DOI:
10.1039/D4TC04614E
(Paper)
J. Mater. Chem. C, 2025,
13, 5287-5294
Voltage-control of the in-plane magnetic anisotropy in hybrid magnetoelectric Ni90Fe10/BaTiO3(011) heterostructures†
Received
29th October 2024
, Accepted 16th January 2025
First published on 16th January 2025
Abstract
This study focuses on hybrid magnetoelectric Ni90Fe10/BaTiO3(011) heterostructures, which enable the control of the in-plane magnetization of the magnetostrictive layer through electric voltage. The heterostructure is both Pb- and rare-earth-free, thus enhancing environmental sustainability. We show that the BaTiO3(011) orientation enables higher deformations in the piezoelectric regime compared to the commonly studied (001) orientation. In the as-grown state, the electrodeposited 200 nm-thick Ni90Fe10 film presents uniaxial in-plane anisotropy aligned with the [100] in-plane crystallographic direction of the BaTiO3(011) substrate. X-ray magnetic circular dichroism photoemission electron microscopy images, along with hysteresis loops obtained by the magneto-optical Kerr effect, confirm the converse magnetoelectric coupling between Ni90Fe10 and BaTiO3(011). The obtained converse magnetoelectric coupling constant of 0.205 μs m−1 is significant considering it is achieved in the piezoelectric regime of the BaTiO3 substrate and using an electrodeposited magnetostrictive film, making this heterostructure more viable for future applications. This value represents an increase of more than double compared to that previously reported for Ni/BTO(001) and, to the best of our knowledge, is the first value reported for the BTO(011) orientation.
Introduction
Recent advancements in the manipulation of magnetic systems through electrical control have revolutionized the field, in a hot topic known as voltage control of magnetic anisotropy (VCMA).1–5 Multiferroics stand out for their converse magnetoelectric coupling, αE = μ0ΔM/ΔE, allowing an electric field (E) to change the magnetization (M).6–8 Multiferroics can be classified into two categories: single-phase and hybrid. In single-phase multiferroics, multiple ferroic orders can coexist in the same material, i.e., ferroelectricity and ferromagnetism, whereas hybrid multiferroics consist of different materials with distinct ferroic orders combined in one heterostructure to produce multiferroic behavior, e.g. the magnetoelectric effect. Single-phase multiferroics, such as TbMnO3 with a converse magnetoelectric coupling constant of 10−5 μs m−1,9 Ni3B7O13I with 10−3 μs m−1,10 and Cr2O3 with 4.1 × 10−6 μs m−1,11 have been reported. On the other hand, there are multiferroics based on hybrid systems like BiFeO3/BaTiO3, which has an αE of 0.2 μs m−1,12 NFO/PZT with 0.8 μs m−1,13 and FeRh/BaTiO3 with 16 μs m−1.14 Hybrid multiferroics outperform single-phase ones due to their notably higher αE values at room temperature, while single-phase multiferroics exhibit lower or null values at room temperature.15 Therefore, in this work, we focus on hybrid multiferroics based on strain-coupled systems, which achieve strong magnetoelectric coupling at room temperature and are ideal for applications such as memory devices, low-power logic devices, and energy harvesters.15,16
The mechanism of operation in these hybrid systems is the interaction between a ferroelectric substrate and a magnetostrictive layer. Applying an electric field across the ferroelectric substrate induces mechanical strain, which transfers to the magnetostrictive layer. This strain modifies the magnetic anisotropy of the top layer through inverse magnetostriction. To improve control over the magnetic anisotropy, it is crucial to use ferroelectrics with high piezoelectric coefficients and ferromagnets with both low anisotropy and large magnetostriction.5 Typical ferroelectrics used in strain-coupled systems are based on Pb, such as PZT17,18 and PMN-PT,19–23 due to their ability to achieve high strain with low voltage. For ferromagnets, the highest magnetostriction can be achieved by using rare-earth ferromagnets like Terfenol or Terfenol-D.24 In this work, we aim to develop an environmentally friendly heterostructure capable of controlling magnetic anisotropy without relying on Pb or rare-earth elements.
Magnetostrictive ferromagnets, particularly NixFe100−x alloys, have been shown to be effective in VCMA systems.5 Previously, we demonstrated control of the magnetic anisotropy of electrodeposited NixFe100−x alloys by means of growth parameters such as stirring, applied magnetic field during deposition, film thickness, or alloy composition.25–27 The Ni90Fe10 composition exhibits a negative magnetostriction of −22 ppm, and in particular, a 200 nm thick film shows an in-plane (IP) isotropic magnetic behavior,25 which is compatible with strain-induced systems due to its absence of IP anisotropy and low coercivity. This layer thickness, 200 nm, is chosen to avoid charge-mediated effects that cause magnetoelectric coupling at the ferroelectric/ferromagnetic interface. As reported by Zhou and collaborators in ref. 28, they calculated that the charge-mediated effect decays beyond 1.2 nm and later experimentally showed in ref. 29 that for a 50 nm film the effect is negligible. Similarly, Paudel et al.30 reported a comparable result, measured by polarized neutron reflectometry, showing that the effect extends to only 2 nm. On the other hand, strain is a volumetric effect that affects all the upper layers if they are thinner than the substrate.30,31 Thus, a 200 nm NiFe layer is sufficiently thick to eliminate charge-mediated effects, and the primary mechanism governing the magnetoelectric coupling will be the strain-coupling mechanism.
