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

Spin-orbit torque switching of Néel order in two-dimensional \chCrI3

Fei Xue Physical Measurement Laboratory, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA Institute for Research in Electronics and Applied Physics & Maryland Nanocenter, University of Maryland, College Park, MD 20742    Paul M. Haney Physical Measurement Laboratory, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA
Abstract

Spin-orbit torque enables electrical control of the magnetic state of ferromagnets or antiferromagnets. In this work we consider the spin-orbit torque in the 2-d Van der Waals antiferromagnetic bilayer \chCrI3, in the nn-doped regime. In the purely antiferromagnetic state, two individually inversion-symmetry broken layers of \chCrI3 form inversion partners, like the well-studied CuMnAs and \chMn2Au. However, the exchange and anisotropy energies are similar in magnitude, unlike previously studied antiferromagnets, which leads to qualitatively different behaviors in this material. Using a combination of first-principles calculations of the spin-orbit torque and an analysis of the ensuing spin dynamics, we show that the deterministic electrical switching of the Néel vector is the result of dampinglike spin-orbit torque, which is staggered on the magnetic sublattices.

Introduction.— Spin-orbit torque is a mechanism for electrically switching thin-film magnets, and has the potential to enable scalable magnetic random access memory and devices for next-generation computing  [1]. The effect occurs in magnetically ordered systems that lack inversion symmetry - such as heavy metal-ferromagnet bilayers [2, 3] - when a DC current or electric field is applied. Spin-orbit torque can be decomposed into a component that is even under time-reversal, which is also known as the “dampinglike” torque, and a component that is odd under time-reversal, known as the “fieldlike” torque [4]. Knowledge of the dominant component of spin-orbit torque can help to identify the microscopic source of the torque and can assist in optimizing the effect [5].

In addition to switching ferromagnets, spin-orbit torque has been shown to switch antiferromagnets [6, 7, 8, 9, 10, 11]. Antiferromagnets are of particular interest due to their insensitivity to stray magnetic fields and the fast time scales of their excitations [8, 10, 1]. It was shown [6] that spin-orbit torque is present in bulk antiferromagnets in which inversion symmetry is locally broken on individual magnetic sublattices, while the crystal lattice retains global inversion symmetry. More precisely, in antiferromagnets that are invariant under the combined operations of inversion and time-reversal, the spin-orbit torques acting on the magnetic sublattices cooperatively switch the antiferromagnetic Néel vector 𝐋{\bf L} [6, 7]. In these types of materials studied so far, such as CuMnAs and \chMn2Au, the magnetic exchange energy is much larger than other energy scales, and the mechanism for switching is a uniform fieldlike torque present on both magnetic sublattices [6, 7].

Refer to caption
Figure 1: (Color online) (a) shows a top-down view of one layer of \chCrI3. The second layer (not shown) is displaced along the xx-direction by a nearest-neighbor distance. (b) Side view of \chCrI3. Note a lack of symmetry with respect to xxx\rightarrow-x. Other symmetries depend on the spin configuration: for a purely antiferromagnetic state, the system is invariant under inversion+time-reversal. For a state with canting in the yy direction, the system has a 2-fold rotational symmetry about the yy-axis. (c) Spin configurations on magnetic sublattices A and B considered in this work, with finite canting in the yy-direction. (d) Mixed representation of system spin in 𝐍^=(Lx,My,Lz)\hat{{\bf N}}=(L_{x},M_{y},L_{z}) space, showing the spin-orbit torque switching trajectory of 𝐍^\hat{{\bf N}} for applied electric field in the yy-direction.

In this work we consider the spin-orbit torque in the recently discovered class of two-dimensional Van der Waals magnetic materials [12, 13, 14], exemplified by antiferromagnetic bilayer \chCrI3. In this semiconducting material, the two magnetic \chCrI3 layers are antiferromagnetically coupled and the ground state Néel vector is oriented perpendicular to the plane [12] (see Fig. 1). There is intense recent interest in this material due to its unique and potentially useful properties, including the tunability and control of its magnetic state through gating and doping [15, 16, 17] and its spin-filtering effects [18, 19, 20]. From the structure shown in Fig. 1(b), it’s clear that bilayer \chCrI3 shares some characteristics with previously mentioned antiferromagnets, such as CuMnAs: Inversion symmetry is locally broken on the magnetic sublattices (denoted A and B), while in the purely antiferromagnetic state, the bilayer is invariant under the combined operations of inversion plus time-reversal.

