Download PDF
Review  |  Open Access  |  13 Jun 2022

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Views: 41157 |  Downloads: 11355 |  Cited:   37
Energy Mater 2022;2:200020.
10.20517/energymater.2022.24 |  © The Author(s) 2022.
Author Information
Article Notes
Cite This Article

Abstract

Alloying materials (e.g., Si, Ge, Sn, Sb, and so on) are promising anode materials for next-generation lithium-ion batteries (LIBs) and sodium-ion batteries (SIBs) due to their high capacity, suitable working voltage, earth abundance, environmental friendliness, and non-toxicity. Although some important breakthroughs have been reported recently for these materials, their dramatic volume change during alloying/dealloying causes severe pulverization, leading to poor cycling stability and safety risks. Although the nanoengineering of alloys can mitigate the volumetric expansion to some extent, there remain other drawbacks, such as low initial Columbic efficiency and volumetric energy density. Porous microscale alloys comprised of nanoparticles and nanopores inherit micro- and nanoproperties, so that volume expansion during lithiation/sodiation can be better accommodated by the porous structure to consequently release stress and improve the cycling stability. Herein, the recent progress of porous microscale alloying-type anode materials for LIBs and SIBs is reviewed by summarizing the Li and Na storage mechanisms, the challenges associated with different materials, common fabrication methods, and the relationship between the structure and electrochemical properties in LIBs and SIBs. Finally, the prospects of porous microscale alloys are discussed to provide guidance for future research and the commercial development of anode materials for LIBs and SIBs.

Keywords

Alloy-type materials, microsized porous materials, lithium-ion batteries, sodium-ion batteries

INTRODUCTION

Environmental pollution from the increasing demand for energy and the use of fossil fuels is attracting enormous attention[1,2], and the development of green and renewable energy sources is therefore urgently required. High-performance energy storage devices are crucial to clean energy because of their special functions, such as the peak load regulation of electricity grids, electric vehicles, and mobile electronics[1,3]. Batteries composed of alkaline and alkaline earth elements (e.g., Li, Na, K, Mg, Ca, and so on) are now ubiquitous because of their high energy densities, suitable working potentials, good cycling stability, and environmental friendliness. Among them, lithium-ion batteries (LIBs) and sodium-ion batteries (SIBs) are the most widely used. A battery comprises the cathode, anode, electrolyte, and separator[4,5], and the electrode materials are the core of the battery and determine the energy density[6-9]. However, the theoretical capacity of the aforementioned cathode materials is relatively low, resulting in low energy densities in LIBs and SIBs[10].

Compared to cathode materials, the anode materials in LIBs or SIBs have greater potential for attaining higher energy densities because of their diversity and high theoretical capacity. The emerging anode materials can be categized into three types according to their reaction mechanisms with Li and Na ions, namely, intercalation, conversion, and alloying[11,12]. Currently, the most extensively used LIB anode material is intercalating graphite with a theoretical specific capacity of 372 mAh g-1[13]; however, it only delivers a reversible capacity of 35 mAh g-1 in SIBs because of the insufficient interlayer space[14,15]. Hard carbon derived from biomass or other organic precursors with a larger disordered layer distance can store extra Na ions, but the capacity is still far from satisfactory. Furthermore, the reaction potential of carbonaceous anodes with Li and Na is relatively low, thus leading to the dendritic deposition of Li and Na metals in the electrochemical process and safety issues[16]. Conversion-type anode materials mainly consist of transition metal oxides[17], sulfides[18-21], selenides[22] and phosphides. Although conversion-type compounds have higher specific capacities than carbonaceous anodes, they are still plagued by low conductivity and appreciable irreversible capacities[23,24]. Different from intercalation- and conversion-type anodes, alloying materials react with Li and Na ions to form alloys, giving rise to higher theoretical capacities, larger molar ratios, safer potentials, and lower costs [Figure 1A-C] and have thus received immense attention for LIBs and SIBs.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 1. (A) Crustal abundance (ppm) and cost of alloying-type anodes (data for Sb, Ge, Sn, and Si costs taken from Shanghai Metals Market, Kitco Metals, London metal exchange, and Deutsche Rohstoffagentu, respectively). (B) Theoretical lithiation capacities of graphite, Si, Ge, and Sn for LIBs and SIBs. (C) Average potential of various alloy materials in LIBs and SIBs. (D) Crystal structures of Si, Ge, Sn, and Sb.

Currently, alloys containing Si, Ge, Sn, and Sb are desirable anode materials for LIBs and SIBs [Tables 1-2]. As shown in Figure 1D, Si and Ge have a diamond structure that provides space to accommodate Li+ and Na+ ions. Sn has a cubic structure and can also hold more Li+ and Na+. In comparison, Sb has a puckered layered structure and the interlamellar spacing can effectively accommodate the insertion and alloying of Li and Na. The electrochemical reaction of alloys is as follows: Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries (M = Si, Ge, Sn, Sb, and so on), which is different from insertion/desertion in carbon-based anodes. However, the alloying reaction results in a serious volume expansion and internal stress, leading to serious pulverization of the electrode materials [Figure 2A]. The frequent volumetric change during cycling also results in repeated formation and destruction of the solid electrolyte interface (SEI), which is the main reason for fast electrolyte consumption and capacity degradation[25,26]. Moreover, some nonmetallic alloys composed of Si and P show relatively low conductivity and poor rates.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 2. (A) Overview of bulk electrode and porous alloy-based anode materials. (B) Structural design for electrodes comprised of alloying-type anode materials.

Table 1

Electrochemical parameters of major alloying-type anode materials for LIBs

ElementAlloying productTheoretical capacity
(mAh g-1)
Volumetric capacity (mAh cm-3)Volume expansion
ratio (%)
Conductivity
(S m-1)
Density
(g cm-3)
SiLi22Si542002190> 3002.52 × 1042.35
GeLi22Ge5162421803702.175.35
SnLi22Sn599419913409.09 × 1067.28
SbLi3Sb66018892002.45 × 1066.68
Table 2

Electrochemical parameters of major alloying-type anode materials for SIBs

ElementAlloying productTheoretical capacity
(mAh g-1)
Volumetric capacity (mAh cm-3)Volume expansion
ratio (%)
Conductivity
(S m-1)
Density
(g cm-3)
SiNaSi9542190> 2442.52 × 1042.35
GeNaGe36921803002.175.35
SnNa15Sn484719915259.09 × 1067.28
SbNa3Sb66018893902.45 × 1066.68

To address the issues of bulk alloying materials, nanoengineering is adopted to increase the active sites, decrease the Li- and Na-ion diffusion distance, and mitigate the volume expansion and resulting stress[27]. Nanostructured alloys can be endowed with high gravimetric capacities and rates and superior fracture toughness. However, because of the large specific Brunner-Emmet-Teller (BET) surface, low tap density, and complex fabrication process, some obstacles, such as the low initial Coulombic efficiency (ICE), inferior volumetric capacity, and high manufacturing costs, hinder the development and subsequent replacement for traditional electrode materials. Nevertheless, research into microscale alloying anode material continues because compared to nanoscale materials, they have higher ICE, volumetric capacity, and mass loading due to the high tap density and small BET surface. However, microscale alloys with large particle sizes have drawbacks, such as large ion and electron diffusion distances, pulverization, large volumetric expansion, and poor rates [Figure 2B]. Ideal alloying-type anode materials should combine the advantages of both nano- and microscale materials by incorporating an overall microscale morphology and built-in nanoscale features.

On this basis, porous microscale materials consisting of nanoblocks and nanopores have aroused significant interest as a result of the following metrics[28]. First, alloys with controllable porosity and pore size can accommodate the volume change during lithiation/sodiation and decrease the volume expansion at the particle level [Figure 2A]. The smaller volume change further impedes the swelling of the electrode and disconnection of the active materials from the current collector. Second, the porous structure enhances electrolyte infiltration, reduces the ion/electron transfer distance, and decreases the internal impedance of the electrode to improve the kinetics [Figure 3A]. Third, they have a high tap density and small specific BET surface area, leading to a high ICE, mass loading, and energy density. In addition, microsized porous materials show obvious advantages in mass transfer because of the three-dimensional (3D) electrolyte permeating channels. Therefore, microsized alloying materials may show utilization even at a mass loading of the active materials [Figure 3B]. Fourth, the interconnected pores offer more active sites for a higher capacity. In general, porous microscale alloys containing nanoparticles deliver better comprehensive performance than either nano- or microscale alloys in LIBs and SIBs [Figure 4].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 3. (A) Basic configuration and working mechanism of a cell. A typical battery device consists of a porous anode and a cathode film sandwiched between two current collectors and isolated by an insulating separator. (B) Plot showing that the capacity of an electrode is positively proportional to the mass loading of active materials on it[4] (Copyright 2017, The American Association for the Advancement of Science).

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 4. Comparison of critical practical metrics required for commercial LIB anodes.

Alloys are promising materials in rechargeable Li-, Na-, Mg- and K-ion batteries because of their high theoretic capacity. To overcome the intrinsic problems of such alloys, including the large volumetric expansion and unstable interfaces, nanoengineering and their combination with carbonaceous materials are two major strategies that have been summarized in recent reviews[29-33]. Porous microscale particles with engineered nano-units and nanoporosity enable relatively stable cycling and low volumetric expansion. However, reviews focusing on these materials for advanced batteries are scarce, and a timely and comprehensive review to facilitate future development is still lacking. This review summarizes the recent development of porous microscale alloying anode materials for LIBs and SIBs. First, the structural design and fabrication methods are described. Focus is given to materials containing continuous nano-units and adjacent nanopores to take advantage of the merits of nanoscale materials, such as the high activity and small ion diffusion distance, as well as microscale materials, including the high tap density and ICE. Representative alloys consisting of Si, Ge, Sn, and Sb and their electrochemical properties in LIBs and SIBs are described, and their future directions and prospects are discussed. This review offers a critical summary of recent advances in the field and will spur the future development of high-efficiency energy storage devices.

SYNTHETIC METHODOLOGIES OF MICROSIZED POROUS ALLOYING MATERIALS

The controllable synthesis of porous alloying microscale materials consisting of nano-ligaments and surrounding nanopores has been reported. In this section, the fabrication protocols are described, and Tables 3-5 summarize the various preparation techniques, including dealloying, templating, chemical etching, and self-assembly.

Table 3

Synthesis strategies and electrochemical properties of porous Si electrode in LIBs

StructuresMethodMorphologyInitial capacityICECapacity retentionReference
NP-SiDealloying in metallic meltRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries3,200 mAh g-1 at 4.2 A g-177%2,100 mAh g-1 after 100 cycles at 4.2 A g-1 [48]
3D porous SiWet alkaline chemical etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,919 mAh g-1 at 0.5 A g-176.5%1,222 mAh g-1 after 200 cycles at 500 mA g-1[38]
Porous SiMagnesiothermic reducedRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,321 mAh g-1 at 0.2 mA cm-272.7%633 mAh g-1 after 30 cycles at 0.2 mA cm-2[94]
MP-SiMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,638.3 mAh g-1 at 0.1 A g-185.1%1,000 mAh g-1 after 360 cycles at 500 mA g-1[95]
Porous SiBall-milling and acid etching Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,957.6 mAh g-1 at 50 mA g-188.1%1,250 mAh g-1 after 100 cycles at 500 mA g-1[96]
Bulk porous SiCu-assisted etching methodRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,126 mAh g-1 at 0.2 A g-192%1,938.3 mAh g-1 after 50 cycles 200 A g-1[97]
3D mp-SiTemplate methodRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,945 mAh g-1 at 0.84 A g-173.6%1,500 mAh g-1 after 100 cycles at 4.2 A g-1[98]
Mesoporous Si Template methodRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,586 mAh g-1 at 0.2 A g-148.3%1,038 mAh g-1 after 170 cycles at 200 mA g-1[99]
Porous Si nanosheetsTemplate methodRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,551.5 mAh g-1 at 240 mA g-174%800 mAh g-1 after 900 cycles at 8,400 mA g-1[100]
3D porous SiMetal assisted chemical etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,390 mAh g-1 at 0.2 A g-194.4%2,079.3 mAh g-1 after 200 cycles at 0.4 A g-1[101]
MG-SiNWsMetal-assisted chemical etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,111 mAh g-1 at 0.84 A g-152.8%2,111 mAh g-1 after 50 cycles at 0.84 A g-1[72]
Mesoporous Si spongeElectrochemical etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries750 mAh g-1 at 0.1 A g-1156%640 mAh g-1 after 1,000 cycles at 1 A g-1[102]
Porous SiCu-assisted chemical etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,521.9 mAh g-1 at 50 mA g-157.5%996.5 mAh g-1 after 100 cycles at 50 mA g-1[103]
Mesoporous SiMagnesium thermalRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,216 mAh g-1 at 400 mA g-166%1,400 mAh g-1 after 100 cycles at 400 mA g-1[104]
Macroporous SiAluminothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries3,385 mAh g-1 at 0.21 A g-186.7%1,500 mAh g-1 after 100 cycles at 0.84 A g-1[105]
Nanoporous SiAir-Oxidation DemagnesiationRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries3,291 mAh g-1 at 0.36 A g-188%1,200 mAh g-1 after 400 cycles at 1.8 A g-1[106]
Nanoporous SiHCl etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,244 mAh g-1 at 0.06 A g-186%1,400 mAh g-1 after 60 cycles 0.24 A g-1[107]
Porous SiMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,731 mAh g-1 at 0.1 A g-164.8%1,030 mAh g-1 after 300 cycles at 1 A g-1[108]
Porous SiSolvothermal reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,679 mAh g-1 at 0.5 A g-164.6%1,025 mAh g-1 after 700 cycles at 3 A g-1[109]
Mesoporous SiMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries830 mAh g-1 at 0.5 A g-187%1,199 mAh g-1 after 500 cycles at 0.5 A g-1[110]
Porous SiElectrochemical etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,073 mAh g-1 at 4.2 A g-192%1,453.5 mAh g-1 after 100 cycles at 4.2 A g-1[111]
Mesoporous SiZeolite-templatedRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,184 mAh g-1 at 0.1 A g-173.1%888 mAh g-1 after 300 cycles at 1 A g-1[112]
Porous SiAcidic ionic liquidRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,036 mAh g-1 at 0.1 A g-172.9%1,000 mAh g-1 after 100 cycles at 1 A g-1[113]
Porous Si spheresMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries3,191 mAh g-1 at 0.42 A g-180.9%1,620 mAh g-1 after 20 cycles at 420 mA g-1[114]
Macroporous SiElectrochemical anodizationRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,425 mAh g-1 at 0.04 A g-181%600 mAh g-1 after 20 cycles at 200 mA g-1[115]
Porous SiHeat-treated and HCl etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,182 mAh g-1 at 0.1 A g-171.4%750 mAh g-1 after 250 cycles at 2 A g-1[116]
nSiElectrochemical Mg dealloyingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,920 mAh g-1 at 1 A g-190%1,430 mAh g-1 after 300 cycles at 3 A g-1[46]
NP-SiMg-Bi-Si dealloyingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries3,500 mAh g-1 at 2.1 A g-185%2,000 mAh g-1 after 150 cycles at 2.1 A g-1[117]
Porous SiBall-milling and stain-etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,900 mAh g-1 at 0.42 A g-183%1,100 mAh g-1 after 600 cycles at 2.1 A g-1 [118]
Porous SiMagnesiothermic reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,045.6 mAh g-1 at 1 A g-168.2%697.7 mAh g-1 after 200 cycles at 1 A g-1[119]
1D-PSiNWMetal assisted chemical etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,534.5 mAh g-1 at 1 A g-156.5%586.7 mAh·g-1 after 5,000 cycles at 16 A g-1[120]
Porous SiFe-Si dealloyingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries3,185 mAh g-1 at 50 mA g-189%2,120 mAh g-1 after 140 cycles at 0.5A g-1[121]
Porous Si spheresMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries3,105 mAh g-1 at 0.21 A g-164%1,500 mAh g-1 after 500 cycles at 2.1 A g-1[122]
Porous SiAl-Si dealloyingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries2,606 mAh g-1 at 0.53 A g-176.7%1,000 mAh g-1 after 300 cycles at 1.4 A g-1[123]
Mesoporous Si Self-assembly methodsRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,917.2 mAh g-1 at 20 mA g-169.9%1,147.7 mAh g-1 after 100 cycles at 0.2 A g-1[88]
Mesoporous SiMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,300 mAh g-1 at 100 mA g-175%1,053 mAh g-1 after 100 cycles at 0.1 A g-1[124]
Table 4

