1. Introduction
Polyvinylidene fluoride (PVDF) is a commercially available polymer in various forms, including sheet form. While the melting temperature is 170 °C, its glass transition temperature is around −40 °C, which makes the material very flexible in room temperature conditions. In homopolymer form, the primary chains in the polymer have alternating structures of CH
2 and CF
2 groups. During polymerization, defect groups form in the polymer so that the monomer units are reversed in some regions, influencing the crystalline structure and relating to the piezoelectric properties observed in these materials [
1]. Creating a piezoelectric material (poling) from a PVDF film consists of stretching the film at an elevated temperature and subjecting it to an electric field. This process leaves the film with orthotropic mechanical and electro-mechanical properties. A thorough understanding of the in-plane properties of this material relative to the electric field is vital for the design of the metamaterial cells. Limited research has studied the orthotropic properties of the PVDF film.
PVDF is a versatile polymer with applications in metamaterials for sensors, actuators, and energy harvesting devices [
2,
3]. PVDF-based materials are used in electromagnetic interference shielding, wearable biosensors, and acoustic sensors [
4,
5]. The polymer’s flexibility and biocompatibility enable its use in smart scaffolds and biomedical applications [
6]. Research on polyvinylidene fluoride (PVDF) has focused on characterizing its complex electromechanical properties and developing constitutive models. Studies have measured the elastic, dielectric, and piezoelectric constants of poled PVDF films, revealing their frequency dependence and orthorhombic symmetry [
7,
8]. Nonlinear elastic behavior has been observed and modeled using the Mooney–Rivlin and neo-Hookean approaches [
9]. Time-dependent viscoelastic properties have been incorporated into constitutive models [
10,
11,
12]. Molecular modeling has provided insights into PVDF’s piezoelectric effects and switching phenomena [
13,
14]. Experimental studies have demonstrated nonlinear and time-dependent electromechanical behavior, including pre-stress dependence of electromechanical coefficients [
15]. Several groups have characterized various properties of poled PVDF films. Seminara [
16] and Vinogradov [
17] have developed experimental setups and procedures for the electromechanical and mechanical characterization of these films.
Jones [
1] has reviewed the degradation and performance of PVDF under various stress environments for possible application to space mirror concepts. The environmental effects in low-orbit space environments were investigated, and the significant resiliency of the material in these environments was shown. Lee [
18] conducted nanoscale characterization using atomic force microscopy of PVDF under stress to measure nanoscale properties and show that the behavior was scale-independent under stress. Wang [
19] reviewed the effects of various nanofillers on the piezoelectric response and energy-harvesting properties of PVDF composite films showing that nanofillers can be added without a reduction in the flexibility of the material. Fortunato [
20] and Dodds [
21] have investigated the piezoelectric properties of PVDF composite films with enhanced piezoelectric response, and PVDF-TrFE thin films enhanced with ZnO nanoparticles, respectively. Fortunato increased the piezoelectric coefficient of PVDF, avoiding the poling process, by inducing an increased β-phase fraction in the PVDF film by adding suitable quantities of nanofillers. Dodds et al., on the other hand, showed that piezoelectric ZnO nanoparticles can enhance the piezoelectric behavior without impacting the flexibility of the PVDF films.
PVDF films have also been extensively used in medical fields, such as smart auxetic stents, biomedical scaffolds, and health monitoring devices. Islam et al. [
22] developed a smart auxetic stent, based on a perforated PVDF membrane that features ultrasonic powering, blood flow sensing, and integrated wireless electronics for accessible and continuous post-EVAR (endovascular aneurysm repair) surveillance. Recently, Pan et al. [
23] proposed a piezoelectric smart stent for electricity generation driven by blood pressure fluctuation. It was fabricated by fused deposition modeling 3D-printing with a built-in electric field, where the stents are made from composite wires of potassium sodium niobate (KNN) particles and polyvinylidene fluoride-co-hexafluoropropylene (PVDF-HFP) matrix. A comprehensive review of PVDF biomedical scaffolds for tissue engineering is provided in Mokhtari et al. [
24] and Ribeiro et al. [
25]. In particular, PVDF polymer-based piezoelectric scaffolds that mimic tissue’s electrical microenvironment have been fabricated for bone tissue regeneration [
26] and neural tissue engineering [
27]. Moreover, in recent years, numerous PVDF-based piezoelectric generators have been developed for healthcare monitoring, such as human sound recording [
28], arterial pulse monitoring [
29], muscle behavior monitoring [
30], and intracochlear sound pressure measurement [
31].
