Han-Yue
Zhang
*
State Key Laboratory of Bioelectronics, School of Biological Science and Medical Engineering, Southeast University, Nanjing 210096, People's Republic of China. E-mail: zhanghanyue@seu.edu.cn
First published on 7th April 2022
Piezoelectric materials that generate electricity when deforming are ideal for many implantable medical sensing devices. In modern piezoelectric materials, inorganic ceramics and polymers are two important branches, represented by lead zirconate titanate (PZT) and polyvinylidene difluoride (PVDF). However, PVDF is a nondegradable plastic with poor crystallinity and a large coercive field, and PZT suffers from high sintering temperature and toxic heavy element. Here, we successfully design a metal-free small-molecule ferroelectric, 3,3-difluorocyclobutanammonium hydrochloride ((3,3-DFCBA)Cl), which has high piezoelectric voltage coefficients g33 (437.2 × 10−3 V m N−1) and g31 (586.2 × 10−3 V m N−1), a large electrostriction coefficient Q33 (about 4.29 m4 C−2) and low acoustic impedance z0 (2.25 × 106 kg s−1 m−2), significantly outperforming PZT (g33 = 34 × 10−3 V m N−1 and z0 = 2.54 × 107 kg s−1 m−2) and PVDF (g33 = 286.7 × 10−3 V m N−1, g31 = 185.9 × 10−3 V m N−1, Q33 = 1.3 m4 C−2, and z0 = 3.69 × 106 kg s−1 m−2). Such a low acoustic impedance matches that of the body (1.38–1.99 × 106 kg s−1 m−2) reasonably well, making it attractive as next-generation biocompatible piezoelectric devices for health monitoring and “disposable” invasive medical ultrasound imaging.
The key obstacles to realizing a desirable piezoelectric sensor material include a large g33 for better detecting the generated signal above the background noise, a high Curie temperature (Tc) that is crucial towards extending the usage temperature range, and especially the adaptability to varying environmental conditions.13 Since the 1950s, inorganic ferroelectric ceramics represented by barium titanate (BTO) and lead zirconate titanate (PZT) have been the leading piezoelectric materials.14–16 Nevertheless, their g33 values are relatively limited (<40 × 10−3 V m N−1), while the high acoustic impedance poses problems when operated in a hydrostatic environment and bodily tissues.17 If the difference in the acoustic impedance of the two media is large, most of the energy will not be transmitted and will be reflected. The obvious comparison is ferroelectric polymers like polyvinylidene difluoride (PVDF), which has been widely investigated as a sensor material because of the much higher g33 coefficient up to 286.7 × 10−3 V m N−1, as well as enhanced acoustic impedance matching with water or the human body.18–22 As the representative and the most important organic ferroelectrics without toxic metals, their attractive advantages of solution-based low-temperature processing, low fabrication cost, biocompatibility, and integrability with silicon technology also show great promise for next-generation wearable and biomedical devices.23 Notably, there are also inevitable challenges for PVDF. Although much effort has been made to reduce the large coercive field (∼0.5 MV cm−1 for PVDF),24,25 due to the nature of the polymer, switching polarization still needs to overcome the large energetic barrier. Moreover, the usage of PVDF will be further hindered by the poor crystallinity and nondegradable properties of plastic, and the degraded polar properties after being heated to about 353 K.26
Actually, the family of organic ferroelectrics is not just limited to those well-known and extensively studied polymers, but includes small-molecule ferroelectric materials. In addition to most of the above attractive superiorities of organic ferroelectrics, these simple small-molecule candidates can also possess potentially easier, less expensive, and lower power processing than their polymeric counterparts.27–31 More importantly, as compared to PVDF, they also give added large benefits like the ability to form a long-range crystalline order and the freedom of molecular design, offering a rich platform for engineering high-g33 piezoelectric materials. A renewed surge of interest in the field of organic electronics has occurred in the past decade, and there is a strong impetus to explore replacement metal-free small-molecule materials with comparable piezoelectric or ferroelectric properties to those of commercial inorganic and polymeric species.32–34 Herein, on the basis of the substitution strategy of hydrogen by F atoms in the prototype compound cyclobutanammonium hydrochloride ((CBA)Cl), we obtained a metal-free small-molecule ferroelectric, 3,3-difluorocyclobutanammonium hydrochloride ((3,3-DFCBA)Cl) (Scheme S1 and Table S1†).35–37 The introduction of F atoms changes the hydrogen bonding network and results in a polar crystal structure. Its unprecedentedly high g33 coefficient of 437.2 × 10−3 V m N−1 of the (3,3-DFCBA)Cl thin film far outperforms that of the state-of-art ferroelectric ceramics and is two times larger than that of PVDF. Moreover, this purely organic component endows (3,3-DFCBA)Cl with a low acoustic impedance (2.25 × 106 kg s−1 m−2), comparable to that of PVDF (3.69 × 106 kg s−1 m−2) and an order of magnitude lower than that of inorganic ferroelectric ceramics (PZT, 2.54 × 107 kg s−1 m−2), which matches that of the body (1.38–1.99 × 106 kg s−1 m−2) reasonably well.18 This combination of outstanding piezoelectric properties and good biocompatibility suggests that (3,3-DFCBA)Cl may replace PVDF in some applications such as health monitoring and medical ultrasound imaging, with benefits in terms of easiness and environment-friendly processing.
