Abstract
Whenever the elastic energy of a solid depends on magnetic field, there is a magnetostrictive response. Field-linear magnetostriction implies piezomagnetism and vice versa. Here, we show that Mn3Sn, a non-collinear antiferromanget with Weyl nodes, hosts a large and almost perfectly linear magnetostriction even at room temperature. The longitudinal and transverse magnetostriction, with opposite signs and similar amplitude are restricted to the kagome planes and the out-of-plane response is negligibly small. By studying four different samples with different Mn:Sn ratios, we find a clear correlation between the linear magnetostriction, the spontaneous magnetization and the concentration of Sn vacancies. The recently reported piezomagnetic data fits in our picture. We show that linear magnetostriction and piezomagnetism are both driven by the field-induced in-plane twist of spins. A quantitative account of the experimental data requires the distortion of the spin texture by Sn vacancies. We find that the field-induced domain nucleation within the hysteresis loop corresponds to a phase transition. Within the hysteresis loop, a concomitant mesoscopic modulation of local strain and spin twist angles, leading to twisto-magnetic stripes, arises as a result of the competition between elastic and magnetic energies.
Similar content being viewed by others
Introduction
Following the discovery of a large room-temperature anomalous Hall effect (AHE) in Mn3Sn1 and in Mn3Ge2, this family of noncolinear antiferromagnets3 became a subject of experimental4,5,6,7 and theoretical8,9,10,11,12,13,14 attention. The AHE signal is present at room temperature and can be inverted with a low magnetic field. Therefore, the discovery has a technological potential in the field of antiferromagnetic spintronics15,16,17,18,19. Very recently, Ikhlas et al.20 reported that Mn3Sn displays a large piezomagnetic effect at room temperature and the sign of the AHE can be modified by a sufficiently large uniaxial stress. By contrasting the response of magnetization and AHE, they demonstrated that the ultimate origin of the AHE is the residual Berry curvature. This confirmed a theoretical prediction21, which argued that AHE in a non-collinear antiferromagnet does not require spontaneous magnetization.
Piezomagnetism, the generation of magnetic moment upon the application of strain, σ, is intimately linked to magnetostriction, ϵ, the field-induced contraction or elongation of the lattice22. Thermodynamics imposes a strict equivalence between the piezomagnetic response (\(\frac{\partial M}{\partial \sigma }\)) and the field slope of magnetostriction (\(\frac{\partial \epsilon }{\partial B}\)). Both quantities represent the second derivative of the free energy with respect to magnetic field and stress (∂2F/∂σ∂B) and their unavoidable equality is a specific case of Maxwell relations23.
Magnetostriction, first discovered in ferromagnetic iron by Joule24, is technologically attractive25, because it can be used to conceive devices converting magnetic to mechanical energy and vice versa. Fundamentally, it arises when the magnetic field plays a role in setting the magnitude of the elastic energy. This can happen in metals26 where it is delectably large (that is >10−6 T−1)whenever the electronic density of states depends on magnetic field. Known cases include heavy-fermion systems near field-induced instabilities27 and dilute metals subject to quantizing magnetic fields28. In ferromagnets, the dominant magnetostrictive response is due to field-induced displacement and rotation of magnetic domains and the boundary between them. When all domains align and magnetization saturates, magnetostriction collapses29,30 (See also our data on cobalt below). In antiferromagnets, magnetostriction, drawing significant interest recently20,31,32, is usually small and quadratic in magnetic field33,34,35. Save for a limited group of 35 antiferromagnetic classes (out of a total of 122 magnetic point groups), symmetry considerations forbid linear magnetostriction33,36,37.
Here, we report on a detailed study of magnetostriction in Mn3Sn, which belongs with a \(mm{\prime} m{\prime}\) magnetic point group12, not of the 35 antiferromagnetic classes allowing linear magnetostriction33,36,37). The linear magnetostriction is allowed by symmetry, because of the residual ferromagnetism.
By studying four different samples, we find that the slope of linear magnetostriction, Λ, and the spontaneous magnetization, M0 both depend on the concentration of Sn vacancies. The dependence of Λ on Sn concentration allows us to reconcile the amplitude of the piezomagnetic coefficient resported by Ikhlas et al.20, with our data.
We argue that the field-induced in-plane distortion of the spin texture and the competition between magnetic and elastic38 energies leads to the emergence of linear magnetostriction. In this anti-chiral39 spin texture, a magnetic field oriented along the kagome planes steadily twists the spins, as seen by torque magnetometry40. The field-induced rotation of spins generates an almost perfectly linear magnetostriction from 0.02 T to 9 T, which is to be contrasted with the magnetostriction caused by re-orientation of spins close to a ferromagnetic-antiferromagnetic phase boundary41.
