Operando elucidation of all six reversible Li–Ga phase transitions, inverted hysteresis, and interfacial dynamics in a nanoconfined CMK-3/Ga anode

Mano Ajayana, Che-an Linbc and Shih-kang Lin*abde
aProgram on Smart and Sustainable Manufacturing, Academy of Innovative Semiconductor and Sustainable Manufacturing, National Cheng Kung University, Tainan, 70101, Taiwan. E-mail: linsk@mail.ncku.edu.tw
bDepartment of Materials Science and Engineering, National Cheng Kung University, Tainan, 70101, Taiwan
cLaboratory for Chemistry and Life Science, Institute of Integrated Research, Institute of Science Tokyo, Japan
dHierarchical Green-Energy Materials (Hi-GEM) Research Center, National Cheng Kung University, Tainan City, 70101, Taiwan
eCore Facility Center, National Cheng Kung University, Tainan City 70101, Taiwan

Received 15th February 2025 , Accepted 24th July 2025

First published on 24th July 2025


Abstract

Stabilizing phase transitions in alloy-type anodes remains a critical challenge for advancing high-performance lithium-ion batteries (LIBs). Herein, we present a gallium-infused mesoporous carbon composite (CMK-3/Ga), synthesized via a facile ball-milling method, as a promising anode material for LIBs. Leveraging the intrinsic properties of gallium(Ga) and its nanoscale confinement within CMK-3, we unveil unprecedented electrochemical and structural insights into Li–Ga alloying dynamics. For the first time in a carbon-confined system, comprehensive operando cyclic voltammetry (CV) combined with in situ X-ray diffraction (XRD) enables complete identification of all six Li–Ga alloying and dealloying phases (L1–L6 and D1–D6), with direct structural validation. This high-resolution mapping significantly advances the understanding of complex alloy electrochemistry. Notably, we observe a unique inverted hysteresis behavior, where delithiation initiates at lower potentials than lithiation, converging at a distinct low-voltage “pinning point” (∼0.05 V). This kinetically favorable pathway is directly correlated with pronounced structural rearrangements observed via in situ XRD, providing new mechanistic insights into lithium extraction processes. Furthermore, we identified the spontaneous formation of a reversible CuGa2 interphase at the current collector. This self-regulating buffer layer, validated by operando CV-EIS through suppressed Cu2O/CuO formation and enhanced kinetics, effectively maintains electrode contact and reduces capacity degradation. The optimized CMK-3/Ga anode delivered a high reversible capacity of 527 mA h g−1 over 100 cycles, exhibiting enhanced rate capability and significantly lower irreversible capacity loss, outperforming pristine CMK-3. This work provides real-time mechanistic understanding of robust alloy anodes, advancing the rational design of high-performance, self-protective energy storage materials.


1. Introduction

The development of lithium-ion batteries (LIBs) has been crucial for advancing portable electronics, electric vehicles, and renewable energy storage due to their high energy density and long cycle life.1,2 Despite their widespread use, conventional graphite anodes, with a theoretical specific capacity of 372 mA h g−1, present inherent drawbacks for future advanced LIB applications. These include significant irreversible capacity loss and poor cycling stability, which limit the overall performance of LIBs. The primary causes of large irreversible capacity loss in anode materials, such as active material detachment due to volume change, particle cracking leading to conductive network disruption, and solid electrolyte interphase (SEI) formation that consumes active Li ions and increases reaction impedance, must be addressed for the development of advanced LIBS. As a result, there is an increased demand for more durable and high energy density anode materials.3 Several advanced carbon-based materials have emerged as promising anodic materials for next-generation high-performance LIBs.4 Such materials often exhibit good electronic conductivity, a large surface area, and an organized porous structure which are ideal characteristics for achieving high specific capacities. These qualities also contribute to increased energy density, longer cycle life, cost-effectiveness, superior reversible capacity, and exceptional cycling performance.

Among various carbons, mesoporous carbon having pore size within the range of 2–50 nm is an excellent material to enhance lithium-ion diffusion kinetics both within the pores and along the pore walls during intercalation.5 Recent studies have also demonstrated that various element-doped mesoporous carbon materials deliver exceptional cycling stability, making them promising candidates for advanced lithium-ion battery applications.6,7 Highly ordered mesoporous carbons like CMK-3 have emerged as promising alternatives, offering high surface area, tunable pore structure, chemical stability, and advantageous textural properties. While such materials offer potential improvements for future LIBs, they also encounter challenges, including initial irreversible capacity loss, poor cycling stability, surface reactivity during cycling8–12 and weakness in volume expansion.13,14 Although CMK-3 is known to be a rigid structure, it is not entirely immune to volume changes during lithiation and delithiation. This can lead to microcracks, resulting in partial or total structural breakdown.14 Carbon-based anode materials typically undergo approximately a 12% volume change during lithiation and delithiation. This, combined with SEI formation and ion transportation challenges, contributes to cycling instability.3,15

Recent advances have proposed dual-confinement strategies and surface engineering approaches to address these structural and interfacial challenges. For instance, Sn4P3 particles embedded within mesoporous CMK-3 and coated with PEDOT have demonstrated enhanced stability and rate performance by simultaneously mitigating volume expansion and SEI accumulation.16 Similarly, conversion-type anodes such as Co9S8 integrated into carbon nanofiber networks show improved conductivity, reversible capacity, and structural resilience, as demonstrated through in situ XRD and magnetometry techniques.17 Moreover, single-atom modifications of hard carbon, such as Ni–N, P co-doped frameworks, have been shown to significantly improve both sodium storage kinetics and interfacial stability by tuning the SEI composition and promoting ion diffusion.18 Recent studies not only offer promising solutions but also underscore that performance degradation arising from electrode instability and volume changes remains one of the most prominent challenges in battery development. These volume changes, along with SEI formation, can also lead to irreversible structural alterations, complex phase formations, and lithium trapping within the carbon network due to sluggish ion release kinetics, ultimately resulting in significant initial capacity loss and long-term performance degradation.3 Repeated electrode expansion and contraction during cycling induce microcracks, exposing the current collector to the electrolyte and accelerating surface degradation, further compromising electrochemical stability and conductivity.19–23 As Guo and colleagues emphasize, such degradation is driven by interactions with the electrolyte, mechanical stress, and other contributing factors that can impact the battery's overall performance.24 Building on this, the presence of hybrid organic/inorganic electrolytes or active materials can lead to electrochemically induced interface amorphization, a significant aspect of cell degradation, directly impacting performance.25

The investigation of lithium intercalation phases in carbon anodes focuses on both ordered and disordered structures to assess their structural changes.26 Furthermore, broader reviews have provided valuable insight into these challenges. For instance, Electrochem. Energy Rev. has critically summarized the root cause of low initial coulombic efficiency (ICE) in carbonaceous anodes and outlined strategies for overcoming these issues based on structural and interfacial control.27 Additionally, emerging deposition techniques like atomic layer deposition (ALD) have been studied as powerful tools to engineer ultrathin protective films on electrode surfaces, further enhancing cycling stability and interface durability.28

Recent research has highlighted the potential of integrating liquid metals (LMs) such as gallium(Ga) with carbon to enhance battery performance.29–33 Ga, a liquid metal, has attracted considerable interest due to its low melting point and lower toxicity. Ga can alloy with many metals including lithium,34–36 thereby enhancing the anodic capacity and safety by reducing volume expansion which is advantageous to LIBs.37,38

Deshpande and colleagues investigated the electrochemical lithiation of a Ga film, revealing that it can accommodate two Li atoms per Ga atom when fully lithiated. They demonstrated its potential as high-capacity recovery anode in LIBs, due to its electronic conductivity in both liquid and solid states.39,40 Ga also exhibits a high theoretical volumetric capacity (4545 mA h cm−3) due to its high density at ambient temperature (5.91 g cm−3), making it a potential candidate for anode material in batteries.41 Owing to these aforementioned benefits, Ga has been extensively explored in combination with carbon-based anode materials to improve overall performance. Wang et al. demonstrated the fabrication of carbon fibers with Ga nanodroplets encapsulated using ultrasonic and electrospinning methods.42 These nano-sized Ga nanoparticles lead to excellent self-healing processes that maintain electrode structural stability for long-term cycles.43 The integration of Ga within a porous matrix has also been reported in the literature as a strategy to prevent irreversible capacity loss, Ga melting, agglomeration, and cracking during LIB cycling.44,45

In bulk Ga, the long diffusion length of Li-ion results in poor rate performance. However, the direct use of liquid metals as active material presents challenges of practical application due to the need for specialized current collectors. To overcome these issues, developing innovative and constructive integration strategies is necessary for the future adaptation of liquid metals in LIBs. Thus, a critical study of the structural and textural properties of Ga-based composite anode materials, along with their alloying–de-alloying mechanisms elucidated through in situ measurements, is essential to further improve LIB performance.40,41,44,46,47 Despite these efforts in material design, a fundamental challenge persists in the complex electrochemical behavior of gallium itself.

The electrochemical behavior of gallium is inherently complex, governed by a series of lithium-rich intermetallic phases that form across a wide compositional window. Foundational insights into Li–Ga intermetallic compounds date back to pioneering thermodynamic studies by Schneider and Hilmer in 1956,48 which laid the groundwork for understanding intermetallic formation, including the important NaTl-type Li–Ga phase. More recently, comprehensive thermodynamic reassessments, notably by Azza et al.49 using advanced CALPHAD methods, have rigorously established the existence of up to eight distinct Li–Ga intermetallic compounds, each forming at characteristic potential and compositions. Such meticulous thermodynamic modelling of phase equilibria and transformations is critical for understanding complex material behavior under various conditions, including novel phenomena like the nano-volcanic eruption of silver driven by Ag–O interactions and transient phase changes, as explored through CALPHAD-type analysis.50 While electrochemical studies on bulk systems, such as detailed investigations by Saint et al.,51 have partially confirmed many of these transformations, and even recent work on carbon-dispersed liquid Ga–In eutectic alloys has revealed a strikingly similar array of CV peaks indicative of multiple complex alloying reactions within a ternary system,52 the comprehensive, real-time electrochemical identification and sequential resolution of all these intricate binary Li–Ga phases within carbon-confined Ga composites (derived from liquid Ga precursors) has remained largely unexplored.

Previous investigations on Ga-based composites, while confirming phases such as Li3Ga14, LiGa, and Li2Ga, largely depend on ex situ characterization,41 which hides transient or overlapping intermediates, such as Li5Ga4 and Li3Ga2. The lack of operando insights not only limits understanding of the full electrochemical alloying pathway but also hinders rational design of reversible, high-capacity anodes. Moreover, intriguing kinetic signatures like inverted hysteresis, “where delithiation occurs at lower potentials than lithiation,” remain unexplored for intermediate phases such as Li3Ga2 and Li5Ga4. Such an inversion can arise from dynamic changes in interface kinetics or internal stress evolution within the confined material, where repeated volume changes and Li-ion diffusion drive structural rearrangement and defect annihilation to lower overall strain energy. This may signal an overlooked barrier to reversibility or a unique energetic structural rearrangement intrinsic to confined systems. Indeed, understanding and mitigating such complex electrochemical degradation mechanisms, including those driven by interfacial dynamics, are critical for battery stability, as highlighted by work on thermal recovery in all-solid-state lithium batteries.25

Despite the theoretical foundation, the electrochemical realization and sequential identification of all six key phases, Li3Ga14, Li5Ga9, LiGa, Li5Ga4, Li3Ga2, and Li2Ga have never been comprehensively demonstrated in carbon-supported architecture. To date, no report has revealed this full progression within a single, continuous operando experiment. This gap leaves critical questions unanswered: Can all Li–Ga intermetallics emerge electrochemically in a confined matrix, and what governs their transitions under real cycling conditions?

Herein, we present a definitive electrochemical and structural mapping of the Li–Ga alloying pathway within mesoporous CMK-3 confined Ga composites. We meticulously investigate the structural and textural properties of Ga-based composites with varying Ga/CMK-3 mass ratios, achieved through ball milling, including surface area analysis, and morphological characterization (SEM, TEM, EDX). Our research further examines the structural properties of infused Ga via low-temperature X-ray diffraction, observing its transition from amorphous to solid below room temperature (300–150 K). Through a comprehensive electrochemical performance characterization, including cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), cycling stability, and rate capability, we demonstrate the robust performance of these composites. Post-cycling analysis, including varying loading conditions, provides further insights into long-term stability and capacity retention.

Crucially, this study employs a multi-modal operando strategy to elucidate the intricate alloying–de-alloying mechanisms. We report the first operando in situ X-ray diffraction (XRD) analysis of a Ga-confined mesoporous carbon CMK-3 composite anode during electrochemical galvanostatic charge/discharge (GCD) cycling, providing real-time insights into its structural evolution over an extended period (65 hours, 10 cycles), specifically monitoring the formation of the Li–Ga alloy and its subsequent dealloying phases. Such a reversible process monitored over this duration represents one of the longest in situ analyses in liquid metal-supported carbon composite anode LIB research to date. Complementing this, sequential operando cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) were performed for various Ga/CMK-3 ratios (e.g., C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) to provide comprehensive insights into evolving redox processes and interfacial impedance characteristics over multiple cycles. These synergistic operando approaches provide unprecedented real-time insights into both the dynamic structural evolution and the intricate electrochemical mechanisms within the anode.

Our findings show that the CMK-3/Ga composite can deliver a reversible 527 mA h g−1 capacity at a current rate of 100 mA g−1 over 100 cycles, exhibiting discharge capacities of 232 and 147 mA h g−1 at the 840th cycle under high current densities of 1600 and 3200 mA g−1, respectively. Furthermore, a capacity recovery phenomenon is observed over a prolonged electrochemical cycling process. For the first cycle, the coulombic efficiency is 52%, and it can be enhanced to 67% by further increasing the amount of Ga. Ultimately, this breakthrough not only establishes a new benchmark for phase-tracking in alloy anodes by resolving the complete six-phase lithiation/delithiation sequence and uncovering evidence of inverted hysteresis but also offers mechanistic clarity essential for advancing gallium-based storage systems, demonstrating CMK-3/Ga as a promising electrode for future energy storage applications.

2. Experimental

2.1. Materials

The triblock copolymer poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(ethylene glycol) (EO20PO70EO20, Pluronic P123, M.wt. 5800), tetraethyl orthosilicate (TEOS), sucrose, sulfuric acid (H2SO4) and hydrochloric acid (HCl; 37 wt%), hydrofluoric acid and gallium(Ga) metal grade 99.9% were used for the fabrication of mesoporous carbon/Ga materials. Lithium foil metal grade 99.9%, N-methyl-2-pyrrolidone, poly(vinylidene fluoride, super p carbon black, electrolyte with lithium hexafluorophosphate (LiPF6) as the Li-salt in ethylene carbonate (EC)–diethyl carbonate (DEC) (1[thin space (1/6-em)]:[thin space (1/6-em)]1 by volume) solvent were used for the electrode material and coin cell preparation.

2.2. Material synthesis

2.2.1. Mesoporous silica SBA-15. Hexagonally ordered mesoporous silica SBA-15 was synthesized using a specific amphiphilic triblock copolymer, Pluronic P123 following the synthesis procedure reported by Zhao et al.8 A typical synthesis was performed by dispersing 4 g of the copolymer in 30 g of water and stirring the mixture for 3 h. Afterwards, 120 g of 2 M HCl solution was introduced, the temperature was raised to 40 °C, and the mixture was stirred for an additional 2 h. Then 9 g of TEOS was added, and the mixture was continuously stirred for 24 h at 40 °C before heating it for 48 h at 100 °C in a polypropylene (PP) bottle. The resulting solid product was filtered, washed multiple times with water, and dried overnight in an oven at 100 °C. Adjusting the synthesis reaction temperature makes it possible to regulate the pore diameter of the materials. Three samples were synthesized, with temperature variations set at 100, 130, and 150 °C in a Teflon-lined autoclave. These samples are labeled as SBA-15-X, with X representing the respective synthesis temperature. Finally, the polymer template was removed via calcination at 540 °C. The synthesized SBA-15 materials were used for the modification of mesoporous carbon CMK-3.
2.2.2. Mesoporous carbon CMK- 3. Mesoporous carbons with varying pore diameters, referred to as CMK-3-X, were synthesized using SBA-15-X materials as templates and sucrose as the carbon source following the synthesis procedure reported by Jun et al.9 The synthesized materials were named CMK-3-100, CMK-3-130, and CMK-3-150, respectively. In a typical synthesis, 1 g of SBA-15 template was added to a solution containing 1.25 g of sucrose, 0.14 g of H2SO4, and 5 g of water. The mixture was then heated in an oven at 100 °C for 6 h, followed by an additional 6 h at 160 °C. A mixture of 0.8 g sucrose, 0.09 g H2SO4, and 5 g water was added to the pre-treated sample to facilitate the complete polymerization and carbonization of sucrose within the silica template pores. This mixture was also subjected to the same thermal treatment as previously described. The polymer–template composites were pyrolyzed at 900 °C for 5 h in a nitrogen atmosphere to carbonize the polymer. To remove silica from the carbonized samples, these were treated with the 5 wt% hydrofluoric acid, and then mesoporous carbon was recovered by filtration after dissolving the silica framework. The product was washed with ethanol several times and dried at 100 °C. The materials obtained are called mesoporous carbons, and the samples are designated as CMK-3-X, where X indicates the synthesis temperature of the silica templates.
2.2.3. Preparation of Ga-infused CMK-3 (CMK-3/Ga). Gallium(Ga) metal of 99.9% purity (metal grade) was used as the active material precursor. CMK-3, a mesoporous carbon, synthesized at various temperatures (e.g., 100 °C, 130 °C, 150 °C), served as the carbon host matrix. Zirconia (ZrO2) milling balls (5 mm diameter, ∼0.4 g each) and a Teflon milling jar were employed for composite fabrication. In a typical process, CMK-3/Ga composites were prepared via a unique liquid-phase incorporation followed by mechanical milling, a method amenable to various carbon host materials. In a typical procedure, gallium metal, solid at room temperature, was gently heated to ∼60 °C to achieve complete melting (Ga melting point ≈ 29.8 °C). Precise quantities of this molten gallium were then carefully weighed and directly poured onto pre-weighed amounts of CMK-3 powder within a Teflon ball milling jar. The Ga/CMK-3 mass ratios were systematically varied to 0.5[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1.5[thin space (1/6-em)]:[thin space (1/6-em)]1, 2[thin space (1/6-em)]:[thin space (1/6-em)]1, and 3[thin space (1/6-em)]:[thin space (1/6-em)]1 (where the CMK-3 weight remained constant for each series, and the Ga amount was adjusted accordingly). The two components were mixed through a controlled ball milling procedure using four 5 mm zirconia balls. The milling process was conducted at 400 revolutions per minute (rpm) for a total duration of 2 hours. To ensure homogeneous dispersion and prevent material adhesion to the jar walls, the milling process was interrupted twice: after the initial 30 minutes of milling, the jar was opened, and any material adhered to the walls was meticulously scraped down and mixed back into the bulk using a spatula. This mixing step was repeated after an additional 1 hour of milling (i.e., after 90 minutes total milling time). The milling was then resumed for the final 30 minutes. This intermittent mixing strategy, combined with the initial liquid-phase incorporation of gallium, facilitates the effective infusion of Ga into the mesoporous structure of CMK-3, leading to the formation of stable and homogeneously dispersed CMK-3/Ga composites. The resulting CMK-3/Ga composites were subsequently characterized using a suite of techniques to evaluate their morphology, porosity, crystallinity, and electrochemical performance. This fabrication method offers a robust and scalable route for embedding low-melting-point metals into conductive carbon hosts, crucial for next-generation Li-ion storage applications.

