Introduction

The interaction of light with polarization and magnetization in matters could create profound nonlinear magneto-optical (NLMO) phenomena, such as the magnetization-induced second-harmonics generation (MSHG) and the photocurrent generation1,2,3. Among them, NLMO effects related to second-harmonics generation (SHG) have distinctive advantages in both magnetization detection and light modulation. As an all-optical probe characterized by higher-rank tensors, SHG-related NLMO measurement is nondestructive with high spatial and temporal resolution4,5,6,7,8 and shows great promise to characterize structural and magnetic signals in 2D magnets9,10,11. Moreover, the rotation angle of the SHG-related NLMO effect is independent of sample thickness12, which is in clear contrast to the thickness-dependent linear magneto-optical Faraday angle and, therefore, has the potential to be used in miniature devices.

The most significant SHG signal originates from electric dipole transitions8, in which the breaking of the inversion (\({{{\mathcal{P}}}}\)) symmetry is necessary. \({{{\mathcal{P}}}}\) symmetry can be broken by either crystal or magnetic structures, generating the corresponding crystal SHG and MSHG. These two types of SHG are described by an i-type tensor χN, which has even-parity under time-reversal (\({{{\mathcal{T}}}}\)) operation, and a c-type tensor χM, which is \({{{\mathcal{T}}}}\)-odd13,14,15,16,17, respectively. Therefore, in non-centrosymmetric materials with \({{{\mathcal{T}}}}\) symmetry, only χN survives, and in non-centrosymmetric magnetic materials with space-time inversion (\({{{\mathcal{PT}}}}\)) symmetry, only χM survives (Detailed derivations are in Supplementary Note 1A). When both types of SHG coexist, that is, in materials with the simultaneous breaking of \({{{\mathcal{P}}}}\), \({{{\mathcal{T}}}}\) and \({{{\mathcal{PT}}}}\) symmetries, a class of specific NLMO effects arises due to the interference of χN and χM under \({{{\mathcal{T}}}}\)-operation as

$${{{\bf{P}}}}(2\omega )\propto ({\chi }_{{{{\rm{N}}}}}\pm {\chi }_{{{{\rm{M}}}}}){{{\bf{E}}}}(\omega ){{{\bf{E}}}}(\omega )\,,$$
(1)

where E(ω) is the electric field of the incident light and P(2ω) is the nonlinear polarization. The ± represents the influence of \({{{\mathcal{T}}}}\)-operation. We denoted this interference effect of χN and χM as the NLMO effect in the rest of the context. As illustrated in Fig. 1, The interference between the two types of SHG can change the intensity of second harmonic (SH) light through I(2ω) P(2ω)2χN ± χM2. Thus, in magnets with comparable crystal SHG and MSHG, notable NLMO effects can emerge.

Fig. 1: Concept of NLMO effects.
figure 1

Magnets with non-centrosymmetric crystal structures exhibit both the crystal SHG and the MSHG due to the simultaneous breaking of \({{{\mathcal{P}}}}\), \({{{\mathcal{T}}}}\) and \({{{\mathcal{PT}}}}\) symmetries. The crystal SHG is unaffected by the magnetization direction reversal while the MSHG changes the sign. Thus, the two terms can constructively (left panel) or destructively (right panel) interfere with each other for the positive and negative magnetization directions, respectively, which leads to NLMO effects. The red and blue arrows represent magnetic orders related by \({{{\mathcal{T}}}}\) operation.

However, experimentally observed NLMO effects in bulk materials are usually small6,12,18,19,20,21,22,23, which seriously hinders the relevant research and applications of these effects. Additionally, a more profound theoretical understanding to explain and control the magnitude of NLMO effects is also absent. In contrast to bulk materials, recent experiments reveal many unusual SHG responses in two-dimensional (2D) materials, such as the large and tunable i-type SHG in MoS224,25, NbOI2/Cl2 layers26,27 and twisted h-BN28,29, the giant pure c-type SHG in \({{{\mathcal{PT}}}}\)-symmetric ultrathin CrI31, MnPS330 and CrSBr31, and the anomalous SHG of uncertain origin in MnBi2Te4 thin films32 and monolayer (ML) NiI233. These inspiring advances, combined with the great tunability of atomically thin materials, made 2D materials an excellent platform for both fundamental research and the design of miniature devices based on NLMO effects.

