This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Electrically controllable exchange bias via interface magnetoelectric effect

Adam B. Cahaya1,2,*, Ansell Alvarez Anderson1, Anugrah Azhar3,4 and Muhammad Aziz Majidi1 1Department of Physics, Faculty of Mathematics and Natural Sciences, Universitas Indonesia, Depok 16424, Indonesia 2Research Center for Quantum Physics, National Research and Innovation Agency (BRIN), South Tangerang 15314, Indonesia 3 Department of Physics and Astronomy, The University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom. 4Physics Study Program, Syarif Hidayatullah State Islamic University Jakarta, South Tangerang 15412, Indonesia *adam@sci.ui.ac.id
Abstract

Exchange bias is a unidirectional magnetic anisotropy that often arise from interfacial interaction of a ferromagnetic and antiferromagnetic layers. In this article, we show that a metallic layer with spin-orbit coupling can induces an exchange bias via an interface magnetoelectric effect. In linear response regime, the interface magnetoelectric effect is induced by spin-orbit couplings that arises from the broken symmetry of the system. Furthermore, we demonstrate that the exchange bias can be controlled by electric field.

Index Terms:
exchange bias, Rashba spin-orbit coupling, spin-electric effect

I Introduction

Magnetization manipulation in magnetic memory is one of the objectives of spintronics research area [1]. Due to low power consumption of voltage-driven magnetization dynamics [2], electrical control of magnetization has the potential for a more efficient manipulation of magnetic memories [3, 4]. The mechanism to couple voltage and magnetization includes the control of exchange bias using electric field [5].

Exchange bias has attracted much attention due to its applications on magnetic sensors and spintronic devices [6]. In magnetic heterostructure, exchange bias is a unidirectional anisotropy that can occur due to the hard magnetization behavior of an adjacent antiferromagnet [7]. The anisotropic exchange interaction at the interface of antiferromagnet and ferromagnet induces an exchange bias on the magnetization of the system [8]. The signature of exchange bias is the shift of the center of magnetic hysteresis loop from the origin[9]. The manipulation of exchange bias in magnetic heterostructure motivates innovative designs for spintronics devices [10]. The exchange bias in magnetic heterostructure is mediated by the conduction electron of a neighboring metallic layer [11]. The spin-orbit coupling due to a non-magnetic metallic layer can be utilized for manipulating exchange bias of the magnetic heterostructure [12].

The spin-orbit coupling may arise from noncentrosymmetry of the bulk [13] and Rashba effect at the interface [14, 15]. Furthermore, materials with noncentrosymmetry structure [16] and Rashba effect [17] has been shown to have magnetoelectric effect. The spin–orbit coupling leads to spin-dependent electric dipole moments of the electron orbitals, which results in magnetoelectric effect [16]. Magnetoelectric effect enables electric control of magnetic phase of multiferroic materials [18]. Moreover, magnetoelectric effect in a multiferroic heterostructure can lead to ultralow power magnetic memory [19].

Here we show that the spin-orbit coupling at the interface can induces an exchange bias in the neighboring ferromagnetic layer via an interface magnetoelectric effect. Sec. II discusses the linear response theory of interface magnetoelectric effect in a system with linear spin-orbit coupling due to noncentrosymmetry structure and Rashba effect. Sec. III discusses the exchange bias that arise from the interface magnetoelectric effect. Lastly, Sec. IV summarizes our findings.

II Interface magnetoelectric effect due to spin-orbit coupling

In second quantization, the interactions in the metallic system near the interface can be written with the following Hamiltonian

H=\displaystyle H= H0+Hint,\displaystyle H_{0}+H_{\rm int},
H0=\displaystyle H_{0}= kβγakβakγ[εkδβγ+αb𝝈βγk+αR𝝈βγ(k×z^)],\displaystyle\sum_{\textbf{k}\beta\gamma}a_{\textbf{k}\beta}^{\dagger}a_{\textbf{k}\gamma}\left[\varepsilon_{k}\delta_{\beta\gamma}+\alpha_{b}\boldsymbol{\sigma}_{\beta\gamma}\cdot\textbf{k}+\alpha_{R}\boldsymbol{\sigma}_{\beta\gamma}\cdot\left(\textbf{k}\times\hat{\textbf{z}}\right)\right],
Hint=\displaystyle H_{\rm int}= d3r[ρ(r)ϕ(r)M(r)B(r)].\displaystyle\int d^{3}r\left[\rho(\textbf{r})\phi(\textbf{r})-\textbf{M}(\textbf{r})\cdot\textbf{B}(\textbf{r})\right]. (1)

