PreprintPDF Available

Spin-frame field theory of a three-sublattice antiferromagnet

Authors:
Preprints and early-stage research may not have been peer reviewed yet.

Abstract and Figures

We present a nonlinear field theory of a three-sublattice hexagonal antiferromagnet. The order parameter is the spin frame, an orthogonal triplet of vectors related to sublattice magnetizations and spin chirality. The exchange energy, quadratic in spin-frame gradients, has three coupling constants, only two of which manifest themselves in the bulk. As a result, the three spin-wave velocities satisfy a universal relation. Vortices generally have an elliptical shape with the eccentricity determined by the Lam\'e parameters.
Content may be subject to copyright.
Spin-Frame Field Theory of a Three-Sublattice Antiferromagnet
Bastian Pradenas and Oleg Tchernyshyov
William H. Miller III Department of Physics and Astronomy,
Johns Hopkins University, Baltimore, Maryland 21218, USA
(Dated: August 1, 2023)
We present a nonlinear field theory of a three-sublattice hexagonal antiferromagnet. The order
parameter is the spin frame, an orthogonal triplet of vectors related to sublattice magnetizations and
spin chirality. The exchange energy, quadratic in spin-frame gradients, has three coupling constants,
only two of which manifest themselves in the bulk. As a result, the three spin-wave velocities satisfy
a universal relation. Vortices generally have an elliptical shape with the eccentricity determined by
the Lam´e parameters.
Theory of magnetic solids is often described as a lattice
problem, exemplified by the Heisenberg model of atomic
spins. However, discrete models are notoriously difficult
to solve outside of the simplest tasks, such as finding
the spectrum of linear spin waves. An analytic theory
of nonlinear solitons—such as domain walls or vortices—
in the framework of a lattice model is often not feasible
or cumbersone [1]. Continuum theories have the clear
advantage of being more amenable to analytic treatment.
By design, they focus on the physics of long distances
and times, capturing the universal aspects of low-energy
physics at the expense of microscopic details.
In magnetism, a well-known example is micromagnet-
ics, the continuum theory of a ferromagnet going back to
Landau and Lifshits [2]. It is usually formulated through
the equation of motion for the magnetization field mof
unit length parallel to the local direction of spins,
Sm
∂t =m×δU
δm.(1)
Here Sis the spin density, U[m] = RddrUis a potential-
energy functional, whose functional derivative δU/δm
acts as an effective magnetic field. The energy density is
usually dominated by the Heisenberg exchange interac-
tion of strength A,
U=A
2im·im+. . . (2)
Doubly repeated indices imply summation. The omitted
terms represent weaker anisotropic interactions of dipolar
and relativistic origin. We use the calligraphic font to
indicate intensive quantities (densities).
The Landau-Lifshitz equation (1) provides a starting
point for understanding the dynamics of ferromagnetic
solitons. A further coarse-graining eliminates fast inter-
nal modes of a soliton and focuses on its slow collective
motion whose seminal achievements; primary examples
are Thiele’s equation of rigid motion [3] and Walker’s
dynamical model of a domain wall [4].
The continuum approach has also been applied to sim-
ple antiferromagnets, in which adjacent spins are (nearly)
antiparallel and can be split into two magnetic sublattices
1 and 2, each with its own magnetization field m1and
m2of unit length. Because at low energies, the sublattice
magnetizations are (nearly) antiparallel, both can be ap-
proximated by a single field of staggered magnetization
nmA mB, whose dynamics is described by an
O(3) σ-model with the Lagrangian density [57]
L=K U =ρ
2tn·tnA
2in·in. . . (3)
The first term K=ρ
2tn·tnis the kinetic energy of
staggered magnetization and ρis a measure of inertia.
The second, potential term comes from the Heisenberg
exchange energy and has the same functional form as in
a ferromagnet (2). The omitted terms represent various
weak anisotropic interactions. Minimization of the action
with the constraint n2= 1 yields the equation of motion
ρ t(n×tn) = n×δU
δn.(4)
As with ferromagnets, the antiferromagnetic Landau–
Lifshitz equation (4) can be translated into equations of
motion for solitons [8,9] and extended to include the
effects of spin transfer and dissipation [10,11].