There is a key difference between this and previously studied materials, namely that the magnetic exchange constant is similar in magnitude to the magnetic anisotropy energy [21, 22]. This allows for significant current-induced canting of the sublattice magnetization and the development of a net magnetization, which breaks the combination of inversion plus time-reversal symmetry. For this material, we show that the important symmetry is instead a 2-fold rotational symmetry about the yy-axis; this symmetry constrains the magnetic dynamics to a subspace of spin configurations and enables spin-orbit torque switching of the Néel vector. We also show that the dominant mechanism for spin-orbit torque switching of the Néel vector is the dampinglike torque, which is staggered on the two sublattices of the antiferromagnet. These conclusions are based on first-principles calculations of the spin-orbit torque as a function of the magnetic configuration, and an analysis of the ensuing spin dynamics.

Spin dynamics in CrI3 — We first consider the spin dynamics in \chCrI3 and show that the crystal symmetry confines the spins a subspace of magnetic configurations. In the general case, the time evolution of the spin orientations 𝐦^A,B\hat{\mathbf{m}}^{\rm A,B} are described by the coupled set of Landau-Lifshitz-Gilbert (LLG) equations [23, 24, 25]:

d𝐦^A,Bdt\displaystyle\frac{d{\hat{\mathbf{m}}}^{\rm A,B}}{dt} =\displaystyle= 𝐦^A,B×(γmδEδ𝐦^A,B+αd𝐦^A,Bdt)+𝓣A,B,\displaystyle\hat{\mathbf{m}}^{\rm A,B}\times\left(\frac{\gamma}{m}\frac{\delta E}{\delta\hat{\mathbf{m}}^{\rm A,B}}+\alpha~{}\frac{d{\hat{\mathbf{m}}}^{\rm A,B}}{dt}\right)+\boldsymbol{\mathcal{T}}^{\rm A,B},

where mm is the magnitude of the magnetic moment (assumed equal on both sublattices), γ\gamma is the absolute value of the gyromagnetic ratio, and α\alpha is the Gilbert damping parameter. The energy EE is comprised of an easy-axis anisotropy (along 𝐳^\hat{\bf z}) and Heisenberg exchange coupling: E(𝐦^A,𝐦^B)=12mHA[(𝐦^A𝐳^)2+(𝐦^B𝐳^)2]+mHE(𝐦^A𝐦^B)E(\hat{\mathbf{m}}^{\rm A},\hat{\mathbf{m}}^{\rm B})=-\frac{1}{2}mH_{A}[\left(\hat{\mathbf{m}}^{\rm A}\cdot\hat{\bf z}\right)^{2}+\left(\hat{\mathbf{m}}^{\rm B}\cdot\hat{\bf z}\right)^{2}]+mH_{E}\left(\hat{\mathbf{m}}^{\rm A}\cdot\hat{\mathbf{m}}^{\rm B}\right), where HAH_{A} and HEH_{E} are the effective magnetic fields from anisotropy and exchange, respectively. 𝓣A,B\boldsymbol{\mathcal{T}}^{\rm A,B} is the spin-orbit torque on the A,B sublattice, which we discuss in more detail in the next section.

The dynamics of the coupled spin system can be quite complex in the general case and analytical treatments are often not possible. A commonly studied limit is HEHAH_{E}\gg H_{A}, which is applicable to many metallic antiferromagnets. In this case the system can be simplified to a single vector equation of motion for the Néel vector 𝐋=(𝐦A𝐦B)/2{\bf L}=({\bf m}^{\rm A}-{\bf m}^{\rm B})/2, while the net magnetization 𝐌=(𝐦A+𝐦B)/2{\bf M}=({\bf m}^{\rm A}+{\bf m}^{\rm B})/2 is no longer an independent variable [24, 25]. For \chCrI3, HEHAH_{E}\approx H_{A} [21] and this simplification is no longer valid. We next discuss an alternative simplification that follows the system crystal symmetry.