Here H0H_{0} is the unperturbed Hamiltonian [13]. HintH_{\rm int} is the interaction Hamiltonian, which represents the potential energy of electric charge density ρ\rho due to electric potential ϕ\phi and magnetization M due to magnetic field B [20]. ajβ(ajβ)a_{j\beta}^{\dagger}(a_{j\beta}) is the creation (annihilation) operator of conduction electron with wave vector k and spin β\beta, 𝝈\boldsymbol{\sigma} is Pauli vectors, εk=2k2/2m\varepsilon_{\textbf{k}}=\hbar^{2}k^{2}/2m is the energy dispersion of conduction electron. z^\hat{\textbf{z}} is normal to the interface. αb\alpha_{b} is the coupling constant of spin-orbit coupling of noncentrosymmetric metals [13, 21]. αR\alpha_{R} is the coupling constant of Rashba spin-orbit coupling due to broken symmetry at the interface [22, 23]. The spin-orbit coupling strength usually αb,RkF0.01\alpha_{b,R}k_{F}\sim 0.01 eV.

Magnetoelectric effect focuses on how magnetization

M(r)=\displaystyle\textbf{M}(\textbf{r})= μBkqβγeiqrak+qβ𝝈βγakγμBs(r)\displaystyle-\mu_{B}\sum_{\textbf{kq}\beta\gamma}e^{i\textbf{q}\cdot\textbf{r}}a_{\textbf{k}+\textbf{q}\beta}^{\dagger}\boldsymbol{\sigma}_{\beta\gamma}a_{\textbf{k}\gamma}\equiv-\mu_{B}\textbf{s}(\textbf{r}) (2)

and electric polarization densities P are coupled to magnetic B and electric E=ϕ(r)\textbf{E}=-\nabla\phi(\textbf{r}). μB\mu_{B} is the Bohr magneton. Here, P is related to charge density

ρ(r)=P=ekqβeiqrak+qβakβes0(r).\displaystyle\rho(\textbf{r})=-\nabla\cdot\textbf{P}=-e\sum_{\textbf{kq}\beta}e^{i\textbf{q}\cdot\textbf{r}}a_{\textbf{k}+\textbf{q}\beta}^{\dagger}a_{\textbf{k}\beta}\equiv-es_{0}(\textbf{r}). (3)

M and P due to B and E can be determined using linear response theory, in term of charge-spin response matrix XX

[s0(r,t)sx(r,t)sy(r,t)sz(r,t)]=\displaystyle\left[\begin{array}[]{c}s_{0}(\textbf{r},t)\\ s_{x}(\textbf{r},t)\\ s_{y}(\textbf{r},t)\\ s_{z}(\textbf{r},t)\end{array}\right]= d3r𝑑tX(rr,tt)[eϕ(r,t)μBBx(r,t)μBBy(r,t)μBBz(r,t)]\displaystyle\int d^{3}r^{\prime}dt^{\prime}X(\textbf{r}-\textbf{r}^{\prime},t-t^{\prime})\left[\begin{array}[]{c}-e\phi(\textbf{r},t^{\prime})\\ -\mu_{B}B_{x}(\textbf{r},t^{\prime})\\ -\mu_{B}B_{y}(\textbf{r},t^{\prime})\\ -\mu_{B}B_{z}(\textbf{r},t^{\prime})\end{array}\right] (12)
[s0(q,ω)sx(q,ω)sy(q,ω)sz(q,ω)]=\displaystyle\left[\begin{array}[]{c}s_{0}(\textbf{q},\omega)\\ s_{x}(\textbf{q},\omega)\\ s_{y}(\textbf{q},\omega)\\ s_{z}(\textbf{q},\omega)\end{array}\right]= X(q,ω)[eϕ(q,ω)μBBx(q,ω)μBBy(q,ω)μBBz(q,ω)],\displaystyle X(\textbf{q},\omega)\left[\begin{array}[]{c}-e\phi(\textbf{q},\omega)\\ -\mu_{B}B_{x}(\textbf{q},\omega)\\ -\mu_{B}B_{y}(\textbf{q},\omega)\\ -\mu_{B}B_{z}(\textbf{q},\omega)\end{array}\right], (21)

where f(q,ω)f(\textbf{q},\omega) is the Fourier transform of f(r,t)f(\textbf{r},t). The jkjk-component of XX is

