Next Article in Journal
Preparation and Properties of Chitosan/Graphene Modified Bamboo Fiber Fabrics
Previous Article in Journal
Preparation of a Novel Fracturing Fluid System with Excellent Elasticity and Low Friction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Dielectric Performance of P(VDF-HFP) Composites with Satellite–Core-Structured Fe2O3@BaTiO3 Nanofillers

School of Materials Science and Engineering, State Key Lab Silicon Mat, Zhejiang University, Hangzhou 310027, China
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(10), 1541; https://doi.org/10.3390/polym11101541
Submission received: 27 July 2019 / Revised: 26 August 2019 / Accepted: 19 September 2019 / Published: 21 September 2019

Abstract

:
Polymer dielectric materials are extensively used in electronic devices. To enhance the dielectric constant, ceramic fillers with high dielectric constant have been widely introduced into polymer matrices. However, to obtain high permittivity, a large added amount (>50 vol%) is usually needed. With the aim of improving dielectric properties with low filler content, satellite–core-structured Fe2O3@BaTiO3 (Fe2O3@BT) nanoparticles were fabricated as fillers for a poly(vinylidene fluoride-co-hexafluoropropylene) (P(VDF-HFP)) matrix. The interfacial polarization effect is increased by Fe2O3 nanoparticles, and thus, composite permittivity is enhanced. Besides, the satellite–core structure prevents Fe2O3 particles from directly contacting each other, so that the dielectric loss remains relatively low. Typically, with 20 vol% Fe2O3@BT nanoparticle fillers, the permittivity of the composite is 31.7 (1 kHz), nearly 1.8 and 3.0 times that of 20 vol% BT composites and pure polymers, respectively. Nanocomposites also achieve high breakdown strength (>150 KV/mm) and low loss tangent (~0.05). Moreover, the composites exhibited excellent flexibility and maintained good dielectric properties after bending. These results demonstrate that composite films possess broad application prospects in flexible electronics.

Graphical Abstract

1. Introduction

Polymer dielectric materials are extensively applied in flexible electronics and energy storage devices owing to their merits of outstanding flexibility, ease of processing, light weight, and low cost [1,2,3,4]. Despite these merits, the dielectric permittivity (εr) of most polymers is quite low (<10). Two main strategies have been developed by researchers to enhance the dielectric permittivity [5,6,7,8,9,10,11,12]. One is incorporating ceramic fillers with intrinsically high dielectric constants (e.g., BaTiO3, BaxSr1-xTiO3, CaCu3Ti4O12) [13,14,15,16,17,18,19,20] into the polymer matrix; the other strategy is employing conductive fillers, including metals (e.g., Ag, Ni, Al) [21,22,23,24,25], carbon materials (e.g., carbon nanotubes, graphene) [26,27,28,29,30,31,32], semiconductors (e.g., ZnO) [33], and conductive polymers (e.g., polyaniline (PANI)) [34,35,36,37]. With ceramic/polymer composites, the merits of high εr from ceramic fillers and high breakdown strength from polymers are combined. However, ceramic fillers with a large amount of additions (>50 vol%) are usually needed to obtain a high εr, which can seriously affect flexibility and mechanical properties. Therefore, it is worth studying this problem to further improve the dielectric properties of composites with low filler content.
One type of n-type semiconductor is α-Fe2O3 (band gap: 2.1 eV). It has been studied extensively in pigments, lithium-ion batteries, gas sensors, and photoelectrochemical water splitting [38,39,40]. It has also attracted the attention of researchers for optimizing the dielectric performance of pure polymers by utilizing α-Fe2O3. Thakur et al. [41] reported in-situ-synthesized Fe2O3 in poly(vinylidene fluoride) (PVDF), which promotes the formation of β-PVDF and enhances εr through interfacial polarization. Hayashida [42] incorporated α-Fe2O3 into ten kinds of polymer matrices and studied its influences on the dielectric properties at 40–160 °C. The research showed that the εr of composites could be raised due to interfacial polarization induced by free electrons in α-Fe2O3 particles. In addition, constructing satellite–core-structured fillers for the polymer matrix was considered to be an effective approach for enhancing the dielectric performance. This structure combines two kinds of fillers by loading one filler onto the surface of another. For example, Ag@BT fillers [43] and SnO2@BT [44] fillers were fabricated by former researchers and enhanced dielectric properties were obtained compared with pristine BT fillers.
In this work, with the aim of improving dielectric properties with low filler content, Fe2O3@BT nanoparticles were fabricated as fillers to prepare Fe2O3@BT/P(VDF-HFP) and (FB/P(VDF-HFP)) composites. Satellite–core-structured Fe2O3@BT introduces extra interfaces, so the interfacial polarization and εr of composites are enhanced. Besides, the satellite–core structure of Fe2O3@BT prevents the direct contact of Fe2O3 particles with each other in the polymer matrix, so the loss tangent remains relatively low.

2. Materials and Methods

2.1. Materials

N, N-dimethylformamide (DMF) and Barium titanate (BaTiO3, BT) were bought from Aladdin (Shanghai, China). P(VDF-HFP) and Ferric nitrate nonahydrate were supplied by Sinopharm (Shanghai, China) and Sigma-Aldrich ( Shanghai, China), respectively.