We consider an applied electric field along the yy-axis and spin configurations that are staggered in (x,z)(x,z) and uniform in yy: mx/zA=mx/zBm^{\rm A}_{x/z}=-m^{\rm B}_{x/z} and myA=myBm^{\rm A}_{y}=m^{\rm B}_{y}. In this case, the system retains 2-fold rotation symmetry about the yy-axis (see Fig. 1). Any torque on the spins is therefore symmetry-constrained to satisfy 𝒯x/zA=𝒯x/zB\mathcal{T}^{\rm A}_{x/z}=-\mathcal{T}^{\rm B}_{x/z} and 𝒯yA=𝒯yB\mathcal{T}^{\rm A}_{y}=\mathcal{T}^{\rm B}_{y}. The xx and zz components of the spins then remain staggered and the yy components remain uniform. The trajectory of the spins is thus symmetry-confined to the subspace (Lx,My,Lz)(L_{x},M_{y},L_{z}). This motivates the definition of a “mixed” order parameter 𝐍(Lx,My,Lz){\bf N}\equiv\left(L_{x},M_{y},L_{z}\right) [26]. Eq. Spin-orbit torque switching of Néel order in two-dimensional \chCrI3 leads to the following equation of motion for 𝐍^\hat{{\bf N}}:

d𝐍^dt\displaystyle\frac{d{\hat{\mathbf{N}}}}{dt} =\displaystyle= 𝐍^×(γmδEδ𝐍^+αd𝐍^dt)+𝒯odd(𝐍^×𝐩^)\displaystyle{\hat{\mathbf{N}}}\times\left(\frac{\gamma}{m}\frac{\delta E}{\delta{\hat{\mathbf{N}}}}+\alpha\frac{d{\hat{\mathbf{N}}}}{dt}\right)+\mathcal{T}^{\rm odd}\left({\hat{\mathbf{N}}}\times\hat{\bf p}\right) (2)
+𝒯even(𝐍^×(𝐩^×𝐍^)),\displaystyle~{}~{}~{}~{}+\mathcal{T}^{\rm even}({\hat{\mathbf{N}}}\times(\hat{\bf p}\times{\hat{\mathbf{N}}})),

where the energy is comprised of the easy-axis anisotropy along 𝐳^\hat{\bf z} and an effective hard-axis anisotropy along 𝐲^\hat{\bf y}, which encodes the magnetic exchange:

E(𝐍^)\displaystyle E(\hat{\mathbf{N}}) =\displaystyle= 12mHA(𝐍^𝐳^)2+mHE(𝐍^𝐲^)2.\displaystyle-\frac{1}{2}mH_{A}\left(\hat{\mathbf{N}}\cdot\hat{\mathbf{z}}\right)^{2}+mH_{E}\left(\hat{\mathbf{N}}\cdot\hat{\mathbf{y}}\right)^{2}. (3)

The last two terms on the right-hand-sided of Eq. 2 are the spin-orbit torques, consisting of terms that are odd and even under time reversal, as indicated (note the prefactors 𝒯odd,even\mathcal{T}^{\rm odd,even} are assumed to be constants, independent of 𝐍^\hat{{\bf N}}). 𝐩^\hat{\bf p} is a direction related to the system symmetry, and the spin-orbit torque vanishes when 𝐍^\hat{{\bf N}} is aligned to 𝐩^\hat{\bf p}. We discuss the spin-orbit torque terms in detail in the next section; for now they can be taken as a general representation of a torque which is spanned by the two vectors perpendicular to 𝐍{\bf N}. We have verified that fluctuations away from the 𝐍{\bf N} subspace do not alter the steady state dynamics [27]. One important feature of this system which enables this simplification is that the easy-axis anisotropy is perpendicular to the axis of 2-fold rotation symmetry. If this were not the case, then the anisotropy torque would immediately drive the spin configuration out of the subspace.

The simple form of the time evolution of 𝐍^\hat{{\bf N}} allows for an analytical treatment of fixed points and an intuitive description of the dynamics. For 𝐩^\hat{\bf p} along 𝐳^\hat{\bf z}, the dampinglike torque 𝒯even\mathcal{T}^{\rm even} competes with γα(HE+HA)\gamma\alpha(H_{E}+H_{A}) while the fieldlike torque 𝒯odd\mathcal{T}^{\rm odd} competes with γHA\gamma H_{A}. For \chCrI3, HA1.77TH_{A}\approx 1.77~{}\rm T, HE0.76TH_{E}\approx 0.76~{}\rm T, and α0.04\alpha\approx 0.04 [21], so that the dampinglike torque is the dominant mechanism for switching. In the next section we show that 𝐩^\hat{\bf p} has the standard xx component, and due to additional mirror plane symmetry breaking in \chCrI3, also has a zz component. For 𝐩^=(px,0,pz)\hat{\bf p}=\left(p_{x},0,p_{z}\right), the fixed points to lowest order in spin-orbit torque are 𝐍^=(𝒯oddpxγHA,±𝒯evenpxγ(2HE+HA),±1)\hat{{\bf N}}=(-\frac{\mathcal{T}^{\rm odd}p_{x}}{\gamma H_{A}},\pm\frac{\mathcal{T}^{\rm even}p_{x}}{\gamma(2H_{E}+H_{A})},\pm 1). The instability threshold to switch between fixed points is:

|𝒯evenpz|>γα(HE+HA).|\mathcal{T}^{\rm even}p_{z}|>\gamma\alpha(H_{E}+H_{A}). (4)