BaTiO3 (BTO), a well-studied ferroelectric material with a perovskite-like structure,32 is a viable substrate for strain-coupled systems. It can generate strain through three main mechanisms: photostriction,33–35 and thermal36–38 or voltage induction.39–41 We specifically investigated voltage induction among these three possibilities, as it enables operation at room temperature and device integration. Although BTO is ferroelectric, in this work, we exploit its piezoelectric regime because it allows for operation at several frequencies and it is independent of polarization switching, leading to a longer lifespan for future applications.42 This is advantageous because polarization switching can be challenging,19 and ferroelectric materials may fatigue after many cycles, leading to unreliable performance.43 The lifespan of a perovskite that switches its polarization is around 104 cycles before experiencing fatigue,44 while in the piezoelectric regime, it can withstand fatigue after 108 cycles.45 The BTO(011) surface has been scarcely explored in the context of strain-coupled systems. However, this BTO cut takes advantage of the piezoelectric coefficient d15, which is one order of magnitude larger than the other coefficients. Therefore, the use of d15 can potentially increase the capability of BTO(011) for VCMA, as it is shown in this work in which αE = 0.205 μs m−1 is obtained. In previous studies exploiting other BTO orientations in which the switching of the polarization is needed, αE of 0.1–0.2 μs m−1 was reported.46–48
Experimental section
We have used a single-crystal BTO(011) substrate onto which a metallic Ti/Au bilayer was deposited on both sides using an e-beam evaporator under high vacuum conditions. The bottom metallic bilayer serves as an electrode for voltage application, while the top bilayer provides the necessary electrical conductivity for electrodeposition, given the non-conductive nature of BTO. Electrical isolation between both sides creates a plane-parallel capacitor around the substrate. On the polished side of the BTO, a 200 nm-thick Ni90Fe10 layer was electrodeposited over the conductive Ti/Au bilayer. Therefore, the hybrid magnetoelectric heterostructure consists of Au(50 nm)/Ti(15 nm)/BTO(500 μm)/Ti(15 nm)/Au(50 nm)/Ni90Fe10(200 nm) as illustrated in Fig. 1a where the electric voltage applied during measurements is also indicated. Electrodeposition was conducted using a PalmSens EmStat3 setup in a controlled environment with a three-electrode cell configuration at room temperature. The top Au capping layer served as the working electrode, a platinum mesh as the counter electrode, and Ag/AgCl (3 M NaCl) as the reference electrode. The electrolyte, a water-based solution with a pH of 2.4, consisted of 0.4 M H3BO3, 0.017 M saccharine, 0.7 M NiSO4, and 0.02 M FeSO4. A growth voltage of −1.20 V was applied, and film thickness was monitored by measuring the electric charge at the working electrode. Film composition was analyzed using energy dispersive X-ray spectroscopy (EDS) with a JEOL JSM 6400 instrument operating at 15 kV acceleration voltage.
 |
| Fig. 1 (a) Schematic drawing of the studied heterostructure showing the applied electric voltage. (b) BTO(011) surface (yellow) showing the IP directions, [100] and [0 1], and the electric field applied in the OOP direction for the BTO(011) orientation studied in this work. The green axes represent the coordinate axes for the piezoelectric matrix. Axes 2′ and 3′ are obtained after a rotation of θ = 45° around axis 1. | |
We studied the room temperature magneto-optic Kerr effect (MOKE) in the longitudinal configuration using s-polarized light. Our home-made experimental setup includes a laser operating at 650 nm with a power output of 5 mW and a frequency of 10 kHz, along with two polarizers, a photodetector, and an electromagnet capable of generating a magnetic field of up to 600 mT. The applied magnetic field is continuously monitored using a Hall probe. The signal detected by the photodetector is processed by a lock-in amplifier model SR830m. To apply an electric field perpendicular to the sample during MOKE measurements, a custom holder was designed and connected to a voltage power supply capable of applying 600 V. The voltage was continuously applied during the in situ MOKE measurements to ensure that the hysteresis loops were recorded in a sequential manner. Data acquisition was carried out using a custom Python program.
To determine the orientation and lattice parameters of the BTO substrate, X-ray diffraction (XRD) was performed using the X'Pert PRO MRD equipment with a Cu Kα wavelength. The ESI,† includes the reciprocal space maps (RSMs) used to obtain the pristine lattice constants of BTO.