2.2. Synthesis of Satellite-Core-Structured Fe2O3@BT Nanoparticles

The 0.303 g ferric nitrate nonahydrate was first dissolved in deionized water (100 mL). Then, 0.700 g BT nanoparticles were dispersed into this solution via sonicating and stirring. The molar ratio of BT/Fe was 4:1. The solution was stirred at 75 °C for 5 h, and cleaned with deionized water. After drying under vacuum, FeOOH@BT nanoparticles were obtained. The generated powder was heated at 550 °C for 2 h in air. Satellite–core-structured Fe2O3@BT nanoparticles were then generated.

2.3. Fabrication of Fe2O3@BT/P(VDF-HFP) Composites

A stoichiometric amount of Fe2O3@BT nanoparticles were distributed into dimethylformamide (DMF) via stirring and ultrasound. P(VDF-HFP) was then added and vigorously stirred for 12 h. The feeding ratio of P(VDF-HFP)/DMF was 1 g:15 mL. The composite films were then prepared through drop casting onto clean glass plates. The composites were kept at 60 °C to eliminate DMF, and then heated to 200 °C (5 min) and quenched in ice water. BT/P(VDF-HFP) and pure polymer were also generated.

2.4. Characterization

Scanning electron microscopy (SEM) (SU-8010, Hitachi, Japan) and transmission electron microscopy (TEM) using a Tecnai G2 F20 (FEI, Hillsboro, OR, USA) (accelerating voltage: 200 kV) with energy dispersive spectroscopy (EDS) were applied to examine the morphology of composites and particles. The elemental composition of nanoparticles was observed using X-ray photoelectron spectroscopy (XPS) with an Escalab 250Xi. XRD (X’ Pert PRO, PANalytical, Netherlands) using Cu Kα radiation was performed to identify the components of particles and composites. Differential scanning calorimetry (DSC) was tested by TA-Q200 at 90–190 °C (10 °C/min, nitrogen atmosphere). Dielectric performances were measured with an 4294 impedance analyzer (Agilent, Palo Alto, CA, USA) from 102–106 Hz (silver electrode, diameter: 4 mm, thickness: 100 nm). A dielectric strength tester (CS2674AX, Nanjing Changsheng, Nanjing, China) was employed to test the Direct Current (DC) breakdown strength under a direct current voltage ramp of 200 V s−1 at 25 °C.

3. Results and Discussion

3.1. Morphology and Structure of Fe2O3@BT Nanoparticles

Figure 1 presents the TEM photos of Fe2O3@BT nanoparticles, as well as the EDS elemental mapping photos. The pure BT nanoparticles are spherical, with a diameter of about 50–100 nm. Fe2O3 nanoparticles (5–10 nm) decorated on BT and the satellite–core-structured Fe2O3@BT nanoparticles are formed. As shown in the High Resolution Transmission Electron Microscope (HRTEM) image, the lattice fringe areas with 0.221 nm and 0.282 nm spacing are assigned to (113) and (110) planes of α-Fe2O3 and BT (JCPDS 75-0462, 33-0664), respectively [45,46]. The structure of Fe2O3@BT nanoparticles is illustrated in Figure 1d. EDS results further reveal the distribution of Fe2O3. It is shown that Ba, Ti, and O are homogenously distributed on the surface of nanoparticles. However, the amount of Fe is much less and its distribution is locally concentrated, corresponding to the satellite–core structure.
Figure 2 presents XRD patterns of BT and hybrid particles. Characteristic peaks of BaTiO3 (Joint Committee on Powder Diffraction Standards (JCPDS) 75-0462) are obviously shown in hybrid particles. Moreover, some weak peaks at 24.1°, 33.2°, 35.6°, 49.5°, and 54.1° are also observed, corresponding to the (012), (104), (110), (024), and (116) planes of α-Fe2O3, respectively. No other phases of Fe2O3 are shown, which indicates that only α-Fe2O3 is obtained after calcination at 550 °C [47,48,49].
To further analyze the elemental composition, XPS is conducted on Fe2O3@BT nanoparticles. As shown in Figure 3a, characteristic peaks of Ba, O, Fe, C, and Ti are shown in survey scan spectra. In Figure 3b, the peak at 724.6 eV and 710.9 eV correspond to Fe3+ 2p1/2 and Fe3+ 2p3/2 peaks, together with two satellite peaks at 733.5 eV and 719.2 eV. The binding energy difference between 2p1/2 and 2p3/2 is 13.7 eV. Besides, characteristic peaks are not observed for Fe2+ [50,51,52,53]. These results indicate that the element Fe in nanoparticles exists in the form of Fe3+, which means Fe2O3 is synthesized. In addition, the color of the powders is red-brown, which is consistent with that of Fe2O3.

3.2. Structure and Morphology of Fe2O3@BT/P(VDF-HFP) Composites

Figure 4 presents cross-section morphologies of composites. Numerous nanoparticle fillers are shown in the polymer. According to the XRD results of the composites, it can be seen that these nanoparticles are Fe2O3@BT. The nanoparticles are distributed well in P(VDF-HFP) and no apparent void or pore can be observed. In addition, the inset shows the digital photograph of 20 vol% composites, which can still be easily bent and rolled.
Figure 5 demonstrates XRD patterns of composites. The three peaks at 18.2°, 19.9°, and 26.5° correspond to the (020), (110), and (021) planes of α-P(VDF-HFP), respectively [54,55]. The hybrid nanofillers peaks can be observed, as well as the matrix peaks. The relative intensity of the matrix peaks decreases as the Fe2O3@BT increases.

