Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Non-Hamiltonian dynamics of indirectly coupled classical impurity spins

Simon Michel1 and Michael Potthoff1, 2


1
I. Institute of Theoretical Physics, Department of Physics,
University of Hamburg, Jungiusstraße 9, 20355 Hamburg, Germany
2
The Hamburg Centre for Ultrafast Imaging, Luruper Chaussee 149, 22761 Hamburg, Germany
We discuss the emergence of an effective low-energy theory for the real-time dynamics of two clas-
sical impurity spins within the framework of a prototypical and purely classical model of indirect
magnetic exchange: Two classical impurity spins are embedded in a host system which consists of
a finite number of classical spins localized on the sites of a lattice and interacting via a nearest-
arXiv:2009.06296v1 [cond-mat.mes-hall] 14 Sep 2020

neighbor Heisenberg exchange. An effective low-energy theory for the slow impurity-spin dynamics
is derived for the regime, where the local exchange coupling between impurity and host spins is
weak. To this end we apply the recently developed adiabatic spin dynamics (ASD) theory. Besides
the Hamiltonian-like classical spin torques, the ASD additionally accounts for a novel topological
spin torque that originates as a holonomy effect in the close-to-adiabatic-dynamics regime. It is
shown that the effective low-energy precession dynamics cannot be derived from an effective Hamil-
ton function and is characterized by a non-vanishing precession frequency even if the initial state
deviates only slightly from a ground state. The effective theory is compared to the fully numerical
solution of the equations of motion for the whole system of impurity and host spins to identify
the parameter regime where the adiabatic effective theory applies. Effective theories beyond the
adiabatic approximation must necessarily include dynamic host degrees of freedom and go beyond
the idea of a simple indirect magnetic exchange. We discuss an example of a generalized constrained
spin dynamics which does improve the description but also fails for certain geometrical setups.

I. INTRODUCTION the derivation of simple effective models is only possible


if there is a clear separation of energy scales.
The coupling between two magnetic moments can be Effective low-energy exchange couplings, like the
a so-called direct coupling, such as the usually short- RKKY interaction, are also employed to predict the real-
ranged quantum Heisenberg exchange interaction or the time spin dynamics in atomistic spin-dynamics theories
long-ranged classical dipole interaction, or an indirect [18]. This is justified, for instance, if only the impurity
coupling [1–3]. All indirect coupling mechanisms, e.g., spin degrees of freedom are driven out of equilibrium so
Anderson’s superexchange [4], double exchange [5, 6], that one stays in the low-energy sector. In other words,
the Ruderman-Kittel-Kasuya-Yosida (RKKY) interac- effective low-energy magnetic couplings also govern the
tion [7–9], or more exotic mechanisms [10, 11], have in real-time dynamics, if the dynamics of the impurity spins
common that they are derived perturbatively. They rep- is slow compared to the fast electron dynamics and if
resent effective interactions between the magnetic mo- only the former are excited initially. Generally, this argu-
ments or spins, generically of the form Jeff S 1 S 2 , where ment exploits that a separation of energy scales obviously
the effective interaction strength Jeff is typically more translates into a separation of time scales.
than an order of magnitude smaller than the typical en- A sufficiently weak coupling K not only leads to a sep-
ergy scales of the host, in which the spins S 1 and S 2 are aration of energy or time scales but also implies that
embedded. linear-response theory applies, i.e., in first-order-in-K
In the RKKY case, for example, two impurity spins are time-dependent perturbation theory [19–21]. Apart from
embedded in an electronic host system, typically a metal- setups which intrinsically prepare non-equilibrium states
lic Fermi liquid. To avoid complications due to Kondo [22, 23], such as transport through nano-structures cou-
effect [12] and its intertwining with the RKKY coupling pled to leads, linear-response theory predicts that effec-
[13–17], the impurity spins are represented by classical tive impurity-spin couplings are ground-state properties
spin vectors S 1 and S 2 . If the local exchange coupling of the host. In the RKKY case, it is the ground-state
K of the impurity spins to the local magnetic moments magnetic susceptibility χ12 (ω = 0) that determines the
of the electron system is weak as compared to the en- RKKY effective interaction. Besides this, full linear-
ergy scales of the host, e.g., the Fermi energy, one may response theory and ground-state response functions also
use standard perturbation theory to derive the effective describe other effects, such as Gilbert damping or inertia
RKKY Hamilton function Heff (S 1 , S 2 ) = JRKKY S 1 S 2 . effects [20, 21, 24–27], but those come at higher order in
The RKKY interaction JRKKY = K 2 χ12 (ω = 0) is given an expansion in the typical memory time scale and can
in terms of the nonlocal retarded static (zero-frequency) thus be classified as being of secondary importance.
magnetic susceptibility χ12 (ω = 0), which is an oscilla- Closely related to linear-response theory is the idea
tory and decaying function of the inter-impurity distance. that the state of the host system, at any instant of time t,
The condition Heff (S 1 , S 2 ) = min then provides us with is the ground state for the given configuration of the clas-
the impurity-spin ground state configuration. Obviously, sical impurity spins (S 1 (t), S 2 (t)) at this time. This is
2

full dynamics effective determined by Heff (S 1 , S 2 ), as the main paradigm sug-


impurity-spin dynamics
gests, we get the following equations of motion: Ṡ r =
S1 S2 ∂Heff /∂S r × S r for r = 1, 2, i.e.,
?

impurity spins
Ṡ 1 = Kf 0 (S 1 S 2 ) S 2 × S 1 , Ṡ 2 = Kf 0 (S 1 S 2 ) S 1 × S 2 ,
si (1)
host spins where f 0 is the derivative of f . One easily sees that the
scalar product S 1 S 2 is a constant of motion. Hence, the
FIG. 1. Left: Two classical impurity spins S 1 and S 2 weakly impurity-spin dynamics could equivalently be deduced
coupled to a system of classical host spins si localized at the from an effective Hamiltonian of the form Heff (S 1 , S 2 ) =
sites i of some lattice. Right: Which type of coupling governs Jeff S 1 S 2 with effective interaction Jeff = Kf 0 (S 1 S 2 ).
the resulting effective impurity-spin dynamics? We call this the naive approach.
Even in the limit K  J, the naive approach is shown
to fail in many cases, depending on the system geome-
reminiscent of the Born-Oppenheimer approach in molec- try. This is worth mentioning since the approach is very
ular dynamics with the nuclei treated classically [28]. tempting and, furthermore, the reason for its failure is
Adiabatic dynamics represents another consequence of very interesting and instructive: The pitfall is that the
the weakness of K and the resulting separation of time consequences of the assumption that the motion can be
scales. In case of a host system with a gapped electronic described as adiabatic are not taken seriously. Assum-
structure, the adiabatic theorem rigorously enforces per- ing that the host system is at time t in its momentary
fect adiabatic dynamics. In other cases, it is expected ground state corresponding to the impurity-spin config-
to represent an excellent approximation, which is moti- uration (S 1 (t), S 2 (t)) means that the state of the whole
vated physically by the idea that the host state should system (impurity and host spins) lives in a very small ac-
be close to the momentary ground state, if the typical cessible configuration space, parameterized by a product
relaxation times of the host dynamics are much shorter of two Bloch spheres (S 1 , S 2 ) ∈ S2 × S2 . This may lead
then the time scale on which the impurity-spin dynamics to holonomy effects [29, 30].
takes place. In particular, as we have shown in a recent paper [31],
With the present paper we reconsider this paradigm by this leads to the appearance of an additional topological
studying an even simpler problem: We still focus on two spin torque. The topological spin torque is given in terms
classical impurity spins but replace the electronic host of a topological charge density which, when integrated,
system by a system that also consists of classical spins. takes quantized values only. In Ref. [31] we have worked
This setup is illustrated in Fig. 1. The host-spin system out the general adiabatic spin dynamics (ASD) theory
is given by a classical Heisenberg model with nearest- for classical spin systems. The application of ASD to the
neighbor interaction J between host spins si that are case of a single impurity spin (R = 1), coupled to a host-
localized on the sites i of some lattice. In addition, there spin environment and subjected to a local magnetic field
are two impurity spins coupled to two host spins at sites has shown that the novel topological spin torque leads
i1 and i2 via a local Heisenberg interaction K. We as- to an anomalous precession frequency. In the present
sume a bipartite lattice, such that the K = 0 host-system paper we work out the ASD for the case of R = 2 impu-
ground states are easily found, and we assume a separa- rity spins and analyze the impact of the topological spin
tion of energy scales in the form |K|  |J|. torque on the time-dependent indirect exchange. For the
Formally, the static problem is then treated easily: The two-spin case one may expect a simple precessional dy-
total energy for a given impurity-spin configuration is namics, similar to Eq. (1), but possibly with a renormal-
E0 (S 1 , S 2 ) = min H({si }, S 1 , S 2 ), where the minimiza- ized frequency. The goal of the present paper is to answer
tion over all host-spin configurations {si } becomes trivial this question, to derive, if possible, the correct effective
in the limit |K|  |J|. Therewith we have an effec- Hamilton function, and to check the applicability of ASD
tive Hamiltonian Heff (S 1 , S 2 ) = E0 (S 1 , S 2 ) which, for theory.
an SO(3) spin-symmetric situation must have the form The rest of the paper is organized as follows: The next
Heff (S 1 , S 2 ) = Kf (S 1 · S 2 ). Here, one should note section introduces the model, some basic notations and
that, opposed to the RKKY setup, our model system is the fundamental equations of motion. Sec. III briefly re-
equipped with essentially a single model parameter K/J views the general ASD theory for R impurity spins cou-
only, such that in the weak-coupling limit |K|  |J| the pled to a classical host spin system. The ASD is spin
strength of the effective exchange must scale with K, dynamics subject to a formal constraint that enforces
while the scalar (smooth) function f : R → R depends adiabaticity. We have to carefully specify this constraint.
on the system geometry only. We will formally derive This is done in Sec. IV and used in Sec. V to set up the
Heff (S 1 , S 2 ) in the body of the paper. effective Hamiltonian and to discuss the resulting naive
Our main interest, however, is focussed on the emer- impurity-spin dynamics. In Sec. VI we then compute the
gent real-time adiabatic dynamics in case of time-scale topological spin torque, and we work out the implications
separation 1/|K|  1/|J|. Assuming that this is fully in Sec. VII. The predictions of the ASD, and of the naive
3

approach, can be compared with the numerical solution that the fast host spins almost instantaneously follow the
of the full set of equations of motions. This is done in motion of the slow impurity spins. In this adiabatic limit,
Sec. VIII. We discuss the parameter regimes, where the one can expect a strong conceptual simplification, pro-
impurity-spin dynamics is close to adiabatic. In Sec. IX viding us with an effective theory for the slow degrees of
we finally discuss an approach which goes beyond the adi- freedom only.
abatic approximation and beyond an effective two-spin
dynamics. Conclusions are given in Sec. X.
III. ADIABATIC SPIN DYNAMICS THEORY