Xjk(rr,tt)=iθ(tt)[sj(r,t),sk(r,t)].\displaystyle X_{jk}(\textbf{r}-\textbf{r}^{\prime},t-t^{\prime})=\frac{i}{\hbar}\theta(t-t^{\prime})\left[s_{j}(\textbf{r},t),s_{k}(\textbf{r}^{\prime},t^{\prime})\right]. (22)

One can see that X00X_{00} is related to the electric susceptibility. XjkX_{jk} (j,k0j,k\neq 0) is the magnetic susceptibility [24, 25, 26], its diagonal terms induces an anisotropic response [27, 28]. X0jX_{0j} and Xj0X_{j0} (j0j\neq 0) is related to the magnetoelectric susceptibility.

By evaluating the time derivative of XX using the unperturbed terms in II, one can show that the Fourier transform of X(r,t)X(\textbf{r},t) is

Xjk(r,t)=δjkkfkfk+qεk+qεk+ω+iτ1+δXjk,\displaystyle X_{jk}(\textbf{r},t)=\delta_{jk}\sum_{\textbf{k}}\frac{f_{\textbf{k}}-f_{\textbf{k}+\textbf{q}}}{\varepsilon_{\textbf{k}+\textbf{q}}-\varepsilon_{\textbf{k}}+\hbar\omega+i\tau^{-1}}+\delta X_{jk}, (23)

where fkf_{\textbf{k}} is the Fermi - Dirac distribution for electron with energy εk\varepsilon_{\textbf{k}}, τ\tau\to\infty is scattering time. δXjk\delta X_{jk} is the linear order correction due to the spin - orbit coupling

Refer to caption
Figure 1: A localized magnetic field B at the origin induces electric polarization P. The relative angle is determined by the ratio of spin-orbit coupling strengths αR\alpha_{R} and αb\alpha_{b}. The direction and magnitude are indicated by arrow and color. The polarization is localized due to the localization of x2sin2xx^{-2}\sin^{2}x function.
{widetext}
δX(q,ω)=kfkfk+q(εk+qεk+ω+iτ1)2\displaystyle\delta X(\textbf{q},\omega)=\sum_{\textbf{k}}\frac{f_{\textbf{k}}-f_{\textbf{k}+\textbf{q}}}{(\varepsilon_{\textbf{k}+\textbf{q}}-\varepsilon_{\textbf{k}}+\hbar\omega+i\tau^{-1})^{2}}
[0αbqxαRqyαbqy+αRqxαbqzαbqxαRqy0iαb(2kz+qz)iαb(2ky+qy)iαR(2kx+qx)αbqy+αRqxiαb(2kz+qz)0iαb(2kx+qx)iαR(2ky+qy)αbqziαb(2ky+qy)+iαR(2kx+qx)iαb(2ky+qy)+iαR(2kx+qx)0].\displaystyle\cdot\left[\begin{array}[]{cccc}0&-\alpha_{b}q_{x}-\alpha_{R}q_{y}&-\alpha_{b}q_{y}+\alpha_{R}q_{x}&-\alpha_{b}q_{z}\\ -\alpha_{b}q_{x}-\alpha_{R}q_{y}&0&-i\alpha_{b}(2k_{z}+q_{z})&i\alpha_{b}(2k_{y}+q_{y})-i\alpha_{R}(2k_{x}+q_{x})\\ -\alpha_{b}q_{y}+\alpha_{R}q_{x}&i\alpha_{b}(2k_{z}+q_{z})&0&-i\alpha_{b}(2k_{x}+q_{x})-i\alpha_{R}(2k_{y}+q_{y})\\ -\alpha_{b}q_{z}&-i\alpha_{b}(2k_{y}+q_{y})+i\alpha_{R}(2k_{x}+q_{x})&i\alpha_{b}(2k_{y}+q_{y})+i\alpha_{R}(2k_{x}+q_{x})&0\end{array}\right]. (28)

Substituting 28 to 21, one can show that αb\alpha_{b} and αR\alpha_{R} generate symmetric and antisymmetric responses, respectively