We also find that magnetostriction displays a discontinuous jump at the threshold magnetic field for domain nucleation. This is a signature of a second-order phase transition. The amplitude of the jump implies that, during the passage between single-domain regimes of opposite polarities, there are structural stripes with a mesoscopic width. We dub these stripes twisto-magnetic, since they refer to concomitant modulations of local strain and local spin orientation. Pinning down their structural details emerges as a subject for future experimental and theoretical studies.
Experimental
The single crystals used in this study were grown by the vertical Bridgman method, as detailed in the ref. 42. Mn3Sn crystallizes into hexagonal DO19 structure. It is known that the single crystals are not stoichiometric and the Mn:Sn ratio is larger than 343. As we will argue below, this is because a small fraction of Sn sites remain unoccupied and therefore the chemical formula becomes Mn3Sn1−δ44, with a number of physical properties depending on δ. The four samples, which were extensively studied this had different concentrations of Sn vacancies: Mn3Sn0.871 (#1), Mn3Sn0.891 (#2), Mn3Sn0.827 (C1) and Mn3Sn0.875 (C2). The samples dimensions were 0.82 × 1.02 × 1.42 mm3 (#1), 1.01 × 1.70 × 2.05 mm3 (#2), 0.93 × 1.80 × 4 mm3 (C1) and 4 × 0.5 × 1.8 mm3 (C2). Samples #1 and #2 were from one batch, while samples C1 and C2 were from another batch.
The dilatometer used in this study is sketched in Fig. 2a, b. It consists of two metallic plates forming a capacitor. One is kept still, while the other, to which the sample is clamped can move when the length of the sample changes. As the field is swept, this change gives rise to the variation of the capacitance between the two plates45,46, which can be measured by a capacitance bridge. Please refer to the Methods section for more measurement details.
Results
Figure 1a, b shows the crystal and magnetic structure of Mn3Sn. Each spin located on a Mn site is oriented 120 degrees off its adjacent spin on the same triangle in a kagome layer. This is an antichiral structure where the rigid clockwise rotation of the triangle with three spins on its vertices would lead to an anti-clockwise rotation of spins on each site39. The magnetic moment on each Mn site is ~3 Bohr magnetons, μB. However, the net magnetic moment is only ~0.002–0.003 μB per Mn atom, indicating almost total compensation between moments oriented 120-degree off each other. This inverse triangular spin structure emerges below TN = 420 K1,3 and is destroyed below a temperature, which can be tuned by changing the Mn:Sn ratio.1,3,4.
Zero-field thermal expansion
Figure 1c shows the thermal expansion data along the three crystal-axis orientations in sample #1. The thermal expansion coefficient is positive along the three orientations. It is larger in the kagome planes and there is a modest anisotropy which decreases with warming. Our thermal expansion data is agreement with the results of a previous study restricted to temperatures above 200 K47. In Fig. 1c, for each orientation, the length at 2 K is taken as the reference. The inset of the Fig. 1c shows the thermal expansion coefficients at low temperatures. The destruction of the inverse triangular spin state around 50 K, indexed by a dashed line, does not cause any detectable anomaly.
Longitudinal and transverse magnetostriction
Figure 2c compares the longitudinal magnetostriction along the three crystalline orientations in low fields up to 0.3 T at 200 K. In our convention, \({\epsilon }_{j}^{i}\) refers to the magnetostriction measured along j when the magnetic field is along i. Longitudinal magnetostriction is contractile, and almost identical along the x and y axes. On the other hand, when the field is along the z-axis, there is no detectable signal. Figure 2d shows the magnetization in the same sample. There is a finite spontaneous magnetization for the two in-plane configurations, but not for the out-of-plane one. Magnetization has a finite slope along the three orientations, as reported previously1. The absence of out-of-plane longitudinal magnetostriction implies that uni-axial stress along z-axis does not affect out-of-plane magnetization. In contrast, the in-plane longitudinal magnetostriction is finite, implying that in-plane stress would affect the magnetization. Moreover, since the magnetostriction is linear in magnetic field, stress should shift magnetization without changing its slope as a function of magnetic field. This is indeed what was found by the recent study of piezomagnetism20.
As seen in Fig. 2c, d, magnetostriction and magnetization behave concomitantly. When the magnetic field exceeds a critical threshold (B0 = 0.02 T marked by two dashed vertical lines for the two sweeping orientations), magnetostriction peaks and magnetization starts to increase. B0 is the field at which domains (with a polarity set by magnetic field and inverse to the prevailing one) nucleate. When the magnetic field exceeds a much larger amplitude (≈0.2 T), the system becomes single domain again42. We will discuss what happens when B = B0, in more detail later below.