2. Design and Experimental Methods
Measuring the piezoelectric constants of PVDF is crucial because these constants quantify the material’s ability to convert mechanical stress into electrical charge and vice versa. Verifying these constants is essential for applying these models to sensors, actuators, or energy-harvesting devices. In poled PVDF, the alignment of molecular dipoles enhances its piezoelectric response. Still, its mechanical properties, such as stiffness and plasticity, strongly influence this behavior, especially in details with high-stress concentrations such as hinges. Thus, the piezoelectric performance of PVDF is often directionally dependent due to its anisotropic nature. A good understanding of mechanical properties (e.g., Young’s modulus, Poisson’s ratio) and piezoelectric constants allows for precise prediction of how the material will deform under mechanical loads and how much electrical charge it will generate. This relationship is essential for optimizing PVDF’s performance in applications that require specific mechanical and electrical coupling behaviors. The stiffness parameters and piezoelectric strain constants were measured using mechanical testing and non-contact 3D digital image correlation. The key parameter, d31, was measured using three different test methods.
Stiffness parameters: The tensile testing experiments used 50 µm and 75 µm thick metalized PVDF films (PolyK Technologies, Philipsburg, PA, USA). The PVDF specimens were magnetron sputter coated with aluminum to a sheet resistance of 1 Ω/□, corresponding to a roughly 100 nm thick metal film. The specimens were tested in tension according to the ASTM D638 [
32] standard for tensile properties of polymer composites. The experiments were performed in a displacement control mode at a rate of 0.5 cm/min. A UV laser cutter machined the dog-bone specimens from flat sheets. The optical strain measurements were performed using a 3D DIC system (Q-400; Dantec Dynamics GmbH, Ulm, Germany, and Skovlunde, Denmark) using 8.1 MP resolution cameras and 35 mm lenses. Before testing, the dog-bone specimens were coated with a flat white paint to conceal the electrodes’ shiny surface and reduce reflections. A pattern was created using black paint applied through a stencil. The cameras were calibrated using a glass calibration plate from the vendor.
Piezoelectric strain coefficients using digital image correlation (DIC): Due to the recent developments in poled PVDF films, there is no standardized testing structure for collecting strain and deflection data on the piezoelectric material. Two methods were employed using the DIC system. A Semiprobe PSM-1000 microscope (Winooski, VT, USA) set to 20× magnification was used to view the material with an AMSCOPE MU300 microscope camera (Irvine, CA, USA) affixed with a 0.5× lens to optimize the field of view. AMSCOPE’s software (Version 4.8) allowed for 1536 × 2048 resolution images to be captured. A 50 µm thick film with metallization (electrodes) on top and bottom was fabricated into a test piece with a 100 × 20 mm active area (area in between electrodes). The long axis of the piece was oriented in the machine direction and was secured along the short end to a non-moving, flat substrate (see
Figure 1). The remainder of the test piece was free to expand and contract. The specimen was coated with a flat white paint to reduce reflections, and an irregular pattern was created using a faint layer of air-brushed black paint.
One test method directly measured strain in the 5 × 5.5 mm DIC viewing area and was used to find d31, d32, and d36 coefficients. The other method only measured d31 by tracking the total displacement from the secured edge. Voltages were applied to the electrodes to create expansion and contraction. Then, a secondary voltage was applied between the bottom electrode and the substrate, separated by a thin dielectric. This secured the film to the substrate by electro-adhesion, thus creating a repeatable test by eliminating out-of-plane motion. DIC pictures were taken at voltages ranging from ± 1200 V, corresponding to electric fields ranging over ±24 MV/m.