[P(q)M(q)]=ki(fkfk+q)(εk+qεk+iτ1)[3e2q2+𝒪(αb,αR)eμB(αbαRz^×)(εk+qεk+iτ1)eμB(αb+αRz^×)(εk+qεk+iτ1)μB2+𝒪(αb,αR)][E(q)B(q)]\displaystyle\left[\begin{array}[]{c}\textbf{P}(\textbf{q})\\ \textbf{M}(\textbf{q})\end{array}\right]=\sum_{\textbf{k}}\frac{i\left(f_{\textbf{k}}-f_{\textbf{k}+\textbf{q}}\right)}{(\varepsilon_{\textbf{k}+\textbf{q}}-\varepsilon_{\textbf{k}}+i\tau^{-1})}\left[\begin{array}[]{cc}\frac{3e^{2}}{q^{2}}+\mathcal{O}(\alpha_{b},\alpha_{R})&\displaystyle\frac{e\mu_{B}\left(-\alpha_{b}-\alpha_{R}\hat{\textbf{z}}\times\right)}{(\varepsilon_{\textbf{k}+\textbf{q}}-\varepsilon_{\textbf{k}}+i\tau^{-1})}\\ \displaystyle\frac{e\mu_{B}\left(-\alpha_{b}+\alpha_{R}\hat{\textbf{z}}\times\right)}{(\varepsilon_{\textbf{k}+\textbf{q}}-\varepsilon_{\textbf{k}}+i\tau^{-1})}&\mu_{B}^{2}+\mathcal{O}(\alpha_{b},\alpha_{R})\end{array}\right]\left[\begin{array}[]{c}\textbf{E}(\textbf{q})\\ \textbf{B}(\textbf{q})\end{array}\right] (35)

When there is no spin-orbit coupling, B and E only responsible for M and P, respectively. When αb,αR0\alpha_{b},\alpha_{R}\neq 0, one find that E and B also generates M and P, respectively

P(r)\displaystyle\textbf{P}(\textbf{r}) =d3reμBm224π3sin2kF|rr||rr|2(αbαRz^×)B(r),\displaystyle=\int d^{3}r^{\prime}\frac{e\mu_{B}m^{2}}{2\hbar^{4}\pi^{3}}\frac{\sin^{2}k_{F}\left|\textbf{r}-\textbf{r}^{\prime}\right|}{\left|\textbf{r}-\textbf{r}^{\prime}\right|^{2}}\left(-\alpha_{b}-\alpha_{R}\hat{\textbf{z}}\times\right)\textbf{B}(\textbf{r}^{\prime}),
M(r)\displaystyle\textbf{M}(\textbf{r}) =d3reμBm224π3sin2kF|rr||rr|2(αb+αRz^×)E(r).\displaystyle=\int d^{3}r^{\prime}\frac{e\mu_{B}m^{2}}{2\hbar^{4}\pi^{3}}\frac{\sin^{2}k_{F}\left|\textbf{r}-\textbf{r}^{\prime}\right|}{\left|\textbf{r}-\textbf{r}^{\prime}\right|^{2}}\left(-\alpha_{b}+\alpha_{R}\hat{\textbf{z}}\times\right)\textbf{E}(\textbf{r}^{\prime}). (36)

We can see that the polarizations due to bulk spin-orbit coupling contribution are parallel to the field. On the other hand, the polarizations induced by interface Rashba spin-orbit coupling are perpendicular to the fields. The relative angle of polarizations to the fields is determined by the ratio of interface and bulk spin-orbit coupling strength. Fig. 1 illustrates electric polarization densities due to a localized in-plane magnetic fields B=By^\textbf{B}=B\hat{\textbf{y}}, which can arise from an exchange interaction with a localized spin. The induced polarization is localized due to the localization of x2sin2xx^{-2}\sin^{2}x function. Because of that, it can be assumed that the leading terms of the polarizations is weakly influenced by the periodicity of the system. To avoid divergences when integrated over large volume, x2sin2xx^{-2}\sin^{2}x will be approximated using its steepest descent [29]

sin2kFr(kFr)2ekF2r2/3=d3q(2π)3eiqr(3π)3/2e3q2/4kF2kF3.\frac{\sin^{2}k_{F}r}{(k_{F}r)^{2}}\approx e^{-k_{F}^{2}r^{2}/3}=\int\frac{d^{3}q}{(2\pi)^{3}}e^{i\textbf{q}\cdot\textbf{r}}\frac{(3\pi)^{3/2}e^{-3q^{2}/4k_{F}^{2}}}{k_{F}^{3}}.