Figure 3a–d shows the longitudinal and the transverse magnetostriction for both x and y orientations at various temperatures. For both orientations, longitudinal magnetostriction is contractile and transverse magnetostriction is expansive. As seen from the figures, it is linear in all four configurations. The field slope of \(\frac{\partial \epsilon }{\partial B}\) is extracted from this data and is plotted as a function of temperature in Fig. 3e, f. For all four configurations, \(\frac{\partial \epsilon }{\partial B}\) are flat above 150 K and begin to increase with decreasing temperature with the approach of the 50 K phase transition (for a brief glimpse to the magnetostriction below 50K, see Supplementary Fig. 11 in the supplementary materials48).
We conclude that even at room temperature, for both x and y directions, there is a large longitudinal \(\frac{\partial \epsilon }{\partial B}\) of comparable magnitude (≈− 1.5 × 10−5 T−1) slightly larger than a transverse signal of opposite sign (≈+1.1 × 10−5 T−1). We also measured transverse magnetostriction for other configurations and found a tiny response within the margin of our experimental resolution (see Supplementary Fig. 3 in the supplementary materials48).
Linear magnetostriction up to 9 T
Figure 4a illustrates the evolution of magnetostriction up to 9 T. One can express the magnetostriction as the sum of a linear and a quadratic term : ϵ = alB + aqB2. Table 1 lists the best fits to these coefficients for two different samples.
The combination of a dominant linear (al) and a sub-dominant quadratic (aq) term up to ≈ 9 T is reminiscent of the case of UO2 in its magnetically ordered state32. Interestingly, the amplitude of the linear magnetostriction is roughly similar in the two cases ( ≈ 10−5 T−1). However, here it is a room-temperature phenomenon, while in UO2 it emerges below a Néel temperature of 30 K32. No saturation is visible in our data up to 9 T in contrast with what is observed in ferromagnets (For an example, see the case of Co in the supplementary materials48). Figure 4b shows a sketch of mechanism. We will discuss the microscopic origin of this robust field-linear magnetostriction in the next section.
Discontinuity at the threshold of domain nucleation
Figure 5a, b is a zoom on the low-field behavior near B0. One can see a discontinuity in longitudinal magnetostriction at B0 ≈0.02 T. This is the field at which magnetic domains nucleate. When the magnetic field is swept down to zero, the magnetostriction linearly decreases to a null value. It changes sign with the inversion of the magnetic field. But the magnetization and the anomalous Hall effect do not change sign when the magnetic field inverts. They continue to have a sign opposite to the magnetic field. At B0, both the magnetization and the AHE begin to evolve along the orientation imposed by the magnetic field42. At this field, longitudinal magnetostriction shows a discontinuous jump (Fig. 5a, b), in contrast with the weaker anomaly seen in transverse magnetostriction. The observed jump represents a discontinuity in the second derivative of the free energy and implies a second-order phase transition.
The magnetostriction hysteresis loop in Mn3Sn differs qualitatively from other cases of linear magnetostriction such as MnF2 or UO2, where the response is set by domain rotation (see Supplementary Fig. 6 in the supplementary materials48). We will come back to the implications of this observation and also the Fig. 5e, f in the next section.
Multiplying the amplitude of the jump in magnetostriction by the Young Modulus (E = 123 GPa38) allows to quantify the change in elastic energy, shown in the Fig. 5c. Similarly, we can identify the energy cost of domain nucleation with the product of B0 and magnetization, shown in the Fig. 5d. The two energies differ by more than three orders of magnitudes and both tend to vanish at TN = 420 K as expected.
Sample dependence
Having found that the amplitude of linear magnetostriction differs from one sample to the other, we undertook an extensive examination of magnetization and magnetostriction of additional samples. Figure 6a portrays the magnetization of several samples at 300K. In addition to four different samples used in our study, the figure includes the sample studied by Ikhlas et al. in their piezomagnetic study20. One can see that the amplitude of the jump in magnetization, M0 varies from sample to sample.
Figure 6b plots the amplitude of this spontaneous magnetization as a function of δ, the concentration of Sn vacancies. In addition to our samples, we include three samples studied by Kurosawa et al.49. There is a clear correlation between the amplitude of M0 and δ. This implies that in-plane ferromagnetism is, at least partially, caused by the presence of Sn vacancies.
Figure 6c compares the magnetostriction of four different samples. There is a visible difference in slope of \({\epsilon }_{x}^{x}\). In other words, linear magnetostriction has not the same amplitude. Plotting \(\Lambda=\frac{\partial }{\partial B}{\epsilon }_{x}^{x}\) as a function of M0 (Fig. 6d) of the four crystals shows a clear quasi-linear correlation between the two. The figure includes another data point, Λ reported by the piezomagnetic study20. Thus, residual ferromagnetism is not only indispensable for the existence of linear magnetostriction, but also sets its amplitude. As seen in Supplementary Fig. 1248, the presence of Sn vacancies leads to a detectable change in the lattice parameter along the a-axis. Along the c-axis, the variation is undetectably small. Note that the difference in lattice parameter among the samples is at least an order of magnitude larger than the change induced by a 10 T magnetic field in each sample.