Piezoelectric strain coefficient,
d31, using a direct mechanical approach: The third method to measure the piezoelectric coupling coefficient
d31 used a mechanical testing system (UStretch; CellScale, Waterloo, ON, Canada). The test piece was a 50 µm thick PVDF film. The active region was 100 × 40 mm, with the long axis oriented in the machine direction. Non-metalized ends extended beyond the active region to provide a clamping location that does not interfere with the piezoelectric strain motion (see
Figure 2). The mechanical tests were run under a constant tension force. A small force (1.0 N) was chosen to minimize stretching (ϵ ~0.01%) of the sample but high enough for the CellScale system to maintain a stable force. Voltage is applied to the electrodes at small protrusions in the middle of the test piece, minimizing the mechanical effect on the test film. Potentials ranged over ±2100 V, corresponding to electric fields as high as ±42 MV/m. At each high voltage (HV) input, the sample was cyclically loaded, starting at 0 V and then going to +HV, back to 0 V, then −HV, back to 0 V, and repeating. Cycling in this manner generated the most piezoelectric strain and would potentially show any hysteresis effect.
3. Theoretical Analysis of Mechanical and Electrical Behavior
Poled polyvinylidene fluoride (PVDF) exhibits orthotropy due to its anisotropic structure from the alignment of molecular chains during the poling process. This alignment causes PVDF to have direction-dependent mechanical properties, such as differing stiffness and strength along its principal axes. To accurately capture the material behavior of poled PVDF, particularly in both the elastic and plastic regimes, models must account for this orthotropy. Without considering the directional dependence, predictions of the stress–strain response and deformation under various loads would be inaccurate, leading to errors in design and performance assessments of components made from poled PVDF. Thus, to model the experimental values obtained, we propose to use the classical strain-charge relationship accounting for elastic orthotropy coupled with an orthotropic plasticity model based on Hill’s orthotropic yield function.
The off-axis stiffness can be related to the stiffness in the material directions using the following relationship based on the transformation of the compliance matrix [
33]:
In this equation,
is the machine direction modulus of elasticity,
is the transverse modulus,
is Poisson’s ratio, and
is the shear modulus. Next, the strain-charge relationship for an electric field perpendicular to the film,
is proposed as follows to account for both the elastic stresses and the strains induced due to the piezoelectric effect:
where S is the compliance,
are piezoelectric strain coefficients and ϵ and σ are the inplane strain and stress components, respectively. The strain in the 1′–2′ direction, as the material is rotated, can then be described as follows, where
with
,
, where
is measured from the reference (1–2) to the (1′–2′) directions. The anisotropic plasticity is simulated with Hill’s orthotropic yield function. Hill’s orthotropic yield function is widely used to simulate anisotropic plasticity in materials with directionally dependent properties. This yield criterion extends the classical von Mises isotropic yield function by introducing distinct parameters for different principal material directions. This allows the model to capture variations in yield strength and plastic deformation behavior across these directions. By incorporating the orthotropic yield function, we can capture the responses to stresses along their various axes. For the PVDF material studied, the yield function
is given by
In the equation, Hill’s function involves multiple parameters, F, G, H, L, M, and N, that correspond to the materials state of anisotropy, allowing for a more accurate prediction of plastic deformation. In terms of volume, the model assumes incompressibility during plastic deformation, meaning the material’s volume remains constant.
Finite element simulation: An axisymmetric dog-bone geometry with dimensions shown in
Figure 3 is created along the y-direction and modeled using the Solid Mechanics module in the COMSOL Multiphysics finite element software (Version 6.2, COMSOL, Boston, MA, USA). For capturing the plasticity, Hill’s orthotropic yield function is considered, for which the Hill’s parameters are chosen as (50, 100, 50, 25, 50, 25) MPa, along with 1.07 GPa isotropic tangent modulus and 3.82 GPa for the initial modulus. The values correspond to the orthotropic properties relative to the machine direction, with the highest values corresponding to the machine direction. The lower end of the dog-bone specimen is designated as a fixed constraint boundary whereas the opposite end receives a prescribed displacement along the positive y-direction. To study the mechanical stress–strain response for the variation of the machine direction orientation relative to the applied displacement only, a rotated coordinate system (Z-X-Z sequence) is introduced to the linear elastic material.
Including the geometric nonlinearity and physics-controlled extra fine mesh (
Figure 3), a stationary analysis is carried out to compute the stress–strain response at the center point (0, 0, 0) of the axisymmetric test specimen by linearly increasing the prescribed displacement from 0 to 4 mm.