A typical switching trajectory is shown in Fig. 1(d): the spin-orbit torque drives 𝐍^\hat{{\bf N}} from north pole to the fixed point close to south pole. For previously studied antiferromagnets, HEH_{E} is sufficiently large so that switching with 𝒯even\mathcal{T}^{\rm even} is not feasible, and the switching relies instead on 𝒯odd\mathcal{T}^{\rm odd} exceeding HAH_{A}. The utilization of 𝒯even\mathcal{T}^{\rm even} for switching the Néel vector of \chCrI3 is one of the material’s distinguishing features.

Microsopic calculations of spin-orbit torques — Having established the relevant degrees of freedom for the spin configuration in \chCrI3 as 𝐍{\bf N}, we next present microscopic calculations of the spin-orbit torque per applied electric field - a quantity known as the “torkance” - as a function of 𝐍^\hat{{\bf N}}. The procedure for this calculation is well-established [28, 29], and we briefly provide a description here and refer the reader to the Supplementary Information for more technical details. We first obtain the Hamiltonian in a localized atomic orbital basis using a combination of Quantum Espresso [30] and Wannier90 [31]. We then utilize linear response theory to compute the torkance on each magnetic sublattice. The torkance is classified as even or odd according to its behavior under time-reversal. We denote the jthj^{\rm th} component of the torkance on atom A,B\rm A,B in response to an electric field along the ii-direction with τijA,B\tau_{ij}^{\rm A,B}. The even and odd components of the torkance are given by:

(τijA,B)even=2eImn,mnfn(Hki)n,m(𝒯jA,B)m,n(EmEn)2+η2,\left(\tau^{\rm A,B}_{ij}\right)^{\rm even}=2e~{}\text{Im}\sum_{\begin{subarray}{c}n,m\neq n\end{subarray}}f_{n}\frac{\left(\frac{\partial H}{\partial k_{i}}\right)_{n,m}\left(\mathcal{T}^{\rm A,B}_{j}\right)_{m,n}}{(E_{m}-E_{n})^{2}+\eta^{2}}, (5)
(τijA)odd=en12ηfnEn(Hki)n,n(𝒯jA,B)n,n\left(\tau^{A}_{ij}\right)^{\rm odd}=-e\sum_{\begin{subarray}{c}n\end{subarray}}\frac{1}{2\eta}\frac{\partial f_{n}}{\partial E_{n}}\left(\frac{\partial H}{\partial k_{i}}\right)_{n,n}\left(\mathcal{T}^{\rm A,B}_{j}\right)_{n,n} (6)

The sum in Eqs. 5-6 is over eigenstates |ψn|\psi_{n}\rangle of the 𝐤{\bf k}-dependent Hamiltonian H𝐤H_{\bf k}, where 𝐤{\bf k} is the Bloch wave vector and the eigenstate label nn includes 𝐤{\bf k} and band index. (O)n,m=ψn|O|ψm\left(O\right)_{n,m}=\langle\psi_{n}|O|\psi_{m}\rangle is the matrix element of the operator OO, and fn=(e(Enμ)/kBT+1)1f_{n}=(e^{(E_{n}-\mu)/k_{\rm B}T}+1)^{-1} is the equilibrium Fermi-Dirac distribution function. μ\mu is the Fermi level, η\eta is the broadening parameter, and ee is the electron charge. The atom-resolved torque operator is 𝓣A,B=i2{[𝐒,Δ],PA,B}\boldsymbol{\mathcal{T}}^{\rm A,B}=\frac{\rm i}{2\hbar}\left\{[\mathbf{S},\Delta],P^{\rm A,B}\right\}, where 𝐒\mathbf{S} is the spin operator, Δ\Delta is the spin-dependent exchange-correlation potential, and PA(B)P^{\rm A(B)} is the projection operator onto the orbitals centered on atomic site A (B). To compute the torque as a function of 𝐍^\hat{{\bf N}}, we manually rotate the spins on A and B sublattices.

