Download PDF
Article  |  Open Access  |  9 Dec 2025

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Views: 4040 |  Downloads: 1 |  Cited:  0
Energy Mater. 2025, 5, 500155.
10.20517/energymater.2025.88 |  © The Author(s) 2025.
Author Information
Article Notes
Cite This Article

Abstract

The long-term stability of lithium-ion batteries is a critical factor limiting their broader adoption in multifunctional and structural energy storage systems. However, conventional metallic current collectors tend to be heavier and less mechanically adaptable than fiber-based materials such as quartz woven fabrics (QWFs), particularly when structural integration is required. Quartz fabrics, composed primarily of silica, offer high thermal stability, mechanical robustness, and low areal weight, making them attractive candidates for multifunctional electrode platforms. In this study, carbon nanotubes (CNTs) were directly grown on QWFs via chemical vapor deposition, using Ni nanoparticles as catalysts and C2H4 as the carbon source. The growth process was optimized by varying temperature over a 2-h duration to form uniform, conductive CNT networks. The resulting CNT-coated QWFs functioned dually as current collectors and active electrode supports, delivering an initial discharge capacity of 201.54 mAh g-1 at a 0.1 C-rate. The electrodes retained 89.8% of their initial capacity after 50 cycles at a 0.5 C-rate, demonstrating excellent rate capability and cycling stability, with performance nearly equivalent to that of conventional Al foil-based electrodes. Although quartz fabrics are inherently insulating, the CNT coating formed an integrated conductive network, enabling efficient charge transport while maintaining the structural integrity required for load-bearing applications. This work presents a novel fiber-based electrode design for structural lithium-ion batteries, offering a promising route toward lightweight, mechanically integrated energy storage systems suitable for advanced electronics, electric vehicles, and aerospace technologies.

Keywords

Structural lithium-ion batteries, carbon nanotube, quartz woven fabric, chemical vapor deposition

INTRODUCTION

Structural batteries, which combine mechanical load-bearing capabilities with energy storage functions, represent a transformative approach to lightweight system design, particularly in applications such as electric vehicles (EVs), unmanned aerial vehicles (UAVs), and aerospace systems[1-3]. In conventional energy storage architectures, lithium-ion batteries (LIBs) are merely passive components, contributing a significant portion of the total system weight without bearing any structural loads. As the demand for higher energy densities, safety, and material efficiency grows, the development of multifunctional materials that can simultaneously store energy and reinforce structural integrity is becoming increasingly crucial[4,5]. A key challenge in the realization of structural batteries is the identification of electrode materials that exhibit both excellent electrochemical properties and mechanical robustness. Fiber-reinforced composites have emerged as promising candidates due to their tunable mechanical properties, formability, and ability to host active materials. Among the various fibers explored, carbon fibers offer high electrical conductivity but relatively lower chemical and thermal stability, whereas silica-based fibers, such as quartz woven fabrics (QWFs), provide superior thermal resistance, chemical inertness, and dimensional stability[5-8]. These properties make QWFs especially attractive for multifunctional energy storage in extreme or thermally demanding environments.

Despite their mechanical and thermal advantages, silica-based fibers exhibit inherently poor electrical conductivity, limiting their role as current collectors or active electrode components in LIBs[9]. To address this issue, several strategies have been explored, including the deposition of metal nanoparticles (e.g., Ni, Cu) or conductive carbon layers onto the fiber surfaces[10]. These modifications aim to bridge the gap between mechanical integrity and electrochemical performance, enabling the development of structurally integrated electrodes. Among conductive additives, carbon nanotubes (CNTs) have been widely investigated due to their outstanding electrical conductivity, high aspect ratio, mechanical resilience, and ability to form percolated electron transport networks[11]. CNTs not only enhance charge transport across the electrode but also improve the interfacial contact between active materials and current collectors. In addition, surface modification of cathode materials such as NCM811 via carbon coating has been actively explored to enhance their structural and interfacial stability[12-15]. Moreover, the one-dimensional nanostructure of CNTs enables mechanical interlocking and reinforcement within the electrode matrix, which is particularly valuable in structural battery designs where mechanical integrity must be maintained during battery operation. Recent advancements have demonstrated the feasibility of growing CNTs directly on fiber surfaces using chemical vapor deposition (CVD) techniques, which enable uniform, vertically aligned CNT arrays anchored via metal catalysts such as nickel[16-19]. This method ensures strong bonding between the CNTs and the underlying fabric, eliminating the need for additional binders or conductive agents. It also allows for precise control over CNT growth density and morphology by tuning reaction parameters such as temperature, gas composition, and catalyst distribution[20-23].