Synthesis strategies and electrochemical properties of porous Ge electrodes in LIBs

StructuresMethodMorphologyInitial capacityICECapacity retentionReference
3D Porous GeEtching SiO2 templateRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,380 mAh g-1 at 0.5C90%1,415 mAh g-1 after 100 cycles at 1C[141]
Mesoporous GeMechanochemical reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries944 mAh g-1 at 0.1C43.8%789 mAh g-1 after 20 cycles at 0.1C[142]
Porous GeSelf-assembledRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,450 mAh g-1 at 0.03C75.5%1,500 mAh g-1 after 40 cycles at 0.3C[143]
Microporous GeMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,299 mAh g-1 at 0.1C80.3%1,131 mAh g-1 after 200 cycles at 1C[63]
Macro-GeHydrothermal reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,263 mAh g-1 at 0.2C84%.755 mAh g-1 after 3,500 cycles at 1C [144]
Meso-porous GeMetallothermal reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,340 mAh g-1 at 0.2C65.5%1,217 mAh g-1 after 400 cycles at 0.8C[39]
3D-pGeHydrogen reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,341.1 mAh g-1 at C/2092.3%770 mAh g-1 after 250 cycles at 1C[145]
Porous GeModified Stöber methodRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,568 mAh g-1 at C/2085.3%1,140 mAh g-1 after 300 cycles at 0.5C[146]
Mesoporous GeRedox transmetalation reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,585 mAh g-1 at 0.05C89%1,400 mAh g-1 after 300 cycles at 0.5C[147]
Ge microcubesHydrogen reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,275 mAh g-1 at 0.1C91.8%1,204 mAh g-1 after 500 cycles at 1C[148]
Mesoporous GeMetathesis reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,479.5 mAh g-1 at 0.2C80.4%1,048 mAh g-1 after 1,000 cycles at 1.0C[149]
Table 5

Synthesis strategies and electrochemical properties of porous Sn and Sb electrode in SIBs

StructuresMethodMorphologyInitial
capacity
ICECapacity retentionReference
Porous SnDealloyingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries500 mAh g-1 at 40 mA g-190.9%275 mAh g-1 after 95 cycles at 40 mA g-1[166]
Porous SnPhase-inversion techniqueRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries674 mAh g-1 at 423.5 mA g-163%519 mAh g-1 after 500 cycles at 423.5 mA g-1[161]
3D porous SnReplacement reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries823 mAh g-1 at 200 mA g-1700 mAh g-1 after 400 cycles at 2.5A g-1[162]
Microsized SnBall-milledRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries847 mAh g-1 at 0.847 A g-192%768 mAh g-1 after 100 cycles at 0.25A g-1[164]
Porous SnGalvanic replacement reactionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries1,100 mAh g-1 at 0.1A g-169%480.8 mAh g-1 after 100 cycles at 1A g-1[167]
Hollow SnSolution-based reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries680 mAh g-1 at 20 mA g-178.5% 220 mAh g-1 after 50 cycles at 20 mA g-1[168]
Porous SbMagnesiothermic reductionRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries570.3 mAh g-1 at 50 mA g-181.4%538 mAh g-1 after 50 cycles at 50 mA g-1[169]
Porous SbSb-Mn dealloyingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries545 mAh g-1 at 66 mA g-1541 mAh g-1 after 200 cycles at 330 mAg-1[170]
Porous Sb/CBall-millingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries480 mAh g-1 at 0.1A g-1283 mAh g-1 after 3,000 cycles at 5.0 A g-1[171]
Mesoporous SbMg-Sb dealloyingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries640 mAh g-1 at 0.2 A g-167.9%658 mAh g-1 after 200 cycles at 0.2 A g-1[172]
NP-Sb70Chemically etchingRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries420 mAh g-1 at 3.3 A g-1573.8 mAh g-1 after 200 cycles at 1.0 A g-1[173]
Sb porous microspheresTemplates methodRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries634.6 mAh g-1 at 0.1 A g-164.6%617 mAh g-1 after 100 cycles at 1.0 A g-1[174]
Sb hollow spheresGalvanic replacementRecent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries603.5 mAh g-1 at 50 mA g-177%622.2 mAh g-1 after 50 cycles 50 mA g-1[175]

Dealloying methods

Dealloying is often expressed as a bimetallic or multi-metal alloy corrosion process, in which more active metals in the precursors react selectively with the medium and are removed operando in the following steps to leave a more stable component. During the reaction, the residual component diffuses between the parent alloy and reactive medium to form a bicontinuous porous structure with 3D open pores and maximum interfaces[34]. Dealloying has attracted intensive attention for the fabrication of porous alloys because pore creation does not require additives and adjustable structures. The 3D porous alloying metal supplies rich Li and Na storage and active sites and enables rapid mass transfer, which is vital for LIBs and SIBs. According to different alloy precursors, the porosity of the structure can be adjusted from 40% to 80% by taking advantage of the alloy parting limit[34,35]. The dealloying of porous structures is carried out by chemical, electrochemical, low-melting metal, and other dealloying methods.

Chemical dealloying is one of the most common methods to obtain porous structures from the alloy precursor via spontaneous chemical reactions based on the different standard electrode potentials in an acidic or alkaline solution. The precursor alloy must be homogeneous or have an intermetallic phase comprising a relatively active metal and another more inert one. Porous metals containing Si, Ge, Sn, Sb, and so on have been prepared by dealloying routes. For example, Feng et al. fabricated porous Si consisting of nanodendrites by direct dealloying of a commercial Al-Si eutectic alloy with microparticles in 5 M HCl for 24 h[36]. Zhang et al. prepared porous microscale antimony from Mg-Sb alloys by one-step chemical vapor dealloying with a HCl vapor as the dissolution medium at 600 °C [Figure 5A][37]. Sohn et al. prepared porous Si via the dealloying of Si/Al-Cu-Fe nanoparticles obtained from smelting, roller melt-spinning, and attrition milling via dual-step acidic and alkaline wet chemical etching[38]. Lin et al. reported a low-temperature molten salt system for the metallothermic (Mg or Zn) reduction of GeO2 to obtain mesoporous hollow Ge microspheres, and the molten ZnCl2 promotes the reaction at a low temperature. In the etching process, part of the metal alloy matrix is consumed due to acid leaching to generate abundant pores, and subsequent alkaline etching decreases the Si domain size and enlarges the pores[39].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 5. Schematic illustrations of porous metal preparation via dealloying fabrication methods. (A) Schematic illustration of fabrication procedures of P-Sb. Reproduced with permission[37] (Copyright 2020, Elsevier B.V.). (B) Preparation of bulk 3D porous-Si by dealloying in a metallic melt. Reproduced with permission[47] (Copyright 2014, ACS Publications). (C) Schematic illustration of fabrication procedures of 3D porous Si. Reproduced with permission[48] (Copyright 2016, Elsevier B.V.). (D) Schematic illustration for the multi-step synthesis of ultrafine Si porous structures. Reproduced with permission[49] (Copyright 2020, Kluwer Academic Publishers).

In addition to single metals, porous bimetallic materials have also been obtained by chemical dealloying using ternary alloys as the raw materials. Yang et al. produced Si-Ge bimetallic porous materials via the dealloying of Al-Ge-Si microparticles with an Al content of 80% in HCl[40]. Other bimetallic compounds, such as Bi-Sn[41], Ag-Ge[42], SnSb[43], and SnNi[44], have been formed by similar dealloying methods. Although chemical dealloying using acidic and alkaline media as the dissolving agents is an emerging route, there are some shortcomings, such as its time-consuming nature, the production of waste acidic or alkaline solutions, and structural collapse of the porous architecture. To solve these problems, electrochemical dealloying methods have been developed to obtain porous Si, Ge, and Ge-Si alloys by adjusting the applied potential. Yin et al. conducted electrochemical dealloying in a molten CaCl2-NaCl eutectic electrolyte at 600 °C to produce porous Ge[45]. Yuan et al. adopted Mg alloying/dealloying using a molten MgCl2-NaCl-KCl electrolyte at 500 °C to fabricate porous Si[46]. Although the porous structure can be retained, the conditions in electrochemical dealloying are relatively harsh and an oxygen-free environment is needed to prevent oxidation of the porous materials.

Liquid metal dealloying has been proposed to prepare porous alloys because of the controllable porosity. Wada et al. reported the fabrication of bulk 3D porous Si via a metallic melting dealloying route with Mg2Si as the precursor and Bi metal as the dealloying melting medium[47]. Mg dissolves selectively from the Si-Mg alloy into the Bi melt to produce porous Si at high temperatures [Figure 5B]. Similarly, Wada et al. fabricated porous 3D Si by immersing a Mg-Si alloy precursor in a Bi melt and washing it with 1 M nitric acid [Figure 5C][48]. However, electrochemical and melting metal dealloying have some disadvantages, including environmental issues and economic efficiency, which hinder their wider application.

Alternative dealloying methods have been developed to obtain more desirable porous structures efficiently. For instance, Chen et al. and Wang et al. synthesized porous Si structures with various particle sizes via the oxidation of ball-milled Mg2Si in Ar/O2 or air in conjunction with pickling reactions in HCl to remove the MgO impurity [Figure 5D][49,50]. We recently developed a new strategy to fabricate ant-nest-like nanoporous Si microparticles by nitriding Mg2Si particles under nitrogen at 750 °C to generate a Mg3N2/Si composite, followed by the dissolution of Mg3N2 in HCl. The ant-nest-like porous Si comprises 3D-interconnected Si nano-ligaments and bicontinuous nanopores[51]. More recently, An et al. fabricated uniform porous Si particles by treating a commercial Mg-Si alloy in CO2 at 500-700 °C and etching MgO in HCl[52]. In this process, CO2 is reduced into the carbon layer, which is in situ coated on the surface of porous Si. The gas dealloying method involving a reactive gas, such as O2, N2, or CO2, reacting with the Si-Mg alloy produces phase separation and generates the 3D composite comprising Si nano-ligaments and the surrounding intermediate product. The porous 3D structure is obtained after dissolving the intermediate product. Compared to the chemical/electrochemical dealloying method, the gas dealloying technique not only can avoid structural collapse during etching but is also more environmentally friendly and faster.

Template method

The template method is one of the most common methods to construct porous materials because the product has the inverse structure of the porous template. Moreover, the size of the porous structure can be adjusted by controlling the template parameters. The common templates are opals, organic spheres, and SiO2. The template method proceeds by the following steps. Infiltration of the precursor by chemical vapor deposition (CVD) or the sol-gel method is performed, followed by elimination of the matrix either by thermal annealing or chemical etching[53]. For example, Song et al. used silica opal as a porous template to prepare Ge inverse opal with porous walls by CVD and a Ge wall layer on the silica opal[54]. The silica template was then etched selectively with HF. Geier et al. produced porous Ge structures by etching and annealing the Ge precursor-loaded poly(methyl methacrylate) template[55]. The template was eliminated by heating at 500 °C for 5 min under a vacuum. In addition to Ge and Si, regular porous structures can be prepared by the deposition of a Si layer on a SiO2 opal template with Si2H6 as the gas precursor, followed by dissolution in HF[56]. However, in addition to the complicated preparation process and high-cost template, the template method for porous structures often involves highly toxic acids or emits a large amount of CO2[57,58].

To overcome these drawbacks, the self-template synthesis of porous materials has been developed using metal oxides, such as SiO2, GeO2, and SnO2, as both the raw materials and self-sacrificial templates[59,60]. For instance, Bao et al. converted microscale diatomite into porous 3D microscale Si with continuous nano-ligaments at a relatively low temperature of 650 °C by the following reaction[59]:

2Mg + SiO2 → 2MgO + Si (1)

Some natural or synthetic porous SiO2 materials, such as zeolite SSZ-13[61], have been used in the fabrication of porous Si. The by-product of MgO is usually used as the pore-forming agent, which can generate uniform nanoscale pores after dissolution in diluted HCl[62]. Similarly, porous Ge can be obtained by magnesiothermic reduction (MgR) with GeO2 as the self-sacrificial template[63]. In addition to MgR, aluminothermic reduction in molten salt can convert the SiO2 template into porous Si at a low temperature below 300 °C. For example, porous crystalline Si has been synthesized by the aluminothermic reduction of an AlCl3 salt at 250 °C[64], and other reducing agents, such as Zn and NaBH4, have been employed to convert GeO2 into porous Ge[65].