4. Results and Discussion
Mechanical behavior: The stress versus strain response for the four specimens studied is shown in
Figure 4. The results show a large variability in the response based on the stretching direction relative to the applied loading. Strains over 18% were observed when the stretching direction was 0°, 15°, or 45° from the machine direction, with a significant drop in the ultimate strength for the 45° and 90° specimens. The 15° and 45° specimens showed a higher strain to failure and were comparable to the 90° and 0° specimens, which may be attributed to the axial-shear coupling response. A significant reduction in failure strain and ultimate stress is also observed when the machine direction is at 90° to the applied load. A closer look at the stress–strain response below 1% strain shows a linear elastic behavior and is shown in
Figure 4b and also illustrates the orthotropic behavior of the material. The elastic modulus ranged from 3.82 GPa in the machine direction to 1.64 GPa in the transverse direction. A comparison of the stiffness predicted from Equation (1) and the slope from the experimental results is shown in
Figure 5 showing a relatively good approximation of the initial stiffness using the equation provided. This equation can be used to understand the relative stiffness of the material when loaded away from the stretching direction, shows the relative sensitivity of the elastic stiffness to the loading orientation, and can be used in the linear regime of behavior.
For the plastic region, the stress–strain response using Hill’s orthotropic yield function is shown in
Figure 6, showing the ability to capture the plastic region directional dependence of yield strength in the poled PVDF material relative to the loading direction. In the figure, the symbols represent the experimental data, and the line patterns follow the scaled model results. The results show that Hill’s orthotropic yield function can capture the variation in the yield initiation point in the loading relative to the machine direction. The corresponding stress–strain response in the plastic region is shown in
Figure 7 which depicts the ability to capture directional dependence of yield strength in the poled PVDF dog-bone model. Further, the Von Mises stress contours for four distinct orientations (0°, 15°, 45°, 90°) at 3% strain show shear-axial strain coupling in the off-axis specimens, which can also be seen from the modeling results.
Figure 8 shows the optical micrographs of the failed region of all four specimens. From these experiments and observations of the failure locations, we observe the dependence of the failure mode on the direction of the applied load to the machine direction. Most deformation is observed when the applied load is 15° and 45° from the machine direction. In the specimen loaded with the machine direction at 90° to the applied load, the failure plane started at 45° relative to the loading direction. Still, it quickly transitioned to fracture along the transverse to the films’ stretching direction, as seen in
Figure 8d.
Piezoelectric Strain Coefficients
d31,
d32,
d36: Three separate test methods were used on two specimens to measure the piezoelectric strain coefficient
d31.
Figure 9 shows electric field-induced strain in the machine direction as a function of the electric field applied across the film. The slope of a linear fit gives the value for
d31 in pC/N. The pull test (linear fit shown on plot) and DIC displacement methods yielded the same value with different precisions,
d31 = 24.4 ± 1.0 pC/N and
d31 = 24.4 ± 0.1 pC/N, respectively. Directly measuring strain with the DIC system resulted in an 8% lower value,
d31 = 22.4 ± 0.3 pC/N. For reference, the manufacturer reports a value of 28 ± 1.4 pC/N, while another manufacturer (Measurement Specialties, Inc., Hampton, PA, USA) reports a
d31 value of 23 pC/N. No measurable hysteresis effect was observed in the data at these electric fields.
As seen in the
d31 data, directly measuring the strain coefficient with the DIC is the most challenging. The small strains and a limited field of view make detecting differential movement in the sample difficult. Nevertheless, the DIC system measured the much smaller motion corresponding to the
d32 and
d36 coefficients. The DIC detected a small strain in the traverse direction (see
Figure 10). A value of
d32 = 3.3 ± 0.3 pC/N was found, roughly an order of magnitude less than
d31. Although the estimated error in the data points is large, the
p-value of the slope is less than 0.01%, indicating a statistically significant effect. The shear data associated with
d36 is shown in
Figure 11. No statistically significant shear strain was detected; thus,
d36 = 0 ± 0.3 pC/N.