2.3. Material characterization

The textural properties of the CMK-3/Ga composites were analyzed using N2 adsorption isotherms that were obtained on a Micromeritics ASAP 2020, Micromeritics Instrument Corp., USA. The adsorption isotherms were used to calculate the specific surface area, pore size, and pore volume.

Wide-angle XRD patterns were collected using two different diffractometers: a D-8 Bruker (Bruker AXS GmbH, Karlsruhe, Germany) with Cu Kα radiation (λ = 1.542 Å, operating at 40 kV and 25 mA) and a Panalytical Empyrean diffractometer utilizing Cu Kα radiation (λ = 1.542 Å). These measurements were conducted to identify crystalline phases and observe structural change outcomes from Ga infusion into CMK-3. Small Angle X-ray Scattering (SAXS) was recorded using the NanoStar U SAXS system (Bruker AXS GmbH, Karlsruhe, Germany) to provide information on the mesostructure of the materials. Before this measurement, all samples were stained with Ag+ ions, a VANTEC-2000 detector equipped with the Cu Kα radiation with a wavelength of λ = 1.542 Å was provided by an Incoatec Microfocus source (IμS) operating at 45 kV and 650 μA. Low-temperature XRD patterns were recorded with a D8 Discover X-ray diffraction system (Bruker AXS GmbH, Karlsruhe, Germany) utilizing Cu Kα radiation (λ = 1.542 Å) at various temperatures (150 K, 200 K, 250 K, 280 K, and 300 K) to analyze phase transitions under different thermal conditions.

The morphological characteristics before and after Ga infusion, post-cycling electrode surface changes, and elemental distribution of the CMK-3/Ga composite were visualized through Scanning Electron Microscopy (SEM) using a Hitachi SU-3500 (Japan). High-resolution field-emission Transmission Electron Microscopy (TEM) images, Energy Dispersive X-ray (EDX) mapping, and diffraction patterns were obtained using a JEOL JEM-2100F TEM (Japan) operating at 200 kV. These techniques provided detailed insights into the structural and compositional changes at the nanoscale. The post-analysis discharge/charge cycling rate performance, coin cells of pristine CMK-3 and CMK-3/Ga ratios of 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 were disassembled, and the surface was washed in dimethyl sulfoxide (DMSO) to remove the electrolyte residue and dried in an argon-filled glove box until the sample surface was completely dried and then characterized through an SEM Hitachi SU-3500 (Japan).

2.4. Electrochemical measurements

The mixture containing active material CMK-3/Ga, super P carbon black, and polyvinyl fluoride (PVDF) binder was used to form a slurry at the weight ratio 80[thin space (1/6-em)]:[thin space (1/6-em)]10[thin space (1/6-em)]:[thin space (1/6-em)]10. The electrode was prepared by casting the slurry on copper (Cu) foil using a doctor blade followed by heating at 120 °C in a vacuum oven overnight. A similar process was carried out to fabricate the pristine CMK-3 electrode. The thickness of the coating on Cu was roughly 50 μm. The electrode was cut into circular pieces with a diameter of about 12 mm for coin cell testing. Li-ion batteries were assembled in CR2032 stainless steel cell with Li metal foil (about 12 mm in diameter) as the counter electrode, 1 M LiPF6 in a mixture of ethylene carbonate (EC)–diethyl carbonate (DEC) (EC–DEC, 1[thin space (1/6-em)]:[thin space (1/6-em)]1 by volume) as the electrolyte, and the porous polypropylene film Celgard®2400 (Celgard, LLC Corp., USA) as the separator inside an Ar-filled glove box with O2 and H2O less than 0.5 ppm.

The cyclic voltammogram (CV) conducted at a scan rate of 0.1 mV s−1 within the potential range of 0.01–3.0 V (versus Li/Li+) was used to investigate the reduction and oxidation processes of CMK-3/Ga, and electrochemical impedance spectroscopy (EIS) with 1 MHz–10 mHz frequency range was employed to confirm the electrochemical resistance of the CMK-3/Ga anode using a Bio-Logic VSP 300. The assembled standard half-cell was electrochemically tested at a cycling rate 100 mA g−1 except for rate capability tests through a galvanostatic charge/discharge process using a Neware battery testing system (BTS4000, Neware Technology Limited).

2.4.1. Operando in situ XRD. To evaluate the phase transformation of gallium during lithiation/delithiation, in situ X-ray diffraction (XRD) techniques were performed during the electrochemical charge–discharge process. In the cell assembly, a side-by-side electrode arrangement was utilized into an ECC-Opto-Std-Aqu electrochemical test cell holder from EL-CELL test equipment GmbH, Hamburg, Germany, with a beryllium window top cover to pass the X-ray without interference, and 1 M LiPF6 in EC[thin space (1/6-em)]:[thin space (1/6-em)]DEC (1[thin space (1/6-em)]:[thin space (1/6-em)]1 volume ratio) electrolyte was used. Detailed in situ cell preparation settings of real-time photos and video are shown in Fig. S1 of the ESI. The CMK-3/Ga sample was charged and discharged with a potential window of 0.01–3 V at a constant current density of 100 mA g−1. At the same time, XRD measurement was continuously recorded in the 2θ range from 10° to 80° at a scan rate of 0.5° min−1 for the entire range, and the measurement was repeated for several hours.

In situ XRD patterns were collected using the D-8 Bruker (Bruker AXS GmbH, Karlsruhe, Germany) diffractometer with Cu Kα radiation (λ = 1.542 Å, operating at 40 kV and 25 mA); at the same time the galvanostatic charge/discharge process was performed with the Bio-Logic SP-300. The in situ technique allows for the direct characterization of phase changes in electrode materials during real-time electrochemical measurements.

2.4.2. Operando sequential CV-EIS protocol. The dynamic electrochemical behavior of the CMK-3/Ga anode, including the evolution of reduction/oxidation processes and electrochemical resistance, was comprehensively investigated over multiple cycles within a single experimental run using a sequential Cyclic Voltammetry (CV) and Electrochemical Impedance Spectroscopy (EIS) protocol. This operando methodology allowed for a real-time assessment of evolving interfacial kinetics and charge transfer characteristics in direct correlation with the redox activity throughout continuous cycling, and importantly, provided direct correlation with the in situ XRD findings presented previously.

Coin cells, fabricated as described in the general electrochemical performance section, were used for these measurements. The programmed sequence for each cycle was meticulously designed. It commenced with a charge (lithium deintercalation/de-alloying) via Linear Sweep Voltammetry (LSV) from 0.01 V to 3.0 V (versus Li/Li+). An EIS measurement was then performed immediately at the fully charged state (3.0 V). Following this, a discharge (lithium intercalation/alloying) was conducted via LSV from 3.0 V to 0.01 V. Finally, another EIS measurement was performed immediately at the fully discharged state (0.01 V).

This “Charge–EIS–Discharge–EIS” sequence constituted one full electrochemical cycle and was precisely programmed to repeat for a total of 15 cycles. All CV scans were conducted at a consistent scan rate of 0.1 mV s−1 within the potential range of 0.01–3.0 V. EIS measurements were performed in the 1 MHz–10 mHz frequency range and with a 20 mV AC amplitude using a Bio-Logic VSP 300 potentiostat. This (LSV–PEIS–LSV–PEIS) sequence, looped for 15 cycles, provides an invaluable dataset for understanding the interplay between redox activity and impedance changes in real-time.

3. Results and discussion

3.1. Characterization of surface area and porosity

To evaluate the surface area, pore size, and pore volume before and after the nanoconfinement of Ga into CMK-3, the N2 adsorption isotherms were recorded at 77 K. Fig. 1a and c present the nitrogen adsorption–desorption isotherms and pore size distributions of pristine CMK-3-100, CMK-3-130, and CMK-3-150, respectively. These isotherms belong to type IV with H1 hysteresis loops, typical of mesoporous materials. The pore size distributions of all three materials fell within a range of 2–7 nm with a single maximum observed for each: 3.7 nm (CMK-3-100), 3.9 nm (CMK-3-130), and 5.0 nm (CMK-3-150).53,54 The increase in synthesis temperature led to an increase in surface area, pore size, and pore volume of CMK-3 materials,53 as shown in Table S1.
image file: d5ta01253h-f1.tif
Fig. 1 (a and b) Nitrogen adsorption–desorption isotherms of pristine CMK-3, and CMK-3/Ga. (c and d) Pore size distribution of the pristine CMK-3, and CMK-3/Ga.

The infusion of Ga into CMK-3-130 significantly decreased the surface area and pore volume of the materials; this reduction varied in direct proportion to the amount of Ga used for loading (Table S1 and Fig. 1b). The pore size distribution of these materials did not show any considerable change in the pore diameter (∼4.0 nm), although the overall volume of pores was lower in Ga-incorporated materials due to the blockage of pores with Ga (Fig. 1d).

The quantification of the textural features revealed that pristine CMK-3-130 possesses the highest micropore surface area and micropore volume compared to CMK-3-100 or CMK-3-150. Furthermore, the reduction in surface area and pore volume after confinement demonstrates similar effects as observed in other carbon materials used for metal confinement.55,56

Similar effects of Ga infusion were observed in CMK-3-100 and CMK-3-150. Fig. S2a and b present the N2 adsorption–desorption isotherms, and Fig. S2c and d show the pore size distribution of CMK-3-100 and CMK-3-150 with various mass loadings of Ga. Surface area and pore volumes decreased with increasing Ga impregnation of these materials (Table S1). CMK-3-100 did not show an appreciable change in the pore diameter after incorporation with Ga; however, CMK-3-150 showed reduced pore diameter from 5.0 nm to 4.6 nm in Ga loaded materials.

Building upon these prior investigations, including studies demonstrating CMK-3's capacity to host diverse guests, such as the adsorption of large molecules like proteins owing to its substantial pore size,53 our current study explores the feasibility of Ga infiltration into CMK-3 carbon using a ball milling process. Our research outcomes, together with the findings presented by Lee et al. regarding carbon black,41 collectively suggested favourable possibilities for the successful integration of Ga into carbon matrices. The diverse properties of carbon materials, including mesopores, micropores, high surface area, large pore volume, and tunable pore size distribution, have been identified as critical factors in facilitating nanoconfinement and improving electrochemical performance in various applications.44,47,53,54,57–59

3.2. Surface morphology analysis of CMK-3/Ga

SEM and TEM were used to observe the microstructure of Ga-infused CMK-3 before and after nanoconfinement. Fig. 2a and b shows similar morphology, and the rod-like pristine CMK-3-130 morphology remained intact after Ga nanoconfinement with the lowest amount (CMK-3-130–Ga 1:0.5).
image file: d5ta01253h-f2.tif
Fig. 2 (a) SEM images of pristine CMK-3. (b) CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5. TEM images of CMK-3/Ga (c) low-resolution, (d) high-resolution. (e) HAADF image of CMK-3/Ga. (f–h) TEM EDX mapping. (i) Diffraction pattern of CMK-3/Ga and inset circle in figure (e) show the diffraction spot area.

Fig. 2c and d show the TEM images of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, which indicate that Ga droplets (dark regions Ga marked with a circle) are evenly distributed into rod-like mesochannels of CMK-3-130. The presence of nanodroplets of Ga in CMK-3-130 was further observed through HAADF-STEM imaging (Fig. 2e), which revealed the consistent encapsulation of Ga nanodroplets (bright regions) into the matrix of CMK-3-130. TEM EDX mapping over CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 (Fig. 2f–h) confirms the uniform distribution of carbon, gallium, and oxygen. The electron diffraction image in Fig. 2i shows an amorphous phase diffraction pattern. The microscopic studies provide strong evidence for the presence of Ga in the form of amorphous nanodroplets. The successful confinement of Ga at the nanoscale in CMK-3 carries the potential to contribute to the kinetics of the electrochemical conversion reaction since it is known that nano-scale electrochemical active materials provide short Li-ion diffusion pathways and many reactive sites which are beneficial to good electrochemical reaction kinetics.

3.3. X-ray diffraction analysis

The incorporation of gallium (Ga) into CMK-3-100, 130, and 150 was further confirmed through wide-angle XRD. Fig. 3a shows wide-angle XRD of CMK-3-130 and CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5. CMK-3 showed two typical amorphous carbon bands which can be indexed as [002] and [100] planes.9 Two new broad peaks were observed for CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 which can be indexed to the [021] and [111] planes of monoclinic metallic Ga (JCPDS no. 73–0811) revealing that the Ga in CMK-3 was rather amorphous.39–42 The electron diffraction image in Fig. 2i shows an amorphous pattern that matches the XRD analyses (Fig. 3a) showing liquid-phase Ga. Similar observations are also encountered for all combinations of Ga with all mesoporous carbons (Fig. S3a–d).
image file: d5ta01253h-f3.tif
Fig. 3 (a) Wide angle XRD pattern of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5. (b) SAXS pattern of pristine CMK-3. (c) CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 ratios. (d) Temperature-dependent wide-angle XRD pattern of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 operated at temperatures ranging from 300 to 150 K. (e) Reversal of 150 to 300 K.

The XRD patterns show that the structure of CMK-3 remained intact for all mass loadings of Ga in all materials. However, in CMK-3-100, the peak intensity at ∼20° increased with the increasing amount of Ga, as shown in Fig. S3b which suggests that Ga may not be completely infused into the pores but rather accumulated on the surface of CMK-3-100. To observe structural distortion in the ordered mesoporous carbon hexagonal structure before and after adding Ga, small-angle XRD patterns were recorded. CMK-3-100, 130, and 150 show typical mesoporous carbon peaks indexed as [100], [110], and [200] which demonstrate an ordered mesoporous structure (Fig. 3b). The main peak in these materials occurs at 2θ positions of 1.08, 1.02, and 0.98, respectively; the minor shift corresponds to an increase in lattice spacing.53 Fig. 3c presents a comparison of small angle XRD of pristine CMK-3-130 and Ga-infused samples with various mass ratios, demonstrating that the CMK-3-130 structure remained intact, and the Ga infusion was successful. Fig. S3e and f shows the small-angle XRD CMK-3-100 and CMK-3-150 and their comparison with various mass ratios of Ga addition. The results demonstrated that the hexagonal structure of CMK-3-150 remained unchanged following the addition of Ga. However, CMK-3-100 exhibited changes in peak width and intensity with increasing amounts of Ga, indicating that Ga likely covers the particle surface, resulting in external structural modifications.60 Such changes are consistent with the wide-angle XRD pattern of CMK-3-100/Ga as shown in Fig. S3b.

While previous reports indicated that carbon-encapsulated Ga composites exist in an amorphous state,41,42,44 our study aimed to investigate deeper. We conducted temperature-dependent XRD analyses to investigate the potential for Ga's liquid-to-solid phase transition within a mesoporous carbon composite framework. Wide-angle XRD patterns were collected over the temperature range of 150–300 K to reveal the phase transition of Ga (Fig. 3d and e).

3.3.1. Low-temperature XRD analysis of gallium phase transitions in CMK-3/Ga composites (150–300 K). Our study demonstrated that amorphous liquid-phase Ga gradually transformed into solid-phase Ga as the temperature decreased, with the reverse transformation occurring upon heating. During the cooling process (Fig. 3d), as the temperature was reduced from 300 K to 150 K, the transition from liquid to solid-phase Ga occurred between 200 K and 150 K, where distinct crystalline peaks began to appear. This observation indicates that the nanoconfined gallium exhibits significant supercooling, crystallizing at temperatures substantially lower than bulk gallium's melting point, with notable crystalline peaks becoming prominent as the temperature approaches 150 K.

Conversely, during the heating process (Fig. 3e), as the temperature increased from 150 K to 300 K, a slow reversible phase transition from solid to liquid Ga was detected between 250 K and 280 K. This melting transition was characterized by the gradual broadening and eventual disappearance of crystalline peaks, signifying the complete liquefaction of gallium. The distinct temperature ranges for crystallization (200K–150 K) highlight a pronounced thermal hysteresis and the broad nature of the phase transitions, typical characteristics of nanoconfined materials.

These findings align with the results of Yarema et al., who reported that Ga nanoparticles (approximately 24 nm in size) remained liquid at room temperature and for several months of storage in the fridge (ca. −10 °C (263.15 K)).61 Our low-temperature analysis corroborates their findings, as we similarly observed Ga crystallite peaks between 200 K and 250 K during the cooling process.

The insights gained from this study are highly significant, as the CMK-3/Ga composite uniquely exhibits reversible amorphous-to-crystalline phase transition at low temperatures. This inherent thermomechanical property, where the nanoconfined gallium reversibly solidifies and melts over a specific temperature range, is crucial. It suggests that such Ga-infused mesoporous carbon could fundamentally enhance electrode structure stability, potentially reduce volume expansion through self-healing capability, prevent dendrite growth, and find novel applications in phase change devices.44,52,61 This distinctly demonstrates that the CMK-3/Ga composite material itself holds beneficial thermomechanical features, extending beyond its electrochemical performance, making it a compelling material for advanced battery architecture and thermal management applications.

3.4. Electrochemical characterization

3.4.1. EIS analysis of interfacial charge transfer and lithium-ion diffusion dynamics. Electrochemical impedance spectroscopy (EIS) was conducted to evaluate the charge transfer characteristics and ion transport dynamics of CMK-3-130 and CMK-3-130/Ga composites with varying Ga content (Fig. 4a). The Nyquist plots exhibit the typical features of porous carbon-based electrodes, with a high-frequency semicircle representing the combined resistance from the solid electrolyte interphase (SEI) layer formation and charge transfer at the electrode–electrolyte interface, and a sloping line in the low-frequency region corresponding to Warburg diffusion behavior, indicative of ion transport resistance within the mesoporous matrix. A partially enlarged view (inset of Fig. 4a) further highlighting the high-frequency region, shows a well-defined semicircle and the distinct onset of the Warburg region, features of the impedance interpretation.
image file: d5ta01253h-f4.tif
Fig. 4 Electrochemical characterization of CMK-3/Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 low loading composite anodes. (a) Nyquist EIS spectrum. (b) Second cycle cyclic voltammetry (CV) curves obtained at a scan rate of 0.1 mV s−1 within the potential window of 0.01 V to 3 V (vs. Li/Li+), with numbered peaks indicating identified Li–Ga alloying/de-alloying phase transitions: Lithiation (L#): L1: Ga → Li3Ga14 (∼0.7–0.71 V), L2: Li3Ga14 → Li5Ga9 (∼0.65 V), L3: Li5Ga9 → LiGa (∼0.59 V), L4: LiGa → Li5Ga4 (∼0.53 V), L5: Li5Ga4 → Li3Ga2 (∼0.44 V), L6: Li3Ga2 → Li2Ga (∼0.10–0.12 V); delithiation (D#): D1: Li2Ga → Li3Ga2 (initial), D2: Li3Ga2 → Li5Ga4 (∼0.30–0.31 V), D3: Li5Ga4 → LiGa (∼0.46–0.47 V), D4: LiGa → Li5Ga9 (∼0.75–0.77 V), D5: Li5Ga9 → Li3Ga14 (∼0.77–0.80 V), D6: Li3Ga14 → Ga (∼0.88–0.9 V) and the reversible A1 CuGa2 alloying interphase around 1.2 V. (c) Galvanostatic charge/discharge (GCD) cycling performance. (d) Rate capability.