In this study, we propose and demonstrate a multifunctional electrode architecture based on CNT-integrated QWFs for application in structural LIBs. CNTs were directly synthesized on SiO2 fabrics via a CVD process using nickel as a catalyst layer[24-27]. This integration significantly improves the electrical conductivity and active surface area of the electrode, enabling enhanced electrochemical performance. Specifically, the CNT layer facilitates efficient charge transport and provides a robust scaffold for active material deposition, allowing for higher areal loading and improved energy density. Compared to conventional aluminum foil-based current collectors, the CNT-QWF electrode demonstrates comparable or superior electrochemical performance while offering added mechanical functionality. Furthermore, we explore the effect of CVD process conditions on CNT morphology and electrode performance. Through systematic optimization, we show that CNTs grown at optimal conditions result in enhanced capacity retention, rate capability, and cycling stability. Additionally, the unique morphology of the CNT-coated fabric allows for reduced usage of traditional conductive agents, potentially lowering production costs and increasing the mass fraction of active materials. This work presents a scalable and effective approach for fabricating structurally integrated electrodes that meet the dual requirements of mechanical strength and electrochemical performance. The C-QWF-700-based electrode presents a viable alternative to conventional metal foils, combining mechanical robustness with sufficient electrochemical performance for structural battery applications. The findings offer insights into the design of next-generation structural batteries and open new avenues for integrating energy storage functions into load-bearing materials for lightweight, high-efficiency systems.

EXPERIMENTAL

Carbon nanotube growth mechanism

The CNT growth mechanism of SiOx follows the vapor-liquid-solid growth mechanism. To grow CNTs on a QWF, Ni metal catalyst particles with a thickness of 10 nm were deposited on the QWF using a high-vacuum sputter and carbon evaporation coater (ACE600, Leica Microsystems). Then, C2H4 gas as a carbon precursor was circulated with N2 gas and heated to two different temperatures: 600 and 700 °C for CNT growth optimization (flow rate C2H4:N2 = 25:125). The heating rate was 5 °C min-1. The CVD process was performed for 2 h. Subsequently, the heater was switched off and the sample was cooled to room temperature. These CNT-coated QWF samples were named C-QWF-600 and C-QWF-700, respectively.

Material characterizations

X-ray diffraction (XRD) patterns were determined using the X-ray diffractometer (MiniFlex 600, BD71000481-01) in the wavelength range of 1.5405 Å, angle range of 10° ≤ 2θ ≤ 70°, and step width of 0.01°. The microstructure and morphology of the sample were characterized by scanning electron microscopy (SEM; EM 30, GSEM) and transmission electron microscopy (TEM; TF30ST, FEI, NEOARM, JEM-ARM200F, JEOL). Electrochemical characterization was performed using Raman spectrometry (DXR3 Raman Microscope, Thermo Fisher Scientific) with a 532 nm laser in the wavenumber range of 0-4,300 cm-1. X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific, NEXSA) analysis was performed to investigate the bonding state of the C atoms and the surface state of Ni on the CNT surface. The measurements were performed without etching, and the C 1s peak was calibrated to 284.6 eV. Thermogravimetric analysis (TGA) was performed using a Simultaneous Thermal Analyzer (SDT 650, TA Instruments) to investigate the thermal degradation characteristics of the samples. Measurements were performed in air at a constant heating rate of 5 °C min-1 from 25 °C to 900 °C. The tensile test was performed utilizing a 10 kN general-purpose testing machine (LR-10K, Lloyd Instruments). The specimens were rectangular with dimensions of 1 cm in width and 4 cm in length. The tests were performed at a strain rate of 10 mm min-1.

Coin cell fabrication

Al foil electrode supports were used to prepare the electrode supports for C-QWF-600 and C-QWF-700 and were also used as a comparison group for CNT growth. The electrode slurry was mixed with cathode active materials (NCM811), conductive material (super P), and polymer binder (polyvinylidene fluoride) in the weight ratio of 80:10:10 using a mixer; then, slurry casting was performed using a doctor blade at a thickness of 250 μm on C-QWF-600 and C-QWF-700 and a thickness of 200 μm on Al foil. Then, the electrode was dried in a vacuum oven at 50 and 80 oC overnight. The electrodes were punched into 12 mm circular disks. For evaluating the electrochemical performance, a lithium coin cell was fabricated as a CR2032 coin cell in an Ar-filled glove box. For the half-cell tests, lithium metal was used as the counter electrode, and the fabricated C-QWF-600, C-QWF-700, and Al foil electrodes were used as the working electrodes. Glass microfiber filter (Grade GF/D, Whatman) was used as the separator and 1.15 M LiPF6 in ethylene carbonate/diethyl carbonate (1:1 vol%) was used as the electrolyte. The lithium metal was punched into circular disks with diameters of 14 mm.

Electrochemical test

Electrochemical tests were performed with 2032 coin-type half-cell with Li metal as the counter. The cell was assembled in a glove box filled with argon gas containing 0.1 ppm less than H2O and O2. The electrolyte was prepared by dissolving 1.15 M LiPF6 in ethylene carbonate/diethyl carbonate (1:1 vol%). The glass fiber membranes (GF/D, Whatman) were used for the separators with a thickness of 1.55 mm. The cyclic voltammetry (CV) was performed using WonATech (WBCS3000 L) in a cut-off voltage range of 3.0 V to 4.3 V (vs. Li/Li+) at a scan rate of 0.1 mV s-1. The galvanostatic measurements were carried out in a cut-off voltage range from 3.0 V to 4.3 V (vs. Li/Li+) using an automatic battery cycler (MIHW-200-160CH-B, Neware). Electrochemical impedance spectroscopy (EIS) was performed in the frequency range of 100 mHz to 1MHz with a signal peak to peak amplitude of 5 Mv at the open circuit potential of the cell. For impedance spectroscopy, the ESR (Equivalent Series Resistance) was measured as the real part of the impedance (Zreal) at a frequency of 1 kHz using a multi-channel battery tester (BioLogic VMP3, France).