5. Conclusions
Poled (piezoelectric) PVDF material properties were measured and compared with theory and other sources. The mechanical behavior of the specimens exhibited significant variability in stress–strain responses based on the orientation of the machine direction relative to the applied load. Strains over 18% were observed for specimens stretched at 0°, 15°, and 45°, with the 45° specimen showing a notable reduction in ultimate strength. The 0° and 15° specimens demonstrated comparable strength, with the 15° specimen achieving higher strain to failure. However, when stretched at 90°, a considerable reduction in both failure strain and ultimate stress was observed, with the failure mode transitioning along the transverse direction. The initial stiffness prediction from the provided model showed a good approximation to the experimental data, validating its utility for assessing stiffness sensitivity to loading orientation. Hill’s orthotropic yield function can model the anisotropic yield behavior.
In the piezoelectric strain coefficient measurements, the d31 coefficient was measured using three different methods, yielding similar results, though slightly lower than the manufacturer’s reported value. The DIC displacement method provided the most precise measurements. For the d32 coefficient, the DIC system detected small but significant strains in the transverse direction, while no significant shear strain was observed for d36, indicating it is negligible. These findings highlight the complexities in measuring piezoelectric strain coefficients, particularly for smaller coefficients like d32 and d36, where statistical significance was harder to achieve.
PVDF can endure up to 18% strain before failing. Compare this to a typical ceramic-based piezoelectric material, PZT, which cannot reliably exceed 0.1% strain without fracturing [
34]. Although the strain coefficient,
d31 = 24.4 pC/N, is 3.5–11 times less than ceramics [
35], it is the authors’ opinion that the advantages in flexibility, biocompatibility, and micro-manufacturability of PVDF will enable a wide range of applications inaccessible to ceramic devices.
Author Contributions
D.S.: Conceptualization, Methodology, Formal Analysis, Supervision. O.S.: Investigation. M.S.C.: Formal analysis. D.G.: Formal analysis, A.L.: Investigation, C.T.L.: Writing-Review and Editing. R.E.: Writing—Original Draft, Conceptualization, Methodology, Formal Analysis. All authors have read and agreed to the published version of the manuscript.
Funding
This research received no external funding.
Data Availability Statement
The data presented in this study are available on request from the corresponding author.
Conflicts of Interest
The authors declare no conflict of interest.
References
- Jones, G.D.; Assink, R.A.; Dargaville, T.R.; Chaplya, P.M.; Clough, R.L.; Elliott, J.M.; Martin, J.W.; Mowery, D.M.; Celina, M.C. Characterization, Performance and Optimization of PVDF as a Piezoelectric Film for Advanced Space Mirror Concepts; Sandia National Laboratories (SNL): Albuquerque, NM, USA; Livermore, CA, USA, 2005. [Google Scholar]
- Hadas, Z.; Rubes, O.; Tofel, P.; Machu, Z.; Riha, D.; Sevecek, O.; Kastyl, J.; Sobola, D.; Castkova, K. Piezoelectric PVDF Elements and Systems for Mechanical Engineering Applications. In Proceedings of the 2020 19th International Conference on Mechatronics-Mechatronika (ME), Prague, Czech Republic, 2–4 December 2020; IEEE: Piscataway, NJ, USA, 2020; pp. 1–8. [Google Scholar]
- Ruan, L.; Yao, X.; Chang, Y.; Zhou, L.; Qin, G.; Zhang, X. Properties and Applications of the β Phase Poly(vinylidene fluoride). Polymers 2018, 10, 228. [Google Scholar] [CrossRef] [PubMed]
