Introduction

Chirality is a common property in nature, describing an asymmetry of an object upon its mirroring. In condensed matter physics, it has been widely known that mirror symmetry breaking in materials with a chiral crystal structure leads to functional properties such as natural optical activity and piezoelectric effects. When time-reversal symmetry is further broken by magnetism, unconventional physical phenomena can emerge. One of the most prominent examples is nonreciprocal propagation and dichroism of elementary particles or excitations, such as photons1,2, electrons3,4, magnons5, and phonons6. In this case, time-reversal symmetry is broken either by spontaneous macroscopic magnetization (i.e., ferromagnetic order) or by an external magnetic field.

Meanwhile, time-reversal symmetry breaking can occur even in antiferromagnets without macroscopic magnetization. Recent theoretical studies7,8 have shown that such time-reversal-broken antiferromagnetic states are fully described by an ordering of particular magnetic multipoles and magnetic toroidal multipoles in terms of their symmetry details. Among them, magnetic quadrupole, which is a second-rank magnetic multipole, is one of the most fundamental multipoles that break not only time-reversal but also space-inversion symmetries. The magnetic quadrupoles have attracted growing interest in the study of a linear magnetoelectric (ME) effect9,10,11,12,13, nonreciprocal optical responses14, and magnetotransport phenomena15 as well as their possible relevance to superconductivity16,17. By combining the magnetic quadrupoles with chirality, one can expect further exotic phenomena distinct from those arising due to the magnetization and chirality. To our knowledge, however, such phenomena have been much less explored so far.

In this paper, based on symmetry considerations, we propose that magnetic quadrupole of a QXY-type (magnetic analogue of the dxy electron orbital, see Fig. 1a) and chirality together give rise to unusual induction of magnetization whose direction is orthogonal to an applied magnetic field and reversed by a switching of chirality (see Fig. 1 for a conceptual illustration of this proposal). We substantiate this proposal through measurements of ME effects in an insulating chiral-lattice antiferromagnet Pb(TiO)Cu4(PO4)4 exhibiting a ferroic order of the QXY magnetic quadrupoles. Moreover, we demonstrate that this “orthogonal” magnetization allows controlling the sign of QXY with a magnetic field only, which is generally impossible in an achiral counterpart. Furthermore, our cluster mean-field calculations successfully reproduce all the experimental observations. The results are also discussed based on a cluster multipole decomposition method and a phenomenological approach.

Fig. 1: Conceptual illustration of orthogonal magnetization induced by magnetic quadrupole and chirality.
figure 1

a Schematic illustration of the QXY-type magnetic quadrupole. In an applied magnetic field along the X axis (HX), QXY produces the electric polarization along the Y axis through the linear magnetoelectric effect. b Directions of HX-induced PY, the trivial HX-induced magnetization along the X axis (MX), and the toroidal moment along the Z axis defined by T = P × M. The system is achiral because the mirror plane parallel to the YZ plane mYZ is present. c, d A pair of chiral systems C− (c) and C+ (d) obtained by breaking the mirror symmetry of mYZ in the achiral system (b). They are related by mYZ. Combination of TZ and chirality generates the new Z-axis component of the magnetization (MZ) in an applied HX, whose sign is reversed by a chirality switching. This orthogonal magnetization is the subject of this work.

Results

Proposal of orthogonal magnetization by combination of magnetic quadrupole and chirality

Before presenting our proposal of orthogonal magnetization, we summarize transformation properties of the following physical quantities crucial in this work: electric polarization (P), magnetization (M), toroidal moment (T) defined as P × M, and chirality (C). Here, the parameter C = ± 1 represents the sign of chirality (handedness). The vectors P, M, and T are transformed identically under proper rotations, while they are transformed in a different manner under space-inversion, time-reversal, and mirror plane operations. Recently, Hlinka18 pointed out that vector-like and director-like physical quantities can be distinguished by their transformations under operations of space-inversion, time-reversal, and mirror plane parallel to the vector (director). Using a notation {±,±,±}, where the first, second, and third signs represent the invariance {+} or variance {−} under these operators, respectively, we can write transformations {−,+,+} for P; {+,−,−} for M; {−,−,+} for T; and {−,+,−} for C. One can find that the product of T and C is transformed identically with M. Hence, T combined with C can linearly induce M along parallel/antiparallel to T, as recently discussed in ref. 19. This is the key piece for our proposal described below.