The primary goal of this paper is to introduce a univer-
sal field theory for an antiferromagnet with three mag-
netic sublattices whose magnetization fields satisfy the
relation m1+m2+m30. Such magnetic states are
typically realized in antiferromagnetic solids of hexag-
onal symmetry with triangular motifs. Although such
magnets have been studied for decades [12,13], recent
experimental studies of metallic antiferromagnets Mn3Sn
and Mn3Ge [14,15] have rekindled theoretical interest in
these frustrated magnets [1618].
The existing field theory for 3-sublattice antiferromag-
nets by Dombre and Read [12] has a couple of drawbacks.
First, it is formulated specifically for the triangular lat-
tice, which has a higher spatial symmetry than other
hexagonal lattices (such as kagome) and therefore misses
some of the universal features of 3-sublattice antiferro-
magnets. Second, their mathematical formalism repre-
sents the magnetic order parameter as a 3 ×3 rotation
matrix, an abstract, and not very intuitive mathematical
object.
arXiv:2307.16853v1 [cond-mat.str-el] 31 Jul 2023
2
m2
m1m3
(a) (b)
nx
ny
nz
FIG. 1. (a) A ground state of the kagome antiferromagnet
and its sublattice magnetizations m1,m2, and m3. (b) The
corresponding spin-frame vectors nx,ny, and nz.
We derive a nonlinear field theory of a 3-sublattice
antiferromagnet with the order parameter represented
by a spin frame, i.e., a triad of orthonormal vectors
ˆ
n {nx,ny,nz}, directly related to sublattice magne-
tizations m1,m2, and m3. At low energies, the mag-
netization dynamics reduce to rigid rotations of the spin
frame, tni=×ni, at a local angular frequency .
One of our main results is the Landau–Lifshitz equation
for a 3-sublattice antiferromagnet,
ρ t=ni×δU
δni
.(5)
A sum over doubly repeated Roman indices, i=x, y, z, is
implied hereafter. Like its analogs (1) and (4), it equates
the rate of change of the local density of angular momen-
tum with the torque density from conservative forces ex-
pressed by a potential energy functional U(ˆ
n). The trans-
parent physical meaning of the Landau–Lifshitz equation
makes it easy to add other relevant perturbations.
To define the spin frame ˆ
n, we first switch from the
three unit-vector fields of sublattice magnetizations m1,
m2, and m3to uniform magnetization mand two stag-
gered magnetizations nxand ny(Fig. 1):
m=m1+m2+m3,
nx= (m2m1)/3,(6)
ny= (2m3m2m1)/3.
To them, we add the vector spin chirality [19]
nz=2
33(m1×m2+m2×m3+m3×m1).(7)
As long as m= 0, sublattice fields m1,m2, and m3are
coplanar and thus define the spin plane. Staggered mag-
netizations nxand nylie in the spin plane, whereas spin
chirality nzis orthogonal to it. The three unit vectors ni
form a right-oriented orthonormal spin frame:
ni·nj=δij ,ni×nj=ϵijk nk.(8)
To derive the dynamics of the spin frame, we follow the
standard Lagrangian approach [6,12,20] and integrate
(b)
x
y
(c)
z
12
3
33
21
12
3
12312
33
21 12
3
12
FIG. 2. Hexagonal lattices and their magnetic sublattices:
(a) kagome, (b) triangular lattice. (c) Spatial coordinate axes
x,y, and z. Red, green, and blue colors indicate magnetic
sublattices 1, 2, and 3. Filled triangles and dotted lines denote
C3and C2rotation axes, respectively.
out the hard field of uniform magnetization mto obtain
the dynamics of the spin frame. Our starting point is the
Landau–Lifshitz equations for sublattice magnetizations,
Stm1=m1×δV
δm1
,(9)
and similarly for sublattices 2 and 3. Like in 2-sublattice
antiferromagnets [21], the potential energy functional V
is dominated by the antiferromagnetic exchange interac-
tion imposing a penalty for m= 0 [22],
V(m,ˆ
n) = m2
2χ+U(ˆ
n),(10)
where χis the paramagnetic susceptibility. The subdom-
inant term U[ˆ
n], expressing the energy of the antiferro-
magnetic order parameter, will be discussed below.
