Erviani Rusman,
Arunakumari Nulu
and
Keun Yong Sohn
*
Department of Nanoscience and Engineering, Inje University, 197 Imje-ro, Gimhae, Gyeongnam-do 50834, Republic of Korea. E-mail: ksohn@inje.ac.kr
First published on 14th August 2025
Germanium-based materials are emerging as promising anode candidates for lithium-ion batteries (LIBs) due to their high theoretical capacity, excellent electrical conductivity, and efficient lithium-ion diffusion at room temperature. However, their practical application is hindered by substantial volume changes during cycling, leading to poor cycling stability and diminished electrochemical performance. This study successfully created a GeO2–Ge@NCNT composite using a cost-effective and environmentally friendly GeO2 precursor through a hydrothermal method. N-doped carbon nanotubes (NCNT) were utilized to enhance electrochemical activity by providing additional active sites and improving structural stability. The GeO2–Ge@NCNT anode demonstrated remarkable cycling stability with a capacity of 1017 mA h g−1 at a current density of 100 mA g−1 after 100 cycles, maintaining a coulombic efficiency of 98.15% and a capacity retention of 71.15% (vs. 2nd cycle discharge capacity). Moreover, C-rate analysis showed exceptional rate capability and consistent long-term cycling performance. This research presents a sustainable and scalable method for producing high-performance Ge-based anodes for next-generation lithium-ion battery technologies.
Extensive research has been dedicated to hybrid materials to address the intrinsic limitations of Ge-based anodes to enhance structural stability and electrochemical performance. The synergistic integration of Ge and GeO2 has demonstrated effectiveness in mitigating electrode pulverization. GeO2 characterized by a high theoretical capacity of 2152 mA h g−1 has emerged as a promising candidate for anode applications.12 During cycling, GeO2 undergoes both conversion and alloying reactions, where the conversion reaction between Ge and Li2O introduces an irreversible process that restricts lithium availability and contributes to capacity fading. It leads to high capacity, but with kinetic limitations.13 However, the integration of Ge in GeO2 acts as a catalyst for Li2O decomposition, facilitating the formation of a core–shell structure, thereby enhancing lithium recovery and mitigating irreversible capacity loss. The Ge–Ge bond length within the shell remains identical to that observed in GeOx crystals, whereas the oxygen shell exhibits a Ge–O bond length comparable to that in GeO2. Nguyen et al. reported that a GeO2/Ge composite synthesized via the reduction of GeO2 using magnesium effectively mitigated volumetric expansion during cycling, thereby maintaining a stable capacity of 520.2 mA h g−1 after 40 cycles.14
GeO2 + 4Li+ + 4e− → Ge + 2Li2O (conversion) |
Ge + xLi+ + xe− → LixGe (0 ≤ x ≤ 4.4) (alloying) |
Several strategies have been explored to improve structural stability and mitigate the adverse effects of volumetric expansion during cycling, including carbonaceous composites, nanostructure, and morphological reconstruction of nanomaterials.15,16 The integration of carbon matrices has emerged as a particularly effective and scalable approach for enhancing cycle stability. Carbon functions as an electrolyte-blocking layer and a dispersing agent, preventing nanosized Ge particle aggregation during repeated charge/discharge cycles. Carbonaceous materials serve as volumetric buffers, alleviating mechanical stress, improving electrical conductivity, and suppressing electrode pulverization.17–19 The integration of conductive carbon within nanocomposite materials significantly enhances charge transport kinetics by shortening lithium-ion diffusion pathways, thereby contributing to superior cycling stability.20–22
Among various carbon-based materials, carbon nanotubes (CNT) have garnered significant attention due to their high thermal stability, large surface area, and exceptional electrical conductivity.23,24 Kaewraung et al. demonstrated the effectiveness of CNT incorporation in a Ge–Fe/MWCNT anode synthesized via a milling-assisted covalent-bonding method, achieving a reversible capacity of 925 mA h g−1 at 100 mA g−1 with an initial coulombic efficiency of 95% and outstanding capacity retention of 75%.25 Furthermore, heteroatom doping using boron (B) and/or nitrogen (N) in CNT frameworks, has been shown to significantly enhance lithium-ion storage capacity, surface reactivity, and conductivity.26,27 Specifically, nitrogen doping effectively tailors the intrinsic physicochemical properties of the carbon host, including structure surface and electronic characteristics, thereby improving electrochemical stability and long cycle life.28 Despite the promising advantages of nitrogen-doped carbon in Ge-based anodes, research in this area remains limited, necessitating further investigations into this emerging strategy.