The EIS spectrum of the pristine CMK-3-130 electrode (Fig. 4a, blue dot) displays a large semicircular arc in the high-frequency region, corresponding to a charge-transfer resistance of approximately 248 Ω. This elevated resistance suggests sluggish interfacial kinetics at the electrode–electrolyte interface, possibly arising from the formation of a resistive or unstable SEI layer. Additionally, the high surface area and defective pore architecture of CMK-3 may contribute to uneven current distribution and interfacial discontinuities that further hinder effective charge transfer. After the semicircle, the impedance response transitions into a short, nearly horizontal segment before gradually rising into a less steep Warburg slope. This two-step low-frequency behavior reflects the buildup of double-layer capacitance at pore entrances, followed by bulk ion diffusion within the mesoporous channels. The combination of high interfacial resistance and delayed ionic diffusion indicates that pristine CMK-3 suffered from limited charge transport kinetics, inefficient Li+ mobility, and a higher tendency toward side reactions, such as continuous SEI growth.

Although no equivalent circuit fitting was performed in this work, the observed impedance behavior aligns well with established models of mesoporous carbon-based electrodes.5 To support and validate our interpretations, we further refer to the work of Saikia et al.,62 where detailed Nyquist analysis of CMK-3 and CMK-8 electrodes was performed using equivalent circuit modeling. In that study, the authors attributed the depressed semicircles to the combined resistance of the SEI and charge-transfer processes, where the inclined Warburg line was linked to lithium-ion diffusion within mesoporous channels. Their results showed that differences in pore geometry significantly influenced both charge-transfer resistance and ion transport properties. Through correspondence, the large semicircle and shallow Warburg slope observed in our pristine CMK-3-130 sample are consistent with the diffusion and interfacial limitations described in their analysis, further confirming the interpretation of our general Nyquist features even in the absence of fitting.

The EIS spectrum reveals substantial changes in charge-transfer and ion transport characteristics upon the incorporation of Ga into CMK-3. For the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 composite (Fig. 4a, black dot), the Nyquist plot exhibits a significantly reduced semicircle diameter, corresponding to a lower combined resistance of the SEI layer and charge transfer (∼98 Ω). This marked decrease suggests that low Ga content facilitates enhanced charge-transfer kinetics, likely due to the optimal nanoconfinement of Ga within the mesoporous channels of CMK-3. Such confinement improves electronic connectivity across the carbon–Ga interface, forming more efficient electron transport pathways. In the low-frequency region, the impedance trace of the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 composite displays a well-defined and steeper Warburg slope, accompanied by a leftward shift in the impedance response. This indicates a shorter efficient diffusion pathway of Li+, attributed to the homogeneous dispersion of Ga within the interconnected mesoporous matrix. The combination of reduced resistance and more favorable ion diffusion pathways reflects a highly synergistic interaction between Ga and CMK-3 at this optimal loading, enabling rapid and efficient lithium-ion transport throughout the electrode structure.

In contrast, the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 composite (Fig. 4a, red dots) presents a moderately sized semicircle, corresponding to a higher interfacial resistance of approximately 213 Ω. While this value remains significantly lower than that of pristine CMK-3 (∼248 Ω), it is notably higher than that observed for the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 composite. The increased resistance at higher Ga loading suggests that excessive Ga may agglomerate or partially block the mesoporous channels of CMK-3, thereby diminishing electronic and ionic transport efficiency. Additionally, the low-frequency region of the 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 composite shows a rightward shift, and a steeper Warburg slope compared to the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 sample, indicating increased ion diffusion resistance. This suggests that Li+ mobility was hindered due to reduced accessibility of the active surface area, likely caused by the partial pore obstruction or aggregation of Ga, which elongates the effective diffusion pathways and limits transport through the porous network.

The critical role of Ga nanoconfinement in enhancing charge transport and diffusion was further supported by recent findings from Guo et al., who demonstrated that in room-temperature liquid metal alloy systems, improved Li+ transport kinetics and reduced interfacial resistance were strongly correlated with the alloy's structural homogeneity and phase composition. Specifically, they reported that the Li diffusion coefficient in Ga-based systems reached up to ∼1.01 × 10−7 cm2 s−1 when in a liquid phase, several orders of magnitude higher than in typical solid-state electrodes, typically due to favorable chemical environments and improved phase interfaces within the alloy matrix.52 Although their study was on bulk LM systems, it directly underscores how phase structure, alloying behavior, and interface design influence ionic transport. Our findings revealed that nanoconfined Ga within CMK-3 channels can likewise promote fast ion movement by preserving conductive and accessible pathways during alloying/de-alloying transitions.

Overall, EIS results underscore the critical role of Ga loading in modulating the charge transport and diffusion behavior of CMK-3/Ga electrodes. Optimal low Ga content (CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5) leads to significantly improved electrochemical kinetics, manifesting as reduced charge-transfer and diffusion resistances, thereby facilitating more effective Li+ diffusion kinetics. This precise control over charge transport and diffusion is intrinsically linked to the enhanced phase resolution observed in the subsequent Cyclic Voltammetry (CV) profiles (Fig. 4b). Improved kinetics evidenced in EIS for lower Ga loadings directly correlate with the ability to resolve distinct Li–Ga alloying/de-alloying transitions in CV, reinforcing the conclusion that an optimal balance of Ga nanoconfinement offers the favorable pathways for both electron and ion transport within the composite anode.

3.4.2. Cyclic voltammetry (CV) analysis. The study of lithium–gallium (Li–Ga) intermetallic phases has a rich history, with foundational research establishing the existence and properties of various Li–Ga compounds. Early pioneering calorimetric work, such as that by Schneider and Hilmer in 1956, provided critical thermodynamic insights into NaTl-type intermetallic phases, including Li–Ga, laying the groundwork for understanding these complex alloy systems.48 Subsequent comprehensive studies, including those by Saint et al.51 on the Li–Ga room temperature phase diagram and Azza et al.49 through thermodynamic reassessments, and the Li–Ga alloying sequence in electrochemical features explored by Deshpande et al.40 have further enriched our understanding of gallium-based alloys. Electrochemical investigations into gallium's behavior, including side reactions with copper current collectors to form phases like CuGa2, have also contributed significantly.63

Despite these advances, the electrochemical identification of the full spectrum of Li–Ga phases in carbon-based composites remains a significant challenge. Previous electrochemical analyses on similar Ga-containing carbon composites typically reported the formation of a limited number of phases, such as Li3Ga14, LiGa, and Li2Ga.41,42,44,45

To overcome these limitations and fully elucidate the redox reaction mechanisms and phase transformations occurring within the CMK-3/Ga anodes during lithiation and delithiation, operando Cyclic Voltammetry (CV) was conducted over a potential range of 0.01–3.0 V (vs. Li/Li+) at a scan rate of 0.1 mV s−1. Representative second-cycle CV profiles are presented in Fig. 4b, with the first and fifth cycles shown in Fig. S4 a and b for CMK-3-130–Ga1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 and CMK-3-130–Ga-1[thin space (1/6-em)]:[thin space (1/6-em)]1.5, respectively. The second cycle was chosen as it typically provides a more stable and representative profile of the reversible redox response, reflecting the material's behavior after initial SEI layer formation and stabilization of the electrode surface. For comparison, the CV data of the pristine CMK-3-130 are also presented (Fig. 4b, blue line). It exhibits broad reduction features (below 0.4 V) and oxidation peaks (0.1–0.5 V) characteristic of carbonaceous materials, which emerge during lithium insertion and extraction.11,62 However, CMK-3 itself possesses intrinsic drawbacks, such as a low initial reversible capacity ratio of approximately 34%11 and poor cycling performance, which limit its practical applications. The incorporation of active liquid metals, such as gallium (which forms several compounds with lithium, including Li3Ga14, Li5Ga9, LiGa, Li5Ga4, Li3Ga2, and Li2Ga51), presents a capable alternative for energy storage systems. The inherent phase transition from reversible liquid to solid in such metals can enhance the durability, cycle life, and stability of rechargeable alkali-ion batteries.

In stark contrast to pristine CMK-3, the successful incorporation of gallium into the CMK-3 mesopores is reflected in the distinct features of the CV profiles (CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, black dashed line; CMK-3-130-–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5, red line in Fig. 4b). The CV of the CMK-3/Ga composite reveals the formation of the SEI and clear electrochemical signatures corresponding to a complete set of Li–Ga alloying and de-alloying reactions observed across various Ga loadings (Fig. S4a and b).


3.4.2.1 Lithiation phase evolution. Our CV examination enabled the comprehensive identification of all six expected Li–Ga phases, owing to the effective nanoconfinement of Ga within the CMK-3 matrix. During the lithiation (reduction) process (Fig. 4b), a small peak at 1.2 V, annotated as A1, was observed prior to the main Li–Ga alloying. This feature is attributed to the interfacial alloying of Ga with the copper current collector, forming the CuGa2 phase. While barely visible in the first cycle, this A1 peak became more prominent from the second cycle onwards and exhibited a reversible counterpart during the subsequent delithiation process. This increased prominence suggests that after initial structural conditioning of the electrode in the first cycle, Ga becomes more relaxed and its penetration through surface defects is facilitated, leading to enhanced surface alloying. The improved electrochemical accessibility of Ga to the copper surface subsequently facilitates the formation of a protective CuGa2 alloying phase on the current collector. The low intensity of the A1 peak further indicated it was less likely a typical redox reaction and more characteristic of a localized surface alloying with Cu, a process potentially aided by the inherent self-healing properties of Ga (Fig. S4b).63

In the continuous progression of the reaction, CMK-3/Ga composites exhibit distinct cathodic peaks (L#) during lithiation, representing the sequential alloying formation of binary intermetallic Li–Ga phases. The high-intensity peak observed at ∼0.71 V annotated as L1, corresponds to the formation of Li3Ga14. This peak indicates a strong and rapid lithium uptake reaction at this potential. The prominent intensity of this initial peak indicates substantial lithium uptake into the Ga framework, marking the onset of Li–Ga alloy formation.

Subsequently, a series of closely spaced reduction transitions emerge at progressively lower potentials (∼0.65 V, ∼0.59 V, ∼0.53 V, and ∼0.44 V), corresponding to the stepwise formation of Li5Ga9 (L2), LiGa (L3), Li5Ga4 (L4), and Li3Ga2 (L5), respectively (Fig. 4b). The final reduction peak (L6), formation of the most lithiated Li2Ga phase, is prominently observed at ∼0.1 V for the low-Ga loading (CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5). For the higher-Ga loading (CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5), this peak appears at a slightly higher potential of ∼0.12 V with increased intensity (Fig. 4b, S4a and b).

Recent studies on carbon-dispersed liquid Ga–In eutectic alloys have reported a comparable series of CV features, reflecting similar complex alloying behaviors within the ternary Li–Ga–In system.52 Other reports also present analogous CV performance for Li–Ga systems,43 consistent with the established Li–Ga phase diagram.49,51


3.4.2.2 Delithiation phase transition and inverted hysteresis. During delithiation (oxidation, lithium extraction), our CV curves display four distinct de-alloying anodic peaks (D#) that are marked for delithiation, corresponding to the sequential transformation of the lithiated phases back towards Ga. The initial delithiation from the fully lithiated Li2Ga phase, annotated as D1, presents as a broad electrochemical feature. Its broadness suggests a solid–solution transition towards the subsequent less lithiated phase, indicating a kinetically accessible initial delithiation pathway. The lower potential peaks at ∼0.3 V and ∼0.46 V are attributed to the de-alloying of Li3Ga2 and Li5Ga4 phases annotated as D2 and D3, respectively. Interestingly, these delithiation potentials are lower than the corresponding lithiation potentials (0.44 V and 0.53 V, respectively), reflecting inverted hysteresis characteristics across all Ga loadings (Fig. 4b, S4a and b). The subsequent broader de-alloying peaks at ∼0.75–0.79 V are associated with the overlapping transitions of LiGa and Li5Ga9 (annotated as D4 and D5). The narrower, lower-intensity peak observed at ∼0.9 V, annotated as D6, corresponds to the dealloying of the Li3Ga14 phase,43,52 consistent with the established Li–Ga phase diagram.49,51

The consistency of redox peaks in successive CV cycles for low Ga loading CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 demonstrates the reproducibility of the electrochemical performance (Fig. S4a). In contrast, higher Ga loading CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 exhibits a gradual decrease in peak intensity in subsequent cycles (Fig. S4b). The notably larger peak intensity difference between the 1st and 5th cycles for the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 sample, compared to the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 configuration, strongly suggests that low Ga loadings provide superior reversibility due to enhanced nanoconfinement effects.47

Recent reports have highlighted the importance of nanoconfinement in other alloying-type anodes, such as Li–Si nanowires,64 small particle multiphase Li-alloys,65 and nano alloy n-Si,66 and fundamental studies have also illustrated the profound influence of nanoscale effects on material structural dynamics, as exemplified by phenomena such as the nano eruption of silver nanoparticles.50 Our comprehensive observation of all six binary intermetallic Li–Ga phases in our CMK-3/Ga composites notably distinguishes our work from previous studies on gallium-containing carbon composites, which typically report only three phases (e.g., Li3Ga14, LiGa, Li2Ga).44,45

By contrast, our operando CV and subsequent in situ XRD analyses provide unprecedented real-time insights into the full phase evolution during cycling, revealing the complete progression of all six binary intermetallic Li–Ga phases in both lithiation and delithiation. This enhanced resolution, we propose, is a direct consequence of the unique nanoconfinement offered by the ball-milled CMK-3/Ga composite, which stabilizes otherwise transient or difficult-to-detect intermediate phases like Li5Ga4 and Li3Ga2 that are often missed in bulk or less confined systems.

The electrochemical galvanostatic cycling performance and rate capability of the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 are presented in Fig. 4c and d respectively. At a current density of 100 mA g−1, this electrode delivers a discharge capacity of 400 mA h g−1 at the 120th cycle with 98% capacity retention (Fig. 4c). This highlights the remarkable long-term cycling stability achieved by incorporating Ga into the CMK-3 structure. Furthermore, Ga confinement significantly reduces the initial irreversible capacity, as evidenced by a substantial improvement in the initial coulombic efficiency (ICE) to 52% for the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 ratio, compared to 44% for CMK-3-130 (Table S2). The aforementioned result differs from pristine CMK-3, which typically exhibits high irreversible capacity during the initial cycle of 34%.11

Further investigation into the impact of Ga loading on (ICE) reveals that increasing Ga content generally reduces the first-cycle irreversible capacity. For instance, the CMK-3-130/Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]3 composite exhibits an ICE of 67% (Table S2). While higher Ga loadings may yield higher ICE values,45 our results demonstrate that low Ga loading consistently delivers superior long-term capacity retention across different mass ratios (Fig. S5). The enhanced ICE in the low Ga-loaded CMK-3-130/Ga composites is attributed to an optimal balance between surface area and active Ga sites, providing sufficient reaction zones while minimizing irreversible lithium loss and excessive SEI formation. This is corroborated by the stable CV response observed in Fig. S4a. Furthermore, the long-term cycling stability is supported by the presence of liquid-phase Ga, which likely facilitates a self-healing or structural recovery mechanism, maintaining electrode integrity over extended cycling.

In contrast, the higher Ga-loaded CMK-3-130/Ga composite displays a notably weakened CV intensity (Fig. S4b), indicating reduced electrochemical activity and corresponding capacity fade despite a slightly improved initial ICE in GCD measurements. The excessive Ga content may lead to pore blockage and reduced electrolyte accessibility, limiting active surface area and compromising structural robustness. Additionally, higher Ga content likely reduces the extent of liquid-phase Ga during operation, diminishing the self-repair behavior and accelerating electrode degradation.

Importantly, low Ga loading facilitates remarkable capacity recovery over extended periods of cycling, evident at both low and high current densities. For instance, CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 exhibits discharge capacities of 232 and 147 mA h g−1 at the 840th cycle even at high current densities of 1600 and 3200 mA g−1, respectively (Fig. S5a). The reason for cycling-induced capacity increment could be due to electrochemical grinding, which leads to smaller particle sizes, as observed in oxide-based anodes.67

While this study primarily focuses on the half-cell configuration to systematically investigate the Li–Ga alloying behavior and the influence of nanoconfined Ga within the CMK-3 framework, further investigation into full-cell configurations represents a critical step toward realizing practical applications. Pairing the CMK-3/Ga composite anode with a compatible cathode would enable the evaluation of overall cell performance, energy density, and cycling stability under more realistic conditions, particularly in systems with limited lithium reservoirs. Several parameters, such as cathode selection, electrolyte compatibility, optimized electrode design, and the stability of the Ga-based anode at higher operating potentials, will play pivotal roles in determining full-cell efficiency. We believe that the unique advantages of Ga nanoconfinement, including improved kinetics, structural integrity, and potential self-healing behavior, can be strategically extended to the full-cell architecture. These considerations are part of our ongoing and future research efforts aimed at translating the fundamental insights presented here into practical energy storage technologies.

Fig. 4d displays the rate capability performance of the material with the lowest loading of Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5) at different current densities ranging from 25 to 3200 mA g−1. Notably, following high-current density cycling, the capacity of CMK-3-130/Ga quickly recovers to 400 mA h g−1 upon reduction to 100 mA g−1, demonstrating excellent cycling stability. This robust rate performance is directly attributed to its rapid Li+ diffusion kinetics and strong structural resilience, which collectively ensure the maintenance of its good electronic and ionic conductivity.