II. CLASSICAL SPIN MODEL Using the notation n(t) ≡ (..., ni (t), ...) and m(t) ≡
(..., mr (t), ...) to characterize the configurations of the
We consider a set of R impurity spins embedded in fast and of the slow spins at a time t by the respective
a lattice of L host spins. The host system consists of unit vectors, one can state that the time evolution is
classical spins si = sni of length s and directions given adiabatic, if, at any instant of time t, the configuration
by unit vectors ni = si /s. They are localized at the of the fast spins n(t) is the ground-state configuration
sites i = 1, ..., L of a D-dimensional lattice, and spins for the present configuration m(t) of the slow spins at
si and si0 interact via an antiferromagnetic Heisenberg time t:
exchange Jii0 . The characteristic time scale of the host
n(t) = n0 (m(t)) . (4)
spin system is set by 1/J (we choose units with ~ ≡ 1).
Impurity spins are given by classical vectors S r for We expect adiabatic spin dynamics to be realized for
r = 1, ..., R, and each impurity spin is written in the K  J (and Br  J). The condition n = n0 (m)
form S r = Smr with lengths S = |S r | and unit vectors specifies a hyper surface {n = n0 (m) | m arbitrary} in
mr = S r /S. We will focus on the case of R = 2 impurity n-space, see Eq. (4), i.e., in the product of Bloch spheres,
QL
spins but develop the theory for the general case of an n ∈ i=1 S2 . Upon approaching the weak-coupling limit
arbitrary number R. The impurity spins are locally ex- in parameter space, the fast-spin configuration will be
change coupled to the host spins at sites i1 , ..., iR of the more and more constrained to this hyper surface. This
lattice, and the strength of the local antiferromagnetic means that a strongly simplified description, i.e., adia-
(“Kondo”) coupling is denoted as K. batic spin dynamics (ASD), should be possible in this
A particular state of this classical spin model limit.
is specified by a configuration (s(t), S(t)) ≡ Using the constraint (4), one can define an effective
(s1 (t), ..., sL (t), S 1 (t), ..., S R (t)) of host and impu- Hamilton function,
rity spins at a time t. Its time evolution is governed by
the Hamilton function H(s, S): H(s, S) 7→ H(sn0 (m), Sm) ≡ Heff (m) , (5)

L R R
which depends on the slow-spin degrees of freedom only.
1 X X X It is therefore tempting to derive the slow-spin dynamics
H= Jii0 si si0 + K sir S r − Sr Br . (2)
2 0 solely from this effective Hamiltonian via the R remaining
i,i =1 r=1 r=1
differential equations
Generically, we have Jii0 = Ji0 i = J between nearest ∂Heff (m)
neighbors i and i0 only. We have also added a term de- S ṁr = × mr (6)
∂mr
scribing a local magnetic field B r coupling to the impu-
rity spin S r . In most cases, however, we set B r = 0. for m(t), while n(t) can be obtained from Eq. (4). This
The Hamilton function leads to the following coupled set constitutes an approach which we will refer to as the
of non-linear ordinary differential equations of motion, naive theory of adiabatic dynamics.
We have recently shown that the naive approach may
∂H ∂H lead to incorrect results [31]. To eliminate the fast spin
ṡi = × si , Ṡ r = × Sr , (3)
∂si ∂S r degrees of freedom correctly, one must rather switch
to a Lagrangian formulation and employ the general
which determine the time evolution of an arbitrary given R
action principle, δ dt L(n, ṅ, m, ṁ) = 0. In this
initial spin configuration. Note that the lengths of si and
framework one may safely make use the holonomic con-
of S r are conserved. This allows us to absorb constants,
straints (4) to eliminate the host degrees of freedom
like gyromagnetic ratios, in si and Sr . The model (2) can
and to set up an effective Lagrangian, Leff (m, ṁ) ≡
be seen as the classical isotropic multi-impurity Kondo-
L(n0 (m), (d/dt)n0 (m), m, ṁ), for the slow-spin degrees
necklace model [13] or simply as a classical Heisenberg
of freedom only. In terms
R of the effective Lagrangian, the
model with a special multi-impurity geometry. There is
action principle reads δ dt Leff = 0, where δ is variation
no direct coupling of the impurity spins but an indirect
of the slow-spin configuration m only. This provides us
coupling is mediated via the host.
with the ASD equations of motion for the slow spins mj :
We will study the model in the limit of weak local cou-
pling K  J. In this parameter regime the system ex- ∂Heff (m)
hibits two very different time scales, K −1 and J −1 , such S ṁr = × mr + T r × mr . (7)
∂mr
4

As compared to the “naive” adiabatic theory, Eq. (6), impurity spin configuration m, and on the coupling con-
there is an additional term due to a field stants Jii0 , the host ground state can strongly differ from
X Rn0 and must be determined numerically in general. For
Trµ = Trµ (m, ṁ) = Ωrµ,sν (m)ṁsν , (8) weak Kondo coupling K  J, however, the impurity
sν spins basically act as infinitesimally weak external local
where s = 1, ..., R and µ, ν ∈ {x, y, z}. Here, magnetic fields, which merely break the host SO(3) in-
variance and typically favor exactly one state out of the
K = 0 ground-state manifold, without further disturbing
X
Ωrµ,sν (m) = 4π se(i)
rµ,sν (m) , (9)
i the spin configuration of that state. At the same time,
K  J just specifies a limit where the ASD is expected
with to apply – as will be discussed later by comparing with
1 ∂n0,i (m) ∂n0,i (m) results from the numerical solution of the full set of equa-
e(i)
rµ,sν (m) = × · n0,i (m) (10) tions of motion given by Eq (3). Hence, we will focus on
4π ∂mrµ ∂msν
the weak-coupling limit.
is a rank-2 tensor for each pair of impurities r, s. It de- The host ground state for K 6= 0 and K  J
scribes certain topological properties of the ground state is obtained by minimization of the Hamilton function
of the fast-spin subsystem on the hyper surface specified H(n, m) ≡ H(sn, Sm), where ni = Rn0,i , for given
by Eq. (4). In fact, each tensor element for fixed r, s de- fixed m, and with respect to all R ∈ SO(3). Note that
fines a topological charge density, which becomes a quan- this minimization problem is much simpler as compared
tized homotopy invariant, namely a topological winding to a high-dimensional minimization with respect to arbi-
(i) (i)
number ers with a quantized value ers ∈ Z when inte- trary host-spin configurations, which would be necessary
grated. There is a close analogy to the concept of the beyond the weak-coupling limit. For the minimization,
skyrmion density [32–34]. Here, however, the skyrmions we can also disregard the magnetic field term in Eq. (2)
live on a product of Bloch spheres rather than in Eu- as this is independent of n and thus of R. Furthermore,
clidean space. Note that the resulting topological spin due to the invariance of the inner product si · si0 under
torque T r × mr in Eq. (7) involves the time derivative spin rotations, also the Heisenberg term in H is constant.
ṁs . It nevertheless respects total-energy conservation, Hence, it is sufficient to focus on the Kondo term only:
unlike a Gilbert damping term [35, 36]. Details of the R
derivation of Eq. (7) and its interpretation are given in !
X
K (Rn0,ir ) · mr = min. . (11)
Ref. 31. r=1

We make use of the fact that R ∈ SO(3) has the form


IV. HOLONOMIC CONSTRAINT R = Ra (ϕ) = exp(ϕLa), where L with components
Lα ∈ so(3) are the real and antisymmetric generators of
For the computation of the topological spin torque, SO(3), and where the unit vector a specifies the rotation
the topological charge density (10) is required. This is a axis and ϕ the rotation angle. Let us now assume that
ground-state property of the host system. Similar to a n0 is the desired ground-state configuration for given m.
two-point response function, it depends on two fixed posi- This implies that the Hamilton function reaches its min-
tions ir and is , i.e., on the two sites at which the r-th and imum at ϕ = 0 for any rotation axis a. With the general
the s-th impurity spin couple to the host. For a concrete relation
calculation, we need the explicit form of the constraint
(4). This means to find the ground-state configuration ∂
exp(ϕLa) (·) = a × (·) , (12)
of the host spins for an arbitrarily given configuration ∂ϕ ϕ=0

m = (m1 , ..., mR ) of all impurity spins.