III Exchange bias due to interface magnetoelectric effect

In this section, we focus on a bilayer system that consists of a magnetic layer and a non-magnetic metallic layer. Near the interface, there is a localized magnetic field [30]

B(r)=JMnδ3(rrn)\textbf{B}(\textbf{r})=J\textbf{M}\sum_{n}\delta^{3}(\textbf{r}-\textbf{r}_{n}) (37)

due to the sds-d exchange interaction between localized magnetic moment M at position rn\textbf{r}_{n} and the spin of conduction electron. J=2μ0/3J=2\mu_{0}/3 is the exchange constant, which can be estimated from the localized term of dipolar magnetic field [31, 32].

Bdipolar=2μ03Mδ(r)μ03(Mr^)r^M4πr3,B_{\rm dipolar}=-\frac{2\mu_{0}}{3}\textbf{M}\delta(\textbf{r})-\mu_{0}\frac{3\left(\textbf{M}\cdot\hat{\textbf{r}}\right)\hat{\textbf{r}}-\textbf{M}}{4\pi r^{3}},

μ0\mu_{0} is the vacuum permeability.

Substituting 37 to 36, we can find that P depends only on zz

P(r)\displaystyle\textbf{P}(\textbf{r}) [αb+αRz^×]JMneμBm2kF224π3ekF2|rrn|2/3\displaystyle\simeq-\left[\alpha_{b}+\alpha_{R}\hat{\textbf{z}}\times\right]J\textbf{M}\sum_{n}\frac{e\mu_{B}m^{2}k_{F}^{2}}{2\hbar^{4}\pi^{3}}e^{-k_{F}^{2}\left|\textbf{r}-\textbf{r}_{n}\right|^{2}/3}
=[αb+αRz^×]JM3eμBm2N24π2AekF2z2/3,\displaystyle=-\left[\alpha_{b}+\alpha_{R}\hat{\textbf{z}}\times\right]J\textbf{M}\frac{3e\mu_{B}m^{2}N}{2\hbar^{4}\pi^{2}A}e^{-k_{F}^{2}z^{2}/3}, (38)

where N/AN/A is number of magnetic moment per unit area. Additionally, from 36 we can see that a uniform electric field E induces a uniform M

M(r)\displaystyle\textbf{M}(\textbf{r}) [αb+αRz^×]Ed3reμBm2kF224π3ekF2|rr|2/3\displaystyle\simeq\left[-\alpha_{b}+\alpha_{R}\hat{\textbf{z}}\times\right]\textbf{E}\int d^{3}r^{\prime}\frac{e\mu_{B}m^{2}k_{F}^{2}}{2\hbar^{4}\pi^{3}}e^{-k_{F}^{2}\left|\textbf{r}-\textbf{r}^{\prime}\right|^{2}/3}
=[αb+αRz^×]E332eμBm224π32kF.\displaystyle=\left[-\alpha_{b}+\alpha_{R}\hat{\textbf{z}}\times\right]\textbf{E}\frac{3^{\frac{3}{2}}e\mu_{B}m^{2}}{2\hbar^{4}\pi^{\frac{3}{2}}k_{F}}. (39)

Substituting 38 and 39 to the magnetic and electric interaction terms in II, one can arrive at exchange bias Hamiltonian

Heb=\displaystyle H_{eb}= d3rM(r)JnM(rrn)d3rP(r)E\displaystyle-\int d^{3}r\textbf{M}(\textbf{r})\cdot J\sum_{n}\textbf{M}(\textbf{r}-\textbf{r}_{n})-\int d^{3}r\textbf{P}(\textbf{r})\cdot\textbf{E}
=\displaystyle= MB0,\displaystyle\textbf{M}\cdot\textbf{B}_{0}, (40)

where

B0=[αbαRz^×]E352eμBm2JN44π32kF,\displaystyle\textbf{B}_{0}=\left[\alpha_{b}-\alpha_{R}\hat{\textbf{z}}\times\right]\textbf{E}\frac{3^{\frac{5}{2}}e\mu_{B}m^{2}JN}{4\hbar^{4}\pi^{\frac{3}{2}}k_{F}}, (41)

correspond to an exchange bias shift of the center of the magnetic hysteresis. The dependency of B0\textbf{B}_{0} to E indicates that the exchange bias is controllable by electric field.