Our proposal is illustrated in Fig. 1. We consider the QXY-type magnetic quadrupole (Fig. 1a) and, first of all, an achiral case. Note that QXY itself is achiral because it is invariant under mirror operations such as the one parallel to the YZ plane (mYZ). As a result of breaking of time-reversal and space-inversion symmetries, QXY can lead to a linear ME effect (Pi = αijHj, where i, j = X, Y, Z) with finite ME tensor components, αXY = αYX. It indicates that an application of a magnetic field (H) along the X axis (HX) linearly induces P along the Y axis (PY), besides a trivial M along the X axis (MX) (HX = χXXMX, where χXX is the susceptibility). Accordingly, the system in the applied HX possesses the orthogonal configuration of P and M, and, hence, T = P×M along the Z axis (TZ = χXXαYXHX2). The system remains achiral because mYZ survives (Fig. 1b). Now, we introduce chirality by breaking mirror symmetries. Recalling that T combined with C can induce M parallel/unparallel to T, one can expect a new Z-axis component of M (MZ), which is phenomenologically expressed as

$$M_Z = \lambda CT_Z$$
(1)

where λ is a proportionality constant. Combined with TZ = χXXαYXHX2, Eq. (1) is rewritten as

$$M_Z = \lambda \chi _{XX}\alpha _{YX}CH_X^2$$
(2)

This equation predicts that, first of all, the combination of QXY and chirality gives rise to the unusual induction of second-order nonlinear magnetization MZ whose direction is orthogonal to the applied HX. Moreover, this “orthogonal” magnetization is reversed between the chiral pairs (C = −1 and C = +1) related by mYZ (compare Fig. 1c, d) and is also reversed by a switching of a sign QXY that reverses the αYX sign. In the following, we will examine this proposal.

Crystal chirality, antiferromagnetic state, and in-field states of Pb(TiO)Cu4(PO4)4

Our target antiferromagnet Pb(TiO)Cu4(PO4)4 (hereafter, abbreviated as PbTCPO) has a chiral crystal structure belonging to the space group P421220. It is possible to grow homo-chiral or enantiopure single crystals of both left-handed and right-handed forms, which can be distinguished by a natural optical activity. We define the chirality of levo and dextro rotatory crystals as C− and C+ , respectively. The crystal structure of PbTCPO consists of characteristic Cu4O12 square cupolas, as shown in Fig. 2a, and the unit cell contains one upward and one downward Cu4O12 square cupola. Here, the X, Y, and Z axes are taken parallel to [110], \([\bar 110]\), and [001], respectively. Figure 2b, c shows the structures of C− and C+ crystals viewed along the Z axis, respectively, in which only Cu ions are depicted for clarity. In the C− crystal, the Cu4 square of the upward cupola [Cu4(U)] rotates in a counter-clockwise direction, while that of the downward cupola [Cu4(D)] rotates in a clockwise direction. A mirror operation parallel to the YZ plane (mYZ) mutually converts between the C+ and C− crystals. Accordingly, the rotation directions of Cu4(U) and Cu4(D) in the C+ crystal are opposite to those in the C− crystal. For convenience, we define φ as a rotation angle for Cu4(U), such that φ > 0 and φ < 0 represent the C− and C+ crystals, respectively. According to the structural analysis at room temperature, |φ | is estimated to be 4.5°20.

Fig. 2: Crystal structure, chirality, and magnetic domains of Pb(TiO)Cu4(PO4)4.
figure 2

a A three-dimensional view of the crystal and magnetic structures. Pb, Ti, and P atoms are omitted for clarity. A unit cell (solid line) contains one upward and one downward Cu4O12 square cupola unit each formed by four Cu2+ ions (large blue balls) and twelve O2− ligands (black balls). Magenta and green arrows represent inwards and outwards Cu2+ spins, respectively. b, c Staggered rotation of Cu4 squares of upward [Cu4(U)] and downward [Cu4(D)] units of the C− (b) and C+ (c) crystals. φ represents the rotation angle of Cu4(U). φ > 0 and φ < 0 correspond to C+ and C− crystals, respectively. The C+ and C− crystal structures are mutually converted by the mirror plane parallel to the YZ plane (mYZ). dg Z-axis views of the magnetic structures of a pair of time-reversed domains Q+ and Q− for each chirality. In each panel of dg, geometrical configuration of MX, PY, MZ, and TZ induced by an application of a magnetic field along the X axis (HX) is illustrated. Here, MX and MZ are the magnetization along the X and Z axes, respectively, PY is the electric polarization along the Y axis, and TZ is the toroidal moment along the Z axis. The sign of the linear magnetoelectric coefficient αYX that governs the relation of PY and HX is also indicated.

