Next Article in Journal
Simulations of Femtosecond-Laser Near-Field Ablation Using Nanosphere under Dynamic Excitation
Previous Article in Journal
Configurational Isomerism in Bimetallic Decametalates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improved Energy Density at High Temperatures of FPE Dielectrics by Extreme Low Loading of CQDs

by
Huan Wang
,
Hang Luo
*,
Yuan Liu
,
Fan Wang
,
Bo Peng
,
Xiaona Li
,
Deng Hu
,
Guanghu He
and
Dou Zhang
State Key Laboratory of Powder Metallurgy, Central South University, Changsha 410083, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(14), 3625; https://doi.org/10.3390/ma17143625
Submission received: 13 June 2024 / Revised: 15 July 2024 / Accepted: 19 July 2024 / Published: 22 July 2024

Abstract

:
Electrostatic capacitors, with the advantages of high-power density, fast charging–discharging, and outstanding cyclic stability, have become important energy storage devices for modern power electronics. However, the insulation performance of the dielectrics in capacitors will significantly deteriorate under the conditions of high temperatures and electric fields, resulting in limited capacitive performance. In this paper, we report a method to improve the high-temperature energy storage performance of a polymer dielectric for capacitors by incorporating an extremely low loading of 0.5 wt% carbon quantum dots (CQDs) into a fluorene polyester (FPE) polymer. CQDs possess a high electron affinity energy, enabling them to capture migrating carriers and exhibit a unique Coulomb-blocking effect to scatter electrons, thereby restricting electron migration. As a result, the breakdown strength and energy storage properties of the CQD/FPE nanocomposites are significantly enhanced. For instance, the energy density of 0.5 wt% CQD/FPE nanocomposites at room temperature, with an efficiency (η) exceeding 90%, reached 9.6 J/cm3. At the discharge energy density of 0.5 wt%, the CQD/FPE nanocomposites remained at 4.53 J/cm3 with an efficiency (η) exceeding 90% at 150 °C, which surpasses lots of reported results.

1. Introduction

One of the distinctive features of electrostatic capacitors is their ability to rapidly release stored energy in a very short period of time (on the order of microseconds) to generate powerful power pulses, which makes them play a key role in modern electronic devices and power systems [1,2,3]. Ceramic capacitors are prone to cracking under temperature changes and mechanical stress due to the brittleness of ceramic materials and are large in size [4,5,6]. Film capacitors with polymer dielectrics showing the advantages of high dielectric strength, good processability, and excellent self-healing ability are receiving attention from researchers [7]. However, in extreme environments, especially at high temperatures (more than 150 °C), the energy storage density and efficiency of polymer dielectrics are significantly reduced due to the inevitably increasing conductivity, which cannot meet the demand for electronic devices in emerging applications such as hybrid vehicles, oil and gas exploration, and advanced propulsion systems [8,9]. Currently, commercially available dielectric biaxially oriented polypropylene (BOPP) cannot be operated above 105 °C [10,11]. This is due to the fact that when operating under high electric fields and temperatures exceeding 85 °C, BOPP will result in a significant increase in charge injection and conduction losses, which causes a rapid increase in the leakage current density, leading to a significant decrease in the energy density and storage efficiency [12,13]. Therefore, it is of great significance to develop capacitor dielectric materials that can meet the demands of applications at high temperatures and high electric fields.
Polymers with high T g , such as FPE, PEI, PI, PEN, PC, etc., are demonstrated as potential dielectrics that can be applied in high-temperature environments [14,15]. FPE (fluorene polyester) is a class of amorphous poly aryl esters with a T g as high as 330 °C, and a dielectric loss of less than 0.003 over a temperature range of 25~250 °C. It is an ideal choice for high-temperature dielectric materials attracting increasing attention [11,16,17]. The escalation of thermally activated charge transport under high electric fields at elevated temperatures results in an exponential surge in leakage current, thereby precipitating a significant decline in the energy density and efficiency of pure FPE at high temperatures [2]. Therefore, polymer dielectrics with a high glass transition temperature alone are not enough to solve the above problem. The key point for improving the performance of energy storage in extreme application environments lies in how to address the conduction loss of polymer dielectrics under high temperatures and high fields [9]. Studies show that the loss mechanism of polymer dielectrics under high temperatures and high fields is mainly due to the increase in charge injection and carrier migration. Therefore, finding solutions for suppressing carrier injection and transmission is urgent [18]. Among the various methods, constructing nanocomposite dielectrics has been proven to be an effective strategy [19]. Researchers believe that the introduction of wide bandgap inorganic nanofillers, such as BNNS, Al2O3, MgO, and HfO2, into the polymer matrix can create energy traps in composite dielectrics, thereby suppressing conduction loss [7,19,20,21]. Ren et al. [22] prepared nanocomposites featuring fillers with ZrO2 cores and Al2O3 shells into a polyetherimide (PEI) matrix, significantly reducing the leakage current density by an order of magnitude compared to pure PEI. This innovation enabled an energy storage density of 5.19 J/cm3 (η > 80%) at 150 °C.
Carbon quantum dots (CQDs) are small, dispersed carbon particles less than 10 nanometers in size, featuring functional groups such as hydroxyl, amino, carboxyl, or aromatic rings on their surface [23,24]. These particles have demonstrated excellent photostability, low cytotoxicity, good biocompatibility, ease of surface modification, and high chemical stability [24,25]. Carbon quantum dots (CQDs) are formed by the random polymerization of linear polymers or precursors that are highly cross-linked and slightly graphitized, combining the properties of both polymers and graphitic carbon [26]. Neutral quantum dots effectively trap electrons, while charged quantum dots can create a potential barrier of e2/2C in insulating systems [27]. When this potential barrier significantly exceeds the thermal energy of the electron, it prevents other electrons from passing through, resulting in a Coulomb blockade effect [28,29]. Consequently, uniformly dispersed quantum dots in a polymer matrix can impede carrier migration, leading to enhanced breakdown strength and energy storage capacity. The introduction of carbon quantum dots into a polymer matrix to construct composite dielectric materials exhibits advantages. (1) Compared with inorganic nano-fillers, the rich functional groups on the surface of carbon quantum dots can solve the dispersion and compatibility problems of fillers in polymer matrix, which is conducive to the uniform distribution of electric field; (2) Due to the high electron affinity energy of carbon quantum dots, the carrier transport and migration can be inhibited through the construction of traps; (3) Compared with the semiconductor small molecules with high electron affinity energy, the synthesis process for carbon quantum dots is more simple, economical, and green [30,31,32].
In this study, CQDs with functional groups such as carboxyl and hydroxyl on their surface were synthesized using a simple, mass-producible hydroxyl-acetaldehyde condensation method and incorporated into an FPE matrix for high-temperature capacitive energy storage. The high electron affinity energy of CQDs allows them to capture injected and excited electrons through strong electrostatic attraction, introducing a large trap energy level ( φ e = 1.54) that makes it difficult for trapped electrons to escape from the trap sites, thereby limiting carrier transport and migration. This reduction in conduction loss significantly enhances the energy storage performance of the composites at high temperatures. For instance, at 150 °C, 0.5 wt% CQD/FPE nanocomposites can achieve a discharge energy density of 4.53 J/cm3, with an energy efficiency maintained at 90.4% and a breakdown electric field reaching 596.2 MV/m.