In this work, based on the computational method developed in refs. 34,35 and symmetry analysis, we investigated the NLMO effects of representative 2D magnets possessing both χN and χM, including trilayer ABA stacking CrI3, ML Janus Cr2I3Br3, and ML H-VSe2. We analytically derived and numerically calculated the NLMO angle for linearly polarized incident light and the NLMO intensity asymmetry for circularly polarized incident light at different frequencies. Remarkably, we uncovered giant NLMO effects due to the maximal interference between comparable χN and χM in CrI3-based 2D magnets, including a nearly 90° polarization rotation or an on/off switching of certain light helicity of SH light upon magnetization reversal and a maximal SHG circular dichroism (SHG-CD) effect within fixed magnetization. We also found that these NLMO effects are extremely sensitive to subtle changes in complex magnetic orders. Moreover, we revealed the influence of interlayer interaction, spin-orbit coupling (SOC), and synergistic effect of stacking and magnetic orders to NLMO effects and proposed strategies to achieve these giant NLMO effects in more 2D materials. Our work not only uncovered a series of giant NLMO effects and corresponding candidate materials but also provides strategies to manipulate these effects in other material systems, and further shed light on the application of NLMO effects in subtle magnetic orders detection and ultra-thin optical devices.

Results and discussion

Structures and SHG of representative 2D magnets

NLMO effects exist in magnetic materials without \({{{\mathcal{PT}}}}\) symmetry and therefore, we considered two typical ferromagnetic (FM) semiconductors in the monolayer limit, CrI3 and H-VSe236,37,38,39,40,41,42. ML VSe2 is non-centrosymmetric with magnetic symmetry \(\bar{6}{m}^{{\prime} }{2}^{{\prime} }\) when the magnetization is along the \(\hat{z}\) axis. ML CrI3 is centrosymmetric with magnetic symmetry \(\bar{3}{m}^{{\prime} }\)43 and AB stacking bilayer CrI3 has \({{{\mathcal{PT}}}}\) symmetry as shown in Supplementary Fig. 2; therefore, the ML does not have SHG, and the bilayer only has the MSHG χM. In the following, we used the strategy of multi-layer stacking and element replacement to break \({{{\mathcal{P}}}}\) and \({{{\mathcal{PT}}}}\) to enable NLMO effects in CrI3-related materials. The simplest examples are ABA stacking trilayer CrI3 with antiferromagnetic (AFM) interlayer coupling (denoted by ABA-AFM CrI3) and ML Janus Cr2I3Br3. Figure 2a, c, e shows the atomic and magnetic structures of ABA-AFM CrI3, ML Cr2I3Br3, and ML H-VSe2 with their magnetic symmetries summarized in Table 1. Their band structures are shown in Supplementary Note 5.

Fig. 2: Atomic structures and SHG of the representative 2D magnets.
figure 2

a, c, e Top and side views of ABA CrI3 with AFM (↑↓↑) magnetic order, ML Cr2I3Br3 and ML H-VSe2 with FM (↑) magnetic order. The red/blue arrows represent the magnetic moment of each atom. b, d, f The real parts and the norm of SHG susceptibility tensor components. The red and blue lines represent the results of AFM (↑↓↑)/FM (↑) orders and their \({{{\mathcal{T}}}}\)-related tAFM (↓↑↓)/tFM (↓) orders, respectively.