For ferroelectric materials, compounds must crystallize in the ten polar point groups (C1, C2, Cs, C2v, C4, C4v, C3, C3v, C6, and C6v). Therefore, second harmonic generation measurements were performed on (3,3-DFCBA)Cl and (CBA)Cl to detect whether the compound has inversion symmetry. For convenience, the phase below Tc1 can be labeled as the low-temperature phase (LTP), the phase above Tc2 as the high-temperature phase (HTP), and the phase between Tc1 and Tc2 as the intermediate-temperature phase (ITP). As shown in Fig. 1B and S2,† (3,3-DFCBA)Cl and (CBA)Cl show a clear second harmonic generation (SHG) signal at room temperature, representing the SHG activity. It is worth noting that the SHG intensity of (3,3-DFCBA)Cl is stronger than that of (CBA)Cl, which can be attributed to the enhanced dipole moment of an organic cation after H/F substitution. For (3,3-DFCBA)Cl, the SHG intensity drops sharply from 1.2 to 0.14 around Tc1 and then drops to about 0 around Tc2. While for (CBA)Cl, the SHG intensity drops sharply from 0.78 to 0.41 around Tc1. The variable-temperature SHG measurements indicate that (3,3-DFCBA)Cl undergoes phase transitions from noncentrosymmetry to noncentrosymmetry and then to centrosymmetry, while (CBA)Cl undergoes a phase transition from noncentrosymmetry to noncentrosymmetry.
The polarization–electric field (P–E) hysteresis loops were measured to characterize the switchable polarization of (3,3-DFCBA)Cl by using the double-wave method. A typical ferroelectric J–E (current density–electric field) curve is shown in Fig. 1C with two opposite peaks at 293 K. According to the J–E curve, we obtained the P–E hysteresis loop by measuring the current. Under the influence of the applied voltage, the measured saturated polarization (Ps) value is about 6.4 μC cm−2, which is close to the calculated value (6.9 μC cm−2) depending on the point charge model (Fig. S3†).
In the vicinity of Tc, dielectric permittivity as a function of temperature generally shows noticeable anomalies in ferroelectrics. As shown in Fig. 1D, the real part ε′ of the dielectric constant of the (3,3-DFCBA)Cl polycrystalline sample at 1 kHz maintains a very low value about 7.88 at 300 K and its value is almost constant in the temperature range of 300–370 K (LTP). Near Tc1, the ε′ shows a significant increase with a step type, entering into ITP, and then increasing gradually. When the temperature increases to Tc2, the ε′ exhibits a remarkable λ-shape dielectric response, and finally stabilizes in the HTP. Similarly, (CBA)Cl also shows obvious dielectric anomalies near the phase transition temperature, which is consistent with the DSC measurement (Fig. S4†). The real part ε′ of the dielectric constant for the (CBA)Cl polycrystalline sample is about 2.6 at 1 kHz and 1.9 at 1 MHz at 300 K, respectively. Specifically, the ε′ measured here is the relative dielectric constant εr.
