Wang
Yan
a,
Wang
Hanbo
a,
Xu
Yahui
a,
Zhu
Dongyu
a,
Wang
Ziming
a,
Li
Yiduo
ad,
Tian
Yumei
*a and
Lu
Haiyan
*abc
aCollege of Chemistry, Jilin University, Changchun, 130012, People's Republic of China
bCollege of Chemistry, State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, Key Laboratory of Physics and Technology for Advanced Batteries, Ministry of Education, Jilin University, Changchun 130012, PR China
cAdvanced Energy Systems and Intelligent Detection Technological Innovation University Enterprise Joint Laboratory of Jilin Province, PR China
dCollege of Instrumentation and Electrical Engineering, Jilin University, Changchun, 130021, China
First published on 11th October 2025
The development of sustainable energy storage technologies is critical in addressing the global challenges posed by climate change. Supercapacitors, while offering exceptional power density and cycling stability, suffer from relatively low energy density, limiting their widespread use in large-scale energy storage systems. To overcome this limitation, we designed a novel composite electrode material featuring a core–shell structure. The core derived from well-defined ZIF-67 nanocubes (NCs) was innovatively processed into a hollow structure, which enhanced ion diffusion and increased the overall energy storage capacity by reducing internal resistance. Meanwhile, the shell consisted of 3D hierarchical Ni–Co layered double hydroxides (NiCo-LDH) grown in situ employing an ambient-temperature method, offering high electrochemical activity and abundant active sites for efficient charge storage. The ultimately synthesized multi-scale hollow core–shell material, Co3O4-HNC@NiCo-LDH, integrated the respective merits of the shell and core materials, while simultaneously addressing issues that arise when these materials exist in isolation. It effectively mitigated problems such as volume expansion and agglomeration that materials might encounter during electrochemical reactions, thereby further enhancing the materials’ performance and service life. Notably, in situ Raman spectroscopy was utilized to trace the dynamic redox processes and structural changes occurring during electrochemical cycling, thereby validating the stability and effectiveness of the charge storage mechanism. The resulting material, Co3O4-HNC@NiCo-LDH, demonstrated impressive capacitance (1862.4 F g−1 at 2 A g−1), high energy density (76.8 Wh kg−1 at 2 A g−1), and excellent cycling stability (98.38% after 15
000 cycles at 15 A g−1), offering a promising solution for next-generation supercapacitors.
Metal–organic frameworks (MOFs) and their derivatives, along with polymetallic hydroxides and transition metal oxides/sulfides, have emerged as frontrunners among pseudocapacitive materials, owing to their exceptional electrochemical properties.9,10 These materials, prized for their high surface areas, tunable structures, and remarkable charge storage capabilities, are at the forefront of modern energy storage research.11 Despite their promising electrochemical potential, they often face inherent limitations, including poor electrical conductivity, structural instability under prolonged cycling, and suboptimal ion diffusion rates, all of which degrade their overall performance in supercapacitor applications.12,13 In addressing the aforementioned challenges, we meticulously engineered and refined both the core and shell components of these sophisticated architectures. Through a nuanced design approach, we enhanced the core material to optimize its structural stability and electrochemical activity, while simultaneously improving the shell material to elevate its conductive properties and surface area. This approach not only addressed the inherent challenges but also maximized the potential of core–shell architectures for advanced energy storage applications, offering unprecedented improvements in charge storage capacity, cycling stability, and overall material efficiency.14
We undertook a meticulous design and refinement process for both the core and shell components, aiming to address the challenges outlined above. By strategically enhancing the properties of each phase, we were able to optimize their individual contributions, ensuring that the core material provided structural integrity and the shell material maximized electrochemical activity. An especially effective strategy for enhancing the electrochemical performance of core materials was the creation of hollow structures, which reduced internal resistance and promoted more efficient ion diffusion. This approach maximized the available surface area for electrochemical reactions, while simultaneously increasing the overall ion storage capacity by providing ample space for ion intercalation. Hollow structures mitigated internal resistance by facilitating ion diffusion through interconnected cavities, which significantly improved charge/discharge efficiency and boosted the rate capability of materials.15 In the case of shell structures, the in situ growth of nanosheet architectures, particularly those based on layered double hydroxides (LDHs), has proven to be a highly effective strategy for significantly enhancing the performance of composite materials, offering