we thus find the following necessary condition for the
We start by considering the ground state at K = 0. In
minimum:
this case, the host subsystem is describedPby a classical
Heisenberg Hamiltonian Hhost (n) = 21 s2 i,i0 Jii0 ni ni0 , R
X
see Eq. (2). For any choice of the matrix of coupling con- a × n0,ir · mr = 0 . (13)
stants Jii0 , the Hamiltonian is invariant under SO(3) spin r=1
rotations such that the ground-state manifold is highly
degenerate. We pick an arbitrary ground-state config- Since this must be satisfied for rotations around an arbi-
uration n0 = (n0,1 , ..., n0,L ) which minimizes Hhost (n), trary axis a, we get
i.e., Hhost (n0 ) = E0 = min., where E0 is the ground- R
state energy. The SO(3) symmetry implies that Rn0 ≡
X
n0,ir × mr = 0 . (14)
(Rn0,1 , ..., Rn0,L ) for any rotation matrix R ∈ SO(3) is r=1
a ground state as well: Hhost (Rn0 ) = E0 .
At finite K, the degeneracy of the ground-state energy This means that the total torque on the impurity spins
is lifted. Depending on the strength of K, on the given must vanish for the ground-state configuration n0 .
5

We now specialize to the case of host spins on a bipar- host-spin degrees of freedom, which yields Heff (m) =
tite lattice with nearest-neighbor interactions Jii0 , where H(n0 (m), m). Concretely, for the considered case of a
the ground-state spin structure is collinear, i.e., we have collinear host-spin configuration, Eq. (15), we have

n0,i = zi η (15) R
X R
X
Heff (m) = E0 + KsS zir ηmr − S mr B r , (20)
with zi ∈ {+1, −1} and with a unit vector η to be de- r=1 r=1
termined from Eq. (14). The host spin structure is in-
PL
variant under the simultaneous transformation zi 7→ −zi where E0 = (s2 /2) i,i0 =1 Jii0 zi zi0 is the m-independent
and η 7→ −η. Hence, we can choose the overall sign ground-state energy of the host-spin Hamiltonian
of (z1 , ..., zL ) as is convenient, e.g., such that zi1 = +1. Hhost (n). With η = zK m0 /m0 , see Eq. (18), we find
Physical properties do not depend on this choice. To fix
η, we insert Eq. (15) in the equilibrium condition (14). R
X
This yields Heff (m) = E0 + zK KsSm0 − S mr B r . (21)
r=1
R
X
η× zir mr = 0 . (16) Note that the total energy is at a minimum for zK K =
r=1 −|K| < 0. This justifies the above choice zK = −signK.
As mentioned before, it is very tempting to derive the
We define the (staggered) sum of the impurity-spin unit slow-spin dynamics solely from this effective Hamilto-
vectors mr nian, see Eq. (6) and the related discussion. This con-
X stitutes the naive adiabatic theory. The corresponding
m0 = zir mr (17) equations of motion of the naive theory are easily de-
r
rived. We note that ∂m0 /∂mr = zir m0 /m0 and find
and m0 = |m0 |. With this we have η × m0 = 0 and thus
∂Heff (m) m0
m0 = −|K|sSzir − SB r . (22)
η = zK . (18) ∂mr m0
m0
This yields
The total energy is minimized for the sign zK =
−signK = −K/|K| as argued below, see Eq. (21). With
 
1 ∂Heff (m) m0
Eq. (15) the explicit constraint Eq. (4) finally reads as ṁr = × mr = −|K|szir − B r × mr
S ∂mr m0
P (23)
zi mr
n0,i (m) = zi zK Pr r . (19) or
| r zir mr |
|K|s X
ṁr = − zi zi mr0 × mr − B r × mr . (24)
PNote that the function n0,i (m) is singular for m0 = m0 r 0 r0
r
r zir mr = 0. The condition m0 = 0 specifies a
submanifold embedded in the full configuration space. This is a comparatively simple nonlinear system of R
Though this has zero measure, we have to keep this in differential equations of motion for the R impurity spins.
mind and must exclude trajectories m(t) crossing the The naive adiabatic theory is in fact conceptually in-
submanifold. More importantly, one cannot choose ini- correct, as the constraint is directly used to simplify the
tial conditions with m0 (t = 0) = 0 within the ASD the- Hamiltonian, which is generally not justified. Before we
ory. Consider the case of two impurity spins R = 2 proceed with the (conceptually correct) adiabatic spin
as an example. Since zi1 = +1 is already fixed (see dynamics (ASD), however, let us discuss some special
above), there are two cases to be taken into account: cases and consequences of the naive theory.
zi2 = ±1. For zi2 = −zi1 = −1, the parallel configura-
For B r = 0, the effective slow impurity-spin dynamics
tion m1 = m2 is singular, while for zi2 = zi1 = +1 the
takes place on the time scale set by 1/K, as is obvious
antiparallel configuration m1 = −m2 is singular must
from Eq. (24). One also easily verifies that the total
P en-
be excluded. The physical meaning of the singularity
ergy Heff and the total impurity spin Smtot ≡ S r mr ,
becomes obvious at a later stage (see Sec. VIII).
and also m0 are conserved quantities. Multiplying Eq.
(24) with zir and summing over r, we see that m0 pre-
V. EFFECTIVE HAMILTONIAN AND NAIVE
cesses around mtot :
ADIABATIC THEORY
|K|s
ṁ0 = mtot × m0 (25)
m0
Having the explicit form of the constraint, Eq. (19),
at hand, we proceed with derivation of the effective with frequency
Hamiltonian Heff (m) for arbitrary R. This is obtained
from H(n, m) by using the constraint to eliminate the ωp = |K|smtot /m0 . (26)
6

For two impurity spins, R = 2, we have mtot = m1 + expanding, exploiting the condition (14), and carrying
m2 , and with mtot the angle ϑ that is enclosed by m1 out the sums over ρ, σ, we find
and m2 is conserved as well. In this case, as is directly
1 X 1
seen from Eq. (24), m1 and m2 precess with the same (i)
erµ,sν (m) = zi zK zir zis εµντ m0τ 3 . (31)
frequency ωp around mtot . 4π τ
m0
The precession frequency decisively depends on the ge-
ometry. Let us assume that zi1 = −zi2 (note that we Summation over i yields the tensor field defined in Eq.
fixed zi1 = +1). This happens to be the case, e.g., for an (9):
antiferromagnetic ground-state configuration of the host X 1
spins, n0,i = ±(−1)i η, if the distance i1 − i2 between the Ωrµ,sν (m) = zir zis zK s∆ εµντ m0τ . (32)
m30
√ is odd. Then m0 = m
two impurity√ spins − m2 and τ
√1 √
thus m0 = 2 1 − cos ϑ. With mtot = 2 1 + cos ϑ Here, we have defined
we find
X
∆≡ zi . (33)
r
1 + cos ϑ
ωp = |K|s = |K|s cot(ϑ/2) . (27) i
1 − cos ϑ
For zi1 = +zi2 = +1, on the other hand, we have m0 = We see that, in case of a collinear host-spin ground
mtot by definition, and Eq. (25) is useless. The same state, Eq. (15), a nonzero field Ωrµ,sν (m) requires a
holds for arbitrary R and zir = +1 for all r, such that ground Pstate withP a finite total host-spin magnetization
again m0 = mtot . Going back to Eq. (24), we have ntot = i n0,i = i zi η = ∆ η. Its modulus ntot = ∆ is
ṁr = −(|K|s/mtot )mtot × mr in this case, i.e., each nonzero, for instance, in case of ferromagnetic exchange
impurity spin precesses with frequency couplings J < 0, or in case of antiferromagnetic couplings
J > 0 when L is odd.
ωp = |K|s (28) Inserting the result for the tensor field into Eq. (8)
yields:
around the total impurity spin. X X
Let us finally emphasize that the naive theory violates Tr = Trµ (m, ṁ)eµ = Ωrµ,sν (m)ṁsν eµ
total spin conservation (for B r = 0). The total impu- µ sµν
rity spin S tot = Smtot is a constant of
Pmotion asPstated s∆ X X
above. The total host spin stot = i si = s i zi η,
= zir 3 zis εµντ m0τ ṁsν eµ
m0 s
however, has a nontrivial precession
P dynamics in the case µντ
zi1 = −zi2 and for ∆ ≡ i zi 6= 0 or, equivalently, for s∆
stot = s∆η 6= 0. Hence, S tot + stot 6= const. This short- = zir zK ṁ0 × m0 , (34)
m30
coming is cured by the adiabatic spin dynamics theory,
see below. so that the additional topological spin torque on the
right-hand side of Eq. (7) reads as:
s∆
VI. TOPOLOGICAL CHARGE DENSITY AND T r × mr = zir zK (ṁ0 × m0 ) × mr . (35)
SPIN TORQUE m30
Note that the torque is independent of the coupling
In addition to the term on the right-hand side of Eq. parameters and depends on the system geometry only.
(24), there is an additional contribution to the equations Combining this with Eq. (7) and Eq. (24), we arrive at
of motion of the ASD resulting from the topological spin the ASD equations of motion:
torque, see Eq. (7). To derive this contribution, we must  
calculate the topological charge density Eq. (10) from the |K|s s∆
ṁr = −zir m0 + zir zK ṁ0 ×m0 − B r × mr .
explicit form of the constraint, Eq. (19). m0 Sm30
This is done straighforwardly. First, we have: (36)
  The first term is the same as in the naive approach, the
∂n0,iµ (m) 1 m0µ m0µ0 second one is due to the topological spin torque.
= zi zK zir0 δµµ0 − , (29)
∂mr0 µ0 m0 m20
P
where the m dependence is due to m0 = r zir mr . VII. ADIABATIC SPIN DYNAMICS
Inserting this expression in Eq. (10),
 