- Kujawa, J.; Boncel, S.; Al-Gharabli, S.; Koter, S.; Kaczmarek-Kędziera, A.; Korczeniewski, E.; Terzyk, A.P. Current and future applications of PVDF-carbon nanomaterials in energy and sensing. Chem. Eng. J. 2024, 492, 151856. [Google Scholar] [CrossRef]
- Zahertar, S.; Wu, J.; Chatzipirpiridis, G.; Ergeneman, O.; Canyelles-Pericas, P.; Tao, R.; Fu, Y.Q.; Torun, H. A Flexible PVDF-based Platform Combining Acoustofluidics and Electroma gnetic Metamaterials. In Proceedings of the 2020 IEEE International Conference on Flexible and Printable Sensors and Systems (FLEPS), Manchester, UK, 16–19 August 2020; IEEE: Piscataway, NJ, USA, 2020; pp. 1–4. [Google Scholar]
- Nivedhitha, D.M.; Jeyanthi, S.; Venkatraman, P.; Viswapriyan, A.S.; Guru Nishaanth, S.; Manoranjith, S. Polyvinylidene Fluoride as an advanced polymer for multifunctional app lications—A review. Mater. Today Proc. 2023, in press. [Google Scholar] [CrossRef]
- Yongrae, R.; Varadan, V.V.; Varadan, K.V. Characterization of all the elastic, dielectric, and piezoelectric con stants of uniaxially oriented poled PVDF films. IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2002, 49, 836–847. [Google Scholar] [CrossRef] [PubMed]
- Varadan, V.; Roh, Y.; Varadan, V.; Tancrell, R. Measurement of all the elastic and dielectric constants of poled PVDF films. In Proceedings of the Proceedings—IEEE Ultrasonics Symposium, Montreal, QC, Canada, 3–6 October 1989; IEEE: Piscataway, NJ, USA, 1989. [Google Scholar] [CrossRef]
- Naarayan, S.; Rao, C.; Srinivasan, S. Nonlinear Elastic Constitutive Model for Pvdf. J. Aerosp. Sci. Technol. 2003, 55, 205–210. [Google Scholar] [CrossRef]
- Vinogradov, A. Constitutive model of piezoelectric polymer PVDF. In Proceedings of the Smart Structures and Materials 1999: Mathematics and Control in Smart Structures, Newport Beach, CA, USA, 1–4 March 1999; SPIE: Cergy-Pontoise, France, 1999. [Google Scholar] [CrossRef]
- Vinogradov, A. Nonlinear Characteristics of Piezoelectric Polymers. In Proceedings of the ASME 2001 International Mechanical Engineering Congress and Exposition. Adaptive Structures and Material Systems, New York, NY, USA, 11–16 November 2001; ASME: New York, NY, USA, 2001. [Google Scholar] [CrossRef]
- Vinogradov, A.; Schumacher, S.; Rassi, E. Dynamic response of the piezoelectric polymer PVDF. Int. J. Appl. Electromagn. Mech. 2005, 22, 39–51. [Google Scholar] [CrossRef]
- Carbeck, J.; Rutledgde, G. Material Behavior of Poly(Vinylidene Fluoride) Deduced from Molecular Modeling. In Fluoropolymers 2: Properties; Hougham, G., Cassidy, P.E., Johns, K., Davidson, T., Eds.; Springer: Boston, MA, USA, 2002. [Google Scholar] [CrossRef]
- Bystrov, V.; Pullar, R.; Kholkin, A.; Gevorkyan, V.; Avakyan, L.; Bdikin, I.; Kopyl, S.; Paramonova, R.; Bystrova, A. Modeling of switching and piezoelectric phenomena in polyvinylideneflu oride (PVDF). In Proceedings of the 2013 Joint IEEE International Symposium on Applications of Ferroelectr ic and Workshop on Piezoresponse Force Microscopy (ISAF/PFM), Prague, Czech Republic, 21–25 July 2013; IEEE: Piscataway, NJ, USA, 2013. [Google Scholar] [CrossRef]
- Sathiyanarayan, S.; Sivakumar, S.; Lakshmana, C. Nonlinear and time-dependent electromechanical behavior of polyvinylidene fluoride. Smart Mater. Struct. 2006, 15, 767. [Google Scholar] [CrossRef]
- Seminara, L.; Capurro, M.; Cirillo, P.; Cannata, G.; Valle, M. Electromechanical characterization of piezoelectric PVDF polymer films for tactile sensors in robotics applications. Sens. Actuators A Phys. 2011, 169, 49–58. [Google Scholar] [CrossRef]