Refer to caption
Figure 2: (Color online) Angular dependence of the dampinglike (a) and fieldlike (b) torkance on the 𝐍^\hat{\mathbf{N}} direction (θ,ϕ)(\theta,\phi) for one layer of bilayer \chCrI3 under an external electric field along the y^\hat{y} direction at Fermi level μ=50meV\mu=50~{}\rm{meV} above the conduction band minimum. The arrow (color) on the sphere indicates the direction (magnitude) of the torkance at the given 𝐍{\bf N} direction. We use kBT=3meV,η=25meVk_{B}T=3~{}\rm{meV},\eta=25~{}\rm{meV}.

Figure 2 shows the 𝐍^\hat{{\bf N}}-dependence of the torkance with (a) and (b) showing the dampinglike (time-reversal even) and fieldlike (time-reversal odd) torkance, respectively. The fixed points of both dampinglike and fieldlike torkance lie in the LxLzL_{x}-L_{z} plane, away from the Lz=0L_{z}=0 equator. This is an important feature and is a consequence of the lack of mirror symmetry with respect to the yzyz plane. This position of the fixed point ensures that the spin-orbit torque drives 𝐍^\hat{{\bf N}} to a point in the northern or southern hemisphere; after the spin-orbit torque is removed, 𝐍^\hat{{\bf N}} then relaxes to the nearest easy-axis along +𝐳^+\hat{\bf z} or 𝐳^-\hat{\bf z}. Previous studies on systems with similar in-plane mirror symmetry breaking, such as \chWTe2-Py heterostructures [32, 33, 34, 29], have verified that this symmetry breaking results in a spin-orbit torque that drives the magnetic order parameter to a point away from the equator. Exploiting this property has emerged as an approach for deterministically switching perpendicularly magnetized thin films with spin-orbit torque, and we show here that this also enables switching of the perpendicular Néel vector.

We note that the 𝐍^\hat{{\bf N}}-dependence of the torkance is quite complex, deviating substantially from the simple, lowest order form used in the analysis of the previous section. In the Supplementary Information, we provide the full symmetry-allowed expansion of the torkance and quantify the substantial contribution from higher order terms. We additionally find that the fixed points for even and odd torkance are different. These features of the microscopically computed torkance have important consequences for the details of the dynamics of 𝐍^\hat{{\bf N}} under spin-orbit torque, which we show in the next section.

Refer to caption
Figure 3: (Color online) Torkance as a function of chemical potential relative to the conduction band edge. The applied electric field is in 𝐲^\hat{{\bf y}} direction. The 𝐍^\hat{\mathbf{N}} vector is in 𝐳^\hat{{\bf z}} (a) and 𝐱^\hat{{\bf x}} (b). Red and blue lines represent staggered time-reversal even torkance and uniform time-reversal odd torque, respectively. The torque is in 𝐱^\hat{{\bf x}} (𝐳^\hat{{\bf z}}) direction when 𝐍^𝐳^(𝐱^)\hat{\mathbf{N}}\parallel\hat{{\bf z}}(\hat{{\bf x}}). The torkance for Fermi energies in the valence band are substantially smaller and not shown here.

We next consider the torkance versus Fermi level for 𝐍^\hat{{\bf N}} along 𝐳^\hat{\bf z} and 𝐱^\hat{{\bf x}} directions, shown in Figs. 3 (a) and (b), respectively. Both even and odd components are peaked for Fermi energies near the conduction band minimum. For 𝐍^=𝐳^\hat{{\bf N}}=\hat{\bf z}, the even torkance is approximately 1ea0/1~{}ea_{0}/\hbar (a00.0529nma_{0}\approx 0.0529~{}{\rm nm} is the Bohr radius) at 0.1eV0.1~{}\rm{eV} above the conduction band minimum, which is larger than the even torkance in the ferromagnetic Pt/Co bilayer (0.6ea0/\approx 0.6~{}ea_{0}/\hbar) [28]. This large magnitude is due to band crossings in the conduction band from pp-orbitals of the heavy Iodine atoms. For 𝐍^=𝐱^\hat{{\bf N}}=\hat{\bf x}, the even torkance magnitude is around 0.4ea0/0.4~{}ea_{0}/\hbar. The even torkance for this 𝐍^\hat{{\bf N}} configuration is solely a consequence of the in-plane mirror symmetry breaking. This value is notably larger than the corresponding torkance derived from in-plane mirror symmetry breaking in the ferromagnetic 1T’-\chWTe2/Co bilayer (0.1ea0/\approx 0.1~{}ea_{0}/\hbar) [29].