RESULTS AND DISCUSSION

SiO2 fabrics for structural batteries were prepared as electrode supports through Ni sputtering followed by CVD. Since SiO2 fabric has inherently low electronic conductivity, Ni nanoparticles were first deposited onto the surface of QWF to enhance its conductivity. Subsequently, the CVD process was conducted at 600 and 700 °C for 2 h to determine the optimal temperature for CNT growth. The morphology and CNT growth were confirmed using SEM imaging. A simplified schematic of the experimental process is presented in Figure 1. Figure 1A-C illustrates the SiO2 fabric (QWF) along with SEM images at low and high magnifications. Figure 1C confirms that the silica fiber structure of pristine QWF (P-QWF) remains intact, which correlates with its highest tensile strength [Supplementary Figure 1]. After Ni sputtering, the Ni-coated QWF (N-QWF) exhibited a more silvery-white luster on its surface, as shown in Figure 1D-F. Additionally, Figure 1F reveals that the N-QWF electrode has a finer surface roughness compared to the P-QWF electrode, indicating a well-distributed Ni catalyst layer. The silica fiber structure of N-QWF also remained intact, with its tensile strength similar to that of P-QWF [Supplementary Figure 1]. Following the thermal CVD process, the surface of N-QWF darkened, indicating uniform carbon deposition across the electrode, forming the CNT-coated QWF (C-QWF), as shown in Figure 1G-I. Figure 1I shows that CNTs growth is more uniform and abundant on the surface of C-QWF-700 compared to C-QWF-600 [Supplementary Figure 2], which can be attributed to the higher thermal decomposition efficiency of C2H4 gas at elevated temperature[28,29]. However, C-QWF-700 exhibited a decrease in tensile strength compared to P-QWF and N-QWF, likely due to the reduction of SiOx under the N2 atmosphere at 700 °C. Nevertheless, C-QWF-700 retained a significantly higher tensile strength (360 ± 72 MPa) than that of Al foil (181 ± 21 MPa), confirming the structural robustness of the electrode support.

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Figure 1. Schematic illustration of (A) pristine quartz woven fabric (P-QWF) electrode, (D) Ni-sputtered quartz woven fabric (N-QWF) electrode, and (G) quartz woven fabric (QWF) electrode. Low- and high-magnification optical images of the top view of (B) P-QWF, (E) N-QWF, and (H) C-QWF-700. Scanning electron microscopy (SEM) images of the top view of the (C) P-QWF, (F) N-QWF, and (I) C-QWF-700.

Figure 2 shows the SEM image of the electrode surface from the top view, including Al foil Figure 2A-C, C-QWF-600 Figure 2D-F, and C-QWF-700 Figure 2G-I. Given the inter-woven fabric structure of QWF, the loading mass was initially expected to be significantly lower than that of the Al foil. However, A comparative analysis of the surface morphology between electrodes using Al foil as the current collector and those utilizing QWF reveals that the loading was uniformly distributed with no substantial voids, even when compared to Al foil. Supplementary Figure 3 shows that the measured loading masses for Al foil, C-QWF-600, and C-QWF-700 were approximately 5.46, 3.77, and 4.03 mg, respectively, indicating that the loading mass of QWF was slightly lower than that of Al foil. The standard deviations are 0.57, 0.53, and 1.18 mg, respectively. The QWF as a current collector offers the advantage of maintaining a lighter weight and enhanced mechanical flexibility compared to Al foil while ensuring adequate active material loading, highlighting its potential as a structural electrode support. Additionally, the bending resistance of the electrode was qualitatively assessed by manually bending it, and the corresponding observations were recorded in Supplementary Figure 4. After growing CNTs on QWF electrodes and coating the electrodes, we found that they still have a reasonable bending ability, which shows promise for future applications in structural batteries.

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Figure 2. SEM image of electrodes (A-C) Al foil, (D-F) C-QWF-600, and (G-I) C-QWF-700. SEM: Scanning electron microscopy; QWF: quartz woven fabric.