2. Materials and Methods

2.1. Materials

The NaOH, acetaldehyde solution (40% in H2O), and HCl solution (1 M) were purchased from Sinopharm Chemical Reagent Beijing Co., Beijing, China. FPE particles were supplied by PolyK Technologies, LLC, Philipsburg, PA, USA. The 1-methyl-2-pyrrolidone (NMP) solvent was supplied by Sinopharm and was used without further purification.

2.2. Preparation of CQDs

A total of 12 g of sodium hydroxide (NaOH) was mixed with 40 mL of an acetaldehyde solution and stirred for 2 h. Subsequently, dilute hydrochloric acid was added to facilitate dissolution. The mixture was then subjected to centrifugation and washed repeatedly with deionized water until a neutral pH was achieved. The resulting powder was dried at 70 °C for 12 h, yielding approximately 4 (±0.2) g of carbon quantum dot (CQD) powder. The reaction yield was determined to be between 23% and 26%.
Sample purification: Ensuring the purity of CQDs synthesized by this method is paramount. To validate this, we employed the following analytical strategies. Initially, We performed an XPS test on the CQDs powder and detailed results are given later, wherein no Na peaks were discernible in the CQDs. This finding attests to the effectiveness of aqueous washing in removing sodium ion residues. Building on this, we redissolved the CQDs in anhydrous ethanol for additional dialysis and quantified the residual Na content using atomic absorption spectroscopy (AAS). The results of the AAS analysis are summarized in Table 1 below.
The data indicated that the Na content in the carbon dots was exceedingly low. Furthermore, the Na content in the solution outside the dialysis bag remained virtually unchanged, suggesting that the dialysis method did not notably enhance the purity of the carbon dots. The presence of Na in the external solution may also be attributed to the prolonged use of glassware during the dialysis process. In light of these findings, we conclude that the carbon dots synthesized and purified using our method fully satisfy the requirements of our study.

2.3. Preparation of Composites

The preparation of CQD/FPE composites utilized the solution casting method. Initially, the CQDs were ultrasonicated for 15 min to ensure their dispersion in NMP solvent. Subsequently, FPE particles were added to the solution and stirred at 60 °C for 12 h. The resulting mixture was then cast onto a glass plate and placed in a 70 °C blast drying oven for 1 h to partially evaporate the NMP. It was then transferred to a vacuum drying oven at 100 °C and held for 3 h, heated to 150 °C and held for 3 h, heated to 200 °C, and finally left for 18 h to completely remove residual solvents to form a composite film.

2.4. Characterizations

The samples were subjected to a comprehensive array of advanced characterization techniques. Transmission electron microscopy (TEM, JEM-1210F, JEOL, Tokyo, Japan) and scanning electron microscopy (SEM, Nova NanoSEM230, FEI, Hillsboro, OR, USA) were utilized to scrutinize the micro-morphology. X-ray photoelectron spectroscopy (XPS, ESCALAB 250xi, Thermo Fisher Scientific, Waltham, MA, USA), X-ray diffraction (XRD, D/max 2550, Rigaku Corporation, Tokyo, Japan), and Fourier Transform Infrared Spectroscopy (FTIR, IS-50, Thermo Fisher Scientific, Waltham, MA, USA) were employed to investigate the structural and compositional attributes. UV–Vis spectroscopy (TU-1901, Shanghai Xipu Instrument Co., Ltd., Liushi, China) provided the raw absorption spectra, from which the band gaps of the carbon quantum dots (CQDs) were determined using the Tauc plot method. The formula in Equation (1) was applied, where α is the absorbance index, h is Planck’s constant, v is the frequency, A is a constant, Eg is the bandgap energy, and n is 1/2 for direct bandgap semiconductors such as CQDs [3].
( α h v ) 1 / n = A ( h v E g )
The mechanical properties were evaluated using an ultra-nanometer hardness tester (UNHT, CSM, Peseux, Switzerland), which yielded load-displacement curves. We calculated the Young’s modulus from these curves employing the Oliver–Pharr method. The indenter is a diamond Berkovich indenter. The geometry of the indenter was a three-sided pyramid, and the loading protocol was as follows: acquisition rate of 10.0 Hz, maximum load of 1.00 mN, loading rate of 6.00 mN/min, unloading rate of 6.00 mN/min, and a pause time of 2.0 s. For electrical property testing, gold electrodes with a sputtered area of 3.14 mm² were deposited on each side of the samples. The dielectric characteristics were assessed with an Agilent 4294A LCR meter over a frequency range of 1 kHz to 10 MHz. The capacitance value (Cp) and loss value (D) were measured, and the dielectric constant (εr) was calculated using the formula in Equation (2), where ε0 is the vacuum dielectric constant (8.85 × 10−12 F/m), S is the electrode area, and d is the sample thickness [5].
ε r = C p × d ε 0 × S
Pulsed discharge characterization was conducted with a DCQ-20A discharge measurement system (PolyK Technologies, North Philipsburg, PA, USA). High-temperature dielectric properties were measured using a DMS-500 high-temperature dielectric spectroscopy system, which provided data at 1 kHz over a temperature range of 30 to 220 °C. The displacement hysteresis return line (D-E loop) was analyzed using a TF Analyzer 2000 series (aixACCT, Aachen, Germany) at 10 Hz. The charged energy density (U) was derived by integrating the area between the charge curve and the ordinate, while the discharge energy density (Ue) was determined by the area between the discharge curve and the ordinate. The efficiency (η) was calculated as the ratio of Ue to U [33]. Cyclic voltammetry curves were obtained using a CHI electrochemical workstation, and the lowest unoccupied molecular orbital (LUMO) and highest occupied molecular orbital (HOMO) energies were calculated from the reduction potential (Ered) using the Equations (3) and (4) [34].
E L U M O = e ( E r e d + 4.40 )
E H O M O = E L U M O E g
The spatial distribution of the electrostatic potential for the FPE model was elucidated through Density Functional Theory (DFT) computations grounded in first-principles calculations. The wavefunction of the molecule under investigation was derived from the solution of the fundamental Schrödinger equation. In this study, DFT calculations were conducted using the B3LYP hybrid functional and the 6-311G(d) basis set without imposing restrictive symmetry constraints within the Gaussian 16 computational package.

