Brought to you by:
Rapid Communication

On the Onsager–Casimir reciprocal relations in a tilted Weyl semimetal

, , , , , and

© 2022 Chinese Physical Society and IOP Publishing Ltd
, , Citation Bingyan Jiang et al 2022 Chinese Phys. B 31 097306 DOI 10.1088/1674-1056/ac754a

1674-1056/31/9/097306

Abstract

The Onsager–Casimir reciprocal relations are a fundamental symmetry of nonequilibrium statistical systems. Here we study an unusual chirality-dependent Hall effect in a tilted Weyl semimetal Co3Sn2S2 with broken time-reversal symmetry. It is confirmed that the reciprocal relations are satisfied. Since two Berry curvature effects, an anomalous velocity and a chiral chemical potential, contribute to the observed Hall effect, the reciprocal relations suggest their intriguing connection.

Export citation and abstract BibTeX RIS

Onsager's reciprocity relations state that in thermodynamic systems out of equilibrium, the linear response coefficient, linking a thermodynamic driving force and the resultant flow, is a symmetric tensor. That is, ${ {\mathcal L} }_{ij}={ {\mathcal L} }_{ji}$. In the presence of time-reversal symmetry breaking fields, it becomes the Onsager–Casimir reciprocal relations, ${ {\mathcal L} }_{ij}({\mathcal B})={ {\mathcal L} }_{ji}(- {\mathcal B})$, where $ {\mathcal B} $ denotes all time-reversal symmetry breaking fields. These relations, rooted in the principle of microscopic reversibility, are a fundamental symmetry of nonequilibrium statistical systems. When multiple coupled pairs of flows and thermodynamic forces are involved, the relations can offer insights into the connection between different response coefficients that seem, at first glance, unrelated. For instance, in his seminal papers in 1931, Onsager derived the relations for several coupled irreversible processes and related different coefficients to one another.[1,2] The reciprocal relations have also played an important role on understanding various spin-related transport.[3]

Recently, an unusual chirality-dependent Hall effect and magnetoresistance that are antisymmetric in both magnetic field and magnetization were predicted to occur in tilted topological Weyl semimetals.[410] It was suggested that such an antisymmetry violates the Onsager–Casimir reciprocal relations.[8] Soon, the antisymmetric Hall and magnetoresistance were experimentally observed in a magnetic Weyl semimetal, Co3Sn2S2.[11] The phenomenon is originated from the Berry curvature of energy bands and involves several effects. In particular, the longitudinal linear magnetoresistance results from the Berry curvature correction to the phase space volume, while the Hall effect consists of contributions from the chiral chemical potential and the anomalous velocity. Here, we study the chirality-dependent Hall effect in Co3Sn2S2, with a focus on the validity of the Onsager–Casimir reciprocal relations. Two methods were employed and it was found that the reciprocal relations hold for this Hall effect. The relations connect the effect of the chiral chemical potential to that of the anomalous velocity. Despite only one pair of flow and force at work, our results reveal an intriguing connection between two effects, which deepens our understanding of Berry-curvature-related electrical transport.

Co3Sn2S2 single crystals were grown by the chemical vapor transport method[11] and polished to a bar shape for the transport measurements. We utilized silver paste to fabricate the electrical contacts. Before applying the paste, the crystal was treated with Ar plasma to improve the contact. The device is shown in the inset of Fig. 1(a) and the x axis in the diagram is along the crystal [100] direction. The electrical transport measurements were performed using a low frequency lock-in method. A typical current of 2 mA–3 mA was employed. The angular dependence experiment was carried out using a motorized rotator stage with a resolution of 0.02°. Our samples exhibit transport properties similar to previous reports.[1114] The temperature dependence of the longitudinal resistivity ρyy is shown in Fig. 1(a), where a ferromagnetic phase transition manifests as a clear kink at the Curie temperature Tc ∼ 175 K. Below Tc, the anomalous Hall effect is observed, as shown in Fig. 1(b). The Hall resistivity ρxy jumps at the coercive field Bc and shows a linear dependence on the magnetic field above Bc, from which we obtain an ordinary Hall coefficient RH = 0.51 μΩ⋅cm/T and a corresponding carrier density of 1.2 × 1021 cm−3.

Fig. 1.

Fig. 1. Basic transport properties of Co3Sn2S2 sample #1 and the chirality-dependent Hall effect. (a) Temperature dependence of the longitudinal resistivity ρyy . A clear kink indicates a Curie temperature of 175 K. The inset shows the Hall bar device and the scale bar is 300 μm. (b) Anomalous Hall effect at T = 50 K. The linear dependence of ρxy on the magnetic field above Bc suggests a single carrier type. (c) Hall resistivity as the magnetic field scans along ±70° in the xz plane at T = 50 K. The inset depicts the measurement setup and the magnetic field direction. (d) The chirality-dependent Hall resistivity ${\rho }_{xy}^{\chi }$ obtained by symmetrizing the data in panel (c). ${\rho }_{xy}^{\chi }$ is linear in magnetic field and changes its sign as the magnetization reverses.

