Abstract
Van der Waals integration of freestanding perovskite-oxide membranes with two-dimensional semiconductors has emerged as a promising strategy for developing high-performance electronics, such as field-effect transistors. In these innovative field-effect transistors, the oxide membranes have primarily functioned as dielectric layers, yet their great potential for structural tunability remains largely untapped. Free of epitaxial constraints by the substrate, these freestanding membranes exhibit remarkable structural tunability, providing a unique material system to achieve huge strain gradients and pronounced flexoelectric effects. Here, by harnessing the excellent structural tunability of PbTiO3 membranes and modulating the underlying substrate’s elasticity, we demonstrate the tip-pressure-induced polarization switching with an ultralow pressure (down to 0.06 GPa). Moreover, as an application demonstration, we develop a prototype non-volatile ferroelectric field-effect transistor integrated on silicon that can be operated mechanically and electrically. Our findings underscore the great potential of oxide membranes for utilization in advanced non-volatile electronics and highly sensitive pressure sensors.
Subject terms: Ferroelectrics and multiferroics; Surfaces, interfaces and thin films
The authors realize low-pressure-driven polarization switching in PbTiO3 membranes by leveraging their structural tunability and substrate elasticity, enabling ferroelectric field-effect transistors on silicon, operatable mechanically and electrically.
Introduction
Recently, the rich properties and diverse applications of freestanding oxide membranes have attracted great attention1–3. By breaking the chemical bond between films and substrates, oxide membranes exhibit great structural tunablility, achieving huge strain and strain gradient far exceeding the range which can be reached in bulk materials and epitaxial films3–7. Additionally, the freestanding membranes can also be freely transferred onto any desired substrates to construct novel interfaces and devices. For example, high-density polar nanodomains8, tunable ferroelectric domain walls9, ferroelectric tunnel junctions10,11, were demonstrated by integrating ferroelectric membranes with silicon wafers. Excitingly, van der Waals integration of perovskite-oxide membranes and two-dimensional semiconductors has become a promising strategy to construct high-performance field-effect transistors (FETs)12–14, showing a high on/off current ratio, a near-ideal subthreshold swing value, and more. In these FETs, freestanding oxide membranes have primarily used as dielectric layers due to their excellent dielectric or ferroelectric properties12,13, yet their remarkable structural tunability remains largely untapped. The great structural tunability in membranes should enable effective mechanical switching of the polarization and ferroelectric domains via flexoelectric effect, providing promising routes to design novel mechanically-operated electronic devices.
Flexoelectricity is an electromechanical coupling between polarization and strain gradient, which enables the mechanical control of ferroelectric polarization. The flexoelectric effect finds wide applications in flexoelectronics15–17, sensing and actuating18,19, data storage20,21, and energy harvesting22. Although flexoelectric effect is a weak effect of little practical significance in bulk materials, its role can be more pronounced in thin films23–25, especially in freestanding oxide membranes and in van der Waals layered materials26–28 due to their remarkable structural tunability, superior elasticity and flexibility2,3,5. By bending the membranes, giant strain gradients and flexoelectric response has been reported in BiFeO36,29 and ferroelectric 2D material CuInP2S626–28,30. However, whether such pronounced flexoelectric effects can be utilized in realizing low-pressure-mechanical polarization switching and applied in designing non-volatile FeFETs remains to be explored.
In this work, benefiting from the excellent structural tunability of freestanding PbTiO3 membranes and the adjustment of underlying substrate’s elasticity, we achieved ultralow pressure (down to 0.06 GPa) for mechanical polarization switching. Moreover, as an application demonstration, we developed a prototype non-volatile ferroelectric field-effect transistor (FeFET) on silicon that can be operated mechanically and electrically for the first time. By constructing freestanding PbTiO3/MoS2 heterostructures, tip-pressure-induced flexoelectric effect can switch the polarization state in PbTiO3 layer and thus further influence the transport behavior in the MoS2 channel.