Discussion
Origin of the linear magnetostriction
Both piezomagnetism and magnetostriction arise because any deviation of the elementary units of Mn3Sn from a perfect equilateral geometry disrupts the cancellation of the net moment of the three spins therein. Microscopically, strain transforms the equilateral triangle to an isosceles one. The cancellation is spoiled because the two most nearby spins have a stronger antiferromagnetic coupling, and hence twist slightly toward a more antiparallel alignment. Consequently, a distortion of the triangle induces a net moment (piezomagnetism), and a net moment caused by field-induced twisting of the spins39,40 induces a distortion (magnetostriction).
In the ideal system, where all triangles are identical, one can understand this effect very simply from both a microscopic calculation and a symmetry point of view. In this case, we can consider the spins on a single triangle, and define three linear combinations:
Here, M is a real vector representing the total spin on a triangle, and Φ is a complex vector which defines the antiferromagnetic order parameter. For spins of fixed length S, we have ∣M∣2 + 2Φ* ⋅ Φ = 9S2. When the three spins are at 120 degrees to one another, M = 0 and \({{{{\mathbf{\Phi}}}}}^{ * } \cdot {{{\mathbf{\Phi}}}}=\frac{9}{2}{S}^{2}\). Note that when all spins rotate by 180 degrees, Φ changes sign, but keeps its absolute value.
Nearest-neighbor exchange in these variables becomes
One can see readily this favors the antiferromagnetic state. Now we include the simplest exchange-striction effect, which corresponds to the modification of the exchange J on a bond proportional to the change in its length. This is modeled by the term
where S3 = S0, ε is the strain tensor, and \({\hat{{{{\boldsymbol{f}}}}}}_{n}\) is the unit vector along the bond connecting spin n and n + 1, and g is proportional to \(-\frac{r}{J}\frac{dJ}{dr}\). Using Eq. (1), one obtains, dropping a trivial volume expansion contribution,
where \({\varepsilon }_{+}=\frac{1}{2}({\epsilon }_{xx}-{\epsilon }_{yy}+2i{\epsilon }_{xy})\).
In a pure spiral state (either chiral or antichiral), the Φ ⋅ Φ = 0 term vanishes. The remaining term in Eq. (4) describes an in-plane anisotropic stress arising in the presence of magnetization and antiferromagnetic order. The stress (and consequent strain) is induced without spin-orbit coupling, because the antiferromagnet order itself couples spin and orbital symmetries. Specifically, a 120 degree ordered state breaks both spin rotation and C3 spatial rotation symmetries, but preserves the combination of the two – in the anti-chiral case, the two rotations are made in the opposite sense. One might think of the antiferromagnet itself as a state in which spin-orbit coupling is generated spontaneously!
To understand how this effect arises without SOC, we need to think about the presence of both magnetization and AF order. In Mn3Sn, the AF order is long-range and well-established at room temperature. When we think about any single triangle of spins, the AF order parameter Φ is therefore imposed upon it by other triangles coupled to it, and related to the order far away. Supposing this global AF order is somehow pinned (typically by an applied field), then the spin orientations on the triangle in question are almost fixed. If a small uniform magnetization is imposed, the spins rotate slightly, and then in turn leads to a difference of bond energies, which depends on the relative orientation of the AF order and the magnetization. That relative orientation determines the axes for the anisotropic stress, which will induce a strain that seeks to strengthen the exchange on the bond with the most anti-aligned pair of spins.
While this effect does not require SOC, other effects can occur when SOC is present. To allow for all possibilities, we consider the general form of linear couplings of an anisotropic in-plane strain ε+ to the magnetic order parameters. Here we specialize to the anti-chiral state, whose order parameter is the complex scalar Φ = Φx − iΦy. We assume Φz = Φx + iΦy = 0, so there is pure anti-chiral order. We also express the in-plane magnetization as a second complex scalar M = Mx − iMy. Then the most general free energy density linear in ε+ is
where γ1,2,3 are phenomenological coupling constants. By comparing to Eq. (4), we can see that \({\gamma }_{1}=\frac{gJ}{3{v}_{{{{\rm{u.c.}}}}}}\), where vu.c. is the volume per unit cell used to convert to an energy density. The other two couplings γ2, γ3 are zero in the absence of SOC, but are generally present and can be obtained for example from single ion anisotropy (SIA)20.