Table 1 Magnetic symmetries and symmetry properties of SHG tensor components of representative 2D magnets

Figure 2b, d, f shows the influence of \({{{\mathcal{T}}}}\) operation to different SHG components χabc(2ω; ω, ω) of the above mentioned materials. In the rest of the article, we used the shorthand notation χabc to represent χabc(2ω; ω, ω), where a, b, c are Cartesian directions. The two magnetic orders related by \({{{\mathcal{T}}}}\) symmetry are denoted as AFM/FM and tAFM/tFM, respectively. For SHG susceptibilities of trilayer CrI3, we scissored the band gap to 1.5 eV, and our results show good agreement with the previous work43 (Supplementary Fig. 3). Due to the C3z symmetry in all three representative materials, each material has only two independent in-plane SHG components, that is χ111 = −χ122 = −χ212 and χ222 = −χ211 = −χ112, where the subscripts 1 and 2 denote the Cartesian direction \(\hat{x}\) and \(\hat{y}\). Due to the presence of \({m}^{{\prime} }\) or \({2}^{{\prime} }\) symmetry in all three materials, one of the SHG components is \({{{\mathcal{T}}}}\)-odd while the other is \({{{\mathcal{T}}}}\)-even, as shown in the upper and middle panels of Fig. 2b, d, f. Generally, the even and odd quantities can coexist in the same tensor component, and in this case, \({\chi }_{{{{\rm{N}}}}}=(\chi +{{{\mathcal{T}}}}\chi )/2\), and \({\chi }_{{{{\rm{M}}}}}=(\chi -{{{\mathcal{T}}}}\chi )/2\), where \({{{\mathcal{T}}}}\chi\) is the SHG susceptibility of the time-reversal pair. The detailed expressions of χN and χM, as well as their symmetry requirements, are summarized in Supplementary Note 1. The parity of each component under \({{{\mathcal{T}}}}\) is also summarized in Table 1.

The bottom panel of Fig. 2b, d, f clearly shows that the relative values of χN and χM are distinctive in the three materials. Although χM exceeds χN in ABA-AFM CrI3 at a wide frequency range and the opposite is observed in ML Cr2I3Br3, there are still several intersections of χN and χM. In contrast, χN is much larger than χM in ML H-VSe2 without any intersections.

Giant NLMO effects of representative 2D magnets

The coexistence of χN and χM can induce a variety of NLMO effects. In the following, we calculated the SHG responses under the linearly and the circularly polarized light (LPL and CPL), with an emphasis on the role of different ratios of χN and χM to NLMO effects. Considering a normal incidence geometry where the material is in the xy-plane and the incident light propagates in the \(-\hat{z}\) direction, the SHG polarization P(2ω) of 2D materials has been given in ref. 43. As the intensity of the emitted SH light I(2ω) is proportional to P(2ω)2 and is commonly measured, we calculated P(2ω)2 to represent I(2ω).

For LPL characterized by \({{{\bf{E}}}}=E(\cos \theta ,\sin \theta )\), where θ denotes the angle between light polarization and the x-axis of the sample, we investigated the polarization-resolved SHG as a function of θ. The experiment measures the parallel () and perpendicular () components of the SHG signal with respect to the direction of the incoming light polarization while the sample rotates with the angle θ. In the three representative materials, the SHG intensity at the parallel and perpendicular polarization directions can be written as43

$$\begin{array}{ll}{I}_{| | }(\theta )&\propto | {\chi }^{111}\cos 3\theta -{\chi }^{222}\sin 3\theta {| }^{2}{E}^{4}\,,\\ {I}_{\perp }(\theta )&\propto | {\chi }^{111}\sin 3\theta +{\chi }^{222}\cos 3\theta {| }^{2}{E}^{4}.\end{array}$$
(2)

Both the parallel and perpendicular components exhibit sixfold sunflower-like patterns and differ only by a ±π/6 angle, as the black and red lines shown in Fig. 3b. In addition, the angle θm corresponding to the maximum (minimum) of I also corresponds to the minimum (maximum) of I.