Refer to caption
Figure 4: (Color online) Magnetization dynamics under spin-orbit torque, for applied electric fieldin the 𝐲^\hat{\bf y} direction. (a) and (b) show the Néel and magnetization vector components as a function of time with applied electric field strength 1.2-1.2 V/µm and 3.5-3.5 V/µ, respectively. The initial configuration is Lz=1L_{z}=1. Red lines, black lines, and blue lines represent the dynamics of LxL_{x}, LzL_{z}, and MyM_{y} respectively. (c) and (d) show the final steady state of 𝐍^\hat{\mathbf{N}} as a function of applied field with staring point at the Lz=+1L_{z}=+1 and Lz=1L_{z}=-1, respectively. The spread in the yy coordinate indicates the oscillation amplitude, and the color of the spread represents the oscillation frequency.

Magnetization dynamics — Given the significant deviation of the 𝐍^\hat{{\bf N}}-dependence of the microscopically computed spin-orbit torque from the simple form given in the earlier analysis, it’s worthwhile to compute the spin dynamics with the ab initio spin-orbit torque (Fig. 2) as input into the coupled LLG equations (Eq. Spin-orbit torque switching of Néel order in two-dimensional \chCrI3). 𝐍^\hat{\bf N} is parameterized by spherical coordinates (θ,ϕ)(\theta,\phi), and we use a bilinear interpolation of a dense 80×8080\times 80 mesh of spin-orbit torque obtained from first-principles to obtain the full 𝐍^\hat{\bf N}-dependence.

Figure 4 shows the spin-orbit torque driven dynamics. We find that the spin-orbit torque can either induce switching or induce steady state oscillations of 𝐍^\hat{{\bf N}}. Figure 4(a) shows that for an applied electric field E=2.3V/μmE={\rm-2.3~{}V/\mu m}, the spin-orbit torque switches the Néel order LzL_{z} from the north pole to the southern hemisphere within 100 ps and generates a finite in-plane magnetization MyM_{y}. By separately turning off the fieldlike (odd) or dampinglike (even) contributions to the spin-orbit torque, we find that the switching of LzL_{z} originates from the dampinglike torque, while the fieldlike torque helps to accelerate the switching dynamics and reduce the E-field threshold. Figure 4(b) shows an oscillating steady state for E=3V/μmE={\rm 3~{}V/\mu m}, with a frequency of approximately 80 GHz. We find that both dampinglike and fieldlike torque are required to induce steady state oscillation.

We summarize the final steady states as a function of the applied field EE for two initial magnetization configurations Lz=+1L_{z}=+1 and Lz=1L_{z}=-1 in Fig. 4 (c) and (d), respectively. The switching of the Néel vector occurs at approximately |E|=2V/μm|E|=2~{}{\rm V/\mu m}. This threshold compares well with the estimate provided by Eq. 4. Reaching the larger scale oscillations at large applied EE will rely on the material to sustain large power dissipation, which depends in turn on factors such as the carrier mobility. The flatness of the conduction bands implies a low mobility, as seen experimentally [35], which should enable larger applied electric fields. Fig. 4 (c) and (d) demonstrate hysteretic switching of the Néel vector, and are related by mirror symmetry about the xzxz plane.

Discussion— The experimental detection of the Néel vector reversal is challenging. For bilayer \chCrI3, out-of-plane magnetic-optical Kerr effect (MOKE) imaging has previously been used to discriminate between Lz=+1L_{z}=+1 and Lz=1L_{z}=-1 [16], and transport effects such as nonlinear anisotropic magnetoresistance can also detect 𝐍{\bf N} [11]. We also note that the moderate exchange energy leads to the development of a substantial steady state in-plane magnetization of the driven system, which may be detected experimentally with in-plane MOKE. Aside from the particulars of \chCrI3, we show generally that antiferromagnets in the weak to moderate exchange coupling regime exhibit different behaviors from their more commonly studied large HEH_{E} counterparts. The switching criteria for these antiferromagnets is reduced by a factor of magnetic damping, offering potentially easier routes to electrical manipulation. Continued progress in the field of Van der Waals antiferromagnets should provide further opportunities for unique modes of electrical control of these materials.