With the aid of Eqs. (6), (9), and (10) we find that, like
in 2-sublattice antiferromagnets [5], staggered magneti-
zationss nxand nyprecess about the direction of uniform
magnetization mat the angular velocity [20]
m
χS.(11)
The linear proportionality between the precession fre-
quency and uniform magnetization (11) can be derived
as the equation of motion for uniform magnetization m
from the following Lagrangian for fields mand ˆ
n[20]:
L(m,ˆ
n) = Sm·m2
2χ U(ˆ
n).(12)
The angular velocity can be expressed explicitly in terms
of ˆ
nvia the kinematic identity
=1
2ni×tni,(13)
The first term in the Lagrangian (12) is linear in the ve-
locities tni, so its action represnts the spin Berry phase.
It yields the expected density of angular momentum Sm.
3
Lagrangian (12) is quadratic in uniform magnetization
m. Integrating out this field with the aid of its equation
of motion (11) yields an effective Lagrangian for the re-
maining fields ˆ
nendowed with kinetic energy of rotation
with the inertia density ρ=χS2:
L(ˆ
n) = ρ2
2 U(ˆ
n) = ρ
4tni·tni U(ˆ
n).(14)
Minimizing the action in the presence of holonomic
constraints (8) yields equations of motion with undeter-
mined Lagrange multipliers [23] Λij = Λji:
I
22
tni=δU
δniΛij nj.(15)
Finally, we take a cross product with ni, sum over i, and
use Eq. (13) to obtain the Landau–Lifshitz equation (5).
The energy functional U[ˆ
n] is usually dominated by
the Heisenberg exchange interaction. The latter respects
the SO(3) symmetry of global spin rotations and there-
fore depends not on the orientation of the spin frame ˆ
n
but rather on its spatial gradients. Like in ferromagnets
and 2-sublattice antiferromagnets, the exchange energy
is quadratic in αnβ, where Greek indices take on values
α=xand yonly. The form of these quadratic terms
is restricted by the D3point-group rotational symmetry
of a hexagonal lattice (Fig. 2), including ±2π/3 spatial
rotations about a C3axis normal to the xy plane and
πspatial rotations about C2axes lying in the xy plane.
Under these transformations, the staggered magnetiza-
tions (nx,ny) transform in terms of each other in the
same way as the in-plane components (kx, ky) of a spatial
vector kdo; chirality nztransforms as kz. This obser-
vation helps to construct energy terms quadratic in the
gradients of staggered magnetizations and invariant un-
der both spin rotations and lattice symmetries. To that
end, we may start with a rank-4 spatial tensor and spin
scalar αnβ·γnδand contract its spatial indices pair-
wise to form a spatial scalar. This procedure yields our
second main result, the three possible gradient terms for
the exchange energy density,
U=λ
2αnα·βnβ+µ
2αnβ·αnβ+ν
2αnβ·βnα.(16)
This expression resembles the elastic energy density of
an isotropic solid [24], albeit with 3 Lam´e constants.
A triangular lattice has an extra spatial symmetry.
Under primitive lattice translations, sublattice indices
undergo cyclic permutations, see Fig. 2(b). Staggered
magnetizations nαeffectively undergo ±2π/3 spatial ro-
tations, whereas gradients αdo not. Thus translational
symmetry forbids the λand νterms for a triangular lat-
tice.
The Landau–Lifshitz equation (5) for a 3-sublattice
Heisenberg antiferromagnet reads
ρ t= (λ+ν)nα×αβnβ+µnα×ββnα.(17)
Note that the exchange coupling constants λand νenter
the equation of motion through a combination λ+ν,
rather than individually. More on that below.