The soluble salt-template method has been developed to prepare porous alloys because the soluble salts can be easily removed. The pores are produced by removing the salt by-products of the template rather than chemical etching, thus providing both high porosity and controllability. In addition, the surface area and pore size of the materials can be adjusted by selecting appropriate alloy precursors. Some inorganic salts, such as NaCl, KCl, and LiCl, have been utilized as templates[66]. Dai et al. reported a bottom-up approach for mesoporous crystalline Si with tunable particle/pore size by SiCl4 reduction [Figure 6A][67]. In this process, SiCl4 is the Si precursor and NaK is the reducing agent. Si and the salt by-products (NaCl and KCl) are considered to precipitate simultaneously to form a composite nano-matrix due to the poor solubility of NaCl/KCl and Si in the solvent. In the subsequent heat treatment, the amorphous Si is transformed into a crystalline framework composed of interconnected primary particles. The product is washed with water to remove the by-products of salts to form porous Si. Tang et al. designed a series of mesoporous Ge materials by the salt-templating method with NaCl and KCl as the in situ templates [Figure 6B][68]. The surface area and pore size and volume can be adjusted by changing the reaction time, temperature, and calcination temperature, and the salt by-products as internal self-forming templates can be removed without an etchant. Compared to traditional template methods, this method is simpler, faster, and more environmentally friendly.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 6. (A) Schematic of synthesis route of mesoporous crystalline Si. Reproduced with permission[67] (Copyright 2014, Nature Publishing Group). (B) Synthetic route of mesoporous Ge. Reproduced with permission[68] (Copyright 2020, Elsevier Inc.).

Chemical etching method

Chemical etching is an effective method to fabricate porous alloys based on different galvanic displacements or chemical activities[69]. Etching methods can be mainly divided into electrode and electroless etching. Electrode etching is controlled by the current, whereas the electroless process often involves other metal ions, such as Ag+ and Cu2+[70-72]. For example, porous Si nanowires have been produced with silver nitrate in HF and doped Si[73,74]. Typically, two reactions take place during etching[69]:

4Ag+ + 4e- → 4Ag (2)

Si + 6F- → [SiF6]2- + 4e- (3)

The porosity and pore size of porous Si can be controlled by the current density and concentration of the etchant. Zhong et al. studied the mechanism of the formation of porous Si nanowires in a two-step Ag-assisted electroless chemical etching method [Figure 7A][75]. The H-terminated Si was coated with Ag by electroless deposition or physical vapor deposition and then placed into a solution containing HF and H2O2 with variable concentrations for different time durations. Zhang et al. synthesized porous Si materials by the ferrite-assisted chemical etching of the Si substrate [Figure 7B][76]. In this reaction, ferrite catalyzes the reaction of Si particles with glycol to produce porous Si particles. To avoid the use of the noble metal Ag, some groups have developed a less expensive Cu ion etching method. For example, porous Si has been prepared by the chemical etching of Cu nanoparticles in HF/H2O2[77]. In addition to Si, porous Ge is also easy to fabricate by this method[78] using an electrolyte of HF:H2O2:H2O (2:3:1). Chemical etching can produce porous Si/Ge with different porosities, but the formation is relatively slow and involves highly corrosive HF.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 7. Schematic illustrations of porous metal preparation via chemical etching method. (A) Schematic illustration of porous Si nanowires. Reproduced with permission[75] (Copyright 2012, American Chemical Society). (B) Schematic illustration of the formation process of porous Si wafer. Reproduced with permission[76] (Copyright 2015, Wiley-VCH Verlag).

Electrochemical etching can also produce porous alloys by applying a current[79]. The starting materials are often wafers, such as Si and Ge, and the electrolyte is typically HF. Uhlirs at Bell Laboratories accidentally found that porous Si could be generated by the electrochemical etching of a Si wafer[80]. The pore formation mechanism and adjustment of the pore parameters were investigated[69]. Thakur et al. synthesized a macroporous Si film by electrochemical etching in 48% HF and dimethylformamide using a constant current density of 2 mA cm-2[81]. The porosity and depth of porous Si can be adjusted by the current density and HF concentration[82]. Bioud et al. reported the fabrication of porous Ge films by galvanostatic etching using a two-electrode system with a Ge wafer and Pt plate as the working and counter electrodes, respectively[83]. The electrolyte consisted of HF and anhydrous ethanol. Porous materials obtained by electrochemical etching have high density and purity. Although the electrochemical etching method can produce porous Si and Ge with different porosities using an applied current, it involves the use of highly corrosive HF and can cause environmental concerns[84].

Self-assembly methods

Self-assembly is attractive in the preparation of porous microstructures with nanostructures, such as nanoparticles, nanowires, and nanosheets. This method is cost-effective, scalable, simple, and can produce particles with a high tap density. Jia et al. designed porous microscale Si spheres composed of nanoscale crystalline Si with a uniform porous structure by self-assembly using the bottom-up microemulsion approach [Figure 8A][85]. Jeong et al. adopted a cyclic iodide process to assemble microclusters comprising many kinked Si nanowires with a diameter of less than 20 nm[86]. The secondary microscale particles had considerable internal porosity, good thermodynamic stability, and high tap density. Jia et al. produced hierarchical porous Si microspheres of 6-8 μm in diameter consisting of Si-coated CNTs by self-assembly [Figure 8B][87]. By introducing molten salts, such as KCl and NaCl, a liquid environment was created to absorb heat and increase ion moving ability at a high temperature. Yao et al. prepared mesoporous diatomite-derived Si microspheres with NaCl as the molten liquid interface to grow Si nanocrystals by the self-assembly of mesoporous Si nanocrystals in MgR[88]. Nevertheless, there is a rate-limiting step in which the conversion of an irregular shape to a nearly spherical shape depends on the concentration of fused salts and liquid magnesium, which may dissolve each other by the interactions of ions and the negative ions in the cavity of these positive ions depend on the vibration frequency of Mg2+, Na+, and Cl-[89].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 8. (A) Schematic illustration of preparation process of porous Si. Reproduced with permission[85] (Copyright 2018, Elsevier B.V.). (B) Schematic showing synthesis of CNT@Si@C microspheres. Reproduced with permission[87] (Copyright 2020, Nature Publishing Group).

MICROSIZED POROUS ALLOYING MATERIALS FOR LIBS AND SIBS

Alloys containing Si, Ge, Sn, and Sb are promising anode materials for next-generation LIBs and SIBs because of their high theoretical capacity and volumetric capacities [Tables 1 and 2]. The porous alloy architecture comprising continuous nano-units surrounded by pores can accommodate the volume expansion and offer paths for electrolyte permeation to increase the stability of the electrodes. Moreover, porous microscale alloys have a high tap density and volumetric energy density in LIBs and SIBs. In this section, different alloys with porous structures for LIBs and SIBs are described, and the relationship between their structures and electrochemical characteristics is discussed.

Si-based porous structures for LIBs and SIBs

Si has the unique cubic space group of Fd3m and a large interplanar crystal spacing of 5.472 Å. It is promising anode material for LIBs on account of its high theoretical capacity, abundant resources, low cost, and non-toxicity. Si also has a low delithiation potential of ~0.4 V vs. Li+/Li, which is higher than that of graphite (below 0.2 V). In addition, Si reacts electrochemically with Li to form alloys during cycling, thus offering space to increase the Li-ion storage[90]. Chevrier et al. carried out first principles calculations on the Si-Li system[91]. Shenoy et al. derived the optimal lattice parameters of the crystalline Li-Si phases using density functional theory (DFT) calculations according to the crystal structures from the Inorganic Crystal Structure Database[92]. According to different reaction conditions, the phases in the Li-Si system were confirmed as LiSi, Li12Si7, Li7Si3, Li13Si4, and Li22Si5. Among them, Li22Si5 showed the highest total theoretical capacity of 4,200 mAh g-1[93]. By combining as many as 4.4 Li atoms, Si shows a volume expansion (~300%) and huge stress during lithiation/delithiation, which causes pulverization, unstable SEI layers, structural collapse, and fast capacity fading. Porous microscale Si-based structures can be divided into nano-unit self-assembled porous Si and continuous and connected skeleton porous Si, as shown in Table 3[38,46.48,72,88,94-124].

Nano-unit self-assembled microporous Si

Si nanomaterials with a small size and high specific surface area also exhibit high capacity and improved cycling stability but suffer from low ICE and tap density. Although it has a high gravimetric capacity of above 2,000 mAh g-1, the volumetric energy density of nano-Si is even lower than that of commercial graphite. The assembly Si nanostructures into microscale secondary particles can combine the merits of nano- and micromaterials to produce a high tap density and capacity simultaneously[125]. Moreover, the nano-units self-assembled in microscale porous Si have many pores to accommodate volume expansion and reduce the volume change on the particle level in LIBs[112]. Porous Si consists of nano-units for lithiation from end-to-end, whereas solid larger Si particles are surface-to-center[126].

Liu et al. fabricated yolk-shell spherical Si pomegranate microbeads with a diameter of up to 10 μm via the self-assembly of Si@SiO2 nanoparticles with the aid of emulsifiers [Figure 9A][127]. The spherical Si pomegranate microbeads show a high reversible capacity of 2,350 mAh g-1 at a rate of C/20 (1 C = 4,200 mA g-1) in conjunction with a capacity retention of 97% after 1,000 cycles at C/2, significantly better than that of Si NPs [Figure 9B]. The enhanced performance of these composites stems from the incorporation of carbon and the smaller particle sizes, with the well-defined internal space allowing Si to expand without changing the secondary particle size. Jia et al. reported porous Si (p-Si) comprised of Si nanoparticles[85]. The unique porous structure endows the materials with good electrical contact and provides a fast route for the transportation of lithium ions, as well as good internal stress accommodation. As a result, p-Si exhibits a reversible capacity of 2,500 mAh g-1 at 0.26 A g-1 with an ICE of 75% and a cycling stability capacity of 1,467 mAh g-1 at 2.6 A g-1 after 370 cycles with a capacity retention of 83%. Furthermore, Shi et al. designed a porous microscale structure made of TiO2-coated Si nanoparticles in building blocks that were further filled with hierarchical Si@TiO2@C (SA-SiTC)[128]. In this composite, the carbon filler and TiO2 layer act as a mechanical buffer and electrolyte reservoir, respectively. The SA-SiTC electrode displayed long cycling stability (842.6 mAh g-1 after 1,000 cycles at 2 A g-1), high volumetric capacity (174 mA h cm-3) and ICE (80.9%), and a stable SEI layer.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 9. (A) Structural evolution of pomegranate Si microparticles before and after electrochemical cycling. (B) Cycling performance of the Si pomegranate and other structures at C/2. Reproduced with permission[127] (Copyright 2014, Springer Nature). (C) Microscale Si particles incorporating self-kinked Si nanowire clusters. (D) Cycling performance of samples at 0.5 C. Reproduced with permission[86] (Copyright 2020, Wiley-VCH Verlag). (E) SEM images of Si nanowires and mapping of Si. Reproduced with permission[129] (Copyright 2020, Elsevier B.V.).

In addition to Si nanoparticles, Si nanowires can be assembled into porous secondary microscale particles. Jia et al. applied self-assembly to prepare hierarchical porous Si microspheres of 6-8 μm in diameter with Si-coated CNTs (CNT@Si@C)[87]. The CNT@Si@C electrode with a practical mass loading (3 mAh cm-2) showed a specific capacity of 750 mAh g-1, an initial swelling of < 20% at 100% state-of-charge and a capacity retention of 92% over 500 cycles. Owing to the unique highly porous yarn-ball-like structure with high porosity and good mechanical strength, the CNT@Si@C microspheres showed the expected effect of suppressed swelling of porous Si particles upon lithiation to maintain the overall structural integrity. Jeong et al. fabricated Si clusters with an average size of 2.7 μm by a silicon iodide (SiI4) decomposition method, as shown in Figure 9C[86]. The porous Si was composed of kinked nanowires with diameters of less than 20 nm. The kinked Si nanowires not only formed voids but also contacted each other to yield electrolyte transfer channels for better ionic conductivity and interconnections for better electronic conductivity. The nanowire-assembled porous particles have ample voids between the nanowires and can accommodate large volume changes. As a result, they showed a capacity of 2,784.7 mAh g-1 at 0.1 C with an ICE of 82.4%, together with a superior capacity retention of 83.6% for 1,000 cycles [Figure 9D]. Recently, Choi et al. produced porous microscale Si consisting of nanowires with a diameter of 300 nm and a length of > 1 μm by self-assembly using a molten salt electrochemical reduction method from RH-derived SiO2[129]. As shown in Figure 9E, the product has a spherical shape with a porous structure. The porous Si anode has a high capacity of 3,161 mAh g-1 at 0.05 C, corresponding to an ICE of 84% and a capacity of 2,550 mAh g-1 after 100 cycles at 0.2 C. Even at a high rate of 1 C, the capacity is maintained at 1,371 mAh g-1.

Continuous microporous Si

Different from porous particles made of nano-units, 3D continuous porous Si networks containing continuous nano-skeletons may have better structural stability to alleviate fracture during volume changes. Continuous nanoporous Si microparticles have been demonstrated to improve the cycling characteristics because the connected 3D ligaments disperse the stress during the lithiation/delithiation or sodiation/desodiation of Si. Kim et al. reported 3D porous bulk Si microparticles fabricated by thermal annealing and etching of the composites of butyl-capped Si gels and SiO2 nanoparticles[130]. Because the thin pore wall of ~40 nm can accommodate a large strain without pulverization, the porous Si particles maintain a high charging capacity of 2,800 mAh g-1 at a current density of 2,000 mA g-1 after 100 cycles, together with excellent rate performance. Jia et al. designed lotus-root-like 3D mesoporous Si by MgR and acid etching[98]. Porous Si has uniform mesopores and the particles are composed of 30 nm Si crystallites that facilitate fast electrode kinetics. The 3D mesoporous Si has high electrochemical reversibility with a capacity retention of 94.4% at 100 cycles.