Fig. S5b and c present the long-term cycling and rate capability performance of the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 composite. While the initial cycles exhibit relatively higher capacity, a gradual decline in capacity retention is observed over 200 cycles (Fig. S5b), indicating structural degradation during prolonged operation. The rate capability (Fig. S5c) further highlights this limitation, as the electrode delivers lower reversible capacities at different current densities ranging from 50 to 1600 mA g−1 and fails to recover its initial capacity upon returning to 100 mA g−1. This performance degradation is likely due to excessive Ga loading leading to agglomeration, pore blockage, and structural instability, which disrupt uniform Li-ion transport and electron conduction. These findings align with prior observations that reversible capacities in Ga-confined carbon materials tend to diminish with cycling at high Ga contents.45 In contrast, our results demonstrate that lower Ga loading enables more stable cycling and better capacity recovery. The homogeneous dispersion of Ga within the mesoporous CMK-3 framework at lower loading facilitates uniform Li alloying/de-alloying, mitigates volume fluctuations, and preserves the electrode architecture, ultimately contributing to enhanced long-term electrochemical performance. To further elucidate the influence of Ga content on electrochemical performance, GCD cycling and voltage profiles were evaluated for CMK-3-130 samples with varying Ga loadings (Fig. S5). Fig. S5d–g and h–k display the long-term cyclability and voltage profiles, respectively, at a current density of 100 mA g−1 for carbon-to-Ga ratios of 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5, and 1[thin space (1/6-em)]:[thin space (1/6-em)]2. Among these, the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 sample exhibited the best performance, maintaining a reversible specific capacity of ∼550 mA h g−1 after 120 cycles with a retention of 59%, along with clear capacity recovery during extended cycling (Fig. S5d). Corresponding voltage profiles (Fig. S5h) also show this capacity rebound behavior, which is attributed to the optimized nanoconfinement of Ga within the CMK-3 matrix. This effect is supported by HR-TEM images (Fig. 2c and d), indicating that finely dispersed Ga enables more uniform alloying reactions and enhances electrode structural stability and reversibility.

In contrast, higher Ga loadings (1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5, and 1[thin space (1/6-em)]:[thin space (1/6-em)]2) led to noticeable capacity fading during extended cycling (Fig. S5e–g), despite initially delivering higher first-cycle efficiencies. The associated charge–discharge profiles (Fig. S5i–k) reveal diminished voltage features and the capacity, suggesting that excessive Ga content may lead to agglomeration, limited ion-accessible surfaces, and structural instability. These effects hinder the long-term cycling stability and kinetic reversibility of the electrode. As shown in Fig. S5l, a clear performance improvement was observed for CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 over pristine CMK-3, confirming the beneficial role of nanoconfined Ga.

The galvanostatic discharge profiles for CMK-3/Ga composites consistently show a distinct plateau near 0.7 V, corresponding to the initial alloying of Ga with lithium Li3Ga14 (L1). Beyond this, a continuous sloping region was observed, reflecting multiple subsequent lithiation steps (e.g., Li5Ga9 → LiGa → Li5Ga4 → Li3Ga2 → Li2Ga). These later steps are not well-resolved in GCD due to their closely spaced potential and overlapping voltage regions. In contrast, CV measurement distinctly captured all six sequential lithiation (L1–L6) and delithiation (D1–D6) transitions, benefiting from higher sensitivity to subtle redox processes and sharper current response. Notably, the charge profiles during delithiation exhibit multiple well-defined plateaus, indicative of kinetically favorable stepwise dealloying reactions. This observation aligns with both the CV data and the established Li–Ga phase diagram reported by Saint et al.,51 and is consistent with our peak assignments detailed in Fig. 4b, confirming the reproducibility of Li–Ga phase transitions over repeated cycles.

The lithiation (alloying) process typically requires higher overpotentials due to the nucleation barriers and structural reordering associated with the formation of Li–Ga intermetallic phases, especially under nanoconfinement. Consequently, while CV can resolve individual alloying steps, the GCD lithiation process appears as a merged sloping profile due to kinetic limitations and potential overlap. In contrast, delithiation proceeds more readily, resulting in clearly separated charge plateaus in both CV and GCD, highlighting the kinetic asymmetry between lithiation and delithiation.

Interestingly, two of the Li–Ga phase transitions (Li3Ga2 and Li5Ga4) exhibit inverted hysteresis, where the delithiation potential is lower than the corresponding lithiation potential. This suggests reduced overpotential during delithiation, potentially arising from structural rearrangements, electrochemical annealing, or phase stabilization effects under nanoconfinement. Such behavior supports the notion that confined architecture not only improves reversibility but also enables more energetically favorable lithium extraction pathways.

In addition to the primary focus on CMK-3-130 variants, the electrochemical performance of Ga-infused composites synthesized using CMK-3-100 and CMK-3-150 templates was also systematically investigated at Ga loading ratios of 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 (Fig. S6) and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 (Fig. S7).

The galvanostatic charge–discharge (GCD) performance of CMK-3-100–Ga and CMK-3-150–Ga composites with 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 Ga loading was evaluated at a constant current density of 100 mA g−1 to assess long-term cycling performance. Corresponding voltage profiles were analysed at the 1st, 50th, and 100th cycles to monitor capacity evolution and structural stability. Rate capability was also investigated across a wide current density range (25–3200 mA g−1) to assess high-rate tolerance (Fig. S6).

Both electrodes exhibit excellent cycling performance over 100 cycles (Fig. S6a). CMK-3-100–Ga maintains a stable capacity of ∼300 mA h g−1, while CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 achieves a higher capacity of ∼400 mA h g−1 with 98% retention. Although CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 displays slight fluctuations in the early cycles, these quickly stabilize. This suggests a structural rearrangement induced by nanoconfined Ga, which contributes to enhanced long-term performance.

The GCD voltage profiles (Fig. S6b and c) support these observations. CMK-3-100–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 shows gradual capacity fading with minimal recovery, whereas CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 demonstrates clear capacity restoration and sustained performance. Both exhibit a distinct discharge plateau around 0.7 V followed by a sloping region, consistent with Li–Ga alloying. Charge profiles reveal multiple plateaus, particularly in CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, indicative of sequential and kinetically favourable dealloying processes.

Rate capability measurements (Fig. S6d and e) further highlight the electrodes' resilience. Upon returning to 100 mA g−1, CMK-3-100–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 and CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 recover capacities of ∼300 mA h g−1 and ∼400 mA h g−1, respectively, confirming robust Li+ transport and structural integrity under high-rate cycling. The superior performance of CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 is attributed to its larger pore size, which enhances electrolyte infiltration, facilitates Li+ diffusion, and effectively buffers volume changes during cycling.

We postulate that the initially observed capacity fluctuation in CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 stems from increased electrolyte decomposition or SEI formation over its larger surface area, followed by structural stabilization that enables subsequent capacity recovery. Despite some irregularities in early cycles, the long-term capacity retention and overall performance of CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 become comparable to the optimized CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 system. In contrast, CMK-3-100–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 lacks such capacity recovery, likely due to less effective Ga confinement and structural buffering stemming from its smaller pores.

While this study primarily focused on CMK-3-130 due to its optimized structural and electrochemical performance across various Ga loadings, the promising results from CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 highlight the significant potential of larger pore systems when combined with controlled Ga nanoconfinement. These findings underscore the complex interplay between pore architecture and Ga nanoconfinement in dictating overall electrochemical performance.

In contrast to the promising performance observed with lower Ga loadings (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, Fig. S6), CMK-3-100–Ga and CMK-3-150–Ga composites with a higher Ga loading ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 exhibited significantly poorer electrochemical performance (Fig. S7a–d). Similar to the behavior of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5, this reduced performance can be primarily attributed to structural instability within the electrode, likely exacerbated by the increased presence of excess Ga beyond optimal nanoconfinement. This observation underscores a consistent trend irrespective of the CMK-3 pore size, higher Ga content (e.g., 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) generally compromises the electrode's long-term stability and capacity retention.

The electrochemical performance of the CMK-3-130/Ga anode, characterized by its high initial coulombic efficiency, impressive reversible capacity (ranging from 400 to 550 mA h g−1 at 100 mA g−1), remarkable capacity recovery, and excellent rate capability, is among the best reported to date for similar materials.45

While the primary focus of this study was on CMK-3-130 due to its consistently optimized performance across varying Ga loadings, the extended investigation into CMK-3-100 and CMK-3-150 (particularly at 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 Ga loading; Fig. S6) has broadened our understanding of how pore architecture influences electrochemical behavior. The successful infusion of Ga into the carbon host's mesopores, along with partial surface attachment, plays a dynamic role in enhancing cycling performance over pristine CMK-3. Notably, low Ga loading effectively suppresses the severe volume expansion associated with alloying reactions while simultaneously promoting lithium-ion transport, as supported by consistent EIS, cycling, and rate performance data. These findings highlight the importance of pore size and Ga distribution, with CMK-3-130 and CMK-3-150 (featuring 4–5 nm pores) emerging as optimal frameworks for achieving stable, high-capacity alloying anodes through controlled nanoconfinement.

The unique confinement of Ga within the CMK-3 matrix creates a stable electrochemical environment that mitigates mechanical degradation and enhances electrode longevity. This structural refinement enables the reproducible formation and transformation of all six Li–Ga intermetallic phases, including the rare observation of inverted hysteresis behavior, an uncommon but insightful feature in alloying-type anodes, offering unique insight into their sequential progression during cycling. Beyond phase stabilization, nanoconfinement minimizes active material loss and suppresses side reactions, protects the copper current collector, thereby improving reversibility and capacity retention, especially at lower Ga contents. Moreover, it facilitates rapid charge transfer and efficient lithium-ion diffusion, resulting in superior rate capabilities. Together, these advantages demonstrate that the synergy between optimized Ga loading and mesoporous carbon architecture is critical for advancing high-energy-density lithium-ion batteries and establishes a clear direction for future full-cell development and practical integration.

To gain a deeper understanding of the observed electrochemical performance disparities, particularly the capacity degradation at higher Ga loadings, the subsequent sections delve into detailed post-cycling morphological and structural analyses. Specifically, scanning electron microscopy (SEM) was employed to investigate the microstructural integrity of pristine CMK-3 and CMK-3-130–Ga samples with varying loadings (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) after extended cycling. Furthermore, in situ X-ray diffraction (XRD) was conducted on both pristine CMK-3-130 and CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 to provide real-time insights into their phase evolution and structural stability during electrochemical cycling.

3.5. Post-cycling analysis and in situ XRD characterization

In the following section, we analyze post-cycling SEM observations and in situ XRD measurements to investigate morphological changes and structural transformations of Ga-loaded CMK-3 materials during electrochemical cycling.

To further elucidate the origins of capacity fading, post-cycling SEM analysis was conducted on the electrode surfaces following the rate capability tests. Fig. 5a–h presents a comparative structural evaluation of pristine CMK-3-130 and CMK-3-130/Ga composites with various Ga loadings (C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5), both before and after galvanostatic cycling. Before cycling, the pristine CMK-3-130 electrode (Fig. 5a) displays a uniformly distributed porous carbon structure, while the Ga-loaded composites (Fig. 5b–d) exhibit a denser, surface-confined morphology indicative of successful Ga infiltration.


image file: d5ta01253h-f5.tif
Fig. 5 SEM images of electrode microstructural changes of pristine CMK-3-130, and various Ga-loaded CMK-3-130–Ga (C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) anode surface examination. (a–d) Before cycling. (e–h) After GCD cycling. (i and j) Low and high-resolution image comparison of the coin cell electrode surface examination after GCD cycling for the pristine CMK-3-130 electrode. (k–l) CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 electrode.

After cycling, the pristine CMK-3-130 electrode (Fig. 5e) shows signs of surface roughening and the formation of microcracks, likely due to structural stress and unstable SEI formation. In contrast, the low Ga-loaded CMK-3-130/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5) electrode (Fig. 5f) maintains a smoother, crack-free surface, demonstrating enhanced structural stability. However, with increased Ga loading (1[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5), the electrode surfaces (Fig. 5g–h) exhibit an uneven, island-like morphology, suggesting localized Ga aggregation and possible volume expansion-induced mechanical degradation.

Fig. S8a–d show the enlarged view of SEM images of pristine CMK-3-130 and low amount of Ga CMK-3-130/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5), respectively. Both materials show a typical rod-like morphology before cycling (Fig. S8a and c). After cycling, structural changes were observed in pristine CMK-3-130 such as structural deformation, and a cracked, uneven morphology as shown in Fig. S8b. On the other hand, Fig. S8d shows that a low Ga content improves the electrode surface morphology by maintaining the dispersion of Ga nanoparticles, preventing their accumulation even after the high-rate performance.

Fig. 5i–l presents high-resolution SEM images comparing pristine CMK-3-130 (Fig. 5i and j) and low Ga-loaded CMK-3-130/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5) electrodes (Fig. 5k and l) after cycling. The pristine CMK-3-130 exhibits clear signs of structural breakdown, including internal cracking and surface deformation. Such damage likely impairs Li+ diffusion pathways, leading to increased irreversible capacity due to trapped lithium within fractured regions.

Fig. 5k and l shows the SEM images of the low Ga loaded material (CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5) after cycling, wherein an intact surface and a homogeneous Ga confinement were observed. The findings indicate that Ga plays a role in evenly stabilizing the surface, as no cracks were observed even after vigorous operation at the high current density. This reveals that, without Ga agglomeration, a more stable structure is maintained during extended cycling. In the low Ga loading sample, reduced surface roughness and uniform Ga confinement help minimize stress accumulation, effectively preventing crack formation. Moreover, nano-sized active Ga particles provide a better pathway for Li-ion conductivity leading to an even lithiation/delithiation process. Yoon et al. had similar observations wherein it was revealed that the reduced size of the Ga particle during electrochemical cycling actively stabilizes the electrode structure, preventing agglomeration within the carbon matrix.45 To provide further insight, SEM images of higher Ga loading electrodes (C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 ratios) after cycling are presented in Fig. S9a–d.

At higher loadings, the electrode surfaces exhibit island-like morphologies and greater roughness, accompanied by signs of mechanical stress and surface instability. This degradation is attributed to the accumulation of solid-phase Ga, which undergoes significant volume changes during cycling, leading to loss of active surface area and performance fading, consistent with capacity degradation trends observed in Fig. S5.

To further verify the spatial distribution of Ga and its integration within the CMK-3 matrix, SEM-EDX mapping was performed on CMK-3-130/Ga composites before and after cycling (Fig. S10). For the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 sample before cycling (Fig. S10a–c), the SEM image reveals a rod-like carbon morphology with dense surface features, while the EDX elemental maps show the even distribution of Ga with irregular accumulation on the surface. In contrast, the post-cycling SEM-EDX mapping of the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 sample (Fig. S10d–f) reveals a remarkably uniform gallium distribution embedded within the CMK-3 framework, with minimal aggregation even after prolonged electrochemical cycling. The strong overlap of carbon and gallium signals consistently supports a homogeneous dispersion throughout the carbon matrix. This observation reinforces the effectiveness of the ball-milling process in promoting Ga infusion and supports the hypothesis that lower Ga-loaded samples retain their uniform nanoconfined structure under high-rate cycling. Furthermore, combined with TEM data presented in Fig. 2, these results confirm the superior homogeneity of Ga distribution in low-loaded samples, which is pivotal for their enhanced structural stability and electrochemical performance.

The key to this improved stability lies in the higher retention of liquid-phase Ga at lower loadings. This state facilitates a more uniform active material distribution across the electrode surface, which in turn helps maintain electrical contact and minimize detrimental SEI formation or surface cracks. Moreover, Ga's intrinsic fluidity allows it to intrinsically accommodate the large volume changes inherent to alloying reactions. This sustained stability provides numerous accessible Ga active sites for efficient electrochemical reactions with Li+, thereby contributing to the excellent long-term capacity retention observed (Fig. 4c and d, S5d and h).

Besides, post-cycling mapping of the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 sample (Fig. S10g–i) reveals significant Ga accumulation and a roughened surface, indicating bulk agglomeration likely due to volume expansion or unstable SEI formation.

3.5.1. In situ XRD of pristine CMK-3-130. To investigate phase transitions during discharge/charge cycles (0.01–3 V, 100 mA g−1), in situ XRD was conducted on pristine CMK-3-130 and CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5. The cell assembly for these measurements is illustrated in Fig. S1 and the video. Distinct phase evolution was observed for pristine CMK-3-130, with newly emerging peaks in the lower angle regions during lithiation (Fig. S11a–e), highlighting irreversible changes and the potential versus time of this process indicated in Fig. S11f. These peaks become more prominent and continue to increase in intensity over time, with insignificant reversible peak shifts observed during delithiation. Extracted patterns at 2θ ranges of 10 to 80° for the 0–11 and 11–22 h cycling periods (Fig. 6a and b) confirm the formation of compounds such as Cu2O (PDF cards: 96-101-0942, 111-35-36, and 110-295-30), CuO (PDF cards 45-0937, 05-0661), Cu(OH)2 (PDF cards 13-0420, 72-0140, 35-0505), Li2O2 (PDF cards 01-073-1640, 09-0355), LiOH (PDF cards 01-085-0777, 01-085-1064), Li2CO3 (PDF cards 87-0729, 22-1141), and LixC6. These XRD peaks were referenced using multiple pdf cards providing additional verification of the observed diffraction patterns.22,68–76 Peaks corresponding to the Cu collector cubic Cu (PDF card no. 003-1018), porous separator, and Be window were differentiated through control XRD measurements (Fig. S12).
image file: d5ta01253h-f6.tif
Fig. 6 In situ XRD patterns with the 2θ ranges from 10° to 80° highlighting various compounds observed during GCD lithiation/delithiation of pristine CMK-3-130 at 100 mA g−1. (a) Phase evaluation for 0–11 hours. (b) Phase evaluation for 11–22 hours. (c) Comparison of in situ XRD patterns extracted at selected 2θ ranges from 34.75° to 35.3° and 43°–44° during various time frames (0, 11, 22, 33, 44, and 49 hours) of GCD cycling highlighting the phase transformation and providing insight into the redox process, corresponding to CuO, Cu(OH)2, Li2O2 and LiOH. (d) Corresponding to Cu, Cu2O, and Cu(OH)2.

Fig. 6c and d further isolates time-specific redox reactions showing Cu surface change alongside irreversible compound formation highlighting Cu surface dynamics. Fig. 6c displays a peak at approximately 2θ = 35°, which can be indexed to the [111] plane of CuO. The shoulder peak at 35° corresponds to the [111] plane and the peak at 34.9° to the [002] plane of Cu(OH)2. Additionally, the peak at 35.14° can be indexed to the [101] plane of Li2O2, while the peak at 34.8° is associated with the [110] plane of LiOH. Fig. 6d shows a peak at around 2θ = 43.5°, which can be indexed to the [111] plane of Cu. The shoulder peaks at 43.5° can be attributed to the [200] plane of Cu2O and the [131] plane of Cu(OH)2.