Magnetic imaging was conducted using X-ray magnetic circular dichroism photo emission electron microscopy (XMCD-PEEM) at the CIRCE beamline of the ALBA Synchrotron.49In situ electrical poling and magnetic field were applied to the sample utilizing custom-designed sample holders.50 Magnetic images were acquired at the Fe L3 absorption edge (∼707 eV photon energy) by imaging with low-energy secondary electrons (Ekin around 1–2 eV) and performing pixel-wise subtraction of images taken with left and right handed circularly polarized X-rays.
Results and discussion
The BTO(011) cut can take advantage of the remarkably high piezoelectric coefficient d15 = 564 pm V−1.51 This value is an order of magnitude greater than other piezoelectric coefficients, such as d33 = 90 pm V−1 and d31 = −33.4 pm V−1.51 For 4mm point group crystals, the reverse piezoelectric effect can be described as follows:52 |  | (1) |
where εi are the strain components, di the piezoelectric coefficients, and Ei the applied electric field components. The c-axis of BTO is chosen, as usual, as axis 3 (Fig. 1b). When an out-of-plane (OOP) electric field is applied to the BTO(011) surface, the unit electric vector aligns with the surface, resulting in Ê = (0, 1, 1). From Eqn 1 it is obtained that ε4 = d15E2, indicating a strain component proportional to d15. As the BTO(011) surface is not aligned with the main axes (i.e., 1, 2, and 3), shear stress components (i.e., 4, 5, and 6) are transferred to the surface. This behavior is particularly significant for BTO(011), where shear components related to d15 will play a crucial role in the surface strain. When compared with the generally studied BTO(001) orientation, it becomes evident that for this orientation, there are no strain components dependent on d15 (E1 = E2 = 0) in the piezoelectric regime. This difference highlights the potential of the BTO(011) orientation compared to the (001) orientation to achieve large strain and, consequently, a higher αE.
To calculate the dependence of the IP strain on the piezoelectric coefficients for BTO(011), we consider a∼c. The OOP electric field is given by
, to satisfy that |
| = E. Fig. 1b shows the coordinate axes of our system (green), where it is straightforward to see that the strain along the BTO[100] axis is equal to the strain along axis 1. To calculate the strain along the BTO[0
1] direction, we need to rotate the coordinate axes around axis 1 by an angle θ to achieve the new axes 2′ and 3′, where 3′ is along the [0
1] direction. As we consider that a∼c, the angle of rotation will be θ = 45°. The strain along axis 3′ can be calculated using the following equation:53
|  | (2) |
Taking into account the results from eqn (1), we finally have that:
|  | (3) |
|  | (4) |
The comparison between eqn (3) and (4) shows that the strain along the [0
1] direction will be an order of magnitude higher than that along the [100] direction. Also, the signs are opposite as expected, as tension applied in one direction induces compression in the perpendicular direction to compensate for the volume change. The magnetostriction is proportional to the strain,54 so a linear magneto-electric coupling dependence on the applied electric field is expected.
An intermediate conductive bilayer between BTO and NiFe is required for the electrodeposition of the latter on top of insulator BTO, which may affect strain transfer effectiveness. The strain transfer function depends on the thickness ratio between the layers and the substrate.31 Since the substrate thickness is much greater than the thickness of the other layers, and assuming a rigid interface, no slipping between layers and negligible edge effects, we can conclude that the strain is uniformly transferred from the substrate to the upper layers. However, the stress is applied differently across the layers, and we can calculate the stress transfer through the layers according to31 as:
|  | (5) |
where
σl and
σs are the stresses, and
El and
Es are the Young's moduli of the layer and substrate, respectively. The ratio of Young's moduli (
Yl/
Ys) determines the effectiveness of the stress transfer. Considering the Young's modulus of each material:
YTi = 115 GPa,
YAu = 79 GPa,
YNiFe = 96.4 GPa and
YBTO = 99 GPa,
55–58 we find that the stress transfer ratios are
YTi/
YBTO = 1.16,
YAu/
YBTO = 0.79,
YNiFe/
YBTO = 0.97. Note that we are using Young's modulus values from the literature, which may differ from those of our actual system due to clamping effects and/or the polycrystallinity of the layers. From the stress transfer ratios, we can infer that the stress transfer between the layers is efficient. Therefore, the introduction of intermediate metallic layers with a Young's modulus similar to BTO makes the heterostructure still viable for strain-coupling systems.