clear advantages over traditional manufacturing techniques.16–18 LDHs are renowned for their exceptional mechanical strength, multifunctional ion-exchange properties, and superior theoretical capacitance, which make them ideal candidates for in situ growth applications, facilitating the development of materials with superior electrochemical characteristics.19 In particular, bimetallic LDHs exhibit inherent synergistic effects, optimal active sites, and excellent redox reversibility, further enhancing their electrochemical performance.20,21 Furthermore, the design and development of advanced synthetic methods for fabricating novel environmentally friendly, room-temperature-controllable nanostructures with tailored properties proved to be an effective strategy in materials science.22 NiCo-LDH, as a battery-type material, benefits from the synergistic coupling of Ni2+/Ni3+ and Co2+/Co3+/Co4+ redox reactions, which enhance both charge storage capacity and cycling stability.23
In this work, we integrated the various advantages mentioned above by optimizing and designing a novel composite electrode material with a core–shell structure. The material was composed of NiCo-LDH nanosheets (NSs), which were grown in situ on Co3O4 hollow nanocubes (Co3O4-HNCs) derived from ZIF-67 to form a hierarchical structure. The synthesis followed a multi-step process, beginning with precise ratio control to produce uniformly structured and well-defined ZIF-67 nanocubes (ZIF-NCs). An etching process was then employed to create hollow ZIF-67 nanocubes (HZIF-NCs), followed by calcination to yield Co3O4-HNCs. This etching process generated a hollow and uniform mesoporous structure, significantly improving the overall electrochemical performance by facilitating more efficient reactions and ion diffusion. After carbonization, the fragile HZIF-NCs were transformed into robust Co3O4-HNCs with a nitrogen-doped carbon framework, which enhanced both the structural stability and electrochemical properties of the core material. In the final phase of the synthesis, NiCo-LDH NSs were grown in situ under room temperature conditions. This step was meticulously fine-tuned to regulate the quantity of the coating, ensuring a precise balance between the core and shell layers. Meanwhile, in situ Raman spectroscopy was utilized to investigate the redox dynamics of the Co3O4-HNC@NiCo-LDH core–shell structure, revealing reversible transition metal redox processes and stable framework retention, thereby confirming its excellent electrochemical resilience. The resulting robust structure, with abundant active sites, effectively promoted the intercalation of ions, exhibiting a remarkable increase in specific capacitance, achieving 1862.4 F g−1 at 2 A g−1. The assembled asymmetric supercapacitor (ASC) established an energy density of 76.8 Wh kg−1 at a power density of 829.2 W kg−1. Moreover, the ASC delivered exceptional stability, retaining 98.38% of its initial capacitance after 15
000 cycles, showcasing its potential for reliable performance in real-world applications. The floating test is a powerful tool for decoupled voltage tracking of the positive and negative electrodes of the ASC, facilitating precise analysis of potential windows and internal losses, which is crucial for optimizing electrode pairing and improving the electrochemical stability of ASCs. By employing environmentally friendly, room-temperature-controllable methods for LDH shell growth, this strategy elevated both sustainability and electrochemical performance, offering improved long-term stability under high current densities and positioning these materials as key candidates for next-generation energy storage technologies.
:
1 volume ratio of water to ethanol. The resulting solution was stirred vigorously at room temperature for 10 min to ensure complete dispersion and interaction. Subsequently, the HZIF NCs were separated by centrifugation, thoroughly washed with ethanol and water multiple times, and then dried at 80 °C overnight. With the incorporation of 0.9 to 1.5 g of tannic acid, the resulting samples were labeled as HZIF-0.9, HZIF-1.2, and HZIF-1.5.
:
0.5
:
1. This composite mixture was then dispersed in absolute ethanol to form a homogeneous slurry. The resultant slurry was uniformly applied to nickel foam (1 × 2 cm2), followed by air-drying at 60 °C overnight to ensure optimal binding and a uniform distribution. The final electrodes exhibited a mass loading of the active material of approximately 1.6 mg cm−2.
Electrochemical evaluation was performed in a 6.0 M KOH aqueous electrolyte. The electrochemical characterization included a suite of advanced techniques, namely, cyclic voltammetry (CV), galvanostatic charge/discharge (GCD) testing, and electrochemical impedance spectroscopy (EIS). For all measurements, a standard three-electrode configuration was employed, wherein platinum foil (1 × 1 cm2) acted as the counter electrode and an Hg/HgO electrode was utilized as the reference electrode.
Based on the calculations using the formula (C1V1M1 = C2V2M2), the mass of the positive electrode was determined to be 1 mg, while the mass of the negative electrode was calculated to be 3 mg to match the capacity.