Prior studies have been conducted to enhance the structural stability and electrochemical performance of Ge-based anodes in LIB. Arro et al. synthesized a GeO2/Ge/r-GO hybrid composite via a ball milling method, achieving a stable capacity of 300 mA h g−1 at a current density of 100 mA g−1 after 100 cycles.29 Wang et al. reported that carbon doping in Ge-based anodes significantly enhances structural integrity, thereby improving cycling stability. This enhancement was attributed to the synergistic interaction between carbon and oxygen, which facilitated a more efficient lithiation process.2 Additionally, Mei et al. employed electrospinning technology to fabricate a GeO2–carbon nanofiber (GeO2–CNF) composite. The result exhibited a high reversible capacity of 1034 mA h g−1 after 200 cycles at a current density of 100 mA g−1.11 Although electrospinning technology enables the synthesis of Ge-based anode materials with superior capacity, its widespread application is constrained by high production costs. Consequently, the development of cost-effective fabrication methods, such as hydrothermal synthesis represents a viable alternative for scalable production of Ge-based anode materials.
Herein, we propose the synthesis of a novel GeO2–Ge@NCNT hybrid anode material using a cost-effective and non-toxic GeO2 precursor via a hydrothermal method. Structural characterization confirmed the presence of Ge and GeO2 lattices within the synthesized material. A comparative analysis between GeO2–Ge and GeO2–Ge@NCNT was conducted to evaluate the role of NCNT in enhancing cycling stability. The electrochemical results demonstrated that GeO2–Ge@NCNT exhibited superior reversible capacity and cycling stability. These findings highlight the potential of GeO2–Ge@NCNT as a potential high-capacity and durable anode material for next-generation lithium-ion batteries.
Fig. 2a presents the XRD pattern of the prepared nitrogen-doped carbon nanotubes (NCNT), exhibiting characteristic diffraction peaks at 25.83° and 42.54°, which correspond to (002) and (100) crystal plane of CNT (JCPDS: 96-101-1061). A small shift was observed at the (200) plane and a small impurity peak disappeared after N-doping identified no side reactions happened during the N-doped process.30,31 Fig. 2b confirms the presence of GeO2 and Ge in both GeO2–Ge@NCNT and GeO2–Ge samples. The diffraction peaks associated with Ge are observed at 28.28°, 53.44°, and 66.02°, corresponding to the crystal planes (111), (311), and (400), respectively (JCPDS: 98-004-4841). Furthermore, the sharp peaks attributed to GeO2 were found at lattice index (100), (101), (110), (102), (111), and (200) which were identified as GeO2 hexagonal (JCPDS: 98-005-3870).14 Additionally, the diffraction peaks of NCNT in the GeO2–Ge@NCNT composite overlap with the GeO2 peaks at 2θ = 25.83° and 42.54°. The presence of NCNT in the composite facilitates rapid electron transfer and enhances electrical conductivity, thereby contributing to the high electrochemical stability and capacity.32
The TGA result reveals the GeO2–Ge@NCNT composite comprises approximately 12.7% carbon, as illustrated in Fig. 3a. An initial gradual weight loss of 1.66% is observed below 450 °C, which is attributed to the evaporation of residual organic solvents and the desorption of physically adsorbed water molecules. A rapid weight reduction occurs within the temperature range of 615–700 °C due to the thermal oxidation of carbon. Beyond 700 °C, the composite mass stabilizes, signifying the complete combustion of carbon and the absence of further decomposition processes.2,33 The Raman spectrum of GeO2–Ge@NCNT in Fig. 3b depicts two peaks at 1342 cm−1 and 1581 cm−1, corresponding to D-band and G-band, respectively. The D-band indicated the structural defect, and the G-band related to the graphitic carbon.20 The intensity ratio of the D-band (ID) and G-band (IG) of GeO2–Ge@NCNT was calculated to be 1.57. It suggests the composite contains more defect sites due to the interruptions sp2 hybridized carbon framework, resulting in more active sites for Li+ insertion which enhances specific capacity. Furthermore, the presence of 2D and D + D′ bands were reported at 2600–3000 cm−1 for the second-order overtone of the D band, suggesting the graphitized sp2 carbon and combination mode revealing the structural defect in the carbon network. It indicates high electrical conductivity and the more active side of GeO2–Ge@NCNT.29