Using the equations of elasticity, we have ε = C−1σ, where σ = − ∂fe−s/∂ε is the strain induced by the spin-lattice coupling. For the case of polar anisotropy and a three-fold rotation axis along z, we find
where C66 is the shear modulus in Voigt notation. For fields which are larger than the hysteresis field, but still small enough that the magnetization is small, one can approximate M ≈ Ms + χH, and \(\Phi \sim | \Phi | {\hat{H}}^{ * }\), where H = Hx − iHy and \(\hat{H}=H/| H|\). We then have
The second line describes linear in field magnetostriction:
We see that it arises from two contributions. The “intrinsic” term obtained above gives γ1, which is independent of the spontaneous magnetization Ms. By contrast, the “extrinsic” term γ2 gives a linear contribution proportional to ∣Ms∣.
Consistency with the experiment would be achieved if γ2 were the dominant contribution. This is surprising and in disagreement with naïve expectations, and γ2 ≪ γ1 for the uniform system in the weak anisotropy limit. It seems that non-stoichiometry has the effect of dramatically enhancing γ2.
Neumann’s principle
From a symmetry point of view, linear magnetostriction may occur in certain antiferromagnetic point groups33, but occurs generically in ferromagnets. While Mn3Sn is dominated by antiferromagnetism, it is in the symmetry sense a ferromagnet: it belongs to the ferromagnetic point group \(mm{\prime} m{\prime}\) (see the supplementary materials48). Consequently, both a spontaneous zero field magnetization and magnetostriction are generically expected – neither are prohibited by Neumann’s principle. The ferromagnetism in Mn3Sn is weak, however, because of the weakness of SIA. This enables the linear magnetostriction to extend over a very wide field range, simply because of the extended linearity of the magnetization due in turn to the large antiferromagnetic exchange J. Nevertheless, it is to be expected that residual ferromagnetism accompanies the presence of Λ.
Piezomagnetism and linear magnetostriction
Our magnetostriction data should match the piezomagnetic data20. Maxwell relations imply a thermodynamic identity between the results of two probes29:
The left hand side is the slope of field-linear magnetostriction and the right hand side is the piezomagnetic coefficient. Both represent the components of the same third-rank tensor Λijk22. Measuring the stress dependence of magnetization, Ikhlas et al.20 reported that \(\frac{dM}{d\sigma }\simeq -0.055\,\)Gauss MPa−1 for the x-axis orientation (and slightly smaller in another sample along the y-direction). Now, magnetization in S.I. units is expressed in A/m and their results corresponds to \(\frac{dM}{d\sigma }\simeq -4.4\times 1{0}^{-6}\) T−1 in S. I. units. This is significantly smaller that what we found in four different samples for both longitudinal and transverse magnetostriction. The absolute amplitude of \(\frac{\partial \epsilon }{\partial B}\) was always larger than 1 × 10−5 T−1).
However, as seen in Fig. 6d the discrepancy fades away by considering the fact that the spontaneous magnetization in the sample subject to the piezomagnetic study was smaller than our samples, indicating that its stoichiometry was different from ours. The calculated piezomagnetic coefficient of Mn3Sn1−δ is in the same order of magnitude as the experimental results(see Supplementary Fig. 10 in the supplementary materials48). However, any quantitative account of Λ should explain the linear proportionality between piezomagnetic/magnetostrictive response and the spontaneous magnetization remains unexplained. This requires to consider the role of Sn vacancies.
Unoccupied Sn vacancy sites
Mn3Sn crystal have a deficit of Sn atoms43,44. A number of Sn sites are vacant. As seen in Fig. 6b, the amplitude of spontaneous magnetization increases with increasing density of these Sn vacancies. Moreover, the spontaneous magnetization calculated by first-principle theory can be perfectly matched with our experimental data. In contrast, if vacant Sn sites were occupied by Mn atoms (giving rise to a chemical formula of Mn3+xSn1−x) the calculated spontaneous magnetization is two orders of magnitude larger than the experimental data (see Supplementary Fig. 9 in the supplementary materials48). Note that this calculation was performed for T= 0 K (See More Computational Details in the supplementary materials48)). The agreement between theory and experiment leads to conclude that the chemical formula is Mn3Sn1−δ.
One expects two important effects introduced by these Sn vacancy sites. First, they disrupt the local antiferromagnetic spin cancellation, and induce spin twisting and a local magnetic moment (Fig. 7). Second, they break the high symmetry of the nearby Mn clusters, which may change the orbital character of these moments, enhancing spin-orbit effects and SIA. These two effect provide a basis for understanding the observed trends. Spontaneous magnetization increases with non-stoichiometry because each Sn vacancy brings an uncompensated moment. The coupling γ2, which was shown in Eq. (7) to drive to Λ ∝ Ms, can only arise from enhanced SIA. These observations are strongly suggestive. A quantitative account of Λ in this picture, requires a Hamiltonian for the non-stoichiometric compound, a substantial theoretical effort, beyond the scope of the present work.
