Introduction

High permittivity materials have received significant attention in recent years due to their potential for application in high energy density capacitors; these are a key technology for the development of portable electronic devices, stationary power systems and high power microwave systems1,2,3,4,5. Ferroelectric ceramics such as lead zirconate titanate (PZT) and barium titanate (BaTiO3) are the preferred choice for application in capacitors because of their high permittivity6,7,8. In general, the energy density (Ue) of dielectric materials is defined as the integral Ue = ∫ EdD, where E is the electric field and D is the electric displacement9,10. Therefore, a high breakdown strength is a key factor to obtain a high energy density, and high permittivity ferroelectric ceramics often exhibit a relatively low dielectric strength and high dielectric loss which limits the energy density11,12.

In recent decades, an number of efforts have been made to fabricate ceramics/polymer nanocomposites to overcome this limitation13,14,15. The nanocomposite systems combine the high permittivity of ceramics with the high breakdown strength, low dielectric loss, and mechanical flexibility of polymers. To achieve a high energy density, researchers have developed a variety of approaches to improve the dielectric properties of nanocomposites. For example, (i) the use of high permittivity polymer matrixes such as polyvinylidene fluoride (PVDF) and its copolymers to reduce the contrast of dielectric properties between ceramic nanoparticles and polymer matrixes16,17,18 and (ii) chemical modification on the surface of ceramic nanoparticles to enhance nanoparticle dispersion in the polymer matrix and improve interfacial adhesion between matrix and particle19,20,21.

In addition, recent studies have demonstrated that the morphology and geometry of ceramic nanoparticles influence the dielectric properties of nanocomposites22,23,24. For 0–3 type ceramic-polymer nanocomposites, where spherical zero-dimensional ceramic nanoparticles are embedded in a three-dimensionally connected polymer matrix, a high volume fraction (>50 vol%) of nanoparticles is necessary to achieve a high permittivity nanocomposite. However, this can lead to a reduced breakdown strength and low mechanical flexibility25. For 1–3 type ceramic-polymer nanocomposites, where one-dimensional ceramic nanoparticles are dispersed in a three-dimensionally connected polymer matrix, a number of studies have indicated that high aspect ratio ceramic fillers can improve the dielectric properties and energy density of nanocomposites more efficiently compared to spherical ceramic fillers26,27,28. Andrews et al. developed a micromechanics approach and finite element models to study the effect of ceramic filler aspect ratio on the electro-elastic properties of nanocomposites29. The results showed the electromechanical coupling can increase up to 60 times compared to its initial values when the aspect ratio was increased from 1 to 10 at 30 vol % of ceramic filler. Tang et al. demonstrated that nanocomposites based on high aspect ratio PZT nanowires exhibited an increased energy density, which was 77.8% higher than lower aspect ratio PZT nanowires30.

In this study, we have prepared P(VDF-HFP) nanocomposites with a range of aspect ratio BaTiO3 nanofibers (NFs) synthesized by a two-step hydrothermal method. There have been a variety of reported methods for synthesizing BaTiO3 NFs, such as the two-step hydrothermal method, one-step hydrothermal method, topochemical solid-state reaction and anodic aluminum oxide template method25,31,32,33. The two-step hydrothermal method has attracted attention due to its superiority in terms of the synthesis of single crystalline nanofibers, morphology control, homogeneity at the molecular level, low temperature processing and simple experimental approach. The effects of aspect ratio and volume fraction of the BaTiO3 NFs on the dielectric constant (relative permittivity), breakdown strength and energy storage density of the nanocomposites were investigated systematically. On increasing the BaTiO3 NF volume fraction or aspect ratio, the dielectric constant and maximum electric displacement of the nanocomposites increased monotonically while the breakdown strength decreased monotonically. Under the same electric field, the nanocomposites with higher volume fraction or higher aspect ratio BaTiO3 NFs possessed higher discharged energy density. The maximal energy storage density reached 15.48 J/cm3 in nanocomposites containing 7.5 vol% BaTiO3 NFs synthesized at 210 °C for 2 h under the electric field of 300 kV/mm.