Liu and co-workers reported a sub-micrometer mesoporous Si sponge prepared by anodization, and the porous Si showed a capacity of 750 mAh g-1 based on the total electrode weight with an 80% capacity retention over 1,000 cycles[102]. In situ TEM demonstrated that the mesoporous Si sponge limits the volume expansion of the active particles and prevents pulverization of the bulk Si particles [Figure 10A]. Choi et al. fabricated porous Si microparticles via an aluminothermic reaction and subsequent MgR [Figure 10B][105]. Because of the special porous structure, which can accommodate large volume changes and facilitate ion transportation during cycling, the porous Si/Al2O3 microcomposite exhibited outstanding stability with a capacity of 1,500 mAh g-1 after 100 cycles at C/5 and a small volume expansion of 34% [Figure 10C]. Tao et al. adopted chemical dealloying to prepare a series of porous Si materials by etching Al-Si alloys [Figure 10D][131]. The porous Si maintained a spherical structure and the pores were uniformly distributed. As a result, the porous Si had a high capacity of 3,004 mAh g-1 at 500 mA g-1 in the first cycle with an ICE of 81% [Figure 10E]. The reversible capacity of porous Si was maintained at 1,525 mAh g-1 after 100 cycles at 1,000 mA g-1. This remarkable stability was attributed to the porous structure, which alleviates the volume change during repetitive lithiation/delithiation [Figure 10F].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 10. (A) TEM images and selected area electron diffraction patterns of mesoporous Si sponge particles at different lithiation states. Reproduced with permission[102] (Copyright 2014, Nature Publishing Group). (B) SEM images of macroporous Si structure. (C) Cycling performance of porous Si/Al2O3 composite at C/5. Reproduced with permission[105] (Copyright 2015, Elsevier B.V.). (D) SEM images of porous Si. (E) Cycling performance of different porous Si samples at 500 mA g-1. (F) Volume changes can be restricted by porous construction during cycling. Reproduced with permission[131] (Copyright 2019, Elsevier Inc.).

Mu et al. investigated the impact of commercial Si nanoparticles (C-SiNP) and mesoporous Si microparticles (MP-Si)[95]. The MP-Si showed outstanding cycling stability and a high reversible capacity of 1,793 mAh g-1 after 120 cycles at 200 mA g-1. The Li-ion storage properties benefit from the two-dimensional (2D) nanoplates and abundant mesopores in the Si microparticles, which not only promote immersion of the electrolyte and the transportation of lithium ions, but also buffer the volume expansion and ensure the structural stability of the electrode. Sohn and co-workers prepared 3D porous Si microparticles by alkali and acid etching[38]. The 3D porous Si electrode showed a high initial charge capacity of 2,919 mAh g-1 with an ICE of 76.5% and a capacity of 1,222 mAh g-1 after 200 cycles with an average CE of ~99%. The excellent electrochemical properties of the 3D porous Si suggest that the etching-induced morphological modification, including pore enlargement and reduction of the Si size, is highly effective in improving the cycling characteristics.

Porous ant-nest-like microscale Si (AMPSi) has been produced by a low-cost and scalable top-down approach, involving thermal nitridation of a Mg-Si alloy and removal of the Mg3N2 by-product in an acidic solution [Figure 11A and B][51]. The high porosity of AMPSi allows the inward expansion of the 3D interconnected Si nano-ligaments during cycling, which gives rise to high structural stability and negligible particle-level outward expansion [Figure 11C]. Therefore, after carbon deposition, the AMPSi and C composites (AMPSi@C) showed a high ICE of 80.3%, reversible capacities of 2,134 and 1,271 mAh g-1 at 0.1 and 0.5 C (1 C = 4,200 mA g-1), respectively, and a 90% capacity retention from 20 to 1,000 cycles in a lithium-ion half-cell. The AMPSi@C//NCM full cell exhibited a high reversible capacity of 134 mAh g-1 at 0.5 C (1 C = 160 mA g-1) with a capacity retention of 84% for over 400 cycles together with a high energy density of 502 Wh kg-1. In situ TEM indicated that AMPSi@C remained stable during lithiation and delithiation, with volumetric expansion introduced to the pores to restrict outward expansion [Figure 11D].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 11. (A) Schematic showing preparation of AMPSi and AMPSi@C. (B) Photograph of an ant nest (scale bar = 20 cm). (C) Schematic of lithiation/delithiation process of ant-nest-like microscale porous Si particles. (D) Time-resolved TEM images depicting the lithiation process of AMPSi@C electrode (scale bar = 200 nm). Reproduced with permission[51] (Copyright 2019, Nature Publishing Group). (E) Cycling performance of Mp-Si with different surface-engineering treatments. Time-lapse TEM images of lithiation of (F) dense Si and (G) Mp-Si@Si@G microparticles. Reproduced with permission[132] (Copyright 2019, Elsevier B.V.).

Wang et al. designed mesoporous Si microparticles by the thermal disproportionation of low-cost SiO microparticles and subsequent removal of the SiO2 by-product[132]. After graphene cage-encapsulated Si skin-sealed mesoporous Si (Mp-Si@Si@G) microparticles were obtained by a surface-engineering strategy by depositing a dense silicon skin onto each mesoporous Si (Mp-Si@Si) microparticle, encapsulation was performed to form a conformal graphene cage. Mp-Si@Si@G showed a high reversible specific capacity of 1,246 mAh g-1 after 300 cycles at a rate of C/2, which can be ascribed to the optimal specific surface area and suitable porosity of Mp-Si@Si@G, consequently decreasing the electrolyte accessible area, SEI formation and irreversible Li consumption [Figure 11E]. The morphology of Mp-Si@Si@G was monitored by in situ TEM during lithiation and an apparent volume expansion of the dense Si microparticles was observed [Figure 11F]. Cracks appeared after only 32 s and widened with time. At 160 s, the dense Si microparticles fractured abruptly into smaller particles. In contrast, the structures of the Mp-Si@Si@G microparticles were well maintained throughout the whole lithiation process [Figure 11G] and remained intact even after full lithiation because of the sufficient inner pore space and the mechanically strong graphene cage that limited volume expansion.

Si-based alloys have been developed as anode materials for SIBs with a capacity of 954 mAh g-1[133], and the volume change of NaSi formed from Si is ~244%, which is much smaller than that for Ge-, Sn- and Sb-based anodes in SIBs. In general, the sodiation of Si has different phase behaviors, insertion voltages, and kinetic barriers compared to the reaction with Li. For instance, because of the difference in the ionic radii of Na+ (0.97 Å) and Li+ (0.68 Å), the insertion of Na into crystalline Si is limited[134,135]. Xu et al. obtained amorphous Si particles by thermal plasma CVD of silane, which showed reversible Na-ion storage with a capacity of 270 mAh g-1 at 20 mA g-1[134]. Wu and co-workers demonstrated the fabrication of amorphous Si-rich interior cores surrounded by a Sn shell (a-SiSN) via mechanical alloying[136]. The amorphous sponge-like a-SiSN has pores [Figure 12A], leading to a reversible capacity of 151 mAh g-1 after 100 cycles at 36.25 mA g-1 and a high rate capacity of 143 mAh g-1 at 725 mA g-1Figure 12B and C]. Recently, Qiu et al. prepared mesoporous Si microspheres (MSMs) with homogeneously distributed mesopores ranging from 1 to 10 nm via the MgR of SiO2[137]. As shown in Figure 12D, the mesoporous Si microspheres with abundant mesopores exhibited a high reversible capacity of 390 mAh g-1 at 100 mA g-1 after 100 cycles [Figure 12E] and a high rate capacity of 160 mAh g-1 at a high current density of 1,000 mA g-1 [Figure 12F]. The good cycling and excellent rate capability can be attributed to the small particle size and mesoporous structure. However, for the practical application of Si materials in SIBs, several issues still need to be addressed, including structural optimization, electrolyte design, and the development of novel binders[138-140].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 12. (A) SEM of a-SiSN. (B) Long-term cycling properties of a-SiSN electrode. (C) Rate performance of a-SiSN. Reproduced with permission[136] (Copyright 2016, Elsevier Ltd.). (D) SEM images of MSMs. (E) Discharge capacity retention of MSMs and Si microspheres. (F) Rate performance of MSMs. Reproduced with permission[137] (Copyright 2018, Springer Verlag).

Ge-based porous structures for LIBs and SIBs

Ge is in the same chemical group as Si and is regarded as a promising alloying anode material because of its high theoretical capacities [Tables 1 and 2] and suitable operating voltages for LIBs and SIBs. Moreover, the conductivity of Ge is 100 times higher than that of Si. Similarly, Ge anodes also suffer from severe volume expansions, leading to pulverization during lithiation/delithiation. Porous Ge has garnered significant attention since the pores in porous Ge can act as volume expansion buffers to improve the cycling stability. Table 4[39,63,141-149] lists the synthesis strategies and compares the electrochemical properties of different porous Ge anode materials for LIBs.

Yang and co-workers prepared mesoporous Ge microparticles by a mechanochemical reaction of GeO2 and Mg powders and subsequent etching in HCl[142]. The mesoporous Ge showed a reversible capacity of 950 mAh g-1 and a capacity of 789 mAh g-1 after 20 cycles at a current density of 150 mA g-1. Jia et al. prepared 3D macroporous Ge (p-Ge) microparticles with a hexagonal-like morphology by MgR and HCl etching[63]. The p-Ge anode had a high reversible capacity of 1,131 mAh g-1 at 1 C after 200 cycles and a high rate capability with a capacity of 717 mAh g-1 at 5 C (1 C = 1,624 mA g-1). The internal space in 3D p-Ge buffers the volume change of Ge during cycling and preserves the structural stability of the anode materials. Cui and his team synthesized Ge microcubes with a hierarchical porous structure directly on a titanium foil by a simple hydrogen reduction method [Figure 13A][148]. The Ge microcubes showed good cycling and rate characteristics due to the voids, such as a high ICE of 91.8% stemming from the high crystallinity of Ge in reversible lithium insertion and extraction, a less adverse side reaction for irreversible Li loss, and unique hierarchical porous structure for easier electrolyte penetration. The Ge microcubes exhibited a high retention capacity of 1,204 mAh g-1 after 500 cycles at a rate of 1 C, corresponding to a capacity retention of 94.1% from the first cycle, and a good rate of 1,121 mAh g-1 at a high rate of 50 C [Figure 13B].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 13. (A) SEM images of porous Ge microcubes. (B) Cycling performance of porous Ge microcubes. Reproduced with permission[148] (Copyright 2015, American Chemical Society). (C) SEM of mesoporous Ge. (D) Cycling performance of Ge (ZnRR) electrodes. Reproduced with permission[150] (Copyright 2017, Wiley-VCH Verlag GmbbH & Co.). (E) Top view of pGe with 18.05% porosity and one surface Li. In the lower panel, the process of the gradual lithiation is shown with one, four, and twenty Li surface atoms in sequence. Green, white and purple balls represent Ge, H, and Li atoms, respectively. (F) Band gap evolution as a function of the porosity of fully H passivated pGe (dark squares) and pGe with one (red circles) and two (blue triangles) Li atoms. Reproduced with permission[152] (Copyright 2020, John Wiley & Sons Inc.).

Choi and co-workers synthesized porous Ge by metallothermic reduction using two different metal reductants by the zincothermic reduction reaction (ZnRR) and MgR, respectively[150]. Figure 13C shows the SEM image of the ZnRR-synthesized Ge particles [Ge (ZnRR)], which have a mesoporous structure with numerous mesopores. With mesopores in the inner parts, the porous Ge provides extra spaces for the expansion of particles and accommodates the large volume change during lithiation/delithiation. The Ge (ZnRR) electrode showed a high ICE of ≈85%, excellent cycling stability at a low temperature of -20 °C, and a high specific capacity of 566 mAh g-1 at a rate of 0.5 C after 50 cycles [Figure 13D]. The Ge//LiFePO4 full cell exhibited an ICE value of 98% in the first cycle and retention of 80% at 0.5 C after 50 cycles at -20 °C. Mishra et al. synthesized porous microscale Ge by the hydrogen reduction of GeO2[151]. The porous Ge electrode showed a reversible capacity of 1,300 mAh g-1 after 340 cycles at 1,000 mA g-1 and a high capacity of 437 mAh g-1 at 16 A g-1. The excellent long-term cycling can be attributed to the porous morphology of the Ge electrode, which can maintain the stress induced by Li ions without crushing, provide good electrolyte infiltration and shorten the Li-ion diffusion length. Sosa et al. studied the effects of surface and interstitial Li on the electronic properties of porous Ge (pGe) using DFT calculations and showed that diffusion is easier due to the ion conductivity [Figure 13E and F][152]. The porous framework of Ge not only facilitates the fast diffusion of ions and electrons but also buffers the volume change during cycling.

Kohandehghan et al. reported that amorphous Ge can reduce the nucleation barrier of the NaxGe phase and improve solid-state ion diffusion to enhance Na-ion storage[153]. Lu et al. also demonstrated that Ge can be readily and reversibly sodiated after turning crystalline Ge nanowires into amorphous ones in the initial lithiation step[154]. The amorphous Ge nanowires exhibited a 300% volumetric expansion upon sodiation, corresponding to the formation of Na1.6Ge. Li et al. prepared porous Ge microparticles with a Mn-Ge alloy by selective corrosion[155]. The porous Ge microparticles showed a broad ligament size distribution in the range between 100 and 300 nm, as shown in Figure 14A. The porous Ge microparticle electrode showed a reversible capacity of 260 mAh g-1 in the initial cycle and good cycling stability over 50 cycles at C/5 [Figure 14B]. Yi et al. synthesized mesoporous amorphous Ge microparticles by acid etching the Zintl-phase Mg2Ge at room temperature and studied the Na-ion storage mechanism in meso-porous amorphous Ge (A-Ge) by in situ Raman scattering and ex situ XRD[156]. As shown in Figure 14C, A-Ge with a rough surface showed a high reversible charge capacity of 600 mAh g-1, a capacity retention of 90.2% after 50 cycles at 0.2 C, and a rate capacity of 273 mAh g-1 at 5 C [Figure 14D]. The in situ Raman spectrum in Figure 14E obtained during sodiation/desodiation reveals no distinct peak shift or peak intensity change in C-Ge during the first charging/discharging cycle. This is because Na ions cannot be easily inserted into the interior of C-Ge during discharging. However, the mesoporous A-Ge showed an apparent signal change during charging/discharging. As shown in Figure 14F, the in situ Raman spectrum discloses that the mesoporous amorphous Ge is active with Na in the SIB. The amorphous phase of the Ge anode lowers the sodiation barrier and increases the Na+ storage. The mesoporous architecture also provides fast electronic/ionic transportation channels to enhance the rate and provides space for the volume expansion.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 14. (A) SEM images of porous Ge microparticles. (B) Galvanostatic charge-discharge curves of porous Ge microparticles. Reproduced with permission[155] (Copyright 2019, Elsevier). (C) SEM image of mesoporous A-Ge microparticles. (D) Rate capability from 0.2 to 5 C of A-Ge. In-situ Raman spectrum of (E) crystalline and (F) amorphous Ge during the first discharge/charge curve. Reproduced with permission[156] (Copyright 2019, Tsinghua University Press).

