 2024, 4, 400079. http://dx.doi.org/10.20517/energymater.2024.34
Download Citation
Export Citation File:
Type of Import
Tips on Downloading Citation
Citation Manager File Format
Type of Import
Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.
Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.
Comments
Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.