In situ XRD measurements reveal that the interaction of the Cu surface with the electrolyte initiates a dynamic redox process. These processes suggest electrolyte-induced Cu degradation or oxidation initiated via microcracks, as supported by post-cycling SEM images (Fig. 5). This process can lead to the formation of various copper oxide and hydroxide phases, as demonstrated by XRD analysis. This lack of reversibility could be a sign of irreversible structural modification or the trapping of lithium within the carbon matrix limits the material's ability to return to its original after delithiation fully. Pristine CMK-3's low coulombic efficiency (34%) indicates irreversible reactions, including SEI formation, surface oxidation at [H], [O] sites, and Li trapping in the carbon matrix and corrosion-like reactions of LixC6.11 Findings by Zhang et al. demonstrate that Cu oxidation that led to Cu redeposition on the anode surface during cycling can result in performance degradation due to an increase in SEI layer thickness.77 Our analysis of pristine CMK-3 revealed structural and textural features relevant to its performance as an anode material. Previously, direct confirmation of the mechanism behind copper current collector oxidation or degradation was challenging due to limitations in post-cycling surface analysis. However, in the current study, our comprehensive electrochemical analysis, particularly through operando CV-EIS measurements, provided direct evidence indicating surface oxidation of the copper current collector at higher potentials. This phenomenon aligns with observations by Wang et al.,20 who elucidated the oxidation of Cu to Cu2O and CuO during delithiation, manifested as characteristic peaks at 1.5 V, 2.5 V, and a shoulder at 2.7 V in their CV analysis. While direct evidence of redeposition remained elusive in our current surface characterization, the confirmed Cu oxidation in our system is a potential factor contributing to capacity fading. The specific high-potential features observed in our operando sequential CV-EIS analysis validate this Cu oxidation, along with their inferences for electrochemical performance, discussed in the CV-EIS section.

Our in situ XRD findings link the observed poor electrochemical efficiency and irreversible capacity loss in the pristine CMK-3 system to the surface instability of the carbon host. This instability leads to the degradation and subsequent oxidation of the copper current collector, resulting in the formation of various copper oxide and hydroxide phases. Furthermore, the emergence of irreversible compounds such as Li2O2 and Li2CO3 is observed. These cumulative factors collectively hinder lithium-ion cycling and significantly contribute to the system's capacity fading.

Building upon this understanding of the pristine CMK-3, the subsequent section presents the detailed in situ XRD analysis of the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 composite. This investigation thoroughly elucidates the sequential formation of all six Li–Ga alloying phases (L1–L6) during lithiation, their corresponding dealloying behavior (D1–D6) during delithiation, the unique inverted hysteresis features observed, and the concurrent formation of CuGa2.

3.5.2. In situ XRD of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5. Our in situ XRD analysis reveals significant insights into the phase transitions of the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 composite during lithiation and delithiation processes. We observed the appearance of new diffraction peaks which are marked with a solid black circle at the onset of lithiation, and it progressively shifts towards lower angles over time, indicating the expansion of the lattice as lithium ions are intercalated into the Ga-infused CMK-3 structure (Fig. 7(a–d)). In contrast, during delithiation, these peaks shift back towards higher angles, reflecting the contraction of the lattice as lithium is deintercalated. By the end of the delithiation process (solid dashed line), the XRD peaks return to their original positions, consistent with the material's initial state before cycling. These reversible processes were observed over prolonged periods of time of up to 65 h (Fig. S13a–d).
image file: d5ta01253h-f7.tif
Fig. 7 (a–c) In situ XRD patterns with the 2θ ranges from figure left to right (2θ = 24°–30°, 40°–45°, and 45°–50°) during GCD lithiation (solid black circle)/delithiation (black dashed line) of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 at 100 mA g−1. (d) In situ XRD measurement discharge/charge potential vs. time (h) for 0–43.5 h.

In situ XRD measurements were conducted on the 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 Ga-loaded sample due to the insufficient signal intensity observed at the lower 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 loading. The higher Ga content produced well-defined diffraction peaks, enabling clear tracking of lattice expansion and contraction during lithiation/delithiation, and unambiguous identification of Li–Ga phase transitions. This provided critical insight into alloying–dealloying mechanisms but revealed peak intensity decreases associated with capacity fading.

Although in situ data focused on the 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 sample, operando CV (Fig. 4b) and sequential CV-EIS analysis offer complementary mechanistic insights for both loadings. The 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 sample showed stable and reproducible redox features, while the 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 sample exhibited progressive peak reduction. Together, these results capture the structural dynamics, diffusion kinetics, phase evolution (L1–L6, D1–D6), unique inverted hysteresis features, and interfacial behavior across different Ga concentrations.

The appearance of these new XRD peaks and their corresponding voltage plateaus aligns well with recent ex situ XRD reports on carbon black confined Ga anodes.41 In such a case, Ga reacts with Li from the initial to the end of the lithiation, forming three electrochemically stable phases, including Li3Ga14 (rhombohedral), LiGa (cubic), and Li2Ga (orthorhombic). Deshpande et al. also demonstrated the Ga solid–liquid transition state of the self-healing electrode surface.40 Our in situ XRD outcomes prove these dynamic changes in the real-time lithiation process and reveal the appearance of the intermetallic phase through the gradual peak intensity changes over an extended time of discharging. The linear shifting of XRD peaks in both lithiation and delithiation suggests a strong correlation with the molar fraction of lithium within the lattice as discussed in the previous literature on similar systems40–42,51,78,79 and is consistent with our in situ XRD findings. This indicates that a complete alloying (liquid to solid)–dealloying (solid to liquid) transition occurs during the discharge/charge multi-step reaction process.41 The consistent, linear shift to lower angles during lithiation implies that the lattice parameter increases proportionally as lithium concentration rises, likely due to the incorporation of lithium into the Ga lattice.

Notably, Guo et al.52 identified several lithiation and delithiation phases in their ex situ study at specific voltages (0.23 V, 0.42 V, 0.7 V, 0.83 V, 0.97 V during charge and 0.8 V, 0.6 V, 0.4 V, 0.2 V during discharge), corresponding to Li2Ga, LiGa, and Li2Ga7. Their interpretations, based on the Li–Ga phase diagram by Saint et al.51 distinguished two structural families: Li-rich phases (Li2Ga, Li3Ga2, Li5Ga4) with layered 2D structures and Ga-rich phases (LiGa, Li5Ga9, Li2Ga7) with 3D frameworks. While these ex situ investigations provided valuable insights into peak evolution and some phase transitions, our in situ XRD uniquely offers direct and continuous identification of all six Li–Ga alloying and dealloying phases (L1–L6, D1–D6) throughout the electrochemical process, providing a comprehensive understanding beyond previously reported observations.

Extracted patterns at 2θ ranges of 10 to 80° for the 0–11 (lithiation), and 11–22 h (lithiation and delithiation) GCD cycling periods (Fig. 8a and b) confirm all six Li–Ga binary phase formation, and the copper electrode surface contact with Ga might form a surface level alloy of CuGa2.23,80,81 It reveals a significant improvement in the structural and electrochemical stability during the lithiation/delithiation process. The formation of Li2Ga alloy was evident, suggesting that a robust interaction between Ga and lithium and CuGa2 proves Ga contact with the copper current collector. During the initial lithiation (Fig. 8a) 0–11 hours, the patterns indicate the suppression of Cu oxidation, as no significant CuO or Cu(OH)2 peaks emerge compared to pristine CMK-3. This result underscores Ga's ability to mitigate copper oxidation by reducing direct Cu surface exposure to the electrolyte. Unlike the pristine sample, which exhibits peaks corresponding to irreversible compounds such as Li2O2 and Li2CO3 (Fig. 6b), the CMK-3/Ga sample shows no such signals during the 11–22 hour cycling period (Fig. 8b), highlighting the stabilizing effect of Ga in mitigating their formation. These findings indicate that the CMK-3/Ga coating stabilizes the electrode's structural integrity and prevents the capacity loss associated with irreversible side reactions. The observed improvements highlight the role of Ga in enhancing the efficiency and lifespan of CMK-3-130–Ga electrodes. This contrasts with the pristine CMK-3-130, where Cu oxidation and irreversible phase formation dominate, leading to poor electrochemical performance.


image file: d5ta01253h-f8.tif
Fig. 8 In situ XRD patterns of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 recorded at 100 mA g−1 showing 2θ ranges from 10° to 80° during lithiation/delithiation. (a) Phase evaluation during 0–11 hours; the inset shows the partially enlarged view of phase transition from amorphous Ga to Li3Ga14 (L1), Li5Ga9 (L2), and the initial new peaks' appearance. (b) Phase evolution during 11–22 hours; the inset shows the partially enlarged view of progressive lithiation and its lithium-rich transition phases LiGa (L3), Li5Ga4 (L4), Li3Ga2 (L5), and Li2Ga (L6) shown with a green line, for delithiation reverse phase transformation is shown with a red line Li2Ga (D1), Li3Ga2 (D2), Li5Ga4 (D3), LiGa (D4), Li5Ga9 (D5), and Li3Ga14 (D6). Comparison of in situ XRD patterns of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 in the 2θ range of 43°–44° during GCD cycling (0–66 hours), CuGa2 phase evolution characteristics. (c) Dynamic CuGa2 peak shape changes during lithiation (black line) and delithiation (red line) for 0–33 hours, intensity shifted. (d) Without peak intensity shift for 0–66 hours.

Fig. S14 illustrates the in situ XRD patterns of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 during GCD cycling (0–66 h) in the 2θ range of 34.8–35.5°. Unlike the pristine CMK-3-130, which exhibits significant copper oxidation and unstable electrode response, CMK-3-130/Ga demonstrates the absence of Cu oxidation phases, attributed to the protective role of Ga. Ga's unique properties, such as its ability to form stable alloys and its inherent electrochemical activity, may provide a protective effect, preventing Cu from oxidizing. Such an action was proven by the initial coulombic efficiency improvement, stabilizing the Cu surface, reducing irreversible capacity losses, and enhancing system performance.

Insets in Fig. 8a and b, focusing on the 2θ ranges of 24.5–26.5° and 40.75–42.2°, display the full progression of Li–Ga phases. These insets highlight the evolution of all six Li–Ga alloying phases (L1–L6, solid green lines) during lithiation, corresponding to transitions from Ga-rich to lithium-rich compositions: Li3Ga14 (L1), Li5Ga9 (L2), LiGa (L3), Li5Ga4 (L4), Li3Ga2 (L5), and Li2Ga (L6). Subsequent delithiation, presented as dealloying transition phases D1–D6 (solid red lines), corresponds to transition from lithium-rich back to Ga-rich compositions: Li2Ga (D1), Li3Ga2 (D2), Li5Ga4 (D3), LiGa (D4), Li5Ga9 (D5), Li3Ga14 (D6). Table S3 provides a detailed summary of the lithiation/delithiation potentials and corresponding key structural observations.


3.5.2.1 Lithiation phase evolution. Fig. 8a illustrates the structural changes during the first lithiation cycle (0 to 11 hours). Initially, the pattern (black line) shows the characteristic amorphous signature of gallium, consistent with its nanoconfined state within the CMK-3 mesopores. As lithiation commences and progresses, we observe a gradual appearance of new diffraction peaks at around 2θ = ∼25°, ∼41.3°, and ∼49°. These initial peaks broaden and exhibit narrow intensity, indicating a major structural change in Ga as vast lithium insertion forms the initial alloying phases, Li3Ga14 (L1) and Li5Ga9 (L2). As lithiation further progresses, more sequential formation of lithium-rich Ga alloyed intermetallic phases LiGa (L3), Li5Ga4 (L4), Li3Ga2 (L5), and Li2Ga (L6) shows a gradual increment in intensity and a characteristic shift towards lower 2θ angles (Fig. 8b), inset corresponding to lithiation. This shift signifies continuous lattice expansion as lithium is incorporated into the gallium host.

While individual phase transitions (L1–L6) compose a broad, single feature in the XRD pattern due to their closely spaced potential (Fig. 4b) and nanocrystalline nature, subtle fluctuations in peak intensity on top of this broad signal were observed, supporting the underlying, sequential phase transitions. At the final stage of lithiation, the overall broad peak appears to be slightly narrowed, with its composite nature becoming a more defined, ordered structure. This collectively demonstrates a well-defined sequential order of lithiation phases, evidenced by the gradual appearance of new peaks, their consistent increment in intensity, and a characteristic shift towards lower angles. As the second lithiation cycle commences, new L1 and L2 peaks emerge (dashed blue line, Fig. 8b). These phases exhibit similar structural dynamics to the first lithiation, with new peaks gradually emerging and shifting towards lower angles, affirming the high reversibility of the alloying process over multiple cycles. Prominently, although the individual differences in XRD peaks for these six phases (L1–L6) might be composed into one broad signal, conclusively distinguishable peaks detected for all six lithiation phases in our sequential CV-EIS operando analysis (as detailed in Fig. 9) strongly support the complete identification of all six Li–Ga lithiation phases. The similarity between these two powerful characterization techniques provides robust evidence for the full range of Li–Ga alloying events.


image file: d5ta01253h-f9.tif
Fig. 9 Operando sequential cyclic voltammograms (CVs) of CMK-3-150/Ga composites recorded over 15 cycles within a potential window of 0.01–3 V at a scan rate of 0.1 mV s−1. A partially enlarged view (0.01–1.5 V) highlights representative cycles 1, 2, 3, and 6 to capture key phase transitions. (a) Delithiation profiles for C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, the inset shows the single cycle D4 and D5 peaks partially highlighted with distinguishable humps. (b) Lithiation profiles for C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5. (c) Delithiation profiles for C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5. (d) Lithiation profiles for C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5. Annotated peak positions correspond to Li–Ga phase transitions, where D1–D6 indicate delithiated states and L1–L6 denote lithiated states, in accordance with the Li–Ga binary phase diagram. Al represents the reversible alloying (lithiation) and dealloying (delithiation) of the interfacial CuGa2 phase.

3.5.2.2 Delithiation phase transition and inverted hysteresis. Fig. 8b captures the structural dynamics from the end of the first lithiation to the start of delithiation (11 to 22 hours). Upon initiation of delithiation, a rapid structural rearrangement occurs, with peaks emerging at approximately 2θ = ∼25.2°, ∼41.5°, and ∼49.2°. These initial delithiation phase transitions, particularly from the highly lithiated Li2Ga, corresponding to D1, D2, and D3 (Li2Ga, Li3Ga2, and Li5Ga4), are marked by the appearance of distinct, high-intensity peaks exhibiting split and relatively narrow features (red lines with white diamonds). This initial structural evolution is crucial, as it directly supports the inverted hysteresis observed in our electrochemical CV data (Fig. 4b) and further evidenced by our CV-EIS sequential data over 15 cycles (Fig. S19 and S20). The sharp, high-intensity peaks in this region indicate significant structural rearrangements or electrochemical annealing that kinetically favor lithium extraction (dealloying) at lower potentials (0.3 V for D2, 0.46 V for D3), contrasting with their kinetically hindered formation (alloying) at higher lithiation potentials (0.44 V for L5, 0.53 V for L4). Crucially, our real-time in situ XRD analysis powerfully corroborates this intriguing inverted hysteresis, continuously capturing structural changes over prolonged cycling (up to 65 h for these initial delithiation phases, Fig. 7 and S13). As delithiation continues, the subsequent dealloying of higher potential Li–Ga phases LiGa (D4), Li5Ga9 (D5), and Li3Ga14 (D6) exhibits a progressive broadening and a gradual decrement in intensity (red lines), indicating the sequential removal of lithium and the eventual transformation back towards amorphous gallium. Peaks consistently shift towards higher 2θ angles, reflecting lattice contraction. Upon complete delithiation, the peaks diminish, indicating a return towards the amorphous gallium initial state.
3.5.2.3 Dynamic evolution of CuGa2 interphase at the current collector interface. Beyond the major Li–Ga alloying transformations, our in situ XRD patterns also provide critical insights into the dynamic behavior of the CuGa2 phase at the current collector interface, a factor often overlooked but crucial for long-term battery performance. Fig. 8c and d specifically highlight the evolution of the CuGa2 diffraction peak within the 2θ range of 43–44° during extended GCD cycling (0–66 hours).

Fig. 8c, which focuses on the 0–33 hour period with intensity shifts for clarity, reveals the subtle yet consistent changes in the CuGa2 peak shape during lithiation (black line, annotated as A1 with green) and delithiation (red line, annotated as A1 with red). Upon lithiation (alloying), the CuGa2 peak tends to become narrower, and its intensity shows a slight, gradual increase. This suggests that the formation of CuGa2 leads to a more coherent or ordered crystalline structure. Conversely, during delithiation (de-alloying), the peak tends to become broader, and its intensity shows a slight decrement. This broadening, despite minimal overall intensity change, indicates that the decomposition process may lead to a comparatively more disordered, nanocrystalline, or strained interface.

The extended observation over 0–66 hours in Fig. 8d, without intensity shifts, illustrating long-term stability further emphasizes that these subtle shape and intensity fluctuations for the CuGa2 phase are consistently reversible across multiple cycles. While the changes are minute, their reproducibility signifies that the CuGa2 layer is dynamically forming and deforming at the interface in a way that effectively maintains its overall structural integrity and protective function. This ongoing yet subtle evolution of the CuGa2 peak provides robust structural evidence for its role as a stable, buffering interphase layer, accommodating volume changes without undergoing bulk changes that would otherwise lead to significant pulverization or rapid capacity decay.23

The formation and behavior of copper–gallium intermetallic compounds (IMCs) like Cu9Ga4 and CuGa2 are well-established21,82 in fields such as microelectronics and solders, where they're often associated with interfacial reactions and diffusion characteristics. For instance, studies have shown how Ga addition can even suppress detrimental IMC growth in other systems by forming dense protective layers.36 However, our findings, particularly within the nanoconfined environment of CMK-3, reveal a remarkable transformation; the CuGa2 phase exhibits dynamic stability and reversibility under electrochemical cycling, evolving from a potentially problematic material into a beneficial component of the battery anode.

The stability of the current collector interface is a critical, long-standing challenge in high-performance Li-ion battery anodes, primarily due to the significant volume changes of anode materials during cycling, which lead to capacity fading and reduced cycle life. While existing solutions range from surface engineering (e.g., CuO nanowires, Cu foams)20 to self-healing liquid metal interfaces,23,43,80,81 and extensive research has explored architecture (e.g., 3D porous) or surface-modified copper current collectors to enhance interfacial stability and control Li deposition in high-performance anodes,83,84 the broader concept of in situ interphase engineering offers a powerful alternative to address diverse interfacial challenges across various battery chemistries. For instance, a recent ab initio study demonstrated in situ interphase engineering in an oxide-based all-solid-state Li battery cathode, where a computationally identified dopant (Al) enabled the formation of a stable protective layer that simultaneously improved sinterability and interfacial stability.85 Analogously, our work introduces an up-and-coming alternative, the spontaneous electrochemical formation of a reversible, protective CuGa2 layer, uniquely enabled by nanoconfined Ga.

Our comprehensive in situ XRD analysis provides fundamental structural evidence for this phenomenon. It details both the primary Li–Ga alloying/de-alloying sequences and the subtle, yet crucial, stable behavior of the interfacial CuGa2 phase. As shown in Fig. 8c and d, the reversible narrowing and broadening of the CuGa2 peak over extended cycling indicate its function as a flexible, self-adjusting buffer layer that maintains interfacial contact. This exceptional flexibility, unlike the typically brittle nature of bulk CuGa2, is attributed to nanoconfinement, which likely restricts macroscopic crack propagation and promotes localized, reversible atomic rearrangements.

Further support for this dynamic and reversible interphase comes from our sequential CV-EIS analyses. Our CV data (Fig. S15) show small, reversible alloying/de-alloying features around 1.2 V, consistent with CuGa2 formation and dissolution.63 EIS measurements, in turn, reveal stable interfacial charge transfer resistance, confirming a robust and continuously active interface. This electrochemical signature, along with evidence from the pristine sample surface evolution showing that the Ga surface is intrinsically mitigated by this protective layer, highlights the self-regulating nature of our system.