PbTCPO undergoes an antiferromagnetic transition at TN ( = 7.0 K at 0 T)20, as indicated by the magnetic phase diagram in Fig. 3a. In the antiferromagnetic phase, the four Cu spins of each square cupola exhibit a noncoplanar arrangement (Fig. 2a), whose in-plane (Z-plane) and the out-of-plane spin components form two-in two-out and up-down–up-down arrangements, respectively (see Fig. 2d). The magnetic point group of the antiferromagnetic state is 4′22′, which breaks both space-inversion and time-reversal symmetries, and allows for an ordering of QXY and the corresponding linear ME effect with finite αYX (= αYX). As depicted in Fig. 2d–g, each chiral crystal can possess a pair of magnetic quadrupole domains, Q− (QXY < 0) and Q+ (QXY > 0), with opposite signs of αYX, which are mutually converted by time-reversal operation (1′) [compare Fig. 2d(2f) and 2e(2g)]. (The sign of QXY is directly calculated based on the magnetic structure; see Supplementary Figs. 1 and 2, and Supplementary Note 1) According to the previous theoretical analysis of the ME coupling in PbTCPO21, the signs of αYX are negative and positive for the Q+ and Q− domains, respectively, while they remain unchanged under the chirality switching. Because the linear ME effect PY = αYXHX was experimentally confirmed20, PbTCPO realizes the aforementioned situation depicted in Fig. 1, where the anomalous induction of second-order nonlinear magnetization MZ by applying HX can be anticipated. (This is the lowest-order induction of MZ by HX because a linear induction of MZ by HX is prohibited by the 4′22′ symmetry22,23). Recalling that MZ is reversed by a switching of chirality as well as a switching of QXY [Eq. (2)], we explicitly indicate in Fig. 2d–g the sign of MZ for the four distinct states, together with the configurations of PY, MX, and TZ. We assume λ is positive for the sign of MZ to be consistent with experimental results, which will be described below.

Fig. 3: Fundamental magnetoelectric properties of Pb(TiO)Cu4(PO4)4.
figure 3

a Temperature (T) versus magnetic field (H) phase diagram for H applied along the X axis (HX). The Q-phase represents an antiferromagnetic phase with a magnetic quadrupole order, which is a subject in this work. FI and PM denote field-induced and paramagnetic phases, respectively. b Ferroelectric hysteresis loops at temperature (T) of 4.2 K measured in HX whose strength is varied between 0 to 12 T with a 3 T interval. The electric polarization along the Y axis (PY) was measured as a function of the electric field along the Y axis (EY). The positions of the measurement temperature and HX are marked by red crosses in the phase diagram in a. Shown in the insets of b are the geometrical configurations of HX, PY, MX, TZ, and MZ for bistable states at EY = 0. Here, MX and MZ are the magnetization along the X and Z axes, respectively, and TZ is the toroidal moment along the Z axis.

Magnetic-field-induced ferroelectric hysteresis loops

To examine the proposed orthogonal magnetization MZ in PbTCPO, we perform measurements of PY in various conditions in terms of external E and H, since the sign of PY in finite HX directly probes the sign of MZ (see Fig. 2d–g). Figure 3b shows EY dependence of PY for a C+ crystal measured at 4.2 K (< TN) in an applied HX. The strength of μ0HX was varied between 0 and 12 T with a 3 T interval (red crosses in the phase diagram of Fig. 3a). As seen in Fig. 3b, a finite spontaneous PY (PY at EY = 0) emerges in the finite HX, whose magnitude increases almost linearly with increasing HX, indicating that the spontaneous PY originates from PY = αYXHX. Moreover, we observe a clear ferroelectric hysteresis loop, a reversal of PY at a certain threshold magnitude of EY. This reversal corresponds to the switching between the Q− and Q+ domains with opposite signs of MZ, as depicted in the inset of Fig. 3b.