Results
Design for low-pressure-driven polarization switching
To achieve low-pressure-driven polarization switching, we propose to tune the energy barrier height and potential profile by utilizing the excellent structural tunability in freestanding membranes and optimizing the elasticity of the underlying substrate. Our experimental design is illustrated in Fig. 1b. In case I, PTO membranes are epitaxially grown on STO substrates, where the double well (blue line) in the energy profile becomes asymmetric (red line) when incorporating flexoelectricity into Landau free energy28. Thus, the upward polarization state is destabilized and can be switched to the downward polarization state when the flexoelectric field is strong enough by applying large tip pressure to bend the film and substrate. In case II, freestanding membranes are released from the original substrate and transferred onto a rigid oxide substrate. In this case, the in-plane lattice parameters (a and b) are free to relax since there is no chemical bonding between the membrane and substrate. As shown in the theoretical calculations (Fig. 1c), allowing relaxation of a and b lattice constants can apparently lower the polarization switching energy barrier even fixing the switching path to be 180° polarization switching. Effective Hamiltonian calculations also show that this energy barrier can be further lowered by allowing the polarization in freestanding membranes to tilt and flip to the in-plane direction as an intermediate step31, which means the 180° switching consists of two 90° rotations. In fact, similar two-step polarization switching has been commonly observed in many ferroelectrics, including BiFeO320, Pb(Zr0.2Ti0.8)O332, Pb(Zr0.1Ti0.9)O333, etc. In case III, we further manipulate the substrate elasticity, by transferring PTO membranes on a softer substrate. Thus, the tip-pressure will cause more obvious deformation in membranes, which leads to a larger strain gradient and a more pronounced flexoelectric effect. Although the energy barrier height is the same as it is in case II under homogenous strain (blue solid and dotted lines), a larger strain gradient would cause a more inequivalent energy double-well (red solid and dotted lines), which maximizes the mechanical switching efficiency. Finite-element simulation (Fig. 1d) shows that vertical flexoelectric field is widely distributed in nearly the entire region beneath the tip in case III, also indicating a more pronounced flexoelectric effect when transferring membranes on soft substrate. During the FEM calculation, for the epitaxial PTO film (case I), the displacement continuity boundary was adopted due to the strong covalent bonding in PTO/LNO interfacial layer. the Young’s modulus and the Poisson’s ratio of LNO were set as ELNO = 299 GPa34 and νLNO = 0.2734, respectively. For freestanding PTO membrane (cases II and III), the contact boundary and Coulomb friction model were adopted to simulate the possible sliding at interface. The underlying substrates in cases II and III are LNO and Au, respectively. The Young’s modulus and the Poisson’s ratio of Au were set as EAu = 78 GPa35 and νAu = 0.4235. Details of theoretical calculations can be found in Methods.
Experimental demonstration of ultralow-pressure-induced polarization switching
By using reactive molecular beam epitaxy, high-quality PTO films were epitaxially grown on (001) SrTiO3 (STO) substrates with 8 nm thick LaNiO3 (LNO) as a bottom electrode denoted as Epi(27 nm)/LNO(8 nm)/STO. PTO membranes (12, 18, 27 nm) were grown on (001) SrTiO3 (STO) substrates with a Sr3Al2O6 (SAO) sacrificial layer. Subsequently, freestanding PTO membranes were prepared by dissolving SAO and then transferred onto LNO/STO substrates (denoted as FS | | LNO/STO), platinum-coated silicon substrate (denoted as FS | |Pt/Si) and gold-coated silicon substrate (denoted as FS | |Au/Si). Schematic images of these samples are shown in Fig. 2a. The spring constant and inverse optical lever sensitivity of AFM tip were calibrated using the force curve and thermal noise method for obtaining the accurate tip force28,36–40. More details about the membrane growth, transfer, and characterizations can be found in the Methods and Fig. S2.
Piezoresponse force microscopy (PFM) measurements demonstrate that both the epitaxial and freestanding PTO membranes exhibit upward spontaneous polarization. The vertical-PFM (VPFM) phase images of Epi/LNO/STO and FS | |Au/Si are shown in Figs. S3d, h, and the polarization state of the region polarized with −10 V sample bias is similar to that of the as-grown region, implying that the latter has an upward self-polarization41. During mechanical writing process, the downward polarization grows gradually at the expense of the upward one with the increasing flexoelectric field with larger mechanical load pressure (Figs. 2b and S4). Here, the scanning mode is used in our experiments to induce polarization switching by AFM tip. The percentage of the switching area as a function of the tip pressure is shown in Fig. 2c. The threshold tip pressure Pth becomes lower in FS||LNO/STO sample (1.83 GPa) than that in Epi/LNO/STO sample (2.14 GPa), which demonstrates that the “freestanding” state indeed lowers the switching barrier since it enables greater freedom in lattice deformation. As illustrated in Fig. 1, in freestanding membranes, the relaxation of in-plane lattice parameters (a and b) can lower the switching barrier and this energy barrier can be further decreased when polarization is tilted and flipped to in-plane direction as an intermediate step. The two-step switching is confirmed by LPFM amplitude measurements as shown in Fig. S1. The LPFM amplitude of the data taken under the intermediate pressure shows the maximum value, indicating the polarization is lying along the in-plane directions. In contrast, the initial (Pup) and final (Pdown) states exhibit much weaker LPFM amplitudes, which is consistent with the two-step model.