By providing strong structural and electrochemical evidence for a stable, dynamically reconfigurable interphase at the Cu surface, our results lay critical groundwork for next-generation current collector strategies. This in situ electrochemically formed CuGa2 interface, with its inherent flexibility and self-adjusting properties, holds significant promise for extending cycle life, minimizing interfacial resistance, and inspiring future anode designs that integrate self-regulating intermetallic protection layers. This innovative approach offers a powerful strategy to overcome a key limitation in battery technology, paving the way for more robust and long-lasting energy storage devices.

The operando in situ XRD analysis, extended to 65 hours of continuous cycling (Fig. 7 and S13), was instrumental in directly capturing the dynamic structural evolution of the CMK-3/Ga composite. Unlike prior ex situ reports that inferred or partially resolved Li–Ga phase transitions, our real-time tracking identifies all six Li–Ga alloying (L1–L6) and dealloying (D1–D6) phases. This unprecedented resolution is largely attributed to the nanoconfinement effect of mesoporous CMK-3, which effectively accommodates volume changes, restricts bulk crystal growth, and enhances surface-driven reaction kinetics. Notably, the in situ XRD results also provide structural validation for the ‘inverted hysteresis’ observed in the electrochemical profiles (Fig. 8b), particularly during the early dealloying transitions (D1–D3), suggesting a possible electrochemical annealing or rearrangement effect. These comprehensive structural insights offer a strong foundation for understanding the complex redox behavior and long-term stability of nanoconfined gallium-based anodes.

While in situ XRD reveals the crystalline phase transitions with high reliability, a complete understanding of the electrode's electrochemical behavior and kinetic pathways requires complementary time-resolved electrochemical techniques. In the following section, we present extensive operando sequential cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) analyses over 15 cycles (Fig. 9 and 10). This integrated approach enables high-resolution differentiation of subtle and overlapping electrochemical features, providing deeper insights into phase-specific redox transitions, kinetic reversibility, and the influence of Ga loading on charge transport dynamics. Additionally, these techniques allow us to evaluate interfacial stability, identify CuGa2 formation at the Cu interface, and compare the performance of pristine and Ga-infused CMK-3 under different synthetic conditions and compositions. Together, these findings further elucidate the mechanisms governing electrochemical stability and efficiency in nanoconfined alloy anodes.


image file: d5ta01253h-f10.tif
Fig. 10 Operando sequential cyclic voltammograms (CVs) and corresponding electrochemical impedance spectra (EIS) for CMK-3-130/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 mass ratio) within a potential window of 0.01–3 V at a scan rate of 0.1 mV s−1. (a) Delithiation, inset arrow demonstrates absence of Cu2O/CuO peaks, bottom shows the corresponding EIS spectrum of (a), inset shows the partially enlarged view of 1, 2, 5, and 6 cycles. (b) Lithiation, the bottom shows the corresponding EIS spectrum of (b).

3.6. Operando CV-EIS analysis of Li–Ga phase dynamics and interfacial stability

To further elucidate the unique behaviors and kinetic contributions of individual Li–Ga phases during alloying and dealloying, we performed detailed operando analyses using sequential cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS). These measurements provide critical insight into the electrochemical reversibility, interfacial stability, and lithium-ion transport dynamics of the CMK-3/Ga composite system. In particular, the influence of nanoconfinement and Ga loading on phase-specific redox kinetics and electrode resistance is explored in depth, revealing how these factors govern overall cycling stability and performance.

Sequential CV and EIS measurements were conducted on pristine CMK-3-130 and CMK-3/Ga composites synthesized at 130 °C and 150 °C. The CV measurements were recorded over 15 continuous cycles within a potential window of 0.01–3.0 V at a scan rate of 0.1 mV s−1. Corresponding EIS measurements were acquired at two key states of charge (SoC): at 3.0 V for the fully delithiated state (SoC ∼0%) and at 0.01 V for the fully lithiated state (SoC ∼100%). This dual-mode operando approach allows us to decouple lithiation and delithiation behaviors and correlate phase transitions with real-time changes in impedance characteristics. A comprehensive set of results is presented in Fig. 9 and 10, with supporting data in Fig. S15–S20.

3.6.1. Effect of Ga loading on interfacial CuGa2 phase reversibility and stability. At the initial stages of lithiation and delithiation, a small, reversible peak near ∼1.2 V marked as A1 in Fig. 9 and S15 emerges in the partially enlarged voltage window (0.01–1.5 V), particularly in cycles 1, 2, 3, and 6. This feature is consistently observed across both CMK-3 types (130 °C and 150 °C) with Ga loadings, though it appears with slightly higher intensity at increased Ga content. Notably, this peak, initially absent during the first cycle, becomes progressively visible in subsequent cycles, indicating the gradual formation of a surface alloy phase.

This electrochemical signature is attributed to the reversible alloying/dealloying of Ga with the copper current collector, which forms the intermetallic CuGa2 phase at the interface. The confined Ga within the CMK-3 matrix likely promotes localized alloying while simultaneously suppressing copper oxide formation. The consistent potential position and reversible nature of this A1 peak over multiple cycles suggest the dynamic but stable evolution of a protective CuGa2 layer, which buffers mechanical stress and maintains electrical contact.

This interpretation is strongly supported by in situ XRD data (Fig. 8c and d), where CuGa2 diffraction peaks are observed and remain stable over extended cycling. Together, these findings confirm the role of nanoconfined Ga in facilitating the reversible formation of a structurally robust interfacial phase, critical for current collector protection and long-term cycling performance.

3.6.2. Effect of Ga loading on Li–Ga lithiation phase reversibility. The operando cyclic voltammetry (CV) profiles of CMK-3-150/Ga composites at both mass ratios (C[thin space (1/6-em)]:[thin space (1/6-em)]Ga = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) closely match those of CMK-3-130/Ga samples (Fig. 9a–d and S15a–d), confirming the robustness and reproducibility of the six-step Li–Ga alloying sequence (L1–L6) over 15 cycles. This consistency highlights the effectiveness of CMK-3 nanoconfinement in stabilizing distinct Li–Ga phases otherwise challenging to resolve due to overlapping potentials and fast kinetics.

During lithiation (Fig. 9b, d, S15b and d), six well-defined cathodic peaks corresponding to Li3Ga14 (L1), Li5Ga9 (L2), LiGa (L3), Li5Ga4 (L4), Li3Ga2 (L5), and Li2Ga (L6) appear sequentially. Notably, the L1 peak, corresponding to the initial formation of the Li3Ga14 phase, emerges as a narrow, high-intensity signal. A subtle, transient shift of this peak to slightly higher potentials is observed in the earliest cycles, characteristic of an initial activation or structural re-arrangement within the nanoconfined system, before consistently returning to stable cycling characteristics for subsequent cycles. In situ XRD confirms this phase formation through a broad peak that progressively narrows and shifts to lower angles, indicative of lattice expansion. For low Ga loading, the L1 peak remains stable in both intensity and position over cycling, correlating with superior coulombic efficiency and long-term performance. In contrast, at higher loadings, L1 shows greater initial intensity but rapidly declines with cycling, indicating poorer phase stability and increased irreversibility. The L2 and L3 peaks are closely spaced but distinguishable, with L2 being smaller and broader in low-loading samples. The L4 peak becomes sharper and more intense with increased Ga content, while the L5 and L6 phases associated with deeper lithiation exhibit contrasting trends. In low-loading systems, L5 is broad, and L6 grows sharper with continued cycling, reflecting improved deep lithiation. In high-loading samples, however, L6 becomes less distinct and often overlaps, indicating hindered phase formation and lithium transport. These findings align with EIS results, where low-loading samples display enhanced diffusion and reduced interfacial resistance.

While CV reveals distinct phase transitions, in situ XRD provides structural confirmation. The alloying-induced XRD peaks broaden initially and shift to lower angles with progressive lithiation, reflecting lattice expansion and nanocrystalline phase formation (Fig. 7, 8a and b). Despite overlap in some XRD patterns due to nanoconfinement effects, the high resolution of CV allows clear electrochemical identification of each Li–Ga intermetallic phase.

3.6.3. Impact of Ga content and Li–Ga delithiation phase reversibility and inverted hysteresis. Conversely, during delithiation (Fig. 9a, c, S15a and c), lithium is extracted from the Li–Ga alloys, exhibiting six characteristic anodic peaks (D1 to D6). These peaks correspond to the sequential delithiation of Li2Ga (D1), Li3Ga2 (D2), Li5Ga4 (D3), LiGa (D4), Li5Ga9 (D5), and Li3Ga14 (D6), demonstrating the remarkable reversibility of these phase transitions for both low and high Ga loadings. Notably, a striking deviation from typical electrochemical hysteresis is observed for the early delithiation transitions involving Li2Ga, Li3Ga2, and Li5Ga4 (D1–D3).

Fig. S16 CV demonstrates the reproducibility of the D1 delithiation transition and the presence of potential ‘pinning points’ in the CMK-3/Ga composites over 15 cycles. Specifically, the CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 sample shows a highly consistent D1 transition at approximately 0.05 V, with strong convergence across cycles (marked in circle), indicating a stable and confined reaction environment (Fig. S16a). Similarly, the CMK-3-150–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 sample exhibits a sharp and reproducible D1 transition at around 0.04 V, also demonstrating marked convergence throughout the cycles (Fig. S16b). Conversely, for samples with a higher loading (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) of both CMK-3-130 and CMK-3-150, the D1 transitions occur near 0.05 V but show reduced convergence (Fig. S16c and d). This suggests increased structural variability and a less confined phase evolution in these higher loading samples.

This striking feature is the distinct convergence of all 15 delithiation curves at approximately 0.05 V. From this point, a consistent formation of a ‘thick line’ across all cycles highlights its role as a highly robust and reproducible thermodynamic potential for an initial delithiation step. This effectively acts as a ‘pinning point’ that the system consistently achieves regardless of subtle kinetic fluctuations in previous cycles, indicating a kinetically favorable and highly stable delithiation sequence. For the higher Ga loading composites, such convergence at 0.05 V is notably absent, with individual cycle lines remaining distinguishable, suggesting increased kinetic hindrance or structural variability.

This remarkable electrochemical stability and the manifestation of a ‘pinning point’ in nanoconfined alloying materials find strong parallels in other well-studied high-performance systems. For instance, Chan et al.'s seminal work on silicon nanowires demonstrated how nanoscale architecture can effectively accommodate the massive volume changes inherent to Si during lithiation/delithiation, leading to significantly improved cycle life and stable electrochemical cycling. Their foundational insights underscored the vital role of nanoscale confinement in mitigating pulverization and preserving structural and electrical integrity.64 Building upon this understanding of structural stability, our work further demonstrates that nanoconfinement enables striking voltage profile convergence and highly reproducible electrochemical behavior, even under substantial volume changes. In the CMK-3 confined Ga system, the superior nanoconfinement at low loading composites provides a similar buffering effect, allowing the Ga particles to relax and move within the confined environment without losing electrical contact or suffering irreversible pulverization. This sustained structural integrity is pivotal for the observed capacity recovery after initial cycles and the overall robust electrochemical response. We hypothesize that during cycling, the continuous volume changes and electrochemical cycling lead to the dynamic emergence of smaller, more active Ga particles from the initially confined material. These newly formed, highly dispersed active sites, benefiting from the cushioned nanoconfinement provided by the CMK-3 matrix, can then more effectively and consistently participate in the subsequent electrochemical reactions, contributing to the observed long-term stability and the characteristic 0.05 V delithiation pinning point. Conversely, higher Ga loadings likely hinder this dynamic particle refinement and effective buffering action, resulting in less consistent electrochemical features and limiting the capacity recovery.

Our observation of inverted hysteresis for the D2 (∼0.30 V) and D3 (∼0.46 V) delithiation peaks, being lower than their lithiation counterparts (L5 and L4, respectively), offers a critical insight into the electrochemical behavior of these phases. As shown in Fig. 9c, d, S15c and d, these specific delithiation peaks exhibit distinct characteristics compared to their lithiation. This phenomenon suggests a structural reordering or “electrochemical annealing” process occurring during the delithiation cycle within the nanoconfined environment. The initial formation (lithiation) of Li3Ga2 and Li5Ga4 may lead to a more disordered or strained structure, requiring a higher overpotential. However, during delithiation, the confined environment could facilitate a more relaxed or ordered structural rearrangement, thereby lowering the kinetic barrier for Li extraction and shifting the de-alloying potential to a seemingly ‘lower’ (more favorable) value than expected based on lithiation. This finding highlights the complex interplay between electrochemistry, structural dynamics, and nanoconfinement in determining overall battery performance. Understanding these precise kinetic and structural details is paramount for designing next-generation Li-ion battery anodes with improved stability and performance.

For these phases, the delithiation peak potentials are found to be lower than their corresponding lithiation potentials (e.g., Li3Ga2 delithiates at ∼0.31 V vs. ∼0.44 V for its lithiation; Li5Ga4 delithiates at ∼0.46 V vs. ∼0.53 V for its lithiation). This inverted hysteresis is a powerful indicator of unique kinetic and structural factors at play. Our in situ XRD analysis directly corroborates this finding: upon initiation of delithiation, the corresponding XRD peaks (for Li3Ga2 and Li5Ga4) exhibit a notable increase in intensity and a slight shift towards higher 2-theta angles (Fig. 8b), inset. This suggests that the lithiated phases (Li3Ga2 and Li5Ga4) undergo a favorable structural rearrangement or electrochemical annealing after formation, allowing for a more facile and kinetically advantageous lithium extraction, despite the higher kinetic barriers encountered during their initial formation via lithiation. The D2 peak (Li3Ga2) appears with high intensity and a narrow shape, while D3 (Li5Ga4) exhibits shared higher intensity and a smooth, narrower profile, indicating a progressively more distinct electrochemical signature as lithium is extracted from the more Li-rich states.

However, a subtle yet significant difference emerges in the delithiation profiles of CMK-3-150/Ga, particularly for the low loading (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5). While the D4 and D5 peaks often appear as a broad, merged feature in the CMK-3-130/Ga samples (especially in initial cycles), the CMK-3-150/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5) distinctively reveals a more pronounced and distinguishable hump on top of this broadened peak (Fig. 9a), inset. This observation provides solid evidence for the clear resolution of both the D4 and D5 phases, which might otherwise appear merged, and this phenomenon becomes even more evident in progressive cycles. The subtle increase in pore size (as CMK-3-150 typically has slightly larger pores than CMK-3-130, though still within the 4–5 nm range) likely led to a more effective nanoconfinement effect. This refined nanoconfinement may subtly refine the local environment for these phase transformations, contributing to better separation or stability of the intermetallic phases over cycling and allowing these distinct humps to emerge more clearly. While peak separation is most visibly pronounced in the low-loading CMK-3-150 sample, strong evidence for the complete phase resolution, albeit less distinct, is also present across other loadings and in the CMK-3-130 series upon close examination of progressive cycles. The final delithiation D6 peak appears broader but with sharper intensity compared to its nearest D5 phase. Furthermore, the intensity of this D6 peak decreases with progressive cycling, which suggests a smoother transition back to its initial Ga states. The reversible appearance of the A1 peak during de-alloying further confirms the dynamic nature of the CuGa2 interfacial layer.

This distinct behavior is further validated by the first-cycle GCD profile of CMK-3-130–Ga 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 shown in Fig. S17. The discharge (lithiation) segment displays a sloping curve beyond the initial L1 plateau (∼0.7 V), reflecting kinetically hindered alloying transitions that result in merged signals rather than well-separated plateaus. Conversely, the charge (delithiation) profile exhibits exceptionally well-resolved, stepwise plateaus (D1–D6). These clear electrochemical fingerprints correspond to the sequential dealloying of all six Li–Ga binary intermetallic phases (from Li2Ga to Li3Ga14), providing compelling evidence of their distinct electrochemical potentials. This striking asymmetry merged lithiation slope versus precisely resolved delithiation plateaus is characteristic of advanced alloying systems and strongly reinforces our sequential CV and real-time in situ XRD findings, which also distinguish all six redox pairs. Collectively, these GCD results profoundly underscore the high degree of phase reversibility and structural stability facilitated by nanoconfined Ga, especially under optimal loading, offering critical mechanistic insight into the performance of alloy-type anodes.

This detailed electrochemical analysis, driven by high-resolution operando CVs and powerfully corroborated by real-time in situ XRD, unequivocally demonstrates the successful formation and remarkable electrochemical reversibility of all six predicted Li–Ga binary intermetallic phases within the nanoconfined CMK-3 host (Table S3). Our work represents the first comprehensive electrochemical resolution of these intricate Li–Ga phase transitions, providing critical mechanistic insights into the alloying processes that govern the performance of Ga-based anode materials. Crucially, the Ga-engineered composites effectively suppress copper current collector oxidation, thereby mitigating the irreversible capacity loss commonly observed in pristine CMK-3 electrodes. The observed phase-specific kinetics, including intriguing inverted hysteresis and distinct pinning points, highlight the profound impact of nanoconfinement. Specifically, optimal nanoconfinement (e.g., in CMK-3-130 and CMK-3-150 with ∼4–5 nm pores) at lower Ga loadings enables kinetically favorable delithiation processes, leading to superior long-term capacity retention and unique capacity recovery. Conversely, higher Ga loadings introduce increased kinetic hindrance and structural variability, limiting their long-term electrochemical performance. These comprehensive findings collectively lay a robust foundation for the rational design of highly stable and efficient Li-ion battery anodes with broader implications for next-generation electrochemical energy storage devices.

3.6.4. Electrochemical performance and the critical role of Ga loading. The CV profiles for both the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 CMK-3/Ga mass ratios confirm the excellent reversibility of the Li–Ga alloying/de-alloying process over 15 cycles. The consistent appearance and relative positions of the lithiation (L1–L6) and delithiation (D1–D6) peaks across multiple cycles underscore the structural stability of the nanoconfined Ga.

However, a closer examination reveals the critical and differential role of Ga loading in the overall electrochemical performance. For instance, while the higher loading CMK-3-130/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) composite exhibits a notably higher initial coulombic efficiency (ICE) of approximately 57% for its initial cycles, compared to the lower ICE of around 50–52% for the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 loading, this initial advantage does not translate to superior long-term stability for higher loading samples.

In terms of capacity retention, the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 mass ratio composite demonstrates superior long-term cycling stability, retaining approximately 56% of its capacity after 100 cycles. Strikingly, this capacity retention value is even higher than its initial ICE, indicating a unique capacity recovery phenomenon over prolonged cycling. This is a significant claim, suggesting that nanoconfined gallium becomes increasingly active and accessible with cycling. This capacity recovery is consistent with our CV observations for low loading, where the deep lithiation peak (L6, near 0.1 V) remains distinguishable, even showing a minute shift towards lower potentials and an increase in intensity over cycles, suggesting a smoother and more complete lithiation process over time. In stark contrast, the higher loading (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) composite suffers from significant capacity fading, retaining only about 16% of its capacity after 100 cycles, primarily due to the diminished intensity of its deep lithiation peaks and increased irreversible reactions.