The crystal data for (3,3-DFCBA)Cl and (CBA)Cl are summarized in Table S2.† At 293 K, (CBA)Cl crystallizes in the orthorhombic chiral space group P212121 with cell parameters of a = 7.8522(6) Å, b = 8.8201(6) Å, c = 8.9227(6) Å, and V = 617.96(8) Å3, which means that it is impossible for (CBA)Cl to have ferroelectricity. Fortunately, when the F atom replaces the H atom in the para position of the CBA molecule, (3,3-DFCBA)Cl crystallizes in the orthorhombic polar space group Pmn21 with cell parameters of a = 6.0473(4) Å, b = 7.0878(5) Å, c = 7.6493(5) Å, and V = 327.87(4) Å3, meaning the possibility of ferroelectricity. As shown in Fig. 2A and C, all the crystal structures of (CBA)Cl and (3,3-DFCBA)Cl consist of monoprotonated cations and Cl anions, all in an ordered state. In particular, the introduction of F atoms has subtly changed the molecular structure of CBA. As shown in Fig. S5 and Table S3,† the plane angle of the CBA molecule decreases from 32.6(2)° in CBA to 22.6(2)° in 3,3-DFCBA, and the N1–C1–C3 bond angle changes from 146.0(2)° to 139.5(2)°. The different molecular sizes and the degree of twist in the molecular structure between CBA and 3,3-DFCBA directly affect the hydrogen bonding interaction, and then change the packing mode of the crystal structure. In the (CBA)Cl and (3,3-DFCBA)Cl crystals, each H atom on the N atom forms N–H⋯Cl hydrogen bonding interaction with its adjacent Cl atoms with a donor–acceptor distance from 3.17(2) Å to 3.22(1) Å (Fig. S5 and Table S4†). As shown in Fig. 2B and S6C,† CBA cations arrange in an anti-parallel manner in the ac plane and form a three-dimensional network through hydrogen bonding interaction. However, (3,3-DFCBA)Cl forms a two-dimensional hydrogen bonding network in the ac plane (Fig. 2D and S6D†). Different from (CBA)Cl, 3,3-DFCBA cations align head-to-head in the ac plane and tilt towards the same direction along the c-axis, which should lead to a spontaneous polarization (Ps).
In order to explore the ferroelectric phase transition of (3,3-DFCBA)Cl, we have tried to determine the crystal structure of the paraelectric phase. Unfortunately, the crystal of (3,3-DFCBA)Cl becomes nontransparent at 363 K before Tc1, resulting in poor X-ray diffraction. Thanks to the low phase transition temperature of the prototype compound (CBA)Cl (Tc = 318.6 K), we successfully obtained the HTP crystal structure of (CBA)Cl at 343 K. (CBA)Cl crystallized in the cubic chiral space group P213 at 343 K. As shown in Fig. S7,† and the CBA cations are located in a special position of the three-fold rotation axes in the HTP, which are highly disordered compared to the ordered state in the LTP. Thus, it can be inferred that the phase transitions of (3,3-DFCBA)Cl are also caused by the ordered-disordered transition of cations. This kind of phase transition caused by ordered-disordered molecular motion is very common in the molecular-based ferroelectric system, such as 3-hydroxyquinucline hydrochloride.39 To further determine the HTP of (3,3-DFCBA)Cl, variable-temperature powder X-ray diffraction (PXRD) measurements were carried out to provide some information about the phase transition behaviors (Fig. S8†). The pattern in the HTP of (3,3-DFCBA)Cl shows obvious changes in the number of diffraction peaks. The Pawley refinements indicate a few possible tetragonal systems with the point group 4/mmm for (3,3-DFCBA)Cl. This simulation result is consistent with the SHG measurement. According to the Aizu rule, the phase transition from the 4/mmm to mm2 point group in (3,3-DFCBA)Cl is definitely a ferroelectric one with the Aizu notion of 4/mmmFmm2, revealing a multiaxial ferroelectric characteristic.