(i) 1 X 1 m0ρ m0µ For the discussion of the equations of motion of adia-
erµ,sν (m) = ερστ zi zK zir δρµ −
4π ρστ m0 m20 batic spin dynamics, see Eq. (36), we will consider sev-
  eral cases. Let us first check the case R = 1, K > 0,
1 m0σ m0ν m0τ J > 0, i1 = 1, and odd L, i.e., there is a single impu-
× zi zK zis δσν − zi zK ,
m0 m20 m0 rity spin only, m ≡ m1 , which couples antiferromagnet-
(30) ically at the first site of an antiferromagnetic host with
7

an odd number of sites. We have m0 = mtot = m and With m0 ṁ0 = 0 we can also simplify the double cross
∆ = +1. The naive equation of motion, Eq. (24), reads product in Eq. (38):
ṁ = m × B, and thus predicts precession around the ex-
ternal magnetic field with Larmor frequendy ωp = B. In- s∆
ṁtot = −zK ṁ0 . (41)
cluding the additional topological spin torque, however, Sm0
we find S ṁ = Sm×B −s(ṁ×m)×m = Sm×B +sṁ. Multiplying with mtot and using ṁ0 mtot = 0, we see
This can be written in the form of the Landau-Lifschitz that the norm of mtot is conserved. Furthermore, ṁ0
equation but leads to an anomalous precession frequency on the right-hand side can be eliminated using Eq. (40),
B such that we are finally left with:
ωp = . (37)
1 − s/S s∆ 1 |K|s
ṁtot = zK s∆
m0 × mtot .
Sm0 1 + zK Sm 3 m0 mtot m0
This is precisely the result derived in Ref. [31]. 0

Next we discuss constants of motion for the general (42)


case (but we assume B r = 0). We start by checking Summing up, for B r = 0, we have, besides energy and
that the equation of motion respects the conservation of total spin conservation, the following conserved quanti-
|mr | = 1. This is the case (also for finite B r ) since ties:
mr ṁr = 0, see Eqs. (7) or (36).
m0 = const. , mtot = const. , m0 mtot = const.
Total energy energy conservation is ensured on gen-
(43)
eral grounds as the equation of motion is derived within
Furthermore, there are two coupled nonlinear ordinary
the standard Lagrange formalism and employing a scle-
differential equations, Eq. (40) and Eq. (42), for m0 and
ronomic holonomic constraint. This has also been proven
mtot .
explicitly and discussed in detail in Ref. [31].
There is a link to the naive adiabatic theory, namely
Conservation of the total
P spin, i.e., the sum of the total in the topologically trivial case where ∆ = 0, i.e., where
impurity P spin S tot = S r mr and the total host spin the topological spin torque vanishes. Here, the total
stot = s i n0,i (m), can be shown for the case of an impurity spin is conserved, ṁtot = 0, and according
SO(3) symmetric effective Hamiltonian Heff (m). This is to Eq. (40), m0 precesses around mtot with frequency
detailed in Appendix A. ωp = |K|smtot /m0 . This precisely recovers the results of
PConservation of the total impurity spin mtot = the naive theory, see Eq. (26).
r zir mr is not expected in general. Summing both In the nontrivial case for ∆ 6= 0, an analytical solution
sides of Eq. (36) over r = 1, ..., R we get (for B r = 0): of Eq. (40) and Eq. (42) is obtained easily. To this end,
s∆ we rewrite the equations as
ṁtot = zK (ṁ0 × m0 ) × m0 , (38)
Sm30 ṁ0 = c0 mtot × m0 , ṁtot = ctot m0 × mtot , (44)
which is nonzero for a finite topological spin torque, i.e., with constants c0 and ctot , and employ a scaling trans-
for ∆ 6= 0. Note that this immediately implies ṁtot m0 = formation to new variables m f0 = α0 m0 and m ftot =
0. αtot mtot such that the prefactors in the transformed
Summing both sides of Eq. (36) over r after multiplying equations of motion are equal (this is the case for zK ∆ >
with zir , we can derive an equation
P for the “staggered” 0) or differ by a sign only (zK ∆ < 0). Details are given
sum of the impurity spins m0 = r zir mr . For B r = 0 in Appendix B. We find that both, m0 and mtot , are
we get: precessing around the conserved total spin stot + S tot ,
and the corresponding precession frequency is:
|K|s s∆
ṁ0 = − m0 × mtot + zK (ṁ0 × m0 ) × mtot . q
m0 Sm30 ωp = c2tot m20 + c20 m2tot ± 2 |c0 ctot | m0 mtot . (45)
(39)
We immediately have ṁ0 mtot = 0. Together with the With the solutions m0 , mtot at hand, the dynamics of
above relation ṁtot m0 = 0, this implies that the in- the individual impurity moments mr can be obtained
ner product m0 mtot is conserved. Furthermore, we have from a numerical solution of Eq. (36) for each mr sep-
(ṁ0 × m0 ) × mtot = −ṁ0 (mtot m0 ) and therewith we arately. This situation is different from but reminiscent
can convert Eq. (39) into an explicit differential equation: of gyroscope theory, since mr precesses around a mo-
mentary axis specified by m0 and ṁ0 , while m0 itself is
1 |K|s precessing around an axis fixed in space.
ṁ0 = s∆
mtot × m0 . (40)
1 + zK Sm 3 m0 mtot m0 Finally, we consider the special case of two impurity
0
spins R = 2 and vanishing external fields B r = 0. Here,
Multiplying both sides of the equation with m0 , yields it is sufficient to analyze the two coupled equations Eq.
m0 ṁ0 = 0 and hence m0 = const (if B r = 0). Hence, (40) and Eq. (42), rather than reverting to Eq. (36) again,
the prefactor of m0 × mtot in Eq. (40) is a constant of since the dynamics of m1 and m2 is fully determined via
motion. mtot = m1 + m2 and m0 = zi1 m1 + zi2 m2 .
8

For the case zi1 = zi2 = 1, we have m0 = mtot , such impurity spins
S1 S1
that Eq. (40) reduces to Eq. (42). Eq. (42) implies that S2 ⌘
the total impurity spin is conserved, ṁtot = 0, and thus K
S2
the topological spin torque vanishes. This is the topolog- J #
m0
ically trivial case. Both impurity spins precess around
host spins si
mtot with frequency ωp = |K|s, as discussed above, see
Eq. (28).
FIG. 2. System geometry: L = 5 host spins (length s = 1),
For the nontrivial case zi1 = −zi2 = 1, we have m0 = R = 2 impurity spins (S = 1) coupled to the host spins at sites
m1 − m2 . This implies m0 mtot = (m1 − m2 )(m1 + i1 = 1, i2 = 4, antiferromagnetic exchange J, K > 0, hence:
m2 ) = 0, since m1 and m2 are unit vectors, such that zi1 = +1, zi2 = −1, and ∆ = 1. Initial angle enclosed by
the last statement of Eq. (43) becomes trivial. Further, the two impurity spins: ϑ. Initially the host-spin system is in
the first and second one imply m1 m2 = cos ϑ = const. its antiferromagnetic ground state for the given impurity-spin
Therewith, the equations of motion for m0 and mtot configuration, i.e., η = −m0 /m0 .
simplify and read:
|K|s
ṁ0 = mtot × m0 (46) regime the adiabatic approximation is justified. We
m0
therefore compare the predictions of the ASD with the
and numerical solution of the full set of equations of motion
s∆ |K|s (3). For the sake of simplicity, we first pick a geometry
ṁtot = zK m0 × mtot . (47)
Sm0 m0 with L = 5 host spins and R = 2 impurity spins with
The precession frequency is: antiferromagnetic couplings J, K > 0, see Fig. 2. The
r impurity spins are coupled to the host at sites i1 = 1 and
1 s2 |∆|2 i2 = 4. Since L is odd and zi1 = −zi2 = 1, this is a
ωp = |K|s + m2tot . (48)
m0 S2 realization of the topologically nontrivial case.
Assuming that s = S and that |∆| = 1 (antiferromag- With Fig. 3 we give an example result which is charac-
netic host-spin configuration and odd L), we have (see teristic of the real-time spin dynamics if K and J are of
Appendix B): the same order of magnitude, and if the initial configura-
p r tions of impurity spins is far from the (antiferromagnetic)
1 + m2tot 1 ϑ ground state configuration. The initial host-spin config-
ωp = |K|s = |K|s 1 + 4 cos2 . uration is taken to be the ground-state configuration for
m0 2 sin /2
ϑ 2
(49) the given impurity-spin directions. As is demonstrated
This also applies to the individual impurity spins: For with Fig. 3, we find an extremely complex dynamics as it
R = 2, the dynamics of the impurity spins, m1 and m2 , is characteristic for a nonlinear classical Hamiltonian sys-
is simple: They precess around the axis specified by the tem with several degrees of freedom. For long times, the
total spin with the same frequency as m0 and mtot . trajectories cover the entire phase space that is accessible
Note that the precession frequency approaches ωp → under total energy and total spin conservation.
|K|s/2 for ϑ → π, i.e., when one approaches the global Clearly, adiabatic spin dynamics is only expected to
ground state. This must be compared with the result be realized in the weak-coupling limit K  J. Fig. 4
ωp → 0 that is obtained by the naive adiabatic theory. provides an example for the same setup and parame-
For ϑ → 0, on the other hand, i.e., for √ m0 → 0, the ters as in Fig. 3 but for K/J = 10−5 . The motion is
precession frequency diverges as ωp → 5|K|s/ϑ. This mainly precessional but there is an additional nutation
divergence originates from the above-mentioned singular- visible. This nutation effect is not captured by the ASD
ity, cf. the discussion following Eq. (19). but gets weaker and finally almost disappears when the
We would like to emphasize that already the simple initial impurity-spin configuration is chosen closer and
R = 2 case demonstrates that an effective impurity- closer to a ground-state configuration.
spin dynamics is not Hamiltonian. There is no effective If the initially enclosed angle ϑ ≈ π, the dynamics is
Hamilton function Heff (m1 , m2 ) with which the equa- even more regular. Fig. 5 displays an example, where
tions of motion (46) and (47) can be reproduced. Any ϑ = 0.95π and where the host is in the corresponding
nontrivial two-impurity-spin Hamiltonian model would ground state initially. Still, for K/J = 1, the individual
be of the form Heff (m1 , m2 ) = Kf (m1 m2 ) with some impurity spins mr for r = 1, 2 show a rather compli-
smooth real function f , which would immediately, and cated time
P evolution (not shown). The staggered sum
incorrectly, imply that m1 + m2 is conserved. m0 = r zir mr = m1 − m2 , on the other hand, is al-
ready much closer to a purely precessional motion. The
figure shows the time dependence of the azimuthal angle
VIII. NUMERICAL RESULTS ϕ, modulo 2π, of m0 with respect to the total conserved
spin stot + S tot . This azimuthal angle more or less grows
It remains to check the validity of the adiabatic spin- linearly in time but with some weak additional struc-
dynamics theory, i.e., to find out in which parameter ture superimposed. With decreasing ratio K/J, the ad-
9

tmax = 20 tmax = 100 tmax = 1000

z z z

x x x

y y y

FIG. 3. Trajectories of the two impurity spins on the Bloch sphere as obtained by numerical solution of the full set of equations
of motion (3) for K = J, ϑ = π/2 for different maximal propagation times tmax as indicated (in units of K −1 ). At time t = 0
the system is prepared as indicated by the arrows. m1 : blue, m2 : red. The host spins are in their ground-state configuration
for given impurity spins. The total spin is parallel to the z-axis.