Fig. 3: NLMO effects in representative 2D magnets.
figure 3

a The NLMO angle and c NLMO intensity asymmetry as a function of frequency. The vertical dashed lines in the curves indicate the NLMO rotation is ill-defined at the corresponding frequencies, and the polarization-resolved SHG close to those frequencies in ABA-CrI3 and ML Cr2I3Br3 are shown in Supplementary Figs. 5a and 7a. b Change of polarization-resolved SHG under \({{{\mathcal{T}}}}\) operation at a LPL characteristic frequency \({\omega }_{{{{\rm{LPL}}}}}^{{\prime} }=1.71\) eV of ML Cr2I3Br3. The yellow line marks one of the LPL incident angles where the maximal polarization direction change of SHG happens under \({{{\mathcal{T}}}}\) operation. d CPL SHG intensity of FM and tFM states of ML Cr2I3Br3. The upper and lower panels are for incident helicity σ+ and σ. The difference between the red and blue curves in each panel reflects the NLMO intensity asymmetry. For a particular magnetic order, the difference between the upper and lower panels reflects the magnitude of SHG-CD. The dashed yellow line indicates one of the CPL characteristic frequencies ωCPL = 1.57 eV of ML Cr2I3Br3. e Multi-degree controlling of NLMO in ML Cr2I3Br3. For linearly-polarized incident light at polarization angle θ = π/12 + nπ/6 and frequency \({\omega }_{{{{\rm{LPL}}}}}^{{\prime} }\), ML Cr2I3Br3 can act as a magneto-optical polarization switcher. For circularly polarized incident light at frequency ωCPL, ML Cr2I3Br3 can act as a magneto-optical switch for circularly polarized SH light or an optical filter for circularly polarized SH light with a particular helicity.

NLMO effects under LPL are reflected by comparing the polarization-resolved SHG patterns before and after \({{{\mathcal{T}}}}\) operation, which results from the sign reversal of χM. Under the symmetry of the three representative materials, this corresponds to a rotation of the polarization-resolved SHG pattern (Supplementary Figs. 5b and 7b), denoted as NLMO angle \({\theta }_{{{{\mathcal{T}}}}}\) (see details in Supplementary Note 2A), with the expression

$${\theta }_{{{{\mathcal{T}}}}}=2{\theta }_{{{{\rm{m}}}}}=-\frac{{\tan }^{-1}\left(\tan 2\delta \cos \Delta \varphi \right)}{3}+\frac{n\pi }{3}\,,$$
(3)

where n is an integer originated from the rotation symmetry of the pattern. \(\tan \delta =\frac{| {\chi }^{111}| }{| {\chi }^{222}| }\in (0,+\infty )\) is the magnitude ratio of χ111 and χ222, and \(\Delta \varphi =-i\left(\ln \frac{{\chi }^{111}}{| {\chi }^{111}| }-\ln \frac{{\chi }^{222}}{| {\chi }^{222}| }\right)\in (-\pi ,+\pi ]\) is the phase difference between χ111 and χ222. The range of δ is (0, π/2). Eq. (3) does not hold when χ111 = χ222 and Δφ = ± π/2. In this situation, the polarization-resolved SHG is isotropic according to Eq. (2), therefore, it does not show any change under \({{{\mathcal{T}}}}\) operation.

The maximum NLMO angle can be achieved is \({\theta }_{{{{\mathcal{T}}}}}=\pi /6+n\pi /3\) with the condition

$$\left\{\begin{array}{l}| {\chi }^{111}| =| {\chi }^{222}| \,,\quad \\ \Delta \varphi \,\ne\, \pm \pi /2\,,\quad \end{array}\right.$$
(4)

and we denoted the corresponding frequency as the LPL characteristic frequency ωLPL. At ωLPL, the patterns of I and I are swapped exactly, as shown in Fig. 3b, which results in a remarkable polarization change of the SH light. If the magnitude of I and I differs greatly, the polarization of the SH light can rotate nearly π/2 under \({{{\mathcal{T}}}}\). The condition to achieve the π/2 polarization rotation at θm = π/12 + nπ/6 is

$$\left\{\begin{array}{l}| {\chi }^{111}| =| {\chi }^{222}| \,,\quad \\ \Delta \varphi =0\,{{{\rm{o}}}}r\,\pi \,,\quad \end{array}\right.$$
(5)

and the corresponding frequency is denoted as \({\omega }_{{{{\rm{LPL}}}}}^{{\prime} }\), which is a subset of ωLPL. This condition is also graphically shown in the complex plane in Supplementary Fig. 1.