In order to gain deep insight into the ferroelectric polarization reversal, density functional theory (DFT) calculations were carried out to evaluate the origin of polarization.40,41 According to the modern theory of polarization,42,43 the necessity of constructing a polarization change path lies in selecting the polarization quantum properly to avoid a wrong estimation of the polarization value. A dynamic path between two ferroelectric states is constructed based on the crystal structure obtained from single crystal X-ray diffraction. Accordingly, the structure of the room temperature ferroelectric phase at 293 K is used as one ferroelectric configuration, while the other states are obtained from the matrix transformation of the coordinates considering both the rotation of the 3,3-DFCBA cations and displacement of anionic Cl− (Fig. 3). In the dynamic path, the sense of 3,3-DFCBA cation rotation is defined as a pair of clockwise and counterclockwise rotations to keep the polarization along the c axis during the ferroelectric reversal (canceling each other perpendicular to the c-axis). The variation of polarization as a function of the dynamic path is shown in Fig. S9A,† from which the ferroelectric polarization with 7.62 μC cm−2 along the c-axis can be extracted from two equivalent ferroelectric configurations (λ = ±1). During the ferroelectric switching process (−1 < λ < 1), the polarization value changes monotonously, and becomes zero at λ = 0, which indicates a reference phase with zero polarization. On the other hand, the energies of two ferroelectric states with an antiparallel polarization direction (λ = ±1) are symmetric (Fig. S9B†). The energy barrier for the polarization reversal reaches a maximum at theλ = 0 state. The variation of the energy path shows a typical ferroelectric double-well potential with two opposite polarization states located at two symmetric energy minimums, which provides direct proof for the appropriate polarization explanation.
The ferroelectric properties of (3,3-DFCBA)Cl are further characterized by performing piezoresponse force microscopy (PFM) measurements in its thin-film samples. Fig. 4A–D shows the domain structure observed in the as-grown (3,3-DFCBA)Cl thin film, where the images are constructed by overlaying the lateral and vertical PFM phase and amplitude mappings on three-dimensional (3D) topography. No topographic features or crosstalk are observed in the PFM images. The same domain patterns can be clearly observed in the lateral (Fig. 4C) and vertical (Fig. 4D) PFM amplitude images. For PFM phase imaging which reflects the domain orientation, we observed that the adjacent domains have a phase contrast of 180° in the lateral PFM mode (Fig. 4A), while it is an almost unique color tone in the vertical PFM mode (Fig. 4B). Therefore, the polarization directions of adjacent domains are the same in the out-of-plane component, and different in the in-plane component, suggesting that domain walls should be non-180° ones. These findings indicate that (3,3-DFCBA)Cl is a multiaxial ferroelectric, where the non-180° domain wall is permitted.44
Fig. 4 PFM of (3,3-DFCBA)Cl. Lateral PFM (A) phase and (B) amplitude mappings overlaid on 3D topography. Vertical PFM (C) phase and (D) amplitude mappings are overlaid on 3D topography. (E) PFM phase hysteretic loops and (F) butterfly-shaped amplitude loops for a selected point. Vertical phase images of a single domain region (G) in the initial state, (H) after electric writing over the red-box region with a voltage of +50 V and (I) back switching operation proceeded by applying an opposite voltage of −50 V over the smaller red-box region. The corresponding topography and amplitude images for (G)–(I) are shown in Fig. S10.† |
PFM is also an efficient tool for probing and switching the local ferroelectric polarization at the nanoscale. We detected a DC bias-dependent PFM amplitude and phase response underneath the PFM tip. As shown in Fig. 4E and F, the typically butterfly-shaped amplitude curves and hysteretic phase curves demonstrate the switchable polarization in the (3,3-DFCBA)Cl thin film. Then, we selected a single-domain region for domain switching measurements (Fig. 4G). The red-box region in Fig. 4H was first poled with a voltage of +50 V, and followed with second electric writing with an opposite voltage of −50 V in the smaller red-box region in Fig. 4I. Finally, a box-in-box switched domain pattern was generated, providing robust evidence for the ferroelectricity of (3,3-DFCBA)Cl.