Refer to caption
Figure 2: (a) In La2/3Sr1/3MnO|3{}_{3}|LaAlO|3{}_{3}|SrTiO3 structure, (a) Exchange bias is not observed when there is no surface charges. (b) an out-of-plane electric field generated by 2 dimensional electron gas system at the interfaces [33] induces exchange bias. (c) The exchange bias is observed as shift in the magnetic hysteresis curve.

Fig. 2 illustrates the magnetoelectric coupling of out-of-plane electric field and magnetization in a bilayer system with αbαR\alpha_{b}\gg\alpha_{R}. The hysteresis curve is illustrated using Stoner-Wohlfarth model with anisotropy KK, saturation magnetization MSM_{S} and angle between easy axis and field θ=45\theta=45^{\circ}. The inverse dependency of B0\textbf{B}_{0} to kFk_{F} is preferable for lightly-doped semiconductor, such as LaAlO3 with kF0.3a1k_{F}\sim 0.3a^{-1}, a=3.8a=3.8Å is the lattice constant [34]. The out-of-plane electric field can be generated from exchange bias or from charge transfer in La2/3Sr1/3MnO|3{}_{3}|LaAlO|3{}_{3}|SrTiO3 [33]. For αbkF=0.01\alpha_{b}k_{F}=0.01 eV, the magnitude B0B_{0} can be estimated to be

B0=αbE352eμBm2JN44π32kF=αb332QeμBμ0m224π32kFa2εrε0=18mTB_{0}=\alpha_{b}E\frac{3^{\frac{5}{2}}e\mu_{B}m^{2}JN}{4\hbar^{4}\pi^{\frac{3}{2}}k_{F}}=\alpha_{b}\frac{3^{\frac{3}{2}}Qe\mu_{B}\mu_{0}m^{2}}{2\hbar^{4}\pi^{\frac{3}{2}}k_{F}a^{2}\varepsilon_{r}\varepsilon_{0}}=18\ \mathrm{mT}

Here we used E=Q/(εrε0a2)E=Q/(\varepsilon_{r}\varepsilon_{0}a^{2}), ε\varepsilon is vacuum permittivity, εr=25\varepsilon_{r}=25 is the dielectric constant of LaAlO3[35], Q=0.5Q=0.5 e [33]. This result is in agreement with experiment by Ref. [36], which observed that there is an exchange bias in La2/3Sr1/3MnO|3{}_{3}|LaAlO|3{}_{3}|SrTiO3 structure when the thickness LaAlO3 is more than 4 unit cell, with B020B_{0}\sim 20 mT. This phenomena occurs because 2 dimensional electron gas emerges at the interface of LaAlO|3{}_{3}|SrTiO3 when the thickness of LaAlO3 is more than 4 unit cell [37, 33].

Refer to caption
Figure 3: Manipulation of exchange bias by in-plane electric field via interface spin-orbit coupling. (a) Relative directions of electric field EE and magnetization M induces (b) exchange bias of the magnetic heterostructure.

Fig. 3 illustrates a bilayer system with αRαb\alpha_{R}\gg\alpha_{b} and an in-plane magnetization. In this case, the exchange bias of in-plane magnetization can be controlled by in-plane electric fields. The direction of electric field is perpendicular to magnetization direction. A more efficient manipulation of magnetic memory can be further developed by combining electrical control of exchange bias and spin-orbit torque [38], because magnetizations also manipulated by transverse electric field in spin-orbit torques devices [39].

IV Summary and conclusion

To summarize, we study the origin of interface magnetoelectric and its application for electrical control of exchange bias. The interface magnetoelectric arises from the spin-orbit couplings due to broken symmetry of the magnetic structure. We consider spin-orbit couplings that arise from noncentrosymmetry of the bulk structure and Rashba effect at the interface. The magnetoelectric effect associated with spin-orbit coupling due to noncentrosymmetry can describe the exchange bias observed in La2/3Sr1/3MnO|3{}_{3}|LaAlO|3{}_{3}|SrTiO3 structure, as illustrated in Fig. 2. On the other hand, the magnetoelectric effect associated with spin-orbit coupling due to interface Rashba effect describes the coupling of in-plane magnetization with transverse electric, as illustrated in Fig. 3. The manipulation of exchange bias by transverse electric may be combined with spin-orbit torque to for a more efficient magnetic memory and spintronic devices.