We numerically traced the ω-dependence of the NLMO angle \({\theta }_{{{{\mathcal{T}}}}}\) in the three representative materials, as shown in Fig. 3a. As highlighted by the dashed horizontal line, \({\theta }_{{{{\mathcal{T}}}}}\) close to π/6 + nπ/3 corresponds to the largest NLMO angle. Remarkably, \({\theta }_{{{{\mathcal{T}}}}}\) is giant in ABA-AFM CrI3 and ML Cr2I3Br3 at several ωLPL frequencies. In contrast, \({\theta }_{{{{\mathcal{T}}}}}\) is always tiny in ML H-VSe2. This distinction is consistent with their different ratio of χN and χM susceptibilities observed in Fig. 2b, d, f. Figure 3b shows the dramatic change of SHG light from perpendicular to parallel polarization direction for incident light at θm = π/12 (indicated by the yellow line) under \({{{\mathcal{T}}}}\) in this atomically thin Cr2I3Br3 at \({\omega }_{{{{\rm{LPL}}}}}^{{\prime} }=1.71\) eV.

Next, we investigated the polarization and intensity of SH light under the illumination of the CPL. For the CPL characterized by E = E(1, ± i), where ±represents the left/right-handed (σ±) CPL, the polarization of the SHG signal is

$${{{\bf{P}}}}(2\omega )=2{E}^{2}({\chi }^{111}\,\mp\, i{\chi }^{222})(1,\,\mp\, i)\,$$
(6)

according to Table 1. The SH light is also CPL with helicity opposite to the incident light. Furthermore, the intensity of the SH light depends on the helicity of the incident light, which is called the SHG-CD effect44,45,46. The SHG-CD intensity asymmetry can be described by

$${\eta }_{{{{\rm{h}}}}}^{{{{\bf{M}}}}}=\frac{{I}^{+}({{{\bf{M}}}})-{I}^{-}({{{\bf{M}}}})}{{I}^{+}({{{\bf{M}}}})+{I}^{-}({{{\bf{M}}}})}\,,$$
(7)

where I+(M) is the SHG intensity generated by the magnetic order M when the helicity of the incident light is σ+.

\({{{\mathcal{T}}}}\) operation can also change the intensity of the SH light with a certain helicity. We can define the NLMO intensity asymmetry at certain incident light helicity σ as

$${\eta }_{{{{\mathcal{T}}}}}^{\sigma }=\frac{{I}^{\sigma }(+{{{\bf{M}}}})-{I}^{\sigma }(-{{{\bf{M}}}})}{{I}^{\sigma }(+{{{\bf{M}}}})+{I}^{\sigma }(-{{{\bf{M}}}})}\,,$$
(8)

where the ±M states are \({{{\mathcal{T}}}}\)-related, and Iσ(+M) is the SHG intensity generated by the +M magnetic order. The range of \({\eta }_{{{{\mathcal{T}}}}}^{\sigma }\) is [−1,1] and \({\eta }_{{{{\mathcal{T}}}}}^{\sigma }=\pm 1\) represents the situation in which an incident CPL with helicity σ can only generate SH light in the ±M magnetic states and is completely blocked in the opposite magnetic state. According to Eq. (6), the reversal of helicity and the reversal of magnetization are equivalent, and therefore the change of intensity under \({{{\mathcal{T}}}}\) operation can also be reproduced by changing the helicity of CPL at a fixed magnetization, i.e., \({\eta }_{{{{\mathcal{T}}}}}^{+}={\eta }_{{{{\rm{h}}}}}^{+{{{\bf{M}}}}}\).

The expression of the NLMO intensity asymmetry in our case is (derivation is in Supplementary Note 2B)

$${\eta }_{{{{\mathcal{T}}}}}^{\pm }=\mp \sin 2\delta \sin \Delta \varphi \,.$$
(9)

As long as Δφ ≠ nπ, the NLMO intensity asymmetry is always nonzero, and the intensity of the SH light is always magnetization-dependent. In general, the maximum \(| {\eta }_{{{{\mathcal{T}}}}}^{\pm }| =1\) requires

$$\left\{\begin{array}{l}| {\chi }^{111}| =| {\chi }^{222}| \,,\quad \\ \Delta \varphi =\pm \pi /2\,,\quad \end{array}\right.$$
(10)

and we denoted the corresponding frequency as the CPL characteristic frequency ωCPL. This condition is also graphically shown in the complex plane in Supplementary Fig. 1. It is worth noting that the ωCPL is just where the \({\theta }_{{{{\mathcal{T}}}}}\) for LPL is ill-defined.