ditional superimposed oscillations get weaker and weaker within the ASD and the full theory. Furthermore, the
and are only hardly visible when K/J = 0.01. angular velocity is by far too small or, as can be seen in
The green line in Fig. 5 shows the result of the ASD the figure, the period 2π/ωp is by far too large.
theory, which predicts a purely precessional motion and With decreasing K/J also other quantities appear to
correspondingly a linear increase of ϕ as a function of t. converge to the predictions of the ASD (not shown). We
We see that with decreasing ratio K/J, the trajectory of find, for example, that the modulus of the staggered and
m0 , obtained from the full theory, appears to converge of the total impurity spin, m0 and mtot , and the scalar
to the ASD result. Not only the additional structure di- product m0 mtot or, equivalently, the enclosed angle ϑ
minishes further and further but also the ASD prediction approach constants when K/J → 0, as is stated in Eq.
of the angular velocity dϕ/dt seems to be approached in (43). Furthermore, also the dynamics of individual im-
the K/J → 0 limit. purity spins seem to more and more approach a purely
On the contrary, the prediction of the naive adiabatic precessional motion with the same precession frequency.
theory, see Eq. (25), which is directly obtained from the There is, however, a finite residual difference between
effective Hamiltonian, Eq. (21), is completely off. The the full theory, Eq. (3), and the ASD persisting in the
approach does yield a purely precessional motion but
mistakenly around the total impurity spin mtot , which
is a constant of motion within the naive theory but not
K/J=10-2 K/J=100 naive
ASD 10-1 theory

tmax = 100

x
FIG. 5. Time dependence (in units of K −1 ) of the azimuthal
angle (mod 2π) in the precession dynamics of m0 around the
y
conserved total spin for the geometry displayed in Fig. 2 and
for ϑ = 0.95π. Numerical solution of the full set of equations
FIG. 4. The same as Fig. 3 but for K/J = 10−5 . Maximal of motion (3) for various ratios K/J as indicated: blue. Naive
propagation time tmax = 100K −1 . theory: orange. Adiabatic spin dynamics (ASD): green.
10

ASD

full theory
ASD
full theory

naive theory
naive theory

FIG. 6. Precession frequency ωp of the staggered sum m0


as a function of the coupling strength K/J at ϑ = 0.95π. FIG. 7. ϑ dependence of the precession frequency ωp as
Results obtained from the numerical solution of the full set predicted by the ASD (green), see Eq. (49), and compared to
of equations of motion, Eq. (3), compared to the predictions the numerical data obtained at K/J = 10−5 by solving the
of the ASD and of the naive adiabatic spin dynamics in the full set of equations of motion (blue points), Eq. (3), and to
limit K/J → 0. the naive theory (orange), Eq. (27).

limit K/J → 0. This is demonstrated in Fig. 6 where


the precession frequency in the real-time dynamics of the ASD. This singularity
√ leads to the divergence of the
m0 is plotted as a function of K/J. Since ϑ = 0.95π, frequency ωp → 5|K|s/ϑ for ϑ → 0, see Eq. (49). It
the impurity-spin configuration is close to a ground-state results from the fact that the ground-state host-spin con-
configuration, such that there is a frequency with dom- figuration cannot be determined unambiguously for the
inant weight in the Fourier analysis of the data. This maximally excited impurity-spin configuration, and that
frequency smoothly depends on K/J and approaches the there is two-dimensional manifold of degenerate host-spin
frequency of the almost pure precessional dynamics that configurations in this case.
remains in the limit K/J → 0. Already at K/J ≈ 10−3 Vice versa, a divergent precession frequency implies a
it approaches saturation, although at a level that differs fast impurity-spin dynamics, i.e., a violation of the cen-
from the ASD result (green arrow) by about 15%. This tral assumption of a slow, adiabatic or close-to-adiabatic
implies that even in the weak-coupling limit, the host motion. We note that this kind of inherently built-
spins do not completely adiabatically follow the impurity- in breakdown of the theory is already known from the
spin dynamics. The observation of a close-to-adiabaticity single-spin (R = 1) case with a finite external magnetic
dynamics has already been made earlier for the single- field [31], where it shows up, however, at a different point
spin (R = 1) case [21]. Turning to the naive adiabatic in parameter space, namely for s → S, see Eq. (37).
theory, see orange arrow in Fig. 6, the predicted preces-
sion frequency is again by far too small, aside from the Finally, the naive adiabatic spin-dynamics theory is
fact that the precession axis is predicted incorrectly. neither correct in the ϑ → π nor in the ϑ → 0 limit.
While the ASD theory is at least qualitatively correct In the latter case, the precession frequency diverges as
at small K/J and ϑ → π, it must break down for initial ωp → 2|K|s/ϑ.
impurity-spin configurations that are far from a ground- The discussion of the results and the conclusions also
state configuration. This is demonstrated with Fig. 7, apply to larger system sizes. This is demonstrated with
where the analytical result (49) for the ASD precession Fig. 8, where the normalized difference between the ASD
frequency is plotted against ϑ and compared to the nu- precession frequency and the precession frequency of the
merical data for a coupling strength K/J = 10−5 deep in numerical solution of the full set of Hamilton equations
the weak-coupling limit. For ϑ → π the ASD is close to of motion is plotted against ϑ for different L. Since L is
the numerical data and correctly predicts a finite nonzero odd in all cases and since the impurity spins are coupled
frequency ωp = |K|s/2 in the limit, while there is a re- to the host at i1 = 1 and i2 = L−1, we have the topolog-
maining discrepancy visible, as discussed above. ically non-trivial case at hand. We see that the difference
With decreasing ϑ and increasing parametric distance diverges for ϑ → 0, as discussed above. For ϑ → π, on the
to the ground state, however, the ASD is less reliable. other hand, the residual difference becomes larger with
This is understood easily and eventually results from the increasing L. This means that it becomes more and more
singular constraint, see Eq. (19), on the m0 = 0 man- difficult to enforce close-to-adiabatic dynamics. Obvi-
ifold. For ϑ → 0, i.e., for m1 = m2 initially, we have ously, this is due to the necessity to communicate the
m0 = 0 initially, and thus m0 = 0 at all times within relative impurity-spin configuration over large distances.
11

L=5 attempts to use the constraint (50) directly for a simpli-


L = 15 fication of the full set of equations of motion (3). The
L = 105 proper way is rather to start from the action principle
again, to set up the Lagrangian of the full theory yield-
101 ing the equations of motion (3), and to treat Eq. (50)
F)/ F

as a holonomic constraint to simplify the Lagrangian to


an effective Lagrangian Leff (η, η̇, m, ṁ) with a strongly
reduced number of degrees of freedom.
A