The micromorphology of the samples was characterized using SEM and TEM. According to Fig. S1(a–f), the synthesized ZIF-67 NCs exhibited a highly regular cubic structure, while the etched HZIF-1.2 formed a cavity cube with reduced dead space, which facilitated ion penetration and transport within the material. Subsequently, NiCo-LDH was then grown in situ on Co3O4-HNCs that were produced by calcining HZIF-1.2 at ambient temperature. During this process, the original morphology of the Co3O4-HNCs was preserved, serving as the core structure for the surface growth of NiCo-LDH.
SEM images of Co3O4-HNC@NiCo-LDH-0.5, Co3O4-HNC@NiCo-LDH-1, and Co3O4-HNC@NiCo-LDH-2 are presented in Fig. 1(a–c). The low magnification SEM images corresponding to Co3O4-HNC@NiCo-LDH are shown in Fig. S1(g–i). The NiCo-LDH layer enveloping the Co3O4-HNCs expanded the material in all directions, thereby providing a larger specific surface area and more active sites. This unique core–shell structure offered rapid ion transport pathways while combining the benefits of both components to optimize the sample's electrochemical performance. Notably, Co3O4-HNC@NiCo-LDH-1 demonstrated the most favorable morphology with more uniform lamellar growth. In contrast, Co3O4-HNC@NiCo-LDH-0.5 exhibited sparse and less-loaded lamellae, resulting in a contraction of the specific surface area and scarcity of active sites. The surface of Co3O4-HNC@NiCo-LDH-2 was completely covered with NiCo-LDH, resulting in significant agglomeration and the lowest specific surface area, which was detrimental to electrolyte diffusion. The advantages of the Co3O4-HNC@NiCo-LDH-1 sample were further highlighted by the corresponding TEM images (Fig. 1(e)), and the TEM images for the other samples are shown in Fig. 1(d and f). The Co3O4-HNC width was approximately 500 nm, providing an ideal environment for the growth of the NiCo-LDH lamellar structure (width ∼140 nm). The shell thicknesses of all composite material products are listed in Table S1 according to TEM. Fig. 1(g) shows the HRTEM image of the optimally proportioned Co3O4-HNC@NiCo-LDH-1. The interplanar spacings of the three lattice fringes were 0.22 nm, 0.24 nm, and 0.28 nm, corresponding to the (015) and (012) crystal faces of NiCo-LDH and the (220) crystal face of Co3O4-HNCs, respectively, further confirming the successful synthesis of the Co3O4-HNC@NiCo-LDH-1 sample.24,25 Additionally, the SAED pattern in Fig. 1(h) reveals diffraction spots and rings. To elucidate the elemental distribution within the sample, Fig. 1(i–n) presents the HAADF-STEM image and the corresponding EDS mapping, indicating that C, Co, Ni, N, and O elements are uniformly distributed throughout the sample. The elemental spectra and compositional ratios of the Co3O4-HNC@NiCo-LDH-1 sample are presented in Fig. S2 and Table S2, respectively.
XRD analysis was utilized to characterize the crystal structures and phase information. Fig. 2(a) and Fig. S3(a–e) show the XRD patterns of all materials. The characteristic peaks of the prepared ZIF-67 NCs, HZIF-1.2, Co3O4-HNCs, and NiCo-LDH aligned with those previously reported. For Co3O4-HNC@NiCo-LDH, the diffraction peaks at 19.1°, 31.3°, 36.9°, 38.5°, 44.8°, 59.4°, 65.2°, and 78.4° corresponded to the (111), (220), (331), (220), (222), (400), (511), and (440) crystal planes of Co3O4-HNCs (PDF#42-1467), respectively. The diffraction peaks at 11.7°, 22.8°, 34.2°, and 63.0° for NiCo-LDH are attributed to the (003), (006), (009), and (110) crystal planes.26,27 The intensity of the characteristic peaks of Co3O4-HNCs in Co3O4-HNC@NiCo-LDH was lower than that of pure Co3O4-HNCs. This reduction in peak intensity was due to the coating of Co3O4-HNCs with NiCo-LDH, which attenuated the diffraction signals from the underlying Co3O4-HNCs. Fig. 2(b) presents the Raman spectra of the composites. The peak at 524 cm−1 was characteristic of NiCo-LDH, while the peak at 662 cm−1 corresponded to the vibration of M–O (Ni–O, Co–O).28 The Raman peak observed at 1040 cm−1 was ascribed to the vibrational mode associated with the CO32− groups that were intercalated between the NiCo-LDH layers.29 All composites exhibited a D band and a G band; the comparable intensities of the D and G peaks indicated a well-balanced structure featuring both graphitized domains and defect sites introduced during the calcination process. These results further confirmed the successful integration of NiCo-LDH onto the surface of Co3O4-HNCs. The FTIR spectra of Co3O4-HNCs, NiCo-LDH, and Co3O4-HNC@NiCo-LDH composites were recorded to confirm the successful hybridization. Fig. S4 reveals that Co3O4-HNCs exhibited a characteristic absorption band at ∼570 cm−1, corresponding to the Co–O stretching vibration of the spinel phase.30 For NiCo-LDH, distinct bands at ∼3440 cm−1 and ∼1620 cm−1 were attributed to the O–H stretching and bending vibrations of interlayer water molecules, while the peak at ∼1380 cm−1 was attributed to intercalated NO3− anions.31,32 In the spectrum of the Co3O4-HNC@NiCo-LDH hybrids, the coexistence of all major absorption bands confirmed the formation of a core–shell architecture without significant disruption of the original frameworks.