Figure 3 shows the carbon properties of the grown CNTs and changes in the crystal structure of the support corresponding to the changes in heat treatment time during the CVD process. Figure 3A shows the XRD patterns of the QWF electrode supports (P-QWF, N-QWF, C-QWF-600, and C-QWF-700). The XRD results indicate that the broad peak in the range of approximately 2θ = 15°-30° in P-QWF is related to amorphous silica, which is maintained without significant change even after the Ni deposition process. The intensity of the amorphous silica peak decreases in the C-QWF-600 and C-QWF-700 electrode supports owing to the growth of CNTs on the Ni catalyst, with a greater reduction in the C-QWF-700 electrode support than in the C-QWF-600 support. Moreover, the peak at 2θ = 27.64° corresponds to the (002) plane of the graphite lattice. This confirmed the graphitization at 700 °C. This result is consistent with the SEM images. Figure 3B shows the Raman spectra of the QWF electrode supports (P-QWF, N-QWF, C-QWF-600, and C-QWF-700). The D band is related to intracrystalline defects and is located at 1,350 cm-1, while the G band is caused by the sp2 bonding of carbon and is located at 1,600 cm-1. The Raman spectra of the C-QWF-600 and C-QWF-700 electrode supports revealed the intensity ratio of the D and G bands (ID/IG), which indicates the disorder in the carbon structure and the degree of graphite crystallization. ID/IG of C-QWF-600 is 0.865, whereas ID/IG of C-QWF-700 is reduced to 0.797. The decrease in the ID/IG ratio indicates that the CNTs grown through the 700 °C thermal process have relatively few intracrystalline defects and stable graphite crystals. Figure 3C shows the TGA curves of CNT quantity of C-QWF-600, C-QWF-700. Figure 3D-F presents the XPS spectra results used to analyze the surface composition of P-QWF, N-QWF, and C-QWF-700. The analysis was performed without an etching process. Figure 3D-F displays the C 1s spectra obtained from the CNT surface. In the C 1s spectrum of C-QWF-700, multiple subpeaks were identified at 284.5 eV (C-C/C-H), 285.3 eV (C-O), and 289.5 eV (C=O). During the CVD process, where ethylene (C2H4) served as the carbon source for CNT growth, the thermal decomposition of ethylene likely generated hydrogen atoms, which bonded with carbon atoms on the CNT surface, forming C-H bonds. Supplementary Figure 5A-C illustrates the Ni 2p spectra. In Supplementary Figure 5A-C, distinct Ni peaks were observed in both N-QWF and C-QWF-700. The peaks at 851.8 eV and 869.5 eV in the N-QWF sample indicate the presence of metallic Ni (Ni0) on the QWF surface. Additionally, in both N-QWF and C-QWF-700, peaks at approximately 852 and 854 eV correspond to Ni 2p1/2 and Ni 2p3/2, respectively, accompanied by satellite peaks at 870 and 872 eV[30]. These peaks suggest the presence of Ni in oxide and hydroxide forms[31].

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Figure 3. (A) X-ray diffraction patterns of the QWF electrodes support (10°-70°), (B) Raman spectroscopy of QWF electrodes support, (C) TGA curves of CNT quantity of C-QWF-600, C-QWF-700; C 1s XPS spectra of (D) P-QWF, (E) N-QWF, and (F) C-QWF-700. QWF: Quartz woven fabric; TGA: thermogravimetric analysis; CNT: carbon nanotube; XPS: X-ray photoelectron spectroscopy; P-QWF: pristine QWF; N-QWF: Ni-coated QWF; C-QWF: CNT-coated QWF.

Figure 4 presents the results of the TEM analysis, conducted to examine the surface morphology of the particles and the distribution of Ni and CNTs. As shown in Supplementary Figure 6, TEM images were utilized to evaluate the uniformity and consistency of the deposited Ni layer. The results confirmed that the Ni layer closely adhered to the intended thickness of 10 nm, ensuring the reliability of the deposition process. Figure 4A-C and Figure 4E-G illustrates the distribution of CNTs on the C-QWF-600 and C-QWF-700 electrode supports, respectively. The crystal structure of Ni is identified in Figure 4A and E. The top-right inset image shows the selective area electron diffraction (SAED) pattern of Ni nanoparticles. In this pattern, a characteristic diffraction ring corresponding to the (111) plane is observed. Figure 4D shows the Energy dispersive X-ray spectroscopy (EDS) mapping results for C-QWF-600, confirming the Ni catalyst-mediated growth and distribution of CNTs on the 366 μm thick SiO2 fabric. Figure 4H shows the EDS mapping results for C-QWF-700. It shows that the Ni catalyst-mediated growth of CNTs on the SiO2 fabric was 1.689 μm thick and more abundant compared to that on C-QWF-600. The results also indicate that Ni and C are uniformly distributed above the QWF, which is consistent with the XRD and SEM analyses.

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Figure 4. (A and B) Transmission electron microscopy (TEM) images of C-QWF-600 electrode support, (C and D) Energy dispersive X-ray spectroscopy (EDS) mapping results of C-QWF-600 showing distributions of Si, Ni, O and C, (E and F) TEM images of C-QWF-700 electrode, (G and H) EDS mapping results of C-QWF-700 showing the distributions of Si, Ni, O and C. C-QWF: CNT-coated QWF; QWF: quartz woven fabric; CNT: carbon nanotube.