3. Results

Figure 1a presents TEM images obtained from the CQD powder at various magnifications. Figure 1b displays photographs of the yellow CQD powder and its solution in NMP, revealing a homogeneous and transparent state, indicative of excellent dispersion in the NMP solvent.
Figure 2a depicts the FTIR spectrum of the CQDs, highlighting absorption bands at 3424 cm−1 (associated with -OH stretching), 2971 cm−1 (-CH stretching), 1716 cm−1 (-C=O stretching), 1681 cm−1 (-C=C stretching), 1378 cm−1 (-CH3 stretching), and 1073 cm−1 (-CO stretching). XPS plots (Figure 2c,d) confirmed the presence of two types of carbon: graphitized carbon (C=C and C-C) with binding energy at 284.7 eV and aliphatic carbon (C-O, C=O, and O=C-OH) with binding energies at 285.8 eV, 287.3 eV, and 289.0 eV. Both FTIR and XPS analyses suggest the presence of oxygen-containing functional groups, such as hydroxyl and carboxyl, which contribute to the good dispersion and interfacial compatibility of the CQDs within the polymer matrix. The XRD pattern (Figure 2b) shows a single broad diffraction peak centered around 2θ = 18°, indicative of a highly disordered carbon structure consistent with the TEM observations. The UV–visible absorption spectra (Figure 2e) were used to calculate the ( α h ν ) 2 h ν curve (Figure 2f) using the equation ( α h ν ) 2 = A ( h ν E g ) , where α is the absorption coefficient, h is Planck’s constant, ν is the optical frequency, and Eg is the band gap energy [35]. This analysis determined a band gap ( E g ) of approximately 2.33 eV for the CQDs, confirming their semiconductor nature.
The efficacy of CQDs in mitigating charge accumulation within dielectrics and curbing conductivity losses stems from their elevated electron affinity energy. This property enables the formation of a trap energy level ( φ e = E A C Q D s E A p o l y m e r ) within the polymer matrix, facilitating the capture of substantial charge injected by metal electrodes [35,36]. Electrostatic potential maps and energy band values of the FPE, as derived from density functional theory (DFT) simulations, are depicted in Figure 3b. The lowest unoccupied molecular orbital (LUMO) energy level of the FPE was determined to be −2.60 eV. To ascertain the electron affinity energies of the carbon quantum dots (CQDs), cyclic voltammetry tests were conducted using a conventional three-electrode electrochemical testing setup, yielding the cyclic voltammetry curves presented in Figure 3c. The onset reduction potential ( E r e d ) of the polymer dot, discernible from the figure, was −0.98 eV. The bandgap of CQDs was calculated by Equations (3) and (4); the LUMO energy level of the polymer was calculated to be −4.14 eV. Considering the definition of electron affinity energy (the energy necessary for electrons to reach the vacuum energy level from the bottom of the conduction band) the value is the energy difference from the conduction band’s base to the vacuum level (Evac) [9]. Hence, EACQDs = Evac − ELUMO = 4.14 eV [35]. The integration of CQDs thus facilitates the establishment of a profound trap energy level, as illustrated in Figure 3a, with a depth of 1.54 eV.
The variation of dielectric constants and loss in the composites with frequency is depicted in Figure 4a. The increase in the dielectric constant correlated with the rising content of filler CQDs, attributed to the expansion of the interfacial region within the composites, which enhances interfacial polarization due to the incorporation of CQDs [37]. The dielectric loss initially decreased and then increased as the CQD content rose. Notably, the dielectric loss for all composites was lower than that of pure FPE (0.008) at 1 kHz, primarily due to the Coulomb-blocking effect of CQDs [28,29]. Neutral CQDs act as traps in the electron transport pathway, while charged CQDs create an energy barrier impeding electron flow through the polymer matrix, thereby reducing conductive loss. At lower filler concentrations, the Coulomb-blocking effect intensifies with increasing filler content, impacting the dielectric loss more significantly than interfacial polarization. However, once the filler content exceeds a certain value, the reduced spacing between CQDs facilitates electron conduction, and the inhibition of electron migration diminishes, resulting in an increase in conductive loss [38]. Consequently, samples with 1.0 wt% CQDs exhibited higher dielectric losses than those with 0.5 wt% CQDs. The dielectric constant and loss of all thin film samples remained stable across the frequency range from 102 Hz to 104 Hz, with the dielectric loss beginning to rise only after the frequency exceeds 104 Hz.
Figure 4b illustrates the variation of dielectric properties with the temperature at 1 kHz, and the dielectric loss remained below 0.01 across the temperature range from 30 °C to 220 °C, a characteristic attributed to the high glass transition temperature of FPE [39]. Notably, the dielectric loss decreased with increasing temperatures at 120 °C, a trend that is advantageous to achieve a superior high-temperature energy storage performance. To facilitate a clearer comparison of the dielectric properties among all samples, the values of dielectric constant and loss measured at room temperature and 150 °C for pure FPE and composite dielectrics at 1 kHz are presented in Figure 4c,d. At 150 °C, the dielectric constant of the 0.5 wt% CQD/FPE composite increased from 3.36 to 3.56 compared to pure FPE, and the dielectric loss decreased by 18.8% at high temperatures relative to that of pure FPE.
To examine the impact of CQDs on the breakdown strength of the polymers, the breakdown field strength and stability of the nanocomposites were analyzed using the Weibull distribution [20].
P ( E ) = 1 e x p ( E E b ) β
The text discusses the cumulative failure probability, P(E), where E represents the tested breakdown strength value, β is a shape parameter reflecting the sample data dispersion, and Eb is the characteristic breakdown strength for a specimen with a 63.2% breakdown probability [40]. Figure 5a,b illustrates the Weibull distribution of nanocomposites with varying filler contents at room temperature and 150 °C. To better depict the temperature-dependent variation in the breakdown field strength of the composites, Figure 5c,d presents a bar graph summarizing the trend of breakdown field strength with filler content. It is evident that the breakdown strength of the nanocomposites initially increased and then decreased as the CQDs content rose. Notably, the breakdown strength peaked at 632.8 MV/m for a 0.5 wt% CQD content, a substantial improvement over the pure FPE’s 567.86 MV/m. At 150 °C, the 0.5 wt% CQD/FPE exhibited a breakdown strength of 596.15 MV/m, which was 11% higher than that of the pure FPE. This trend is attributed to the small number of CQDs enhancing the breakdown strength due to their high electron affinity energy and distinctive Coulomb-blocking effect [41]. However, increasing the filler content beyond this point resulted in reduced inter-filler distances, which can form conductive paths and diminish the material’s breakdown strength [42].
Figure 6a,b displays the results of leakage current density and conductivity tests for pure polymer and composite dielectrics at 150 °C. The introduction of CQDs significantly reduced these values by an order of magnitude. For instance, the leakage current density and conductivity of the 0.5 wt% CQD/FPE dropped from 1.02 × 10−6 A cm−2 and 5.52 × 10−6 S m−1 for the FPE to 4.26 × 10−7 A cm−2 and 6.40 × 10−7 S m−1, respectively. In addition, we found that the leakage current density of all nanocomposites increased exponentially with an increasing electric field, indicating that the main charge transfer mechanism in the materials was hopping conduction. We derived the jump distance (average spacing between trap points) by fitting the J-E data to the jump conduction model, utilizing the expression for jump conduction:
J = n c v λ e × exp W a K B T sinh ( λ e E 2 K b T )
Here, nc denotes the carrier concentration, λ represents the hopping distance, v signifies the attempt-to-escape frequency, Wa is the activation energy expressed in electron volts (eV), e is the elementary charge of the carriers, T is the temperature, and Kb is the Boltzmann constant [28]. The 0.5% CQD/FPE, exhibiting the smallest λ (1.36 nm), demonstrated the highest trap density compared to the composite film with the pure FPE and other composites containing varying filler concentrations, demonstrating the largest trap density, thereby enhancing the Eb [29]. The plot of ln(J) versus 1/T is provided, illustrating the temperature dependence of the current. The activation energies determined for the pure FPE and the composites with 0.5% CQDs were 0.65 eV and 0.81 eV, respectively. This indicates that CQDs with higher electron affinity energies can establish trap energy levels within the polymers, effectively trapping and binding charges [9]. The breakdown mechanisms of polymer dielectrics encompass mechanical breakdown alongside electrical and thermal breakdown. When the applied stress surpasses the yield strength of the nanocomposite film, the gold electrodes, formed through sputtering, will fail and initiate a discharge, ultimately leading to the film’s breakdown [20,43]. To delve deeper into the impact of the material’s mechanical strength on the breakdown strength, nanoindentation tests were conducted on the nanocomposite film, yielding load-displacement curves and Young’s modulus values, as shown in Figure 6c,d. As the CQDs filler content increased, the Young’s modulus of the composites rose. Due to the abundant functional groups on the CQDs surface, CQDs could form a strong bonding network with the FPE matrix. This leads to enhanced interfacial compatibility between the filler and the FPE matrix, resulting in a more uniform filler distribution and an increased Young’s modulus for the nanocomposites. However, as the filler content continues to increase, partial agglomeration of CQDs may occur, leading to uneven forces within the nanocomposite, thus causing Young’s modulus to initially increase and then decrease with rising filler content [44].