Next, we investigate the effect of substrate elasticity on tip-induced polarization switching. For freestanding membranes transferred to different target substrates, Pth is significantly different with the lowest value (0.54 GPa) found in FS||Au/Si sample. Substrate elasticity is a contributing factor to this phenomenon, as the Young’s modulus of LNO, platinum, and gold are 29934, 16842, and 78 GPa35, respectively. This demonstrates that mechanical polarization switching is more efficient when transferring membranes onto a substrate with the lowest Young’s modulus, which means that modulating the underlying substrate’s elasticity can enhance the flexoelectric field significantly. To exclude the potential contribution of metal work function43–45, we performed contrast experiments by inserting monolayer MoS2 into the oxide/metal interfaces. As shown in Fig. S5, Pth in FS | |Au(Pt)/Si and FS | |MoS2/Au(Pt)/Si are the same, thus the existence of MoS2 barely influences mechanical polarization switching behavior. In addition, the gold thickness dependence of the threshold tip pressure in FS | |Au/Si samples is obviously shown in Fig. 2d. Pth decreases with increasing gold thickness and finally reaches its minimum for dAu = 100 nm. This is because the overall substrate elasticity decreases with the increasing thickness of gold coating layer (the Young’s modulus of silicon is 160 GPa46, which is larger than gold) and the Young’s modulus for Au film decreased with increasing film thickness47. This phenomenon further proves that mechanical writing pressure is affected by the substrate elastic properties other than metal work function. Besides, we also investigated the effect of probe size on mechanical writing behavior (Fig. S6). Fig. S6a, b show that the Pth is 2.96 GPa when AFM tip radius is 10 nm, which is much higher than that written by an AFM tip with 25 nm radius (0.54 GPa). This is mainly due to the highly localized flexoelectric field when the tip radius is 10 nm, which may fail to switch the polarization completely along the thickness direction of the film, as constrained by the long-range Coulomb interactions between electric dipoles48. Our FEM simulations (Figs. S6c, d) also prove that vertical flexoelectric field is confined within a small spatial region along the thickness direction (Rtip = 10 nm), however, it is widely distributed into nearly the entire region when the tip radius is 25 nm.
We compare the Pth in our samples with other ferroelectric perovskite oxide membranes from previous works (Figs. 3a and S7), showing that the required tip pressure is getting lower as the membrane thickness decreases. Notably, an ultralow switching pressure down to 0.06 GPa is achieved in 12 nm freestanding PTO membranes transferred on gold (100 nm)-coated silicon, which is the lowest value ever reported36,38,49–54, even lower than those with thinner thickness. From the view of device applications, high-quality ultrathin membranes are technically demanding in fabrication and integration32, thus it will be more attractive to achieve mechanical domain switching in relatively thicker membranes. However, for epitaxial films constrained by hard substrates, it is challenging to mechanically switch the polarization in thicker films due to the contribution of tip-induced flexoelectricity decreases substantially when the film thickness is above ~25 nm32,33,38,55. Here, we successfully achieve low switching pressure (0.54 GPa) in 27 nm freestanding PTO membranes transferred on gold (100 nm)-coated silicon (FS(27 nm)||Au(100 nm)/Si), showing three times smaller than that in epitaxial membrane with the same thickness.
Reversible polarization switching can be achieved by mechanical writing and electrical erasing in FS(27 nm)||Au(100 nm)/Si, as shown in Fig. 3b. The initial downward polarization state (the yellow region) can be switched to the upward state (the violet region) by applying the negative voltage of −8 V (Fig. 3b, left). Next, this upward polarization poled by electric field can be mechanically written to the downward state (the yellow region) (Fig. 3b, middle). Then, to test the repeatability, a negative voltage (−8 V) reverses the polarization to upward again (the violet region)(Fig. 3b right). Moreover, the mechanical polarization switching region is stable and it barely evolves over time (Fig. S8). Scanning kelvin probe force microscopy (SKPFM) is performed to investigate the surface potential of polarization poled by electricity and loading force (Fig. 3c). Section profiles of the SKPFM images demonstrate that the surface potential of electrically poled region (ΔE1) is higher than that by tip pressure (ΔE2), indicating charge injection is inevitable during the electrical switching process36,53.
Flexoelectricity-tuned ferroelectric field-effect transistors
Next, we explore transferrable PTO membranes as a mechanically-tuned ferroelectric gate dielectric for two-dimensional (2D) field-effect transistors, as a preliminary attempt. In ferroelectric field-effect transistors (FeFETs), the ferroelectric polarization state can be tuned by tip pressure via flexoelectricity and detected by the conductance of the channel material. Meanwhile, 2D materials such as MoS2 have been widely used as channel materials due to their atomic thickness, absence of dangling bonds, and potentially high mobility56,57. Thus, integrating PTO with MoS2 offers a promising route for realizing low-power, high-speed operation and reconfigurable functionalities devices.
To fabricate MoS2 transistors with PTO bottom-gate dielectrics, we sequentially transferred PTO, MoS2, and hBN on Au/SiO2/Si substrates (Figs. 4a, b and S9). PTO membrane with a thickness of 27 nm was transferred onto an Au/SiO2/Si substrate. A monolayer MoS2 thin film grown by chemical vapor deposition (CVD) was then transferred on the top of the freestanding PTO membrane. The shape of the MoS2 channel was then defined by electron beam lithography (EBL) and subsequent reactive ion etching (RIE). Next, source and drain electrodes (Ti/Au) were deposited at the end of MoS2 channels. Finally, devices were encapsulated by hBN to avoid sample degradation and more importantly, it can protect MoS2 from damage during mechanically writing. In our experiments, we can effectively avoid electron doping during the fabricating process of MoS2 transistors. The method of releasing monolayer MoS2 from a sapphire is shown in Fig. S10.