3.3. Melting and Crystallization Behavior of Fe2O3@BT/P(VDF-HFP) Composites

Differential scanning calorimetry (DSC) was performed to analyze the crystallization of the polymer. As is shown in Figure 6a, a melting peak appears in the heating curve for each film, corresponding to the melting process of the polymer. The melting temperature (Tm) and crystallization temperature (Tc) decrease as the filler increases. The crystallinity (χc) can be calculated through the formula below:
χ c = Δ H m ( 1 ω ) × Δ H m 0 × 100 %
where ΔHm and Δ H m 0 represent the melting enthalpy of samples and 100% crystallized α-P(VDF-HFP) (93.07 J/g), respectively. Here, ω is the weight fraction of Fe2O3@BT nanoparticles in composites.
As shown in Table 1, when the filler content increases, the crystallization peak moves towards lower temperatures and Tc decreases gradually. This phenomenon is mainly attributed to the hindering effect of nanoparticle fillers [56,57]. During the crystallization process, the Fe2O3@BT nanoparticles retard the movement of the polymer chain and impede the progress of crystallization, leading to the decrease of Tc. Fe2O3@BT can also act as a heterogeneous nucleation site, facilitating the crystallization. However, the hinderance effect dominates the crystallization process and the influences of heterogeneous nucleation are covered up. When more Fe2O3@BT nanoparticles are added, the hinderance effect is further enhanced and Tc continues to decrease. The final χc also reduces gradually because of the accumulation of the hinderance effect during crystallization.

3.4. Dielectric Properties of Fe2O3@BT/P(VDF-HFP) Composites

Figure 7 presents the dielectric characteristics of a pristine polymer, FB/P(VDF-HFP), and 20 vol% BT/P(VDF-HFP). In Figure 7a, the εr of each composite decreases as the frequency gets higher. This phenomenon is due to the interfacial polarization relaxation and dipole polarization relaxation at low and high frequencies. To further analyze the influences of Fe2O3@BT nanoparticles on the dielectric performance of composites, εr and tan δ values at 1 kHz of all samples are compared in Figure 8 (left axis). As the content of nanoparticles increases, the εr of FB/P(VDF-HFP) is increased notably. The enhancement is larger than in BT/P(VDF-HFP) at the same concentration, which is caused by the interfacial polarization induced by Fe2O3@BT particles, an important polarization mechanism that occurs in low frequency ranges because of its relatively long time of establishment. When a dielectric is placed in an electric field, the internal free electrons and holes migrate under the electric field and gather at the interfacial area containing two phases, impurities, and defects. Then, dipole moments are generated, and thus, interfacial polarization is induced. In FB/P(VDF-HFP) composites, the satellite–core-structured Fe2O3@BT nanoparticles introduce extra interfaces, including the Fe2O3/BT interface, Fe2O3/P(VDF-HFP) interface, and BT/P(VDF-HFP) interface; semi-conductive Fe2O3 brings about more charge carriers. Therefore, the interfacial polarization is enhanced by Fe2O3@BT nanoparticles and the dielectric permittivity of composites is raised. Figure S1 exhibits the dielectric performances of BT/P(VDF-HFP) composites. With 20 vol% nanoparticles added, the εr value of Fe2O3@BT/P(VDF-HFP) is 31.7 at 1 kHz, nearly 1.8 and 3.0 times that of 20 vol% BT/P(VDF-HFP) (18.0) and pure polymer (10.6), respectively. Figure S2 shows that the composite maintains good dielectric performances after bending, which proves the potential application in flexible electronics.
Figure 7b shows the dielectric loss of composites. Tan δ declines at first and then increases for each sample as the frequency gets higher. The increase of tan δ is attributed to dipole polarization relaxation at high frequency. In this range, the establishment of dipole polarization cannot follow the electric field, so the relaxation leads to enhanced loss. The tan δ of composites is lower than the pristine polymer and it continues to decrease when the filler content increases. This phenomenon probably occurs because the Fe2O3@BT nanoparticles retard the movement of polymer chains, which can decrease the dipole polarization relaxation loss [58,59]. The loss tangent is derived from electric conduction loss and interfacial polarization relaxation at low frequencies. The tan δ values of all samples at 1 kHz are also compared in Figure 8 (right axis). With the increase of nanofiller content, the tan δ is slightly increased, because Fe2O3 generates many charge carriers. However, the satellite–core structure of Fe2O3@BT could prevent the direct contact of Fe2O3 particles with each other in the polymer matrix and suppress the long-range movement of charge carriers; therefore, the tan δ remains low (<0.06). With the addition of 20 vol% Fe2O3@BT nanoparticles, the tan δ of composites maintains a rather low value of 0.05. The tan δ values of BT/P(VDF-HFP) (20 vol%) and pure polymer are 0.03 and 0.02, respectively (Figure S1). And compared with other BT-based/polymer nanocomposites reported in the previous literature (Table S1), the results of the FB/P(VDF-HFP) nanocomposites reported herein are comparable or better. Figure 7c shows that the conductivity of composites increases when more nanofillers are added. Nevertheless, the conductivity of all composites is lower than 2 × 10−8 S/m, proving that the film provides good insulation.
Breakdown strength (Eb) is also a significant characteristic and determines the energy density and work voltage of composites. Due to the randomness of breakdown events, measured data of Eb is usually further processed by a two-parameter Weibull distribution function [60,61]:
P = 1 exp [ ( E E 0 ) β ]
where P is the cumulative probability of electrical failure, E represents breakdown strength, E0 is the characteristic breakdown strength (cumulative failure probability: 0.632), and β is the shape parameter. As shown in Figure 9, breakdown strength decreases as the nanofiller content increases. This phenomenon results from the electrical mismatch between the polymer and the nanoparticles. However, the satellite–core structure of Fe2O3@BT nanoparticles suppresses the rise of dielectric loss and impedes the formation of conductive paths, so Eb still remains at a relatively high level. The Eb value of the 20 vol% Fe2O3@BT-filled composite is 152.7 MV/m.