Sn-based porous structures for LIBs and SIBs

Like Si and Ge, metallic Sn can alloy with Li and Na as an alternative to traditional graphite anodes due to its high theoretical capacity [Tables 1 and 2], high electrical conductivity, moderate operating voltage, and safety. However, the practical application of metallic Sn to LIBs and SIBs is affected by the huge volume change, aggregation of Sn particles during cycling, serious pulverization, and loss of electrical connection with the current collector. The newly formed surface consumes Li by forming the SEI film, and the aggregation of particles leads to poor kinetics, fast capacity decay, and poor cyclability. Direct or indirect methods to alleviate the volume expansion by reducing the particle size and introducing an accommodation space or buffer are currently being studied.

Pore formation can improve the electrochemical properties of electrode materials. However, porous Sn obtained by a simple method still presents significant challenges due to the low melting point and inevitable electrochemical agglomeration of Sn. Song et al. fabricated porous 3D continuous microscale Sn from immiscible Al-Sn alloys[157]. The microporous Sn anode showed an outstanding initial capacity of 818 mAh g-1 and a high CE of 99%. The porous structure of Sn still retained the porous framework after first cycling [Figure 15A and B]. Ryu et al. prepared porous Sn consisting of homogeneous Sn particles prepared by vacuum deposition[158]. The porosity of the Sn exceeded 98% and the porous Sn anode showed a high initial discharge capacity of 1,939 mAh g-1 and a charge capacity of 714 mAh g-1 at 0.1 C [Figure 15C]. The reversible capacity of the porous Sn anode changes from 603 to 378 mAh g-1 at a rate of 1 C after 100 cycles, corresponding to a capacity retention of 63% [Figure 15D]. As shown in Figure 15E, the volume expansion of the porous Sn anode was only 34% during the first lithiation. Cook et al. prepared porous tin (P-Sn) particles by MgR [Figure 15F][159]. The P-Sn particles exhibited a high capacity of 578 mAh g-1 and the maximum capacity remained at 80% after 300 cycles at 250 mA g-1 [Figure 15G]. The good cycling performance of NP-Sn was attributed to the porous structure that buffers the volume change during alloying. X-ray microscopy showed that the porosity is retained in both the charged and discharged states. As shown in Figure 15H-J, the size distribution of the P-Sn particles was wide (1-20 μm). The pristine particles showed an interconnected network of random mesopores [Figure 15K] arising from dealloying. Figure 15L and M show that the initial random pores and interconnected ligaments still exist even in a fully lithified state with volume expansion (~300%). Therefore, the porous structure not only accommodates Li ions during volume expansion but also retains space for electrolyte diffusion.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 15. SEM images of porous Sn structure (A) before and (B) after lithiation/delithiation process for first cycling. Reproduced with permission[157] (Copyright 2015, Springer). (C) Charge/discharge profiles of porous Sn anode during the first three cycles. (D) Cycling performance of Sn samples. (E) Cross-sectional SEM images of the as-prepared sample and after 100 cycles of porous Sn anode. Reproduced with permission[158] (Copyright 2019, John Wiley & Sons Ltd.). (F) SEM images of porous Sn. (G) Cycling performance of porous Sn. (H) Pristine sample. (I) Electrode recovered after one electrochemical cycle at 1.2 V. (j) Fully lithiated sample at 70 mV. (H), (I) and (J) show the magnified subsurface renderings of the highlighted areas in panels (K), (L), and (M), respectively. Reproduced with permission[159] (Copyright 2016, American Chemical Society).

Metallic Sn with a porous structure is also promising in SIBs. Wang et al. studied the two-step reaction for sodiation of Sn by in situ TEM, which revealed a two-phase reaction mechanism with a modest volumetric expansion of ~60% and a single-phase reaction with a large volume expansion of ~420% after full sodiation[160]. The large volume expansion/contraction exerted strong mechanical pressure on the Sn nanoparticles, resulting in the production of cracks, serious pulverization of Sn, continuous growth of the SEI, slow kinetics of Na ions, low CE, and rapid electrode failure due to the loss of contact between particles. The porous structure not only provided more space for the volume expansion of Sn but also reduced the ion transfer distance to improve the kinetics. Kim et al. prepared porous Sn (PSn) by the phase-inversion technique for anode materials in SIBs[161]. The PSn electrode showed a high initial capacity of 1,066 mAh g-1 and a reversible capacity of 519 mAh g-1 after 500 cycles at 0.5 C [Figure 16A and B].

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 16. (A) Charge-discharge profiles of PSn for 500 cycles. (B) Cycling performance of PSn. Reproduced with permission[161] (Copyright 2016, Elsevier). (C) Comparison of voltage and specific capacity of a full battery. (D) XRD patterns of porous Sn/Cu anode at different states during the first cycle €. (F) Refined crystal structures of intermediates, NaSn3, NaSn, Na9Sn4, and Na15Sn4, in discharging process. Reproduced with permission[162] (Copyright 2018, Wiley-VCH Verlag GmbbH & Co.). (G) In situ XRD patterns of microsized Sn during the first cycle. Reproduced with permission[164] (Copyright 2016, Wiley-Blackwell).

Wang et al. obtained 3D porous Sn by tuning the redox potential of Cu+/Cu with a ligand of thiourea to trigger the replacement reaction between Cu and Sn2+[162]. The porous Sn electrode showed a high reversible capacity of 700 mAh g-1 at 2,500 mA g-1 for 400 cycles, which was significantly better than that of the bare Sn. The full battery of Sn/Cu//Na3(VO0.5)2(PO4)3F2 had high energy and power densities of 311.7 Wh kg-1 and 182.5 W kg-1, respectively [Figure 16C], and good cycling performance for 200 cycles without any obvious degradation observed. As shown in Figure 16D and E, the electrochemical reaction was studied by XRD in selected discharging and charging states. During discharging, the new diffraction peak belonging to the cubic phase of NaSn3 was found in state B (discharging to 0.2 V). After discharging to 0.1 V (state C), tetragonal NaSn with a space group of I41/acd could be observed. Upon discharging to 0.06 V (state D), the characteristic diffraction peaks of Na9Sn4 became evident and after full discharging to 0.01 V, the diffraction peaks at 16.5°, 19.0°, 31.8°, 33.3°, 34.7°, and 37.4° corresponded to the final discharge product of cubic Na15Sn4. As shown in Figure 16(F), metallic Sn was transformed to cubic NaSn3 and then orthorhombic NaSn in the initial Na+ insertion process. Upon further sodiation, the rhombohedral Na9Sn4 containing Sn-Sn dimers with a distance of 2.85 Å and five Na atoms oriented in the same direction was formed. Finally, cubic Na15Sn4 was generated after full sodiation, in which an isolated Sn atom is surrounded by Na atoms in the relatively symmetric 12-fold coordination. In the reverse process, Na9Sn4 was gradually desodiated to Sn with four distinctive plateaus of states E-I, as shown in Figure 16D and E. The diffraction peak of Cu6Sn5 showed no change during cycling, which is indicative of its nonactivity in SIBs[163]. However, Cu6Sn5 acted as a buffer layer to improve the electron conductivity between the active Sn and Cu foil. It was speculated that the reversible reactions of Sn were Sn↔NaSn3↔NaSn↔Na9Sn4↔Na15Sn4. Zhang et al. also observed other sodiation phases, such as NaSn, Na9Sn4, and Na15Sn4, by in situ XRD [Figure 16G][164].

Sb-based porous structures for LIBs and SIBs

Sb has attracted significant attention as an anode material for LIBs and SIBs because of its high theoretical capacity [Tables 1 and 2], abundant resources, low cost, and environmental friendliness[165]. Moreover, Sb has high electrical conductivity, reliable ion diffusivity, and low volume expansion (200%) compared to Si. Sb anodes have a safer Li+/Na+ insertion potential of ~0.5 V, which not only affords stable working voltages in batteries but also shuns the formation of lithium dendrites at low voltages. Sb with a 2D layered structure offers a desirable framework for fast lithium storage. However, the dense Sb still leads to structural fracturing and instability of the SEI layer on the surface of Sb during charging/discharging to accelerate capacity degradation. So far, efforts have been made to overcome these issues. Pore formation in Sb materials can shorten the diffusion length of Li+ and Na+ and accommodate the volume expansion to improve the electrochemical properties. Typical examples of porous Sb-based materials and their associated electrochemical properties are presented in Table 5[161,162,164,166-175].

Hou et al. prepared porous hollow Sb spheres by a facile galvanic replacement method[175]. The hollow Sb spheres showed an initial specific capacity of 629.6 mAh g-1 and retained a reversible capacity of 627.3 mAh g-1 after 50 cycles at 100 mA g-1, corresponding to a capacity retention of 99.6% [Figure 17A]. In addition, such an electrode exhibited a high rate capacity of 435.6 mAh g-1 at 1600 mA g-1 [Figure 17B]. The superior electrochemical properties can be attributed to the unique hollow and porous spherical structure [Figure 17C]. The hollow Sb spheres adapted to the volume change and strain during Li+ insertion and extraction, and the 3D Sb structure ensured efficient and continuous electron transportation. The large electrode/electrolyte contact area also provided many active centers for charge transfer to enhance the electrochemical kinetics. Rodriguez et al. synthesized 3D Sb chains with a rhombohedral phase by reduction [Figure 17D][176]. The 3D Sb electrode exhibited excellent Li storage properties, as exemplified by a highly reversible capacity of 523 mAh g-1 at 0.5 C after 100 cycles and a retention of 92% [Figure 17E]. Moreover, these materials showed more stable rates and a high specific capacity of 430 mAh g-1 at 1 C [Figure 17F]. The sodium storage performance of the 3D Sb chains was better than that of Sb/Gr due to the alleviated volume expansion. Meng and co-workers prepared porous Sb with 3D Sb dendrites by a galvanic replacement reaction between Sb3+ and Zn powder[177]. The porous Sb with 3D dendrites showed a high initial charging capacity of 838.6 mAh g-1 at 0.1 A g-1, corresponding to an ICE of 79.9% and a capacity retention of 603.2 mAh g-1 after 124 cycles.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 17. (A) Cycling performance of Sb hollow spheres. (B) Rate performance of Sb hollow spheres. (C) Schematic of lithiation of Sb hollow spheres. Reproduced with permission[175] (Copyright 2014, American Chemical Society). (D) SEM of porous hollow Sb chains. (E) Cycling performance of porous hollow Sb chains and Sb/Gr. (F) Rate performance of porous hollow Sb chains and Sb/Gr. Reproduced with permission[176] (Copyright 2019, American Chemical Society).

Hou et al. prepared porous hollow Sb microspheres (PHMSs) for SIBs using a replacement reaction employing Zn microspheres as the template[174]. The porous shell was metallic Sb and the inner core was a metallic Zn template. Hollow microspheres with porous shells were obtained after removing the Zn templates. The Sb PHMSs had a high initial charge capacity of 634.6 mAh g-1 at a current density of 100 mA g-1, corresponding to an ICE of 64.6%. In addition, the Sb PHMSs exhibited a high reversible capacity of 617 mAh g-1, a retention of 97.2% after 100 cycles at 100 mA g-1, and excellent rate performance (312.9 mAh g-1 at 3,200 mA g-1). The superior electrochemical characteristics stemmed from the unique porous and hollow structure of Sb PHMSs, which could accommodate the strain and facilitate Na+ diffusion during repeated alloying/dealloying. Liu et al. studied the effects of the morphology of porous antimony (P-Sb) and the size of Sb particles on the electrochemical properties[173]. The coral-like P-Sb exhibited excellent electrochemical reversibility after 200 cycles at 100 mA g-1. The P-Sb electrode could be reversibly cycled at a high current density of 3,300 mA g-1 and showed a high capacity of 420 mAh g-1, indicating the fast transfer of sodium ions and electrons. The P-Sb could adapt to the volume change of Na+ insertion and improve the capacity retention. The high surface area of the pores allowed for efficient Na+/electron exchange to improve the capacity and cycling life of SIBs. The improvement rendered by the coral-like Sb anode arose mainly from the high porosity, and therefore a suitable porous structure plays an essential role in the electrochemical properties of Sb electrodes.

Ma et al. fabricated mesoporous antimony-based Sb2O3@Sb with an ultrafine bicontinuous mesoporous structure in the process of dealloying the Mg90Sb10 precursor[172]. Sb2O3@Sb with a mesoporous structure exhibited an ultrahigh specific capacity of 659 mAh g-1 at 0.2 A g-1 and excellent cycling stability [Figure 18A]. Operando XRD measurements were executed to study the Na storage mechanism in the mesoporous Sb2O3@Sb electrode [Figure 18B]. At the beginning of the first discharge, only crystalline Sb2O3 and Sb peaks could be observed. With the progress of the first discharge, the peak intensities of Sb2O3 gradually decreased and disappeared, indicating the conversion reaction of Sb2O3 with Na to form Na2O and Sb. In the later stage, the Sb peaks began to weaken and finally vanished. Moreover, the peak of Sb at 28.7° slightly shifted to lower 2θ angles, perhaps due to the insertion of Na into the Sb lattice at the early stage of sodiation. Afterwards, the well-defined crystalline peaks of hexagonal Na3Sb appeared in the final stage of the first discharge. In contrast, Li et al. reported multidimensional porous Sb nanostructures (Sb NS) synthesized by a facile chemical dealloying strategy[178]. Such Sb NS anodes exhibited a high reversible capacity of 620 mAh g-1 and retained 90.2% capacity after 100 cycles at 100 mA g-1 [Figure 18C]. An ex situ XRD analysis of the Sb-NS electrodes at different charge-discharge states was carried out to further understand the mechanisms of sodium storage [Figure 18D]. In the discharge process, the intensities of the Sb peaks gradually weakened and almost disappeared when the cell was fully discharged to 0.01 V. Moreover, new peaks of 33.5° and 34.3° appeared, corresponding to the formation of the Na3Sb phase, indicating the conversion reaction of Sb with Na to form Na3Sb. During the subsequent charge process, the Na3Sb peaks gradually disappeared, while the intensities of the Sb peaks increased, suggesting the transformation of Na3Sb into Sb.

Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries

Figure 18. (A) Cycling performance of Sb2O3@Sb. (B) Operando XRD patterns of mesoporous Sb2O3@Sb electrode. Reproduced with permission[172] (Copyright 2018, Elsevier). (C) Cycling performance of Sb-NS electrodes. (D) Ex situ XRD patterns of Sb-NS electrodes at different charge-discharge states. Reproduced with permission[178] (Copyright 2019, American Chemical Society).

CONCLUSION AND OUTLOOK

Earth-abundant alloying materials (e.g., Si, Ge, Sn, and Sb) with high gravimetric and volumetric capacities have emerged as promising electrode materials to replace low-capacity carbonaceous anode materials in next-generation high energy density LIBs and SIBs. Nevertheless, the large volume change and unstable electrode interface of alloys impede their commercial development. Micro- or sub-microporous alloying materials comprised of nano-units can overcome some of these hurdles by combining the virtues of nano- and microscale materials. In this review, recent advances in porous microscale alloying materials are described from the perspective of high-capacity anodes in LIBs and SIBs. Porous alloys have remarkable properties in Li- and Na-ion storage by combining the structural advantages of nano- and micromaterials. Recent developments pertaining to the composition/structural design, modification, and mechanisms to improve the electron/ion conductivity and structural stability are discussed. Different fabrication methods, including dealloying, templating, chemical etching, and self-assembly, are summarized, and common alloying materials and their relevant properties for LIBs and SIBs are compared. Although significant progress has been made, the practical implementation of porous microscale alloys still faces some challenges.

First, the recognized evaluation parameters of porous structures and the quantitative relationship between the electrochemical characteristics and structures remain unclear. On the one hand, the porosity determines the tap density and internal space to accommodate volume expansion. Higher porosity results in a smaller volume expansion at the particle level but leads to a lower tap density and volumetric capacity of the active materials and vice versa. On the other hand, the size and arrangement of the nano-units in the porous materials affect the active sites, Li- and Na-ion insertion distance, and mechanical stability of the active materials. Nano-ligaments with a smaller size are beneficial for Li- and Na-ion storage and transmission, but the mechanical strength of the framework may not be sufficient to endure the large stress during Li/Na alloying. The particle size determines the tap density and electron/ion distance of the active materials and is crucial to the optimization of the electrochemical properties. The various alloys have different properties. For instance, research into the alloys with high conductivity, such as Sn, Sb, and Bi, should focus on improving the electron conductivity, whereas that of other alloys, such as Si, need to balance the electron/ion conductivity. Overall, the quantitative assessment of the structural parameters must be improved, and the relationship with the structure of porous alloying anode materials must be better understood to expedite their application in LIBs and SIBs.

Second, large-scale and low-cost fabrication methods for different materials are highly desirable for LIBs and SIBs. Currently, porous alloys are prepared by chemical/noble-metal etching dealloying, liquid-metal dealloying, and templating. Chemical/electrochemical dealloying using an acidic solution to selectively dissolve one active metal component and leave another metallic ligament is one of the most common methods for porous alloys. However, this method is time-consuming and not suitable for active metals, and the parameters of bicontinuous metallic frameworks are difficult to control. The liquid metal dealloying method has problems, such as the high temperature, metal residues, and high cost rendering it to be not economical or sustainable. The template technique requires costly porous templates that are difficult to implement at a large scale. Although other methods, such as the gas dealloying of Mg2Si, have been used to obtain porous Si comprising nanoscale pores and metallic ligaments, the large-scale and low-cost fabrication of porous alloying materials with adjustable structures requires more research and development.

Third, issues, such as the mechanical stability and electrolyte absorption, affect the cycling lifetime of batteries. In the typical electrode production process, most of the high-capacity anodes, such as Si, should be mixed with commercial graphite particles, and the rolling procedure under tens of MPa pressure is essential to obtaining high-quality electrode films with a high tap density and surface evenness. The porous structure of alloys can be destroyed easily during ball milling with graphite and rolling of the electrode. Hence, optimizing the structural parameters to enhance the mechanical strength should be considered in practical applications. In contrast, the porous structure with a higher specific surface area and more pores consume more electrolytes, resulting in the nonuniform dispersion of electrolytes and more side reactions, which may produce inferior stability and raise safety concerns. Therefore, it is urgent to understand the relationship between electrolyte adsorption/consumption and materials properties, such as porosity.

Fourth, better comprehension of the phase/morphological evolution of materials at the atomic/molecular level and the failure mechanism of porous structures is crucial to the preparation and development of high-performance alloying electrodes. Although the results obtained by in situ characterization techniques are helpful in providing a better fundamental understanding, they are not sufficient to clearly explain how to improve the electrochemical properties of actual batteries. Furthermore, the formation mechanism and interfacial reactions of the SEI layers between alloying materials and electrolytes are still not well understood, and further research, including quantitative analysis, is necessary in order to fathom the effects of different materials under actual application conditions.

DECLARATIONS

Authors’ contributions

Conceived the manuscript: Li G, Guo S, Gao B

Wrote the manuscript: Li G, Guo S, Gao B

Reviewed the manuscript: Gao B, Chu P, Huo K

Contributed to the discussion of the manuscript: Li G, Xiang B, Mei S, Guo S, Zheng Y, Zhang X, Gao B, Chu P, Huo K

Availability of data and materials

Not applicable.

Financial support and sponsorship

This work was financially supported by the National Natural Science Foundation of China (U2004210, 51974208, U2003130, 21875080, and 52002297), the Outstanding Youth Foundation of Natural Science Foundation of Hubei Province (2020CFA099), the Special Project of Central Government for Local Science and Technology Development of Hubei Province (2019ZYYD024), the Innovation group of Natural Science Foundation of Hubei Province (2019CFA020), and the City University of Hong Kong Strategic Research Grants (7005505).

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) 2022.

REFERENCES

1. Gogotsi Y, Simon P. Materials science. True performance metrics in electrochemical energy storage. Science 2011;334:917-8.

2. Zu L, Zhang W, Qu L, et al. Mesoporous materials for electrochemical energy storage and conversion. Adv Energy Mater 2020;10:2002152.

3. Jiang B, Li J, Luo B, et al. LiPO2F2 electrolyte additive for high-performance Li-rich cathode material. J Energy Chem 2021;60:564-71.

4. Cheng HM, Li F. Charge delivery goes the distance. Science 2017;356:582-3.

5. Zhang Z, Zhang D, Lin H, Chen Y. Flexible fiber-shaped supercapacitors with high energy density based on self-twisted graphene fibers. J Power Sources 2019;433:226711.

6. Pu X, Wang H, Zhao D, et al. Recent progress in rechargeable sodium-ion batteries: toward high-power applications. Small 2019;15:e1805427.

7. Mishra K, Yadav N, Hashmi SA. Recent progress in electrode and electrolyte materials for flexible sodium-ion batteries. J Mater Chem A 2020;8:22507-43.

8. Xu B, Qian D, Wang Z, Meng YS. Recent progress in cathode materials research for advanced lithium ion batteries. Mater Sci Eng R Rep 2012;73:51-65.

9. Choi JU, Voronina N, Sun Y, Myung S. Recent progress and perspective of advanced high-energy Co-less Ni-rich cathodes for Li-ion batteries: yesterday, today, and tomorrow. Adv Energy Mater 2020;10:2002027.

10. Zhai L, Li G, Yang X, et al. 30 Li + -accommodating covalent organic frameworks as ultralong cyclable high-capacity Li-ion battery electrodes. Adv Funct Materials 2022;32:2108798.

11. Wang L, Światowska J, Dai S, et al. Promises and challenges of alloy-type and conversion-type anode materials for sodium-ion batteries. Mater Today Energy 2019;11:46-60.

12. Zhang H, Hasa I, Passerini S. Beyond insertion for Na-ion batteries: nanostructured alloying and conversion anode materials. Adv Energy Mater 2018;8:1702582.

13. Li X, Wang Y, Lv L, Zhu G, Qu Q, Zheng H. Electroactive organics as promising anode materials for rechargeable lithium ion and sodium ion batteries. Energy Mater 2022;2:200014.

14. Zhao L, Hu Z, Lai W, et al. Hard carbon anodes: fundamental understanding and commercial perspectives for Na-ion batteries beyond Li-ion and K-ion counterparts. Adv Energy Mater 2021;11:2002704.

15. Li G, Chen K, Wang Y, et al. Cream roll-inspired advanced MnS/C composite for sodium-ion batteries: encapsulating MnS cream into hollow N,S-co-doped carbon rolls. Nanoscale 2020;12:8493-501.

16. Hou H, Qiu X, Wei W, Zhang Y, Ji X. Carbon anode materials for advanced sodium-ion batteries. Adv Energy Mater 2017;7:1602898.

17. Chang H, Wu Y, Han X, Yi T. Recent developments in advanced anode materials for lithium-ion batteries. Energy Mater 2021; doi: 10.20517/energymater.2021.02.

18. Cao L, Liang X, Ou X, et al. Heterointerface engineering of hierarchical Bi2S3/MoS2 with self-generated rich phase boundaries for superior sodium storage performance. Adv Funct Mater 2020;30:1910732.

19. Xiong X, Yang C, Wang G, et al. SnS nanoparticles electrostatically anchored on three-dimensional N-doped graphene as an active and durable anode for sodium-ion batteries. Energy Environ Sci 2017;10:1757-63.

20. Yang C, Liang X, Ou X, et al. Heterostructured nanocube-shaped binary sulfide (SnCo)S2 interlaced with S-doped graphene as a high-performance anode for advanced Na+ batteries. Adv Funct Mater 2019;29:1807971.

21. Ou X, Cao L, Liang X, et al. Fabrication of SnS2/Mn2SnS4/carbon heterostructures for sodium-ion batteries with high initial coulombic efficiency and cycling stability. ACS Nano 2019;13:3666-76.

22. Chen K, Li G, Hu Z, et al. Simple preparation of baroque Mn-based chalcogenide/honeycomb-like carbon composites for sodium-ion batteries from renewable Pleurotus Eryngii. Energy Fules ;35:6265-71.

23. Wu C, Dou SX, Yu Y. The state and challenges of anode materials based on conversion reactions for sodium storage. Small 2018;14:e1703671.

24. Lu Y, Yu L, Lou XW. Nanostructured conversion-type anode materials for advanced lithium-ion batteries. Chem 2018;4:972-96.

25. Chae S, Ko M, Kim K, Ahn K, Cho J. Confronting issues of the practical implementation of Si anode in high-energy lithium-ion batteries. Joule 2017;1:47-60.

26. Yang Z, Song Y, Zhang C, et al. Porous 3D silicon-diamondyne blooms excellent storage and diffusion properties for Li, Na, and K ions. Adv Energy Mater 2021;11:2101197.

27. Ke C, Liu F, Zheng Z, et al. Boosting lithium storage performance of Si nanoparticles via thin carbon and nitrogen/phosphorus co-doped two-dimensional carbon sheet dual encapsulation. Rare Met 2021;40:1347-56.

28. Cheng X, Shao R, Li D, et al. A self-healing volume variation three-dimensional continuous bulk porous bismuth for ultrafast sodium storage. Adv Funct Mater 2021;31:2011264.

29. Lei K, Wang J, Chen C, et al. Recent progresses on alloy-based anodes for potassium-ion batteries. Rare Met 2020;39:989-1004.

30. Imtiaz S, Amiinu IS, Xu Y, Kennedy T, Blackman C, Ryan KM. Progress and perspectives on alloying-type anode materials for advanced potassium-ion batteries. Mater Today 2021;48:241-69.

31. Niu J, Zhang Z, Aurbach D. Alloy anode materials for rechargeable Mg ion batteries. Adv Energy Mater 2020;10:2000697.

32. Wu X, He G, Ding Y. Dealloyed nanoporous materials for rechargeable post-lithium batteries. ChemSusChem 2020;13:3287.

33. 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.

34. Chen Q, Ding Y, Chen M. Nanoporous metal by dealloying for electrochemical energy conversion and storage. MRS Bull 2018;43:43-8.

35. Artymowicz D, Erlebacher J, Newman R. Relationship between the parting limit for de-alloying and a particular geometric high-density site percolation threshold. Philos Mag 2009;89:1663-93.

36. Feng J, Zhang Z, Ci L, Zhai W, Ai Q, Xiong S. Chemical dealloying synthesis of porous silicon anchored by in situ generated graphene sheets as anode material for lithium-ion batteries. J Power Sources 2015;287:177-83.

37. Zhang H, An W, Song H, et al. Synthesis of micro-sized porous antimony via vapor dealloying for high-performance Na-ion battery anode. Solid State Ionics 2020;352:115365.

38. Sohn M, Lee DG, Park H, Park C, Choi J, Kim H. Microstructure controlled porous silicon particles as a high capacity lithium storage material via dual step pore engineering. Adv Funct Mater 2018;28:1800855.

39. Lin N, Li T, Han Y, Zhang Q, Xu T, Qian Y. Mesoporous hollow Ge microspheres prepared via molten-salt metallothermic reaction for high-performance Li-storage anode. ACS Appl Mater Interfaces 2018;10:8399-404.

40. Yang Y, Liu S, Bian X, Feng J, An Y, Yuan C. Morphology- and porosity-tunable synthesis of 3D nanoporous SiGe alloy as a high-performance lithium-ion battery anode. ACS Nano 2018;12:2900-8.

41. Gao H, Niu J, Zhang C, Peng Z, Zhang Z. A dealloying synthetic strategy for nanoporous bismuth-antimony anodes for sodium ion batteries. ACS Nano 2018;12:3568-77.

42. Yan Y, Shi Y, Wang Z, Qin C, Zhang Y. AlF3 microrods modified nanoporous Ge/Ag anodes fabricated by one-step dealloying strategy for stable lithium storage. Mater Letters 2020;276:128254.