This beneficial behavior of low Ga loading, including the phenomenon of capacity recovery, is consistently observed across different CMK-3 synthesis temperatures, notably for the CMK-3-150/Ga low loading as well. Both CMK-3-130 and CMK-3-150 share similar pore sizes, generally ranging between 4 and 5 nm, which appears to be crucial for providing the optimal nanoconfinement necessary for these stable and recoverable electrochemical processes. The nanoconfinement ensures Ga remains highly active and accessible, preventing detrimental agglomeration and facilitating the observed long-term stability and capacity recovery.

3.6.5. Probing charge transfer and diffusion dynamics during cycling. To gain deeper mechanistic insights into the electrochemical processes governing cycling stability, operando cyclic voltammetry (CV) coupled with electrochemical impedance spectroscopy (EIS) measurements were conducted across various CMK-3/Ga composites and pristine CMK-3 (Fig. 10 and S18–S20). While a comprehensive quantitative fitting of the impedance data was not performed in this study due to the complex evolving nature of the interfaces, the qualitative evolution of key impedance parameters provides crucial insights into charge transfer, solid-electrolyte interphase (SEI) formation, and diffusion dynamics during lithiation and delithiation. Specifically, we focus on changes in solution resistance (Rs), charge transfer resistance (Rct), solid-electrolyte interphase resistance (RSEI), and Warburg impedance (W). It is important to note that for all samples, the Warburg region of the EIS spectra during the lithiation sequence often presented as scattered or inconsistent points, precluding clear interpretation. Therefore, for better visualization and focus on the distinct semicircle regions, the Warburg features for lithiation cycles are generally omitted from the main Nyquist plots, though an explanatory example of such a spectrum is provided in Fig. S20d inset for reference. Conversely, the delithiation Warburg features, when well-defined, are presented and discussed.

The inherently high initial irreversible capacity loss of pristine mesoporous carbons like CMK-3 has been widely attributed to their large surface area, leading to extensive Solid Electrolyte Interphase (SEI) formation and associated irreversible faradaic reactions during the first lithiation cycle.11 Our operando CV-EIS analyses of the pristine CMK-3-130 (Fig. S18) directly demonstrate the severity of this issue and highlight how gallium infusion in CMK-3 significantly mitigates it.

In the pristine CMK-3-130 CV profile (Fig. S18a), representing delithiation cycles 1–15, the initial cycle exhibits a pronounced irreversible capacity, characterized by a large voltage hysteresis between the first and second cycles (marked as 1 and 2). This large irreversible capacity is primarily associated with initial lithium consumption for SEI formation, which tends to be thicker and uneven due to the carbon's large surface area. Furthermore, the inset of Fig. S18a clearly shows the appearance of distinct Cu oxidation peaks (e.g., at ∼1.89 V, 2.31 V, 2.5 V, and 2.7 V) in subsequent delithiation scans after the first cycle. This suggests that the significant volume changes experienced by the pristine carbon during initial lithiation lead to electrode cracking, allowing electrolyte penetration and subsequent oxidation of the copper current collector surface to complex phases such as Cu2O/CuO and other oxides. A small intensity hump observed around 1.2 V during delithiation is consistent with partial SEI decomposition, as reported by other studies on CMK-3.62

The corresponding EIS spectra for pristine CMK-3-130 (Fig. S18a) provide direct evidence for the instability and progressive impedance increase over cycles. The Nyquist plots display a large, often ill-defined semicircle in the high-to-mid frequency range, where the RSEI and Rct largely overlap (indicated by the arrow with the circle mark), signifying a highly resistive and intertwined interfacial impedance during cycling. The inset of Fig. S18a, showing an enlarged view of extracted cycles (1, 2, 5, and 6), further demonstrates that the first delithiation cycle exhibits a large semicircle, and for the second cycle the corresponding EIS shows an even higher and wider semicircle than the first, indicating an immediate increase in resistance. Beyond this, while some reduction in semicircle size is noted (e.g., cycle 5), inconsistent behavior with subsequent increases (e.g., cycle 6) is observed. Moreover, the low-frequency Warburg region during delithiation is poorly defined, and the observed two-step low-frequency behavior with a shallow Warburg slope indicates severely hindered ion diffusion within the pristine carbon. Collectively, these impedance trends, along with the electrochemical features, indicate significant structural irregularities and instability within the electrode, further supported by post-cycling surface morphology observations that reveal a cracked and uneven surface (Fig. 5e, i and j).

For the pristine CMK-3-130 lithiation (Fig. S18b) and its corresponding EIS spectrum, fewer distinct features are observed in the CV, apart from a small nuisance peak around 0.15 V, likely associated with initial Li insertion into the carbon matrix. In the EIS (circled and arrowed mark), the Rs shows a slight shift towards the left, and the semicircle generally exhibits a decreasing trend. Warburg region was not included due to its noisy and inconsistent nature, reflecting complex diffusion dynamics during lithium insertion into this unstable pristine host.


3.6.5.1 CMK-3-130/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 mass ratio): enhanced reversibility and kinetics. Compared to pristine CMK-3-130, the CMK-3-130–Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 mass ratio) composite exhibits markedly improved electrochemical behavior, as demonstrated through the sequential CV-EIS analysis (Fig. 10a and b). The cyclic voltammograms (CVs) over 15 cycles reveal a substantial reduction in first-cycle irreversible capacity, as evidenced by the minimized hysteresis between the first two cycles. Notably, the absence of parasitic Cu2O/CuO oxidation peaks indicated in Fig. 10a persists throughout the entire cycling sequence. This suppression of side reactions is consistent with previous in situ XRD results, which confirmed the formation of a protective CuGa2 interphase at the electrode interface. Although in situ validation focused on higher loadings, it is reasonable to infer similar behavior in the low-loading system due to analogous structural features.

In the delithiation profile (Fig. 10a), a fluctuating peak around 1.2 V initially absent emerges in later cycles, suggesting a reversible deformation of the CuGa2 interphase, consistent with the A1 peak discussed in Section 3.6.3. The lithiation CV (Fig. 10b) shows overlapping, stable peak profiles with increasing or constant intensities, indicating enhanced reversibility. All six Li–Ga alloying phases (L1–L6), as outlined in Section 3.6.2, are resolved and maintain stability throughout the cycles. A prominent peak near 0.1 V further signifies stable deep lithiation behavior.

The corresponding EIS analysis during delithiation (Fig. 10a, bottom) reveals a progressive reduction in both solid-electrolyte interphase (SEI) resistance (RSEI) and charge transfer resistance (Rct), as evidenced by the decreasing diameters of the corresponding semicircles in Nyquist plots. These semicircles are distinctly separated, indicating well-resolved contributions from SEI and charge transfer processes. Concurrently, the Warburg region exhibits a leftward shift and a steeper slope, reflecting enhanced lithium-ion diffusion kinetics and improved interface stability, consistent with observations in Fig. 4a. The inset in Fig. 10a, displaying impedance spectra for cycles 1, 2, 5, and 6, further supports this trend, showing a consistent decrease in impedance and increasingly pronounced Warburg behavior over cycling. These findings align with the electrochemical convergence and pinning effect observed in Fig. S16, suggesting improved interfacial stability and charge transport efficiency. Collectively, these results highlight the CMK-3/Ga's potential for enhanced cycling performance and longevity in lithium-ion batteries.

During lithiation (Fig. 10b bottom), the EIS reveals a progressive decrease in the diameter of the semicircle, indicating a reduction in Rct and RSEI. These overlapping Rct and RSEI contributions suggest the formation of a uniform and highly conductive electrode–electrolyte interface, which remains stable despite repeated volume changes during cycling. The solution resistance (Rs) also remains consistent, underscoring the structural integrity of the Ga-loaded electrode. In contrast, the pristine CMK-3 electrode exhibits large and erratic impedance fluctuations, likely due to its heterogeneous surface and unstable SEI formation. The Warburg region, associated with lithium-ion diffusion, was excluded from the spectra due to significant noise, which obscured a reliable interpretation; however, this does not detract from the observed trends in Rct, RSEI, and Rs. These results highlight the superior electrochemical stability of the Ga-loaded electrode compared to pristine CMK-3, attributed to gallium's role in enhancing interface uniformity and conductivity.

These findings are corroborated by prior studies on alloy-based anodes, which reported enhanced electrochemical performance through reduced charge-transfer resistance and increased electrochemically active surface area. For instance, Yu et al.86 observed a significant reduction in impedance after the initial lithiation cycle in Cu–Ga alloy-based anodes, attributed to stabilized charge-transfer processes. Similarly, Li et al.83 demonstrated that 3D Cu/Li electrodes with high surface area exhibit enhanced lithium-ion diffusion kinetics, leading to lower Rct. Guo et al.52 further reported that removing surface oxides in liquid Ga–In anodes reduces impedance and improves lithium-ion diffusion, aligning with the stable Rs and decreasing Rct observed in our CMK-3/Ga system. The progressive reduction in impedance in our Ga-loaded CMK-3 electrode supports the hypothesis that well-dispersed Ga nanoparticles increase the electrochemically active surface area, as evidenced by post-cycling SEM (Fig. 5k), thereby enhancing lithium-ion diffusion and cycling stability. This carbon-hosted nanoconfinement principle is mirrored in systems like carbon-encapsulated CeO2–Co catalysts for zinc–air batteries, where nanostructured carbon frameworks form a corrosion-resistant interface, mitigating oxidative degradation thus improving structural reversibility.87 These comparisons underscore the role of gallium nanoconfinement in achieving a robust and conductive electrode–electrolyte interface in our system.


3.6.5.2 CMK-3-130/Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 mass ratio): interfacial instability at higher loading. The CV and EIS profiles of the CMK-3-130–Ga electrode with a 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 mass ratio (Fig. S19a and b) reveal distinct electrochemical behavior compared to the optimized 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 counterpart. In delithiation (Fig. S19a), the CV shows reduced first-cycle hysteresis and no Cu2O/CuO peaks, with a minor 1.2 V peak confirming CuGa2 formation, consistent with the low-loading electrode. However, lithiation (Fig. S19b) exhibits high-intensity initial peaks that reduce with cycling, indicating capacity degradation. The 0.12 V peak's reduced resolution suggests lower surface activity for lithium storage, correlating with decreased long-term reversibility. We attribute the capacity degradation in high-loading CMK-3/Ga to larger Ga nanoparticles, which diminish carbon nanoconfinement and promote electrolyte decomposition, as evidenced by post-cycling SEM analysis (Fig. 5h).

EIS spectra for delithiation (Fig. S19a) show an uncertain semicircle after some cycles, also the Warburg region shows scattered data due to high signal-to-noise ratios, reflecting unstable lithium-ion diffusion. During lithiation (Fig. S19b), the semicircle diameter decreases slightly, yet RSEI and Rct resistances remain elevated and poorly resolved. The Warburg component was excluded due to excessive noise, suggesting a kinetic barrier, potentially exacerbated by high state-of-charge (SoC) or surface heterogeneity at elevated Ga loading. These findings highlight the limitations of excessive Ga incorporation, contrasting with the stable interface of the low-loading system.


3.6.5.3 CMK-3-150/Ga series: nanoconfinement enhances reversibility and kinetics. The electrochemical behavior of CMK-3-150/Ga composites (Fig. S20a–h) further validates the robustness of the nanoconfinement strategy across different synthesis temperatures. For the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 Ga loading sample (Fig. S20a and c), the CVs for both delithiation and lithiation closely mirror the superior performance observed in the CMK-3-130–Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5) system. The delithiation CVs exhibit highly overlapping features and significantly reduced first-cycle hysteresis, along with the appearance of subtle humps corresponding to the D4 and D5 phases, as previously discussed in Fig. 9a. The comprehensive 15-cycle dataset further underscores the long-term reproducibility of phase evolution and the consistent suppression of detrimental Cu2O/CuO-related side reactions. Correspondingly, lithiation CVs exhibited nearly overlapping cathodic peaks and stable fluctuations near 1.2 V, indicative of stable CuGa2 interphase formation and transformation. This observation reinforces that the Ga content plays a pivotal role in current collector protection.

The EIS spectra during delithiation for the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 Ga loading (Fig. S20b) consistently exhibit stable solution resistance (Rs), and the progressive reductions in SEI resistance (RSEI), and charge transfer resistance (Rct). The continuous shrinkage of the high-frequency semicircle and a clear leftward shift of the Warburg region signify significantly enhanced lithium-ion diffusion kinetics and stabilized electrode–electrolyte interfaces. These impedance features, particularly the distinct reduction in semicircle width and the well-defined Warburg line displayed in Fig. S20b inset, collectively support kinetically reversible processes, improved charge transfer, the formation of a stable and progressively thinner SEI layer, and optimized lithium-ion pathways. These findings closely match those observed in CMK-3-130–Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5), confirming that the optimized Ga dispersion strategy yields comparable kinetic and interfacial performance across both synthesis temperatures. During lithiation (Fig. S20d), EIS also shows a reduction in semicircle diameter, although the Warburg region appears noisy and undefined, manifesting as circling behavior. This likely stems from dynamic structural strain at high states-of-charge (SoC), suggesting a kinetic barrier, which was further clarified through a magnified view in Fig. S20d inset.

For the higher Ga loading (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) sample (Fig. S20e–h), CV responses are consistent with the CMK-3-130–Ga (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) system. While the initial CV peaks appear with greater intensity, they gradually fade over continued cycling, indicating reduced long-term stability. The CuGa2-related peak at 1.2 V remains apparent but becomes more prominent compared to low-loading samples. EIS analysis (Fig. S20f and h) further confirms increased interfacial resistance and structural instability at this higher loading. Scattered and unstable Warburg behavior, coupled with poorly defined semicircle features, reflect compromised lithium-ion transport and interfacial control, which is consistent with the capacity fading trends observed in electrochemical testing.

Overall, the comparative analysis between CMK-3-150/Ga and CMK-3-130/Ga systems shows that Ga nanoconfinement plays a crucial role in achieving stable, reproducible, and kinetically favorable electrode behavior. Specifically, the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 Ga loading consistently maintained optimal phase evolution, interfacial protection, and diffusion characteristics across both temperature regimes, demonstrating its superior reversibility and high structural stability. These findings highlight the kinetically driven pathways observed at optimal loadings (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5), contrasting with the more significant kinetic hindrance observed in their high-loading counterparts.


3.6.5.4 Operando CV-EIS analysis summary and reaction pathway mapping. Taken together, our operando CV-EIS data combined with in situ XRD structural insights allow for the full mapping of lithiation and delithiation reaction pathways for the Ga-infused CMK-3 anode system. For the first time, all six Li–Ga alloy phases are both electrochemically resolved and structurally confirmed in sequence:
Lithiation: Ga(l) → Li3Ga14 → Li5Ga9 → LiGa → Li5Ga4 → Li3Ga2 → Li2Ga

Delithiation: Li2Ga → Li3Ga2 → Li5Ga4 → LiGa → Li5Ga9 → Li3Ga14 → Ga(l)

This proposed reaction sequence, supported by the reproducibility of sequential CV cycling and in situ XRD validation, can be incorporated into Fig. 4b to contextualize the phase evolution. Importantly, the nanoconfined Ga environment enables kinetic stabilization and clear resolution of otherwise transient or overlapping intermediate phases, particularly Li3Ga2, Li5Ga4, and Li5Ga9, which are often indistinguishable in bulk systems due to phase overlap and rapid transformation kinetics.

Additionally, a comparative overview of the CMK-3/Ga system with other advanced anode materials is summarized in Table S4 highlighting synthesis methods, structural properties, low-temperature behavior, and key electrochemical metrics. This comprehensive comparison allows us to contextualize insights within the broader field.

Across both CMK-3-130 and CMK-3-150 systems, the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 mass ratio consistently exhibits superior electrochemical performance, attributed to optimal Ga dispersion within the carbon framework. Enhanced CV resolution, minimal impedance, and suppressed side reactions confirm the advantage of nanoconfined Ga at lower loadings. Conversely, higher Ga content compromises structural integrity and hinders long-term stability, despite initial improvements in ICE. These findings reaffirm the critical role of nanoconfinement and precise compositional tuning for advancing alloy-type anodes.

4. Conclusions

In summary, we present a gallium-infused CMK-3 composite anode, synthesized via a scalable ball-milling method, which marks a significant advancement in carbon-based alloy-type anodes for lithium-ion batteries. Initial structural analyses confirmed the successful nanoconfinement of Ga nanodroplets within the preserved mesoporous carbon (4–5 nm), while low-temperature XRD revealed Ga's intrinsic reversible liquid–solid phase transition critical to its electrochemical adaptability.

Through operando cyclic voltammetry and in situ X-ray diffraction, we achieved the first complete electrochemical and structural identification of all six Li–Ga alloying and dealloying phases Li3Ga14, Li5Ga9, LiGa, Li5Ga4, Li3Ga2, and Li2Ga (L1–L6, D1–D6) within a confined system. This confinement also kinetically stabilizes and enables clear detection of intermediate phases such as Li3Ga2, Li5Ga4, and Li5Ga9 that are often obscured in bulk Ga systems due to their transient nature or overlapping transformations. This high-resolution mapping not only clarifies longstanding ambiguity in Li–Ga phase evolution but also uncovers a novel inverted hysteresis behavior where delithiation begins at lower potentials than lithiation, due to kinetically driven lithium release and significant structural rearrangements.

Beyond phase tracking, Ga's incorporation plays a multifaceted role in enhancing electrode stability. Operando CV-EIS and in situ XRD confirm that Ga effectively suppresses parasitic Cu/C-based oxide formation seen in pristine CMK-3, while enabling the reversible formation of a CuGa2 interphase. This self-regulating buffer layer stabilizes the electrode–collector interface, mitigates delamination, and enhances Li+ transport kinetics.

Optimization of Ga content proved essential, with the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 mass ratio demonstrating superior cycling stability, high initial coulombic efficiency, and minimal structural degradation, as validated by post-cycling SEM. In contrast, higher Ga loadings (e.g., 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) exhibited notable capacity fading and microstructural instability, attributed to Ga aggregation and increased distortion. These effects likely stem from reduced nanoconfinement efficacy and enhanced irreversible reactions. Nonetheless, even at higher loadings, Ga consistently protected the copper current collector from oxidation, underscoring its critical role in interfacial engineering and structural stabilization, despite diminished electrochemical performance.

Overall, this work establishes a foundational framework for designing next-generation self-protective anodes, advancing both the fundamental understanding and practical application of liquid metal-based electrochemical systems. The approach and insights herein may also be extended to alternative porous carbon hosts and other low-melting-point metals, fostering broader innovation across lithium and beyond-lithium energy storage platforms.

Data availability

The standard XRD diffraction patterns were obtained from the DIFFRAC.EVA software purchased from Bruker corporation. The rest of the data supporting this article have been included as part of the ESI.

Conflicts of interest

The authors declare that there is no conflict of interest regarding the publication of this manuscript.