Fig. 2a presents the XRD characterization of pristine BTO(011). θ–2θ diffraction patterns confirmed that the BTO surface corresponds to either the (011) or (101) orientations. Through additional RSM measurements, we determined that the surface of the studied BTO is (011), with IP directions [0
1] and [100], and lattice parameters: a0 = 3.994 Å and c0 = 4.034 Å (Fig. 2b and ESI†). Fig. 2c shows the IP MOKE hysteresis loops for the as-grown Ni90Fe10, revealing a clear IP uniaxial anisotropy aligned with the BTO [100] direction. Electrodeposited layers on glass substrates of the same thickness are magnetically isotropic in the sample plane.27 However, the appearance of this uniaxial IP anisotropy indicates that this BTO orientation highly influences the magnetic behavior of Ni90Fe10, as the easy magnetic axis aligns with the short side of the rectangular BTO cell on the surface.
 |
| Fig. 2 (a) θ–2θ XRD measurements of pristine BTO(011) (blue line) compared with possible surface orientations (red and green dashed lines). (b) RSM measurement (black line) of the (122) reflection and corresponding fit (red dashed line). (c) MOKE hysteresis loops of as-grown Ni90Fe10 layer along IP hard and easy directions, corresponding to in-plane BTO [0 1] (black line) and [100] directions (red line), respectively. | |
After characterizing the as-grown state, we applied electric fields of ±0.46 MV m−1 to pole the sample at the CIRCE beamline. These electric field values are below the ferroelectric coercive field of BTO (0.60 MV m−1),59 ensuring that the experiment was carried out within the piezoelectric regime. Fig. 3a shows XMCD-PEEM images, and Fig. 3b is a schematic illustration to interpret those results. Whereas for E = −0.46 MV m−1, XMCD-PEEM shows minimal contrast, for E = +0.46 MV m−1, the image displays a clear contrast that reveals IP uniaxial magnetic anisotropy (red arrows indicate the magnetization vectors). These experimental results can be understood thanks to Fig. 3b that depicts the BTO cell with blue arrows for the direction of mechanical stress promoted by poling. For negative bias (left panel), the IP strain tends to make the cell squarer, diminishing the magnetic anisotropy of the magnetostrictive Ni90Fe10. On the other hand, for positive bias (right panel) it elongates the cell in the [0
1] direction, enhancing the uniaxial magnetic anisotropy in the [100] direction according to the negative magnetostriction of Ni90Fe10.
 |
| Fig. 3 (a) XMCD-PEEM images taken at E = ±0.46 MV m−1 with the same scale bar. IP axes (white) as well as the magnetization (red) are indicated. (b) Schematic drawings illustrate how the stress (blue arrows) generated by poling changes the size of the piezoelectric substrate that induces a modification of the magnetization on the magnetostrictive layer (red arrow) as a result. The left panel shows the interpretation of the XMCD-PEEM image at E = −0.46 MV m−1, and the right panel at E = +0.46 MV m−1. | |
MOKE hysteresis loops with the applied magnetic field along the BTO [100] direction were conducted under several applied electric fields between ±0.46 MV m−1 to quantify αE for the Ni90Fe10/BTO(011) heterostructure. Fig. 4a presents a representative set of IP hysteresis loops at four different applied electric fields. Here, +0 MV m−1 is used for zero electric field after a positive bias, while −0 MV m−1 denotes zero electric field after a negative bias. By calculating the squareness (Mr/Ms) for each cycle, we can plot its value at different electric fields, as depicted in Fig. 4b. The higher squareness for positive bias indicates a better alignment of the easy axis in the studied direction, whereas the reduction of the squareness for negative bias points to a loss of magnetization alignment or a reduction in magnetic anisotropy, as observed by XMCD-PEEM in Fig. 3a.
 |
| Fig. 4 (a) IP MOKE hysteresis loops for Ni90Fe10 taken along the BTO [100] direction at several polings. (b) Squareness (Mr/Ms) represented as a function of poling. Descent (black down triangles) and ascent (red up triangles) electric paths are shown differently for clarity. The linear fit of is used to obtain αE according to eqn (6) of the main text. | |
The observed behavior indicates a linear coupling between magnetism and the electric field, as expected. From the linear fit of the squareness as a function of the applied electric field (Fig. 4b), αE is obtained as:
|  | (6) |
where
μ0 is the vacuum permeability, and
Ms is the saturation magnetization of Ni
90Fe
10, which is 652 kA m
−1.
54 From the linear fit it is obtained
αE = 0.205 μs m
−1 = 61.5 mV (cm Oe)
−1 that can be compared with other works performed with the BTO(001) orientation. Researchers found
αE for strain-coupled heterostructures to be 0.09 μs m
−1 for Ni/BTO,
46 0.1 μs m
−1 for Fe
3O
4/BTO,
47 0.23 μs m
−1 for La
0.67Sr
0.33MnO
3/BTO
48 and 16 μs m
−1 for FeRh/BTO.
14 In these cases, the differences in
αE arise from the magnetostrictive layer, with the highest values associated with rare-earth alloys. In our study, the Ni
90Fe
10 alloy exhibits a magnetostriction of −22 ppm,
25 which is comparable to that of Fe
3O
4 (−25 ppm)
60 and even to pure Ni (−37 ppm).
54 Comparing our
αE value is noteworthy, as we achieved a value that is nearly double that of previous studies, demonstrating the enhanced performance of our heterostructure.