4. Conclusions

Satellite–core-structured Fe2O3@BT nanoparticles were fabricated as fillers to prepare FB/P(VDF-HFP) composites. Fe2O3@BT nanoparticles show a hinderance effect on the crystallization process of polymers and the crystallization temperature and crystallinity of composite films both decrease as the content of the filler increases. The interfacial polarization effect is enhanced by Fe2O3 nanoparticles, and thus, the dielectric permittivity of composites is enhanced. The satellite–core structure prevents Fe2O3 particles from directly contacting each other, so the dielectric loss remains low. With the addition of 20 vol% Fe2O3@BT nanoparticles, the permittivity value of the composite is 31.7 at 1 kHz, nearly 1.8 and 3.0 times that of the 20 vol% BT and pristine polymer, respectively. Nanocomposites also demonstrate low loss tangent (~0.05) and high breakdown strength (>150 KV/mm). In addition, the composites also exhibit excellent flexibility and maintains good dielectric performances after bending.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/11/10/1541/s1. Figure S1: Frequency dependence dielectric performances of BT/P(VDF-HFP) composites. Figure S2: Digital photograph of 20 vol% FB/P(VDF-HFP) composites in (a) unbent and (b) bent statuses. (c) Comparison of dielectric properties of the original 20 vol% FB/P(VDF-HFP) composites and bent composites for 1000 cycles. Table S1: Some research studies related to BaTiO3 in PVDF-based composites.

Author Contributions

Conceptualization, Q.Z.; data curation, Y.J., Z.Z. (Zhao Zhang), and Z.Z. (Zheng Zhou); formal analysis, Y.J., Z.Z. (Zhao Zhang), and Zheng Zhou; funding acquisition, Q.Z.; investigation, Y.J., Z.Z. (Zhao Zhang), Z.Z. (Zheng Zhou), H.Y., and Q.Z.; methodology, Y.J., Z.Z. (Zhao Zhang), and Z.Z. (Zheng Zhou); supervision, Q.Z. and H.Y.; writing, original draft, Y.J.; writing, review and editing, Q.Z. and H.Y.

Funding

The authors gratefully acknowledge the financial support from the National Natural Science Foundation of China (Grant No. 51772267), and the Key R&D Program of Zhejiang Province (Grant No. 2019C05001)