Next, the piezoelectricity of the (3,3-DFCBA)Cl thin film was investigated by using the PFM technique, a powerful method to probe electromechanical coupling at the nanoscale.33,45 We first excited the thin film by using PFM tip across the resonant frequency in the vertical PFM mode at a voltage of 2 V and compared it to that of the PVDF film at the same drive voltage. We found that the resonance peak of the (3,3-DFCBA)Cl thin film was comparable with that of the PVDF thin film (Fig. 5A). The vertical PFM measures the z-axis vibration of the tip cantilever, which is proportional to the piezoelectric coefficient d33 of the sample. To estimate the value of the piezoelectric coefficient, d33, of the (3,3-DFCBA)Cl thin film in the out-of-plane component, such resonance measurements have been performed under a set of drive voltages. The peak values which have been corrected by quality factors were plotted against the drive voltages, as shown in Fig. 5B. Both lines show good linearity, indicating that the response comes from intrinsic piezoelectricity.46 Their slopes represent the relative magnitude of d33. Here, the PVDF thin film was set as the benchmark, whose d33 is known as 33 pC N−1.34 We then estimated the d33 of the (3,3-DFCBA)Cl thin film as 30.5 pC N−1 according to the slope values, comparable to those of classical organic piezoelectrics (Table S5†).47–53
Moreover, we observed that the (3,3-DFCBA)Cl thin film also has a strong in-plane piezoresponse when we aligned the PFM cantilever along the y-axis direction, as shown in Fig. S11.† In this case, the measured polarization is along the x component, and the cantilever motion is induced by the piezoelectric coefficient d31.54 Using the same approach, we extract the d31 of the (3,3-DFCBA)Cl film from the comparison with that of PVDF, whose d31 has been determined to be 21.4 pC N−1.55Fig. 5C shows the resonance peaks for both films in the lateral PFM mode under a voltage of 5 V, where a stronger resonance peak is observed in the (3,3-DFCBA)Cl film. Furthermore, we drove both films under a set of voltages up to 5 V, as shown in Fig. 5D. The slope of the (3,3-DFCBA)Cl film is 1.91 times larger than that of the PVDF film. Then, we estimated the d31 of (3,3-DFCBA)Cl to be about 40.9 pC N−1.
For practical application, the piezoelectric voltage coefficient (gij), as a key parameter for piezoelectric sensors, can be evaluated through piezoelectric effects and dielectric permittivity. The piezoelectric voltage coefficient g33 can be obtained through the formula g33 = d33/ε33, in which ε33 can be derived from εr = ε33/ε0. Therefore, we calculated that the g33 for our thin film is 437.2 × 10−3 V m N−1, about two times larger than that of PVDF, 286.7 × 10−3 V m N−1, and much higher than that of high-end PZT-based piezoelectric ceramics (about 20–40 × 10−3 V m N−1) (Table S5†). With regard to the in-plane piezoelectric voltage coefficient, g31 = d31/ε33, we calculated it to be about 586.2 × 10−3 V m N−1. The large in-plane piezoelectric voltage coefficient makes the (3,3-DFCBA)Cl thin film have great merit for in-plane devices. Of particular note is the low acoustic impedance (z0) of (3,3-DFCBA)Cl. The z0 value of (3,3-DFCBA)Cl (2.25–3.26 × 106 kg s−1 m−2) is significantly lower than that of the conventional molecular ferroelectric triglycine sulfate (TGS, 9.74 × 106 kg s−1 m−2) and even that of the ferroelectric polymer PVDF (3.69 × 106 kg s−1 m−2) (Table S6†), an order of magnitude lower than that of inorganic ferroelectric ceramics (PZT, 2.54 × 107 kg s−1 m−2), which suggests good biocompatibility and efficient broadband operation of (3,3-DFCBA)Cl based electroacoustic and electromechanical transducers operating in tissue, water, or other low impedance material.
In addition to a large piezoelectric voltage coefficient (g33), a large electrostriction coefficient (Q33), which can make the materials have the characteristics of a fast response speed, temperature stability, and reducing aging effects, is also critical for good piezoelectrics. Based on the relation of Q33 = g33/2Ps, the electrostriction coefficient Q33 of the (3,3-DFCBA)Cl thin film is calculated to be 4.29 m4 C−2, which is much larger than that of PVDF (1.3 m4 C−2) (Fig. S12†), and two orders of magnitude larger than that of ceramic ferroelectrics.5,20 Herein, an energy harvesting device was fabricated with an electrode-(3,3-DFCBA)Cl-electrode structure based on its pristine polycrystalline sample, which was made with a pressed-powder pellet (Fig. 5E). The generated piezoelectric voltages were recorded after repetitively impacting at a certain mechanical pressure, and measured by using an oscilloscope. As shown in Fig. 5F, the device exhibits an average voltage of about 1.7 V. The good performance of electromechanical energy conversion revealed that (3,3-DFCBA)Cl has great potential in the application of self-powered devices.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1978969–1978973. For ESI and crystallographic data in CIF or other electronic format see https://doi.org/10.1039/d1sc06909h |
This journal is © The Royal Society of Chemistry 2022 |