We numerically traced the ω-dependence of \({\eta }_{{{{\mathcal{T}}}}}^{+}(2\omega )\) in the three representative materials in Fig. 3c. Again, two CrI3-based materials can achieve large \({\eta }_{{{{\mathcal{T}}}}}^{+}\) including ±1 at several frequencies while \(| {\eta }_{{{{\mathcal{T}}}}}^{+}|\) is always below 0.4 in ML H-VSe2. This contrast is another feature stemming from their distinct ratio of χN and χM. We take ML Cr2I3Br3 as an example to illustrate the intensity change of SH light under \({{{\mathcal{T}}}}\) operation for σ± CPL in Fig. 3d. (The CPL SHG intensity of ABA-AFM CrI3 and ML H-VSe2 are shown in Supplementary Figs. 4 and 8). The vertical dash line in Fig. 3d highlights one of the frequencies ωCPL at which the intensity of SHG changes between a considerable value and zero under \({{{\mathcal{T}}}}\) operation. The upper (σ+) and lower (σ) panels of Fig. 3d look exactly the same, except the line colors are swapped, which reflects the equivalency of helicity and magnetization reversal. Other than that, the difference in the red (blue) line between the upper and lower panels in Fig. 3d also indicates that this material exhibits a 100% SHG-CD effect at ωCPL.

The giant NLMO effects discovered in CrI3-based materials enabled them to be used as magnetic-field-controlled atomically thin optical devices, as illustrated in Fig. 3e. Utilizing incident light at frequencies \({\omega }_{{{{\rm{LPL}}}}}^{{\prime} }\), they can serve as optical polarization switchers for LPL as they can rotate the polarization of SH light by nearly π/2. At the incident frequency ωCPL, they are ideal magneto-optical switches for CPL due to their large \({\eta }_{{{{\mathcal{T}}}}}\). In addition, even without an external magnetic field, due to their large SHG-CD effects, they can serve as optical filters for CPL with a particular helicity.

It is worth noting that the above-discussed NLMO and SHG-CD effects are analyzed based on the nonlinear polarization P(2ω) generated in materials, and the intensity of radiation is estimated by I(2ω) P(2ω)2, which is not specific for the reflected or the transmitted SH light. The exact value of reflected and transmitted SH light can be obtained by solving Maxwell equations with corresponding boundary conditions47. Although the SH light in reflection or transmission may have different intensities, the above-discussed NLMO and SHG-CD effects persist.

The materials we investigated so far have several particular symmetry features. In general, NLMO effects are highly sensitive to changes in magnetic orders and can be used as a powerful tool to distinguish subtle magnetic states. As an example, we considered a different magnetic order ↑↑↓ for ABA stacking trilayer CrI3, which we denoted as the MIX state and its \({{{\mathcal{T}}}}\)-pair state as tMIX. ABA-MIX CrI3 and ABA-AFM CrI3 have the same net magnetic moment and tensor components as Table 1 shows, and their linear optical responses are similar, as shown in Supplementary Note 4, which makes them difficult to be distinguished experimentally. However, in ABA-MIX CrI3, each SHG tensor component contains both χN and χM and does not have a definite parity under \({{{\mathcal{T}}}}\), as shown in Fig. 4a. As a result, for LPL, the polarization-resolved SHG shows both rotation and magnitude change under \({{{\mathcal{T}}}}\) operation as shown at the bottom two plots in Fig. 4c. For CPL, the equivalence between helicity and magnetization reversal, i.e., \({\eta }_{{{h}}}^{+\mathbf{M}}={\eta }_{{{{\mathcal{T}}}}}^{+}\), is broken, as reflected by the remarkable difference between the upper and lower panels in Fig. 4b. Therefore, under the illumination of LPL (CPL) at ωLPL (ωCPL) for ABA-AFM CrI3, we can observe notable differences between ABA-AFM CrI3 and ABA-MIX CrI3, as shown in Fig. 4c.