Fig. 2(c) shows the nitrogen adsorption isotherms of the synthesized samples. All composite samples exhibited type IV isotherms, indicating the presence of micropores, mesopores, and macropores.33 This hierarchical porous structure provided effective pathways for electrolyte ion diffusion, facilitating the full utilization of the electrode material. It was found that the SSA initially increased and then decreased with the addition of NiCo-LDH. Co3O4-HNC@NiCo-LDH-1 had the highest SSA at 172.6 m2 g−1, while Co3O4-HNC@NiCo-LDH-0.5 and Co3O4-HNC@NiCo-LDH-2 had SSAs of 141.6 m2 g−1 and 80.7 m2 g−1, respectively. This trend was attributed to the initial increase in SSA due to the presence of NiCo-LDH, which enhanced the overall surface area. However, excessive NiCo-LDH led to agglomeration, reducing the SSA. Fig. 2(d) reveals that the mesopores in Co3O4-HNC@NiCo-LDH-2 are primarily concentrated at 4.3 nm, with macropores at 53.3 nm. In contrast, the macropores in Co3O4-HNC@NiCo-LDH-0.5 and Co3O4-HNC@NiCo-LDH-1 were mainly concentrated at 56.3 nm. The combined nitrogen adsorption and desorption curve and pore size distribution of a single composite sample are shown in Fig. S5(a–c).
To clarify their surface chemical characteristics and the elemental oxidation states, all samples underwent further XPS analysis. The total spectra of ZIF-67 NCs, HZIF-1.2, and Co3O4-HNCs, as well as the spectra of each element, are shown in Fig. S6–S9. The peak positions of C, N, O, and Co in ZIF-67 NCs and HZIF-1.2 were essentially identical. In contrast, partial oxidation of Co2+ to Co3+ was observed in Co3O4-HNCs, which enhanced electrochemical performance. The full-scan spectra of the composite Co3O4-HNC@NiCo-LDH are shown in Fig. 3(a). The electrode materials were primarily composed of Co, Ni, O, N, and C elements. In Fig. 3(b), the peaks at 780.4 eV and 795.7 eV correspond to Co2+, while those at 781.8 eV and 797.3 eV correspond to Co3+. Satellite peaks were located at 785.6 eV and 802.4 eV, respectively.34Fig. 3(c) illustrates that for Ni 2p, the peaks at 855.5 eV and 873.2 eV are attributed to Ni2+, whereas those at 856.7 eV and 874.6 eV are attributed to Ni3+, and satellite peaks are observed at 861.6 eV and 879.5 eV.35 These results confirmed the successful integration of properties from individual materials in Co3O4-HNC@NiCo-LDH. The increased intensity of the Co and Ni peaks with higher feed ratios indicated the formation of the NiCo-LDH layer on the surface. In Fig. 3(d), the C 1s XPS spectra are deconvoluted into three distinct peaks at binding energies of 284.8 eV (C–C/C
C), 286.1 eV (C–O/C–N), and 288.1 eV (O–C
O).36 For N 1s (Fig. 3(e)), the peaks at 398.9 eV (C–N), 399.8 eV (pyrrolic N), 402.4 eV (N–O), and 406.6 eV (nitrate) were identified.37 The O 1s spectra (Fig. 3(f)) showed three peaks at 530.4 eV (M–O), 531.2 eV (M–OH), and 532.3 eV (adsorbed H2O).38 Fig. S10–S12 specifically show the full spectra of Co3O4-HNC@NiCo-LDH and each element.