Conclusions
In conclusion, this study offers valuable insights into the Pb- and rare-earth-free Ni90Fe10/BTO(011) strain-mediated heterostructures, demonstrating their capability to control magnetic anisotropy through an electric field. The as-grown Ni90Fe10 film exhibits uniaxial IP anisotropy promoted during growth by the BTO(011) substrate. By poling the sample within the piezoelectric regime, it is possible to tune this IP magnetic anisotropy, as observed in XMCD-PEEM images. MOKE hysteresis loops measured at various polings demonstrated a noteworthy converse magnetoelectric coefficient of αE = 0.205 μs m−1. The αE value in our work is more than double the previously reported value for Fe3O4/BTO(001) (which has a similar magnetostriction layer) and, to the best of our knowledge, is the first value reported for the BTO(011) orientation. Future research could focus on using deposition systems that can deposit directly onto the substrate, eliminating the need for intermediate layers. This could improve stress transfer from the substrate to the layer, increasing the magnetoelectric coupling value, and broadening its applications in efficient magnetic memories.
Author contributions
The manuscript was written through contributions from all authors. All authors have given approval to the final version of the manuscript. A. Begué: data curation (lead); analysis (lead); investigation (lead); methodology; writing – original draft (lead). M. W. Khaliq: investigation; methodology. N. Cotón: investigation; methodology. M. A. Niño: analysis; investigation; methodology; funding acquisition; writing – review & editing. M. Foerster: analysis; investigation; methodology; funding acquisition; writing – review & editing. R. Ranchal: data curation; analysis; investigation; methodology (lead); conceptualization (lead); funding acquisition; project administration; writing – review & editing.
Data availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Conflicts of interest
The authors have no conflicts to disclose.
Acknowledgements
This work has been financially supported through the projects PID2021-122980OB-C51 (AEI/FEDER) and PID2021-122980OB-C54 (AEI/FEDER) of the Spanish Ministry of Science and Innovation. A. B. would like to acknowledge the funding received from the Ministry of Universities and the European Union-Next Generation for the Margarita Salas fellowship. These experiments were performed on the BL24-CIRCE beamline at the ALBA Synchrotron with the collaboration of ALBA staff. We thank the CAI X-ray Diffraction Unit and Ion Implantation Unit at the Complutense University of Madrid for their valuable technical assistance.
References
- K. Cai, M. Yang, H. Ju, S. Wang, Y. Ji, B. Li, K. W. Edmonds, Y. Sheng, B. Zhang, N. Zhang, S. Liu, H. Zheng and K. Wang, Electric field control of deterministic current-induced magnetization switching in a hybrid ferromagnetic/ferroelectric structure, Nat. Mater., 2017, 16, 712–716 CrossRef CAS PubMed.
- H. Meer, F. Schreiber, C. Schmitt, R. Ramos, E. Saitoh, O. Gomonay, J. Sinova, L. Baldrati and M. Kläui, Direct Imaging of Current-Induced Antiferromagnetic Switching Revealing a Pure Thermomagnetoelastic Switching Mechanism in NiO, Nano Lett., 2020, 21, 114–119 CrossRef PubMed.
- P. Wadley, B. Howells, J. Železný, C. Andrews, V. Hills, R. P. Campion, V. Novák, K. Olejník, F. Maccherozzi, S. S. Dhesi, S. Y. Martin, T. Wagner, J. Wunderlich, F. Freimuth, Y. Mokrousov, J. Kuneš, J. S. Chauhan, M. J. Grzybowski, A. W. Rushforth, K. W. Edmonds, B. L. Gallagher and T. Jungwirth, Electrical switching of an antiferromagnet, Science, 2016, 351, 587–590 CrossRef CAS PubMed.
- T. Koyama, Y. Nakatani, J. Ieda and D. Chiba, Electric field control of magnetic domain wall motion via modulation of the Dzyaloshinskii-Moriya interaction, Sci. Adv., 2018, 4, 0265 Search PubMed.
- C. Song, B. Cui, F. Li, X. Zhou and F. Pan, Recent progress in voltage control of magnetism: Materials, mechanisms, and performance, Prog. Mater. Sci., 2017, 87, 33–82 CrossRef.
- K. F. Wang, J.-M. Liu and Z. F. Ren, Multiferroicity: the coupling between magnetic and polarization orders, Adv. Phys., 2009, 58, 321–448 CrossRef CAS.
- N. A. Spaldin and R. Ramesh, Advances in magnetoelectric multiferroics, Nat. Mater., 2019, 18, 203–212 CrossRef CAS PubMed.
- C. N. R. Rao and C. R. Serrao, New routes to multiferroics, J. Mater. Chem., 2007, 17(47), 4931–4938 RSC.
- N. Hur, S. Park, P. A. Sharma, J. S. Ahn, S. Guha and S.-W. Cheong, Electric polarization reversal and memory in a multiferroic material induced by magnetic fields, Nature, 2004, 429, 392–395 CrossRef CAS PubMed.