Twisto-magnetic stripes
In the previous section, we saw that at a small field of B ≃ 0.02 T, longitudinal magnetostriction shows a discontinuous drop but the transverse magnetization shows just a peak. At this field, the spontaneous magnetization begins to change sign, signaling that the sample is no more single domain42,50. The accumulated magnetic energy (i.e. the product of magnetization and magnetic field) at B0 becomes sufficiently large to pay the energy cost of building domain walls.
As seen in Fig. 5a, b, at B = B0, longitudinal magnetostriction falls to zero. The total volume of the sample is therefore occupied by magnetic domains whose mutual mechanical deformation along the orientation of the magnetic field cancel each other. Since the total magnetization is not zero at B0, the multi-domain regime, restricted to a field window between 0.02 T to 0.2 T, there should be a concomitant modulation of lattice parameter and spin orientation (Fig. 5e) at Mn sites. When two neighboring atomic triangles have slightly different lattice parameters, atomic-bond angles and spin twisted angles, the force exerted by the strain gradient will be compensated by the magnetic force due to the angle between twisted spins (see Fig. 5f). Given that the magnetic energy density (that is the product of magnetization and magnetic field) is only a few pascals42), which is orders of magnitude smaller than the elastic energy density, the differential angle of neighboring spins should be orders of magnitude larger than the differential angle of neighboring atomic bonds. A magnetic field of 13T can distort the spin structure by around two degrees40. This intricate mesoscopic structure emerges as a subject of futures investigation by experimental probes in both real space and in momentum space.
Comparison with other magnets
Table 2 lists solids in which linear magnetostriction and piezomagnetism have been observed22. We saw that in the case of Mn3Sn, the departure from stoichiometry plays a crucial rule in setting the amplitude of Λ. Notice the discrepancy in the reported amplitudes of linear magnetostriction and piezomagnetism in other cases, which is yet to be sorted out. Measuring the same sample with both methods could settle the issue.
In summary, we find that Mn3Sn hosts a large in-plane magnetostriction dominantly linear in magnetic field. We argue that it arises from the field-induced twist of spins and its amplitude can be accounted for given the magnetic and the elastic energy scales. Thermodynamic consistency between the magnetostriction and the piezomagnetic data is achieved only by considering the fact that both, as well as spontaneous magnetization depend on the concentration of Sn vacancies. The sudden vanishing of longitudinal magnetostriction at the onset of domain nucleation implies the existence of twisto-magnetic stripes with a concomitant modulation of strain and local magnetization in a narrow field window.
Methods
Sample
The Mn3Sn single crystals used in this work were grown using the vertical Bridgman method42. The millimeter-size samples were cut from the as-grown samples by a wire saw.
Crystal orientations were determined by a single crystal XRD diffractometer (XtaLAB mini II of Rigaku). The variation of temperature during magnetostriction measurements remained within 0.05%.
Measurement
A capacitive dilatometer with a resolution of ΔL = 0.02 Å made by Innovative Measurement Technology was used to measure magnetostriction in an Oxford Instruments Teslatron PT. We used an Andeen-Hagerling 2550A capacitance bridge for our measurements. By installing the dilatometer perpendicular to the magnetic field, we also measured magnetostriction in the transverse configuration. The reliability of our set-up was checked by measuring magnetostriction in a cobalt single crystal (see the supplementary materials48).
Calculation
The theoretical computation was realized by first principles calculations using Vienna ab initio simulation package (VASP)51 with the projector augmented wave (PAW) pseudopotential52 and the Perdew-Burke-Ernzerhof (PBE) type of the generalized gradient approximation (GGA) of the exchange correlation potential53. See More Computational Details in the supplementary materials48.
Data availability
The data that support the findings of this study are available from the corresponding author upon reasonable request. Source data are provided with this paper.
References
Nakatsuji, S., Kiyohara, N. & Higo, T. Large anomalous Hall effect in a non-collinear antiferromagnet at room temperature. Nature 527, 212 (2015).
Nayak, A. K. et al. Large anomalous Hall effect driven by a nonvanishing Berry curvature in the noncolinear antiferromagnet Mn3Ge. Sci. Adv. 2, e1501870 (2016).
Tomiyoshi, S. & Yamaguchi, Y. Magnetic structure and weak ferromagnetism of Mn3Sn studied by polarized neutron diffraction. J. Phys. Soc. Jpn. 51, 2478 (1982).
Li, X. et al. Anomalous Nernst and Righi-Leduc effects in Mn3Sn: Berry curvature and entropy flow. Phys. Rev. Lett. 119, 056601 (2017).