Results and Discussion

In order to obtain BaTiO3 nanofibers (BT NFs), an initial hydrothermal reaction was used to synthesize Na2Ti3O7 nanofibers (NT NFs) due to its extensive research history and easily controlled nanofiber morphology25. Figure 1a shows the morphology of hydrothermally synthesized Na2Ti3O7 nanofibers (NT NFs) which exhibited a high aspect ratio and favorable dispersibility. Figure 1b shows the XRD pattern of NT NFs indexed to monoclinic Na2Ti3O7 (PDF card NO. 31–1329). The corresponding TEM images were shown in Fig. 1c. The NT NFs possessed a smooth surface with a diameter of approximately 100 nm. The clear lattice fringes shown in Fig. 1c demonstrate that the NT NFs were single-crystalline. The parallel lattice spacings were approximately 0.19 nm and 0.28 nm, which correspond to the (020) and (003) planes, respectively, and reveal that the NT NFs grew in the [010] direction. Figure 1d shows the SEM image of H2Ti3O7 nanofibers (HT NFs), which retained the morphology of the Na2Ti3O7 nanofibers.

Figure 1
figure 1

(a) SEM image, (b) XRD pattern and (c) TEM images of hydrothermally synthesized Na2Ti3O7 nanofibers (NT NFs). (d) SEM image of H2Ti3O7 NFs.

The HT NFs were transformed to BaTiO3 NFs by a second hydrothermal reaction. The H2Ti3O7 phase is a layered titanate, which is a good precursor for soft chemical synthesis because of its open structure that enables ion exchange and topochemical transformation. Figure 2a,b shows schematic diagrams of the crystalline structures of H2Ti3O7 and BaTiO3, respectively. During the second hydrothermal reaction, the Ba2+ ions diffuse into the host lattice of H2Ti3O7 by ion exchange with H+, which leads to a rearrangement of the octahedra sharing and structural transformation to perovskite BaTiO3. The Ba2+ ions possess a higher positive charge compared with H+, thereby promoting the structural transformation. In addition, it is known that the edge sharing octahedra were driven into vertex sharing octahedra during the topochemical transformation34. Therefore, the HT NFs were successfully transformed into BaTiO3 NFs.

Figure 2
figure 2

Schematic diagrams of the crystalline structures of (a) H2Ti3O7 and (b) BaTiO3.

Figure 3a–c show the sizes and morphologies of BaTiO3 NFs hydrothermally synthesized at 210 °C for 2–12 h, respectively. The corresponding XRD patterns are shown in Fig. 3d. The diffraction patterns indicate that tetragonal BaTiO3 (PDF card NO. 75–0462) without any impurity phase can be obtained at 210 °C with a reaction time ranging from 2 h to 12 h and the crystallization of BaTiO3 was improved at longer reaction times. When the reaction time was 2 h, the products were a mass of nanoparticles and few nanofibers, as shown in Fig. 3a. The amount of nanofibers increased significantly when reaction time increased to 6 h. Furthermore as the reaction time increased to 12 h, the nanofibers became dominant. Figure 3e shows a TEM image of BaTiO3 NFs synthesized at 210 °C for 12 h. It can be observed that the surface of BaTiO3 NFs was rough and a HRTEM image of a typical BaTiO3 NF synthesized at 210 °C for 12 h is shown in Fig. 3f. The clear lattice fringes illustrated that the BT NF were single-crystalline. The parallel lattice spacing were about 0.406 nm and 0.286 nm corresponding to (001) and (101) planes of tetragonal BaTiO3 respectively, which revealed that the NFs grew in the [001] direction. The corresponding selected area diffraction pattern (SADP) also exhibited the characteristics of a single crystal.

Figure 3
figure 3

SEM images of BaTiO3 NFs hydrothermally synthesized at 210 °C for (a) 2 h, (b) 6 h, (c) 12 h. (d) XRD patterns of BaTiO3 NFs. (e) TEM images of BaTiO3 NFs synthesized at 210 °C for 12 h. (f) HRTEM image of a typical BaTiO3 NF and corresponding SADP.

It was worth noting that with an increase of reaction time, the length of the BaTiO3 NFs increased to a much greater extent compared to the diameter, leading to an increase in the fibre aspect ratio. The length and diameter of the BaTiO3 NFs synthesized at 210 °C for 2 h, 6 h and 12 h were analyzed using SEM images by ImageJ software, as shown in Fig. 4. The aspect ratio of the BaTiO3 NFs calculated from Fig. 4 were 3.5, 7.4 and 21.0 for the reaction time of 2 h, 6 h and 12 h, respectively and clearly shows that the aspect ratio of the BaTiO3 NFs increased with an increase of reaction time.

Figure 4: BaTiO3 NFs length and diameter distribution for different reaction times.
figure 4

(a,b) 210 °C. 2 h. (c,d) 210 °C. 6 h. (e,f) 210 °C, 12 h.