The XPS analysis was carried out to evaluate the chemical bonding states within the GeO2–Ge@NCNT composite. Fig. 4a exhibits the XPS spectrum with characteristic peaks at binding energies of 32.9 eV, 124.5 eV, 185.2 eV, 284.3 eV, 318.2 eV, 348.9 eV, and 532.8 eV, corresponding to Ge 3d, Ge 3p, Ge 3s, C 1s, Ge KLL, N 1s, and O 1s, respectively. The high-resolution XPS spectrum of O 1s (Fig. 4b) demonstrates three distinct peaks at 530.3 eV, 530.9 eV, and 531.4 eV are attributed to the O–Ge, CC, and C–O functional groups, respectively. The high-resolution C 1s spectrum (Fig. 4c) displays peaks at 283.1 eV, 284.3 eV, 286 eV, and 289.6 eV, corresponding to the C–Ge, C–C, C–O, and C
O bonding states, respectively. The Ge 3d spectrum (Fig. 4d) exhibits multiple peaks at 30.4 eV, 31.2 eV, 31.8 eV, and 32.9 eV, indicating Ge–Ge, Ge–C, Ge–O, and Ge–N, respectively. The Ge–C functional group plays a crucial role in stabilizing GeO2 during the lithiation/delithiation process, thereby improving structural integrity and overall electrochemical performance.34 Furthermore, the high-resolution N 1s spectrum (Fig. 4e) identifies three characteristic peaks of pyridinic N, pyrrolic N, and graphitic N at 397.9 eV, 400.8 eV, and 403.2 eV, respectively. The spectrum of N 1s confirmed the successful N-doping in NCNT elaborating Li+ diffusion properties in the Ge-based composite.
![]() | ||
Fig. 4 XPS spectra (a) survey scan of GeO2–Ge@NCNT; high-resolution XPS spectrum of (b) O 1s, (c) C 1s, (d) Ge 3d, and (e) N 1s in GeO2–Ge@NCNT. |
SEM and TEM analysis was conducted to investigate the morphological characteristics of the GeO2–Ge@NCNT composite. Fig. S1 illustrates that the NCNT's nanotubular structure uniformly encapsulates GeO2–Ge particles. The GeO2–Ge domains have a spherical shape, with an average particle size of 400–700 nm, indicating uniform dispersion throughout the composite matrix. As shown in Fig. 5, the GeO2–Ge particles are surrounded by NCNT, indicating a well-integrated composite structure. Fig. 5b confirms the presence of NCNT within the composite with tubular structures as highlighted by the yellow arrow. The darker regions observed in Fig. 5c indicate oxygen-rich areas, further corroborated by energy-dispersive X-ray spectroscopy (EDX) analysis. The selected regions from high-resolution TEM (HR-TEM) in Fig. 5d was examined in detail to determine the crystal lattice spacing using lattice fringe analysis. As depicted in Fig. 5e and f, the lattice fringe patterns revealed interplanar spacings of 0.327 nm and 0.33 nm, corresponding to the Ge (111) and GeO2 (101) crystal planes, respectively.
![]() | ||
Fig. 5 TEM image of GeO2–Ge@NCNT in different magnifications: (a) 1 μm, (b) 200 nm, (c) 100 nm, (d) HR-TEM image, and (e and f) lattice fringes of a selected area from image (d). |
Fig. 6a–e reports EDX mapping which confirmed the presence of C, N, Ge, and O elements in the synthesized GeO2–Ge@NCNT composite. The uniform distribution of nitrogen which overlaps with carbon verifies the successful incorporation of NCNT in the GeO2–Ge structure. The elemental composition is further supported by the EDX spectrum of GeO2–Ge@NCNT, as depicted in Fig. 6f. The specific percentages of elemental composition of the GeO2–Ge@NCNT composite are summarized in Table 1.