- E. Ascher, H. Rieder, H. Schmid and H. Stössel, Some Properties of Ferromagnetoelectric Nickel-Iodine Boracite, Ni3B7O13I, J. Appl. Phys., 1966, 37, 1404–1405 CrossRef CAS.
- V. J. Folen, G. T. Rado and E. W. Stalder, Anisotropy of the Magnetoelectric Effect in Cr2O3, Phys. Rev. Lett., 1961, 6, 607–608 CrossRef CAS.
- R. Gupta, S. Chaudhary and R. K. Kotnala, Interfacial Charge Induced Magnetoelectric Coupling at BiFeO3/BaTiO3 Bilayer Interface, ACS Appl. Mater. Interfaces, 2015, 7, 8472–8479 CrossRef CAS PubMed.
- Z. Zheng, P. Zhou, Y. Liu, K. Liang, R. G. Tanguturi, H. Chen, G. Srinivasan, Y. Qi and T. Zhang, Strain effect on magnetoelectric coupling of epitaxial NFO/PZT heterostructure, J. Alloys Compd., 2020, 818, 152871 CrossRef CAS.
- R. O. Cherifi, V. Ivanovskaya, L. C. Phillips, A. Zobelli, I. C. Infante, E. Jacquet, V. Garcia, S. Fusil, P. R. Briddon, N. Guiblin, A. Mougin, A. A. Ünal, F. Kronast, S. Valencia, B. Dkhil, A. Barthélémy and M. Bibes, Electric-field control of magnetic order above room temperature, Nat. Mater., 2014, 13, 345–351 CrossRef CAS PubMed.
- R. Gupta and R. K. Kotnala, A review on current status and mechanisms of room-temperature magnetoelectric coupling in multiferroics for device applications, J. Mater. Sci., 2022, 57, 12710–12737 CrossRef CAS.
- A. Nicolenco, A. Gómez, X.-Z. Chen, E. Menéndez, J. Fornell, S. Pané, E. Pellicer and J. Sort, Strain gradient mediated magnetoelectricity in Fe-Ga/P(VDF-TrFE) multiferroic bilayers integrated on silicon, Appl. Mater. Today, 2020, 19, 100579 CrossRef.
- J. Ryu, S. Priya, K. Uchino and H.-E. Kim, Magnetoelectric Effect in Composites of Magnetostrictive and Piezoelectric Materials, J. Electroceram., 2002, 8, 107–119 CrossRef CAS.
- S.-K. Kim, J.-W. Lee, S.-C. Shin, H. W. Song, C. H. Lee and K. No, Voltage control of a magnetization easy axis in piezoelectric/ferromagnetic hybrid films, J. Magn. Magn. Mater., 2003, 267, 127–132 CrossRef CAS.
- A. Begué and M. Ciria, Strain-Mediated Giant Magnetoelectric Coupling in a Crystalline Multiferroic Heterostructure, ACS Appl. Mater. Interfaces, 2021, 13, 6778–6784 CrossRef PubMed.
- L. Leiva, J. L. Ampuero Torres, J. E. Gómez, D. Velázquez Rodriguez, J. Milano and A. Butera, Electric field control of magnetism in FePt/PMN-PT heterostructures, J. Magn. Magn. Mater., 2022, 544, 168619 CrossRef CAS.
- A. Chen, H.-G. Piao, C. Zhang, X.-P. Ma, H. Algaidi, Y. Ma, Y. Li, D. Zheng, Z. Qiu and X.-X. Zhang, Switching magnetic strip orientation using electric fields, Mater. Horiz., 2023, 10, 3034–3043 RSC.
- M. Buzzi, R. V. Chopdekar, J. L. Hockel, A. Bur, T. Wu, N. Pilet, P. Warnicke, G. P. Carman, L. J. Heyderman and F. Nolting, Single Domain Spin Manipulation by Electric Fields in Strain Coupled Artificial Multiferroic Nanostructures, Phys. Rev. Lett., 2013, 111, 027204 CrossRef CAS PubMed.
- S. Finizio, M. Foerster, M. Buzzi, B. Krüger, M. Jourdan, C. A. F. Vaz, J. Hockel, T. Miyawaki, A. Tkach, S. Valencia, F. Kronast, G. P. Carman, F. Nolting and M. Kläui, Magnetic Anisotropy Engineering in Thin Film Ni Nanostructures by Magnetoelastic Coupling, Phys. Rev. Appl., 2014, 1, 021001 CrossRef CAS.
-
R. Grössinger, R. S. Turtelli and N. Mehmood, Materials with high magnetostriction, IOP Conference Series: Materials Science and Engineering, 2014, vol. 60, p. 012002.
- N. Cotón, J. P. Andrés, E. Molina, M. Jaafar and R. Ranchal, Stripe domains in electrodeposited Ni90Fe10 thin films, J. Magn. Magn. Mater., 2023, 565, 170246 CrossRef.