![]() | ||
| Fig. 3 (a) XPS survey spectra of the composites Co3O4-HNC@NiCo-LDH; (b–f) Co 2p, Ni 2p, C 1s, N 1s, and O 1s spectra of the composites Co3O4-HNC@NiCo-LDH. | ||
After the carbonization of HZIF-1.2 with the optimal etching ratio, the fragile HZIF-1.2 was transformed into robust Co3O4-HNCs with a nitrogen-doped carbon framework, and varying proportions of NiCo-LDH were subsequently grown in situ on this foundation. To demonstrate that the enhanced performance of the composite material was not attributable to any single component, Fig. S17 shows the CV and GCD curves of Co3O4-HNCs, respectively, while Fig. S18 displays the corresponding CV and GCD curves of NiCo-LDH. The quantity of shell structures developed on the core structure was modulated by altering the ratio of the reaction solution. Fig. 4(a) shows the CV curves of NiCo-LDH and Co3O4-HNC@NiCo-LDH at 20 mV s−1. Among the samples tested, the CV curve of pure NiCo-LDH enclosed the smallest area, while the CV curve of Co3O4-HNC@NiCo-LDH-1 enclosed the largest area. The CV curves of the samples with the optimal ratios, evaluated at various scanning speeds, are shown in Fig. 4(b). (The CV curves of the remaining samples are presented in Fig. S19.) As the sweep speed increased, the CV curves exhibited no significant distortion, demonstrating excellent rate performance. Additionally, the CV curves displayed pronounced redox peaks, which collectively confirmed the occurrence of redox reactions in all samples. These redox peaks corresponded to the electrochemical reactions of the electrode material in 6.0 M KOH electrolyte as shown in eqn (1)–(4):29,39,40
| Co(OH)2 + OH− ↔ CoOOH + H2O + e− | (1) |
| CoOOH + OH− ↔ CoO2 + H2O + e− | (2) |
| Ni(OH)2 + OH− ↔ NiOOH + H2O + e− | (3) |
| NiOOH + OH− ↔ NiO2 + H2O + e−. | (4) |
Fig. 4(c) shows the GCD curve of NiCo-LDH and Co3O4-HNC@NiCo-LDH at 2 A g−1. At the identical current density, Co3O4-HNC@NiCo-LDH-1 showed the most extended discharge time. Specifically, at 2 A g−1, Co3O4-HNC@NiCo-LDH-1 achieved a specific capacitance of 1862.4 F g−1, which was notably superior to that of Co3O4-HNC@NiCo-LDH-0.5 (1740.0 F g−1) and Co3O4-HNC@NiCo-LDH-2 (1789.6 F g−1). Fig. 4(d) presents the GCD curves of the optimally proportioned sample Co3O4-HNC@NiCo-LDH-1 at various current densities, while the GCD curves of the other samples are shown in Fig. S20. The GCD curves of all Co3O4-HNC@NiCo-LDH samples exhibited nearly symmetrical charging and discharging plateaus. This symmetry in the charge–discharge profiles suggested that the electrochemical reactions occurring at the electrode surface were highly reversible.41
The CV and GCD curves jointly confirmed that Co3O4-HNC@NiCo-LDH exhibited the highest specific capacitance, attributed to the in situ growth of an optimal density of NiCo-LDH on the robust Co3O4-HNC framework. The distinctive architecture facilitated cooperative redox coupling between Ni2+ and Co2+ centers. The synergistic impact greatly increased the overall electrochemical performance by optimizing the ion transport routes and improving the electrode's electrical conductivity and structural stability.42Fig. 4(e) presents the specific capacitance of all samples in this study across varying current densities. As evident from Fig. 4(e), the Co3O4-HNC@NiCo-LDH-1 composite demonstrated remarkable specific capacitance performance at both low and high current densities. These findings highlight the superior rate capability and electrochemical stability of the material, as reflected in its robust capacitance retention from low to high current densities.
To further elucidate the charge storage mechanism of the Co3O4-HNC@NiCo-LDH-1 electrode, eqn (5) and (6) were utilized to differentiate between capacitive and diffusion-limited charge storage mechanisms:43,44
| i = avb | (5) |
log i = b log v + log a. | (6) |
The specific steps for calculating the b value could be found in Scheme S2. As shown in Fig. 4(f), the b values of the Co3O4-HNC@NiCo-LDH-1 electrode, calculated from the anodic and cathodic peak current densities, were 0.58 and 0.56, respectively. The calculated b value was close to 0.5, indicating that the charge storage mechanism of the material was primarily governed by diffusion-controlled behavior during the charge and discharge processes.45 This finding was consistent with the results observed in the CV and GCD curves.