A signature for the existence of MZ is observed when the applied H is slightly tilted from the X axis to the Z axis by an angle Δθ, so that a small Z-axis component of the magnetic field (HZ) is present (see the inset of Fig. 4a). Figure 4a displays the ferroelectric hysteresis loops for the C+ crystal measured at 4.2 K in the tilted H of μ0H = 9 T (−8° ≤ Δθ ≤ +8°). It is seen that introducing the finite HZθ ≠ 0) causes a significant shift of the center of the ferroelectric hysteresis loops along the EY axis, or a bias effect. Moreover, the bias electric field (ΔE) is positive for HZ > 0 (Δθ > 0), whereas negative for HZ < 0 (Δθ < 0). This means that the original degeneracy between the Q− and Q+ domains at EY = 0 is broken by the small HZ component; HZ > 0 stabilizes the Q+ domain (PY < 0, Fig. 2g), whereas HZ < 0 stabilizes the Q− domain (PY > 0, Fig. 2f). Assuming that the Q+ and Q− domains have MZ > 0 and MZ < 0, respectively (i.e., proportionality constant λ > 0), the interaction of HZ and MZ ( = −HZMZ) should differentiate the free energy of the Q− and Q+ domains by 2HZMZ, which naturally explains the observed domain stabilization. This provides the first evidence for the existence of MZ.

Fig. 4: Chirality-magnetism coupling as probed via ferroelectric hysteresis loops.
figure 4

a, b PYEY hysteresis loops for a C+ and b C− crystals in an applied H tilted from the +X direction by an angle Δθ of −8° ≤ Δθ ≤ +8°. Definition of Δθ is illustrated in the inset of a. In the finite H tilting, the ferroelectric hysteresis loops are shifted along the EY axis by ΔE, whose sign depends on the sign of Δθ as well as the chirality switching. The magnetic quadrupole domains (Q+ /Q−) stabilized by the tilted H are indicated by the white and black circles, together with respective signs of orthogonal magnetization MZ.

Second, and conclusive evidence for the existence of MZ is observed in results of the same measurements for a crystal with the opposite chirality (C− crystal). As shown in Fig. 4b, the sign of ΔE for the C− crystal is completely reversed from that of the C+ crystal, meaning that the quadrupole domain stabilized in the same condition of HZ is opposite between the C− and C+ crystals. This is reasonably explained by assuming that the signs of MZ of the respective quadrupole domains are reversed by the chirality switching, in complete agreement with Eq. (2). Therefore, our chirality-specific measurements of the ferroelectric hysteresis loops in the tilted H successfully demonstrate the existence of the proposed MZ arising from the combination of QXY and chirality.

We also confirm that the sign of ΔE does not alter by a reversal of the applied H (i.e., 180° rotation in the ZX plane), as shown in Supplementary Fig. 3. This is consistent with an expected feature of our proposal, as explained below. Let us consider the C+ crystal and a reversal of H from HX > 0 and HZ > 0 to HX < 0 and HZ < 0. As already depicted in Fig. 4a, before the H reversal (HX > 0, HZ > 0), the Q+ domain with PY < 0 and MZ > 0 is stabilized. This is reflected as the positive ΔE in the ferroelectric hysteresis loops. After the reversal of H (HX < 0, HZ < 0), due to the second-order dependence of MZ to HX [Eq. (2)], the Q+ and Q− domains keep in having MZ > 0 and MZ < 0, respectively. Thus, the interaction −HZMZ will stabilize the Q− domain with MZ < 0. In the applied HX < 0, the Q− domain (αYX > 0, see Fig. 2) has PY < 0. To make the PY < 0 state stable at EY = 0, ΔE will be positive, which is thus identical with that before the H reversal. This H-reversal-independent behavior is completely different from a previously reported biased hysteresis loop in a ferroelectric ferromagnet Dy0.7Tb0.3FeO3, where ΔE depends linearly on an applied H24. A more detailed analysis of the observed bias effect in the present compound PbTCPO is given in Supplementary Fig. 4 and Supplementary Note 2.

Single domain state of magnetic quadrupole by tilted magnetic field

The breaking of the degeneracy of the Q+ and Q− domains in tilted H raises an intriguing possibility that a single domain state is achieved by cooling a sample across TN with H only. Figure 5 shows the temperature dependence of PY for the C+ crystal in μ0H = 9 T with Δθ = 0, ±4, or ±8° (see the inset), measured on warming without any electric field. No electric field was applied during a cooling process before each measurement. At Δθ = 0, PY below TN is vanishingly small. This indicates that the Q+ and Q− domains coexist in the sample, in agreement with the degeneracy of the two domains in HZ = 0. By contrast, at Δθ = +8° (HZ > 0) and Δθ = −8° (HZ < 0), finite PY of opposite signs is observed. The values of +40 μC m−2 (−40 μC m−2) at 4.2 K are close to those corresponding to the single domain state of Q+ (Q−) evaluated as the spontaneous PY in the hysteresis loops (see Fig. 4a). Thus, the single domain states of Q+ and Q− are successfully formed by tilted H with HZ > 0 and HZ < 0, respectively. This result is remarkable, because achievement of a single domain state in ME antiferromagnets generally requires not only H, but also E. Therefore, the orthogonal magnetization provides the distinct way of controlling ME antiferromagnetic domains free from E.