Schematic of a bottom-gate MoS2 FET and optical images of the device can be seen in Fig. 4a, b, respectively. PFM results (Fig. 4c) indicate that the PTO polarization state can be mechanically switched by AFM tip even across hBN and MoS2 layers. Vertical PFM signal of PTO region covered by MoS2 is obviously weaker than that without MoS2 capping layer (Fig. S11), due to the screening effect of semiconductor MoS258. In addition, a relatively thick hBN layer (13 nm) on the top of MoS2 also weakens the flexoelectric effect. These might be the reasons for incomplete polarization reversal. In flexoelectricity-tuned MoS2 transistors, the n-type MoS2 channel’s charge state is controlled by an underlying PTO membrane. Tip-induced flexoelectricity can switch the polarization state of the underlying PTO membrane and then a scanning of the drain current (ID) can read the state of the polarization since different states are associated with distinct transport behaviors. Specifically, the initial polarization state of PTO region under the channel area is uniformly points upward, which accumulates electrons into MoS2 and the ID is in a high current state (‘on’ state). Mechanically switching the polarization of PTO to Pdown state depletes electrons from MoS2, leading to a low current state of ID (‘off’ state). Fig. 4d shows the room temperature drain current (ID) vs voltage (VDS) relations of the device, which are obtained when the PTO membrane at initial Pup state and after mechanical writing to Pdown states at zero back-gate voltage (VGS = 0 V), respectively. At VDS = 0.1 V, ID is suppressed, changing from 53 pA in the Pup state to 3.5 pA in the Pdown state, indicating that mechanically-tuned on/off state is achieved effectively in our prototype FeFET device.
Discussion
While ultralow pressure mechanical polarization switching has been realized and flexoelectricity-tuned FeFET has been demonstrated, there are still some challenges to be solved in the near future. At present, for the prototype flexoelectricity-tuned FeFETs, a relatively high mechanical pressure (1.8 GPa) is needed. This is because an encapsulating hBN with a thickness of up to 13 nm is required to protect the underlying MoS2 due to the poor mechanical durability of the monolayer MoS259. Replacing MoS2 with other suitable 2D channel materials that have better mechanical properties should lower the required pressure and will be explored in the future. Moreover, reversible polarization switching still relies on a combination of a mechanical and an electrical switching process so far. Reversible mechanical switching60 has been demonstrated in epitaxial membranes by manipulating the competition between the flexoelectric field and effective dipolar field delicately61. Following the same mechanism, using a bilayer membrane as the ferroelectric layer, reversible mechanical polarization switching should be applied in membrane-based FeFETs.
In summary, we report an ultralow-pressure-driven polarization switching in freestanding PTO membranes. Without epitaxial constraints, the greater freedom in lattice deformation lowers the energy barrier for polarization switching. Besides, tip-pressure can be further reduced by tuning the substrate elasticity. Finally, a prototype non-volatile flexoelectricity-tuned FeFET has been fabricated by integrating PTO with MoS2 and the ‘on’ and ‘off’ states can be controlled by mechanical switching polarization of the freestanding PTO membrane. Mechanical switching with ultralow pressure can effectively avoid sample damage, and benefit low-energy-consumption and high-sensitivity devices. Our work demonstrates great potential in combining ferroelectric memebranes with 2D materials for novel electronic applications, and paves the way for stress-modulated devices with low energy-consumption or high sensitivity.
Methods
Density functional theory (DFT) calculations
The nudged elastic band (NEB) method62,63 is used to estimate the ferroelectric switching energy barrier, which is implemented in the Vienna ab initio simulation package code64,65. The projected augmented wave method66,67 and the Perdew-Burke-Ernzerhof revised for solids exchange correlation68 are used in the calculations. The plane-wave cutoff energy is 520 eV and the k-mesh is 9 × 9 × 8. The geometry structure is relaxed until the force on each atom is less than 0.01 eV Å−1, and the energy convergence criterion is 10−7 eV. As for PTO epitaxially grown on STO substrate, PTO in-plane lattice parameters are fixed (a = b = 3.877 Å) and out-of-plane lattice parameter (c) is free to relax. In freestanding membranes, the lattice constants of a, b, and c are all allowed to relax.
The flexoelectric calculations
The strain and strain gradient distributions in PTO films were calculated using finite element method (FEM). The model of an AFM tip in contact with a 12 nm-thick PTO film on 100 nm-thick substrates was built for FEM calculations. The radius of AFM tip was set as 25 nm. A vertical force of 2 μN was applied on the top surface of tip. Different mechanical boundary conditions at interface were adopted in the epitaxial and freestanding films. For the epitaxial films, the displacement continuity boundary was adopted due to the strong covalent bonding in PTO/LNO interfacial layer. For the freestanding films on different metal-coated substrates, the contact boundary and Coulomb friction model were adopted to simulate the possible sliding at interface. The Young’s modulus and Poisson’s ratio of PTO, LNO, Au were set as EPTO = 142 GPa69, ELNO = 299 GPa34, EAu = 78 GPa35, νPTO = 0.2969, νLNO = 0.2734, νAu = 0.4235, respectively. The vertical flexoelectric field can be calculated as
1 |
where the transverse, longitudinal and shear flexoelectric coupling coefficients were set as f3113 = 5 V, f3333 = 2 V, f3131 = 2 V, respectively70. The flexoelectric field was then incorporated into the Landau free energy to obtain the double-potential well curves.