4. Discussion

The energy storage density and efficiency of the composites at room temperature and 150 °C are depicted in Figure 7a,b. Notably, the sample with 0.5 wt% CQDs achieved the highest breakdown strength of 632 MV/m and an energy density of 9.6 J/cm3 (η > 90%). It was also significant that both the pure FPE and its composite dielectric maintained efficiencies above 88%. At 150 °C, the 0.5 wt% CQD/FPE composite reached an efficiency of 90.34% and an energy storage density of 4.53 J/cm3, which was 2.53 times greater than the 1.79 J/cm3 of pure FPE. Figure 7c illustrates the relationship between composite energy storage density and filler content at room temperature and 150 °C, with efficiencies exceeding 90%. Figure 7d similarly shows the trend of maximum energy storage density with filler content at both temperatures. It is evident that the energy storage density and the maximum energy storage density for composites with efficiencies above 90% exhibited a trend of increasing and then decreasing as filler content increases.
Furthermore, Figure 8a presents the U e and η results after 105 charge/discharge cycles at 150 °C, 200 MV/m, and 1 kHz. The cycling test revealed no degradation in sample performance, with a slight improvement, indicating excellent cycling stability. The 0.5 wt% CQD/FPE nanocomposites outperformed pure FPE in this regard. The actual pulse discharge energy densities were assessed using a 10 kΩ load resistor to create an overdamped state circuit at 150 °C. The 0.5 wt% CQD/FPE nanocomposites underwent cycling tests at electric fields ranging from 50 to 200 MV/m and at 1 kHz. The U e and η results post-105 cycles are shown in Figure 8a. Figure 8b displays the overdamped discharge curves of FPE across the 50 to 200 MV/m electric field range. The pulsed discharge energy density (Wd) was calculated using the formula in Equation (7):
W d = R I 2 t d t / V
where R, I, t, and V represent load resistance, current, time, and sampling volume, respectively [45]. The calculated curves are presented in Figure 8c,d, with a high Wd of 0.61 J/cm3 achieved at 150 °C and an electric field of 200 MV/m, along with a rapid discharge time of 10 μs, surpassing the 0.43 J/cm3 (120 °C) of commercial BOPP films.
Table 2 compares the energy storage properties of 0.5 wt% CQD/FPE composites with those of several polymer matrix composites reported in the literature. At 150 °C, the performance of the composites in this study surpassed that of most reported dielectric materials. The 0.5 wt% CQD/FPE composites exhibited a synergistic enhancement in both the dielectric constant and breakdown strength while maintaining a high energy efficiency.

5. Conclusions

In summary, CQDs were synthesized via a hydroxyl aldehyde condensation method. Due to the abundant surface functional groups, CQDs enable uniform dispersion in the polymer matrix. They also exhibit high electron affinity energy and a unique Coulomb-blocking effect, which can trap migrating carriers and reduce their migration rate. The results indicated that the addition of CQDs significantly improved the dielectric constant and breakdown strength of the composites. At room temperature, the breakdown strength of 0.5 wt% CQD/FPE composites reached 632.89 MV/m, a 13% increase over that of the pure polymer. The energy storage density of the composites was 37% greater than that of pure FPE, and the energy efficiency exceeded 90%. At 150 °C, the incorporation of CQDs effectively mitigated the rapid decline in breakdown strength, energy density, and energy efficiency typically observed in pure FPE as the temperature increased. The 0.5 wt% CQD/FPE composites demonstrated a breakdown strength of 596.15 MV/m, an 11% improvement over the pure polymer, and an energy storage density of 4.53 J/cm3, which represents a 2.53-fold increase compared to that of pure FPE.