An antiferromagnet with nearest-neighbor exchange
interaction Jhas λ+ν= 0 and µ=JS23/4 on a
triangular lattice; on kagome, λ+ν=3JS2/4 and
µ= 0 on kagome. See Supplemental Material [25] for
contributions of further-neighbor interactions.
In what follows, we use the spin-frame formulation of
the field theory to obtain the properties of excitations:
spin waves and vortices.
Spin waves. Linear spin waves on top of a uniform
ground state ni= const can be parametrized in terms
of infinitesimal rotations of the spin frame, δni=δϕ×
ni. Here δϕ=niδϕiis a triplet of infinitesimal rotation
angles δϕiabout the corresponding spin axes; tδϕ=.
Assuming a plane wave with wavenumber ktravelling
along the xdirection, δϕ(t, x) = eδϕ ei(kxωt), we ob-
tain three spin waves with ω=ck and the following
polarizations eand velocities c:
eI=nx, cI=pµ/ρ,
eII =ny, cII =p(λ+µ+ν)/ρ,
eIII =nz, cIII =p(λ+ 2µ+ν)/ρ.
(18)
The velocities satisfy the identity
c2
I+c2
II =c2
III .(19)
Modes I, II are analogs of transverse and longitudinal
sound in a hexagonal solid in two dimensions.
Vortices. The existence of topologically stable point
defects—vortices—in a 3-sublattice Heisenberg antifer-
romagnet was first pointed out by Kawamura and
Miyashita [19]. A 2πrotation of the spin frame corre-
sponds to a loop in the order-parameter space that can-
not be continuously deformed to a point. It is convenient
to parametrize the orientation of the spin frame by start-
ing with a reference uniform configuration nx= (1,0,0),
ny= (0,1,0), nz= (0,0,1), and applying consecutive
Euler rotations through angles ϕabout nz,θabout ny,
and ψabout nz. On a triangular lattice (λ+ν= 0), a
vortex configuration with the lowest energy is described
by the Euler angles ϕ,θ, and ϕgiven by
e =x+iy
|x+iy|, θ =π
2, ψ = const.(20)
This expression agrees well with a numerically obtained
vortex configuration for a triangular lattice, Fig. 3(a).
4
FIG. 3. Vortices in 3-sublattice abtiferromagnets. (a) Triangular lattice with nearest-neighbor interactions only. (b) kagome
lattice with first and third-neighbor interactions, J3=J
3=J1/20. See Supplemental Material [25] for the definition of
further-neighbor interactions. Red, green, and blue arrows are spins of the three magnetic sublattices. The blue spins point
away from the viewer; the red and green ones have components pointing toward the reader. The circle and ellipse reflect the
expected shape of the vortex with the major axis ratio b=p(λ+µ+ν).
Although we have not been able to find an exact vor-
tex solution for a generic hexagonal antiferromagnet, we
can understand the effect of the λ+νterm on the vortex
shape perturbatively. Starting with the isotropic solu-
tion (20) for λ+ν= 0, we keep θ=π/2 and choose
for simplicity ψ=π/2 to obtain the following energy
density:
U=λ+µ+ν
2(xϕ)2+µ
2(yϕ)2.(21)
The vortex acquires an elliptical shape with the major
axis ratio b=p(λ+µ+ν):
e =bx +iy
|bx +iy|, θ =π
2, ψ =π
2.(22)
Although this result is obtained in the limit λ+νµ,
our numerical calculations on a kagome lattice demon-
strate its accuracy even in the opposite limit. The right
panel of Fig. 3(b) shows a vortex in a kagome antifer-
romagnet with the ratio of first and third-neighbor ex-
change interactions J1/J3=20, or (λ+ν) = 5/3.
Its shape agrees with the extrapolated semiaxis ratio
b=p8/3.