Fig. 5: Temperature dependence of PY for the C+ crystal in an applied μ0H of 9 T with Δθ = 0, ±4°, and ± 8°.
figure 5

No electric field was applied during each measurement as well as during a cooling process before each measurement.

Cluster mean-field calculations

The validity of the chirality-induced orthogonal magnetization is further examined theoretically by the cluster mean-field calculations of nonlinear magnetization. According to the previous studies21,25, overall magnetic and ME behaviors of PbTCPO are well reproduced by an effective spin-1/2 model which is described as

$$\begin{array}{ll}H_{eff}=\mathop{\sum}\limits_{\left\langle {i,j} \right\rangle} {\left[ {J_1{\mathbf{S}}_i \cdot {\mathbf{S}}_j - {\mathbf{D}}_{ij} \cdot \left( {{\mathbf{S}}_i \times {\mathbf{S}}_j} \right)} \right]} + J_2\mathop {\sum}\limits_{\left\langle {\left\langle {i,j} \right\rangle} \right\rangle} {{\mathbf{S}}_i \cdot {\mathbf{S}}_j}\\ \qquad\qquad+ \,J^{\prime}\mathop {\sum}\limits_{\left( {i,j} \right)} {{\mathbf{S}}_i \cdot {\mathbf{S}}_j} + J^{\prime\prime}\mathop {\sum}\limits_{\left( {\left[ {i,j} \right]} \right)} {{\mathbf{S}}_i \cdot {\mathbf{S}}_j} - g\mu _B\mathop {\sum}\limits_i {{\mathbf{H}} \cdot {\mathbf{S}}_i}.\end{array}$$
(3)

Here, J1 and J2 are the nearest-neighbor and next-nearest-neighbor intracupola symmetric exchange interactions, respectively, while J′ and Jʺ are the intralayer- and interlayer-intercupola interactions, respectively (see Supplementary Fig. 5). Dij is the so-called Dzyaloshinskii–Moriya (DM) vector describing an antisymmetric DM interaction at J1 bonds that reflects the geometry of the Cu4O12 unit. As depicted in Fig. 6a, the direction of Dij is largely tilted from the Z axis to the X or Y axis by 80°. In the previous calculations, the arrangement of Dij at different J1 bonds was assumed so as to preserve the mirror symmetry mYZ (see Fig. 6a); in other words, the effect of the chirality was neglected. In the present calculation, by contrast, the effects of the chirality are considered by introducing a staggered rotation of Dij about the Z axis by an angle φDM = 5°, which is accompanied by the staggered rotation of Cu4 squares by φ = 5° (corresponding to C+ crystal), as shown in Fig. 6b. We shall call this model the chiral spin-1/2 model. Except for the rotation of Dij, we use the same set of parameters as in the previous calculations: J1 = 1, J2 = 1/7, J′ = 3/4, Jʺ = 1/100, and D = | Dij | = 1.1.

Fig. 6: Calculations based on the chiral spin-1/2 model.
figure 6

a, b Achiral (a) and chiral spin 1/2 model (b) and parameters setting. The red thick arrows represent the DM vector at the J1 bond, D, which is considerably tilted from the Z axis by 80°. A staggered rotation of D by φDM = 5° is introduced in the chiral spin-1/2 model. c Calculated results of the magnetization along the Z axis, mZ, as a function of magnetic field (h) along the X axis, hX, in the absence of h along the Z axis (hZ). The gray symbols represent mZ in the case of φDM = 0 for both q+ and q− , which are zero within the numerical accuracy. The blue squares and the red circles denote mZ for q+ and q− , respectively, in the case of φDM = 5°. d hZ dependence of the free energy f at hX = 0.5 in the case of φDM = 5°.