- N. Cotón, J. P. Andrés, M. Jaafar, A. Begué and R. Ranchal, Tuning the out-of-plane magnetic textures of electrodeposited Ni90Fe10 thin films, J. Appl. Phys., 2024, 135, 093905 CrossRef.
- A. Begué, N. Cotón and R. Ranchal, Magnetic anisotropy evolution with Fe content in electrodeposited Ni100−xFex thin films, J. Mater. Chem. C, 2024, 12, 10104–10109 RSC.
- Z. Zhou, T. X. Nan, Y. Gao, X. Yang, S. Beguhn, M. Li, Y. Lu, J. L. Wang, M. Liu, K. Mahalingam, B. M. Howe, G. J. Brown and N. X. Sun, Quantifying thickness-dependent charge mediated magnetoelectric coupling in magnetic/dielectric thin film heterostructures, Appl. Phys. Lett., 2013, 103, 232906 CrossRef.
- Z. Zhou, B. M. Howe, M. Liu, T. Nan, X. Chen, K. Mahalingam, N. X. Sun and G. J. Brown, Interfacial charge-mediated non-volatile magnetoelectric coupling in Co0.3Fe0.7/Ba0.6Sr0.4TiO3/Nb:SrTiO3 multiferroic heterostructures, Sci. Rep., 2015, 5, 7740 CrossRef PubMed.
- B. Paudel, I. Vasiliev, M. Hammouri, D. Karpov, A. Chen, V. Lauter and E. Fohtung, Strain vs. charge mediated magnetoelectric coupling across the magnetic oxide/ferroelectric interfaces, RSC Adv., 2019, 9, 13033–13041 RSC.
- I. C. Noyan, T. C. Huang and B. R. York, Residual stress/strain analysis in thin films by X-ray diffraction, Crit. Rev. Solid State Mater. Sci., 1995, 20, 125–177 CrossRef CAS.
- M. Acosta, N. Novak, V. Rojas, S. Patel, R. Vaish, J. Koruza, G. A. Rossetti and J. Rödel, BaTiO3-based piezoelectrics: Fundamentals, current status, and perspectives, Appl. Phys. Rev., 2017, 4, 041305 Search PubMed.
- F. Rubio-Marcos, A. Del Campo, P. Marchet and J. F. Fernández, Ferroelectric domain wall motion induced by polarized light, Nat. Commun., 2015, 6, 6594 CrossRef CAS PubMed.
- F. Rubio-Marcos, D. A. Ochoa, A. Del Campo, M. A. García, G. R. Castro, J. F. Fernández and J. E. García, Reversible optical control of macroscopic polarization in ferroelectrics, Nat. Photonics, 2017, 12, 29–32 CrossRef.
- A. Bagri, A. Jana, G. Panchal, D. M. Phase and R. J. Choudhary, Amalgamation of Photostriction, Photodomain, and Photopolarization Effects in BaTiO3 and Its Electronic Origin, ACS Appl. Electron. Mater., 2022, 4, 4438–4445 CrossRef CAS.
- M. Ghidini, F. Maccherozzi, X. Moya, L. C. Phillips, W. Yan, J. Soussi, N. Métallier, M. E. Vickers, N. Steinke, R. Mansell, C. H. W. Barnes, S. S. Dhesi and N. D. Mathur, Perpendicular Local Magnetization Under Voltage Control in Ni Films on Ferroelectric BaTiO3 Substrates, Adv. Mater., 2015, 27, 1460–1465 CrossRef CAS PubMed.
- A. Bagri, A. Jana, G. Panchal, S. Chowdhury, R. Raj, M. Kumar, M. Gupta, V. R. Reddy, D. M. Phase and R. J. Choudhary, Light-Controlled Magnetoelastic Effects in Ni/BaTiO3 Heterostructures, ACS Appl. Mater. Interfaces, 2023, 15, 18391–18401 CrossRef CAS PubMed.
- R. V. Chopdekar and Y. Suzuki, Magnetoelectric coupling in epitaxial CoFe2O4 on BaTiO3, Appl. Phys. Lett., 2006, 89, 182506 CrossRef.
- S. Sahoo, S. Polisetty, C.-G. Duan, S. S. Jaswal, E. Y. Tsymbal and C. Binek, Ferroelectric control of magnetism in BaTiO3 heterostructures via interface strain coupling, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 092108 CrossRef.
- Z. Q. Liu, L. Li, Z. Gai, J. D. Clarkson, S. L. Hsu, A. T. Wong, L. S. Fan, M.-W. Lin, C. M. Rouleau, T. Z. Ward, H. N. Lee, A. S. Sefat, H. M. Christen and R. Ramesh, Full Electroresistance Modulation in a Mixed-Phase Metallic Alloy, Phys. Rev. Lett., 2016, 116, 097203 CrossRef CAS PubMed.