43. Ma W, Yin K, Gao H, Niu J, Peng Z, Zhang Z. Alloying boosting superior sodium storage performance in nanoporous tin-antimony alloy anode for sodium ion batteries. Nano Energy 2018;54:349-59.

44. Zhang H, Zhang M, Zhang M, et al. Hybrid aerogel-derived Sn-Ni alloy immobilized within porous carbon/graphene dual matrices for high-performance lithium storage. J Colloid Interface Sci 2017;501:267-72.

45. Yin H, Xiao W, Mao X, Zhu H, Wang D. Preparation of a porous nanostructured germanium from GeO2 via a “reduction-alloying-dealloying” approach. J Mater Chem A 2015;3:1427-30.

46. Yuan Y, Xiao W, Wang Z, Fray DJ, Jin X. Efficient nanostructuring of silicon by electrochemical alloying/dealloying in molten salts for improved lithium storage. Angew Chem Int Ed Engl 2018;57:15743-8.

47. Wada T, Ichitsubo T, Yubuta K, Segawa H, Yoshida H, Kato H. Bulk-nanoporous-silicon negative electrode with extremely high cyclability for lithium-ion batteries prepared using a top-down process. Nano Lett 2014;14:4505-10.

48. Wada T, Yamada J, Kato H. Preparation of three-dimensional nanoporous Si using dealloying by metallic melt and application as a lithium-ion rechargeable battery negative electrode. J Power Sources 2016;306:8-16.

49. Chen Y, Yuan Y, Xu C, et al. Multi-step low-cost synthesis of ultrafine silicon porous structures for high-reversible lithium-ion battery anodes. J Mater Sci 2020;55:13938-50.

50. Wang J, Huang W, Kim YS, et al. Scalable synthesis of nanoporous silicon microparticles for highly cyclable lithium-ion batteries. Nano Res 2020;13:1558-63.

51. An W, Gao B, Mei S, et al. Scalable synthesis of ant-nest-like bulk porous silicon for high-performance lithium-ion battery anodes. Nat Commun 2019;10:1447.

52. An Y, Tian Y, Wei H, et al. Porosity- and graphitization-controlled fabrication of nanoporous silicon@carbon for lithium storage and its conjugation with MXene for lithium-metal anode. Adv Funct Mater 2019;30:1908721.

53. Míguez H, Meseguer F, López C, et al. Germanium FCC structure from a colloidal crystal template. Langmuir 2000;16:4405-8.

54. Song T, Jeon Y, Samal M, et al. A Ge inverse opal with porous walls as an anode for lithium ion batteries. Energy Environ Sci 2012;5:9028.

55. Geier S, Jung R, Peters K, Gasteiger HA, Fattakhova-rohlfing D, Fässler TF. A wet-chemical route for macroporous inverse opal Ge anodes for lithium ion batteries with high capacity retention. Sustainable Energy Fuels 2018;2:85-90.

56. Esmanski A, Ozin GA. Silicon inverse-opal-based macroporous materials as negative electrodes for lithium ion batteries. Adv Funct Mater 2009;19:1999-2010.

57. Jeong J, Kim K, Jung D, Kim K, Lee S, Oh E. High-performance characteristics of silicon inverse opal synthesized by the simple magnesium reduction as anodes for lithium-ion batteries. J Power Sources 2015;300:182-9.

58. Gowda SR, Pushparaj V, Herle S, et al. Three-dimensionally engineered porous silicon electrodes for Li-ion batteries. Nano Lett 2012;12:6060-5.

59. Bao Z, Weatherspoon MR, Shian S, et al. Chemical reduction of three-dimensional silica micro-assemblies into microporous silicon replicas. Nature 2007;446:172-5.

60. Hwa Y, Kim W, Yu B, Hong S, Sohn H. Mesoporous nano-Si anode for Li-ion batteries produced by magnesio-mechanochemical reduction of amorphous SiO2. Energy Technol 2013;1:327-31.

61. Wang B, Li W, Wu T, Guo J, Wen Z. Self-template construction of mesoporous silicon submicrocube anode for advanced lithium ion batteries. Energy Stor Mater 2018;15:139-47.

62. Entwistle J, Rennie A, Patwardhan S. A review of magnesiothermic reduction of silica to porous silicon for lithium-ion battery applications and beyond. J Mater Chem A 2018;6:18344-56.

63. Jia H, Kloepsch R, He X, et al. Reversible storage of lithium in three-dimensional macroporous germanium. Chem Mater 2014;26:5683-8.

64. Lin N, Han Y, Zhou J, et al. A low temperature molten salt process for aluminothermic reduction of silicon oxides to crystalline Si for Li-ion batteries. Energy Environ Sci 2015;8:3187-91.

65. Manukyan KV, Schools RS, Mukasyan AS. Size-tunable germanium particles prepared by self-sustaining reduction of germanium oxide. J Solid State Chem 2019;270:92-7.

66. Peterson AK, Morgan DG, Skrabalak SE. Aerosol synthesis of porous particles using simple salts as a pore template. Langmuir 2010;26:8804-9.

67. Dai F, Zai J, Yi R, et al. Bottom-up synthesis of high surface area mesoporous crystalline silicon and evaluation of its hydrogen evolution performance. Nat Commun 2014;5:3605.

68. Tang D, Yu H, Zhao J, et al. Bottom-up synthesis of mesoporous germanium as anodes for lithium-ion batteries. J Colloid Interface Sci 2020;561:494-500.

69. Ge M, Fang X, Rong J, Zhou C. Review of porous silicon preparation and its application for lithium-ion battery anodes. Nanotechnology 2013;24:422001.

70. Huang Z, Geyer N, Werner P, de Boor J, Gösele U. Metal-assisted chemical etching of silicon: a review. Adv Mater 2011;23:285-308.

71. Peng K, Wu Y, Fang H, Zhong X, Xu Y, Zhu J. Uniform, Axial-orientation alignment of one-dimensional single-crystal silicon nanostructure arrays. Angew Chem 2005;117:2797-802.

72. Li X, Yan C, Wang J, et al. Stable silicon anodes for lithium-ion batteries using mesoporous metallurgical silicon. Adv Energy Mater 2015;5:1401556.

73. Bai F, Li M, Song D, Yu H, Jiang B, Li Y. One-step synthesis of lightly doped porous silicon nanowires in HF/AgNO3/H2O2 solution at room temperature. J Solid State Chem 2012;196:596-600.

74. Hochbaum AI, Gargas D, Hwang YJ, Yang P. Single crystalline mesoporous silicon nanowires. Nano Lett 2009;9:3550-4.

75. Zhong X, Qu Y, Lin YC, Liao L, Duan X. Unveiling the formation pathway of single crystalline porous silicon nanowires. ACS Appl Mater Interfaces 2011;3:261-70.

76. Zhang Z, Wang Y, Ren W, Tan Q, Zhong Z, Su F. Low-cost synthesis of porous silicon via ferrite-assisted chemical etching and their application as Si-based anodes for Li-ion batteries. Adv Electron Mater 2015;1:1400059.

77. Cao M, Li SY, Deng JX, et al. Preparation of large-area porous silicon through Cu-assisted chemical etching. MSF 2016;847:78-83.

78. Rezvani SJ, Pinto N, Boarino L. Rapid formation of single crystalline Ge nanowires by anodic metal assisted etching. Cryst Eng Comm 2016;18:7843-8.

79. Han X, Zhang Z, Chen H, et al. Bulk boron doping and surface carbon coating enabling fast-charging and stable Si anodes: from thin film to thick Si electrodes. J Mater Chem A 2021;9:3628-36.

80. Uhlir A. Electrolytic shaping of germanium and Silicon. Bell Syst Tech J 1956;35:333-47.

81. Thakur M, Sinsabaugh SL, Isaacson MJ, Wong MS, Biswal SL. Inexpensive method for producing macroporous silicon particulates (MPSPs) with pyrolyzed polyacrylonitrile for lithium ion batteries. Sci Rep 2012;2:795.

82. Turner DR. Electropolishing silicon in hydrofluoric acid solutions. J Electrochem Soc 1958;105:402.

83. Bioud YA, Boucherif A, Belarouci A, et al. Fast growth synthesis of mesoporous germanium films by high frequency bipolar electrochemical etching. Electrochim Acta 2017;232:422-30.

84. Bang BM, Kim H, Lee J, Cho J, Park S. Mass production of uniform-sized nanoporous silicon nanowire anodes via block copolymer lithography. Energy Environ Sci 2011;4:3395.

85. Jia H, Zheng J, Song J, et al. A novel approach to synthesize micrometer-sized porous silicon as a high performance anode for lithium-ion batteries. Nano Energy 2018;50:589-97.

86. Jeong YK, Huang W, Vilá RA, et al. Microclusters of kinked silicon nanowires synthesized by a recyclable iodide process for high-performance lithium-Ion battery anodes. Adv Energy Mater 2020;10:2002108.

87. Jia H, Li X, Song J, et al. Hierarchical porous silicon structures with extraordinary mechanical strength as high-performance lithium-ion battery anodes. Nat Commun 2020;11:1474.

88. Yao R, Xie L, Wu Y, Meng W, He Y, Zhao D. Controllable self-assembled mesoporous silicon nanocrystals framework as anode material for Li-ion battery. Electrochim Acta 2021;390:138850.

89. Liu X, Giordano C, Antonietti M. .

90. Xiang B, An W, Fu J, et al. Graphene-encapsulated blackberry-like porous silicon nanospheres prepared by modest magnesiothermic reduction for high-performance lithium-ion battery anode. Rare Met 2021;40:383-92.

91. Chevrier VL, Zwanziger JW, Dahn JR. First principles studies of silicon as a negative electrode material for lithium-ion batteries. Can J Phys 2009;87:625-32.

92. Shenoy V, Johari P, Qi Y. Elastic softening of amorphous and crystalline Li-Si Phases with increasing Li concentration: a first-principles study. J Power Sources 2010;195:6825-30.

93. Zuo X, Zhu J, Müller-buschbaum P, Cheng Y. Silicon based lithium-ion battery anodes: a chronicle perspective review. Nano Energy 2017;31:113-43.

94. Shen L, Guo X, Fang X, Wang Z, Chen L. Magnesiothermically reduced diatomaceous earth as a porous silicon anode material for lithium ion batteries. J Power Sources 2012;213:229-32.

95. Mu T, Shen B, Lou S, et al. Scalable mesoporous silicon microparticles composed of interconnected nanoplates for superior lithium storage. Chem Eng J 2019;375:121923.

96. He W, Tian H, Xin F, Han W. Scalable fabrication of micro-sized bulk porous Si from Fe-Si alloy as a high performance anode for lithium-ion batteries. J Mater Chem A 2015;3:17956-62.

97. Han X, Zhang Z, Zheng G, et al. Scalable engineering of bulk porous Si anodes for high initial efficiency and high-areal-capacity lithium-ion batteries. ACS Appl Mater Interfaces 2019;11:714-21.

98. Jia H, Gao P, Yang J, Wang J, Nuli Y, Yang Z. Novel three-dimensional mesoporous silicon for high power lithium-ion battery anode material. Adv Energy Mater 2011;1:1036-9.

99. Zhou Y, Jiang X, Chen L, et al. Novel mesoporous silicon nanorod as an anode material for lithium ion batteries. Electrochim Acta 2014;127:252-8.

100. Tang J, Yin Q, Wang Q, et al. Two-dimensional porous silicon nanosheets as anode materials for high performance lithium-ion batteries. Nanoscale 2019;11:10984-91.

101. Bang BM, Lee J, Kim H, Cho J, Park S. High-performance macroporous bulk silicon anodes synthesized by template-free chemical etching. Adv Energy Mater 2012;2:878-83.

102. Li X, Gu M, Hu S, et al. Mesoporous silicon sponge as an anti-pulverization structure for high-performance lithium-ion battery anodes. Nat Commun 2014;5:4105.

103. Zhang Z, Wang Y, Ren W, et al. Scalable synthesis of interconnected porous silicon/carbon composites by the rochow reaction as high-performance anodes of lithium ion batteries. Angew Chem 2014;126:5265-9.

104. Zhong H, Zhan H, Zhou Y. Synthesis of nanosized mesoporous silicon by magnesium-thermal method used as anode material for lithium ion battery. J Power Sources 2014;262:10-4.

105. Choi S, Bok T, Ryu J, Lee J, Cho J, Park S. Revisit of metallothermic reduction for macroporous Si: compromise between capacity and volume expansion for practical Li-ion battery. Nano Energy 2015;12:161-8.

106. Liang J, Li X, Hou Z, Guo C, Zhu Y, Qian Y. Nanoporous silicon prepared through air-oxidation demagnesiation of Mg2Si and properties of its lithium ion batteries. Chem Commun (Camb) 2015;51:7230-3.

107. Jiang T, Zhang R, Yin Q, et al. Morphology, composition and electrochemistry of a nano-porous silicon versus bulk silicon anode for lithium-ion batteries. J Mater Sci 2017;52:3670-7.

108. Liu X, Miao R, Yang J, et al. Scalable and cost-effective preparation of hierarchical porous silicon with a high conversion yield for superior lithium-ion storage. Energy Technol 2016;4:593-9.

109. Jia H, Gao P, Yang J, Wang J, Nuli Y, Yang Z. Novel three-dimensional mesoporous silicon for high power lithium-ion battery anode material. Adv Energy Mater 2011;1:1036-9.

110. Xu Z, Gang Y, Garakani MA, Abouali S, Huang J, Kim J. Carbon-coated mesoporous silicon microsphere anodes with greatly reduced volume expansion. J Mater Chem A 2016;4:6098-106.

111. Kim Y, Lee J, Kim H. Nanoporous silicon flakes as anode active material for lithium-ion batteries. Physica E Low Dimens Syst Nanostruct 2017;85:223-6.

112. Kim N, Park H, Yoon N, Lee JK. Zeolite-templated mesoporous silicon particles for advanced lithium-ion battery anodes. ACS Nano 2018;12:3853-64.

113. Wang F, Zhao B, Zi W, Du H. Ionothermal synthesis of crystalline nanoporous silicon and its use as anode materials in lithium-ion batteries. Nanoscale Res Lett 2019;14:196.

114. Yang Z, Du Y, Hou G, Ouyang Y, Ding F, Yuan F. Nanoporous silicon spheres preparation via a controllable magnesiothermic reduction as anode for Li-ion batteries. Electrochim Acta 2020;329:135141.