First, we have calculated the magnetization along the Z axis, mZ, for the two magnetic domains q+ (qXY > 0) and q− (qXY < 0), in the case of φDM = 0 and 5°. A magnetic field (h = BH) along the X axis, hX, is varied between −1.5 and 1.5, while h along the Z axis, hZ, is fixed at zero. (We use the lowercase symbols q, qXY, h, and m to distinguish them from the respective uppercase symbols for the experiment.) The result in both cases of φDM is shown in Fig. 6c. It is seen that, when the chirality is neglected (i.e., φDM = 0), mX is zero within the numerical accuracy for both the q+ and q− states. By contrast, when the chirality is introduced (i.e., φDM = 5°), the q+ and q− states have a finite mZ with the same magnitude but opposite signs. This is consistent with the experimental results. Next, we examine the effect of the field tilting in the XZ plane on the stability of the domains in the chiral system. Figure 6d shows the calculated free energy f as a function of the applied hZ, at a fixed hX of 0.5. It is found that the original degeneracy between the q+ and q− states at hZ = 0 is lifted by introducing finite hZ, in good agreement with the experimental results. Moreover, the relationship between the stabilized domain and the sign of hZ (q + by hZ > 0 and q− by hZ < 0, respectively) is also consistent with the experimental results (see Fig. 2f, g). Taken together, our chiral spin-1/2 model successfully reproduces all the experimental results, strongly supporting the existence of the chirality-induced orthogonal magnetization.

Cluster multipole analysis and phenomenological approach

Motivated by the recent development of the cluster multipole theory, we analyze the experimental results based on the cluster multipoles. It is indicated by Eq. (2) that the orthogonal magnetization MZ is categorized as a second-order nonlinear magnetization with the general form of Mi(2) = χijkHjHk (i,j,k = X,Y,Z). χijk is the second-order nonlinear magnetic susceptibility and the component χZXX corresponds to λχXXαYXC of Eq. (2). According to the comprehensive theoretical work by Hayami et al.7, Mi(2) is associated with magnetic octupoles and/or magnetic toroidal quadrupoles (Supplementary Note 1). As described in Supplementary Fig. 6 and Supplementary Note 3, our additional calculations based on the chiral spin-1/2 model successfully identify that an \(O_Z^\beta\)-type magnetic octupole (see Fig. 7 for schematic representation) gives a dominant contribution to χZXX. The \(O_Z^\beta\) is expressed as

$$O_Z^\beta = \mathop {\sum}\limits_i {\sqrt {15} \left[ {\frac{1}{2}\left( {X_i^2 - Y_i^2} \right)m_i^Z + Z_i\left( {X_im_i^X - Y_im_i^Y} \right)} \right]}$$
(4)

where the sum is taken over eight Cu ions in the unit cell and \(m_i^n\) represents an n-axis (n = X, Y, Z) component of a magnetic dipole mi of the i-th (i = 1−8) Cu ion located at a position ri = (Xi, Yi, Zi). The origin of the coordinate is taken as the center of the unit cell, which is denoted by “O” in Fig. 2d. We emphasize that the presence of chirality is indispensable for finite \(O_Z^\beta\) in PbTCPO, because a hypothetical magnetic structure (magnetic point group 4′/mmm′) has a PT symmetry \((\bar 1^{\prime})\), which makes \(O_Z^\beta\) strictly zero (see Supplementary Fig. 1 and Supplementary Note 1).

Fig. 7: Conceptual illustration of cross-control of magnetic quadrupole and octupole.
figure 7

a In the case of a hypothetical achiral structure, the antiferromagnetic phase of Pb(TiO)Cu4(PO4)4 would possess only magnetic quadrupoles QXY. + and − represent the signs of QXY. The control of the sign of QXY requires a simultaneous application of electric field EY and magnetic field HX (EYHX). b, c By introducing the chirality of the crystal structure, which corresponds to the actual structure of Pb(TiO)Cu4(PO4)4, the magnetic structure acquires additional octupole moments \(O_Z^\beta\). Here, a pair of chiral structures C– (b) and C+ (c) are related to each other by mirror operation mYZ. The signs of \(O_Z^\beta\) depend on those of QXY. Thus, a switching of the sign of QXY accompanies that of \(O_Z^\beta\) or vice versa (see text for details). The signs of QXY are controllable by a nonlinear magnetic field HZHXHX. Accordingly, introducing the chirality allows for an unusual cross-control of QXY and \(O_Z^\beta\) (represented by thick red and blue arrows); namely, control of QXY with a magnetic field only and control of \(O_Z^\beta\) with an electric field in a bias magnetic field. The sign of QXY and \(O_Z^\beta\) selected by this cross-control is reversed by a switching of the chirality, as experimentally observed.