![]() | ||
Fig. 6 (a) STEM images and corresponding EDX elemental mapping, indicating (b) C, (c) Ge, (d) N, (e) O signal mapping of GeO2–Ge@NCNT image, and (f) EDX pattern of GeO2–Ge@NCNT. |
Element | Mass% | Atom% |
---|---|---|
C | 29.41 | 59.12 |
N | 0.53 | 0.91 |
O | 14.18 | 21.40 |
Ge | 55.88 | 18.57 |
The galvanostatic charge/discharge profile of the GeO2–Ge@NCNT composite was evaluated at a current density of 100 mA g−1 within a voltage window of 0.01–2.00 V vs. Li/Li+, as depicted in Fig. 7a and b. The charge/discharge profile of GeO2–Ge (Fig. 7a) demonstrates initial discharge and charge capacities of 2271 mA h g−1 and 1157 mA h g−1, respectively. In contrast, Fig. 7b displays the corresponding profiles for GeO2–Ge@NCNT, revealing initial discharge and charge capacities of 2198 mA h g−1 and 1351 mA h g−1, respectively, resulting in an initial coulombic efficiency (CE) of 61.47%. The relatively low CE observed in the first cycle is primarily ascribed to the development of the solid electrolyte interphase (SEI) layer and the irreversible Ge–Li alloying reaction.2,34 During the second cycle, the discharge/charge capacities of GeO2–Ge and GeO2–Ge@NCNT are recorded as 1260/1216 mA h g−1 and 1429/1367 mA h g−1, respectively, with corresponding CE values of 96.54% and 95.64%. GeO2–Ge composite exhibits a more capacity loss between the first and second cycles compared to the GeO2–Ge@NCNT composite, highlighting the role of NCNT in improving cycling stability. After 100 cycles, the discharge/charge capacities of GeO2–Ge and GeO2–Ge@NCNT decrease to 344/342 mA h g−1 and 1017/1002 mA h g−1 with capacity retention values of 27.23% and 71.15%, respectively. These results indicate that NCNT effectively stabilizes the active material by mitigating volume expansion and suppressing capacity degradation during Li+ insertion/extraction reaction, thereby enhancing the structural integrity and electrochemical performance.
Fig. 7c presents a comparative analysis of the long-term cycling performance of CNT, GeO2–Ge, and O2–Ge@NCNT electrodes. The GeO2–Ge electrode exhibits the highest initial discharge capacity of 2271 mA h g−1 among the tested electrodes. However, a rapid capacity reduction is observed with the discharge capacity declining to 345 mA h g−1 after 100 cycles, corresponding to a capacity retention of only 27.37%. This significant capacity fading is primarily attributed to the extensive volume expansion during the lithium insertion/extraction process, which induces structural fractures and particle pulverization. In contrast, the NCNT electrode demonstrates superior structural stability over 100 cycles but low capacity. Consequently, NCNT is a promising dopant for GeO2–Ge, reinforcing the composite structure and mitigating large-scale fractures during the lithiation/delithiation process. As a result, the GeO2–Ge@NCNT composite exhibits markedly enhanced cycling performance compared to GeO2–Ge and NCNT. After 100 cycles at a current density of 100 mA g−1 within a voltage range of 0.01–2.00 V, GeO2–Ge@NCNT retains a high reversible discharge capacity of 1017 mA h g−1 with a remarkable coulombic efficiency exceeding 98.58% and a capacity retention of 71.15%.
Fig. 7d exhibits the rate performance tests conducted at varying current densities of 0.1, 0.2, 0.4, 0.8, and 1.6 A g−1. The GeO2–Ge@NCNT composite demonstrates excellent recovery of reversible capacity when the current density is reduced from 1.6 A g−1 to 0.1 A g−1. Following high-rate cycling, GeO2–Ge@NCNT shows an impressive reversible capacity of 1006 mA h g−1 and stabilizes around 1100 mA h g−1 at 0.1 A g−1. These results highlight the effectiveness of NCNT incorporation in stabilizing the Ge-based composite, thereby significantly improving both electrochemical performance and long-term cycle stability.
The rate capability performance of the electrodes at varying current densities was evaluated, as depicted in Fig. 8a and b. The charge–discharge profiles provide a comparative analysis of the rate capacities of GeO2–Ge and GeO2–Ge@NCNT during the first cycle across a range of current densities from 0.1 A g−1 to 1.6 A g−1. As illustrated in Fig. 8a, the reversible capacities of the GeO2–Ge electrode at increasing current densities of 0.1, 0.2, 0.4, 0.8, and 1.6 A g−1 are recorded as 2354, 1094, 823, 589, and 240 mA h g−1, respectively. A sharp decline in capacity at high current densities indicates the structural instability of GeO2–Ge, where severe mechanical stress and structural fracturing lead to rapid loss of capacity. Conversely, the GeO2–Ge@NCNT composite demonstrates significantly improved rate capability and structural stability under high-current conditions. As shown in Fig. 8b, the GeO2–Ge@NCNT electrode exhibits superior reversible capacities of 2165, 1221, 1069, 903, and 699 mA h g−1 at current densities of 0.1, 0.2, 0.4, 0.8, and 1.6 A g−1, respectively. These findings underscore the critical role of NCNT incorporation in enhancing electrical conductivity, facilitating Li-ion diffusion, and stabilizing the electrochemical cycling process, particularly under high-rate conditions.