In Appendix D this program is carried out for an arbi-


(

100 trary function n0,i (η) without further specification. The


form of the resulting equation of motion, Eq. (D13),
∂Heff (η, m)
0= ×η+T ×η, (51)
∂η
0.0 0.2 0.4 0.6 0.8 1.0
/ turns out as quite unusual as it lacks an explicit η̇ term.
However, similar to the ASD, see Eq. (7), there is an ad-
FIG. 8. Normalized difference of the ASD precession fre- ditional topological spin-torque term resulting from the
quency ωA and the frequency ωF obtained numerically from constraint. This has the form T = η̇ × Ω, where Ω
the full solution of Eq. (3) as function of ϑ at K/J = 10−3 . is the pseudo-vector corresponding to an antisymmetric
Results for various system sizes L = 5, 15, 105. The impurity tensor Ωµν that derives from the topological charge den-
spins couple to positions i1 = 1 and i2 = L − 1.
sity [31] or magnetic vorticity [37], see Eqs. (D16) and
(D17). This topological spin torque brings the η̇ depen-
IX. BEYOND THE ADIABATIC
dency back into the theory.
APPROXIMATION Eq. (51) holds generally for constraints of the form
ni = n0,i (η). In Appendix E, we evaluate the topo-
logical charge density and the resulting topological spin
One way to improve the theory and to go beyond torque for the constraint Eq. (50) explicitly.
the adiabatic approximation is to relax the constraint P This leads
to equations of motion, which, for ∆ = i zi 6= 0, have
Eq. (19) defining the ASD. For the weak-coupling limit the familiar Hamiltonian form, cf. Eq. (E10):
K  J, it is tempting to keep the host spins tightly cou- X
pled together but to relax the demand that the host-spin ∆η̇ = SK zir mr × η ,
configuration should be given, at any instant of time, by r
the ground-state configuration for the currently present ṁr = sKzir η × mr − SB r × mr . (52)
configuration of the impurity spins. This idea can be
formalized by substituting Eq. (19) by the constraint Some general properties and conservation laws related to
these equations are discussed in Appendix E as well.
! Here, we consider the setup discussed in the previ-
ni = n0,i (η) = zi η , (50)
ous section, see Fig. 2, and compare the numerical so-
where η is a (three-component) dynamical degree of free- lution of Eqs. (52) with that of the full set of equations
dom normalized to unity, η 2 = 1. If we consider host of motion (3) and with the predictions of the ASD. For
spins on a bipartite lattice or, for the sake of simplic- R = 2 impurity spins, the constrained spin dynamics is
ity, on a one-dimensional chain of sites i = 1, ..., L that in fact more complicated. In particular, there is no sim-
are tightly bound together via a strong antiferromagnetic ple precessional motion at moderate K/J. For a com-
coupling J > 0, we have zi = (−1)i+1 , with the conven- parison with the full spin-dynamics theory, Eq. (3), we
tion z1 = +1. nevertheless concentrate on the dominant peak in the
A conceptual disadvantage of an effective spin- Fourier spectrum and the corresponding precession fre-
dynamics theory under these tight-binding constraints is quency ωp . Fig. 9 demonstrates that spin dynamics un-
that it necessarily involves (with η) dynamical host de- der the tight-binding constraint in fact substantially im-
grees of freedom, such that one will not end up with an proves the description and is reliable in the weak-coupling
effective theory of the impurity-spin degrees of freedom limit K/J  1 for all angles ϑ specifying the initial
only. Clearly, this is the price to be paid when aiming at impurity-spin configuration at time t = 0. This is op-
an improved theory beyond the ASD. On the other hand, posed to the ASD, which requires weaker couplings K/J
a formal advantage is that there is no singularity and that and which captures the full spin dynamics for angles close
no submanifold of spin configurations must be excluded, to ϑ = π only, as is shown in Fig. 10.
as compared to the ASD, cf. the discussion following Eq. The situation is completely different, however, for the
(19). case ∆ = 0, i.e., if the chain of host spins consists of
Again, one must be very careful when imposing the an even number of sites, such that for antiferromagnetic
constraint Eq. (50). In Appendix C, it is demonstrated coupling J the total host spin vanishes. Solving the equa-
that one runs into unacceptable inconsistencies, if one tions of motion (3) of the full theory for R = 2 impurity
12

100 We have analyzed the origin of this inconsistency in


Appendix F. In fact, in the case ∆ = 0, the effective
0.8 Lagrangian of the constrained spin-dynamics theory is
singular. This can be made explicit with a proper gauge
transformation after which Leff (η, η̇, m, ṁ) becomes in-
0.6 dependent of η̇. Hence, it cannot describe situations

F)/ F
10 1
where the η degrees of freedom are dynamic. Actually,
/

this represents a clear example of a non-admissible effec-

C
0.4

(
tive Lagrangian theory.
Let us finally turn to the case ∆ 6= 0 once more. It is
0.2 10 2 worth mentioning that the ASD can be newly derived by
starting from the effective Lagrangian for spin dynamics
under the tight-binding constraint ni = n0,i (η) and by
10 5 10 4 10 3 10 2 10 1 100 imposing the additional constraint η = zK m m0 expressing
0

K/J adiabaticity, see Eq. (18). A heuristic argument is given


in Appendix G for the single-impurity-spin case R = 1.
FIG. 9. Main precession frequency in the Fourier spectrum The formal derivation for the general case is worked out
of the real-time dynamics of m0 obtained from constrained in Appendix H. This must be seen as a successful consis-
spin-dynamics theory ωC as function of ϑ and K/J. Color
tency check of the formal theory.
code: normalized difference with the result of the full spin-
dynamics theory ωF .

X. CONCLUSIONS
spins for K/J  1, one finds a nontrivial spin dynamics
with a precessional motion of η while mtot = const. For Classical Heisenberg spin models are frequently used in
R = 1, there is no spin dynamics at all, since for ∆ = 0 atomistic spin-dynamics studies of condensed-matter sys-
the total spin is solely given by the single impurity spin, tems, nanostructures or molecular systems. From a prag-
and, therefore, the impurity spin is fixed to its initial di- matic point of view, they are quite attractive since the
rection due to total spin conservation. Hence, we note corresponding classical Hamiltonian equations of motion
that there are no special features here. form a nonlinear set of ordinary differential equations,
Turning to the constrained spin dynamics and special- which can be integrated by numerical means, such that,
izing Eq. (52) to the case ∆ = 0, R = 1 and B = 0, as compared to quantum-spin models, long propagation
provides us with the two equations 0 = m × η and times for a large number of spins are easily accessible.
ṁ = sKη × m, which correctly imply m = const. How- Usually, for generic model parameters, the resulting mi-
ever, the first equation is dubious, since it may conflict croscopic spin trajectories are chaotic and cover the entire
with an initial condition where m × η 6= 0. On the other accessible phase space, as it is expected for a nonlinear
hand, such an initial state is perfectly allowed by our con- classical ergodic system.
straint Eq. (50). This clearly implies that the constrained More regular dynamics is obtained for cases with
spin dynamics is inherently inconsistent. strongly varying exchange-coupling parameters or, equiv-
alently, for systems with a clear separation of intrinsic
time scales. Such situations are often quite realistic, and
101
a typical setup has been considered here. We have per-
formed a comprehensive study of a prototypical model
0.8 consisting of two impurity spins that are weakly cou-
pled to an antiferromagnetically coupled host-spin sys-
tem, i.e., slow impurity spins are interacting with fast
0.6
F)/ F

host spins. For initial states with energy close to the


100 ground-state energy, a very regular, mainly precessional
/

dynamics emerges, which calls for an effective low-energy


A

0.4
(

theory.
The purely classical system studied here has turned out
0.2 as highly interesting conceptually since its dynamics is
unexpectedly non-Hamiltonian in many cases, i.e., there
10 1 is no effective RKKY-like Hamiltonian that merely con-
10 5 10 4 10 3 10 2 10 1 100 sists of the impurity-spin degrees of freedom and is able
K/J to reproduce the impurity-spin dynamics. The reason is
that, quite generally, the time-scale separation leads to
FIG. 10. The same as Fig. 9 but comparing the ASD and the emergence of a topological spin torque, which pro-
the full spin-dynamics theory. foundly affects the spin dynamics. This is reminiscent of
13

the Berry phase that emerges in a quantum (host) sys- even in this case there is a good but not fully convincing
tem upon slow variation of classical model parameters agreement with the full theory.
(the impurity spins). An important difference, however, We have therefore studied another version of a con-
is that the (purely classical) topological spin torque ac- strained spin dynamics assuming that the host-spin sys-
tually represents a back-reaction of the local topological tem is tightly bound but not necessarily in the ground
charge density of the host system on the slow impurity state for the present impurity-spin configuration at any
spins. instant of time. Also this constrained spin dynamics
must be worked out carefully within the Lagrange for-
Our main ansatz for constructing an effective low-
malism, and again there is a topological spin torque in-
energy impurity-spin dynamics has been the adiabatic
volved. Spin dynamics under the tight-binding constraint
approximation, which is formulated as a constraint for
somewhat relaxes the ASD constraint. In fact, the ASD
the host-spin configuration. This constraint has to be in-
could be newly derived by enforcing the missing piece
corporated carefully: Making use of the constraint on the
again. Comparing with the full theory, we found that
level of the equations of motion runs into unacceptable
the relaxation of the constraint in fact results in an im-
inconsistencies. A consistent effective theory is obtained
proved effective theory, which now covers the entire weak-
when using the constraint to simplify the original Hamil-
coupling limit. This advantage, however, also comes at a
tonian. This naive effective theory, however, runs the
cost: Spin dynamics under the tight-binding constraint
risk of not respecting certain conservation laws, e.g., to-
necessarily involves host degrees of freedom, i.e., it fails
tal spin conservation and has been explicitly shown to fail
to provide an effective impurity-spin dynamics theory.
in cases, where the host-spin system has a finite total spin
More severely, however, the effective Lagrangian is sin-
moment. A satisfactory effective theory rather requires
gular in the case of a nonmagnetic host with a vanishing
to work in the Lagrange formalism which allows us to
total spin.
include arbitrary constraints in a consistent way. Using
Our present study can be seen as a first step towards an
the adiabatic constraint defines adiabatic spin dynam-
effective theory of RKKY real-time dynamics, i.e., where
ics (ASD). For the relevant weak-coupling limit, we were
impurity spins are coupled to a conduction-electron sys-
able to work out the non-Hamiltonian effective equations
tem, and work on this quantum-classical problem in al-
of motion analytically. The big impact of the topological
ready in progress. Clearly, this problem is more involved
spin torque appearing in the ASD equations becomes ev-
since with the Fermi energy of the electronic system there
ident when comparing the ASD results with those of the
is an additional energy scale to be considered. This also
naive theory.
implies the emergence of a length scale, resulting, e.g., in
From a theoretical perspective, the ASD appears as a the nontrivial distance dependence of the effective RKKY
very attractive approach: It follows a clear construction exchange. Furthermore, we expect to make contact with
principle, it maintains conservation laws resulting from the Berry curvature of the electronic system and a corre-
the symmetries of the original Hamiltonian, it provides sponding topological spin torque, replacing the topologi-
a true effective theory formulated in terms of the slow cal charge density of the purely classical host-spin system
impurity-spin degrees of freedom only, and it brings a studied here. Clearly, the quantum-classical problem is
hidden topological structure to light that substantially more relevant for interpreting experimental findings. We
modifies the slow spin dynamics. On the other hand, the believe that the insights gained from our present classical
applicability of the ASD stands and falls with the validity study will be very helpful for this next step.
of the constraint imposed, and unfortunately, contrary to
quantum systems, there is no direct classical equivalent
of the adiabatic theorem which ensures adiabaticity in ACKNOWLEDGMENTS
certain limits. Comparison of the predictions of the ASD
with those of the full theory treating all, slow and fast This work was supported by the Deutsche Forschungs-
degrees of freedom, is thus necessary. In fact, this has gemeinschaft (DFG) through the Cluster of Excellence
uncovered some deficiencies: While the ASD applies to “Advanced Imaging of Matter” - EXC 2056 - project ID
the weak-coupling limit only, as it was anticipated, it 390715994, and by the DFG Sonderforschungsbereich 925
also requires that the initial impurity-spin configuration “Light-induced dynamics and control of correlated quan-
is not too far from the ground-state configuration, and tum systems” (project B5).