To further elucidate the proportion of capacitance control and diffusion control in the total capacitance of the sample, the following formula was employed for quantitative analysis (eqn (7)):46
| i(v) = k1 + k2v1/2. | (7) |
Here, k1 and k2 represent the contribution coefficients of capacitance control and diffusion control, respectively. Fig. S21(a) depicts the CV curve of the Co3O4-HNC@NiCo-LDH-1 electrode from 1 mV s−1 to 20 mV s−1. By fitting the CV curves at different scan rates, the values of k1 and k2 were determined. The calculated variation trend of the capacitance-controlled ratio of the Co3O4-HNC@NiCo-LDH-1 electrode, increasing from 18% at 1 mV s−1 to 49% at 20 mV s−1, suggests that as the sweep rate increased, the contribution of the capacitance-controlled mechanism to the total capacitance became more significant. However, the diffusion-controlled process remained the dominant factor in determining the overall capacitance. Fig. 4(h) presents the area proportion diagram of diffusion control in the CV curve at 5 mV s−1, while Fig. S21 illustrates the area proportion diagrams of diffusion control at other sweep rates. Concurrently, the area ratio diagrams of capacitive control in the CV curve at various sweep speeds are shown in Fig. S22.
The cycling durability of the Co3O4-HNC@NiCo-LDH-1 electrode was evaluated via prolonged galvanostatic cycling at 15 A g−1, as shown in Fig. 4(i). The comparison chart illustrating the stability tests of other electrodes is shown in Fig. S23. The Co3O4-HNC@NiCo-LDH-1 electrode exhibited an outstanding capacitance retention rate of 98.15% after 10
000 consecutive GCD cycles. Based on the GCD profiles of the initial and final three cycles, the electrochemical performance of the electrode remained virtually unchanged, despite prolonged operation at high current densities. This remarkable cycling stability demonstrates that the electrode material could maintain its efficient charge storage capacity even under high current density conditions, while also exhibiting excellent structural integrity. This superior performance could be attributed to the unique core–shell structure design of the material. Specifically, the etching of the core removed inactive regions, providing a robust internal framework, while the outward growth of the surface NiCo-LDH layer significantly increased the specific surface area and created more active sites for electrochemical reactions. EIS analysis provided insights into the electron/ion transport mechanisms within the hybrid materials. As illustrated in Fig. S24(a–d) and S25(a), the Rs of the ZIF-67 NCs exhibited a slight decrease compared to their pre-etching state. This observation suggested that the overall resistance of the electrode system was partially reduced. Fig. S26(a) presents the EIS diagrams of NiCo-LDH, and the individual samples of Co3O4-HNC@NiCo-LDH are provided in Fig. S27. Fig. 5(a) reveals the comparison diagram of Co3O4-HNC@NiCo-LDH composite materials. Among these, Co3O4-HNC@NiCo-LDH-1 exhibited the lowest internal resistance (Rs) value of 0.62 Ω, while the Rs values of the other samples were relatively higher. This similarity in Rs values indicated that the internal resistance of the entire system was not significantly different among the samples.47Rct represents the charge transfer resistance, which reflects the resistance encountered during the charge transfer process in the electrode reaction.48,49 The Co3O4-HNC@NiCo-LDH-1 composite exhibited the smallest Rct value of 0.32 Ω, followed by Co3O4-HNC@NiCo-LDH-2 (0.34 Ω) and Co3O4-HNC@NiCo-LDH-0.5 (0.85 Ω). The smallest Rct value of Co3O4-HNC@NiCo-LDH-1 indicated superior electrochemical activity compared to other samples. Additionally, the Nyquist plot of Co3O4-HNC@NiCo-LDH-1 showed the steepest linear slope in the low-frequency region, which suggested the least diffusion resistance at the electrode–electrolyte interface.50,51 This further highlighted the excellent electrochemical performance and efficient ion transport capability of Co3O4-HNC@NiCo-LDH-1.
![]() | ||
| Fig. 5 Co3O4-HNC@NiCo-LDH: (a) Nyquist plots; (b) Bode plots of phase angle versus frequency; the normalized real (c) and imaginary (d) part capacitance versus frequency. | ||
Furthermore, analysis of the Bode phase diagram (Fig. 5(b)) revealed that at a phase angle of −45°, the relaxation time (τ), determined from the reciprocal of frequency, could serve as an indicator of the electrode discharge rate.52 The relaxation times for ZIF-67 NCs, HZIF-0.9, HZIF-1.2, HZIF-1.5, and NiCo-LDH were found to be 10.00 s, 9.84 s, 8.24 s, 8.33 s, and 6.67 s, respectively (Fig. S25(b) and Fig. S26(b)). Fig. 5(b) demonstrates that Co3O4-HNC@NiCo-LDH-1 exhibits a value of 1.75 s, outperforming Co3O4-HNC@NiCo-LDH-0.5 (5.56 s) and Co3O4-HNC@NiCo-LDH-2 (2.63 s). This indicates that Co3O4-HNC@NiCo-LDH-1 achieved the most efficient ion transport and charge transfer kinetics during the electrochemical process.
