A 3D-printed molecular ferroelectric metamaterial
Edited by Thomas E. Mallouk, University of Pennsylvania, University Park, PA, and approved September 21, 2020 (received for review July 2, 2020)
Significance
Molecular ferroelectrics, which show the ability to switch the electromechanical activity by an external electric field, establish the basis for mechanical metamaterial technologies. Despite their theoretical promise, such mechanical metamaterials remain hindered by the lack of adaptive stimuli-responsive materials which can be effectively tuned “on demand” across time and length scales. Here, we unravel a printable mechanical metamaterial of imidazolium perchlorate with superior electromechanical coupling and reprogrammable stiffness. We propose a continuous rapid three-dimensional (3D) printing technique which can reduce the manufacturing time of ferroelectrics from hours down to minutes. The printed molecular ferroelectric metamaterial structure is then shown to enable a tunable-frequency vibration-isolating architecture. This study paves the way for rationally designed 3D-printable molecular ferroelectric metamaterials.
Abstract
Molecular ferroelectrics combine electromechanical coupling and electric polarizabilities, offering immense promise in stimuli-dependent metamaterials. Despite such promise, current physical realizations of mechanical metamaterials remain hindered by the lack of rapid-prototyping ferroelectric metamaterial structures. Here, we present a continuous rapid printing strategy for the volumetric deposition of water-soluble molecular ferroelectric metamaterials with precise spatial control in virtually any three-dimensional (3D) geometry by means of an electric-field–assisted additive manufacturing. We demonstrate a scaffold-supported ferroelectric crystalline lattice that enables self-healing and a reprogrammable stiffness for dynamic tuning of mechanical metamaterials with a long lifetime and sustainability. A molecular ferroelectric architecture with resonant inclusions then exhibits adaptive mitigation of incident vibroacoustic dynamic loads via an electrically tunable subwavelength-frequency band gap. The findings shown here pave the way for the versatile additive manufacturing of molecular ferroelectric metamaterials.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
Solution-processable molecular ferroelectrics, which show ferroelectric properties approaching inorganic perovskites, have amassed much recent attention due to their lightweight, tunable electrooptic and electromechanical coupling effects (1–5). Spontaneous polarization and the ability to switch the electromechanical activity by an external electric or mechanical stimulus is of prime importance, establishing the basis for many metamaterial technologies (6–8). Over the past decade, elastic metamaterials with resonant inclusions have gained significant traction owing to their unique response to incident dynamic loads, ranging from subwavelength band gaps (9–11), back-scattering immune wave guides (12, 13), and topological pumps (14–16), to the design of logic gates, nonreciprocity, and diodelike formations (17, 18). Despite their theoretical promise, such mechanical metamaterials remain hindered by the lack of adaptive stimuli-responsive materials which can be effectively tuned “on demand” across the time and length scales dictated by such metamaterials. On the other hand, heterogeneous metamaterial structures and systems have been shown to produce tailorable properties beyond those of the constitutive materials (ferroelectrics) owing to the unique geometrical and topological material reorganization. However, the hallmark feature of such metamaterials is a hierarchical architecture which often exhibits highly complicated internal features, rendering these exotic structures extremely challenging, if not impossible, to achieve with traditional manufacturing processes.
Three-dimensional (3D) printing has been hailed as an emerging advanced manufacturing paradigm, providing a large potential to rapidly fabricate highly complicated metamaterial structures with a wide variety of materials owing to its elegant concept of layer by layer deposition. However, 3D printing is mainly limited to creating complex geometries of nonfunctional structures. The competition between printability and functionality has been identified as the main limitation of 3D printing in functional materials. Generally, a high loading of solid functional component is required to afford functionalities in printed structures (19), while this challenges the homogeneity and viscosity stability for ultimate printability of the feedstock. Additive manufacturing and its design space in functional ferroelectric materials are particularly limited due to the limitation of ferroelectric solid component loading. In addition, the mechanical flexibility and ferroelectricity of printed materials are diminished as a result of randomly organized components (20, 21). Therefore, a seamless integration of additive manufacturing, molecular ferroelectric materials, and metamaterial design would provide the ultimate solution to unlock the functional material and exotic structural properties of molecular ferroelectric metamaterials for unprecedented emerging applications.
Here, we unravel a printable mechanical metamaterial of imidazolium perchlorate (ImClO4) (2, 22, 23), a transparent molecular ferroelectric with superior electromechanical coupling and reprogrammable stiffness. We propose a continuous rapid 3D printing technique which can reduce the manufacturing time of ferroelectrics from hours down to minutes. Different from conventional inorganic or organic–inorganic ferroelectrics (24, 25), the ionic nature of molecular ferroelectrics enables high solubility in water, ensuring a highly dense and transparent precursor. We optically pattern the 3D architecture scaffolds through stereolithography (SLA) 3D printing for selective volumetric crystallization of ionic ImClO4 precursor. The SLA-printed water-soluble ferroelectric precursor and photopolymerizable material with a highly porous yet tough polymer network serve as an excellent carrier to in situ crystalize and organize molecular ferroelectric crystals. The patterned scaffold is then dehydrated under a biased electric field to crystallize ImClO4 with the desired polarization orientation. The printed molecular ferroelectrics also exhibit a self-healing ability from the overloaded mechanical and electric field (26, 27). The printed molecular ferroelectric metamaterial structure is then shown to enable a tunable-frequency vibration-isolating architecture. This study paves the way for rationally designed 3D-printable molecular ferroelectric materials for mechanical metamaterials.
Results
3D Fabrication Process.
Three-dimensional SLA printing enables the creation of volumetric architectures consisting of nanoscale feature size using programmed automation processes (28–30). By analogy, we inferred that this SLA strategy could be incorporated for the volumetric design of ferroelectric metamaterials if the following fundamental prerequisites are met: transparent ferroelectric precursor for the curing light and sufficient curing depth, light-sensitive polymers, and sufficient density to trigger electromechanical coupling and ferroelectricity. If these challenges can be addressed by molecular ferroelectrics, a rapid 3D fabrication strategy would allow control over the geometry, feature size, and polarization direction of molecular ferroelectric metamaterials. Fig. 1A illustrates the schematic diagram of SLA printing for molecular ImClO4 ferroelectrics with high dimensional accuracy, structural complexity, and high throughput. The molecular ferroelectric ImClO4 crystal is water-soluble and can mix with photopolymerizable material to achieve transparent printable precursor ink solution with a viscosity of 1.70 cSt (SI Appendix, Table S1 and Fig. S3) (31). The transparent and highly concentrated ferroelectric precursor solution plays a key role in the SLA printing as its low diffraction index can facilitate light penetration and prevent light scattering. This ultimately allows for a reliable, accurate, and efficient 3D printing process.
Fig. 1.

In the bottom-up SLA process (Fig. 1A), a smaller traction force between the structure being printed and the material container is the vital factor in printing speed and reliability (32), and such force is linearly proportional to viscosity (33). Therefore, the low viscosity of the precursor material enables continuous rapid printing, which reduces the manufacturing time from hours to minutes for the same structure (Movie S1). This represents a game-changing molecular ferroelectric manufacturing technique and offers the unique advantages of the prevention of hydrogel dehydration and printing failure, as well as process efficiency and structural accuracy. During printing (Fig. 1C), the precursor solution is exposed to ultraviolet (UV) light, and poly(ethylene glycol) diacrylate (PEGDA) is then cross-linked to form the scaffold network with the encapsulated ions of the ImClO4 precursor. As shown in Movie S1, a complex geometric structure (Schwarz primitive structure, 25 × 25 × 25 mm3) can be printed in 8 min as opposed to 3 h by conventional methods. The large pore size of the scaffold (about a few micrometers, SI Appendix, Fig. S5) provides higher controllability and flexibility. Additionally, the electrical resistance measurement (SI Appendix, Fig. S6) shows that the as-printed sample is electrically conductive due to the abundance of ions. Dehydration then crystallizes molecular ferroelectric ImClO4, while the polymer skeleton holds the as-printed structure after water evaporation due to the elasticity of the PEGDA polymer network (34). It has to be mentioned that a high loading of molecular ferroelectric component is required to maintain the high polarization in printed structures. The volume ratio of ferroelectric phase can be easily tuned to serve for the desired ferroelectric performance by changing the concentration of ImClO4 or the ratio of PEGDA. Thus, we prepare the precursor with saturated ImClO4 solution and low volume ratio of PEGDA (5 vol %). Finally, dried printed samples with around 50 vol% ratio of ImClO4 are obtained.
Movie S1.
Overall process of printing a 3D hydrogel structure from precursor through SLA printing. Molecular ferroelectric ImClO4 crystal is dissolved in water and mixed with PEGDA in the volume ratio of 9:1 to make the precursor. The pure precursor is transparent. Thus, we add yellow-colored dye to make the precursor yellow in color for recording the printing process. The ultraviolet (UV) light has a wavelength of 385 nm and the control of image projection was achieved through a dynamic micro-mirror device. The precursor was exposed to the UV light, and the PEGDA was then cross-linked to form the scaffold network with the encapsulated ions. As shown in the video, a complex geometric structure (Schwarz primitive structure, 25×25×25 mm3) can be printed in 8 minutes.
In general, the crystal growth in a liquid environment is controlled by two processes: the diffusion of ions through the liquid phase to the growth front and the reorganization of crystal grains into the polycrystal. In our experiments, the diffusion of ions should not be the limiting factor because of the highly concentrated ImClO4 precursor solution and large pore size of the hydrogel scaffold. Thus, the crystal grain reorganization is the limiting step during the growth. By applying a biased electric field, we can organize the orientation of small grains formed right after the nucleation. As shown in Fig. 1C, the dehydration process was conducted under an electric field, which was applied to obtain a high-quality polycrystal with a preferred polarization orientation of ImClO4. The small single-crystal ImClO4 grains can merge into a large polycrystalline structure with preferred crystal orientation as water continuously and slowly evaporates under the biased electric field. Printed ImClO4 with and without the biased electric field are denoted as the printed and control sample, respectively, in the following discussion.
Structural and Optical Properties.
As shown in Fig. 2A, the printed sample maintains its geometry with uniform shrinkage after it is fully dried. Fig. 2B shows its in situ resistance measurement during the drying, while the resistance escalates continuously until the formation of the ImClO4 crystal. As shown in Fig. 2C, if the biased electric field is applied together with drying, the obtained sample shows the preferred orientations with the diffraction peaks at 22.36° and 24.62° representing the () and () planes, respectively. A high diffraction intensity for the () planes suggests that its preferred growth orientation matches with its polarization axis for the pronounced ferroelectric properties (22), confirmed by our theoretical modeling (SI Appendix, Figs. S7–S9). Since the ferroelectric ordering temperature of ImClO4 is far above its crystallization temperature, the anisotropic energy associated with this ferroelectric state is large enough to align the grains under the electric field (SI Appendix, Supporting Text 5). The anisotropic energy leads to the rotation of the crystal nucleus parallel to the external electric field. Moreover, the scaffolds not only have superlative intrinsic properties, such as optical transparency and light weight, but they are also geometrically compatible with the ImClO4 crystal. A cubic ImClO4 sample (SI Appendix, Fig. S4) is printed for optical measurement. Similar to the single crystal (2, 23), the printed ImClO4 is transparent in the visible range (Fig. 2D) with a large energy band gap of 4.1 eV (SI Appendix, Fig. S8). The transparency of the printed sample renders it a promising candidate for electrooptic devices (2).
Fig. 2.

Dielectric, Ferroelectric, and Electromechanical Coupling Properties.
Fig. 3A shows the temperature dependence of dielectric permittivity of the ImClO4 single crystal, the printed sample, and the control sample, which all show the characteristic sharp dielectric anomalies around the phase transition temperature. The second-order phase transition occurs at 375 K from the paraelectric to the ferroelectric phase, suggesting that the transition is not influenced by the crystallization process. However, these samples show distinct dielectric behavior around the first-order phase transition (around 210 K with large hysteresis as shown in SI Appendix, Fig. S10). The peaks of dielectric permittivity for the first-order transition in the printed sample indicate that the applied electric field improves its crystal order with the dielectric anomaly appearing at 237 and 222 K, respectively. More importantly, its dielectric permittivity shows an abrupt jump at 222 K, suggesting the single-crystal–like ferroelectric properties.
Fig. 3.

The ferroelectricity can be attributed to the small permanent dipole moment of the slightly distorted ClO4− anion together with the contribution from the off-center displacement of the imidazolium cations (SI Appendix, Fig. S9). We further carried out the polarization vs. electric-field hysteresis loops (P-E loops) at room temperature. As shown in Fig. 3B, the polarization of the ImClO4 single crystal, the printed sample, and the control sample are 1.49, 0.22, and 0.08 μC/cm2 at 3-kV/cm poling electric field, respectively. As shown in SI Appendix, Fig. S11, the Raman spectrum for printed ImClO4 suggests that the chemical bonding is not influenced by the scaffold when compared with the ImClO4 single crystal. The thermogravimetric analysis (TGA) measurement shows one broad decomposition temperature range from 500 to 700 K; such broad decomposition can be considered as a combination of the decomposition temperature for ImClO4 and PEGDA. The 50% volume ratio for printed ImClO4 is projected to decrease its polarization. In comparison, the 3D-printed PEGDA does not show any P-E loop behavior (SI Appendix, Fig. S12), suggesting that the measured polarization originated from the ImClO4 phase. Often, ferroelectric property degradation could appear in ferroelectrics due to electrical conduction under repetitive cycling or high, local electric field. As shown in Fig. 3C, the degraded sample shows lower polarization when compared with printed ImClO4. But here, the water-based printable ferroelectrics provide an effective way toward self-healing: the printed sample with degraded ferroelectric property could be healed by dissolving into the ImClO4 solution followed by recrystallization. As shown in Fig. 3E and Movie S2, the degraded ImClO4 can be dissolved into ImClO4 solution and lead to the formation of the hydrogel state containing Im+ and ClO4− ions. Additional electrical-field assistant drying process could help to recover its ferroelectric property.
Movie S2.
Self-healing process. Printed ImClO4 with internal cracks is immersed into ImClO4 solution. The printed ImClO4 expands with the formation of the hydrogel state containing Im+ and ClO4− ions. Well-crystalized printed ImClO4 can be obtained after performing the electric field assistant drying process (Fig. 1C).
Metamaterial Simulation.
The printed ferroelectric ImClO4 is expected to show response to an external electric field, particularly a change in material stiffness due to the ferroelectric alignment in the direction of the external field. To investigate this property, the stress–strain curves of the printed sample are recorded under varying electric-field strength and its Young’s modulus is measured, as presented in Fig. 3D and SI Appendix, Table S2. These experiments demonstrate an initial Young’s modulus of 12.0 MPa in the absence of electric field and variation in elastic modulus up to 25% as the field strength increases from 0 to 2,000 V/cm (Fig. 3D and SI Appendix, Table S2), demonstrating significant programmable tunability in the ferroelectric material stiffness. This successful design of a printed molecular ferroelectric material opens up new avenues in the increasingly popular domain of active electroacoustic metamaterials with highly responsive and tunable mechanical properties. To illustrate an application of this tunable material, a “hard–soft–hard” locally resonant metamaterial (LRM) is architected with the printed molecular ferroelectric material as the filler component, in which the soft printed ferroelectric is attached to stiffer matrix and resonator elements to form each unit cell of the periodic structure. This LRM comprises multiple such unit cells in a two-dimensional (2D) array and is schematically depicted in Fig. 4A.
Fig. 4.

An illustration of the behavior of this ferroelectric metamaterial is presented in SI Appendix, Fig. S16, showing the frequency response from 500 to 2,000 Hz at three distinct measurement locations (SI Appendix, Fig. S16A) and illustrating the array’s vibration profile both outside (M1, 630 Hz) and inside (M2, 845 Hz) the emergent frequency band gap (SI Appendix, Fig. S16B) using aluminum and brass as the matrix and resonator materials, respectively (material properties are provided in SI Appendix, Table S3). The metamaterial is excited via transverse z-directional flexural waves at the bottom corner of the array which propagate in the xy plane of the structure, permitted to propagate through the metamaterial at pass-band frequencies but forbidden in the frequency band gap indicated by the shaded region of SI Appendix, Fig. S16A. SI Appendix, Fig. S16A pits the displacement response of the finite metamaterial (top portion; generated from a finite-element model) against the wave dispersion behavior of an infinite array of the constitutive unit cell (bottom portion; predicted via a Bloch-wave band structure). The band gap calculated from the dispersion analysis spans the 759–1,266-Hz range (SI Appendix, Table S2) and provides excellent agreement with the displacement transfer function captured in the frequency response.
Using this passive configuration as a benchmark, the metamaterial’s versatility is then expanded by exploiting the sensitivity of the printed ferroelectric material’s elastic modulus to an externally applied electric field. Different simulated scenarios for electric field strengths of 0, 300, 700, and 2,000 V/cm are depicted in Fig. 4, showing the response and band gap shift at the same measurement locations used in SI Appendix, Fig. S16 (Fig. 4B) and demonstrating the change in the metamaterial’s displacement field at a single fixed frequency of 810 Hz (Fig. 4C). Fig. 4D presents a direct quantitative comparison between the displacement transfer function and the electric-field strength at the three measurement locations, demonstrating an ability to fully control the elastic deformation level in prescribed regions by selectively tuning the external stimulation. Due to the increased stiffness of the soft ferroelectric filler, large electric-field strengths of 700 and 2,000 V/cm drive the band gap to higher frequencies, while little to no electric field strengths of 0 and 300 V/cm yield a lower-frequency band gap for the same geometry and configuration, thus causing an incident excitation of the same frequency to lie both inside and outside the forbidden regime depending on electric-field strength (Fig. 4 C and D and Movie S3). Furthermore, a change in field strength from 0 to 2,000 V/cm instigates a band gap shift from 759–1,266 Hz to 847–1,411 Hz (SI Appendix, Table S2), culminating in an 11.6% increase in center band gap frequency corresponding to the maximum change in electric-field strength.
Movie S3.
Tunable elastic wave dispersion and frequency response in a simulated ferroelectric locally resonant metamaterial. The vibration response of the ferroelectric metamaterial is displayed for an excitation frequency of 810 Hz and simulated electric field strengths of 0, 300, 700, and 2000 V/cm applied around the metamaterial. Movie S3 displays an animated view of the displacement field for each electric field scenario, demonstrating the shift in frequency band gap and animating the metamaterial tunability instigated by the varying applied electric field.
Conclusion
We demonstrate a 3D-printed molecular ferroelectric metamaterial architecture via additive manufacturing, which exhibits optical transparency, ferroelectricity, and a self-healing ability. The ability to restore ferroelectric and electromechanical coupling properties opens up opportunities to greatly enhance the durability and reliability of mechanical metamaterials. The mechanical properties of the printed metamaterial are also shown to be dynamically tunable by an external electric field. Finally, a locally resonant architecture is designed with the printed molecular ferroelectric metamaterial, yielding a tunable range of frequencies within which incident loads are exponentially suppressed. Our study shows that printed molecular ferroelectric metamaterials are of particular interest for mechanical isolation and dispersive wave manipulation under incident dynamic loads, notably without the need for hard-wired setups (35), shunt circuits (36), spinning parts (37), and motor-driven components (38), which cannot be readily scaled to most practical applications.
Materials and Methods
Detailed descriptions of the experimental procedures and analyses are provided in SI Appendix, Materials and Methods. In brief, ImClO4 was synthesized by allowing the imidazole base dissolved in distilled water to react with perchloric acid. UV-vis absorption and transmittance spectra were collected using an Agilent Cary 7000 UV-Visnear-IR spectrophotometer. The crystal structure was characterized by X-ray diffraction (XRD; Rigaku Ultima IV instrument operating with a Cu Ka radiation). The microstructure examination was performed in a field emission gun scanning electron microscope (JSM 7001F, JEOL). Compression experiments for the 3D-printed sample were performed with a Mark-10 universal testing machine. Ferroelectric characterization was performed with a Radiant Ferroelectric Tester Precision LC. The temperature dependence of dielectric constant was measured in a Quantum Design Physical Property Measurement System with a Radiant High Voltage Cryogenic Probe and an Agilent 4294A impedance analyzer. The SLA printing process was performed with a custom-built SLA printer, utilizing a bottom-up configuration.
Data Availability
All study data are included in the article and SI Appendix.
Acknowledgments
Work at the University at Buffalo (S.R.) was supported by the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering, under Award DE-SC0018631 (Physical properties of molecular electronic crystals). The US Army Research Office supports S.R. under Award W911NF-18-2-0202 (Self-assembly of organic crystals). M.N. acknowledges support of this work from the US National Science Foundation through Awards 1904254 and 1847254. M.N. and S.R. acknowledge support from the NY State Center of Excellence in Material Informatics. Y.H. acknowledges support from Yunru Zhang in figure design.
Supporting Information
Appendix (PDF)
- Download
- 1.01 MB
Movie S1.
Overall process of printing a 3D hydrogel structure from precursor through SLA printing. Molecular ferroelectric ImClO4 crystal is dissolved in water and mixed with PEGDA in the volume ratio of 9:1 to make the precursor. The pure precursor is transparent. Thus, we add yellow-colored dye to make the precursor yellow in color for recording the printing process. The ultraviolet (UV) light has a wavelength of 385 nm and the control of image projection was achieved through a dynamic micro-mirror device. The precursor was exposed to the UV light, and the PEGDA was then cross-linked to form the scaffold network with the encapsulated ions. As shown in the video, a complex geometric structure (Schwarz primitive structure, 25×25×25 mm3) can be printed in 8 minutes.
- Download
- 8.94 MB
Movie S2.
Self-healing process. Printed ImClO4 with internal cracks is immersed into ImClO4 solution. The printed ImClO4 expands with the formation of the hydrogel state containing Im+ and ClO4− ions. Well-crystalized printed ImClO4 can be obtained after performing the electric field assistant drying process (Fig. 1C).
- Download
- 12.74 MB
Movie S3.
Tunable elastic wave dispersion and frequency response in a simulated ferroelectric locally resonant metamaterial. The vibration response of the ferroelectric metamaterial is displayed for an excitation frequency of 810 Hz and simulated electric field strengths of 0, 300, 700, and 2000 V/cm applied around the metamaterial. Movie S3 displays an animated view of the displacement field for each electric field scenario, demonstrating the shift in frequency band gap and animating the metamaterial tunability instigated by the varying applied electric field.
- Download
- 25.06 MB
References
1
P.-P. Shi et al., Symmetry breaking in molecular ferroelectrics. Chem. Soc. Rev. 45, 3811–3827 (2016).
2
Z. Zhang et al., Tunable electroresistance and electro-optic effects of transparent molecular ferroelectrics. Sci. Adv. 3, e1701008 (2017).
3
D.-W. Fu et al., Diisopropylammonium bromide is a high-temperature molecular ferroelectric crystal. Science 339, 425–428 (2013).
4
H.-Y. Ye et al., Metal-free three-dimensional perovskite ferroelectrics. Science 361, 151–155 (2018).
5
Y.-M. You et al., An organic-inorganic perovskite ferroelectric with large piezoelectric response. Science 357, 306–309 (2017).
6
W. Y. Kim et al., Graphene-ferroelectric metadevices for nonvolatile memory and reconfigurable logic-gate operations. Nat. Commun. 7, 10429 (2016).
7
W. Zhou, P. Chen, Q. Pan, X. Zhang, B. Chu, Lead-free metamaterials with enormous apparent piezoelectric response. Adv. Mater. 27, 6349–6355 (2015).
8
L. Van Lich, T. Shimada, S. Sepideh, J. Wang, T. Kitamura, Multilevel hysteresis loop engineered with ferroelectric nano-metamaterials. Acta Mater. 125, 202–209 (2017).
9
Z. Liu et al., Locally resonant sonic materials. Science 289, 1734–1736 (2000).
10
M. I. Hussein, M. J. Leamy, M. Ruzzene, Dynamics of phononic materials and structures: Historical origins, recent progress, and future outlook. Appl. Mech. Rev. 66, 040802 (2014).
11
J. Li, C. T. Chan, Double-negative acoustic metamaterial. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 70, 055602 (2004).
12
R. K. Pal, M. Ruzzene, Edge waves in plates with resonators: An elastic analogue of the quantum valley Hall effect. New J. Phys. 19, 025001 (2017).
13
N. Swinteck et al., Bulk elastic waves with unidirectional backscattering-immune topological states in a time-dependent superlattice. J. Appl. Phys. 118, 063103 (2015).
14
Z. Yang et al., Topological acoustics. Phys. Rev. Lett. 114, 114301 (2015).
15
M. Serra-Garcia et al., Observation of a phononic quadrupole topological insulator. Nature 555, 342–345 (2018).
16
R. Chaunsali, F. Li, J. Yang, Stress wave isolation by purely mechanical topological phononic crystals. Sci. Rep. 6, 30662 (2016).
17
R. Fleury, D. L. Sounas, C. F. Sieck, M. R. Haberman, A. Alù, Sound isolation and giant linear nonreciprocity in a compact acoustic circulator. Science 343, 516–519 (2014).
18
B. Liang, B. Yuan, J. C. Cheng, Acoustic diode: Rectification of acoustic energy flux in one-dimensional systems. Phys. Rev. Lett. 103, 104301 (2009).
19
H. Chabok et al., “Ultrasound transducer array fabrication based on additive manufacturing of piezocomposites” in ASME/ISCIE 2012 International Symposium on Flexible Automation, (American Society of Mechanical Engineers Digital Collection, St. Louis, MO, 2012), pp. 433–444.
20
Z. Chen et al., 3D printing of piezoelectric element for energy focusing and ultrasonic sensing. Nano Energy 27, 78–86 (2016).
21
H. Cui et al., Three-dimensional printing of piezoelectric materials with designed anisotropy and directional response. Nat. Mater. 18, 234–241 (2019).
22
Y. Zhang et al., A molecular ferroelectric thin film of imidazolium perchlorate that shows superior electromechanical coupling. Angew. Chem. Int. Ed. Engl. 53, 5064–5068 (2014).
23
H. Ma et al., Ferroelectric polarization switching dynamics and domain growth of triglycine sulfate and imidazolium perchlorate. Adv. Electron. Mater. 2, 1600038 (2016).
24
B. Jiang, X. Pang, B. Li, Z. Lin, Organic–inorganic nanocomposites via placing monodisperse ferroelectric nanocrystals in direct and permanent contact with ferroelectric polymers. J. Am. Chem. Soc. 137, 11760–11767 (2015).
25
B. Jiang et al., Barium titanate at the nanoscale: Controlled synthesis and dielectric and ferroelectric properties. Chem. Soc. Rev. 48, 1194–1228 (2019).
26
Y. Yang et al., Self-healing of electrical damage in polymers using superparamagnetic nanoparticles. Nat. Nanotechnol. 14, 151–155 (2019).
27
L. Gao et al., Autonomous self-healing of electrical degradation in dielectric polymers using in situ electroluminescence. Matter 2, 451–463 (2020).
28
J. Gong et al., Complexation-induced resolution enhancement of 3D-printed hydrogel constructs. Nat. Commun. 11, 1267 (2020).
29
D. Oran et al., 3D nanofabrication by volumetric deposition and controlled shrinkage of patterned scaffolds. Science 362, 1281–1285 (2018).
30
D. Han, Z. Lu, S. A. Chester, H. Lee, Micro 3D printing of a temperature-responsive hydrogel using projection micro-stereolithography. Sci. Rep. 8, 1963 (2018).
31
Y. Yang et al., Recent progress in biomimetic additive manufacturing technology: From materials to functional structures. Adv. Mater. 30, e1706539 (2018).
32
C. Zhou, Y. Chen, Z. Yang, B. Khoshnevis, Digital material fabrication using mask‐image‐projection‐based stereolithography. Rapid Prototyping J. 19, 153–165 (2013).
33
Y. Pan, C. Zhou, Y. Chen, A fast mask projection stereolithography process for fabricating digital models in minutes. J. Manuf. Sci. Eng. 134, 051011 (2012).
34
S. Nemir, H. N. Hayenga, J. L. West, PEGDA hydrogels with patterned elasticity: Novel tools for the study of cell response to substrate rigidity. Biotechnol. Bioeng. 105, 636–644 (2010).
35
G. Trainiti et al., Time-periodic stiffness modulation in elastic metamaterials for selective wave filtering: Theory and experiment. Phys. Rev. Lett. 122, 124301 (2019).
36
Y. Chen, G. Huang, C. Sun, Band gap control in an active elastic metamaterial with negative capacitance piezoelectric shunting. J. Vib. Acoust. 136, 061008 (2014).
37
P. Wang, L. Lu, K. Bertoldi, Topological phononic crystals with one-way elastic edge waves. Phys. Rev. Lett. 115, 104302 (2015).
38
M. Attarzadeh, J. Callanan, M. Nouh, Experimental observation of nonreciprocal waves in a resonant metamaterial beam. Phys. Rev. Appl. 13, 021001 (2020).
Information & Authors
Information
Published in
Classifications
Copyright
© 2020. Published under the PNAS license.
Data Availability
All study data are included in the article and SI Appendix.
Submission history
Published online: October 19, 2020
Published in issue: November 3, 2020
Keywords
Acknowledgments
Work at the University at Buffalo (S.R.) was supported by the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering, under Award DE-SC0018631 (Physical properties of molecular electronic crystals). The US Army Research Office supports S.R. under Award W911NF-18-2-0202 (Self-assembly of organic crystals). M.N. acknowledges support of this work from the US National Science Foundation through Awards 1904254 and 1847254. M.N. and S.R. acknowledge support from the NY State Center of Excellence in Material Informatics. Y.H. acknowledges support from Yunru Zhang in figure design.
Notes
This article is a PNAS Direct Submission.
Authors
Competing Interests
The authors declare no competing interest.
Metrics & Citations
Metrics
Altmetrics
Citations
Cite this article
A 3D-printed molecular ferroelectric metamaterial, Proc. Natl. Acad. Sci. U.S.A.
117 (44) 27204-27210,
https://doi.org/10.1073/pnas.2013934117
(2020).
Copied!
Copying failed.
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.