Appendix A: Total spin conservation within the ASD

Within adiabatic spin dynamics the total spin, i.e., the sum of the total impurity spin S tot and the total host
spin stot , is conserved, if Heff is SO(3) symmetric. Here, we prove total-spin conservation for a collinear host-spin
14

structure. We start by computing the time derivative of the total spin:

d X X d X ∂Heff X X ∂n0,i (m)


(S tot + stot ) = S ṁr + s n0,i (m) = × mr + T r × mr + s ṁrµ , (A1)
dt r i
dt r
∂m r r irµ
∂mrµ

where, in the first step, we have inserted the equation of motion Eq. (7) to eliminate ṁr . Consider the second term
in the last expression. Making use of Eq. (35) we find:
X X s∆ s∆
T r × mr = zir zK (ṁ0 × m0 ) × mr = zK 3 (ṁ0 × m0 ) × m0 . (A2)
r r
m30 m0

To treat the third term, we employ Eq. (29):


 
X ∂n0,i (m) X 1 m0µ
s ṁrµ =s zi zK ṁrµ zir eµ − 2 m0
irµ
∂mrµ irµ
m0 m0
 
X 1 m0 1
= szK zi ṁ0 − (ṁ0 m0 ) 2 = zK s∆ 3 m0 × (ṁ0 × m0 ) . (A3)
i
m0 m0 m 0

This cancels the second term. SO(3) symmetry implies that Heff (m) has the general form

Heff (m) = f ((crr0 )r,r0 =1,...,R ) , (A4)

where f is an arbitrary smooth function of all inner products crr0 ≡ mr mr0 . Since m2r = 1, we have
X ∂Heff XX ∂f ∂cr0 r00 X X ∂f
× mr = × mr = (δrr0 mr00 + δrr00 mr0 ) × mr = 0 . (A5)
∂mr ∂cr0 r00 ∂mr ∂cr0 r00
r r r 0 r 00 r 0 00 r r

This proves that S tot + stot = const.

Appendix B: Solution of coupled ODE’s where s1 = c1 /|c1 | and s2 = c2 /|c2 | are sign factors. We
distinguish the two cases s1 = ±s2 and conclude that
We consider the following system of two (three- there is a conserved vector
component) ordinary differential equations s s
c2 c1
ẋ1 = c1 x2 × x1 , y ≡ y1 ± y2 = x1 ± x2 = const. (B6)
c1 c2
ẋ2 = c2 x1 × x2 , (B1)
Since Eqs. (B5) and (B6) imply
where c1 , c2 are constants. With the scaling transforma- p
tion ẏ 1 = ±s1 |c1 c2 | y × y 1 ,
p
y 1 = α1 x1 , y 2 = α 2 x2 , (B2) ẏ 2 = s2 |c1 c2 | y × y 2 , (B7)
we see that y 1 and y 2 and, thus, x1 and x2 precess with
where α1 , α2 are constants to be determined, we have
equal orientationparound y. To compute the precession
c1 frequency ωp = |c1 c2 | |y| we need the length of y:
ẏ 1 = y × y1 ,
α2 2
c2 |y|2 = |y 1 |2 + |y 2 |2 ± 2y 1 y 2
ẏ 2 = y × y2 . (B3)
α1 1 c2 2 c1 2
= x1 + x ± 2x1 x2 . (B8)
Choosing c1 c2 2
s s Here, we have used Eq. (B4) and α1 α2 = 1 in particular.
c2 c1 1 Note that Eq. (B1) immediately implies that x1 , x2 and
α1 = , α2 = = , (B4)
c1 c2 α1 x1 x2 are constant. The precession frequency
q
we get ωp = c22 x21 + c21 x22 ± 2 |c1 c2 | x1 x2 (B9)
p
ẏ 1 = s1 |c1 c2 | y 2 × y 1 , depends on the lengths of x1 and x2 and on the angle
p
ẏ 2 = s2 |c1 c2 | y 1 × y 2 , (B5) enclosed initially.
15

Let us now consider the differential equations Eq. (40) Appendix C: Using tight-binding constraints to
and Eq. (42) for x1 = m0 and x2 = mtot . The coeffi- simplify the equations of motion
cients corresponding to m0 and mtot read
In an attempt to construct an alternative effective the-
1 |K|s ory, let us start from the fundamental equations of mo-
c0 = s∆
,
1 + zK Sm3 m0 mtot m0 tion (3) for the impurity spins S r = Smr . Using the
0
notations of Sec. II, we have
s∆ 1 |K|s
ctot = zK s∆
, (B10)
Sm0 1 + zK Sm 3 m0 mtot m0 ṁr = Ks nir × mr , (C1)
0

respectively. With ctot /c0 = zK s∆/Sm0 , this means pre- where we have assumed B r = 0, for simplicity. Further,
cession around the axis the equations of motion for sir = snir read
s s X
s |∆| S m0 ṅir = KS mr × nir + s Jir i0 ni0 × nir . (C2)
y= m0 ± mtot = const. (B11) i0
S m0 s |∆| We want to exploit the constraint

where the “+”-sign applies for zK ∆ > 0 and the “−”- ni = zi η (C3)
sign for zK ∆ < 0. After rescaling, we note that y is
collinear to (zi = ±1). This tight-binding constraint expresses that
for K  J all host spins are tightly bound together such
s|∆| that, irrespective of the impurity-spin configuration, all
± m0 + Smtot = const. (B12)
m0 ni are collinear to a unit vector η at all times t.
P It is tempting, but incorrect, to use the constraint to
Since stot = s i zi η = s∆zK m0 /m0 = ±s|∆|m0 /m0 , simplify the equations of motion as will be shown here.
this means that the precession axis is just defined by the Eq. (C3) implies that the second term in Eq. (C2) van-
total spin stot + S tot , as expected on physical grounds. ishes. Using the constraint once more, we can eliminate
A simple result for the precession frequency is obtained the host spins and are left with
in the case of two impurity spins (R = 2) with zi1 =
−zi2 , where we can exploit the relation m0 mtot = (m1 − ṁr = Kszir η × mr , (C4)
m2 )(m1 + m2 ) = 0:
and
r
1 s2 |∆|2 η̇ = KS mr × η (C5)
ωp = |K|s + m2tot . (B13)
m0 S2
for all r = 1, ..., R. This yields
Assuming that s = S and that |∆| = 1 (antiferromag-
netic host-spin configuration and odd L), the precession (mr − mr0 ) × η = 0 , (C6)
frequency is
or
p r
1 + m2tot 1 ϑ mr − mr0
ωp = |K|s = |K|s 1 + 4 cos2 , η=± , (C7)
m0 2 sin ϑ/2 2 |mr − mr0 |
(B14)
where ϑ ∈ ]0, π] is the conserved angle enclosed by m1 for all r, r0 . For arbitrary directions mr and for R ≥ 3,
and m2 . however, this obviously leads to contradictions.

Appendix D: Lagrange formalism using tight-binding constraints

The correct dynamics under a constraint of the form n = n0 (η) (ni = n0,i (η)) can be derived from the effective
Lagrangian

Leff (η, η̇, m, ṁ) ≡ L(n0 (η), (d/dt)n0 (η), m, ṁ) , (D1)
P P
where L(n, ṅ, m, ṁ) = S j A(mj )ṁj + s i A(ni )ṅi − H(n, m) is the full Lagrangian. Here, we use the short-
hand notation n0 = (n0,1 , ..., n0,L ), n = (n1 , ..., nL ) and m = (m1 , ..., mR ). Furthermore, A(r) is a vector field
satisfying ∇ × A(r) = −r/r3 , and which can thus be interpreted as the vector potential of a unit magnetic (Dirac)
monopole located at r = 0. In the standard gauge [38], this is given by A(r) = −(1/r2 )(ez × r)/(1 + ez r/r). The
16

equations of motion deriving from the full Lagrangian are equivalent with the Hamilton equations (3), see Ref. [31]
for further details.
With
d
n0,i (η) = (η̇∇)n0,i (η) (D2)
dt
we find:
X X  
Leff (η, η̇, m, ṁ) = S A(mj )ṁj + s A(n0,i (η)) (η̇∇)n0,i (η) − Heff (η, m) , (D3)
j i

where i = 1, ..., L and j = 1, ..., R, and where Heff (η, m) = H(sn0 (η), Sm). To get the Lagrange equations of motion,
we first compute
∂Leff (η, η̇, m, ṁ) X ∂Heff (η, m)
=S ∇r Aβ (mr )ṁrβ − . (D4)
∂mr ∂mr
β

and
∂Leff (η, η̇, m, ṁ) X X ∂Aβ (n0,i (η)) ∂Heff (η, m)
=s Aβ (n0,i (η)) (η̇∇)∇n0,iβ (η) + s ∇n0,iα (η)(η̇∇)n0,iβ (η) − .
∂η ∂n0,iα ∂η
iβ iαβ
(D5)
Here, ∇r = ∂/∂mr , and Greek indices α, β, ... ∈ {x, y, z}. Furthermore,
∂Leff (η, η̇, m, ṁ) ∂Leff (η, η̇, m, ṁ) X
= SA(mr ) , =s Aα (n0,i (η))∇n0,iα (η) , (D6)
∂ ṁr ∂ η̇ iα

which yields
d ∂Leff (η, η̇, m, ṁ)
= S(ṁr ∇r )A(mr ) (D7)
dt ∂ ṁr
and
d ∂Leff (η, η̇, m, ṁ) X ∂Aα (n0,i (η)) X
=s (η̇∇n0,iβ (η))∇n0,iα (η) + s Aα (n0,i (η)∇ (η̇∇n0,iα (η)) . (D8)
dt ∂ η̇ ∂n0,iβ iα
iαβ