Ikhlas, M. et al. Large anomalous Nernst effect at room temperature in a chiral antiferromagnet. Nat. Phys. 13, 1085 (2017).
Xu, L. et al. Finite-temperature violation of the anomalous transverse Wiedemann-Franz law. Sci. Adv. 6, eaaz3522 (2020).
Li, X., Zhu, Z. & Behnia, K. A monomaterial Nernst thermopile with hermaphroditic legs. Adv. Mater. 33, 2100751 (2021).
Yang, H. et al. Topological Weyl semimetals in the chiral antiferromagnetic materials Mn3Ge and Mn3Sn. N. J. Phys. 19, 015008 (2017).
Kuroda, K. et al. Evidence for magnetic Weyl fermions in a correlated metal. Nat. Mater. 16, 1090 (2017).
Liu, J. & Balents, L. Anomalous Hall effect and topological defects in antiferromagnetic Weyl semimetals: Mn3Sn/Ge. Phys. Rev. Lett. 119, 087202 (2017).
Miwa, S. et al. Giant effective damping of octupole oscillation in an antiferromagnetic Weyl semimetal. Small Sci. 1, 2000062 (2021).
Suzuki, M.-T., Koretsune, T., Ochi, M. & Arita, R. Cluster multipole theory for anomalous Hall effect in antiferromagnets. Phys. Rev. B 95, 094406 (2017).
Zelenskiy, A., Monchesky, T. L., Plumer, M. L. & Southern, B. W. Anisotropic magnetic interactions in hexagonal AB-stacked kagome lattice structures: Application to Mn3X (X = Ge, Sn, Ga) compounds. Phys. Rev. B 103, 144401 (2021).
Park, P. et al. Magnetic excitations in non-collinear antiferromagnetic Weyl semimetal Mn3Sn. npj Quantum Mater. 3, 63 (2018).
Tsai, H. et al. Electrical manipulation of a topological antiferromagnetic state. Nature 580, 608 (2020).
Baltz, V. et al. Antiferromagnetic spintronics. Rev. Mod. Phys. 90, 015005 (2018).
Šmejkal, L., Mokrousov, Y., Yan, B. & MacDonald, A. H. Topological antiferromagnetic spintronics. Nat. Phys. 14, 242 (2018).
Kimata, M. et al. Magnetic and magnetic inverse spin Hall effects in a non-collinear antiferromagnet. Nature 565, 627 (2019).
Higo, T. et al. Perpendicular full switching of chiral antiferromagnetic order by current. Nature 607, 474 (2022).
Ikhlas, M. et al. Piezomagnetic switching of the anomalous Hall effect in an antiferromagnet at room temperature. Nat. Phys. 18, 1086 (2022).
Chen, H., Niu, Q. & MacDonald, A. H. Anomalous Hall effect arising from noncollinear antiferromagnetism. Phys. Rev. Lett. 112, 017205 (2014).
Borovik-romanov, A. S. Piezomagnetism, linear magnetostriction and magnetooptic effect. Ferroelectrics 162, 153 (1994).
Pippard, A. B. Elements of classical thermodynamics: for advanced students of physics, https://books.google.fr/books?id=GVhaSQ7eBQoC (Cambridge University Press, 1964).
Joule, J. P. On the effects of magnetism upon the dimensions of iron and steel bars. Sturgeon’s. Ann. Electricity 8, 219 (1842).
Liu, J., Jiang, C. & Xu, H. Giant magnetostrictive materials. Sci. China Technol. Sci. 55, 1319 (2012).
Chandrasekhar, B. S. & Fawcett, E. Magnetostriction in metals. Adv. Phys. 20, 775 (1971).
Puech, L. et al. High-field magnetostriction in the pseudo-metamagnetic heavy-fermion system CeRu2Si2. J. Low. Temp. Phys. 70, 237 (1988).
Küchler, R. et al. Thermodynamic evidence for valley-dependent density of states in bulk bismuth. Nat. Mater. 13, 461 (2014).
Lee, E. W. Magnetostriction and magnetomechanical effects. Rep. Prog. Phys. 18, 184 (1955).
Chopra, H. D. & Wuttig, M. Non-joulian magnetostriction. Nature 521, 340 (2015).
Ma, H.-Y. et al. Multifunctional antiferromagnetic materials with giant piezomagnetism and noncollinear spin current. Nat. Commun. 12, 2846 (2021).
Jaime, M. et al. Piezomagnetism and magnetoelastic memory in uranium dioxide. Nat. Commun. 8, 99 (2017).
Birss, R. R. & Anderson, J. C. Linear magnetostriction in antiferromagnetics. Proc. Phys. Soc. 81, 1139 (1963).