Figure 7 conceptually illustrates a coupling among \(O_Z^\beta\), QXY, and the chirality C, as well as field responses of \(O_Z^\beta\) and QXY, which is obtained by a phenomenological approach based on the Landau theory26,27 as described below. To take chirality into account, we use a hypothetical achiral crystal structure with the space group P4/nmm, which was recently observed in a sister compound Ba(TiO)Cu4(PO4)4 at high temperatures28. Table 1 summarizes the transformations of various order parameters for PbTCPO based on the generators of the point group 4/mmm22,23,29 for the achiral case. We consider this point group, because all the listed parameters are invariant under the translational symmetry of P4/nmm. We also add the time-reversal operation 1′.

Table 1 Transformations of various order parameters for Pb(TiO)Cu4(PO4)4 based on the generators of 4/mmm point group.

From Table 1, one can find that the invariant E1 ~ (PXMY + PYMX)QXY stands for the achiral case (4/mmm) [and of course also for the chiral case (422) because 422 is a subgroup of 4/mmm]. If we use an electric field E and a magnetic field H to manipulate P and M, we can reformulate E1 ~ (EXHY + EYHX)QXY. This can explain the reason why the sign of QXY is controllable by, e.g., EYHX, and is in one-to-one relation with the sign of αYX = αXY (= −∂2E1/∂EXHY QXY). Similarly, the invariant E2 ~ (HZHXHX − HZHYHY) \(O_Z^\beta\) suggests that the sign of \(O_Z^\beta\) is controllable by, i.e., HZHXHX, and is in one-to-one relation with the sign of χZXX = −χZYY (= −∂3E2/∂HZHXHX \(O_Z^\beta\)). Note that the achiral case (i.e., with mXY) can only support QXY or \(O_Z^\beta\), but not the both, because they transform in opposite ways with respect to mXY. In the experiment, QXY was selected for the achiral case. Conversely, \(O_Z^\beta\) (and corresponding χZXX) was absent for the achiral case.

By contrast, when introducing chirality (i.e., removing mYZ), both QXY and \(O_Z^\beta\) are transformed identically. Hence, they can coexist, and, moreover, the switching of the sign of QXY simultaneously changes the sign of \(O_Z^\beta\). This allows for the cross-control of QXY and \(O_Z^\beta\) by their non-conjugate fields, that is, the control of QXY by the tilted magnetic field (HZHXHX) and the control of \(O_Z^\beta\) by EYHX, as indicated by thick red and blue arrows in Fig. 7b, c. This explains the reason why the quadrupole domains can be controlled by the tilted H. Furthermore, the switching of chirality reverses the sign relation between QXY and \(O_Z^\beta\) (compare Fig. 7b, c), because mXY transforms QXY and \(O_Z^\beta\) in opposite ways. It results in a reversal of the sign of QXY selected by the tilted H, in agreement with the experimental results. Therefore, all the experimental results originate from the combination among magnetic quadrupoles, magnetic octupoles, and chirality.

Discussion

In summary, starting from the simple symmetry considerations depicted in Fig. 1, we have proposed that the QXY magnetic quadrupole and the chirality together give rise to the anomalous induction of orthogonal magnetization MZ in an applied magnetic field HX, which arises from the coupling of the chirality and the HX-induced toroidal moment (TZ) due to the linear ME effect of QXY. Using the model material PbTCPO, we have substantiated this proposal by observing that the domain state of QXY can be controlled not only by the electric and magnetic fields together (EYHX), but also by the tilted magnetic field only (HZHXHX). We have also associated the experimental results with the cluster multipoles, revealing that MZ is viewed as the second-order nonlinear magnetization and originates from the magnetic octupole \(O_Z^\beta\). Moreover, our phenomenological model has revealed that the signs of QXY and \(O_Z^\beta\) are coupled due to the chirality, which allows for the unusual cross-control of QXY and \(O_Z^\beta\) by the EYHX and HZHXHX fields, respectively.