115. Sun X, Huang H, Chu K, Zhuang Y. Anodized macroporous silicon anode for integration of lithium-ion batteries on chips. J Elec Mater 2012;41:2369-75.

116. Wang F, Sun L, Zi W, Zhao B, Du H. Solution synthesis of porous silicon particles as an anode material for lithium ion batteries. Chemistry 2019;25:9071-7.

117. Wada T, Kato H. Preparation of nanoporous Si by dealloying in metallic melt and its application for negative electrode of lithium ion battery. Mater Today: Proc 2017;4:11465-9.

118. Ge M, Lu Y, Ercius P, et al. Large-scale fabrication, 3D tomography, and lithium-ion battery application of porous silicon. Nano Lett 2014;14:261-8.

119. Gao P, Tang H, Xing A, Bao Z. Porous silicon from the magnesiothermic reaction as a high-performance anode material for lithium ion battery applications. Electrochim Acta 2017;228:545-52.

120. Chen X, Bi Q, Sajjad M, et al. One-dimensional porous silicon nanowires with large surface area for fast charge-discharge lithium-ion batteries. Nanomaterials (Basel) 2018;8:285.

121. Guo R, Zhang S, Ying H, Yang W, Wang J, Han WQ. New, Effective, and low-cost dual-functional binder for porous silicon anodes in lithium-ion batteries. ACS Appl Mater Interfaces 2019;11:14051-8.

122. Wang W, Favors Z, Ionescu R, et al. Monodisperse porous silicon spheres as anode materials for lithium ion batteries. Sci Rep 2015;5:8781.

123. 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.

124. Entwistle JE, Patwardhan SV. Enabling scale-up of mesoporous silicon for lithium-ion batteries: a systematic study of a thermal moderator. RSC Adv 2021;11:3801-7.

125. Zeng Y, Huang Y, Liu N, et al. N-doped porous carbon nanofibers sheathed pumpkin-like Si/C composites as free-standing anodes for lithium-ion batteries. J Energy Chem 2021;54:727-35.

126. Shen C, Ge M, Luo L, et al. In situ and ex situ TEM study of lithiation behaviours of porous silicon nanostructures. Sci Rep 2016;6:31334.

127. Liu N, Lu Z, Zhao J, et al. A pomegranate-inspired nanoscale design for large-volume-change lithium battery anodes. Nat Nanotechnol 2014;9:187-92.

128. Shi J, Zu L, Gao H, Hu G, Zhang Q. Silicon-based self-assemblies for high volumetric capacity Li-ion batteries via effective stress management. Adv Funct Mater 2020;30:2002980.

129. Choi JH, Kim HK, Jin EM, et al. Facile and scalable synthesis of silicon nanowires from waste rice husk silica by the molten salt process. J Hazard Mater 2020;399:122949.

130. Kim H, Han B, Choo J, Cho J. Three-dimensional porous silicon particles for use in high-performance lithium secondary batteries. Angew Chem 2008;120:10305-8.

131. Tao Y, Zeng G, Xiao C, Liu Y, Qian Y, Feng J. Porosity controlled synthesis of nanoporous silicon by chemical dealloying as anode for high energy lithium-ion batteries. J Colloid Interface Sci 2019;554:674-81.

132. Wang J, Liao L, Lee HR, et al. Surface-engineered mesoporous silicon microparticles as high-Coulombic-efficiency anodes for lithium-ion batteries. Nano Energy 2019;61:404-10.

133. Morito H, Yamada T, Ikeda T, Yamane H. Na-Si binary phase diagram and solution growth of silicon crystals. JAlloys and Compd 2009;480:723-6.

134. Xu Y, Swaans E, Basak S, Zandbergen HW, Borsa DM, Mulder FM. Reversible Na-ion uptake in Si nanoparticles. Adv Energy Mater 2016;6:1501436.

135. Han Y, Lin N, Xu T, et al. An amorphous Si material with a sponge-like structure as an anode for Li-ion and Na-ion batteries. Nanoscale 2018;10:3153-8.

136. Lim C, Huang T, Shao P, et al. Experimental study on sodiation of amorphous silicon for use as sodium-ion battery anode. Electrochim Acta 2016;211:265-72.

137. Qiu DF, Ma X, Zhang JD, Lin ZX, Zhao B. Mesoporous silicon microspheres produced from in situ magnesiothermic reduction of silicon oxide for high-performance anode material in sodium-ion batteries. Nanoscale Res Lett 2018;13:275.

138. Du FH, Li B, Fu W, Xiong YJ, Wang KX, Chen JS. Surface binding of polypyrrole on porous silicon hollow nanospheres for Li-ion battery anodes with high structure stability. Adv Mater 2014;26:6145-50.

139. Yi Z, Lin N, Zhao Y, et al. A flexible micro/nanostructured Si microsphere cross-linked by highly-elastic carbon nanotubes toward enhanced lithium ion battery anodes. Energy Stor Mater 2019;17:93-100.

140. Son Y, Kim N, Lee T, et al. Calendering-compatible macroporous architecture for silicon-graphite composite toward high-energy lithium-ion batteries. Adv Mater 2020;32:e2003286.

141. Park MH, Kim K, Kim J, Cho J. Flexible dimensional control of high-capacity Li-ion-battery anodes: from 0D hollow to 3D porous germanium nanoparticle assemblies. Adv Mater 2010;22:415-8.

142. Yang L, Gao Q, Li L, Tang Y, Wu Y. Mesoporous germanium as anode material of high capacity and good cycling prepared by a mechanochemical reaction. Electrochem Commun 2010;12:418-21.

143. Ke FS, Mishra K, Jamison L, et al. Tailoring nanostructures in micrometer size germanium particles to improve their performance as an anode for lithium ion batteries. Chem Commun (Camb) 2014;50:3713-5.

144. Liang J, Li X, Hou Z, et al. Honeycomb-like macro-germanium as high-capacity anodes for lithium-ion batteries with good cycling and rate performance. Chem Mater 2015;27:4156-64.

145. Yoon T, Song G, Harzandi A, et al. Intramolecular deformation of zeotype-borogermanate toward a three-dimensional porous germanium anode for high-rate lithium storage. J Mater Chem A 2018;6:15961-7.

146. Kwon D, Ryu J, Shin M, et al. Synthesis of dual porous structured germanium anodes with exceptional lithium-ion storage performance. J Power Sources 2018;374:217-24.

147. Choi S, Kim J, Choi NS, Kim MG, Park S. Cost-effective scalable synthesis of mesoporous germanium particles via a redox-transmetalation reaction for high-performance energy storage devices. ACS Nano 2015;9:2203-12.

148. Zhang C, Lin Z, Yang Z, et al. Hierarchically designed germanium microcubes with high initial coulombic efficiency toward highly reversible lithium storage. Chem Mater 2015;27:2189-94.

149. Liu X, Lin N, Cai W, et al. Mesoporous germanium nanoparticles synthesized in molten zinc chloride at low temperature as a high-performance anode for lithium-ion batteries. Dalton Trans 2018;47:7402-6.

150. Choi S, Cho YG, Kim J, et al. Mesoporous germanium anode materials for lithium-ion battery with exceptional cycling stability in wide temperature range. Small 2017;13:1603045.

151. Mishra K, Liu X, Ke F, Zhou X. Porous germanium enabled high areal capacity anode for lithium-ion batteries. Compos B Eng 2019;163:158-64.

152. Sosa AN, González I, Trejo A, Miranda Á, Salazar F, Cruz-Irisson M. Effects of lithium on the electronic properties of porous Ge as anode material for batteries. J Comput Chem 2020;41:2653-62.

153. Kohandehghan A, Cui K, Kupsta M, et al. Activation with Li enables facile sodium storage in germanium. Nano Lett 2014;14:5873-82.

154. Lu X, Adkins ER, He Y, et al. Germanium as a sodium ion battery material: in situ TEM reveals fast sodiation kinetics with high capacity. Chem Mater 2016;28:1236-42.

155. Li M, Wang Z, Fu J, Ma K, Detsi E. In situ electrochemical dilatometry study of capacity fading in nanoporous Ge-based Na-ion battery anodes. Scr Mater 2019;164:52-6.

156. Yi Z, Lin N, Li T, Han Y, Li Y, Qian Y. Meso-porous amorphous Ge: synthesis and mechanism of an anode material for Na and K storage. Nano Res 2019;12:1824-30.

157. Song T, Yan M, Qian M. A dealloying approach to synthesizing micro-sized porous tin (Sn) from immiscible alloy systems for potential lithium-ion battery anode applications. J Porous Mater 2015;22:713-9.

158. Ryu S, Shim HC, Song JT, et al. High-pressure evaporation-based nanoporous black Sn for enhanced performance of lithium-ion battery anodes. Part Part Syst Charact 2019;36:1800331.

159. Cook JB, Detsi E, Liu Y, et al. Nanoporous tin with a granular hierarchical ligament morphology as a highly stable Li-ion battery anode. ACS Appl Mater Interfaces 2017;9:293-303.

160. Wang JW, Liu XH, Mao SX, Huang JY. Microstructural evolution of tin nanoparticles during in situ sodium insertion and extraction. Nano Lett 2012;12:5897-902.

161. Kim C, Lee K, Kim I, et al. Long-term cycling stability of porous Sn anode for sodium-ion batteries. J Power Sources 2016;317:153-8.

162. Wang L, Ni Y, Lei K, Dong H, Tian S, Li F. 3D porous tin created by tuning the redox potential acts as an advanced electrode for sodium-ion batteries. Chem Sus Chem 2018;11:3376-81.

163. Kim IT, Allcorn E, Manthiram A. Cu6Sn5-TiC-C nanocomposite anodes for high-performance sodium-ion batteries. J Power Sources 2015;281:11-7.

164. Zhang B, Rousse G, Foix D, Dugas R, Corte DA, Tarascon JM. Microsized Sn as advanced anodes in glyme-based electrolyte for Na-ion batteries. Adv Mater 2016;28:9824-30.

165. Liang S, Cheng Y, Zhu J, Xia Y, Müller-buschbaum P. A chronicle review of nonsilicon (Sn, Sb, Ge)-based lithium/sodium-ion battery alloying anodes. Small Methods 2020;4:2000218.

166. Detsi E, Petrissans X, Yan Y, et al. Tuning ligament shape in dealloyed nanoporous tin and the impact of nanoscale morphology on its applications in Na-ion alloy battery anodes. Phys Rev Materials 2018:2.

167. Guo M, Meng W, Zhang X, et al. Electrochemical behavior and self-organization of porous Sn nanocrystals@acetylene black microspheres in lithium-ion half cells. Appl Surf Sci 2019;470:36-43.

168. Cheng Y, Huang J, Li R, et al. Enhanced cycling performances of hollow Sn compared to solid Sn in Na-ion battery. Electrochim Acta 2015;180:227-33.

169. Shi L, Wang W. Synthesis and sodium storage performance of Sb porous nanostructure. J Alloys Compd 2020;846:156369.

170. Li M, Qiu T, Foucher AC, et al. Impact of hierarchical nanoporous architectures on sodium storage in antimony-based sodium-ion battery anodes. ACS Appl Energy Mater 2020;3:11231-41.

171. Yuan Y, Jan S, Wang Z, Jin X. A simple synthesis of nanoporous Sb/C with high Sb content and dispersity as an advanced anode for sodium ion batteries. J Mater Chem A 2018;6:5555-9.

172. Ma W, Wang J, Gao H, et al. A mesoporous antimony-based nanocomposite for advanced sodium ion batteries. Energy Stor Mater 2018;13:247-56.

173. Liu S, Feng J, Bian X, Liu J, Xu H. The morphology-controlled synthesis of a nanoporous-antimony anode for high-performance sodium-ion batteries. Energy Environ Sci 2016;9:1229-36.

174. Hou H, Jing M, Yang Y, et al. Sb porous hollow microspheres as advanced anode materials for sodium-ion batteries. J Mater Chem A 2015;3:2971-7.

175. Hou H, Jing M, Yang Y, Zhu Y, Fang L, Song W, Pan C, Yang X, Ji X. Sodium/lithium storage behavior of antimony hollow nanospheres for rechargeable batteries. ACS Appl Mater Interfaces 2014;6:16189-96.

176. Rodriguez JR, Hamann HJ, Mitchell GM, Ortalan V, Pol VG, Ramachandran PV. Three-Dimensional Antimony Nanochains for Lithium-Ion Storage. ACS Appl Nano Mater 2019;2:5351-5.

177. Meng W, Guo M, Chen J, Li D, Wang Z, Yang F. Porous Sb with three-dimensional Sb nanodendrites as electrode material for high-performance Li/Na-ion batteries. Nanotechnology 2020;31:175401.

178. Li H, Wang K, Zhou M, et al. Facile Tailoring of Multidimensional Nanostructured Sb for Sodium Storage Applications. ACS Nano 2019;13:9533-40.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Li G, Guo S, Xiang B, Mei S, Zheng Y, Zhang X, Gao B, Chu PK, Huo K. Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries. Energy Mater 2022;2:200020. http://dx.doi.org/10.20517/energymater.2022.24

AMA Style

Li G, Guo S, Xiang B, Mei S, Zheng Y, Zhang X, Gao B, Chu PK, Huo K. Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries. Energy Materials. 2022; 2(3): 200020. http://dx.doi.org/10.20517/energymater.2022.24

Chicago/Turabian Style

Li, Gaojie, Siguang Guo, Ben Xiang, Shixiong Mei, Yang Zheng, Xuming Zhang, Biao Gao, Paul K. Chu, Kaifu Huo. 2022. "Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries" Energy Materials. 2, no.3: 200020. http://dx.doi.org/10.20517/energymater.2022.24

ACS Style

Li, G.; Guo S.; Xiang B.; Mei S.; Zheng Y.; Zhang X.; Gao B.; Chu PK.; Huo K. Recent advances and perspectives of microsized alloying-type porous anode materials in high-performance Li- and Na-ion batteries. Energy Mater. 2022, 2, 200020. http://dx.doi.org/10.20517/energymater.2022.24

About This Article

© The Author(s) 2022. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
41157
Downloads
11355
Citations
Comments
0
6

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.

0
Download PDF
Cite This Article 33 clicks
Like This Article 6 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Energy Materials
ISSN 2770-5900 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/