Fig. 4: SHG susceptibility, CPL SHG intensity and polarization-resolved SHG for ABA-MIX CrI3.
figure 4

a Real parts of SHG susceptibility tensor components of ABA-MIX CrI3. Red and blue lines represent results for MIX(↑↑↓) and tMIX(↓↓↑) magnetic orders, respectively. b SHG intensity of MIX and tMIX magnetic orders of ABA-MIX CrI3 under CPL illumination. The upper and lower panels are for incident helicity σ+ and σ. The difference between the red and blue curves in each panel reflects the \({\eta }_{{{{\mathcal{T}}}}}\). For a particular magnetic order, the difference between the upper and lower panels reflects the magnitude of SHG-CD. c Polarization-resolved SHG of AFM, tAFM, MIX, and tMIX magnetic orders at one of the LPL frequency \({\omega }_{{{{\rm{LPL}}}}}^{{\prime} }\)= 1.22 eV of ABA-AFM CrI3. The red/blue arrow represents the magnetic moment in each CrI3 layer.

In addition to the detection of full ordered magnetic ground states, the NLMO effects can also distinguish the influence from spin fluctuation at finite temperatures22 and spin canting caused by an external field, as shown in Supplementary Fig. 6.

Controllable NLMO effects of representative 2D magnets

The giant NLMO effects manifest many intriguing applications as we have learned from the above results, a necessary condition to achieve those effects is to have comparable χN and χM. However, in real materials, χN and χM may differ greatly, which results in negligible NLMO effects. In the following, taking ABA-AFM CrI3 as a model system, we demonstrated several strategies to enhance NLMO effects, in which χ111 is the \({{{\mathcal{T}}}}\)-odd term χM and χ222 is the \({{{\mathcal{T}}}}\)-even term χN as summarized in Table 1.

The strength of interlayer interaction has a great impact on properties of multilayer magnets37,48,49; thus, we first investigated the influence of interlayer distance d in trilayer CrI3, where d0 is the interlayer distance of ABA-AFM CrI3 at equilibrium. As shown in Fig. 5a, both χM and χN change with d, but χN is more sensitive to the variation of d. Therefore, a more comparable χN and χM can be realized by exerting hydrostatic pressure on multilayer CrI3 along the z-axis in experiments, as shown in Fig. 5d.

Fig. 5: Design principles of NLMO effects.
figure 5

The influence of interlayer spacing (a), SOC (b), and layer number (c) to χ111 which is the \({{{\mathcal{T}}}}\)-odd term χM and χ222 which is the \({{{\mathcal{T}}}}\)-even term χN in ABA-AFM CrI3 multilayers. The SHG tensor components of ABA-AFM CrI3 with interlayer distance 0.9d0 (d), ABA-AFM CrBr3 (e) and 7L H-VSe2 (f). The \({{{\mathcal{T}}}}\)-odd component χ222 in even-layer AFM CrI3 is not shown.

Next, we investigated the influence of SOC on NLMO effects by artificially tuning the magnitude of SOC strength λ. As Fig. 5b shows, χM is sensitive to the change of λ while χN changes slightly. As χM is much larger than χN in ABA-AFM CrI3, we replaced I by Br to reduce SOC and found ABA-AFM CrBr3 possesses very comparable χN and χM in a broad frequency range, as shown in Fig. 5e.