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, H.; Liu, F.; Fan, B.; Ai, D.; Peng, Z.; Wang, Q. Nanostructured ferroelectric-polymer composites for capacitive energy storage. Small Methods 2018, 2, 1700399. [Google Scholar] [CrossRef]
  2. Dang, Z.M.; Yuan, J.K.; Yao, S.H.; Liao, R.J. Flexible nanodielectric materials with high permittivity for power energy storage. Adv. Mater. 2013, 25, 6334–6365. [Google Scholar] [CrossRef] [PubMed]
  3. Zhu, L.; Wang, Q. Novel ferroelectric polymers for high energy density and low loss dielectrics. Macromolecules 2012, 45, 2937–2954. [Google Scholar] [CrossRef]
  4. Yao, Z.; Song, Z.; Hao, H.; Yu, Z.; Cao, M.; Zhang, S.; Lanagan, M.T.; Liu, H. Homogeneous/inhomogeneous-structured dielectrics and their energy-storage performances. Adv. Mater. 2017, 29, 1601727. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, Q.; Shen, Y.; Zhang, S.; Zhang, Q.M. Polymer-based dielectrics with high energy storage density. Annu. Rev. Mater. Res. 2015, 45, 433–458. [Google Scholar] [CrossRef]
  6. Huang, X.; Sun, B.; Zhu, Y.; Li, S.; Jiang, P. High-k polymer nanocomposites with 1D filler for dielectric and energy storage applications. Prog. Mater. Sci. 2019, 100, 187–225. [Google Scholar] [CrossRef]
  7. Dang, Z.M.; Zheng, M.S.; Zha, J.W. 1D/2D carbon nanomaterial-polymer dielectric composites with high permittivity for power energy storage applications. Small 2016, 12, 1688–1701. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, X.; Jiang, J.; Shen, Z.; Dan, Z.; Li, M.; Lin, Y.; Nan, C.W.; Chen, L.; Shen, Y. Polymer nanocomposites with ultrahigh energy density and high discharge efficiency by modulating their nanostructures in three dimensions. Adv. Mater. 2018, 30, 1707269. [Google Scholar] [CrossRef] [PubMed]
  9. Thakur, V.K.; Gupta, R.K. Recent progress on ferroelectric polymer-based nanocomposites for high energy density capacitors: Synthesis, dielectric properties, and future aspects. Chem. Rev. 2016, 116, 4260–4317. [Google Scholar] [CrossRef]
  10. Lin, M.F.; Thakur, V.K.; Tan, E.J.; Lee, P.S. Surface functionalization of BaTiO3 nanoparticles and improved electrical properties of batio3/polyvinylidene fluoride composite. RSC Adv. 2011, 1, 827–836. [Google Scholar] [CrossRef]
  11. Thakur, V.K.; Lin, M.F.; Tan, E.J.; Lee, P.S. Green aqueous modification of fluoropolymers for energy storage applications. J. Mater. Chem. 2012, 22, 5951–5959. [Google Scholar] [CrossRef]
  12. Thakur, V.K.; Tan, E.J.; Lin, M.F.; Lee, P.S. Poly(vinylidene fluoride)-graft-poly(2-hydroxyethyl methacrylate): A novel material for high energy density capacitors. J. Mater. Chem. 2011, 21, 3751–3759. [Google Scholar] [CrossRef]
  13. Yu, D.; Xu, N.; Hu, L.; Zhang, Q.; Yang, H. Nanocomposites with BaTiO3-SrTiO3 hybrid fillers exhibiting enhanced dielectric behaviours and energy-storage densities. J. Mater. Chem. C 2015, 3, 4016–4022. [Google Scholar] [CrossRef]
  14. Luo, S.; Shen, Y.; Yu, S.; Wan, Y.; Liao, W.-H.; Sun, R.; Wong, C.-P. Construction of a 3D-BaTiO3 network leading to significantly enhanced dielectric permittivity and energy storage density of polymer composites. Energy Environ. Sci. 2017, 10, 137–144. [Google Scholar] [CrossRef]
  15. Wang, Z.; Wang, T.; Wang, C.; Xiao, Y.; Jing, P.; Cui, Y.; Pu, Y. Poly(vinylidene fluoride) flexible nanocomposite films with dopamine-coated giant dielectric ceramic nanopowders, Ba(Fe0.5Ta0.5)O3, for high energy-storage density at low electric field. ACS Appl. Mater. Interfaces 2017, 9, 29130–29139. [Google Scholar] [CrossRef] [PubMed]
  16. Xie, Y.; Jiang, W.; Fu, T.; Liu, J.; Zhang, Z.; Wang, S. Achieving High Energy Density and Low Loss in PVDF/BST Nanodielectrics with Enhanced Structural Homogeneity. ACS Appl. Mater. Interfaces 2018, 10, 29038–29047. [Google Scholar] [CrossRef]
  17. Wan, W.; Luo, J.; Huang, C.-E.; Yang, J.; Feng, Y.; Yuan, W.-X.; Ouyang, Y.; Chen, D.; Qiu, T. Calcium copper titanate/polyurethane composite films with high dielectric constant, low dielectric loss and super flexibility. Ceram. Int. 2018, 44, 5086–5092. [Google Scholar] [CrossRef]
  18. Liu, S.H.; Xiu, S.M.; Shen, B.; Zhai, J.W.; Kong, L.B. Dielectric properties and energy storage densities of poly(vinylidenefluoride) nanocomposite with surface hydroxylated cube shaped Ba0.6Sr0.4TiO3 nanoparticles. Polymers 2016, 8, 45. [Google Scholar] [CrossRef]
  19. Chen, C.; Wang, L.; Liu, X.; Yang, W.; Lin, J.; Chen, G.; Yang, X. K0.5Na0.5NbO3-SrTiO3/PVDF polymer composite film with low remnant polarization and high discharge energy storage density. Polymers 2019, 11, 310. [Google Scholar] [CrossRef]
  20. Yang, J.; Tang, Z.; Yin, H.; Liu, Y.; Wang, L.; Tang, H.; Li, Y. Poly(arylene ether nitrile) composites with surface-hydroxylated calcium copper titanate particles for high-temperature-resistant dielectric applications. Polymers 2019, 11, 766. [Google Scholar] [CrossRef]