Deposition of epitaxial thin films
The water-soluble SAO layer was grown first on (001) STO single-crystalline substrate followed by the growth of a thin film PTO by reactive MBE system. The SAO films were grown with an oxidant (10% O3 and 90% O2) background partial pressure of oxygen pO2 of 1 × 10−6 Torr and at a substrate temperature Tsubstrate = 950°C. The SAO and PTO films were grown layer by layer, for each of which the thickness was monitored by reflection high-energy electron diffraction (RHEED) oscillations. The PTO films were grown with an oxidant (distilled O3) background pressure of 2×10−5 Torr and at Tsubstrate = 625 °C. Due to the volatility of lead, PTO films were grown in adsorption-controlled mode with a fixed Pb:Ti flux ratio of 13:1 and the thickness was controlled by deposition time of titanium. The perovskite LaNiO3 (LNO) film were grown first on (001) STO single-crystalline substrate followed by the growth of PTO film. The LNO films were grown on the TiO2-terminated SrTiO3 (001) substrates by reactive MBE using a DCA R450 MBE system, with the substrate temperature of 600 °C and an oxidant background pressure of 1×10−5 Torr using the distilled ozone.
Release and transfer PTO films
Freestanding PTO films were directly transferred to desired substrate without any mechanical support from polydimethylsiloxane (PDMS) or other tapes during the whole process, which avoids chemical residue on the film surface.
Structural characterization
The crystal structure was examined by a high-resolution single-crystal X-ray diffractometer using a Bruker D8 Discover instrument. The incident X-ray was from Cu-Kα emission and had a wavelength of 1.5418 Å.
Piezoresponse force microscopy measurements
The local ferroelectric properties were measured on freestanding films on a conducting Si wafer using an Asylum Research MFP-3D Origin+ SPM scanning probe microscope. Olympus AC240TM Pt/Ti-coated tip (2 N m−1 in force constant) were used in the PFM measurements. Hysteresis loops were collected in the dual AC resonance tracking (DART) mode. PFM images were taken in the DART mode with driving voltage (1 V a.c.) applied at the tip. During the domain writing, the voltage was also applied at the tip. The corresponding pressure can be estimated with a tip radius of 25 nm (Olympus AC240TM) under the assumption that the tip-sample contact is represented by a disc with radius equivalent to the tip radius36,51,60,71.
CVD growth and transfer of monolayer MoS2
We grew single-crystal monolayer n-type MoS2 on c/a sapphire (0001) substrates, with αa ranging from 0.2° to 1.0° under 1000 °C annealing for 4 h. To transfer the CVD MoS2 film off the sapphire, PMMA was spin coated at a speed of 600 r.p.m. for 10 s and 1000 r.p.m. for 45 s and then cured at 100 °C for 2 min, followed by lamination of PDMS. Next, the sample was immersed in deionized water and the whole stack was spontaneously delaminated from the sapphire and transferred onto a target substrate. Finally, the PDMS/PMMA can be removed by soaking in acetone.
Device fabrication
On a silicon substrate with 300 nm SiO2, a back-gate was defined by electron-beam lithography (EBL), and 3 nm Ti/60 nm Au was deposited by an electron-beam evaporator. Freestanding PTO film with 27 nm is then transferred on it. Next, assisted by PDMS/PMMA, the MoS2 film was transferred to the freestanding PTO film. Then, the unnecessary MoS2 is etched by using SF6 and O2 plasma in a reactive ion etcher. The last step of EBL defined the S/D pattern, and 10 nm Ti/35 nm Pd/10 nm Ti were deposited by the electron-beam evaporator. Finally, an hBN flake with 13 nm picked up and put down onto the top of the device.
Electrical measurements
Electrical characterization of the FETs was performed using a Keithley 4200 semiconductor parameter analyzer and a Keithley 2450 source meter in a closed-cycle probe station with base pressure ~ 10−6 Torr.
Supplementary information
Acknowledgements
This work is supported by the National Basic Research (Key R&D) Program of China (Nos. 2022YFA1402502 (Y.F.N.), 2022YFB4400100 (X.W.), 2021YFA1400300 (J.W.H.) and 2021YFA1400400 (Y.F.N.), 2023YFF1500500 (X.W.), 2022YFA1402500 (W.L.) and 2021YFA0715600 (W.L.)), the National Natural Science Foundation of China (Grant No. 12434002 (Y.F.N.), 12172047 (J.W.H.), T2221003 (X.W.), 61927808 (X.W.), 52302181 (L.H.), 12402183 (Y.Z.L.), 62322408 (T.L.), 62204113 (T.L.) and 62304101 (W.L.)), Beijing Natural Science Foundation (Grant No. 1244057 (Y.Z.L.)), China National Postdoctoral Program for Innovative Talents (Grant No. BX20220147 (L.H.)), China Postdoctoral Science Foundation (Grant No. 2023M731610 (L.H.)), the Xiaomi foundation (L.H.), the Leading-edge Technology Program of Jiangsu Natural Science Foundation (Grant No. BK20232024 (X.W.) and BK20232001 (Y.S.)), the Natural Science Foundation of Jiangsu Province (Grant No. BK20220773 (T.L.) and BK20230776 (W.L.)) and the Jiangsu Province Key R&D Program (Grant No. BE2023009-3 (W.L.)). The numerical calculations in this paper have been done on the computing facilities in the High Performance Computing Center (HPCC) of Nanjing University (J.Z.). X.W. acknowledges the support by the New Cornerstone Science Foundation through the XPLORER PRIZE. The authors would like to express sincere gratitude to the Interdisciplinary Research Center for Future Intelligent Chips (Chip-X) and Yachen Foundation for their invaluable support.