In order to investigate the effect of Ni as a catalyst on the electrolyte, CV tests were conducted within the voltage range of 3.0-4.3 V (vs. Li/Li+). Supplementary Figure 7 presents a comparative analysis of N-QWF and P-QWF, demonstrating that no significant additional redox peaks or side reactions were observed. This indicates that Ni does not induce noticeable side reactions with the electrolyte under the operating voltage conditions, confirming the absence of irreversible reactions between the Ni catalyst and the electrolyte. Figure 5 presents the electrochemical characteristics of the QWF electrode supports during charge and discharge, based on cycle performance tests conducted at various charge and discharge rates. These tests were performed using NCM811 as the cathode material (theoretical capacity: 275 mAh g-1) to optimize the structural cell for high energy density. Figure 5A displays the discharge capacity of NCM811 applied to Al foil and QWF electrode supports (C-QWF-600 and C-QWF-700) over 50 cycles at a 0.1 C-rate. Additionally, to analyze the charge-discharge mechanisms of the electrodes, CV analysis was conducted to compare the electrochemical behavior of Al foil, C-QWF-600, and C-QWF-700, as shown in Supplementary Figure 8. It shows that two distinct redox couples are observed within a specific voltage range for all three electrodes. The oxidation/reduction peaks for Al foil, C-QWF-600, and C-QWF-700 were observed at approximately 3.76V/3.66V, 3.74V/3.70V, and 3.76V/3.68V, respectively, corresponding to the Ni2+/Ni4+ redox process. Additionally, oxidation/reduction peaks were observed at approximately 4.23V/4.13V, 4.22V/4.14V, and 4.24V/4.10V, respectively, which are attributed to the Co3+/Co4+ redox process. The Al foil and QWF electrode supports exhibited superior discharge capacity up to 200 mAh g-1, and the C-QWF-700 electrode support retained 89.8% of the initial discharge capacity after 50 cycles. The charge and discharge tests at 0.5 and 2 C-rates illustrated in Figure 5B and C show that, in the initial cycle, C-QWF-700 shows discharge capacities of 194.6 and 175.8 mAh g-1 at 0.5 and 2 C, respectively. This is higher than those of the cell with Al foil electrode support, which exhibit discharge capacities of 183.4 and 159.3 mAh g-1 in the initial cycle, respectively. After 500 cycles, C-QWF-700 exhibits a discharge capacity similar to that of a cell with an Al foil collector. Moreover, the C-QWF-700 electrode support showed capacity retention rates of 51.0% and 65.3% of the initial discharge at 0.5 and 2 C-rates, respectively, in a 500-cycle life test, indicating stable performance in continuous cycling. In general, carbon-based electrode collectors have lower conductivity than Al foil and show a rapid decrease in high-rate performance. However, the high CNT content and stable cathode material composite of the C-QWF-700 electrode support resulted in stable electrochemical performance in high-charge-and-discharge tests. The electrochemical reaction zones of the QWF electrode supports are shown in Figure 5D-F, which show the formation of a stable solid electrolyte interphase layer on the cathode surface during the initial cycle. Furthermore, the discharge curves for the second, fifth, and tenth consecutive cycles overlap, indicating reversible charge and discharge processes throughout the cycling. As shown in Figure 5E and F, at 0.5 C, the initial discharge capacities of the Al foil and QWF electrode supports (C-QWF-600 and C-QWF-700) are 178.8, 179.3, and 194.6 mAh g-1, respectively. After 10 cycles, these values decreased to 176.1, 176.9, and 191.8 mAh g-1, respectively. The capacity retention rates of the Al foil and QWF electrode support were similar; however, C-QWF-700 exhibited the highest discharge capacity. Supplementary Table 1 compares the competitive electrochemical performance of CNT-coated QWF-based electrodes with previously reported Ni, Co and Mn-based (NCM-based) cathodes using non-metallic or composite current collectors. As shown in Supplementary Table 1, the C-QWF-700 electrode exhibited superior rate performance and cycle stability compared to similar systems. Notably, it maintained a high capacity of 175.9 mAh g-1 at 2 C and achieved a capacity retention rate of 82.6% after 100 cycles, superior to many reported NCM811-based electrodes using carbon fibers, carbon-polymer composites, or aluminum collectors.

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Figure 5. Cycle performance of QWF electrodes at (A) 0.1 C-rate (1 C = 275 mAh g-1), (B) 0.5 C-rate, and (C) 2 C-rate; Galvanostatic discharge/charge voltage profiles of 3.0 and 4.3 V at 0.5 C-rate for (D) Aluminum foil electrode support, (E) C-QWF-600 electrode support, and (F) C-QWF-700 electrode support. QWF: Quartz woven fabric; C-QWF: CNT-coated QWF; CNT: carbon nanotube.