Discussion. In this paper, we have presented a uni-
versal field theory of a hexagonal antiferromagnet with
3 magnetic sublattices. The order parameter is a spin
frame constructed from the sublattice magnetizations
and vector chirality. Its mechanics is fully specified by
the inertia of the spin frame Iand three Lam´e constants
λ,µ, and ν. The simple and versatile field theory enabled
us to establish a Pythagorean identity for the three spin-
wave velocities (19) and to predict a generally elliptical
shape for vortices (22).
It is worth noting that the three Lam´e constants enter
the equations of motion (17) in the form of two linear
combinations, λ+νand µ. As a result of that, the three
spin-wave velocities are constrained by a Pythagorean
identity (19). The origin of this behavior can be under-
stood by examining the exchange energy density (16).
The λand νterms in it can be obtained from one an-
other through integration by parts. Thus their infinitesi-
mal variations are the same—up to boundary terms—and
so they make identical contributions to classical dynam-
ics. The difference λνis a “silent” coupling constant
that does not manifest itself in the classical dynamics of
magnetization. It can be seen that it has an intriguing
topological nature with the aid of the identity
1
2(αnα·βnβαnβ·βnα) = nz·(xnz×ynz).(23)
We now see that the energy term (23) is a topological
quantity proportional to the skyrmion density of the spin
chirality (7). Its topological nature assures that it gives a
quantized contribution to the exchange energy that does
not change under continuous variations of the order pa-
rameter and therefore does not affect classical dynamics.
However, it may lead to interesting boundary effects such
as the presence of edge modes, as discussed recently in
a different context by Dong et al. [26]. We will address
the topological aspects of this field theory in a separate
publication [27].
Acknowledgments. We thank Boris Ivanov and Se
Kwon Kim for useful discussions. The research has been
supported by the U.S. Department of Energy, Office of
Science, Basic Energy Sciences under Award No. DE-
SC0019331.
5
[1] I. G. Gochev, Zh. Eksp. Teor. Fiz. 85, 199 (1983), [Sov.
Phys. JETP 58, 115–119 (1983)].
[2] L. Landau and E. Lifshits, Phys. Z. Sowjet. 8, 153 (1935),
reprinted in Collected Papers of L. D. Landau, Pergamon,
New York, pp. 101-114 (1965).
[3] A. A. Thiele, Phys. Rev. Lett. 30, 230 (1973).
[4] N. L. Schryer and L. R. Walker, J. Appl. Phys. 45, 5406
(1974).
[5] I. V. Bar’yakhtar and B. A. Ivanov, Fiz. Nizk. Temp. 5,
759 (1979), [Sov. J. Low Temp. Phys. 5, 361 (1979)].
[6] A. F. Andreev and V. I. Marchenko, Usp. Fiz. Nauk 130,
39 (1980), [Sov. Phys. Usp. 23, 21–34 (1980)].
[7] F. D. M. Haldane, Phys. Rev. Lett. 50, 1153 (1983).
[8] I. V. Bar’yakhtar, B. A. Ivanov, and A. L. Sukstanskii,
Zh. Eksp. Teor. Fiz. 51, 757 (1979), [Sov. Phys. JETP
78, 1509–1522 (1979)].
[9] V. G. Bar’yakhtar, B. A. Ivanov, and M. V. Chetkin,
Usp. Fiz. Nauk 146, 417 (1985), [Sov. Phys. Usp. 28,
563–588 (1985)].
[10] K. M. D. Hals, Y. Tserkovnyak, and A. Brataas, Phys.
Rev. Lett. 106, 107206 (2011).
[11] E. G. Tveten, A. Qaiumzadeh, O. A. Tretiakov, and
A. Brataas, Phys. Rev. Lett. 110, 127208 (2013).
[12] T. Dombre and N. Read, Phys. Rev. B 39, 6797 (1989).
[13] A. B. Harris, C. Kallin, and A. J. Berlinsky, Phys. Rev.
B45, 2899 (1992).
[14] S. Nakatsuji, N. Kiyohara, and T. Higo, Nature 527, 212
(2015).