![]() | ||
Fig. 8 Charging and discharging profile of (a) GeO2–Ge and (b) GeO2–Ge@NCNT at different C-rate (0.1–1.6 A g−1) in the first cycle. |
Fig. 9a and b illustrate the cyclic voltammetry (CV) profiles of the GeO2–Ge and GeO2–Ge@NCNT electrodes, recorded over a voltage range of 0.01–2.00 V. The formation of the solid electrolyte interphase (SEI) during the initial lithiation process consequently affected the distinct features observed in the cyclic voltammetry (CV) curves during the first cycle for both electrodes.35 As depicted in Fig. 9a, the GeO2–Ge electrode exhibits two prominent oxidation peaks. During the oxidation process, a sharp peak is observed at 0.55 V and a broader peak at 1.15 V. During the reduction process, a peak emerges at 0.13 V and an additional peak appears at 0.11 V after five cycles. These characteristic peaks correspond to the alloying reaction between Ge and Li+ and the conversion reaction of GeO2, which typically occurs within the voltage range of 0.11–0.50 V.29,34
Fig. 9b exhibits the CV profile of GeO2–Ge@NCNT electrodes. During the delithiation process, two distinct peaks are observed at 0.45 V and 1.15 V. The sharp peak at 0.45 V corresponds to the oxidation of GeO2. The increased intensity is likely due to a substantial transition from the deintercalation of Li+ ions in the NCNT toward the more dominant Li–Ge de-alloying process during delithiation. The broad peak at 1.15 V serves as an indicator of the oxidation of GeO2. It suggests that Ge encapsulated within the NCNT matrix undergoes oxidation.29,36,37
The sharper and broader peaks observed in the GeO2–Ge@NCNT electrode during the lithiation and delithiation processes indicate enhanced electrochemical kinetics facilitated by the NCNT network. This network improves charge transport and promotes structural stability.38 In the reduction reaction, the CV scan reveals a peak at 0.11 V during the initial cycle, associated with the solid electrolyte interphase (SEI) layer formation. As cycling progresses, the intensity of this 0.11 V peak diminishes, while the sharp peak at 0.49 V intensifies, indicating the development of a stable SEI layer. Additionally, the decreasing peak intensity at 0.11 V confirms the lithiation of GeO2, signifying the decomposition of GeO2 and the initial formation of the Li–Ge alloying process.39–41 Fig. 10 depicts the surface morphology change of GeO2–Ge@NCNT before and after 100 cycles at 100 mA g−1. The SEM images identified that the GeO2–Ge@NCNT intensively retains the surface from extreme cracking formation, maintaining the electrode integrity, which leads to the electrode stability during the repeated lithiation–delithiation process.
XPS characterization conducted after electrochemical cycling offers detailed insights into the surface chemistry evolution of the GeO2–Ge@NCNT electrode. Fig. S2 illustrates a pronounced increase in the O 1s peak intensity, along with the emergence of a distinct Li 1s signal, which is indicative of SEI as a key factor in stabilizing electrode–electrolyte interactions and ensuring long-term cycling performance in Li-ion batteries.42,43 Simultaneously, a reduction in the C 1s and Ge 3d signal intensities reflects partial surface passivation due to SEI deposition and reaction product accumulation, which is commonly observed in conversion and alloying anodes.44 The peaks at C 1s signal after long-term cycling were observed to be higher and clearer. The increasing intensity of group C–C and CO in the C 1s signal supports the presence of organic species derived from electrolyte decomposition.45 Moreover, the weak N 1s signal is shown to nearly disappear in the survey spectrum, and from the high-resolution XPS of the N 1s signal, only the pyridinic N function group was observed after long-term cycling. It suggests either encapsulation of the N functional group by SEI or degradation of the N-doped carbon framework. Despite these transformations, the presence of Ge 3d and Li 1s signals post-cycling confirms continued lithium storage activity, underscoring the electrochemical reversibility of the Ge-based phase. These findings collectively demonstrate that the GeO2–Ge@NCNT architecture undergoes expected interfacial modifications while retaining functional lithium-reactive sites.