Furthermore, we found stacking and magnetic orders can have synergistic effects on engineering the relative and absolute values of χN and χM. As shown in Fig. 5c, in multilayer AB stacking AFM CrI3, χN is forbidden in even-layer structures due to the presence of \({{{\mathcal{PT}}}}\) symmetry and remains almost the same in odd-layer structures, while χM increases with the layer number. The above observations can be simply understood as χN from each layer alternates the sign with similar magnitude, while χM from each layer has the same sign due to the \({{{\mathcal{PT}}}}\) symmetry between the A and the B layers24. Inspired by this, as χM is much smaller than χN in ML H-VSe2, we calculated the SHG susceptibility of 7-layer (7L) AA\({}^{{\prime} }\)-AFM H-VSe2, which also has \({{{\mathcal{PT}}}}\) symmetry between A and A\({}^{{\prime} }\) layers. As expected, the peak value of χM in 7L structure shown in Fig. 5f is almost 7 times that in ML H-VSe2 (Fig. 2f) while its χN remains almost the same. χN and χM are comparable in 7L-AA\({}^{{\prime} }\) stacking H-VSe2, indicating it can host giant NLMO effects. Similarly, by tuning the stacking sequence and magnetic order, we can achieve \({{{\mathcal{P}}}}\), \({{{\mathcal{T}}}}\) or neither symmetry between neighboring layers and can realize an arbitrary control of the relative and absolute value of χN and χM (Supplementary Fig. 9).

To summarize, the simultaneous breaking of \({{{\mathcal{P}}}}\), \({{{\mathcal{T}}}}\) and their joint symmetry allows NLMO effects with the electric-dipole origin. In several representative 2D magnets with similar symmetry, we investigated and demonstrated NLMO angle \({\theta }_{{{{\mathcal{T}}}}}\) under LPL, NLMO intensity asymmetry \({\eta }_{{{{\mathcal{T}}}}}\) and SHG-CD ηh under CPL. In particular, we discovered promising candidates with giant NLMO effects, including a near 90° polarization rotation and an on/off switching of certain helicity of SH light upon magnetization reversal, as well as ±1 SHG-CD within a certain magnetic configuration. These giant NLMO effects in candidate 2D magnets can be used not only to design atomically thin NLMO devices such as optical polarization switchers, switches, and filters but also to detect subtle magnetic orders in multilayer magnets such as ABA CrI3. We further derived that the comparable magnitude of χN and χM is indispensable to achieve giant NLMO effects. Lastly and most importantly, we found the interlayer distance, the magnitude of SOC, and the synergistic effect of stacking and magnetic orders could be used to control the relative and absolute magnitude of χN and χM, which provides general design principles to achieve giant NLMO effects in 2D magnets. Our finding not only reveals several intriguing NLMO phenomena but also paves the way to achieve subtle magnetization detection, as well as giant and controllable NLMO effects in ultra-thin magneto-optical devices.

Methods

First-principles calculations

First-principles calculations were performed by Vienna Abinitio Simulation Package50 with SOC included. The exchange-correlation function was parameterized in the Perdew–Burke–Ernzerhof form51, and the projector augmented-wave potential52 was used. For the 3d orbitals in magnetic ions Cr and V, the Hubbard U of 337 and 1.16 eV53 were used. For layered materials, we used DFT-D3 form van der Waals correction without damping54. The cut-off energy of plane waves was set to 450 and 500 eV for CrI3-based materials and H-VSe2-based materials, respectively. The convergence criterion of force was set to 10 and 1 meV Å−1 for CrI3-based materials and H-VSe2-based materials, respectively. Total energy is converged within 10−6 eV. k-point samplings of 13 × 13 × 1 were used for CrI3-based materials and 15 × 15 × 1 for H-VSe2-based materials. Vacuum thickness of about 20 Å was used in the calculations of 2D materials.

SHG calculations

After getting the converged electronic structures, we generated the maximally localized Wannier functions using Wannier9055 to build the tight-binding Hamiltonian and calculate optical responses34. The \({{{\mathcal{T}}}}\) symmetry breaking is carefully treated in the calculations of SHG for magnets35,56. We obtained 56 maximally localized orbitals for each layer of CrI3 or Cr2I3Br3 and 22 for each layer of H-VSe2. The broadening factor of the Dirac delta function is taken to be 0.05 eV. We found that 50 × 50 × 1 and 150 × 150 × 1 k-mesh samplings are enough for the converged SHG susceptibilities for CrI3-based materials and H-VSe2-based materials. The band gap of trilayer CrI3 has been scissored to 1.5 eV in the calculation of SHG susceptibility43.