- Y. Xie, Q. Zhan, Y. Liu, G. Dai, H. Yang, Z. Zuo, B. Chen, B. Wang, Y. Zhang, X. Rong and R.-W. Li, Electric-field control of magnetic anisotropy in Fe81Ga19/BaTiO3 heterostructure films, AIP Adv., 2014, 4, 117113 CrossRef.
- T. Li and P. S. Lee, Piezoelectric Energy Harvesting Technology: From Materials, Structures, to Applications, Small Struct., 2022, 3, 2100128 CrossRef CAS.
- A. V. Ievlev, K. C. Santosh, R. K. Vasudevan, Y. Kim, X. Lu, M. Alexe, V. R. Cooper, S. V. Kalinin and O. S. Ovchinnikova, Non-conventional mechanism of ferroelectric fatigue via cation migration, Nat. Commun., 2019, 10, 3064 CrossRef PubMed.
- I. Stolichnov, A. Tagantsev, E. Colla, S. Gentil, S. Hiboux, J. Baborowski, P. Muralt and N. Setter, Downscaling of Pb(Zr,Ti)O3 film thickness for low-voltage ferroelectric capacitors: Effect of charge relaxation at the interfaces, J. Appl. Phys., 2000, 88, 2154–2156 CrossRef CAS.
- R. Salazar, M. Serrano and A. Abdelkefi, Fatigue in piezoelectric ceramic vibrational energy harvesting: A review, Appl. Energy, 2020, 270, 115161 CrossRef CAS.
- T. Buttlar, T. Walther, K. Dörr and S. G. Ebbinghaus, Preparation and Magnetoelectric Behavior of Ni/BaTiO3 Heterostructures with 0-3 Connectivity, Phys. Status Solidi B, 2020, 257, 1900622 CrossRef CAS.
- R. Revathy, N. Kalarikkal, M. R. Varma and K. P. Surendran, Exotic magnetic properties and enhanced magnetoelectric coupling in Fe3O4-BaTiO3 heterostructures, J. Alloys Compd., 2021, 889, 161667 CrossRef.
- W. Eerenstein, M. Wiora, J. L. Prieto, J. F. Scott and N. D. Mathur, Giant sharp and persistent converse magnetoelectric effects in multiferroic epitaxial heterostructures, Nat. Mater., 2007, 6, 348–351 CrossRef CAS PubMed.
- L. Aballe, M. Foerster, E. Pellegrin, J. Nicolas and S. Ferrer, The ALBA spectroscopic LEEM-PEEM experimental station: layout and performance, J. Synchrotron Radiat., 2015, 22, 745–752 CrossRef CAS PubMed.
- M. Foerster, J. Prat, V. Massana, N. Gonzalez, A. Fontsere, B. Molas, O. Matilla, E. Pellegrin and L. Aballe, Custom sample environments at the ALBA XPEEM, Ultramicroscopy, 2016, 171, 63–69 CrossRef CAS PubMed.
- M. Davis, M. Budimir, D. Damjanovic and N. Setter, Rotator and extender ferroelectrics: Importance of the shear coefficient to the piezoelectric properties of domain-engineered crystals and ceramics, J. Appl. Phys., 2007, 101, 054112 CrossRef.
-
A. L. Kholkin, N. A. Pertsev and A. V. Goltsev, Piezoelectricity and Crystal Symmetry, de Piezoelectric and Acoustic Materials for Transducer Applications, Springer, US, 2008, pp. 17–38 Search PubMed.
-
D. Roylance, Mechanics of materials, Wiley, New York, 1996 Search PubMed.
-
B. D. Cullity, Introduction to magnetic materials, ed. C. D. Graham, Hoboken, N. J: IEEE/Wiley, 2nd ed., 2009 Search PubMed.
-
M. J. Donachie, Titanium: A Technical Guide, ASM International, Ohio, 2nd ed., 2010 Search PubMed.
- J. G. Noel, Review of the properties of gold material for MEMS membrane applications, IET Circuits Dev. Syst., 2016, 10, 156–161 CrossRef.
- X. Li, G. Ding, T. Ando, M. Shikida and K. Sato, Micromechanical characterization of electroplated permalloy films for MEMS, Microsyst. Technol., April 2007, 14, 131–134 Search PubMed.
- S. Jiansirisomboon and A. Watcharapasorn, Effects of alumina nano-particulates addition on mechanical and electrical properties of barium titanate ceramics, Curr. Appl. Phys., 2008, 8, 48–52 CrossRef.
- T. M. Raeder, T. S. Holstad, I.-E. Nylund, M.-A. Einarsrud, J. Glaum, D. Meier and T. Grande, Anisotropic in-plane dielectric and ferroelectric properties of tensile-strained BaTiO3 films with three different crystallographic orientations, AIP Adv., 2021, 11, 025016 CrossRef CAS.
- D. Odkhuu, P. Taivansaikhan, W. S. Yun and S. C. Hong, A first-principles study of magnetostrictions of Fe3O4 and CoFe2O4, J. Appl. Phys., 2014, 115, 17A916 CrossRef.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.