[15] Y. Chen, J. Gaudet, S. Dasgupta, G. G. Marcus, J. Lin,
T. Chen, T. Tomita, M. Ikhlas, Y. Zhao, W. C. Chen,
M. B. Stone, O. Tchernyshyov, S. Nakatsuji, and C. Bro-
holm, Phys. Rev. B 102, 054403 (2020).
[16] J. Liu and L. Balents, Phys. Rev. Lett. 119, 087202
(2017).
[17] Y. Yamane, O. Gomonay, and J. Sinova, Phys. Rev. B
100, 054415 (2019).
[18] S. Dasgupta and O. Tchernyshyov, Phys. Rev. B 102,
144417 (2020).
[19] H. Kawamura and S. Miyashita, J. Phys. Soc. Jpn. 53,
4138 (1984).
[20] H. Ochoa, R. Zarzuela, and Y. Tserkovnyak, Phys. Rev.
B98, 054424 (2018).
[21] S. K. Kim, Y. Tserkovnyak, and O. Tchernyshyov, Phys.
Rev. B 90, 104406 (2014).
[22] B. I. Halperin and W. M. Saslow, Phys. Rev. B 16, 2154
(1977).
[23] L. D. Landau and E. M. Lifshitz, Mechanics, 3rd ed.,
Course of Theoretical Physics, Vol. 1 (Butterworth-
Heinemann, New York, 1976).
[24] L. D. Landau and E. M. Lifshitz, Theory of Elasticity, 3rd
ed., Course of Theoretical Physics, Vol. 7 (Butterworth-
Heinemann, New York, 1986).
[25] Supplemental Material.
[26] Z. Dong, O. Ogunnaike, and L. Levitov, Phys. Rev. Lett.
130, 206701 (2023).
[27] B. Pradenas, E. G. Ozaktas, and O. Tchernyshyov (un-
published), in preparation.
Supplemental Material
Further neighbors exchange interactions
Convention for further-neighbor interactions is shown in Fig. S1.
J2
J3
J1J1
J3J2J
3
FIG. S1. First, second, and third-neighbors exchange interactions. The kagome lattice has two distinct types of third neighbors.
In antiferromagnets with up to third-neighbor exchange interactions, the field-theory parameters are as follows. For
the triangular lattice,
λ+ν= 0, µ =3
4(J16J2+ 4J3)S2, ρ1=93
2(J1+J3)a2.(S1)
Note that λ+νvanishes in accordance with the higher spatial symmetry of the triangular lattice.
6
For kagome,
λ+ν=3
4(J13J22J32J
3)S2, µ =33
4(J2J3J
3)S2, ρ1= 43(J1+J2)a2.(S2)
In both cases, ais the distance between first neighbors.
For a triangular lattice with first-neighbor interactions only, we obtain
cI=cII =33
22J1Sa, cIII =33
2J1Sa, (S3)
in agreement with Dombre and Read [12]. For a kagome antiferromagnet with J
3= 0, we find
cI= 3p(J1+J2)(J2J3)Sa, cII =p3(J1+J2)(J15J3)Sa, cIII =p3(J1+J2)(J1+ 3J28J3)Sa, (S4)
in agreement with Harris et al. [13].
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
We construct a field-theoretic description of spin waves in hexagonal antiferromagnets with three magnetic sublattices and coplanar 120∘ magnetic order. The three Goldstone modes can be separated by point-group symmetry into a singlet α0 and a doublet β. The α0 singlet is described by the standard theory of a free relativistic scalar field. The field theory of the β doublet is analogous to the theory of elasticity of a two-dimensional isotropic solid with distinct longitudinal and transverse “speeds of sound.” The well-known Heisenberg models on the triangular and kagome lattices with nearest-neighbor exchange turn out to be special cases with accidental degeneracy of the spin-wave velocities. The speeds of sound can be readily calculated for any lattice model. We apply this approach to the compounds of the Mn3X family with stacked kagome layers.