To improve the compatibility and dispersibility of BaTiO3 NFs in the P(VDF-HFP) matrix, the BaTiO3 NFs were surface functionalized by dopamine. As highlighted in the introduction, the breakdown strength of the nanocomposite is equally important as the dielectric constant with regard to the energy density and the breakdown strength tends to decrease with increasing volume fraction of ceramic filler. In order to improve the dielectric properties of nanocomposites and maintain a high breakdown strength, the nanocomposites were fabricated at a relatively low volume fraction of BaTiO3 NFs in this study, varying from 2.5% to 7.5%. The upper-surface morphology of the BaTiO3 NFs/P(VDF-HFP) nanocomposites are shown in Fig. 5. Figure 5a–c shows SEM images of the nanocomposites with 5.0 vol% of BaTiO3 NFs synthesized at 210 °C for 2 h, 6 h and 12 h, respectively where it can be seen that the aspect ratio of the BaTiO3 NFs in nanocomposites gradually increased. Figure 5d–f shows the SEM images of the nanocomposites with 2.5 vol%, 5.0 vol% and 7.5 vol% BaTiO3 NFs synthesized at 210 °C for 12 h. The visibility of the BaTiO3 NFs in nanocomposite became more pronounced with increasing BaTiO3 NFs volume fraction. It can be observed that the BaTiO3 NFs exhibited good compatibility and dispersibility in the P(VDF-HFP) matrix and the nanocomposites exhibited limited defects as the volume fraction of BaTiO3 NFs approached 7.5%. This can be attributed to that the fact that the BaTiO3 NFs were surface modified by dopamine and therefore formed a strong adhesive bonding force with the polymer matrix.

Figure 5
figure 5

(ac) SEM images of the nanocomposites with 5 vol% low aspect ratio BaTiO3 NFs synthesized at 210 °C for 2 h, 6 h and 12 h. (df) SEM images of the nanocomposites with 2.5 vol%, 5.0 vol% and 7.5 vol% high aspect ratio BaTiO3 NFs synthesized at 210 °C for 12 h.

Figure 6 shows the variation of the dielectric constant and loss of the nanocomposites with the BaTiO3 NFs aspect ratio and volume fraction for a frequency range of 1 kHz to 10 MHz. The dielectric constant of the nancomposites increased with the increase of volume fraction of BaTiO3 NFs since BaTiO3 possess a higher dielectric constant compared with pure P(VDF-HFP) polymer. At the measurement frequency of 1 kHz, the dielectric constant of the nanocomposite with 7.5 vol% BaTiO3 NFs synthesized at 210 °C for 12 h (aspect ratio 21.0) reached 23.4 while the dielectric constant of the samples with 7.5 vol% BaTiO3 NFs synthesized at 210 °C for 6 h (aspect ratio 7.4) and 2 h (aspect ratio 3.5) was 17.8 and 14 respectively. This demonstrates that on increasing the aspect ratio of the BaTiO3 NFs, the dielectric constant of the nancomposites was significantly improved. There are a number of previous reports and theoretical models that have demonstrated the increased dielectric constant as a result of using high aspect ratio ceramic fillers14,25,26. For instance, the Maxwell-Garnet model can efficiently describe the effect of aspect ratio, as shown in equation 135,36.

Figure 6: Dependency of dielectric constant and loss of the nanocomposites on the aspect ratio (AR) and volume fraction of BT NFs with frequency ranging from 1 kHz to 10 MHz.
figure 6

(a,d) 210 °C 2 h; AR = 3.5, (b,e) 210 °C 6 h; AR = 7.4, (c,f) 210 °C 12 h; AR = 21.0.

where Ni is known as the depolarization factor of ellipsoids in the x, y, z direction. For BaTiO3 NFs, where the radii ax > ay = az. Ni can be expressed as equation 2

The Maxwell-Garnet model indicates that the dielectric constant of the nanocomposites will increase on increasing the aspect ratio of BaTiO3 NFs.

However, there was no apparent variation about the dielectric loss on increasing the aspect ratio of BaTiO3 NFs. For example, the dielectric loss of the nanocomposites with 7.5 vol% BaTiO3 NFs synthesized at 210 °C for 2 h, 6 h and 12 h was 0.032, 0.037 and 0.046 at 1 kHz, respectively. The dielectric loss remained less than 0.06 (<100 kHz) for all samples which was attributed to the relatively low volume fraction of BaTiO3 NFs and their good compatibility and dispersion in the polymer matrix. The dielectric properties of nanocomposites were efficiently improved by increasing the aspect ratio of the BaTiO3 NFs, without the need for additional fillers.