- Vinogradov, A.M.; Holloway, F. Mechanical testing and characterization of PVDF, a thin film piezoelectric polymer. J. Adv. Mater. 1997, 29, 567435. [Google Scholar]
- Lee, H.; Cooper, R.; Wang, K.; Liang, H. Nano-scale characterization of a piezoelectric polymer (polyvinylidene difluoride, PVDF). Sensors 2008, 8, 7359–7368. [Google Scholar] [CrossRef] [PubMed]
- Wang, Y.; Zhu, L.; Du, C. Progress in piezoelectric nanogenerators based on PVDF composite films. Micromachines 2021, 12, 1278. [Google Scholar] [CrossRef] [PubMed]
- Fortunato, M.; Bidsorkhi, H.C.; Chandraiahgari, C.R.; De Bellis, G.; Sarto, F.; Sarto, M.S. PFM characterization of PVDF nanocomposite films with enhanced piezoelectric response. IEEE Trans. Nanotechnol. 2018, 17, 955–961. [Google Scholar] [CrossRef]
- Dodds, J.S.; Meyers, F.N.; Loh, K.J. Piezoelectric characterization of PVDF-TrFE thin films enhanced with ZnO nanoparticles. IEEE Sens. J. 2011, 12, 1889–1890. [Google Scholar] [CrossRef]
- Islam, S.; Song, X.; Choi, E.T.; Kim, J.; Liu, H.; Kim, A. In vitro study on smart stent for autonomous post-endovascular aneurysm repair surveillance. IEEE Access 2020, 8, 96340–96346. [Google Scholar] [CrossRef]
- Pan, F.; Sui, J.; Silva-Pedraza, Z.; Bontekoe, J.; Carlos, C.R.; Wu, G.; Liu, W.; Gao, J.; Liu, B.; Wang, X. 3D-Printed Piezoelectric Stents for Electricity Generation Driven by Pressure Fluctuation. ACS Appl. Mater. Interfaces 2024, 16, 27705–27713. [Google Scholar] [CrossRef]
- Mokhtari, F.; Azimi, B.; Salehi, M.; Hashemikia, S.; Danti, S. Recent advances of polymer-based piezoelectric composites for biomedical applications. J. Mech. Behav. Biomed. Mater. 2021, 122, 104669. [Google Scholar] [CrossRef] [PubMed]
- Ribeiro, C.; Sencadas, V.; Correia, D.M.; Lanceros-Méndez, S. Piezoelectric polymers as biomaterials for tissue engineering applications. Colloids Surf. B Biointerfaces 2015, 136, 46–55. [Google Scholar] [CrossRef] [PubMed]
- Chen, A.; Su, J.; Li, Y.; Zhang, H.; Shi, Y.; Yan, C.; Lu, J. 3D/4D printed bio-piezoelectric smart scaffolds for next-generation bone tissue engineering. Int. J. Extrem. Manuf. 2023, 5, 032007. [Google Scholar] [CrossRef]
- Zaszczynska, A.; Sajkiewicz, P.; Gradys, A. Piezoelectric scaffolds as smart materials for neural tissue engineering. Polymers 2020, 12, 161. [Google Scholar] [CrossRef]
- Liu, Z.; Zhang, S.; Jin, Y.; Ouyang, H.; Zou, Y.; Wang, X.; Xie, L.; Li, Z. Flexible piezoelectric nanogenerator in wearable self-powered active sensor for respiration and healthcare monitoring. Semicond. Sci. Technol. 2017, 32, 064004. [Google Scholar] [CrossRef]
- Yu, J.; Hou, X.; He, J.; Cui, M.; Wang, C.; Geng, W.; Mu, J.; Han, B.; Chou, X. Ultra-flexible and high-sensitive triboelectric nanogenerator as electronic skin for self-powered human physiological signal monitoring. Nano Energy 2020, 69, 104437. [Google Scholar] [CrossRef]
- Yang, T.; Pan, H.; Tian, G.; Zhang, B.; Xiong, D.; Gao, Y.; Yan, C.; Chu, X.; Chen, N.; Zhong, S. Hierarchically structured PVDF/ZnO core-shell nanofibers for self-powered physiological monitoring electronics. Nano Energy 2020, 72, 104706. [Google Scholar] [CrossRef]
- Park, S.; Guan, X.; Kim, Y.; Creighton, F.X.; Wei, E.; Kymissis, I.; Nakajima, H.H.; Olson, E.S. PVDF-based piezoelectric microphone for sound detection inside the cochlea: Toward totally implantable cochlear implants. Trends Hear. 2018, 22, 2331216518774450. [Google Scholar] [CrossRef] [PubMed]
- D638-14; Standard Test Method for Tensile Properties of Plastics. ASTM International: West Conshohocken, PA, USA, 2017.