F.X. acknowledges support under the Cooperative Research Agreement between the University of Maryland and the National Institute of Standards and Technology Physical Measurement Laboratory, Award 70NANB14H209, through the University of Maryland.

References

  • Manchon et al. [2019] A. Manchon, J. Železnỳ, I. M. Miron, T. Jungwirth, J. Sinova, A. Thiaville, K. Garello, and P. Gambardella, Reviews of Modern Physics 91, 035004 (2019).
  • Liu et al. [2011] L. Liu, T. Moriyama, D. Ralph, and R. Buhrman, Physical review letters 106, 036601 (2011).
  • Miron et al. [2011] I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten, M. V. Costache, S. Auffret, S. Bandiera, B. Rodmacq, A. Schuhl, and P. Gambardella, Nature 476, 189 (2011).
  • foo [a] As described in Ref. [36], time-reversal even (odd) torque coincides with dampinglike (fieldlike) torque only for odd values of \ell in the expansion of torque in vector spherical harmonic basis (which includes the lowest order term). Nevertheless, we use these terms interchangeably for ease of reference.
  • Go et al. [2020] D. Go, F. Freimuth, J.-P. Hanke, F. Xue, O. Gomonay, K.-J. Lee, S. Blügel, P. M. Haney, H.-W. Lee, and Y. Mokrousov, Physical Review Research 2, 033401 (2020).
  • Železný et al. [2014] J. Železný, H. Gao, K. Výborný, J. Zemen, J. Mašek, A. Manchon, J. Wunderlich, J. Sinova, and T. Jungwirth, Phys. Rev. Lett. 113, 157201 (2014).
  • Wadley et al. [2016] P. Wadley, B. Howells, J. Železnỳ, C. Andrews, V. Hills, R. P. Campion, V. Novák, K. Olejník, F. Maccherozzi, S. Dhesi, et al., Science 351, 587 (2016).
  • Jungwirth et al. [2016] T. Jungwirth, X. Marti, P. Wadley, and J. Wunderlich, Nature Nanotechnology 11, 231 (2016).
  • Meinert et al. [2018] M. Meinert, D. Graulich, and T. Matalla-Wagner, Phys. Rev. Applied 9, 064040 (2018).
  • Baltz et al. [2018] V. Baltz, A. Manchon, M. Tsoi, T. Moriyama, T. Ono, and Y. Tserkovnyak, Rev. Mod. Phys. 90, 015005 (2018).
  • Godinho et al. [2018] J. Godinho, H. Reichlová, D. Kriegner, V. Novák, K. Olejník, Z. Kašpar, Z. Šobáň, P. Wadley, R. P. Campion, R. M. Otxoa, P. E. Roy, J. Železný, T. Jungwirth, and J. Wunderlich, Nature Communications 9, 4686 (2018).
  • Huang et al. [2017] B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A. McGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo-Herrero, and X. Xu, Nature 546, 270 (2017).
  • Gong et al. [2017] C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao, W. Bao, C. Wang, Y. Wang, Z. Q. Qiu, R. J. Cava, S. G. Louie, J. Xia, and X. Zhang, Nature 546, 265 (2017).
  • Deng et al. [2018] Y. Deng, Y. Yu, Y. Song, J. Zhang, N. Z. Wang, Z. Sun, Y. Yi, Y. Z. Wu, S. Wu, J. Zhu, J. Wang, X. H. Chen, and Y. Zhang, Nature 563, 94 (2018).
  • Jiang et al. [2018a] S. Jiang, J. Shan, and K. F. Mak, Nature Materials 17, 406 (2018a).
  • Huang et al. [2018] B. Huang, G. Clark, D. R. Klein, D. MacNeill, E. Navarro-Moratalla, K. L. Seyler, N. Wilson, M. A. McGuire, D. H. Cobden, D. Xiao, W. Yao, P. Jarillo-Herrero, and X. Xu, Nature Nanotechnology 13, 544 (2018).
  • Jiang et al. [2018b] S. Jiang, L. Li, Z. Wang, K. F. Mak, and J. Shan, Nature Nanotechnology 13, 549 (2018b).
  • Song et al. [2018] T. Song, X. Cai, M. W.-Y. Tu, X. Zhang, B. Huang, N. P. Wilson, K. L. Seyler, L. Zhu, T. Taniguchi, K. Watanabe, M. A. McGuire, D. H. Cobden, D. Xiao, W. Yao, and X. Xu,  360, 1214 (2018).