To further analyze the electrochemical behavior, both the real and imaginary parts of capacitance were examined to reflect the complex impedance characteristics of the electrode system. Specifically, changes in the imaginary part of capacitance were indicative of diffusion-controlled processes within the electrode material, while variations in the real part of capacitance reflected the charge accumulation and release processes occurring at the electrode surface. After normalizing the values of the real and imaginary capacitance as functions of frequency, the performance of the electrode material could be assessed based on the position of the curve.53,54 Specifically, the closer the curve shifted to the right as the real capacitance changed with frequency, the better the performance of the material (Fig. 5(c) and Fig. S26(c)). This shift indicated a longer time constant for charge accumulation and release processes, suggesting enhanced electrochemical stability and efficiency. When the imaginary capacitance reached its maximum value, the reciprocal of the corresponding frequency reflected the key time scale of the electrode material during the charge and discharge process. A shorter time scale indicated that the material could maintain high energy efficiency even during rapid charge and discharge cycles. Fig. 5(d) and Fig. S26(d) show the calculated times for various samples: Co3O4-HNC@NiCo-LDH-0.5 (10.00 s), Co3O4-HNC@NiCo-LDH-1 (3.15 s), Co3O4-HNC@NiCo-LDH-2 (3.16 s), and NiCo-LDH (14.68 s). Fig. S28 presents the linear fitting plots of Z′ versus the inverse square root of angular frequency (ω−1/2) in the low-frequency region. The Warburg region was typically associated with ion diffusion resistance, showing well-defined linearity for all three nanomaterials, indicating that the electrochemical processes were governed by diffusive behavior. The corresponding Warburg slopes (K), extracted from the linear fits, exhibit distinct values. Notably, the Co3O4-HNC@NiMn-LDH-1 electrode exhibited the lowest Warburg slope, suggesting significantly reduced diffusion impedance and enhanced ion transport kinetics. In summary, the Co3O4-HNC@NiCo-LDH-1 electrode exhibited the most superior electrochemical performance, which could be attributed to the optimal etching ratio, the most suitable in situ growth ratio, and the synergistic effects between these two factors.
In situ Raman spectroscopy was conducted to elucidate the electrochemical reaction pathways of the Co3O4-HNC@NiCo-LDH-1 composite. Raman spectra, illustrated in Fig. 6(a), were captured throughout the electrode's oxidation within a potential range from 0 to 0.9 V. A detailed view in Fig. 6(b) highlights the characteristic Raman peak for M–O (M = Ni or Co) at 662 cm−1. As the voltage increased, this peak intensity heightened and shifted to higher wavenumbers, a blue shift indicating enhanced activity of the corresponding vibration mode (M–O).55 This observation was in accordance with the electrochemical reactions outlined in eqn (1)–(4), where Ni(OH)2 and Co(OH)2 were oxidized to their respective oxides. In contrast, as depicted in Fig. 6(c and d), the 662 cm−1 peak intensity diminished during reduction as the voltage decreased. The Raman spectra displayed heightened fluctuations and instability at elevated voltages, stabilizing as the voltage reduced.56 Throughout the redox cycle, the Raman curve preserved a nearly symmetrical configuration. Furthermore, following a single electrochemical cycle, the Raman spectra were substantially unaltered relative to its initial state, thereby validating the reversibility of the material's electrochemical reactions.
Fig. 7(a) presents the CV curves of the positive and negative electrodes, and the Co3O4-HNC@NiCo-LDH-1//AC ASC at 20 mV s−1. The potential window of AC was −1 to 0 V, while that of Co3O4-HNC@NiCo-LDH-1 was 0 to 0.6 V. Based on these ranges, it could be inferred that the potential window of the assembled ASC was approximately 0 to 1.60 V. Therefore, in the potential window exploration of the ASC, the maximum voltage was set at 1.5–1.65 V. During CV testing (Fig. 7(b)), it was observed that significant polarization and water decomposition reactions occurred at 1.65 V. Fig. 7(c) presents the CV curves of the device at various sweep speeds within the 0–1.6 V potential window. Notably, no significant polarization was observed even at low sweep speeds. Importantly, even at 100 mV s−1, the CV curves maintained their shape without obvious distortion, indicating that the ASC maintained exceptional rate capability and stability. Meanwhile, these curves illustrated that the specific capacitance of the ASC was determined by the synergistic contribution of the double-layer capacitance of activated carbon and the pseudocapacitance of the Co3O4-HNC@NiCo-LDH-1 electrode.