Standard image

Co3Sn2S2 is a magnetic Weyl semimetal.[1416] The interplay of band topology and magnetism gives rise to a number of unusual phenomena, attracting intensive research interest.[1220] A very recent experiment has uncovered a chirality-dependent Hall effect that is antisymmetric in both the in-plane magnetic field and the magnetization.[11] It is believed to be associated with the titled Weyl cones. The chirality-dependent Hall effect is observed in our samples. When the magnetic field is rotated from the z axis to the x axis, as depicted in the inset of Fig. 1(c), the in-plane field generates an extra Hall resistivity proportional to Bx . Therefore, the transverse signals at ±θ diverge, shown in Fig. 1(c) (θ = ±70° for instance). The chirality-dependent Hall effect ${\rho }_{xy}^{\chi }$ can be obtained by data antisymmetrizing as ${\rho }_{xy}^{\chi }(\theta)=[{{\rho }^{\prime}_{xy}}(+\theta)-{{\rho }^{\prime}_{xy}}(-\theta)]/2$, where ${{\rho }^{\prime}_{xy}}={\rho }_{xy}-{\rho }_{xy}^{{\rm{AHE}}}$. Considering that the amplitude and orientation of the magnetization remain almost unchanged at 50 K due to a strong magnetocrystalline anisotropy,[17,18] a constant anomalous Hall resistivity ${\rho }_{xy}^{{\rm{AHE}}}$ is subtracted. The obtained ${\rho }_{xy}^{\chi }$ for positive and negative magnetization at θ = ±70° are plotted in Fig. 1(d), which is consistent with the previously reported results of ${\rho }_{xy}^{\chi }\propto {B}_{x}{M}_{z}$.[11]

The chirality-dependent Hall effect satisfies ρxy (B,M) = ρxy (–B,–M), which is consistent with the Onsager–Casimir reciprocal relations.[11] However, to prove that the reciprocal relations hold, one needs to verify ρxy (B,M) = ρyx (–B,–M). One method is to attest the reciprocity theorem

Equation (1)

where the two pairs of indices denote the current leads and voltage leads, respectively.[31,32] Equation (1) can be derived from the Onsager–Casimir reciprocal relations. Figure 2 shows the Hall effect measured under two configurations. In the first configuration, the current is applied through contacts 1 and 2, and the voltage between 3 and 4 is measured to obtain R12,34, as depicted in the bottom inset of Fig. 2(b). In the second configuration, the current and voltage leads are swapped. The current is applied through contacts 3 and 4, and the voltage between 1 and 2 is measured to obtain R34,12, as depicted in the bottom inset of Fig. 2(c). Measurements were performed at several angles. Data are shown in Fig. 2(a). In all cases, R34,12(–B,–M) follows R12,34(B,M) exactly. Consequently, the extracted chirality-dependent Hall effect satisfies the reciprocity theorem, ${R}_{12,34}^{\chi }(B,M)={R}_{34,12}^{\chi }(-B,-M)$, as shown in Figs. 2(b) and 2(c). We further measured the magnetic field angular dependence of two resistances, shown in Figs. 2(a) and 2(b). The magnetic field was rotated in the xz plane. The chirality-dependent Hall resistivity can be extracted as Rχ (θ) = [R(+θ) – R(–θ)]/2 and is presented in Figs. 3(c) and 3(d). The angular dependences of ${R}_{12,34}^{\chi }(B,M)$ and ${R}_{34,12}^{\chi }(-B,-M)$ show almost the same behavior, further supporting the validity of Eq. (1).

Fig. 2.

Fig. 2. Verification of the reciprocal theorem, Eq. (1). (a) R12,34(B) and R34,12(–B) at different angles in the xz plane for T = 50 K. Data are shifted for clarity. (b)–(c) ${R}_{12,34}^{\chi }(B)$ and ${R}_{34,12}^{\chi }(-B)$ at 80° by data antisymmetrizing, respectively. The bottom inset depicts the measurement configuration, while the top inset shows the angle definition.

Standard image
Fig. 3.

Fig. 3. Angular dependence of the Hall resistance. (a)–(b) Angular dependence of R12,34(B) and R34,12(–B) at B = 8.8 T for T = 50 K. (c)–(d) Angular dependence of ${R}_{12,34}^{\chi }(B)$ and ${R}_{34,12}^{\chi }(-B)$ by data antisymmetrizing. The same angular dependence of R and Rχ under two configurations further verifies the reciprocal theorem.