  • Klein et al. [2018] D. R. Klein, D. MacNeill, J. L. Lado, D. Soriano, E. Navarro-Moratalla, K. Watanabe, T. Taniguchi, S. Manni, P. Canfield, J. Fernández-Rossier, and P. Jarillo-Herrero,  360, 1218 (2018).
  • Wang et al. [2018] Z. Wang, I. Gutiérrez-Lezama, N. Ubrig, M. Kroner, M. Gibertini, T. Taniguchi, K. Watanabe, A. Imamoğlu, E. Giannini, and A. F. Morpurgo, Nature Communications 9, 2516 (2018).
  • Zhang et al. [2020] X.-X. Zhang, L. Li, D. Weber, J. Goldberger, K. F. Mak, and J. Shan, Nature Materials 19, 838 (2020).
  • Richter et al. [2018] N. Richter, D. Weber, F. Martin, N. Singh, U. Schwingenschlögl, B. V. Lotsch, and M. Kläui, Physical Review Materials 2, 024004 (2018).
  • Stiles and Miltat [2006] M. D. Stiles and J. Miltat, Spin-transfer torque and dynamics, in Spin Dynamics in Confined Magnetic Structures III, edited by B. Hillebrands and A. Thiaville (Springer Berlin Heidelberg, Berlin, Heidelberg, 2006) pp. 225–308.
  • Gomonay and Loktev [2010] H. V. Gomonay and V. M. Loktev, Phys. Rev. B 81, 144427 (2010).
  • Manchon [2017] A. Manchon, Journal of Physics: Condensed Matter 29, 104002 (2017).
  • foo [b] For spin configurations in the 𝐍{\bf N}-subspace (mx/zA=mx/zB,myA=myBm_{x/z}^{\rm A}=-m_{x/z}^{\rm B},~{}m_{y}^{\rm A}=m_{y}^{\rm B}), it’s easy to show that |𝐍|=1|{\bf N}|=1. This spin configuration subspace can also be represented with 𝐦^A{\hat{\bf m}}^{\rm A} alone, with the value of 𝐦^B{\hat{\bf m}}^{\rm B} implied by symmetry.
  • foo [c] To determine the robustness of the projection of the spin configuration to the 𝐍{\bf N} subspace, we solved for the steady state for 20420^{4} possible initial conditions in the 4-dimensional space (θA,ϕA,θB,ϕB)(\theta^{\rm A},\phi^{\rm A},\theta^{\rm B},\phi^{\rm B}), which characterize the orientations of the spins on A and B. We find that the steady states from all initial conditions belong to the 𝐍{\bf N} subspace.
  • Freimuth et al. [2014] F. Freimuth, S. Blügel, and Y. Mokrousov, Phys. Rev. B 90, 174423 (2014).
  • Xue et al. [2020] F. Xue, C. Rohmann, J. Li, V. Amin, and P. Haney, Phys. Rev. B 102, 014401 (2020).
  • et al. [2017] P. G. et al., Journal of Physics: Condensed Matter 29, 465901 (2017).
  • Mostofi et al. [2014] A. A. Mostofi, J. R. Yates, G. Pizzi, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, Computer Physics Communications 185, 2309 (2014).
  • MacNeill et al. [2016] D. MacNeill, G. M. Stiehl, M. H. D. Guimaraes, R. A. Buhrman, J. Park, and D. C. Ralph, Nature Physics 13, 300 (2016).
  • MacNeill et al. [2017] D. MacNeill, G. M. Stiehl, M. H. D. Guimarães, N. D. Reynolds, R. A. Buhrman, and D. C. Ralph, Phys. Rev. B 96, 054450 (2017).
  • Stiehl et al. [2019] G. M. Stiehl, R. Li, V. Gupta, I. E. Baggari, S. Jiang, H. Xie, L. F. Kourkoutis, K. F. Mak, J. Shan, R. A. Buhrman, and D. C. Ralph, Phys. Rev. B 100, 184402 (2019).
  • Patil et al. [2020] R. A. Patil, H.-W. Tu, M.-H. Jen, J.-J. Lin, C.-C. Wu, C.-C. Yang, D. Van Pham, C.-H. Tsai, C.-C. Lai, Y. Liou, et al., Materials Today Physics 12, 100174 (2020).
  • Belashchenko et al. [2020] K. Belashchenko, A. A. Kovalev, and M. van Schilfgaarde, Physical Review B 101, 020407 (2020).