Author Contributions

Methodology, Y.L., D.H., F.W., X.L. and H.W.; investigation, D.H., B.P., H.L., F.W., Y.L., G.H. and H.W.; data measurement, X.L., D.Z., H.L. and H.W.; data analysis, all authors; writing—original draft preparation, H.W. and B.P.; writing—review and editing, H.L., D.Z. and H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (52172265), a Scientific research project of the Hunan Provincial Department of Education (21B0009), Excellent Youth Science Foundation of Hunan Province (2022JJ20067), the Science and Technology Innovation Program of Hunan Province (2022RC1074), Central South University Innovation-Driven Research Program (2023CXQD010) and State Key Laboratory of Powder Metallurgy, Central South University, Changsha, China.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Luo, F.Y.; Li, Y.T.; Zhang, J.Y.; He, L.; Li, J.L.; Sun, N.; Li, G.L.; Jiang, Y.; Zhou, K.; Liang, Q.Q.; et al. Scalable Dual In Situ Synthesis of Polyester Nanocomposites for High-Energy Storage. Small 2024, 2401308. [Google Scholar] [CrossRef]
  2. Li, X.N.; Luo, H.; Yang, C.C.; Wang, F.; Jiang, X.; Guo, R.; Zhang, D. Enhancing High-Temperature Energy Storage Performance of PEI-Based Dielectrics by Incorporating ZIF-67 with a Narrow Bandgap. ACS Appl. Mater. Interfaces 2023, 15, 41828–41838. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, F.; Cai, J.M.; Yang, C.C.; Luo, H.; Li, X.A.; Hou, H.S.; Zou, G.Q.; Zhang, D. Improved Capacitive Energy Storage Nanocomposites at High Temperature Utilizing Ultralow Loading of Bimetallic MOF. Small 2023, 19, 2300510. [Google Scholar] [CrossRef]
  4. Xie, Z.L.; Le, K.; Li, H.; Pang, X.; Xu, T.L.; Altoe, V.; Klivansky, L.M.; Wang, Y.F.; Huang, Z.Y.; Shelton, S.W.; et al. Interfacial Engineering Using Covalent Organic Frameworks in Polymer Composites for High-Temperature Electrostatic Energy Storage. Adv. Funct. Mater. 2024, 34, 2314910. [Google Scholar] [CrossRef]
  5. He, G.H.; Liu, Y.; Wang, C.; Chen, S.; Luo, H.; Zhang, D. All-organic polymer dielectrics prepared via optimization of sequential structure of polystyrene-based copolymers. Chem. Eng. J. 2022, 446, 137106. [Google Scholar] [CrossRef]
  6. Liu, Y.; Luo, H.; Xie, H.; Xiao, Z.; Wang, F.; Jiang, X.; Zhou, X.; Zhang, D. Trilayer PVDF nanocomposites with significantly enhanced energy density and energy efficiency using 0.55Bi0.5Na0.5TiO3-0.45(Sr0.7Bi0.2)TiO3 nanofibers. Microstructures 2023, 3, 2023008. [Google Scholar]
  7. Ren, L.L.; Yang, L.J.; Zhang, S.Y.; Li, H.; Zhou, Y.; Ai, D.; Xie, Z.L.; Zhao, X.T.; Peng, Z.R.; Liao, R.J.; et al. Largely enhanced dielectric properties of polymer composites with HfO2nanoparticles for high-temperature film capacitors. Compos. Sci. Technol. 2021, 201, 108528. [Google Scholar] [CrossRef]
  8. Chen, J.; Pei, Z.T.; Liu, Y.J.; Shi, K.M.; Zhu, Y.K.; Zhang, Z.C.; Jiang, P.K.; Huang, X.Y. Aromatic-Free Polymers Based All-Organic Dielectrics with Breakdown Self-Healing for High-Temperature Capacitive Energy Storage. Adv. Mater. 2023, 35, 2306562. [Google Scholar] [CrossRef] [PubMed]
  9. Yuan, C.; Zhou, Y.; Zhu, Y.J.; Liang, J.J.; Wang, S.J.; Peng, S.M.; Li, Y.S.; Cheng, S.; Yang, M.C.; Hu, J.; et al. Polymer/molecular semiconductor all-organic composites for high-temperature dielectric energy storage. Nat. Commun. 2020, 11, 3919. [Google Scholar] [CrossRef] [PubMed]
  10. Xu, W.H.; Zhou, C.Y.; Ji, W.H.; Zhang, Y.H.; Jiang, Z.H.; Bertram, F.; Shang, Y.S.; Zhang, H.B.; Shen, C. Anisotropic Semicrystalline Homopolymer Dielectrics for High-Temperature Capacitive Energy Storage. Angew. Chem.-Int. Edit. 2024, 63, e202319766. [Google Scholar] [CrossRef] [PubMed]
  11. Zhang, K.Y.; Ma, Z.Y.; Deng, H.; Fu, Q. Improving high-temperature energy storage performance of PI dielectric capacitor films through boron nitride interlayer. Adv. Compos. Hybrid Mater. 2022, 5, 238–249. [Google Scholar] [CrossRef]
  12. Zha, J.W.; Xiao, M.Y.; Wan, B.Q.; Wang, X.M.; Dang, Z.M.; Chen, G. Polymer dielectrics for high-temperature energy storage: Constructing carrier traps. Prog. Mater. Sci. 2023, 140, 101208. [Google Scholar] [CrossRef]
  13. Xiong, J.; Fan, X.; Long, D.J.; Zhu, B.F.; Zhang, X.; Lu, J.Y.; Xie, Y.C.; Zhang, Z.C. Significant improvement in high-temperature energy storage performance of polymer dielectrics via constructing a surface polymer carrier trap layer. J. Mater. Chem. A 2022, 10, 24611–24619. [Google Scholar] [CrossRef]
  14. Li, H.; Gadinski, M.R.; Huang, Y.Q.; Ren, L.L.; Zhou, Y.; Ai, D.; Han, Z.B.; Yao, B.; Wang, Q. Crosslinked fluoropolymers exhibiting superior high-temperature energy density and charge-discharge efficiency. Energy Environ. Sci. 2020, 13, 1279–1286. [Google Scholar] [CrossRef]
  15. Luo, H.; Wang, F.; Guo, R.; Zhang, D.; He, G.H.; Chen, S.; Wang, Q. Progress on Polymer Dielectrics for Electrostatic Capacitors Application. Adv. Sci. 2022, 9, 2202438. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, W.C.; Guan, F.; Jiang, M.; Li, Y.P.; Zhu, C.C.; Yue, D.; Li, J.L.; Liu, X.X.; Feng, Y. Enhanced energy storage performance of all-organic sandwich structured dielectrics with FPE and P(VDF-HFP). Compos. Part A-Appl. Sci. Manuf. 2022, 159, 107018. [Google Scholar] [CrossRef]
  17. Liu, S.Y.; Zhang, W.C.; Guan, F.; Yao, Y.H.; Yue, D.; Feng, Y. The simple synthesis of all-organic polymer bilayer films: Achieving simultaneous enhancements in energy storage density and efficiency. J. Appl. Polym. Sci. 2023, 140, e53991. [Google Scholar] [CrossRef]