Author contributions
Y.F.N. conceived the idea and directed the project. X.R.Y. and L.H. synthesized the samples and characterized the crystalline structure with the help of Y.Y.L. under the supervision of Y.F.N. and Z.B.G.; X.R.Y. and L.H. performed the PFM measurements and data analysis under the supervision of Y.F.N. and D.W.; S.Q.X. performed finite-element simulation with the help of Y.Z.L. under the supervision of J.W.H.; Y.C.L. performed DFT calculations under the supervision of J.Z.; X.R.Y. and L.H. fabricate devices with the help of B.H., S.J.Y., H.K.N., Y.G. T.T.L., C.Y.G. and W.S.L. under the supervision of Y.F.N., X.R.W., and Y.S.; H.K.N. and Y.G. performed the electrical measurements of FeFETs under the supervision of X.R.W. and Y.S.; X.R.Y. and L.H. wrote the manuscript. All authors discussed the data and contributed to the manuscript.
Peer review
Peer review information
Nature Communications thanks Chang Liu and the other anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
The data supporting the findings of this study have been included in the main text and Supplementary Information. All other information can be obtained from the corresponding author upon request.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Xinrui Yang, Lu Han, Hongkai Ning, Shaoqing Xu, Bo Hao.
Contributor Information
Lu Han, Email: lhan@nju.edu.cn.
Xinran Wang, Email: xrwang@nju.edu.cn.
Yuefeng Nie, Email: ynie@nju.edu.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-53436-6.
References
- 1.Lu, D. et al. Synthesis of freestanding single-crystal perovskite films and heterostructures by etching of sacrificial water-soluble layers. Nat. Mater.15, 1255–1260 (2016). [DOI] [PubMed] [Google Scholar]
- 2.Ji, D. et al. Freestanding crystalline oxide perovskites down to the monolayer limit. Nature570, 87–90 (2019). [DOI] [PubMed] [Google Scholar]
- 3.Dong, G. et al. Super-elastic ferroelectric single-crystal membrane with continuous electric dipole rotation. Science366, 475–479 (2019). [DOI] [PubMed] [Google Scholar]
- 4.Han, L. et al. Giant uniaxial strain ferroelectric domain tuning in freestanding PbTiO3 films. ACS Appl. Mater. Interfaces7, 1901604 (2020). [Google Scholar]
- 5.Hong, S. S. et al. Extreme tensile strain states in La0.7Ca0.3MnO3 membranes. Science368, 71–76 (2020). [DOI] [PubMed] [Google Scholar]
- 6.Cai, S. et al. Enhanced polarization and abnormal flexural deformation in bent freestanding perovskite oxides. Nat. Commun.13, 5116 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Xu, R. et al. Strain-induced room-temperature ferroelectricity in SrTiO3 membranes. Nat. Commun.11, 3141 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Han, L. et al. High-density switchable skyrmion-like polar nanodomains integrated on silicon. Nature603, 63–67 (2022). [DOI] [PubMed] [Google Scholar]
- 9.Sun, H. et al. Nonvolatile ferroelectric domain wall memory integrated on silicon. Nat. Commun.13, 4332 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Lu, D. et al. Freestanding oxide ferroelectric tunnel junction memories transferred onto silicon. Nano Lett19, 3999–4003 (2019). [DOI] [PubMed] [Google Scholar]
- 11.Zheng, N. et al. Perovskite-oxide-based ferroelectric synapses integrated on silicon. Adv. Funct. Mater. 34, 2316473 (2024).