Figure 7a–c show typical electric displacement-electric field (D-E) loops of nanocomposites where the volume fraction of BaTiO3 NFs ranged from 2.5% to 7.5% and the aspect ratio of the BaTiO3 NFs were 3.5, 7.4 and 21.0 respectively. As can be seen from Fig. 6a, the maximum electric displacement (Dmax) increased monotonically with an increase of the volume fraction of BaTiO3 NFs. A similar trend can also be observed in Fig. 6b and c. Figure 6d summarizes the Dmax of nanocomposites with 7.5 vol% for various aspect ratios of BaTiO3 NFs and the Dmax was enhanced with increasing the electric field. Figure 7d also clearly demonstrated that with increasing the aspect ratio of BaTiO3 NFs, the Dmax of nanocomposites improved significantly. For instance, the Dmax of the nanocomposites with 7.5 vol% and higher aspect ratio (AR) BaTiO3 NFs (AR = 21) reached 5.4 μC/cm2 while the value was only 3.9 μC/cm2 for lower aspect ratio BaTiO3 NFs (AR = 3.5) under an electric field of 160 kV/mm. It has been demonstrated in Fig. 6 that the dielectric constant of nanocomposites increases on increasing the aspect ratio of the BT NFs. Since D = ε0εrE, where ε0 and εr are free space and the relative of the permittivity of the nanocomposites respectively, then under the same electric field (E) an increase in Dmax can result from an increase of dielectric constant of the nanocomposites on increasing the volume fraction or aspect ratio of the BaTiO3 NFs.

Figure 7
figure 7

The D-E loops of nanocomposites where the volume fraction of BaTiO3 NFs ranged from 2.5% to 7.5% and the aspect ratio (AR) of BaTiO3 NFs was (a) 210 °C 2 h; AR = 3.5, (b) 210 °C 6 h; AR = 7.4, and (c) 210 °C 12 h; AR = 21.0. (d) The Dmax of the nanocomposites with electric field for 7.5 vol% BT NFs with various aspect ratios.

The electric breakdown strength of the nanocomposites was analyzed by a two parameter Weibull cumulative probability function, as shown in equation 3,

where P(E) is the cumulative probability of electric failure, E is the experimental breakdown strength, EB is a scale parameter that refers to the breakdown strength at the cumulative failure probability of 63.2% which is regarded as the characteristic breakdown strength, and β is the Weibull modulus associated with the linear regressive fit of the distribution. Figure 8a–c shows the electric breakdown strength of nanocomposites with different volume fraction and aspect ratio BaTiO3 NFs. The characteristic breakdown strength (EB) is summarized in Fig. 7d for the different volume fractions and aspect ratios. All samples possessed a relatively high breakdown strength over 180 kV/mm and on increasing the volume fraction of BaTiO3 NFs, the EB of the nanocomposites decreased; see Fig. 8d. The EB of the nanocomposites with 2.5 vol% BaTiO3 NFs and aspect ratio of 21 was 268.7 kV/mm, which reduced to 257.4 kV/mm and 188.4 kV/mm when the volume fraction of BaTiO3 NFs increased to 5.0% and 7.5%, respectively. The introduction of more BaTiO3 NFs into the polymer matrix resulted in defects and electric field concentrations, which increased the risk of failure and decreased the breakdown strength. The nature of the field concentrations with respect to volume fraction and aspect ratio are discussed later. It also can be demonstrated in Fig. 8d that the EB of the nanocomposites decreased with an increase of the aspect ratio of the BaTiO3 NFs.

Figure 8: The electric breakdown strength of nanocomposites with different volume fraction and aspect ratio (AR) BaTiO3 NFs.
figure 8

(a) 210 °C 2 h; AR = 3.5, (b) 210 °C 6 h; AR = 7.4, (c,f) 210 °C 12 h; AR = 21.0.