Alberts, L. & Lee, E. W. Magnetostriction in antiferromagnetic nickel oxide. Proc. Phys. Soc. 78, 728 (1961).
Lines, M. E. Elastic properties of magnetic materials. Phys. Rep. 55, 133 (1979).
Tavger, B. A. Symmetry of piezomagnetic crystals. Sov. Phys. Crystallogr 3, 341 (1958).
Moral, A. D. & Brooks, M. S. S. Field dependence of forced magnetostriction in cubic ferro-, ferri and antiferromagnets. J. Phys. C: Solid State Phys. 7, 2540 (1974).
Theuss, F. et al. Strong magnetoelastic coupling in Mn3X (X = Ge, Sn). Phys. Rev. B 105, 174430 (2022).
Balents, L. Antichiral order and spin reorientation transitions of triangle-based antiferromagnets. Phys. Rev. B 106, L020403 (2022).
Li, X. et al. Free energy of twisting spins in Mn3Sn. Phys. Rev. B 106, L020402 (2022).
Song, Y. et al. The critical role of spin rotation in the giant magnetostriction of La(Fe,Al)13. Sci. China Mater. 64, 1238 (2021).
Li, X. et al. Momentum-space and real-space Berry curvatures in Mn3Sn. SciPost Phys. 5, 063 (2018).
Ikhlas, M., Tomita, T., and Nakatsuji, S. Proceedings of the International Conference on Strongly Correlated Electron Systems (SCES2019), JPS Conference Proceedings, Vol. 30, https://doi.org/10.7566/JPSCP.30.011177 (Journal of the Physical Society of Japan, 2020)
Krén, E., Paitz, J., Zimmer, G. & Zsoldos, E. Study of the magnetic phase transformation in the Mn3Sn phase. Phys. B+C. 80, 226 (1975).
Küchler, R., Bauer, T., Brando, M. & Steglich, F. A compact and miniaturized high resolution capacitance dilatometer for measuring thermal expansion and magnetostriction. Rev. Sci. Instrum. 83, 095102 (2012).
Kübler, J. & Felser, C. Non-collinear antiferromagnets and the anomalous Hall effect. EPL (Europhys. Lett.) 108, 67001 (2014).
Song, Y. et al. Opposite thermal expansion in isostructural noncollinear antiferromagnetic compounds of Mn3A (A = Ge and Sn). Chem. Mater. 30, 6236 (2018).
See Supplementary Materials for more details (2023)
Kurosawa, S. et al. Chiral-anomaly-driven magnetotransport in the correlated weyl magnet Mn3Sn, https://arxiv.org/abs/2204.00882 (2022).
Li, X. et al. Chiral domain walls of Mn3Sn and their memory. Nat. Commun. 10, 3021 (2019).
Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 47, 558 (1993).
Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 17953 (1994).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996).
Borovik-Romanov, A. S., Grimmer, H., Kenzelmann, M. Magnetic properties, in International Tables for Crystallography, https://doi.org/10.1107/97809553602060000904 (Wiley Online Library, 2013).
Baruchel, J. et al. Piezomagnetism and domains in MnF2. J. Phys. Colloq. 49, C8 (1988).
Sandlund, L. et al. Magnetostriction, elastic moduli, and coupling factors of composite Terfenol-D. J. Appl. Phys. 75, 5656 (1994).
Acknowledgements
This work was supported by The National Key Research and Development Program of China (Grant No. 2023YFA1609600, 2022YFA1403500), the National Science Foundation of China (Grant No. 12304065, 12004123, 51861135104 and 11574097) and the Fundamental Research Funds for the Central Universities (Grant No. 2019kfyXMBZ071). K. B. was supported by the Agence Nationale de la Recherche (ANR-19-CE30-0014-04). S. J. acknowledges a PhD scholarship by the China Scholarship Council (CSC). L.B. was supported by the NSF CMMT program under Grant No. DMR-2116515.
Author information
Authors and Affiliations
Contributions
X.L., Z.Z., and K.B. conceived and designed the study. Q.M. helped by P.N, L.X., J. W., S.J., and H.Z. performed the measurements. J.D. and J.Z. performed the abinitio calculations. L.B. carried out a theoretical analysis of the linear magnetostriction. Q.M., X.L., Z.Z., L.B., and K.B. analyzed the data. Q.M., X.L., Z.Z., L.B., and K.B. wrote the manuscript with assistance from all the authors.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Tomoya Higo, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Meng, Q., Dong, J., Nie, P. et al. Magnetostriction, piezomagnetism and domain nucleation in a Kagome antiferromagnet. Nat Commun 15, 6921 (2024). https://doi.org/10.1038/s41467-024-51268-y
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-024-51268-y
Comments
By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.