- 12.Huang, J.-K. et al. High-κ perovskite membranes as insulators for two-dimensional transistors. Nature605, 262–267 (2022). [DOI] [PubMed] [Google Scholar]
- 13.Yang, A. J. et al. Van der Waals integration of high-κ perovskite oxides and two-dimensional semiconductors. Nat. Electron.5, 233–240 (2022). [Google Scholar]
- 14.Puebla, S. et al. Combining freestanding ferroelectric perovskite oxides with two-dimensional semiconductors for high performance transistors. Nano Lett22, 7457–7466 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Wang, L. et al. Flexoelectronics of centrosymmetric semiconductors. Nat. Nanotechnol.15, 661–667 (2020). [DOI] [PubMed] [Google Scholar]
- 16.Guo, D. et al. Silicon flexoelectronic transistors. Sci. Adv.9, eadd3310 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Huang, B. et al. Mechanically gated transistor. Adv. Mater.35, 2305766 (2023). [DOI] [PubMed] [Google Scholar]
- 18.Zubko, P. et al. Flexoelectric effect in solids. Annu. Rev. Mater. Res.43, 387–421 (2013). [Google Scholar]
- 19.Bhaskar, U. K. et al. A flexoelectric microelectromechanical system on silicon. Nat. Nanotechnol.11, 263–266 (2016). [DOI] [PubMed] [Google Scholar]
- 20.Park, S. M. et al. Selective control of multiple ferroelectric switching pathways using a trailing flexoelectric field. Nat. Nanotechnol.13, 366–370 (2018). [DOI] [PubMed] [Google Scholar]
- 21.Wu, J. et al. Manipulation of bilayer MoS2-based MESFET with flexoelectric polarization field. Nano Energy123, 109415 (2024). [Google Scholar]
- 22.Jiang, X. et al. Flexoelectric nano-generator: Materials, structures and devices. Nano Energy2, 1079–1092 (2013). [Google Scholar]
- 23.Catalan, G. et al. Flexoelectric rotation of polarization in ferroelectric thin films. Nat. Mater.10, 963–967 (2011). [DOI] [PubMed] [Google Scholar]
- 24.Catalan, G. et al. The effect of flexoelectricity on the dielectric properties of inhomogeneously strained ferroelectric thin films. J. Phys. Condens. Matter16, 2253 (2004). [Google Scholar]
- 25.Lee, D. et al. Giant flexoelectric effect in ferroelectric epitaxial thin films. Phys. Rev. Lett.107, 057602 (2011). [DOI] [PubMed] [Google Scholar]
- 26.Ming, W. et al. Flexoelectric engineering of van der Waals ferroelectric CuInP2S6. Sci. Adv.8, eabq1232 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Chen, C. et al. Large-scale domain engineering in two-dimensional ferroelectric CuInP2S6 via giant flexoelectric effect. Nano Lett.22, 3275–3282 (2022). [DOI] [PubMed] [Google Scholar]
- 28.Lun, Y. et al. Ultralow tip-force driven sizable-area domain manipulation through transverse flexoelectricity. Adv. Mater.35, 2302320 (2023). [DOI] [PubMed] [Google Scholar]
- 29.Han, L. et al. Freestanding perovskite oxide membranes: a new playground for novel ferroic properties and applications. Adv. Funct. Mater.34, 2309543 (2024). [Google Scholar]
- 30.Liu, H. et al. Reversible flexoelectric domain engineering at the nanoscale in van der Waals ferroelectrics. Nat. Commun.15, 4556 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Tadmor, E. B. et al. Polarization switching in PbTiO3: an ab initio finite element simulation. Acta Mater50, 2989–3002 (2002). [Google Scholar]
- 32.Li, Q. et al. Ferroelastic nanodomain-mediated mechanical switching of ferroelectricity in thick epitaxial films. Nano Lett21, 445–452 (2021). [DOI] [PubMed] [Google Scholar]
- 33.Yuan, G. et al. Ferroelastic-domain-assisted mechanical switching of ferroelectric domains in Pb(Zr,Ti)O3 thin films. Adv. Electron. Mater.6, 2000300 (2020). [Google Scholar]
- 34.Masys, Š. et al. Elastic properties of rhombohedral, cubic, and monoclinic phases of LaNiO3 by first principles calculations. Comp. Mater. Sci.108, 153–159 (2015). [Google Scholar]
- 35.Kiely, J. D. et al. Nanomechanical properties of Au (111), (001), and (110) surfaces. Phys. Rev. B57, 12588–12594 (1998). [Google Scholar]
- 36.Wang, J. et al. Domain switching in BaTiO3 films induced by an ultralow mechanical force. ACS Appl. Mater. Interfaces14, 48917–48925 (2022). [DOI] [PubMed] [Google Scholar]
- 37.Edwards, D. et al. Local probing of ferroelectric and ferroelastic switching through stress-mediated piezoelectric spectroscopy. ACS Appl. Mater. Interfaces3, 1500470 (2016). [Google Scholar]
- 38.Gonzalez Casal, S. et al. Mechanical switching of ferroelectric domains in 33-200 nm-thick sol-gel-grown PbZr0.2Ti0.8O3 films assisted by nanocavities. Adv. Electron. Mater.8, 2200077 (2022). [Google Scholar]
- 39.Cappella, B. et al. Force-distance curves by atomic force microscopy. Surf. Sci. Rep.34, 1–104 (1999). [Google Scholar]
- 40.Hutter, J. L. et al. Calibration of atomic‐force microscope tips. Rev. Sci. Instrum.64, 1868–1873 (1993). [Google Scholar]
- 41.Yang, Q. et al. Ferroelectricity in layered bismuth oxide down to 1 nanometer. Science379, 1218–1224 (2023). [DOI] [PubMed] [Google Scholar]
- 42.Farraro, R. et al. Temperature dependence of the Young’s modulus and shear modulus of pure nickel, platinum, and molybdenum. Metall. Trans. A8, 1563–1565 (1977). [Google Scholar]
- 43.Yi, S.-H. et al. Enhancement of energy storage for electrostatic supercapacitors through built-in electric field engineering. Nano Energy99, 107342 (2022). [Google Scholar]
- 44.Lu, H. et al. Enhancement of ferroelectric polarization stability by interface engineering. Adv. Mater.24, 1209–1216 (2012). [DOI] [PubMed] [Google Scholar]
- 45.Gupta, S. K. et al. Measurement of the work function of some metals using internal voltage in MIM structures. J. Appl. Phys.50, 2852–2855 (1979). [Google Scholar]
- 46.Hopcroft, M. A. et al. What is the Young’s Modulus of Silicon? J. Microelectromech. Syst.19, 229–238 (2010). [Google Scholar]
- 47.Xu, Z.-H. et al. Deriving mechanical properties of soft coatings using nanoindentation: an application of mechanism-based strain gradient plasticity. Surf. Coat. Technol.157, 231–237 (2002). [Google Scholar]
- 48.Stengel, M. Flexoelectricity from density-functional perturbation theory. Phys. Rev. B88, 174106 (2013). [Google Scholar]
- 49.Wen, Z. et al. Mechanical switching of ferroelectric polarization in ultrathin BaTiO3 films: the effects of epitaxial strain. Appl. Phys. Lett. 104, 42907 (2014).