Fig. 11a presents a comparative analysis of the electrochemical resistance of GeO2–Ge and GeO2–Ge@NCNT based on Nyquist plots obtained from AC impedance spectroscopy. The tilted line at the low-frequency range corresponds to the diffusion of lithium kinetics to the electrode. In addition, semi-cycle at the middle and high-frequency range refers to charge transfer resistance in the electrode and electrolyte interface.46 The total resistance of GeO2–Ge@NCNT before cycling is lower (257 Ω) than that of GeO2–Ge (351 Ω), which is primarily attributed to the incorporation of NCNT. The introduction of NCNT in composite enhances electrical conductivity and facilitates more efficient redox kinetics.1,47 The impedance characteristics illustrated in Fig. 11b further depict Nyquist plots for GeO2–Ge and GeO2–Ge@NCNT after 10 cycles. A reduction in resistance is observed for both electrodes after cycling, where GeO2–Ge@NCNT shows a more significant decrease. This result confirms its enhanced structural stability and improved electrical conductivity during redox processes, thereby reinforcing its potential as an advanced anode material.
![]() | ||
Fig. 11 Nyquist plots of AC impedance for GeO2–Ge and GeO2–Ge@NCNT (a) before cycling and (b) after 10 cycles, and (c) equivalent circuit model of EIS fitting curve. |
The equivalent circuit model from Fig. 11c consists of multiple resistance components: Rs represents the total resistance, encompassing electrolyte resistance, separator resistance, and electrical contact resistance. Furthermore, RSEI and Rct correspond to the resistance of the solid electrolyte interphase (SEI) layer and charge transfer resistance, respectively, while the constant phase elements CPE1 and CPE2 denote capacitance and charge transfer capacitance.2,48 According to Table 2, the Rs and Rct values of GeO2–Ge@NCNT before cycling (9.8 Ω and 200 Ω) are less than that of GeO2–Ge (12 Ω and 215 Ω), indicating a more efficient electron transport pathway facilitated by the superior conductivity of NCNT. Moreover, the Rct of GeO2–Ge@NCNT exhibits a substantial reduction after cycling, suggesting that NCNT effectively stabilizes the electrode structure and mitigates the depletion of active material. Thus, NCNT serves as an effective component in reducing electrolyte penetration and reinforcing the SEI layer during charge/discharge cycles, ultimately contributing to enhancing coulombic efficiency and capacity retention.25,30 Compared with previous research related to Ge-based anodes, as reported in Table 3, GeO2–Ge@NCNT provided high reversible capacity and excellent stability performance for long-term cycling which potentially supports it as durable anode material for LiB.
Samples | Before cycling | After 10 cycles | ||||
---|---|---|---|---|---|---|
Rs (Ω) | RSEI (Ω) | Rct (Ω) | Rs (Ω) | RSEI (Ω) | Rct (Ω) | |
GeO2–Ge | 12 | 124 | 215 | 5.6 | 4.8 | 5.3 |
GeO2–Ge@NCNT | 9.8 | 47.2 | 200 | 4 | 1.5 | 3.5 |
Electrode | Specific capacity (mA h g−1) | Current density (mA g−1) | Cycle | Ref. |
---|---|---|---|---|
GeO2 CNF | 1034 | 100 | 200 | 11 |
GeO2/Ge | 502.2 | 100 | 40 | 14 |
3D void/Ge@C | 990 | 100 | 100 | 17 |
Ge–Fe/MWCNT | 491 | 5000 | 1000 | 25 |
GeO2/Ge/r-GO | 300 | 100 | 100 | 29 |
GeOx-coated rGO | 758 | 2000 | 700 | 33 |
GeO2/C | 984.2 | 200 | 200 | 34 |
GeO2/Ge | 665.3 | 500 | 100 | 39 |
GeO2@C | 940 | 50 | 50 | 49 |
CNT–GeO2 | 844.8 | 500 | 500 | 50 |
GeO2–Ge@NCNT | 1017 | 100 | 100 | This work |
EIS studies. See DOI: https://doi.org/10.1039/d5ra02163d.
This journal is © The Royal Society of Chemistry 2025 |