Standard image

However, equation (1) is a global symmetry. In principle, a whole series of four-probe measurements must be carried out and all results must comply with Eq. (1) in order to prove the validity of the Onsager–Casimir relations.[32,33] We now attempt another method, i.e., direct verification of the local symmetry of ρxy (B,M) = ρyx (–B,–M). Note that there is a practical disadvantage in this method, as it is difficult to prepare Hall bars along two directions on the same crystal. In Fig. 2, R12,34(B,M), corresponding to ρxy (B,M), was measured. We had to prepare a Hall bar with another sample to measure ρyx (–B,–M). Since no two samples can have exactly the same resistivity, only a semi-quantitative comparison will be made.

Figure 4(a) displays the measured ρyx (–B) at θ = 80°, 0°, –80° for sample #2. Above Bc, ρyx exhibits a linear B-dependence, from which a carrier density of 8.4 × 1020 cm−3 is obtained. It is slightly smaller than the carrier density of sample #1. The resistivity at zero field is about 125 μΩ⋅cm, compared to 65 μΩ⋅cm of sample #1. With increasing magnetic field or angle θ, ρyx (–B) at ± θ diverge. This behavior qualitatively agrees with that observed in sample #1. The chirality-dependent Hall resistivity ${\rho }_{yx}^{\chi }(-B)$ can be extracted by symmetrizing, depicted in Fig. 4(b). It is antisymmetric in both magnetic field and magnetization, mimicking ρxy (B,M), which confirms the local Onsager–Casimir reciprocal relations. The amplitude of ${\rho }_{yx}^{\chi }$ for sample #2 is nearly twice as much as that for sample #1, which we believe can be attributed to the difference of a factor of 2 in the resistivity.

Fig. 4.

Fig. 4. Verification of the local Onsager–Casimir relations. (a) ρyx (–B) of sample #2 at several angles for T = 50 K. The behavior is similar to sample #1. Data are shifted for clarity. (b) ${\rho }_{yx}^{\chi }(-B)$ at 80° for positive and negative magnetization. The top inset depicts the measurement setup. The bottom inset shows an optical image of sample #2 and the scale bar is 300 μm.

Standard image

To reveal the implication of the reciprocal relations on the chirality-dependent Hall effect, it is necessary to identify various effects of the Berry curvature involved in the phenomenon. Start from the equations of motion for band electrons. The velocity can be expressed as

Equation (2)

where ${{\boldsymbol{v}}}_{{\boldsymbol{k}}}=\displaystyle \frac{1}{\hslash }\displaystyle \frac{\partial \epsilon }{\partial {\boldsymbol{k}}}$ is the band group velocity, ${{\boldsymbol{\varOmega }}}_{{\boldsymbol{k}}}\propto \chi \displaystyle \frac{{\boldsymbol{k}}}{{k}^{3}}$ is the Berry curvature, and $D({\boldsymbol{B}},{{\boldsymbol{\varOmega }}}_{{\boldsymbol{k}}})={[1+\displaystyle \frac{e}{\hslash }({\boldsymbol{B}}\cdot {{\boldsymbol{\varOmega }}}_{{\boldsymbol{k}}})]}^{-1}$ is the modification of the phase space volume.[34,35] e, , E , and B are the elementary charge, the reduced Planck constant, electric field, and magnetic field, respectively. χ = ±1 is the chirality of Weyl cones. $\displaystyle \frac{e}{\hslash }({{\boldsymbol{v}}}_{{\boldsymbol{k}}}\cdot {{\boldsymbol{\varOmega }}}_{{\boldsymbol{k}}}){\boldsymbol{B}}$ is the anomalous velocity related to the Berry curvature. One can calculate the nonequilibrium distribution function by solving the Boltzmann equation under a relaxation time approximation and arrive at[7,8]

Equation (3)

neglecting higher-order corrections. Here, f0 is the equilibrium distribution function, and τ is the mean free time. The first term in the square bracket is the shifted distribution induced by the electric field, while the second term is the chiral chemical potential, resulting from pumping of electrons between Weyl cones of opposite chiralities. The current can be calculated by integration of the velocity

Equation (4)

When calculating the integral, it should be born in mind that Weyl cones appear in pairs. There are three contributions that are linear in magnetic field and electric field, which are non-zero only when Weyl cones are tilted. One is the product of the first term of Eq. (2) and the second term in Eq. (3). It generally contributes to the diagonal current and hence is not of our interest. We write down the other two contributions as

Equation (5)

The first current is the contribution of the anomalous velocity from the shifted distribution, denoted as j A. The second current is the contribution of the velocity from the chiral chemical potential, denoted as j C.