Figure 6A shows the charge/discharge rate performance of the QWF electrode supports with increasing charge/discharge rates of 0.1, 0.2, 0.5, 1, 2, and 5 C. The rate performance of the QWF electrode supports (C-QWF-600 and C-QWF-700) was also evaluated. The discharge capacities of C-QWF-700 at the aforementioned charge/discharge rates are 205.3, 199.9, 187.6, 173.6, 155.5, and 122.4 mAh g-1, respectively; moreover, C-QWF-700 exhibits higher discharge capacity and rate capability at 0.5, 1, 2, and 5 C. Additionally, Supplementary Figure 9 presents the cycling performance of cells at 5 C, 10 C, and 20 C for Al foil, C-QWF-600, and C-QWF-700 electrodes. Supplementary Figure 9A illustrates the cyclability test conducted at a 5 C rate over 100 cycles, demonstrating stable cycling performance with discharge capacities of 84.91, 141.40, and 166.85 mAh g-1 for Al foil, C-QWF-600, and C-QWF-700, respectively. Furthermore, as shown in Supplementary Figure 9B and C, only C-QWF-700 sustained 100 cycles at 10 C and 200 cycles at 20 C. Based on the first cycle, the discharge capacities at 10 and 20 C were measured as 117.87 and 57.94 mAh g-1, respectively. After 100 cycles, the discharge capacities were retained at 77.19, 133.67, and 154.26 mAh g-1. Furthermore, when the current rate was changed to 0.5 C, the discharge capacity was shown at 181.2 mAh g-1, corresponding to a performance restoration rate of 96.5% as compared to that of the initial discharge capacity at 0.5 C. These rate capabilities are closely related to the charge-transfer processes and resistance at the electrode interface which is confirmed by EIS analysis. The Nyquist plots are shown in Figure 6B. Each electrode shows one semicircle, which indicates the charge-transfer resistance (Rct) at the electrode interface. The Rct of the Al foil and QWF electrode supports (C-QWF-600, C-QWF-00) are 94, 103, and 75 cm-2, respectively. C-QWF-700 showed the lowest internal resistance, which is consistent with the results of the previous rate capability test, demonstrating that C-QWF-700 has superior electrochemical properties owing to the high graphite crystallization of the CNTs. As shown in Figure 6C, the first discharge capacities at 0.1 C for Al foil and QWF-based electrode supports (C-QWF-600 and C-QWF-700) were measured as 210.58 mAh g-1 for Al foil, 194.8 mAh g-1 for C-QWF-600, and 205.1 mAh g-1 for C-QWF-700, indicating that Al foil exhibited the highest discharge capacity. In contrast, Figure 6D shows that the discharge capacity at 5 C was 47.2 mAh g-1 for aluminum foil, 105.4 mAh g-1 for C-QWF-600, and 122.1 mAh g-1 for C-QWF-700, indicating that C-QWF-700 exhibited the highest discharge capacity. These results demonstrate the distinct advantage of C-QWF-700 in delivering higher discharge capacity at elevated C-rates compared to Al foil. Supplementary Figure 10 shows that the ex-situ SEM analysis was conducted to evaluate the stability of CNTs and the electrode structure after 10 cycles at a scan rate 0.1 mV s-1. The results revealed no significant morphological changes in the electrode compared to its pre-cycling state. This indicates that the CNT network and electrode structure remained stable throughout the cycling process, demonstrating their structural integrity under electrochemical operation.

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Figure 6. (A) Rate performance of Al foil and QWF electrode supports at 0.1 to 5 C-rate, (B) electrochemical impedance spectra for Al foil and QWF electrode supports. First cycle of galvanostatic discharge/charge voltage profile between 3.0 and 4.3 V at (C) 0.1 C-rate, (D) 5 C-rate. QWF: Quartz woven fabric.

Considering the electrochemical performance, a reaction schematic of the superior electrochemical performance of the C-QWF-700 electrode support in comparison with that of the P-QWF electrode support is shown in Figure 7. The P-QWF is a silica substrate electrode that results in slow electron transfer owing to the low electrical conductivity of silicon [Figure 7A]. In contrast, C-QWF-700 is a silica substrate electrode with a carbon coating on the silica surface and simultaneous CNT growth, which forms a carbon network that facilitates electron transfer [Figure 7B]. This CNT network enables high electron transfer while efficiently fixing the electrode material, which is believed to be favorable for high-rate charging and discharging[32-35]. In conventional structural battery electrodes, CNTs, which exhibit high mechanical strength but can only play a relatively minor role as energy assistants, will expand the possibilities in terms of energy density through the application of multifunctional electrode supports for structural batteries.

Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

Figure 7. Schematic illustration of the cathode structure considering the electron movement pathways: (A) Conventional slurry casting P-QWF cathode substrate and (B) C-QWF-700 cathode substrate. QWF: Quartz woven fabric; P-QWF: pristine QWF; CNT: carbon nanotube; C-QWF: CNT-coated QWF.