Acknowledgement

We thank Indonesia Toray Science Foundation for funding this research through Science & Technology Research Grant.

References

  • [1] B. Dieny et al., “Opportunities and challenges for spintronics in the microelectronics industry,” Nature Electronics, vol. 3, no. 8, pp. 446–459, Aug. 2020.
  • [2] T. Nozaki et al., “Electric-field-induced ferromagnetic resonance excitation in an ultrathin ferromagnetic metal layer,” Nature Physics, vol. 8, no. 6, pp. 491–496, Apr. 2012.
  • [3] A. Chen, D. Zheng, B. Fang, Y. Wen, Y. Li, and X.-X. Zhang, “Electrical manipulation of magnetization in magnetic heterostructures with perpendicular anisotropy,” Journal of Magnetism and Magnetic Materials, vol. 562, p. 169753, Nov. 2022.
  • [4] C. Song, B. Cui, F. Li, X. Zhou, and F. Pan, “Recent progress in voltage control of magnetism: Materials, mechanisms, and performance,” Progress in Materials Science, vol. 87, pp. 33–82, Jun. 2017.
  • [5] S. M. Wu, S. A. Cybart, D. Yi, J. M. Parker, R. Ramesh, and R. C. Dynes, “Full electric control of exchange bias,” Phys. Rev. Lett., vol. 110, p. 067202, Feb 2013.
  • [6] F. Torres, R. Morales, I. K. Schuller, and M. Kiwi, “Dipole-induced exchange bias,” Nanoscale, vol. 9, pp. 17 074–17 079, Oct. 2017.
  • [7] J. Nogués and I. K. Schuller, “Exchange bias,” Journal of Magnetism and Magnetic Materials, vol. 192, no. 2, pp. 203–232, Feb. 1999.
  • [8] R. L. Stamps, “Mechanisms for exchange bias,” Journal of Physics D: Applied Physics, vol. 33, no. 23, pp. R247–R268, Nov. 2000.
  • [9] M. Kiwi, “Exchange bias theory,” Journal of Magnetism and Magnetic Materials, vol. 234, no. 3, pp. 584–595, Sep. 2001.
  • [10] P.-H. Lin et al., “Manipulating exchange bias by spin–orbit torque,” Nature Materials, vol. 18, no. 4, pp. 335–341, Feb. 2019.
  • [11] N. J. Gökemeijer, T. Ambrose, and C. L. Chien, “Long-range exchange bias across a spacer layer,” Phys. Rev. Lett., vol. 79, pp. 4270–4273, Nov. 1997.
  • [12] S. Peng et al., “Exchange bias switching in an antiferromagnet/ferromagnet bilayer driven by spin–orbit torque,” Nature Electronics, vol. 3, no. 12, pp. 757–764, Nov. 2020.
  • [13] S.-X. Wang, H.-R. Chang, and J. Zhou, “Rkky interaction in three-dimensional electron gases with linear spin-orbit coupling,” Phys. Rev. B, vol. 96, p. 115204, Sep. 2017.
  • [14] W. Lin et al., “Interface-based tuning of rashba spin-orbit interaction in asymmetric oxide heterostructures with 3d electrons,” Nature Communications, vol. 10, no. 1, Jul. 2019.
  • [15] G. Bihlmayer, O. Rader, and R. Winkler, “Focus on the rashba effect,” New Journal of Physics, vol. 17, no. 5, p. 050202, May 2015.
  • [16] V. P. Sakhnenko and N. V. Ter-Oganessian, “The magnetoelectric effect due to local noncentrosymmetry,” Journal of Physics: Condensed Matter, vol. 24, no. 26, p. 266002, May 2012.
  • [17] T. Ojanen, “Magnetoelectric effects in superconducting nanowires with rashba spin-orbit coupling,” Phys. Rev. Lett., vol. 109, p. 226804, Nov. 2012.
  • [18] M. Fiebig, “Revival of the magnetoelectric effect,” Journal of Physics D: Applied Physics, vol. 38, no. 8, p. R123, Apr. 2005.
  • [19] S. Fujii et al., “Giant converse magnetoelectric effect in a multiferroic heterostructure with polycrystalline co2fesi,” NPG Asia Materials, vol. 14, no. 1, May 2022.