Acknowledgements

We gratefully acknowledge the support of Taiwan's National Science and Technology Council (NSTC) (Grant No. 113-2628-E-006-019, 113-2622-8-006-020, 113-2923-E-006-002, and 113-2923-E-006-011). We sincerely thank the Hierarchical Green-Energy Materials (Hi-GEM) research centre for accessing in situ XRD measurement equipment. We kindly acknowledge the Nanostar U system (XRD005000), and JEOL JEM-2100F TEM equipment belonging to the core facility centre of National Cheng Kung University (NCKU) and thanks to Mr Kun-Hsu-Lee for assisting with SAXS. We also acknowledge the core facility centre at National Yang Ming Chiao Tung University for its expertise and support in using low-temperature XRD analysis. We thank the Global Innovative Centre for Advanced Nanomaterials at the University of Newcastle, Australia, for their collaboration. We also extend our thanks to Dr Sathish Clastinrusselraj Indirathankam for his assistance with BET and XRD analysis, and to Dr Gurwinder Singh for his corrections on the manuscript. The collaborative efforts of these institutions and facilities played a crucial role in the success of our research, and we are grateful for their contributions.

References

  1. J. M. Tarascon and M. Armand, Nature, 2001, 414, 359–367 CrossRef CAS PubMed .
  2. B. Dunn, H. Kamath and J.-M. Tarascon, Science, 2011, 334, 928–935 CrossRef CAS PubMed .
  3. W.-J. Zhang, J. Power Sources, 2011, 196, 13–24 CrossRef CAS .
  4. Y. P. Wu, E. Rahm and R. Holze, J. Power Sources, 2003, 114, 228–236 CrossRef CAS .
  5. T. Kesavan, T. Partheeban, M. Vivekanantha, N. Prabu, M. Kundu, P. Selvarajan, S. Umapathy, A. Vinu and M. Sasidharan, ACS Appl. Mater. Interfaces, 2020, 12, 24007–24018 Search PubMed .
  6. T. Koketsu, B. Paul, C. Wu, R. Kraehnert, Y. Huang and P. Strasser, J. Appl. Electrochem., 2016, 46, 627–633 CrossRef CAS .
  7. C. Zheng, M. Liu, W. Chen, L. Zeng and M. Wei, J. Mater. Chem. A, 2016, 4, 13646–13651 RSC .
  8. D. Zhao, J. Feng, Q. Huo, N. Melosh, G. H. Fredrickson, B. F. Chmelka and G. D. Stucky, Science, 1998, 279, 548–552 CrossRef CAS PubMed .
  9. S. Jun, S. H. Joo, R. Ryoo, M. Kruk, M. Jaroniec, Z. Liu, T. Ohsuna and O. Terasaki, J. Am. Chem. Soc., 2000, 122, 10712–10713 CrossRef CAS .
  10. R. Ryoo, S. H. Joo and S. Jun, J. Phys. Chem. B, 1999, 103, 7743–7746 CrossRef CAS .
  11. H. Zhou, S. Zhu, M. Hibino, I. Honma and M. Ichihara, Adv. Mater., 2003, 15, 2107–2111 CrossRef CAS .
  12. R. Huang, X. Li, Y. Wu, Z. Huang, H. Ye, Y. Niu, L. Li and J. Wang, Chemosphere, 2022, 294, 133761 CrossRef CAS PubMed .
  13. R. Ryoo, S. H. Joo, M. Kruk and M. Jaroniec, Adv. Mater., 2001, 13, 677–681 CrossRef CAS .
  14. A. Eftekhari, Microporous Mesoporous Mater., 2017, 243, 355–369 CrossRef CAS .
  15. Y. Cheng, C. Wang, F. Kang and Y.-B. He, Nanomaterials, 2022, 12, 3656 CrossRef CAS PubMed .
  16. D. Zhao, Y. Zhang, M. Dang, Y. Liu and S. Guo, ACS Appl. Nano Mater., 2024, 7, 15207–15214 CrossRef CAS .
  17. D. Qiu, B. Zhang, T. Zhang, T. Shen, Z. Fang, W. Zhao, J. Xu and Y. Hou, ACS Mater. Lett., 2023, 5, 1488–1496 CrossRef CAS .
  18. D. Qiu, W. Zhao, B. Zhang, M. T. Ahsan, Y. Wang, L. Zhang, X. Yang and Y. Hou, Adv. Energy Mater., 2024, 14, 2400002 CrossRef CAS .
  19. H. Hao, R. Tan, C. Ye and C. Low, Carbon Energy, 2024, 6, e604 CrossRef CAS .
  20. Z. Wang, Y. Zhang, H. Xiong, C. Qin, W. Zhao and X. Liu, Sci. Rep., 2018, 8, 6530 CrossRef PubMed .
  21. S.-k. Lin, C.-l. Cho and H.-m. Chang, J. Electron. Mater., 2014, 43, 204–211 CrossRef CAS .
  22. X. Lin, R. Yuan, Y. Cao, X. Ding, S. Cai, B. Han, Y. Hong, Z. Zhou, X. Yang, L. Gong, M. Zheng and Q. Dong, Chem, 2018, 4, 2685–2698 CAS .
  23. Y. Shi, M. Song, Y. Zhang, C. Zhang, H. Gao, J. Niu, W. Ma, J. Qin and Z. Zhang, J. Power Sources, 2019, 437, 226889 CrossRef CAS .
  24. L. Guo, D. B. Thornton, M. A. Koronfel, I. E. L. Stephens and M. P. Ryan, JPhys Energy, 2021, 3, 032015 CrossRef CAS .
  25. M. Ihrig, L.-Y. Kuo, S. Lobe, A. M. Laptev, C.-A. Lin, C.-H. Tu, R. Ye, P. Kaghazchi, L. Cressa, S. Eswara, S.-k. Lin, O. Guillon, D. Fattakhova-Rohlfing and M. Finsterbusch, ACS Appl. Mater. Interfaces, 2023, 15, 4101–4112 CrossRef CAS PubMed .
  26. P.-c. Tsai, S.-C. Chung, S.-k. Lin and A. Yamada, J. Mater. Chem. A, 2015, 3, 9763–9768 RSC .
  27. Z. Tang, S. Zhou, Y. Huang, H. Wang, R. Zhang, Q. Wang, D. Sun, Y. Tang and H. Wang, Electrochem. Energy Rev., 2023, 6, 8 CrossRef CAS .
  28. S. Karimzadeh, B. Safaei, C. Yuan and T.-C. Jen, Electrochem. Energy Rev., 2023, 6, 24 CrossRef CAS .
  29. S. B. Patil, I. Y. Kim, J. L. Gunjakar, S. M. Oh, T. Eom, H. Kim and S.-J. Hwang, ACS Appl. Mater. Interfaces, 2015, 7, 18679–18688 CrossRef CAS PubMed .
  30. J. Ma, H. Zhang, Y. Xin, S. Liu, Y. Li, L. Yang, G. Xu, T. Lou, H. Niu and S. Yang, Dalton Trans., 2021, 50, 1703–1711 RSC .
  31. X. Meng, K. He, D. Su, X. Zhang, C. Sun, Y. Ren, H.-H. Wang, W. Weng, L. Trahey, C. P. Canlas and J. W. Elam, Adv. Funct. Mater., 2014, 24, 5435–5442 CrossRef CAS .
  32. Y.-H. Lee, Y. Hwa and C.-M. Park, J. Mater. Chem. A, 2021, 9, 20553–20564 RSC .
  33. I.-S. Hwang, Y.-H. Lee, J.-M. Yoon, Y. Hwa and C.-M. Park, Composites, Part B, 2022, 243, 110142 CrossRef CAS .
  34. B.-W. Zhang, L. Ren, Y.-X. Wang, X. Xu, Y. Du and S.-X. Dou, Interdiscip. Mater., 2022, 1, 354–372 Search PubMed .
  35. H. Wang, S. Chen, X. Zhu, B. Yuan, X. Sun, J. Zhang, X. Yang, Y. Wei and J. Liu, Matter, 2022, 5, 2054–2085 CrossRef CAS .
  36. S.-k. Lin, T. L. Nguyen, S.-c. Wu and Y.-h. Wang, J. Alloys Compd., 2014, 586, 319–327 CrossRef CAS .
  37. M. D. Dickey, ACS Appl. Mater. Interfaces, 2014, 6, 18369–18379 CrossRef CAS PubMed .
  38. B. Xu, W. Peng, J. He, Y. Zhang, X. Song, J. Li, Z. Zhang, Y. Luo, X. Meng, C. Cai, Y. Liu, Z. Wei, S. Wang, S. Nie and Q. Duan, Nano Energy, 2024, 120, 109107 CrossRef CAS .
  39. Y.-T. Cheng, R. D. Deshpande, J. Li and M. W. Verbrugge, MRS Proc., 2011, 1333, 1304–1305 Search PubMed .
  40. R. D. Deshpande, J. Li, Y.-T. Cheng and M. W. Verbrugge, J. Electrochem. Soc., 2011, 158, A845 CrossRef CAS .
  41. Y.-H. Lee, I.-S. Hwang, J.-H. Choi and C.-M. Park, J. Energy Storage, 2023, 63, 107067 CrossRef .
  42. J. Wang, L. Wang, Y. Ma and S. Yang, Mater. Lett., 2018, 228, 297–300 CrossRef CAS .
  43. C. Huang, X. Wang, Q. Cao, D. Zhang and J.-Z. Jiang, ACS Appl. Energy Mater., 2021, 4, 12224–12231 CrossRef CAS .
  44. K. T. Lee, Y. S. Jung, T. Kim, C. H. Kim, J. H. Kim, J. Y. Kwon and S. M. Oh, Electrochem. Solid-State Lett., 2008, 11, A21 CrossRef CAS .
  45. J.-M. Yoon, Y.-H. Lee and C.-M. Park, Mater. Today Energy, 2023, 35, 101327 CrossRef CAS .
  46. W. Liang, L. Hong, H. Yang, F. Fan, Y. Liu, H. Li, J. Li, J. Y. Huang, L.-Q. Chen, T. Zhu and S. Zhang, Nano Lett., 2013, 13, 5212–5217 CrossRef CAS PubMed .
  47. S. Rao, R. Wu, Z. Zhu, J. Wu, Y. Ding and L. Mai, Nano Energy, 2023, 112, 108462 CrossRef CAS .
  48. A. Schneider and O. Hilmer, Z. Anorg. Allg. Chem., 1956, 286, 97–117 CrossRef CAS .
  49. H. Azza, N. Selhaoui, A. Iddaoudi and L. Bouirden, J. Phase Equilib. Diffus., 2017, 38, 788–795 CrossRef CAS .
  50. S.-k. Lin, S. Nagao, E. Yokoi, C. Oh, H. Zhang, Y.-c. Liu, S.-g. Lin and K. Suganuma, Sci. Rep., 2016, 6, 34769 CrossRef CAS PubMed .
  51. J. Saint, M. Morcrette, D. Larcher and J. M. Tarascon, Solid State Ionics, 2005, 176, 189–197 CrossRef CAS .
  52. X. Guo, Y. Ding, L. Xue, L. Zhang, C. Zhang, J. B. Goodenough and G. Yu, Adv. Funct. Mater., 2018, 28, 1804649 CrossRef .
  53. A. Vinu, C. Streb, V. Murugesan and M. Hartmann, J. Phys. Chem. B, 2003, 107, 8297–8299 CrossRef CAS .
  54. K. Yan, X. Sun, S. Ying, W. Cheng, Y. Deng, Z. Ma, Y. Zhao, X. Wang, L. Pan and Y. Shi, Sci. Rep., 2020, 10, 6227 CrossRef CAS PubMed .
  55. C.-P. Yang, S. Xin, Y.-X. Yin, H. Ye, J. Zhang and Y.-G. Guo, Angew. Chem., Int. Ed., 2013, 52, 8363–8367 CrossRef CAS PubMed .
  56. F. Han, W.-C. Li, M.-R. Li and A.-H. Lu, J. Mater. Chem., 2012, 22, 9645–9651 RSC .
  57. M.-L. C. Piedboeuf, A. F. Léonard, G. Reichenauer, C. Balzer and N. Job, Microporous Mesoporous Mater., 2019, 275, 278–287 CrossRef CAS .
  58. T. Koketsu, C. Wu, Y. Huang and P. Strasser, J. Appl. Electrochem., 2018, 48, 1265–1271 CrossRef CAS .
  59. Y. Zhang, W. Lu, P. Zhao, M. H. Aboonasr Shiraz, D. Manaig, D. J. Freschi, Y. Liu and J. Liu, Carbon, 2021, 173, 11–21 CrossRef CAS .
  60. M. Hartmann and A. Vinu, Langmuir, 2002, 18, 8010–8016 CrossRef CAS .
  61. M. Yarema, M. Wörle, M. D. Rossell, R. Erni, R. Caputo, L. Protesescu, K. V. Kravchyk, D. N. Dirin, K. Lienau, F. von Rohr, A. Schilling, M. Nachtegaal and M. V. Kovalenko, J. Am. Chem. Soc., 2014, 136, 12422–12430 CrossRef CAS PubMed .
  62. D. Saikia, T.-H. Wang, C.-J. Chou, J. Fang, L.-D. Tsai and H.-M. Kao, RSC Adv., 2015, 5, 42922–42930 RSC .
  63. X. Liu, Z. Yang, H. Quan, J. Li, Y. Xiang and F. Wu, Electrochem. Commun., 2021, 132, 107145 CrossRef CAS .
  64. C. K. Chan, H. Peng, G. Liu, K. McIlwrath, X. F. Zhang, R. A. Huggins and Y. Cui, Nat. Nanotechnol., 2008, 3, 31–35 CrossRef CAS PubMed .
  65. J. Yang, M. Winter and J. O. Besenhard, Solid State Ionics, 1996, 90, 281–287 CrossRef CAS .
  66. B. Gao, S. Sinha, L. Fleming and O. Zhou, Adv. Mater., 2001, 13, 816–819 CrossRef CAS .
  67. J. Zhang, G.-H. Lee, V. Wing-hei Lau, F. Zou, Y. Wang, X. Wu, X.-L. Wang, C.-L. Chen, C.-J. Su and Y.-M. Kang, Acta Mater., 2021, 211, 116863 CrossRef CAS .
  68. N. Cheng, Y. Xue, Q. Liu, J. Tian, L. Zhang, A. M. Asiri and X. Sun, Electrochim. Acta, 2015, 163, 102–106 CrossRef CAS .
  69. S. Zhou, X. Feng, H. Shi, J. Chen, F. Zhang and W. Song, Sens. Actuators, B, 2013, 177, 445–452 CrossRef CAS .
  70. J. Tian, Q. Liu, N. Cheng, A. M. Asiri and X. Sun, Angew. Chem., Int. Ed., 2014, 53, 9577–9581 CrossRef CAS PubMed .
  71. H. Ahmadi, G. Khalaj, F. Soleymani, M. Moalem, M. Pourabdollah and M. Mahmoudan, J. Solid State Electrochem., 2024, 28, 2269–2281 CrossRef CAS .
  72. S. A. Freunberger, Y. Chen, N. E. Drewett, L. J. Hardwick, F. Bardé and P. G. Bruce, Angew. Chem., Int. Ed., 2011, 50, 8609–8613 CrossRef CAS PubMed .
  73. Z. Li, S. Ganapathy, Y. Xu, J. R. Heringa, Q. Zhu, W. Chen and M. Wagemaker, Chem. Mater., 2017, 29, 1577–1586 CrossRef CAS PubMed .
  74. X. Gao, Y. Chen, L. Johnson and P. G. Bruce, Nat. Mater., 2016, 15, 882–888 CrossRef CAS PubMed .
  75. X. Gao, Z. P. Jovanov, Y. Chen, L. R. Johnson and P. G. Bruce, Angew. Chem., Int. Ed., 2017, 56, 6539–6543 CrossRef CAS PubMed .
  76. H. T. T. Le, D. T. Ngo, V.-C. Ho, G. Cao, C.-N. Park and C.-J. Park, J. Mater. Chem. A, 2016, 4, 11124–11138 RSC .
  77. L. Zhang, Y. Ma, X. Cheng, C. Du, T. Guan, Y. Cui, S. Sun, P. Zuo, Y. Gao and G. Yin, J. Power Sources, 2015, 293, 1006–1015 CrossRef CAS .
  78. C. J. Wen and R. A. Huggins, J. Electrochem. Soc., 1981, 128, 1636 CrossRef CAS .
  79. J. Sangster and A. D. Pelton, J. Phase Equilib., 1991, 12, 33–36 CrossRef CAS .
  80. Y. Zhang, L. Tan, Y. Wu, Y. An, Y. Liu, Y. Wang, C. Wei, B. Xi, S. Xiong and J. Feng, Appl. Mater. Today, 2022, 26, 101300 CrossRef .
  81. S.-J. Hong and C. Suryanarayana, J. Appl. Phys., 2004, 96, 6120–6126 CrossRef CAS .
  82. T. h. Huang, J. w. Huang, Z. f. Lin and S. k. Lin, An Investigation of Cu/Ni/Ga Interfacial Reaction with Different Ni/Ga Ratio, in, 2024 International Conference on Electronics Packaging (ICEP), IEEE, Toyama, Japan, 2024, pp. 135–136,  DOI:10.23919/ICEP61562.2024.10535663 .
  83. Q. Li, S. Zhu and Y. Lu, Adv. Funct. Mater., 2017, 27, 1606422 CrossRef .
  84. Y. Ding, N. Yi, F. Zhang, H. Li, T. Zhan, Z. Guo, J. Zhu and J. Chen, Mater. Sci. Eng., B, 2025, 321, 118459 CrossRef CAS .
  85. C.-a. Lin, M. Ihrig, K.-c. Kung, H.-c. Chen, M. Finsterbusch, O. Guillon and S.-k. Lin, J. Mater. Chem. A, 2024, 12, 9438–9453 RSC .
  86. J. Yu, S. Yin, G. Xiong, X. Guan, J. Xia, J. Li, S. Zhang, Y. Xing and P. Yang, J. Electrochem. Energy Convers. Storage, 2022, 20, 031007 CrossRef .
  87. L. Xu, Z. Mao, J. Liu, M. Bi, T. Tu, Y. Tian, X. Zhou, J. Wu, Y. Wu, J. Su, S. Chen and H. Yin, Adv. Energy Mater., 2025, 2501790 CrossRef .

Footnote

Electronic supplementary information (ESI) available: The real-time photograph of in situ analysis sample preparation and video. Detailed discussions of BET, XRD, and SAXS analyses for various mass-loaded CMK-3/Ga composites (CMK-3-100, 130, and 150 series); comprehensive electrochemical performance data with summary tables for different mass loadings; SEM and EDX mapping observations of post-cycling coin cell electrode surface morphology; in situ XRD analysis of pristine CMK-3-130 and CMK-3-130/Ga composites; operando CV-EIS analysis of Li–Ga phase dynamics and interfacial stability, including an electrochemical assignment summary table; and a comparative table of anode material performance. See DOI: https://doi.org/10.1039/d5ta01253h

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.