  18. Meng, Z.T.; Zhang, T.D.; Zhang, C.H.; Shang, Y.N.; Lei, Q.Q.; Chi, Q.G. Advances in Polymer Dielectrics with High Energy Storage Performance by Designing Electric Charge Trap Structures. Adv. Mater. 2023, e2310272. [Google Scholar] [CrossRef] [PubMed]
  19. Ai, D.; Li, H.; Zhou, Y.; Ren, L.L.; Han, Z.B.; Yao, B.; Zhou, W.; Zhao, L.; Xu, J.M.; Wang, Q. Tuning Nanofillers in In Situ Prepared Polyimide Nanocomposites for High-Temperature Capacitive Energy Storage. Adv. Energy Mater. 2020, 10, 1903881. [Google Scholar] [CrossRef]
  20. Liu, X.J.; Zheng, M.S.; Chen, G.; Dang, Z.M.; Zha, J.W. High-temperature polyimide dielectric materials for energy storage: Theory, design, preparation and properties. Energy Environ. Sci. 2022, 15, 56–81. [Google Scholar] [CrossRef]
  21. Wang, P.J.; Guo, Y.; Zhou, D.; Li, D.; Pang, L.X.; Liu, W.F.; Su, J.Z.; Shi, Z.Q.; Sun, S.K. High-Temperature Flexible Nanocomposites with Ultra-High Energy Storage Density by nanostructured MgO fillers. Adv. Funct. Mater. 2022, 32, 2204155. [Google Scholar] [CrossRef]
  22. Ren, L.L.; Li, H.; Xie, Z.L.; Ai, D.; Zhou, Y.; Liu, Y.; Zhang, S.Y.; Yang, L.J.; Zhao, X.T.; Peng, Z.R.; et al. High-Temperature High-Energy-Density Dielectric Polymer Nanocomposites Utilizing Inorganic Core-Shell Nanostructured Nanofillers. Adv. Energy Mater. 2021, 11, 2101297. [Google Scholar]
  23. Liu, H.X.; Zhong, X.; Pan, Q.; Zhang, Y.; Deng, W.T.; Zou, G.Q.; Hou, H.S.; Ji, X.B. A review of carbon dots in synthesis strategy. Coord. Chem. Rev. 2024, 498, 215468. [Google Scholar] [CrossRef]
  24. Liu, Y.Q.; Liang, F.Y.; Sun, J.L.; Sun, R.; Liu, C.; Deng, C.; Seidi, F. Synthesis Strategies, Optical Mechanisms, and Applications of Dual-Emissive Carbon Dots. Nanomaterials 2023, 13, 2869. [Google Scholar] [CrossRef] [PubMed]
  25. Miao, S.H.; Liang, K.; Zhu, J.J.; Yang, B.; Zhao, D.Y.; Kong, B. Hetero-atom-doped carbon dots: Doping strategies, properties and applications. Nano Today 2020, 33, 100879. [Google Scholar] [CrossRef]
  26. Li, L.; Li, Y.T.; Ye, Y.; Guo, R.T.; Wang, A.N.; Zou, G.Q.; Hou, H.S.; Ji, X.B. Kilogram-Scale Synthesis and Functionalization of Carbon Dots for Superior Electrochemical Potassium Storage. ACS Nano 2021, 15, 6872–6885. [Google Scholar] [CrossRef] [PubMed]
  27. Li, S.; Dong, J.F.; Niu, Y.J.; Li, L.; Wang, F.; Hu, R.C.; Cheng, J.; Sun, L.; Pan, Z.Z.; Xu, X.W.; et al. Enhanced high-temperature energy storage properties of polymer composites by interlayered metal nanodots. J. Mater. Chem. A 2022, 10, 18773–18781. [Google Scholar] [CrossRef]
  28. Cai, H.C.; Wang, R.; Gou, B.; Fu, J.; Zhu, Y.J.; Yang, H.; Zhou, J.A.; Li, M.X.; Zhong, A.; Zhang, D.M.; et al. Significantly improved high-temperature capacitive performance in polymer dielectrics utilizing ultra-small carbon quantum dots with Coulomb-blockade effect. Chem. Eng. J. 2023, 476, 146672. [Google Scholar] [CrossRef]
  29. Wang, L.W.; Huang, X.Y.; Zhu, Y.K.; Jiang, P.K. Enhancing electrical energy storage capability of dielectric polymer nanocomposites via the room temperature Coulomb blockade effect of ultra-small platinum nanoparticles. Phys. Chem. Chem. Phys. 2018, 20, 5001–5011. [Google Scholar] [CrossRef] [PubMed]
  30. Xie, H.R.; Luo, H.; Liu, Y.; Guo, R.; Ji, X.B.; Hou, H.S.; Zhang, D. Extremely low loading of carbon quantum dots for high energy density in polyetherimide nanocomposites. Chem. Eng. J. 2022, 433, 133601. [Google Scholar] [CrossRef]
  31. Jiang, X.; Luo, H.; Wang, F.; Li, X.A.; Xie, H.R.; Liu, Y.; Zou, G.Q.; Ji, X.B.; Hou, H.S.; Zhang, D. Ultralow loading of carbon quantum dots leading to significantly improved breakdown strength and energy density of P(VDF-TrFE-CTFE). J. Mater. Chem. A 2023, 11, 16127–16137. [Google Scholar] [CrossRef]
  32. Ding, J.L.; Zhou, Y.; Xu, W.H.; Yang, F.; Zhao, D.Y.; Zhang, Y.H.; Jiang, Z.H.; Wang, Q. Ultraviolet-Irradiated All-Organic Nanocomposites with Polymer Dots for High-Temperature Capacitive Energy Storage. Nano-Micro Lett. 2024, 16, 59. [Google Scholar] [CrossRef] [PubMed]
  33. Jin, Z.; Lin, Y.; Wang, Y.; Yang, H. Sandwich-structured SrTiO3/PEI composite films with high-temperature energy storage performance. J. Power Sources 2024, 609, 234677. [Google Scholar] [CrossRef]
  34. Zhang, S.C.; Chen, D.; Liu, Z.F.; Ruan, M.N.; Guo, Z.A. Novel strategy for efficient water splitting through pyro-electric and pyro-photo-electric catalysis of BaTiO3 by using thermal resource and solar energy. Appl. Catal. B-Environ. 2021, 284, 119686. [Google Scholar] [CrossRef]
  35. Ding, J.L.; Xu, W.H.; Zhu, X.B.; Liu, Z.; Zhang, Y.H.; Jiang, Z.H. All-organic nanocomposite dielectrics contained with polymer dots for high-temperature capacitive energy storage. Nano Res. 2023, 16, 10183–10190. [Google Scholar] [CrossRef]
  36. Wang, Q.L.; Tanaka, Y.; Miyake, H.; Endo, K.; An, Y.; Du, H.S.; Chen, X.R.; Paramane, A. Enhanced high-temperature electrical properties and charge dynamics of inorganic/organic silicone elastomer nanocomposites via nanostructure grafting and molecular trap construction. Polym. Compos. 2024. [Google Scholar] [CrossRef]
  37. Lin, Q.; Hou, Y.F.; Zheng, Y.Q.; Tan, Y.N.; Du, P.; Luo, L.H.; Li, W.P. Poly(ether imide)-Based Nanocomposites with Low Fraction of Hierarchical Ag@AO Nanofiber for High-Temperature Energy Storage. ACS Appl. Energ. Mater. 2022, 5, 2329–2338. [Google Scholar] [CrossRef]