Furthermore, Fig. 7(d) reveals that the stable operating voltage under the GCD curve was 1.6 V. When the peak voltage reached 1.65 V, a charging plateau emerged in the GCD curve. Based on these combined results, the subsequent voltage test range for the Co3O4-HNC@NiCo-LDH-1//AC ASC device was determined to be 0–1.6 V. Fig. 7(e) presents the GCD curve of the Co3O4-HNC@NiCo-LDH-1//AC ASC device within a voltage range of 0 to 1.6 V (the comparison chart of GCD at high mass loading is shown in Fig. S29.) The nearly symmetrical GCD curves observed at various current densities demonstrated that the device exhibited remarkable reversibility and excellent charge–discharge efficiency. Upon calculation, the specific capacitance of the Co3O4-HNC@NiCo-LDH-1//AC ASC device was 98 F g−1 at 2 A g−1.
Fig. 7(f) displays the EIS diagram of the ASC. The device exhibited low Rs and Rct in the high-frequency region, along with steeply inclined lines in the low-frequency region. These characteristics underscored the system's superior performance. The outstanding performance of the ASC could be primarily attributed to the superior characteristics of both the positive and negative electrodes. In particular, the unique core–shell structure of the Co3O4-HNC@NiCo-LDH-1 electrode played a pivotal role in enhancing the overall electrochemical performance of the device. This innovative architecture not only streamlined the engagement of active materials but also significantly amplified the efficiency of ion migration and charge transfer. The synergistic interaction between the core and shell components ensured that the electrode maintains stability even at elevated scan rates.
Remarkably, as illustrated in Fig. 7(g), the energy density reached 76.8 Wh kg−1 at 2 A g−1 and 68.2 Wh kg−1 at 15 A g−1. The retention rate of 88.80% under these conditions highlighted the system's superior rate capability. Fig. 7(h) further proves the stability of the ASC. The data revealed that even after 15
000 cycles at 15 A g−1, the ASC maintained a retention rate of 98.38% with a coulombic efficiency of 98.94%. These nearly negligible attenuation figures underscored the exceptional stability of the ASC device. The inset SEM image depicts the morphology of the sample after cycling. It is evident that the structure remains largely intact, with negligible collapse, compared to its condition before cycling. The superior core–shell architecture endowed the electrode with remarkable stability. As demonstrated in Fig. 7(h), the assembled button-type supercapacitor successfully powered an LED lamp, showcasing its practical applicability. In Fig. 7(i) a pentagonal star diagram is used to compare the key performance metrics of the current work on supercapacitors, including specific capacitance, retention rate, rate performance, power density, and energy density, with those of similar studies.20,24,27,57–59 The results demonstrated that the current work outperforms most comparable efforts in these critical areas.
The floating test is a sophisticated method for real-time monitoring of the potential fluctuations of both positive and negative electrodes during cyclic charge–discharge testing.60 This technique enables real-time monitoring of electrode responses under varying electrochemical conditions.61,62Fig. 8(a and b) shows the real-time voltages of the positive and negative electrode materials, respectively, within the same system under different current densities. Fig. 8(c) reveals that as the current density increased, the voltage range of the positive material increased, while the voltage range of the negative material decreased. Additionally, the combined voltage of the positive and negative electrodes decreased with the rise in current density. As evidenced in Fig. S30, the sum of these values approached 1.6 V, indicating that the device could fully utilize the entire potential window to deliver its electrochemical performance, regardless of changes in current density. This further implied that the system exhibited negligible internal resistance. Fig. 8(d) illustrates the real-time voltage range of the Co3O4-HNC@NiCo-LDH-1 electrode over 20 cycles at 10 A g−1, while also presenting the energy density profile of the Co3O4-HNC@NiCo-LDH-1//AC ASCs. The system demonstrated excellent stability, with minimal voltage fluctuations across the cycles.
000 cycles. This work underscores the significant potential of the Co3O4-HNC@NiCo-LDH-1 composite for high-performance energy storage applications, highlighting its balanced combination of high specific capacitance, excellent rate capability, and superior cycling stability. The “multi-scale hollow core–shell” structure finds broad application across diverse battery systems, including lithium-ion batteries, sodium-ion batteries, and lithium–sulfur batteries. By augmenting the specific capacity, cycling stability, and rate performance of materials, it offers an effective solution to key challenges in energy storage.
| This journal is © the Partner Organisations 2026 |