- Barbero, E.J. Introduction to Composite Materials Design; CRC Press: Boca Raton, FL, USA, 2010. [Google Scholar]
- Guillon, O.; Thiebaud, F.; Perreux, D. Tensile fracture of soft and hard PZT. Int. J. Fract. 2002, 117, 235–246. [Google Scholar] [CrossRef]
- Hooker, M.W. Properties of PZT-Based Piezoelectric Ceramics Between −150 and 250 °C; Langley Research Center: Hampton, VA, USA, 1998. [Google Scholar]
Figure 1.
The DIC setup shows the test strip clamped on the left side, a speckled pattern, and a microscope.
Figure 1.
The DIC setup shows the test strip clamped on the left side, a speckled pattern, and a microscope.
Figure 2.
100 × 40 mm × 75 µm PVDF specimen in mechanical test system.
Figure 2.
100 × 40 mm × 75 µm PVDF specimen in mechanical test system.
Figure 3.
Geometry of the dog-bone model and COMSOL Multiphysics fine mesh.
Figure 3.
Geometry of the dog-bone model and COMSOL Multiphysics fine mesh.
Figure 4.
Stress versus strain as a function of the loading angle relative to the stretching or main axis, (a) stiffness details of the active range showing variability relative to the stretching or main axis (b).
Figure 4.
Stress versus strain as a function of the loading angle relative to the stretching or main axis, (a) stiffness details of the active range showing variability relative to the stretching or main axis (b).
Figure 5.
Axial stiffness versus loading angle relative to the primary axis using model in Equation (1) for the elastic region of the response.
Figure 5.
Axial stiffness versus loading angle relative to the primary axis using model in Equation (1) for the elastic region of the response.
Figure 6.
Stress–strain response in the plastic region from the FE model using Hill’s orthotropic yield function, showing ability to capture directional dependence of yield strength in the poled PVDF.
Figure 6.
Stress–strain response in the plastic region from the FE model using Hill’s orthotropic yield function, showing ability to capture directional dependence of yield strength in the poled PVDF.
Figure 7.
Von Mises stresses post-yield for specimens with the machine direction oriented at an angle, θ, relative to the applied load at a strain value of 3%.
Figure 7.
Von Mises stresses post-yield for specimens with the machine direction oriented at an angle, θ, relative to the applied load at a strain value of 3%.
Figure 8.
Optical micrographs of failed specimens at various angles to the applied load. Each image was taken using a Keyence VHX, model number OP-88660, showing (a) 0° to the applied load, (b) 15° to the applied load showing more deformation in the specimen, (c) 45° to the applied load also showing multiple regions of large deformations, (d) 90° to the applied load showing fracture starting at 45° and transition to failure transverse to the stretching direction of the film.
Figure 8.
Optical micrographs of failed specimens at various angles to the applied load. Each image was taken using a Keyence VHX, model number OP-88660, showing (a) 0° to the applied load, (b) 15° to the applied load showing more deformation in the specimen, (c) 45° to the applied load also showing multiple regions of large deformations, (d) 90° to the applied load showing fracture starting at 45° and transition to failure transverse to the stretching direction of the film.
Figure 9.
Piezoelectric strain, corresponding to d31, vs. electric field using three different test methods: mechanical testing system, DIC direct strain, and DIC displacement.
Figure 9.
Piezoelectric strain, corresponding to d31, vs. electric field using three different test methods: mechanical testing system, DIC direct strain, and DIC displacement.
Figure 10.
Piezoelectric strain, corresponding to d32, vs. electric field, as measured by the DIC system.
Figure 10.
Piezoelectric strain, corresponding to d32, vs. electric field, as measured by the DIC system.
Figure 11.
Piezoelectric shear strain, corresponding to d36, vs. electric field, as measured by the DIC system.
Figure 11.
Piezoelectric shear strain, corresponding to d36, vs. electric field, as measured by the DIC system.
| 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. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).