Article
Full-text available
These authors contributed equally to this work Quantum materials with strong transport responses to disparate physical quantities are of great fundamental significance and may hold technological potentials. The interplay between 1 arXiv:2001.09495v1 [cond-mat.str-el] 26 Jan 2020 interactions and topology drive such responses through the effects of spontaneous symmetry breaking and the associated domain configurations on quantum transport. Here we provide a comprehensive description of the magnetism of Mn 3 Ge, an antiferromagnetic kagome-based semimetal with room temperature transport anomalies associated with topologically protected Weyl nodes. Using polarized neutron diffraction, we show the all-important magnetic structure is anti-chiral and coplanar carrying the symmetry of a ferromagnet without appreciable magnetization. We probe and classify the long wavelength excitations that determine its macroscopic responses including a set of collective magneto-elastic modes. We develop a phenomenological spin Hamiltonian with exchange, Dzyaloshinskii-Moriya, and crystal field interactions to describe its collective magnetism. The itinerant character of the magnetism that drives quantum transport is apparent in spin wave damping and extended magnetic interactions. Our work provides the scientific basis for manipulation of the chiral antiferromagnetic texture of Mn 3 Ge to control its topological quantum transport.
Article
Full-text available
We present a theoretical formalism to address the dynamics of textured, noncollinear antiferromagnets subject to spin current injection. We derive sine-Gordon type equations of motion for the antiferromagnets, which are applicable to technologically important antiferromagnets such as Mn3Ir and Mn3Sn, and enables an analytical approach to domain wall dynamics in those materials. We obtain the expression for domain wall velocity, which is estimated to reach ∼1 km/s in Mn3Ir by exploiting spin Hall effect with electric current density ∼1011A/m2.
Article
Full-text available
We theoretically study the interplay between bulk Weyl electrons and magnetic topological defects, including magnetic domains, domain walls, and Z6\mathbb{Z}_6 vortex lines, in the antiferromagnetic Weyl semimetals \mnsn\ and Mn3_3Ge with negative vector chirality. We argue that these materials possess a hierarchy of energies scales which allows a description of the spin structure and spin dynamics using a XY model with Z6\mathbb{Z}_6 anisotropy. We propose a dynamical equation of motion for the XY order parameter, which implies the presence of Z6\mathbb{Z}_6 vortex lines, the double-domain pattern in the presence of magnetic fields, and the ability to control domains with current. We also introduce a minimal electronic model which allows efficient calculation of the electronic structure in the antiferromagnetic configuration, unveiling Fermi arcs at domain walls, and sharp quasi-bound states at Z6\mathbb{Z}_6 vortices. Moreover, we have shown how these materials may allow electronic-based imaging of antiferromagnetic microstructure, and propose a possible device based on domain-dependent anomalous Hall effect.
Article
We argue that spin- and valley-polarized metallic phases recently observed in graphene bilayers and trilayers support chiral edge modes that allow spin waves to propagate ballistically along system boundaries without backscattering. The chiral edge behavior originates from the interplay between the momentum-space Berry curvature in Dirac bands and the geometric phase of a spin texture in position space. The edge modes are weakly confined to the edge, featuring dispersion that is robust and insensitive to the detailed profile of magnetization at the edge. This unique character of edge modes reduces their overlap with edge disorder and enhances the mode lifetime. The mode propagation direction reverses upon reversing valley polarization, an effect that provides a clear testable signature of geometric interactions in isospin-polarized Dirac bands.
Article
Spin superfluidity, i.e., coherent spin transport mediated by topologically stable textures, is limited by parasitic anisotropies rooted in relativistic interactions and spatial inhomogeneities. Since structural disorder in amorphous magnets can average out the effect of these undesired couplings, we propose this class of materials as platforms for superfluid spin transport. We establish nonlinear equations describing the hydrodynamics of spin in insulating amorphous magnets, where the currents are defined in terms of coherent rotations of a noncollinear texture. Our theory includes dissipation and nonequilibrium torques at the interface with metallic reservoirs. This framework allows us to determine different regimes of coherent dynamics and their salient features in nonlocal magnetotransport measurements. Our work paves the way for future studies on macroscopic spin dynamics in materials with frustrated interactions.