CONCLUSION

In this study, to analyze the electrochemical performance of CNT growth for robust battery electrodes, optimized QWF electrode supports for LIBs were prepared by sequential sputtering and CVD processes, and C-QWF electrode supports with uniform and abundant CNT growth on the QWF were successfully fabricated. Compared to C-QWF-600, the optimized C-QWF-700 showed a discharge capacity of 201.54 mAh g-1 at 0.1 C-rate. After 500 cycles at 2 C-rate, the Al foil electrode support retained 75.0% of its initial discharge capacity, whereas the C-QWF-700 electrode support retained 65.3% of its initial discharge capacity; however, both exhibited almost equivalent discharge capacities of 119.5 and 114.9 mAh g-1, respectively. This suggests that the C-QWF-700 electrode support can be used to obtain a robust LIB. In addition to its superior electrical properties, the C-QWF-700 electrode support exhibited low internal resistance despite the inherent limitation of slow electrical conductivity owing to its silica-based composition. The advantages of this electrode support, including the ease of mass production and extended cycle life, demonstrate its potential for advancing the development of QWF electrode materials in next-generation LIBs.

DECLARATIONS

Authors’ contributions

Conceived the research and supervised the work: Park, M. Y.; Kim, J. H.

Wrote the manuscript: Hwang, H.

Homogenized the final version and played a major role in the corrections: Choi, H. J.; Baek, D.; Lee, D.; Kwon, D. J.; Jeong, S. H.

All authors have approved the final version of the manuscript.

Availability of data and materials

Some results of supporting the study are presented in the Supplementary Materials. Other raw data that support the findings of this study are available from the corresponding author upon reasonable request.

Financial support and sponsorship

This work was supported by the Glocal University 30 Project Fund of Gyeongsang National University in 2024. This work was supported by the Glocal University 30 Project Fund of Gyeongsang National University in 2025.

Conflicts of interest

We confirm that the author “Park, M. Y.” is affiliated with SASUNG Power Co. and that her contribution to this work does not involve any conflicts of interest. We have verified that her participation complies with the company’s internal policies and does not present any issues regarding affiliation or potential conflict. While other authors declare no competing interests.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2025

Supplementary Materials

REFERENCES

1. Zeng, X.; Li, M.; Abd El-hady, D.; et al. Commercialization of lithium battery technologies for electric vehicles. Adv. Energy. Mater. 2019, 9, 1900161.

2. Larcher, D.; Tarascon, J. M. Towards greener and more sustainable batteries for electrical energy storage. Nat. Chem. 2015, 7, 19-29.

3. Li, M.; Lu, J.; Chen, Z.; Amine, K. 30 years of lithium-ion batteries. Adv. Mater. 2018, 30, 1800561.

4. Galos, J.; Pattarakunnan, K.; Best, A. S.; Kyratzis, I. L.; Wang, C.; Mouritz, A. P. Energy storage structural composites with integrated lithium-ion batteries: a review. Adv. Mater. Technol. 2021, 6, 2001059.

5. Saeed, G.; Kang, T.; Byun, J. S.; et al. Two-dimensional (2D) materials for 3D printed micro-supercapacitors and micro-batteries. Energy. Mater. 2024, 4, 400017.

6. Li, W.; Liu, X.; Xie, Q.; You, Y.; Chi, M.; Manthiram, A. Long-term cyclability of NCM-811 at high voltages in lithium-ion batteries: an in-depth diagnostic study. Chem. Mater. 2020, 32, 7796-804.

7. Andre, D.; Kim, S.; Lamp, P.; et al. Future generations of cathode materials: an automotive industry perspective. J. Mater. Chem. A. 2015, 3, 6709-32.

8. Xu, X.; Liu, J.; Liu, Z.; et al. FeP@C nanotube arrays grown on carbon fabric as a low potential and freestanding anode for high-performance Li-ion batteries. Small 2018, 14, 1800793.

9. Liu, S.; Xiong, L.; He, C. Long cycle life lithium ion battery with lithium nickel cobalt manganese oxide (NCM) cathode. J. Power. Sources. 2014, 261, 285-91.

10. Asp, L. E.; Johansson, M.; Lindbergh, G.; Xu, J.; Zenkert, D. Structural battery composites: a review. Funct. Compos. Struct. 2019, 1, 042001.

11. Park, M. Y.; Kim, J.; Kim, D. K.; Kim, C. G. Perspective on carbon fiber woven fabric electrodes for structural batteries. Fibers. Polym. 2018, 19, 599-606.

12. Wang, M.; Zhang, X.; Guo, Z.; et al. Two-step carbon coating onto nickel-rich LiNi0.8Co0.1Mn0.1O2 cathode reduces adverse phase transition and enhances electrochemical performance. Electrochim. Acta. 2023, 454, 142339.

13. She, S.; Zhou, Y.; Hong, Z.; Huang, Y.; Wu, Y. Surface coating of NCM-811 cathode materials with g-C3N4 for enhanced electrochemical performance. ACS. Omega. 2022, 7, 24851-7.

14. Goh, M. S.; Moon, H.; Shin, H.; et al. Unlocking high-efficiency lithium-ion batteries: sucrose-derived carbon coating on nickel-rich single crystal Li[Ni0.8Co0.1Mn0.1]O2 cathodes. Surf. Interfaces. 2024, 51, 104721.