  21. Huang, X.; Jiang, P.; Kim, C.; Ke, Q.; Wang, G. Preparation, microstructure and properties of polyethylene aluminum nanocomposite dielectrics. Compos. Sci. Technol. 2008, 68, 2134–2140. [Google Scholar] [CrossRef]
  22. Zhou, Y.; Wang, H.; Xiang, F.; Zhang, H.; Yu, K.; Chen, L. A poly(vinylidene fluoride) composite with added self-passivated microaluminum and nanoaluminum particles for enhanced thermal conductivity. Appl. Phys. Lett. 2011, 98, 182906. [Google Scholar] [CrossRef] [Green Version]
  23. Zhou, Y.; Wang, H. An Al@Al2O3@SiO2/polyimide composite with multilayer coating structure fillers based on self-passivated aluminum cores. Appl. Phys. Lett. 2013, 102, 132901. [Google Scholar] [CrossRef]
  24. Dang, Z.; Lin, Y.; Nan, C. Novel ferroelectric polymer composites with high dielectric constants. Adv. Mater. 2003, 15, 1625–1629. [Google Scholar] [CrossRef]
  25. Chen, X.; Liang, F.; Lu, W.; Jin, Z.; Zhao, Y.; Fu, M. High permittivity nanocomposites embedded with Ag/TiO2 core-shell nanoparticles modified by phosphonic acid. Polymers 2018, 10, 586. [Google Scholar] [CrossRef]
  26. Yang, C.; Hao, S.; Dai, S.; Zhang, X. Nanocomposites of poly(vinylidene fluoride)—Controllable hydroxylated/carboxylated graphene with enhanced dielectric performance for large energy density capacitor. Carbon 2017, 117, 301–312. [Google Scholar] [CrossRef]
  27. Iqbal, A.; Lee, S.H.; Siddiqi, H.M.; Park, O.O.; Akhter, T. Enhanced dielectric constant, ultralow dielectric loss, and high-strength imide-functionalized graphene oxide/hyperbranched polyimide nanocomposites. J. Phys. Chem. C 2018, 122, 6555–6565. [Google Scholar] [CrossRef]
  28. Liu, Y.; Zhang, C.; Huang, B.; Wang, X.; Li, Y.; Wang, Z.; Lai, W.; Zhang, X.; Liu, X. Skin–core structured fluorinated MWCNTs: A nanofiller towards a broadband dielectric material with a high dielectric constant and low dielectric loss. J. Mater. Chem. C 2018, 6, 2370–2378. [Google Scholar] [CrossRef]
  29. Li, W.; Song, Z.; Qian, J.; Tan, Z.; Chu, H.; Wu, X.; Nie, W.; Ran, X. Enhancing conjugation degree and interfacial interactions to enhance dielectric properties of noncovalent functionalized graphene/poly(vinylidene fluoride) composites. Carbon 2019, 141, 728–738. [Google Scholar] [CrossRef]
  30. Wang, L.; Dang, Z.-M. Carbon nanotube composites with high dielectric constant at low percolation threshold. Appl. Phys. Lett. 2005, 87, 042903. [Google Scholar] [CrossRef]
  31. He, F.; Lau, S.; Chan, H.L.; Fan, J. High dielectric permittivity and low percolation threshold in nanocomposites based on poly(vinylidene fluoride) and exfoliated graphite nanoplates. Adv. Mater. 2009, 21, 710–715. [Google Scholar] [CrossRef]
  32. Wang, Y.; Xing, C.; Guan, J.; Li, Y. Towards flexible dielectric materials with high dielectric constant and low loss: PVDF nanocomposites with both homogenously dispersed CNTs and ionic liquids nanodomains. Polymers 2017, 9, 562. [Google Scholar] [CrossRef] [PubMed]
  33. Wu, W.; Huang, X.; Li, S.; Jiang, P.; Toshikatsu, T. Novel three-dimensional zinc oxide superstructures for high dielectric constant polymer composites capable of withstanding high electric field. J. Phys. Chem. C 2012, 116, 24887–24895. [Google Scholar] [CrossRef]
  34. Zhang, L.; Lu, X.; Zhang, X.; Jin, L.; Xu, Z.; Cheng, Z.Y. All-organic dielectric nanocomposites using conducting polypyrrole nanoclips as filler. Compos. Sci. Technol. 2018, 167, 285–293. [Google Scholar] [CrossRef]
  35. Zhang, Q.M.; Li, M.; Poh, M.; Xia, F.; Cheng, Z.Y.; Xu, H.; Huang, C. An all-organic composite actuator material with a high dielectric constant. Nature 2002, 419, 284–287. [Google Scholar] [CrossRef] [PubMed]
  36. Huang, C.; Zhang, Q. Enhanced dielectric and electromechanical responses in high dielectric constant all-polymer percolative composites. Adv. Funct. Mater. 2004, 14, 501–506. [Google Scholar] [CrossRef]
  37. Yuan, J.-K.; Dang, Z.-M.; Yao, S.-H.; Zha, J.-W.; Zhou, T.; Li, S.-T.; Bai, J. Fabrication and dielectric properties of advanced high permittivity polyaniline/poly(vinylidene fluoride) nanohybrid films with high energy storage density. J. Mater. Chem. 2010, 20, 2441–2447. [Google Scholar] [CrossRef]
  38. Wang, Z.; Luan, D.; Madhavi, S.; Hu, Y.; Lou, X.W. Assembling carbon-coated α-Fe2O3 hollow nanohorns on the CNT backbone for superior lithium storage capability. Energy Environ. Sci. 2012, 5, 5252–5256. [Google Scholar] [CrossRef]
  39. Sivula, K.; Le Formal, F.; Gratzel, M. Solar water splitting: Progress using hematite (alpha-Fe2O3) photoelectrodes. ChemSusChem 2011, 4, 432–449. [Google Scholar] [CrossRef]
  40. Kevin, S.; Radek, Z.; Florian, L.; Rosa, R.; Anke, W.; Jiri, T.; Jiri, F.; Michael, G. Photoelectrochemical water splitting with mesoporous hematite prepared by a solution-based colloidal approach. J. Am. Chem. Soc. 2010, 132, 7436–7444. [Google Scholar] [CrossRef]