- 50.Guo, E. J. et al. Strain induced low mechanical switching force in ultrathin PbZr0.2Ti0.8O3 films. Appl. Phys. Lett. 105, 12903 (2014).
- 51.Vats, G. et al. Low-pressure mechanical switching of ferroelectric domains in PbZr0.48Ti0.52O3. Adv. Electron. Mater.6, 2000523 (2020). [Google Scholar]
- 52.Tian, S. et al. Effect of mechanical force on domain switching in BiFeO3 ultrathin films. Sci. China Phys. Mech.63, 217711 (2019). [Google Scholar]
- 53.Guan, Z. et al. Mechanical polarization switching in Hf0.5Zr0.5O2 thin film. Nano Lett22, 4792–4799 (2022). [DOI] [PubMed] [Google Scholar]
- 54.Chen, L. et al. Electrical and mechanical switching of ferroelectric polarization in the 70 nm BiFeO3 film. Sci. Rep.6, 19092 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Cao, Y. et al. Pressure-induced switching in ferroelectrics: Phase-field modeling, electrochemistry, flexoelectric effect, and bulk vacancy dynamics. Phys. Rev. B96, 184109 (2017). [Google Scholar]
- 56.Geim, A. K. et al. Van der Waals heterostructures. Nature499, 419–425 (2013). [DOI] [PubMed] [Google Scholar]
- 57.Liu, Y. et al. Promises and prospects of two-dimensional transistors. Nature591, 43–53 (2021). [DOI] [PubMed] [Google Scholar]
- 58.Kalinin, S. V. et al. Surface-screening mechanisms in ferroelectric thin films and their effect on polarization dynamics and domain structures. Rep. Prog. Phys.81, 036502 (2018). [DOI] [PubMed] [Google Scholar]
- 59.Lu, H. et al. Nanodomain engineering in ferroelectric capacitors with graphene electrodes. Nano Lett16, 6460–6466 (2016). [DOI] [PubMed] [Google Scholar]
- 60.Wang, Y. et al. Mechanically driven reversible polarization switching in imprinted BiFeO3 thin films. Adv. Funct. Mater.33, 2213787 (2023). [Google Scholar]
- 61.Liu, J. et al. Bidirectional mechanical switching window in ferroelectric thin films predicted by first-principle-based simulations. Npj Comput. Mater.8, 137 (2022). [Google Scholar]
- 62.Mills, G. et al. Reversible work transition state theory: application to dissociative adsorption of hydrogen. Surf. Sci.324, 305–337 (1995). [Google Scholar]
- 63.Berne, B. J. et al. Classical and Quantum Dynamics in Condensed Phase Simulations. p. 880 (WORLD SCIENTIFIC: 1998).
- 64.Kresse, G. et al. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comp. Mater. Sci.6, 15–50 (1996). [Google Scholar]
- 65.Kresse, G. et al. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
- 66.Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B50, 17953–17979 (1994). [DOI] [PubMed] [Google Scholar]
- 67.Kresse, G. et al. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B59, 1758–1775 (1999). [Google Scholar]
- 68.Perdew, J. P. et al. Restoring the density-gradient expansion for exchange in solids and surfaces. Phys. Rev. Lett.100, 136406 (2008). [DOI] [PubMed] [Google Scholar]
- 69.Li, Z. et al. Dielectric and elastic properties of ferroelectric materials at elevated temperature. J. Phys. Chem. Solids57, 1433–1438 (1996). [Google Scholar]
- 70.Li, Q. et al. Quantification of flexoelectricity in PbTiO3/SrTiO3 superlattice polar vortices using machine learning and phase-field modeling. Nat. Commun.8, 1468 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Kalinin, S. V. et al. Quantitative determination of tip parameters in piezoresponse force microscopy. Appl. Phys. Lett. 90, 212905 (2007).
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data supporting the findings of this study have been included in the main text and Supplementary Information. All other information can be obtained from the corresponding author upon request.