The energy storage density of the nanocomposites were calculated from D-E loops according to the equation 4,

Figure 9a presents the energy storage density of the nanocomposites with low aspect ratio BaTiO3 NFs synthesized at 210 °C for 2 h as a function of electric field and volume fraction of BaTiO3 NFs. This indicates that the energy storage density increased with an increase of electric field. Moreover, the nanocomposites with the higher volume fraction of BaTiO3 NFs possessed the higher energy storage density under the same electric field. Figure 9b shows the energy storage density of the nanocomposites with 7.5 vol% BaTiO3 NFs as a function of electric field and aspect ratio of the BaTiO3 NFs. Similarly, the nanocomposites with the higher aspect ratio BaTiO3 NFs exhibited a higher energy storage density under the same electric field. The higher energy storage density can be attributed to the significantly enhanced dielectric constant and Dmax of the nanocomposites on increasing the volume fraction and aspect ratio of the BaTiO3 NFs. Figure 9b also indicates that the electric breakdown strength of the nanocomposites decreased with increasing aspect ratio of BaTiO3 NFs. Under an electric field of 300 kV/mm, the maximal discharged energy density of 15.48 J/cm3 was obtained in the nanocomposites with 7.5 vol% on low aspect ratio BaTiO3 NFs synthesized at 210 °C for 2 h.

Figure 9
figure 9

(a) Energy storage density of the nanocomposites with BaTiO3 NFs synthesized at 210 °C for 2 h (aspect ratio 3.5) as a function of electric field and volume fraction. (b) Energy storage density of the nanocomposites with 7.5 vol% BaTiO3 NFs as a function of electric field and aspect ratio (AR): 210 °C 2 h (AR = 3.5), 210 °C 6 h (AR = 7.4), 210 °C 12 h (AR = 21.0).

To provide a deeper understanding of the decrease in breakdown strength and increase in dielectric constant with an increase in BaTiO3 NF volume fraction and aspect ratio, finite element modelling (Ansys APDL v15.0) was employed. A two dimensional electrostatic analysis of a single high permittivity BaTiO3 NF (relative permittivity, εr = 1500) embedded in a low permittivity P(VDF-HFP) matrix (εr = 7) was used to investigate the effect of fiber aspect ratio and angle (with respect to applied field direction) on the maximum localized electric field. BaTiO3 composites in the range of the experimentally produced aspect ratios (3.5 to 20.1) and volume fractions (2.5 vol.% to 7.5 vol.%) were analyzed.

The inclusion of a high permittivity BaTiO3 inclusion leads to the electric field concentrating in the low permittivity matrix, and such field concentrations are likely to be sites for initiating dielectric breakdown within these nanocomposites. This can be seen in Fig. 10a, which is a close up image of the field distribution around a BaTiO3 fiber with an aspect ratio of 3.5 at 2.5 vol.%. The blue contours are low field regions and red contours indicate high field concentrations at the tips of the fiber. High aspect ratio fibers aligned almost perpendicular to the applied field direction (angle > 85°) produced slightly lower electric field concentrations compared to lower aspect ratios, however at angles below 85° the high aspect ratio fibers result in local electric fields that are significantly higher than the applied field, see Fig. 10b. The manufactured composites had a random orientation of fibers within the film (Fig. 5) and therefore a nominal fiber orientation of 45° to the applied field was used to demonstrate the effect of BaTiO3 aspect ratio and volume fraction on the relative permittivity (dielectric constant) and field concentration, where an increase in local electric field leads to a reduced breakdown strength. The relative permittivity (Fig. 10c) and localized field concentrations (Fig. 10d) were found to increase with an increase in both fiber aspect ratio and volume fraction. These results agree well with the experimental observations that on increasing the aspect ratio and volume fraction of the BaTiO3 NFs, the dielectric constant of the nanocomposites was improved (Fig. 6a–c) while the breakdown strength of the nanocomposites decreased (Fig. 8d).

Figure 10
figure 10

(a) Contour plot of finite element analysis for BaTiO3 fiber (2.5 vol.%) in P(VDF-HFP) matrix with aspect ratio 3.5 orientated at 30° to the applied field, where blue and red contours represent regions of high and low field, respectively; (b) effect of fiber angle and aspect ratio on maximum local field concentration (plotted as Elocal/Eapplied); (c,d) show the effect of BaTiO3 fiber aspect ratio and volume fraction at a 45° angle to applied field on dielectric constant and maximum local field concentration, respectively; (e) contour plot of two BaTiO3 NFs aligned in field direction, showing field concentration between particles; (f) effect of separation distance on field concentration within P(VDF-HFP) matrix.