15. Zhuang, Y.; Shen, W.; Yan, J.; et al. Regulating the heat generation power of a LiNi0.8Co0.1Mn0.1O2 cathode by coating with reduced graphene oxide. ACS. Appl. Energy. Mater. 2022, 5, 4622-30.

16. Gakis, G. P.; Termine, S.; Trompeta, A. A.; Aviziotis, I. G.; Charitidis, C. A. Unraveling the mechanisms of carbon nanotube growth by chemical vapor deposition. Chem. Eng. J. 2022, 445, 136807.

17. Welna, D. T.; Qu, L.; Taylor, B. E.; Dai, L.; Durstock, M. F. Vertically aligned carbon nanotube electrodes for lithium-ion batteries. J. Power. Sources. 2011, 196, 1455-60.

18. Bitew, Z.; Tesemma, M.; Beyene, Y.; Amare, M. Nano-structured silicon and silicon based composites as anode materials for lithium ion batteries: recent progress and perspectives. Sustain. Energy. Fuels. 2022, 6, 1014-50.

19. Chen, J.; Minett, A.; Liu, Y.; et al. Direct growth of flexible carbon nanotube electrodes. Adv. Mater. 2008, 20, 566-70.

20. Liu, X.; Huang, Z. D.; Oh, S. W.; et al. Carbon nanotube (CNT)-based composites as electrode material for rechargeable Li-ion batteries: a review. Compos. Sci. Technol. 2012, 72, 121-44.

21. Wang, X.; Waje, M.; Yan, Y. CNT-based electrodes with high efficiency for PEMFCs. Electrochem. Solid-State. Lett. 2005, 8, A42.

22. Hwang, T. H.; Lee, Y. M.; Kong, B. S.; Seo, J. S.; Choi, J. W. Electrospun core-shell fibers for robust silicon nanoparticle-based lithium ion battery anodes. Nano. Lett. 2012, 12, 802-7.

23. Peng, H.; Jain, M.; Peterson, D. E.; Zhu, Y.; Jia, Q. Composite carbon nanotube/silica fibers with improved mechanical strengths and electrical conductivities. Small 2008, 4, 1964-7.

24. Landi, B. J.; Ganter, M. J.; Cress, C. D.; Dileo, R. A.; Raffaelle, R. P. Carbon nanotubes for lithium ion batteries. Energy. Environ. Sci. 2009, 2, 638-54.

25. Varzi, A.; Täubert, C.; Wohlfahrt-mehrens, M.; Kreis, M.; Schütz, W. Study of multi-walled carbon nanotubes for lithium-ion battery electrodes. J. Power. Sources. 2011, 196, 3303-9.

26. Pampal, E. S.; Stojanovska, E.; Simon, B.; Kilic, A. A review of nanofibrous structures in lithium ion batteries. J. Power. Sources. 2015, 300, 199-215.

27. Wu, Y.; Wang, J.; Jiang, K.; Fan, S. Applications of carbon nanotubes in high performance lithium ion batteries. Front. Phys. 2014, 9, 351-69.

28. Qian, W.; Tian, T.; Guo, C.; et al. Enhanced activation and decomposition of CH4 by the addition of C2H4 or C2H2 for hydrogen and carbon nanotube production. J. Phys. Chem. C. 2008, 112, 7588-93.

29. Wei, T.; Bao, L.; Hauke, F.; Hirsch, A. Recent advances in graphene patterning. Chempluschem 2020, 85, 1655-68.

30. Shabaker, J.; Simonetti, D.; Cortright, R.; Dumesic, J. Sn-modified Ni catalysts for aqueous-phase reforming: characterization and deactivation studies. J. Catal. 2005, 231, 67-76.

31. Ma, Z.; Jia, R.; Liu, C. Production of hydrogen peroxide from carbon monoxide, water and oxygen over alumina-supported Ni catalysts. J. Mol. Catal. A. Chem. 2004, 210, 157-63.

32. Wang, G.; Ahn, J.; Yao, J.; Lindsay, M.; Liu, H.; Dou, S. Preparation and characterization of carbon nanotubes for energy storage. J. Power. Sources. 2003, 119-121, 16-23.

33. Wang, X.; Wang, J.; Chang, H.; Zhang, Y. Preparation of short carbon nanotubes and application as an electrode material in Li-ion batteries. Adv. Funct. Mater. 2007, 17, 3613-8.

34. Yang, S.; Huo, J.; Song, H.; Chen, X. A comparative study of electrochemical properties of two kinds of carbon nanotubes as anode materials for lithium ion batteries. Electrochim. Acta. 2008, 53, 2238-44.

35. Pushparaj, V. L.; Shaijumon, M. M.; Kumar, A.; et al. Flexible energy storage devices based on nanocomposite paper. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 13574-7.

Cite This Article

Article
Open Access
Fiber-reinforced CNT-integrated quartz fabrics as multifunctional electrodes for structural lithium-ion batteries

How to Cite

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

© The Author(s) 2025. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
4040
Downloads
1
Citations
0
Comments
0
0

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at [email protected].

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

Portico

All published articles are preserved here permanently:

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

Portico

All published articles are preserved here permanently:

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