  41. Thakur, P.; Kool, A.; Bagchi, B.; Das, S.; Nandy, P. Effect of in situ synthesized Fe2O3 and Co3O4 nanoparticles on electroactive beta phase crystallization and dielectric properties of poly(vinylidene fluoride) thin films. Phys. Chem. Chem. Phys. 2015, 17, 1368–1378. [Google Scholar] [CrossRef] [PubMed]
  42. Hayashida, K. Highly improved dielectric properties of polymer/α-Fe2O3 composites at elevated temperatures. RSC Adv. 2016, 6, 64871–64878. [Google Scholar] [CrossRef]
  43. Luo, S.; Yu, S.; Sun, R.; Wong, C.P. Nano Ag-deposited BaTiO3 hybrid particles as fillers for polymeric dielectric composites: Toward high dielectric constant and suppressed loss. ACS Appl. Mater. Interfaces 2014, 6, 176–182. [Google Scholar] [CrossRef] [PubMed]
  44. Zha, J.W.; Meng, X.; Wang, D.; Dang, Z.M.; Li, R.K.Y. Dielectric properties of poly(vinylidene fluoride) nanocomposites filled with surface coated BaTiO3 by SnO2 nanodots. Appl. Phys. Lett. 2014, 104, 072906. [Google Scholar] [CrossRef]
  45. Lu, X.; Zeng, Y.; Yu, M.; Zhai, T.; Liang, C.; Xie, S.; Balogun, M.S.; Tong, Y. Oxygen-deficient hematite nanorods as high-performance and novel negative electrodes for flexible asymmetric supercapacitors. Adv. Mater. 2014, 26, 3148–3155. [Google Scholar] [CrossRef]
  46. Lian, J.; Duan, X.; Ma, J.; Peng, P.; Kim, T.; Zheng, W. Hematite (alpha-Fe2O3) with various morphologies: Ionic liquid-assisted synthesis, formation mechanism, and properties. ACS Nano 2009, 3, 3749–3761. [Google Scholar] [CrossRef]
  47. Jeong, J.M.; Choi, B.G.; Lee, S.C.; Lee, K.G.; Chang, S.J.; Han, Y.K.; Lee, Y.B.; Lee, H.U.; Kwon, S.; Lee, G.; et al. Hierarchical hollow spheres of Fe2O3@polyaniline for lithium ion battery anodes. Adv. Mater. 2013, 25, 6250–6255. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, X.; Niu, Y.; Li, Y.; Hou, X.; Wang, Y.; Bai, R.; Zhao, J. Synthesis, optical and magnetic properties of α-Fe2O3 nanoparticles with various shapes. Mater. Lett. 2013, 99, 111–114. [Google Scholar] [CrossRef]
  49. Zhang, P.; Yu, L.; Lou, X.W.D. Construction of Heterostructured Fe2O3-TiO2 Microdumbbells for Photoelectrochemical Water Oxidation. Angew. Chem. Int. Ed. 2018, 57, 15076–15080. [Google Scholar] [CrossRef]
  50. Yamashita, T.; Hayes, P. Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide materials. Appl. Surf. Sci. 2008, 254, 2441–2449. [Google Scholar] [CrossRef]
  51. Bhargava, G.; Gouzman, I.; Chun, C.M.; Ramanarayanan, T.A.; Bernasek, S.L. Characterization of the “native” surface thin film on pure polycrystalline iron: A high resolution XPS and TEM study. Appl. Surf. Sci. 2007, 253, 4322–4329. [Google Scholar] [CrossRef]
  52. Tahir, A.A.; Wijayantha, K.G.U.; Saremi-Yarahmadi, S.; Mazhar, M.; McKee, V. Nanostructured α-Fe2O3 thin films for photoelectrochemical hydrogen generation. Chem. Mater. 2009, 21, 3763–3772. [Google Scholar] [CrossRef]
  53. Desai, J.D.; Pathan, H.M.; Min, S.-K.; Jung, K.-D.; Joo, O.S. FT-IR, XPS and PEC characterization of spray deposited hematite thin films. Appl. Surf. Sci. 2005, 252, 1870–1875. [Google Scholar] [CrossRef]
  54. Martins, P.; Lopes, A.C.; Lanceros-Mendez, S. Electroactive phases of poly(vinylidene fluoride): Determination, processing and applications. Prog. Polym. Sci. 2014, 39, 683–706. [Google Scholar] [CrossRef]
  55. Huang, X.; Jiang, P.; Kim, C.; Liu, F.; Yin, Y. Influence of aspect ratio of carbon nanotubes on crystalline phases and dielectric properties of poly(vinylidene fluoride). Eur. Polym. J. 2009, 45, 377–386. [Google Scholar] [CrossRef]
  56. Xu, N.; Hu, L.; Zhang, Q.; Xiao, X.; Yang, H.; Yu, E. Significantly enhanced dielectric performance of poly(vinylidene fluoride-co-hexafluoropylene)-based composites filled with hierarchical flower-like TiO2 particles. ACS Appl. Mater. Interfaces 2015, 7, 27373–27381. [Google Scholar] [CrossRef]
  57. Yang, L.; Qiu, J.; Ji, H.; Zhu, K.; Wang, J. Enhanced dielectric and ferroelectric properties induced by TiO2@MWCNTs nanoparticles in flexible poly(vinylidene fluoride) composites. Compos. Part A 2014, 65, 125–134. [Google Scholar] [CrossRef]
  58. Zhu, M.; Huang, X.; Yang, K.; Zhai, X.; Zhang, J.; He, J.; Jiang, P. Energy storage in ferroelectric polymer nanocomposites filled with core-shell structured polymer@BaTiO3 nanoparticles: Understanding the role of polymer shells in the interfacial regions. ACS Appl. Mater. Interfaces 2014, 6, 19644–19654. [Google Scholar] [CrossRef]
  59. Xie, L.; Huang, X.; Yang, K.; Li, S.; Jiang, P. “Grafting to” route to PVDF-HFP-GMA/BaTiO3 nanocomposites with high dielectric constant and high thermal conductivity for energy storage and thermal management applications. J. Mater. Chem. A 2014, 2, 5244–5251. [Google Scholar] [CrossRef]