Article
In ferromagnetic conductors, an electric current may induce a transverse voltage drop in zero applied magnetic field: this anomalous Hall effect is observed to be proportional to magnetization, and thus is not usually seen in antiferromagnets in zero field. Recent developments in theory and experiment have provided a framework for understanding the anomalous Hall effect using Berry-phase concepts, and this perspective has led to predictions that, under certain conditions, a large anomalous Hall effect may appear in spin liquids and antiferromagnets without net spin magnetization. Although such a spontaneous Hall effect has now been observed in a spin liquid state, a zero-field anomalous Hall effect has hitherto not been reported for antiferromagnets. Here we report empirical evidence for a large anomalous Hall effect in an antiferromagnet that has vanishingly small magnetization. In particular, we find that Mn3Sn, an antiferromagnet that has a non-collinear 120-degree spin order, exhibits a large anomalous Hall conductivity of around 20 per ohm per centimetre at room temperature and more than 100 per ohm per centimetre at low temperatures, reaching the same order of magnitude as in ferromagnetic metals. Notably, the chiral antiferromagnetic state has a very weak and soft ferromagnetic moment of about 0.002 Bohr magnetons per Mn atom (refs 10, 12), allowing us to switch the sign of the Hall effect with a small magnetic field of around a few hundred oersted. This soft response of the large anomalous Hall effect could be useful for various applications including spintronics-for example, to develop a memory device that produces almost no perturbing stray fields.
Article
We analyze the dynamics of a domain wall in an easy-axis antiferromagnet driven by circularly polarized magnons. Magnons pass through a stationary domain wall without reflection and thus exert no force on it. However, they reverse their spin upon transmission, thereby transferring two quanta of angular momentum to the domain wall and causing it to precess. A precessing domain wall partially reflects magnons back to the source. The reflection of spin waves creates a previously identified reactive force. We point out a second mechanism of propulsion, which we term redshift: magnons passing through a precessing domain wall lower their frequency by twice the angular velocity of the domain wall; the concomitant reduction of magnons' linear momentum indicates momentum transfer to the domain wall. We solve the equations of motion for spin waves in the background of a uniformly precessing domain wall with the aid of supersymmetric quantum mechanics and compute the net force and torque applied by magnons to the domain wall. Redshift is the dominant mechanism of propulsion at low spin-wave intensities; reflection dominates at higher intensities. We derive a set of coupled algebraic equations to determine the linear velocity and angular frequency of the domain wall in a steady state. The theory agrees well with numerical micromagnetic simulations.
Article
The continuum field theory describing the low-energy dynamics of the large-spin one-dimensional Heisenberg antiferromagnet is found to be the O(3) nonlinear sigma model. When weak easy-axis anisotropy is present, soliton solutions of the equations of motion are obtained and semiclassically quantized. Integer and half-integer spin systems are distinguished.
Article
The up-to-date theoretical and experimental data on the fairly-high-velocity motion of a solitary domain wall in a weak ferromagnet are reviewed. The experimental data pertain to orthorhombic and uniaxial crystals of the orthoferrite and iron-borate types. The theoretical analysis demonstrates the great generality of the nonlinear dynamics of walls in weak ferromagnets of different symmetries. The necessary data on the magnetic structure and the magneto-optical properties of weak ferromagnets are presented. The experimental techniques of investigating domain-wall motion, including non-steady-state motion, are described. Mobility and limiting-wall-velocity data are given, and a quantitative theoretical description of these experiments is presented. The magnetoelastic anomalies observed experimentally in the case when the wall velocity is close to the velocity of sound are described. A theory of this phenomenon is given which is based on the notion of Cherenkov emission of sound by a moving wall. The results of the investigation of the non-steady-state and nonlinear domain-wall motion that arises due to an instability of the wall front are presented, and theoretical models are proposed for such motion.