High BaTiO3 NF aspect ratios are also likely to reduce the average separation distance between high permittivity inclusions for the same volume fraction. A finite element model of two BaTiO3 particles (aspect ratio of unity) within a P(VDF-HFP) matrix was used to demonstrate the effect of separation distance on electric field concentrations within the matrix. The ‘worst-case’ in terms of field concentration was occurred when two particles were aligned in the direction of the field, as in Fig. 10e. The ratio of the maximum local field to the applied field is shown in Fig. 10f as a function of the distance between the high permittivity particles, with the electric field concentration following an inverse separation law. These result indicate that a combination of electric field concentrations due to the presence of high permittivity fibers and fiber separation distance influence the permittivity and dielectric strength and the high aspect ratio fibers lead to a lower breakdown strength, as observed experimentally (Fig. 8d).

Conclusion

In this study, BaTiO3 nanofibers (BT NF) with a variety of aspect ratios were synthesized by a two-step hydrothermal method. The effects of the aspect ratio and volume fraction of the BT NF on dielectric properties and energy storage densities of the P(VDF-HFP) based one-dimensional nanocomposites were investigated and modeled in detail. As the aspect ratio and volume fraction of the BaTiO3 NFs was increased, the dielectric constant and Dmax of the nanocomposites were both increased monotonically while the breakdown strength decreased. The nanocomposites with highest aspect ratio and volume fraction of BT NFs exhibited the highest energy storage density under the same electric field. The maximal energy storage density reached 15.48 J/cm3 in the nanocomposites with 7.5 vol% BT NFs synthesized at 210 °C for 2 h under the electric field of 300 kV/mm. This work provides a potential new route to prepare and tailor the properties of novel high energy density capacitor nanocomposites.

Methods

Synthesis of BaTiO3 nanofibers

The BaTiO3 nanofibers (BT NFs) were synthesized by a two-step hydrothermal method. Firstly, sodium titanate nanofibers (Na2Ti3O7 NFs, NT NFs) were synthesized. A 1.446 g mass of titanium oxide (TiO2, Anatase) was added to 70 ml NaOH solution (10 M) and the mixture was stirred for 2 h to form a homogeneous suspension. Hydrothermal reactions were carried out at 210 °C under an auto-generated pressure for 24 h in a 100 ml Teflon-lined autoclave. The products were washed by distilled water and then soaked in diluted 0.2 M hydrochloric acid (HCl, 37%) for 4 hours to obtain hydrogen titanate nanofibers (H2Ti3O7 NFs). The BT NFs were synthesized by a second hydrothermal reaction where 0.150 g of H2Ti3O7 NFs were dispersed in 70 ml Ba(OH)2∙8H2O solution and the mixture was sonicated for 10 min. The hydrothermal reactions were carried out at 210 °C under an auto-generated pressure for 2–24 h in a 100 ml Teflon-lined autoclave to obtain BT NFs at a range of aspect ratios. The products were soaked in 0.2 M HCl solution briefly, then washed using distilled water several times and dried at 80 °C in an oven.

Fabrication of BT NFs/P(VDF-HFP) nanocomposite

The BT NFs were mixed with a solution of P(VDF-HFP) in N,N-dimethylformamide (DMF) by stirring and sonicating to form a homogeneous suspension. The suspension was then cast onto a clean glass and dried at 80 °C for 12 h under vacuum. The dried nanocomposite sheets were then compressed into films at 200 °C under a pressure of approximately 15 MPa. Gold electrodes were sputtered on both sides of the film using a mask with 2 mm diameter eyelets.

Characterization

The crystalline phases of the NFs were evaluated by X-ray diffraction (XRD, D/max 2550, Japan) with Cu-Kα radiation (λ = 1.5406 Å) at room temperature. The size and morphology of the NFs and the nanocomposites were observed using a scanning electron microscope (SEM, Nova NanoSEM230, USA). High-resolution transmission electron microscopy (HRTEM) images of the NFs were taken with a Titan G2 60–300 instrument, using an accelerating voltage of 300 kV. The frequency-dependent dielectric constant and dielectric loss were measured using an Agilent 4294 A LCR meter with frequency ranging from 1 kHz to 10 MHz. The electric displacement-electric field loops and leakage current were measured by a precision Premier II ferroelectric polarization tester (Radiant, Inc.) at room temperature and 10 Hz.

Additional Information

How to cite this article: Zhang, D. et al. Significantly Enhanced Energy Storage Density by Modulating the Aspect Ratio of BaTiO3 Nanofibers. Sci. Rep. 7, 45179; doi: 10.1038/srep45179 (2017).

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.