  60. Xie, L.; Huang, X.; Wu, C.; Jiang, P. Core-shell structured poly(methyl methacrylate)/BaTiO3 nanocomposites prepared by in situ atom transfer radical polymerization: A route to high dielectric constant materials with the inherent low loss of the base polymer. J. Mater. Chem. 2011, 21, 5897–5906. [Google Scholar] [CrossRef]
  61. Li, Y.; Huang, X.; Hu, Z.; Jiang, P.; Li, S.; Tanaka, T. Large dielectric constant and high thermal conductivity in poly(vinylidene fluoride)/barium titanate/silicon carbide three-phase nanocomposites. ACS Appl. Mater. Interfaces 2011, 3, 4396–4403. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Transmission electron microscopy (TEM) photo of a BT nanoparticle. (b) TEM and (c) HRTEM photos of a Fe2O3@BT nanoparticle. (c) Partially enlarged image of the blue square area in image (b). (d) Schematic illustration of a Fe2O3@BT nanoparticle. (ei) HAADF-STEM image with mapping images of a Fe2O3@BT nanoparticle. The scale of the images (fi) is the same with that of image (e).
Figure 1. (a) Transmission electron microscopy (TEM) photo of a BT nanoparticle. (b) TEM and (c) HRTEM photos of a Fe2O3@BT nanoparticle. (c) Partially enlarged image of the blue square area in image (b). (d) Schematic illustration of a Fe2O3@BT nanoparticle. (ei) HAADF-STEM image with mapping images of a Fe2O3@BT nanoparticle. The scale of the images (fi) is the same with that of image (e).
Polymers 11 01541 g001
Figure 2. X-ray diffraction (XRD) patterns of Fe2O3@BT and BT nanoparticles.
Figure 2. X-ray diffraction (XRD) patterns of Fe2O3@BT and BT nanoparticles.
Polymers 11 01541 g002
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra of Fe2O3@BT nanoparticles: (a) survey scan, (b) Fe 2p.
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra of Fe2O3@BT nanoparticles: (a) survey scan, (b) Fe 2p.
Polymers 11 01541 g003
Figure 4. SEM images of cross-sectional film of: (a) pristine P(VDF-HFP), (b) 5 vol%, (c) 10 vol%, (d) 15 vol%, and (e) 20 vol% composites. Inset is the digital photograph of 20 vol% film.
Figure 4. SEM images of cross-sectional film of: (a) pristine P(VDF-HFP), (b) 5 vol%, (c) 10 vol%, (d) 15 vol%, and (e) 20 vol% composites. Inset is the digital photograph of 20 vol% film.
Polymers 11 01541 g004
Figure 5. XRD patterns of P(VDF-HFP), Fe2O3@BT nanoparticles, and FB/P(VDF-HFP) composites.
Figure 5. XRD patterns of P(VDF-HFP), Fe2O3@BT nanoparticles, and FB/P(VDF-HFP) composites.
Polymers 11 01541 g005
Figure 6. DSC of polymer and FB/P(VDF-HFP) composite (a) heating curves and (b) cooling curves.
Figure 6. DSC of polymer and FB/P(VDF-HFP) composite (a) heating curves and (b) cooling curves.
Polymers 11 01541 g006
Figure 7. Frequency dependence of (a) dielectric constant, (b) dielectric loss tangent, and (c) conductivity of pristine polymer, BT/P(VDF-HFP), and FB/P(VDF-HFP) composites.
Figure 7. Frequency dependence of (a) dielectric constant, (b) dielectric loss tangent, and (c) conductivity of pristine polymer, BT/P(VDF-HFP), and FB/P(VDF-HFP) composites.
Polymers 11 01541 g007
Figure 8. Dielectric properties of the composites filled with Fe2O3@BT and BT as a function of filler content at 1 kHz.
Figure 8. Dielectric properties of the composites filled with Fe2O3@BT and BT as a function of filler content at 1 kHz.
Polymers 11 01541 g008
Figure 9. (a) Weibull distribution for breakdown strength and (b) characteristic breakdown strength of composites.
Figure 9. (a) Weibull distribution for breakdown strength and (b) characteristic breakdown strength of composites.
Polymers 11 01541 g009
Table 1. Melting Temperature (Tm), crystallization temperature (Tc), and crystallinity (χc) of polymer and FB/P(VDF-HFP) composites.
Table 1. Melting Temperature (Tm), crystallization temperature (Tc), and crystallinity (χc) of polymer and FB/P(VDF-HFP) composites.
SampleP(VDF-HFP)5 vol% Fe2O3@BT10 vol% Fe2O3@BT15 vol% Fe2O3@BT20 vol% Fe2O3@BT
Tm (°C)155.5155.0153.9153.7153.0
Tc (°C)127.0126.6126.3125.4124.7
χc (%)37.335.433.433.230.7

Share and Cite

MDPI and ACS Style

Jiang, Y.; Zhang, Z.; Zhou, Z.; Yang, H.; Zhang, Q. Enhanced Dielectric Performance of P(VDF-HFP) Composites with Satellite–Core-Structured Fe2O3@BaTiO3 Nanofillers. Polymers 2019, 11, 1541. https://doi.org/10.3390/polym11101541

AMA Style

Jiang Y, Zhang Z, Zhou Z, Yang H, Zhang Q. Enhanced Dielectric Performance of P(VDF-HFP) Composites with Satellite–Core-Structured Fe2O3@BaTiO3 Nanofillers. Polymers. 2019; 11(10):1541. https://doi.org/10.3390/polym11101541

Chicago/Turabian Style

Jiang, Yongchang, Zhao Zhang, Zheng Zhou, Hui Yang, and Qilong Zhang. 2019. "Enhanced Dielectric Performance of P(VDF-HFP) Composites with Satellite–Core-Structured Fe2O3@BaTiO3 Nanofillers" Polymers 11, no. 10: 1541. https://doi.org/10.3390/polym11101541

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