Our observation of the electric-field-free control of the magnetic quadrupole domains in the chiral-lattice antiferromagnet PbTCPO has various implications for other antiferromagnets with ME multipoles (i.e., magnetic monopole, magnetic quadrupole, and magnetic toroidal dipole). Of the 122 magnetic point groups, there are six enantiomorphic, antiferromagnetic groups [222, 422, 4′22′ (present case), 32, 622, and 23] that allow both the linear ME effect (hence, ME multipoles) and second-order nonlinear magnetization. The present control method might be applicable to antiferromagnets belonging to these magnetic classes. In particular, the application to metallic antiferromagnets is of great interest. In metals, manipulation of ME multipoles by an electric field is severe challenge because of the screening of an applied electric field due to conduction electrons. Although an electric-current control of magnetic toroidal dipole order in metals has been recently demonstrated30,31,32, an analogous electric-current control of other two types of ME multipoles (i.e., magnetic monopole and quadrupole) cannot be anticipated because symmetries of these ME multipoles do not match with that of electric current32. Thus, the present electric-field-free method can be a useful way of controlling these ME multipoles in metals. It is also worthy to note that chirality can be externally induced by a mechanical twist in a nonchiral material, as observed in bismuth3. This can expand the applicability of the method to nonchiral antiferromagnets by simultaneously applying a mechanical twist and a magnetic field.

The present study reveals that PbTCPO possesses two different switchable (i.e., ferroic) orders: magnetic quadrupole and magnetic octupole orders. It can be viewed as a distinct type of multiferroic in which a coupling between these two orders gives rise to the observed unconventional cross-coupled phenomena. Moreover, the coupling originates from the presence of the crystal chirality. Further explorations of chirality-driven phenomena in PbTCPO and other chiral-lattice time-reversal-broken antiferromagnets are of great interest.

Methods

Sample preparation and characterization

Single crystals of Pb(TiO)Cu4(PO4)4 were grown by slow cooling of the melt20. The crystal chirality of grown crystals was distinguished by a sense of optical rotatory measured by means of polarized light microscopy, in which the incident light was parallel to the Z axis and light wavelength was 500 ± 10 nm. We defined the crystal chirality of levo-rotatory crystals as C− and that of dextro-rotatory crystals as C+ . Monodomain crystals with C+ and C− were used for electric polarization measurements described below. The orientation of the crystals was determined using the Laue X-ray photograph. The crystal structure displayed in Fig. 2a was drawn by using VESTA software33.

Electric polarization measurements

Electric polarization measurements on C+ and C− crystals were performed at the High Field Laboratory for Superconducting Materials in Tohoku University. Electric polarization along the Y axis (PY) as a function of temperature was obtained by integrating a displacement current measured with an electrometer (Keithley 6517). Before each measurement, the samples were cooled from 10 K ( > TN) to 2 K ( < TN) with or without a poling electric field EP, after which, if applied, EP was switched off. Subsequently, the displacement current was measured on warming. To obtain ferroelectric hysteresis loops, i.e., the EY dependence of PY, we applied a triangle wave voltage to the samples with a frequency of 5 mHz, which was generated by using a commercial voltage amplifier and a function generator. The electric polarization was obtained by integrating a displacement current induced by the voltage sweeping. In these measurements, a magnetic field was applied in the direction parallel to the XZ plane, which makes an angle −8° ≤ Δθ ≤ +8° from the X axis. The angle Δθ was controlled by rotating a sample stage about the Y axis using a stepping motor.

Theoretical calculations

We employ the cluster mean-field theory of the chiral spin-1/2 model [Eq. (3)] for a microscopic understanding of the nonlinear magnetization in the system. In this treatment, the intercupola interactions are dealt with by the conventional mean-field approximation; namely \({\mathbf{S}}_i \cdot {\mathbf{S}}_j\) is decoupled as \({\mathbf{S}}_i \cdot {\mathbf{S}}_j \simeq \left\langle {{\mathbf{S}}_i} \right\rangle \cdot {\mathbf{S}}_j + {\mathbf{S}}_i \cdot \left\langle {{\mathbf{S}}_j} \right\rangle - \left\langle {{\mathbf{S}}_i} \right\rangle \cdot \left\langle {{\mathbf{S}}_j} \right\rangle\), while all the quantum fluctuations from the intracupola interactions and the Zeeman coupling term are fully taken into account by the exact diagonalization. In the calculation, we considered four cupolas in two unit cells; up and down cupolas in a single unit cell and those in a neighboring unit cell stacked along the Z direction (see Supplementary Fig. 5). However, we always found solutions with the two unit cells identical to the single unit cell shown in Fig. 2a. This theoretical approach is suitable for the model analysis of cluster-based compounds, and has been used to qualitatively explain the magnetoelectric behavior in the series of compounds, Ba(TiO)Cu4(PO4)425, Pb(TiO)Cu4(PO4)421, and Sr(TiO)Cu4(PO4)434.