  38. Fan, M.Z.; Hu, P.H.; Dan, Z.K.; Jiang, J.Y.; Sun, B.Z.; Shen, Y. Significantly increased energy density and discharge efficiency at high temperature in polyetherimide nanocomposites by a small amount of Al2O3 nanoparticles. J. Mater. Chem. A 2020, 8, 24536–24542. [Google Scholar] [CrossRef]
  39. Hassan, Y.A.; Hu, H.L. Current status of polymer nanocomposite dielectrics for high-temperature applications. Compos. Part A-Appl. Sci. Manuf. 2020, 138, 106064. [Google Scholar] [CrossRef]
  40. Ren, W.B.; Pan, J.Y.; Dan, Z.K.; Zhang, T.; Jiang, J.Y.; Fan, M.Z.; Hu, P.H.; Li, M.; Lin, Y.H.; Nan, C.W.; et al. High-temperature electrical energy storage performances of dipolar glass polymer nanocomposites filled with trace ultrafine nanoparticles. Chem. Eng. J. 2021, 420, 127614. [Google Scholar] [CrossRef]
  41. Feng, M.J.; Feng, Y.; Zhang, C.H.; Zhang, T.D.; Chen, Q.G.; Chi, Q.G. Ultrahigh energy storage performance of all-organic dielectrics at high-temperature by tuning the density and location of traps. Mater. Horiz. 2022, 9, 3002–3012. [Google Scholar] [CrossRef] [PubMed]
  42. Li, H.; Ai, D.; Ren, L.L.; Yao, B.; Han, Z.B.; Shen, Z.H.; Wang, J.J.; Chen, L.Q.; Wang, Q. Scalable Polymer Nanocomposites with Record High-Temperature Capacitive Performance Enabled by Rationally Designed Nanostructured Inorganic Fillers. Adv. Mater. 2019, 31, e1900875. [Google Scholar] [CrossRef] [PubMed]
  43. Yu, X.; Yang, R.; Shen, G.Y.; Sun, K.X.; Lv, F.C.; Fan, S.D. Tailoring anisotropic thermal conductivity of 2D aramid nanoribbon-based dielectrics with potential high-temperature capacitive energy storage. J. Mater. Chem. C 2024, 12, 7338–7350. [Google Scholar] [CrossRef]
  44. Yang, M.H.; Zhao, Y.L.; Wang, Z.P.; Yan, H.R.; Liu, Z.R.; Li, Q.; Dang, Z.M. Surface ion-activated polymer composite dielectrics for superior high-temperature capacitive energy storage. Energy Environ. Sci. 2024, 17, 1592–1602. [Google Scholar] [CrossRef]
  45. Yuan, Y.; Wang, X.H.; Ni, X.Y.; Qian, J.; Zuo, P.Y.; Zhuang, Q.X. Excellent high-temperature energy storage capacity for polyetherimide nanocomposites with hierarchically structured nanofillers. Compos. Part A-Appl. Sci. Manuf. 2023, 175, 107791. [Google Scholar] [CrossRef]
  46. Yan, J.J.; Wang, J.; Zeng, J.Y.; Shen, Z.H.; Li, B.W.; Zhang, X.; Zhang, S.J. Modulating the charge trapping characteristics of PEI/BNNPs dilute nanocomposite for improved high-temperature energy storage performance. J. Mater. Chem. C 2022, 10, 13157–13166. [Google Scholar] [CrossRef]
  47. Zhi, J.P.; Wang, J.; Shen, Z.H.; Li, B.W.; Zhang, X.; Nan, C.W. Nanofiber-reinforced polymer nanocomposite with hierarchical interfaces for high-temperature dielectric energy storage applications. Sci. China-Mater. 2023, 66, 2652–2661. [Google Scholar] [CrossRef]
  48. Fan, Z.H.; Gao, S.B.; Chang, Y.F.; Wang, D.W.; Zhang, X.; Huang, H.T.; He, Y.B.; Zhang, Q.F. Ultra-superior high-temperature energy storage properties in polymer nanocomposites via rational design of core-shell structured inorganic antiferroelectric fillers. J. Mater. Chem. A 2023, 11, 7227–7238. [Google Scholar] [CrossRef]
  49. Li, Z.W.; Qin, H.M.; Song, J.H.; Liu, M.; Zhang, X.L.; Wang, S.; Xiong, C.X. Polyimide Nanodielectrics Doped with Ultralow Content of MgO Nanoparticles for High-Temperature Energy Storage. Polymers 2022, 14, 2918. [Google Scholar] [CrossRef] [PubMed]
  50. Ding, X.P.; Pan, Z.B.; Zhang, Y.; Shi, S.H.; Cheng, Y.; Chen, H.X.; Li, Z.C.; Fan, X.; Liu, J.J.; Yu, J.H.; et al. Regulation of Interfacial Polarization and Local Electric Field Strength Achieved Highly Energy Storage Performance in Polyetherimide Nanocomposites at Elevated Temperature via 2D Hybrid Structure. Adv. Mater. Interfaces 2022, 9, 2201100. [Google Scholar] [CrossRef]
Figure 1. (a) TEM images of CQDs under different magnifications, and (b) photographs of CQD powder and solution in NMP solvent.
Figure 1. (a) TEM images of CQDs under different magnifications, and (b) photographs of CQD powder and solution in NMP solvent.
Materials 17 03625 g001
Figure 2. Characterization data of the carbon quantum dots. (a) FTIR spectrum, (b) XRD pattern, (c) full XPS spectrum, (d) high-resolution C1s XPS spectrum, (e) UV–Vis absorbance spectroscopy, and (f) (αhv)2hv plot, which is a common method for determining the bandgap energy of semiconductors.
Figure 2. Characterization data of the carbon quantum dots. (a) FTIR spectrum, (b) XRD pattern, (c) full XPS spectrum, (d) high-resolution C1s XPS spectrum, (e) UV–Vis absorbance spectroscopy, and (f) (αhv)2hv plot, which is a common method for determining the bandgap energy of semiconductors.
Materials 17 03625 g002
Figure 3. (a) Band diagram illustrating the interface between the CQDs filler and the FPE matrix, (b) electrostatic potential diagrams and corresponding energy band values for FPE, (c) cyclic voltammetry (CV) curves of the CQDs, revealing their electrochemical properties.
Figure 3. (a) Band diagram illustrating the interface between the CQDs filler and the FPE matrix, (b) electrostatic potential diagrams and corresponding energy band values for FPE, (c) cyclic voltammetry (CV) curves of the CQDs, revealing their electrochemical properties.
Materials 17 03625 g003