For simplicity, let us consider the particular configurations in our measurements. As depicted in Fig. 5, the tilt of Weyl cones is along the y axis,[11] while the in-plane magnetic field is along the x axis. Various terms in the integrand are sketched to provide an intuitive estimation on the corresponding currents. For ρxy (Since the Hall resistivity is much smaller than the longitudinal resistivity, ${\rho }_{xy}=-{\sigma }_{xy}/({\sigma }_{xx}^{2}+{\sigma }_{xy}^{2})\propto {\sigma }_{xy}$. Thus, σxy calculated by Eq. (5) can be directly compared with ρxy obtained by experiments.), the electric field is in the y direction. j C vanishes because B E = 0. j A is parallel to B (transverse). Although the anomalous velocity $\displaystyle \frac{e}{\hslash }({{\boldsymbol{v}}}_{{\boldsymbol{k}}}\cdot {{\boldsymbol{\varOmega }}}_{{\boldsymbol{k}}}){\boldsymbol{B}}$ are opposite in Weyl cones of opposite chiralities, the shifted distributions are not symmetric in the presence of a tilt, yielding a finite current. For ρyx , the electric field is in the x direction. j A is in the x direction and affects the longitudinal conductivity. The transverse component of current comes from j C. It is worth pointing out that for Weyl cones without a tilt, when considering the effect of the chiral chemical potential, the Fermi level lies horizontally and only shifts in energy. In contrast, for tilted Weyl cones, the Fermi level not only shifts in energy, but also tilts along the y direction, as sketched in Fig. 5(d). When Weyl cones are not tilted, j C vanishes because vk is odd in k within a Weyl cone. A tilt breaks the inversion symmetry of vk within a Weyl cone and the balance between two cones, giving rise to a finite current. The current is along the direction of the tilt (the y axis), as sketched in Fig. 5(f).

Fig. 5.

Fig. 5. Schematic diagram of different contributions to the B-linear transverse current. (a) Weyl cones of ω = ±1 tilt along the y direction. The Fermi level tilts due to the electric field. (b) Anomalous velocity va ∝ ( vk Ωk ) ⋅ B (red arrows) and the distribution function shift ΔfE vk E (contour plot) due to E field, where E y and B x. (c) B-linear current due to va ΔfE (blue arrows on the Fermi surface), the integration of which over Fermi surfaces yields j A. (d) The influence of the chiral chemical potential on the Fermi level. The Fermi level also tilts as vk Ωk varies on the Fermi surface. (e) vk (red arrows) and the influence of the chiral chemical potential on the distribution function Δfca ∝ ( B E )( vk Ωk ) (contour plot) on the Fermi surface, where E ∥ –x and B x. (f) B-linear current due to vk Δfca (blue arrows).

Standard image

Our experiments indicate that the Onsager–Casimir reciprocal relations hold for the chirality-dependent Hall effect, ρxy (B,M) = ρyx (–B,–M). Based on the above analysis on the origin of ρxy and ρyx , it can be therefore concluded that j A(B,M) = j C(–B,–M) in our experiment configurations. Although these two currents are manifests of different effects, our results suggest that they are intricately connected.

Having verified the reciprocal relations under a particular electric and magnetic field configuration and pointed out its implication, we generalize the conclusion to all configurations. Substituting Eqs. (2) and (3) into Eq. (4) and dropping the term for the anomalous Hall effect, one can express the conductivity tensor as

Equation (6)

where vi denotes the i-th component of vk . It was incorrectly concluded that equation (6) violates the reciprocal relations.[8] This misunderstanding stems from the implicit dependence of the anomalous velocity on the magnetization. In a time-reversal-symmetry-broken Weyl semimetal, reversing the magnetization leads to a reversal of the chiralities of Weyl nodes.[36] Thus, the sign of Ωk depends on the sign of the magnetization. Taking the magnetization into consideration, equation (6) indeed satisfies the Onsager–Casimir reciprocal relations, σij (B,M) = σji (–B,–M). When Ω = 0, equation (6) is reduced to the conventional formula. The effect of the Berry curvature is to introduce a correction (the anomalous velocity) to the velocity, besides the change of the phase space volume.

Acknowledgements

We are grateful for discussions with J. Feng, J. R. Shi and H. Z. Lu. Project supported by the National Key Basic Research Program of China (Grant No. 2020YFA0308800) and the National Natural Science Foundation of China (Grant Nos. 11774009 and 12074009), the Natural Science Foundation of Beijing (Grant No. Z200008), and the Youth Innovation Promotion Association of the Chinese Academy of Sciences (Grant No. 2021008).

Please wait… references are loading.