  • [20] D. J. Griffiths, Introduction to electrodynamics, 3rd ed.   Upper Saddle River, NJ: Pearson, Dec. 1998.
  • [21] H. Weng, C. Fang, Z. Fang, B. A. Bernevig, and X. Dai, “Weyl semimetal phase in noncentrosymmetric transition-metal monophosphides,” Phys. Rev. X, vol. 5, p. 011029, Mar. 2015.
  • [22] G. Bihlmayer, O. Rader, and R. Winkler, “Focus on the rashba effect,” New Journal of Physics, vol. 17, no. 5, p. 050202, May 2015.
  • [23] H. C. Koo et al., “Rashba effect in functional spintronic devices,” Advanced Materials, vol. 32, no. 51, p. 2002117, Sep. 2020.
  • [24] A. B. Cahaya, “Paramagnetic and diamagnetic susceptibility of infinite quantum well,” Al-Fiziya: Journal of Materials Science, Geophysics, Instrumentation and Theoretical Physics, Dec. 2020.
  • [25] A. B. Cahaya, R. M. Sitorus, A. Azhar, A. R. T. Nugraha, and M. A. Majidi, “Enhancement of spin-mixing conductance by sds\text{$-$}d orbital hybridization in heavy metals,” Phys. Rev. B, vol. 105, p. 214438, Jun. 2022.
  • [26] A. B. Cahaya, “Adiabatic limit of rkky range function in one dimension,” Journal of Magnetism and Magnetic Materials, vol. 547, p. 168874, Apr. 2022.
  • [27] A. B. Cahaya and M. A. Majidi, “Spin accumulation induced anisotropic rkky interaction,” Journal of Physics: Conference Series, vol. 1816, no. 1, p. 012074, Feb. 2021.
  • [28] ——, “Effects of screened coulomb interaction on spin transfer torque,” Phys. Rev. B, vol. 103, p. 094420, Mar. 2021.
  • [29] A. B. Cahaya, “Improved steepest descent method using modified bessel function K1/4\textit{K}_{1/4} for gamma function evaluation,” Al-Fiziya Journal of Materials Science, Geophysics, Instrumentation and Theoretical Physics, vol. 4, p. 89, Dec. 2021.
  • [30] A. B. Cahaya, A. O. Leon, and G. E. W. Bauer, “Crystal field effects on spin pumping,” Phys. Rev. B, vol. 96, p. 144434, Oct. 2017.
  • [31] W. Kutzelnigg, “Origin and meaning of the fermi contact interaction,” Theoretica Chimica Acta, vol. 73, no. 2-3, pp. 173–200, 1988.
  • [32] A. B. Cahaya, “Antiferromagnetic spin pumping via hyperfine interaction,” Hyperfine Interactions, vol. 242, no. 1, Dec. 2021.
  • [33] T. Asmara et al., “Mechanisms of charge transfer and redistribution in LaAlO3/SrTiO3 revealed by high-energy optical conductivity,” Nature Communications, vol. 5, no. 1, Apr. 2014.
  • [34] A. A. Anderson, A. Azhar, A. B. Cahaya, and M. A. Majidi, “Rkky interaction in spin valve structure with laalo3 spacer,” Materials Science Forum, vol. 1080, pp. 131–137, 2 2023.
  • [35] M. Suzuki, “Comprehensive study of lanthanum aluminate high-dielectric-constant gate oxides for advanced CMOS devices,” Materials, vol. 5, no. 12, pp. 443–477, Mar. 2012.
  • [36] W. M. Lü et al., “Long-range magnetic coupling across a polar insulating layer,” Nature Communications, vol. 7, no. 1, p. 11015, Mar. 2016.
  • [37] A. D. Caviglia et al., “Two-dimensional quantum oscillations of the conductance at laalo3/srtio3{\mathrm{laalo}}_{3}/{\mathrm{srtio}}_{3} interfaces,” Phys. Rev. Lett., vol. 105, p. 236802, Dec. 2010.
  • [38] W. Zhang and K. M. Krishnan, “Epitaxial exchange-bias systems: From fundamentals to future spin-orbitronics,” Materials Science and Engineering: R: Reports, vol. 105, pp. 1–20, Jul. 2016.
  • [39] Q. Shao et al., “Roadmap of spin–orbit torques,” IEEE Transactions on Magnetics, vol. 57, no. 7, pp. 1–39, Jul. 2021.