Figure 4. (a) Frequency dependence and (b) temperature dependence of dielectric constant and dielectric loss of nanocomposites. Changing law of dielectric constant and dielectric loss of nanocomposites with filler content at room temperature (c) and 150 °C (d). The green arrow serves to denote the coordinate axis pertinent to the data, thereby enhancing the clarity of the visual representation.
Figure 4. (a) Frequency dependence and (b) temperature dependence of dielectric constant and dielectric loss of nanocomposites. Changing law of dielectric constant and dielectric loss of nanocomposites with filler content at room temperature (c) and 150 °C (d). The green arrow serves to denote the coordinate axis pertinent to the data, thereby enhancing the clarity of the visual representation.
Materials 17 03625 g004
Figure 5. Weibull distribution of breakdown electric field at RT (a) and 150 °C (b), and characteristic breakdown electric field strength at RT (c) and 150 °C (d).
Figure 5. Weibull distribution of breakdown electric field at RT (a) and 150 °C (b), and characteristic breakdown electric field strength at RT (c) and 150 °C (d).
Materials 17 03625 g005
Figure 6. (a) Leakage current density at an electric field strength of 200MV/m and (b) conductivity variation with temperature of composites. (c) Load-displacement curve and (d) Young’s modulus as a function of filler content of the nanocomposites.
Figure 6. (a) Leakage current density at an electric field strength of 200MV/m and (b) conductivity variation with temperature of composites. (c) Load-displacement curve and (d) Young’s modulus as a function of filler content of the nanocomposites.
Materials 17 03625 g006
Figure 7. The energy density and efficiency of pure FPE and CQD/FPE nanocomposites at (a) RT and (b) 150 °C. (c) The energy density above 90% efficiency, and (d) the maximum energy density. The green arrow serves to denote the coordinate axis pertinent to the data, thereby enhancing the clarity of the visual representation.
Figure 7. The energy density and efficiency of pure FPE and CQD/FPE nanocomposites at (a) RT and (b) 150 °C. (c) The energy density above 90% efficiency, and (d) the maximum energy density. The green arrow serves to denote the coordinate axis pertinent to the data, thereby enhancing the clarity of the visual representation.
Materials 17 03625 g007
Figure 8. (a) Charge–discharge cycle performance of pure FPE and composites. (b) Overdamped discharge curves and (c) time dependence of Wd under different electric fields of the composites with 0.5 wt% CQDs. (d) Time dependence of Wd under an electric field of 200 MV/m 200 MV/m of commercial BOPP and the composites with 0.5 wt% CQDs. The blue circle serves to denote the coordinate axis pertinent to the data, thereby enhancing the clarity of the visual representation.
Figure 8. (a) Charge–discharge cycle performance of pure FPE and composites. (b) Overdamped discharge curves and (c) time dependence of Wd under different electric fields of the composites with 0.5 wt% CQDs. (d) Time dependence of Wd under an electric field of 200 MV/m 200 MV/m of commercial BOPP and the composites with 0.5 wt% CQDs. The blue circle serves to denote the coordinate axis pertinent to the data, thereby enhancing the clarity of the visual representation.
Materials 17 03625 g008
Table 1. Na Content Analysis: Sodium Ion Concentration in Pre- and Post-Dialysis Solutions and External Dialysis Bag Solution.
Table 1. Na Content Analysis: Sodium Ion Concentration in Pre- and Post-Dialysis Solutions and External Dialysis Bag Solution.
SamplesDialysis Time (h)Na Content (μg/mL)
Undialyzed stock solution01.056
Solution outside dialysis bag90.474
Solution outside dialysis bag180.535
Solution outside dialysis bag300.506
Solution outside dialysis bag360.501
Dialyzed 36 h stock solution361.034
Table 2. Comparison of energy storage properties with the other composites at 150 °C [3,11,27,46,47,48,49,50].
Table 2. Comparison of energy storage properties with the other composites at 150 °C [3,11,27,46,47,48,49,50].
MatrixFillerEb (MV/m)Ue (J/cm3)η (%)References
PEIBNNS@ST4604.29>80Compos. Pt. A-Appl. Sci. Manuf. 2023, 175, 9, 107791 [11]
PCAu5184.84>90Mater. Chem. A 2022, 10, 18773 [27]
PEIZIF 8–675642.96>90Small 2023, 19, 12 [3]
PEIBNNPs5804.2>90Mater. Chem. C 2022, 10, 13157 [46]
PEIBN/BaTiO35075.23>90Sci. China-Mater. 2023, 66, 2652 [47]
PEIPLZST@Al2O348910.283.5Mater. Chem. A 2023, 11, 7227 [48]
PIMgO4502.689Polymers 2022, 14, 10, 2918 [49]
PIBNNS4892.58>90Adv. Compos. Hybrid Mater. 2022, 5, 238 [50]
FPECQDs5964.53>90This Work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Luo, H.; Liu, Y.; Wang, F.; Peng, B.; Li, X.; Hu, D.; He, G.; Zhang, D. Improved Energy Density at High Temperatures of FPE Dielectrics by Extreme Low Loading of CQDs. Materials 2024, 17, 3625. https://doi.org/10.3390/ma17143625

AMA Style

Wang H, Luo H, Liu Y, Wang F, Peng B, Li X, Hu D, He G, Zhang D. Improved Energy Density at High Temperatures of FPE Dielectrics by Extreme Low Loading of CQDs. Materials. 2024; 17(14):3625. https://doi.org/10.3390/ma17143625

Chicago/Turabian Style

Wang, Huan, Hang Luo, Yuan Liu, Fan Wang, Bo Peng, Xiaona Li, Deng Hu, Guanghu He, and Dou Zhang. 2024. "Improved Energy Density at High Temperatures of FPE Dielectrics by Extreme Low Loading of CQDs" Materials 17, no. 14: 3625. https://doi.org/10.3390/ma17143625

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop