banner
Casa / Notizia / Vicino
Notizia

Vicino

Mar 31, 2023Mar 31, 2023

Nature Communications volume 13, numero articolo: 3565 (2022) Citare questo articolo

4277 accessi

12 citazioni

Dettagli sulle metriche

Il fattore di accoppiamento elettromeccanico, k, dei materiali piezoelettrici determina l'efficienza di conversione dell'energia meccanica in energia elettrica o dell'energia elettrica in meccanica. Qui, forniamo un approccio fondamentale per progettare materiali piezoelettrici che forniscono una grandezza quasi ideale di k, sfruttando l'anisotropia elettrocristallina attraverso la fabbricazione di ceramiche a grani orientati o strutturate. La simulazione accoppiata del campo di fase e l'indagine sperimentale su ceramiche testurizzate Pb(Mg1/3Nb2/3)O3-Pb(Zr,Ti)O3 mostrano che k può raggiungere la stessa grandezza di quella di un singolo cristallo, ben oltre il valore medio di ceramica tradizionale. Per fornire una comprensione su scala atomistica del nostro approccio, utilizziamo un modello teorico per determinare l'origine fisica di k nei ferroelettrici perovskiti e scopriamo che un forte legame covalente tra il catione del sito B e l'ossigeno tramite ibridazione dp contribuisce maggiormente all'entità di k. Questa dimostrazione di un valore k quasi ideale nelle ceramiche testurizzate avrà un impatto enorme sulla progettazione di dispositivi piezoelettrici a larghezza di banda ultra ampia, alta efficienza, alta densità di potenza e alta stabilità.

I materiali piezoelettrici consentono la conversione elettromeccanica tra energia elettrica ed energia meccanica e viceversa. Sono ampiamente utilizzati in sensori, attuatori, trasduttori, dispositivi di imaging e raccoglitori di energia1,2. Il fattore di accoppiamento elettromeccanico, k, quantifica l'efficacia del materiale piezoelettrico nel fornire la conversione tra energia elettrica ed energia meccanica e viceversa. Il parametro k2 riflette il rapporto tra l'energia meccanica immagazzinata e l'energia elettrica in ingresso, o il rapporto tra l'energia elettrica immagazzinata e l'energia meccanica in ingresso, dato come: k2 = energia meccanica immagazzinata/energia elettrica in ingresso, o k2 = energia elettrica immagazzinata/ energia meccanica in ingresso3. I materiali piezoelettrici con k elevato forniranno un'elevata larghezza di banda massima ottenibile e una densità massima di polvere con elevata efficienza, e quindi k è uno dei parametri più importanti per i dispositivi di trasduzione piezoelettrica3.

Come altro vantaggio particolare, la progettazione dei materiali piezoelettrici con k elevato può fornire un approccio alternativo per aumentare il coefficiente piezoelettrico d secondo la relazione: \(d=k\sqrt{s\cdot \varepsilon }\), dove s è la cedevolezza elastica e ε è la permettività dielettrica3. Questo approccio supera diversi colli di bottiglia dell’approccio tradizionale tramite la progettazione della composizione. Con l'approccio tradizionale, il miglioramento della risposta piezoelettrica d si ottiene aumentando la permettività dielettrica ε in base all'espressione d = 2QPsε, dove Ps è la polarizzazione spontanea, Q è il coefficiente elettrostrittivo e ε è la permettività dielettrica4. L'aumento di ε si realizza appiattendo il panorama dell'energia libera rispetto alla polarizzazione (abbassando la barriera energetica per la rotazione della polarizzazione ferroelettrica), in particolare, mediante la progettazione della coesistenza multifase indotta dalla composizione (come il confine di fase morfotropico e la transizione di fase polimorfica) 5,6 o progettare strutture compositive locali (come l'ordinamento a corto raggio su scala nanometrica7 e l'eterogeneità strutturale locale8,9). I colli di bottiglia associati a questo approccio tradizionale includono: (1) La proprietà piezoelettrica d migliorata è ottenuta a scapito della stabilità della temperatura (temperatura di depolarizzazione inferiore Td o bassa temperatura di Curie Tc), che segue un andamento dato come: d ∝ 1/T1, 10,11; (2) L'aumento di ε diminuisce il coefficiente di tensione piezoelettrica g, con conseguente riduzione della sensibilità come sensore piezoelettrico12; (3) Questo approccio non è efficace nel migliorare la densità energetica dei materiali piezoelettrici, caratterizzati da d·g13. Ad esempio, anche se il d33 dei ceramici casuali Pb(Mg1/3Nb2/3)O3-PbTiO3 (abbreviati come PMN-PT) drogati con Sm può raggiungere 1500 pC N−1 aumentando ε33 a 13.000, d33·g33 è limitato a 19,8 × 10−12 m2 N−1, che rappresenta solo 1/3 del valore delle ceramiche testurizzate Pb(Mg1/3Nb2/3)O3-Pb(Zr,Ti)O3 (abbreviato come PMN-PZT)9,13 ; (4) L'aumento di d non può dar luogo ad un incremento significativo di k, che è uno dei fattori più importanti per i dispositivi piezoelettrici come menzionato sopra; (5) Aumentando la permettività dielettrica ε appiattendo il panorama dell'energia libera generalmente si riduce il campo coercitivo Ec, che indebolisce la stabilità del campo elettrico (depoling) e limita l'uso di materiale in applicazioni ad alta potenza. Qui mostriamo che queste sfide possono essere superate progettando materiali piezoelettrici con k elevato.

oriented textured piezoelectrics can have higher k than those of their random counterparts13,18,19. This provides us initial direction towards addressing several questions: what is the maximum value of k that can be achieved in textured ceramic and is it possible to obtain the same or even higher k in textured ceramics than the values in their single-crystal counterparts? Does the neighboring grain correlation in textured ceramic limit the maximum achievable k? Does this approach of increasing k by microstructure texturing show the advantages over the traditional approach of increasing dielectric permittivity ε by composition design as mentioned above? In order to answer these questions, we carried out phase-field simulation to investigate the effects of crystallographic orientation and grain boundary in textured ceramics on the electromechanical coupling factor k. Results from the simulation were experimentally verified to confirm the increase in k in highly <001> textured PMN-PZT ceramics. A theoretical model is developed to gain an understanding of the physical origin of electromechanical coupling in perovskite ferroelectrics and determine the key correlations./p> 0.9) was observed for relaxor-PT single crystals, and the magnitude of k is orientation-dependent20. Figure 1c shows the simulated k for the rhombohedral PMN-PT single crystal under the electric field with various orientations. The corresponding simulated permittivity εT and εS are shown in Supplementary Fig. 1. Based on the microscopic model, the magnitude of k depends on the competition between chemical energy and elastic energy. Since both energies are polarization-dependent and anisotropic, k must be also anisotropic and thus depend on the electric field orientation. Because of the strong strain constraint, the change of εS is almost negligible in comparison with that of εT. Therefore, the anisotropic behavior of k will be mainly determined by εT. Since the polarization of PMN-PT behaves as a rotator rather than an extender21, the permittivity and thus k will be higher if the angle between electric field and polarization is larger. As shown in Fig. 1c, the highest k occurs when the electric field is perpendicular to polarization, while the lowest k occurs when the electric field is parallel with polarization, which suggests that k15 mode possesses the largest value of k in rhombohedral relaxor-PT single crystals22. However, upon electrical poling, only the values of k with θ ∈ [0,90°] are allowed, indicated by the solid blue line in Fig. 1c. Among those values, the highest k occurs when the electric field is along [001] direction, i.e., [001]-poled single crystal can exhibit the highest k33. In polycrystal ceramics, however, since the grain orientations are randomly distributed, the highest k33 can't be obtained as in single crystals. Nevertheless, the [001]-textured polycrystal can help to realize the highest k33 in ceramics. For a [001]-textured polycrystal, all grains are oriented in [001] crystallographic axis while the other crystallographic axes are completely random. Even though the [001]-textured polycrystal can possess the highest k33 in ceramics, can it be as large as that of a single crystal?/p>-textured ceramic can possess a large k that is comparable to the value of the single crystal, highly <001>-textured PMN-PZT ceramic was prepared by templated grain growth technique with a different volume percentage of BaTiO3 templates. Figure 3a shows the X-ray diffraction patterns for random PMN-PZT ceramic with 0 vol% BaTiO3 templates and textured PMN-PZT ceramic with 3 vol% BaTiO3 templates, respectively. Both samples exhibit a perovskite phase, while textured ceramic shows a remarkable enhancement in the intensities of the {001} diffraction peaks compared to random ceramic. The Lotgering factor of the textured sample is over 98%, indicating a strong [001] preferred grain orientation. Figure 3b shows the grain orientation of random and textured PMN-PZT ceramics via inverse pole figure (IPF) maps measured by the SEM-EBSD technique along thickness (Z) direction (The IPF-X and IPF-Y maps are shown in Supplementary Fig. 3). In order to evaluate the electromechanical coupling, longitudinal 33 mode and transverse 31 mode samples with dimensions according to IEEE standards were prepared, and their impedance spectra are shown in Fig. 3c, d. The k33 and k31 of <001> textured PMN-PZT are surprisingly as high as 0.93 and 0.65, respectively. Figure 3e lists the k33 of representative [001] oriented single crystals and polycrystalline random ceramics. The k33 of <001> single crystals are in the range of 0.90–0.94, while the k33 of random ceramics are limited to below 0.80. The k33 of <001> textured PMN-PZT is the same as the value of PMN-PZT single crystal counterpart. Figure 3f lists the k31 of representative [001] oriented single crystals and polycrystalline random ceramics. In general, the k31 of random ceramics are in the range of 0.30–0.40. The k31 and k31(45o) of [001] oriented single crystals are about 0.43 and 0.80, respectively. Here k31(45o) is the k31 of [001] oriented single crystal sample with 45o cut, where the orientations of the sides are [110], [\(\bar{1}10\)] and [001], respectively. The k31 in textured PMN-PZT ceramic is 0.65, which is slightly higher than the average value of k31 and k31(45o) of [001] oriented single crystals. This characteristic is due to the distribution of grain orientation in <001> textured ceramics, where the [001]-orientation of grains in textured samples are well aligned along the thickness direction (z, out of casting plane) but the [100] and [010] orientations of grains in textured samples are randomly distributed in the casting plane, which is related to the fact that unidirectional shear force was used for aligning the templates. Based on these observations, it can be suggested that the k, either k31 or k33, in textured ceramic is solely dependent on the texture direction and texture degree, and is not limited by the existence of the grain boundary. The <001>-textured ceramic can possess a large k33 that is comparable to those of the <001>-oriented single crystals, which verifies the prediction in Figs. 1 and 2./p> textured PMN-PZT ceramics. c Impedance spectra of <001> textured PMN-PZT ceramics in longitudinal 33 mode, the electromechanical coupling factor, k33, is as high as 0.93. d Impedance spectra of <001> textured PMN-PZT ceramics in transverse 31 mode, the electromechanical coupling factor, k31, is as high as 0.65. e Comparison of k33 among [001] oriented single crystals, random ceramics and <001> textured ceramics. f Comparison of k31 among [001] oriented single crystals, random ceramics and <001> textured ceramics. Here k31(45o) is the k31 of [001] oriented single crystal sample with 45o cut, where the orientations of the sides are [110], [\(\bar{1}10\)] and [001], respectively. The references for k33 and k31 data for single crystals and ceramics in e, f can be found in Supplementary Note 1./p>-textured ceramics suggests that they may possess similar domain configurations. It is well known that the ability of domains to switch in ferroelectric polycrystals depends critically on the crystallographic symmetry of the ferroelectric phase23. The domains in random polycrystal ceramics that are either tetragonal or rhombohedral are difficult to switch due to the constrain by the differently oriented neighboring grains. <001> texture is requisite for non-180o domain switching in tetragonal phase23. Figure 4 shows the in-situ electric field XRD patterns of <001>-textured PMN-PZT-3BT ceramics. It can be observed that the unpoled textured sample has MPB composition with the coexistence of rhombohedral and tetragonal phases (Fig. 4a), characterized by peak splitting near 44o. With increasing the electric field (1st up), the high angle peak (a-axis) is merged into the low angle peak (c-axis) and becomes a single peak, indicating that the a-domains in the tetragonal phase can be fully switched to c-domains. This new domain structure is very stable during the removal of the electric field (1st down, Fig. 4b) and application of electric field (2nd up, Fig. 4c). Based on these observations, it can be suggested that the non-180o domains in <001> textured ceramics are switchable, further confirming that <001>-textured ceramics could have the same electromechanical coupling factor k as [001]-oriented single crystal./p>0.9) prefers the strong covalent bonding effect rather than the ionic effect. These findings suggest an optimal condition for perovskite ferroelectrics that simultaneously possess a large permittivity, large piezoelectricity, and a high electromechanical coupling factor: strong B-O covalent bonding and close to the phase transition boundary. Similar studies based on experimental measurements to analyze the effects of A-site or B-site ions on the electromechanical coupling factor have been reported by Yamashita27,28,29,30./p> textured PMN-PZT is measured to be 0.93, which closes the gap between the ceramic and single crystal. In addition, increasing k via texturing provides an alternative approach of increasing the piezoelectricity, which overcomes the bottlenecks of the traditional approach via composition design. Further, we employed a theoretical model to understand the physical origin of k in perovskite ferroelectrics and found that strong covalent bonding between B-cation and oxygen via d-p hybridization contributes most to k. These findings provide a novel design strategy to develop the next generation of high-performance piezoelectric materials with ultrahigh piezoelectricity and low cost, to fulfill the demands for ultra-wide bandwidth, high efficiency, high power density, and high stability piezoelectric devices./p>