The last term equals the first term on the right-hand side of Eq. (D5) in the Lagrange equations, since ∇ and η̇∇
commute, such that we are left with:
d ∂Leff ∂Leff X ∂Heff (η, m)
0= − = S(ṁr ∇r )A(mr ) − S ∇r Aβ (mr )ṁrβ +
dt ∂ ṁr ∂mr ∂mr
β
∂Heff (η, m)
= S(∇r × A(mr )) × ṁr + , (D9)
∂mr
and
d ∂Leff ∂Leff ∂ X ∂Aβ (n0,i (η))
0= − = Heff (η, m) − s ∇n0,iα (η)(η̇∇)n0,iβ (η)
dt ∂ η̇ ∂η ∂η ∂n0,iα
iαβ
X ∂Aα (n0,i (η))
+s (η̇∇n0,iβ (η))∇n0,iα (η)
∂n0,iβ
iαβ
∂Heff (η, m)
= +T , (D10)
∂η
where T stands for the last two terms. Taking in Eq. (D9) the cross product from the right, (...) × mr , we find
∂Heff (η, m)
S((∇r × A(mr )) × ṁr ) × mr + × mr = 0 . (D11)
∂mr
17

Using ∇r × A(mr ) = −mr /m3r , expanding the remaining double cross product and exploiting that mr is a unit
vector, yields:

∂Heff (η, m)
S ṁr = × mr , (D12)
∂mr
which just recovers the standard form of the equation of motion for mr . On the contrary, the equation of motion for
η, which is obtained from Eq. (D10) by taking the cross product with η, is unconventional:
∂Heff (η, m)
0= ×η+T ×η. (D13)
∂η

Note that actually we should have added Lagrange- Note that T η̇ = 0. Inserting the result for T in the
multiplier terms, Leff (η, η̇, m, ṁ) 7→ Leff (η, η̇, m, ṁ) − equation of motion, we obtain
2 2
P
λ r (mr − 1) − λ(η − 1), to account for the normaliza-
r ∂Heff (η, m)
tion conditions mr = 1 and η 2 = 1. However, this would
2
0= × η + (η̇ × Ω) × η . (D19)
merely have resulted in additional summands 2λr mr and ∂η
2λη on the right-hand sides of Eqs. (D9) and (D10), re- If the pseudovector Ω is interpreted as a magnetic field
spectively, which do not contribute after taking the re- in η-space, T = η̇ × Ω is the Lorentz force (per unit
spective cross products (...) × mr and (...) × η. On the charge) and T × η the corresponding torque. On the
other hand, taking the dot products, (...)·mr and (...)·η, other hand, the analogy cannot be made complete, as
in Eqs. (D9) and (D10), respectively, just yields the nec- the curl of the vector potential of a “magnetic monopole”,
essary conditional equations for λr and λ, if these were ∇i × A(n0,i ) = −n0,i /n30,i , is a field in n0,i -space.
required.
T gives rise to a geometrical spin torque T × η and
can be read off from Eq. (D10): Appendix E: Spin dynamics under tight-binding
X  ∂Aα (n0,i (η)) ∂Aβ (n0,i (η))  constraints
T =s −
∂n0,iβ ∂n0,iα
iαβ Starting from the constraint, ni = n0,i (η) = zi η, the
· (η̇∇)n0,iβ (η) ∇n0,iα (η) . (D14) computation of the topological spin torque is straightfor-
ward. We have:
Exploiting once more the defining property of the vec-
tor potential, ∇i × A(n0,i ) = −n0,i /n30,i , and using the ∂n0,i (η)
normalization n0,i = 1 in the end, we find: = zi eµ (E1)
∂ηµ
X
T =s αβγ ∇n0,iα (η) (η̇∇)n0,iβ (η) n0,iγ (η) and thus
iαβγ X X
XX Ωµν = s (zi eµ ) × (zi eν ) · (zi η) = sη zi eµ × eν ,
=s ∇µ n0,i (η) × ∇ν n0,i (η) · n0,i (η) η̇ν eµ . i i
i µν (E2)
(D15) or in terms of the psuedovector
The scalar triple product defines an antisymmetric tensor Ω = s∆ η , (E3)
of rank two:
X ∂n0,i (η) ∂n0,i (η) with
Ωµν = s × · n0,i (η)
i
∂ηµ ∂ην L
X
X ∆≡ zi . (E4)
= −Ωνµ = µνρ Ωρ , (D16) i=1
ρ

where the last equation For an antiferromagnetic host, e.g., we have ∆ = ±1 if L


P defines the pseudovector Ω with is odd, and ∆ = 0 if L is even. Generally, Ω is just the
components Ωρ = 21 µν µνρ Ωµν :
total host spin:
s XX
Ω= αβγ ∇n0,iα × ∇n0,iβ n0,iγ , (D17) X X
2 i
αβγ
stot = si = szi η = s∆ η = Ω . (E5)
i i
which has precisely the form of the “magnetic vorticity”
[37]. Hence: Now, the topological spin torque reads as

T =
X
Ωµν η̇ν eµ = η̇ × Ω . (D18) T × η = (η̇ × Ω) × η = s∆ (η̇ × η) × η
µν = −s∆ η̇ = −sṅtot = −ṡtot , (E6)
18

and therefore the set of equations of motion is given by Appendix F: Non-admissible effective Lagrangian in
the case ∆ = 0
∂Heff (η, m)
s∆η̇ = ×η,
∂η
We proceed with the discussion of the topologically
∂Heff (η, m)
S ṁr = × mr . (E7) trivial case ∆ = 0. Here, Eq. (E10) implies that m0 ×η =
∂mr 0, and this yields η = ±m0 /m0 , and η = zK m0 /m0
We see that, for ∆ 6= 0 and opposed to the ASD, one if, initially, the dynamics starts with the ground-state
arrives at a standard Hamiltonian dynamics for the re- configuration of the host spins for given m. We would
maining degrees of freedom η and m governed by an thus exactly recover the naive adiabatic theory, see Sec.
effective Hamiltonian which is obtained by making use of V, and Eq. (24) in particular. However, the constrained
the constraint in the original Hamiltonian. Lagrangian theory is inconsistent in general, as m0 ×η =
Explicitly, the effective Hamiltonian, Heff (η, m) = 0 may conflict with an initial state of the system where
H(si = szi η, S r = Smr ), reads m0 × η 6= 0.
R
X R
X In the case ∆ = 0, one can in fact show that con-
Heff (η, m) = E0 +KsSη zir mr −S mr B r . (E8) straining the spin system by imposing Eq. (50) leads to
r=1 r=1 a singular effective Lagrangian. This singularity is sub-
tle. For a discussion, we first start with a short note on
This describes a central spin model: The impurity spins
Lagrange mechanics for a system of point particles de-
S r = Smr couple with strengths zir K to the central spin
scribed by N coordinates q = (q1 , ..., qN ). Consider the
stot = sntot = s∆η. With
Euler-Lagrange equations
∂Heff (η, m) X
= sSK zir mr , ∂L(q, q̇) d ∂L(q, q̇)
∂η r 0= −
∂qj dt ∂ q̇j
∂Heff (η, m)
= sSKzir η − SB r , (E9) ∂L X ∂ 2 L X ∂2L
∂mr = − q˙i − q¨i . (F1)
∂qj ∂qi ∂ q̇j ∂ q˙i ∂ q̇j
the Hamiltonian equations are given by: i i
X
s∆η̇ = sSK zir mr × η , The Lagrangian is called singular, if the Hesse matrix
r Hij = ∂ 2 L/∂ q˙i ∂ q̇j cannot be inverted. This is the typi-
S ṁr = sSKzir η × mr − SB r × mr . (E10) cal case in classical spin dynamics and explains why there
is no simple connection between Hamiltonian and La-
Let us start the discussion with the case ∆ 6= 0 and de-
grangian formalism mediated by a Legendre transforma-
rive some consequences of the equations of motion. First,
tion (see, e.g., the discussion in the supplemental ma-
we note that the total spin is conserved, if B r = 0:
terial of Ref. [31], section B) and why it is convenient
d X
to stay with in Lagrangian framework when discussing
(sntot + Smtot ) = sṅtot + S ṁr
dt r
constrained classical spin systems. In principle, how-
ever, a Hamiltonian formulation can be derived directly
∂Heff (η, m) X ∂Heff (η, m)
= ×η+ × mr = 0 , from a singular Lagrangian with the Dirac-Bergmann for-
∂η r
∂mr malism [39–41]. However, the problem is more severe if
(E11) L(q, q̇) = L(q). In such a case not only the Hesse matrix
is singular (in fact, H = 0), but also the coefficient matrix
exploiting Eq. (E9). Hence, sntot + Smtot = const. En-
Kij = ∂ 2 L/∂qi ∂ q̇j vanishes. This may lead to inconsis-
ergy conservation (d/dt)Heff (η, m) = 0 follows by con-
tencies and, hence, such Lagrangians are not admissible.
struction and can also be verified explicitly. There are
This means that imposing the constraint is unphysical
further conserved quantities. From Eq. (E10) we imme-
and does not lead to a valid effective theory.
diately find |mr | = const. = 1, |η| = const. = 1, and for
The latter exactly applies to our spin system when
B r = 0 we can derive
imposing the constraint (50) in the case ∆ = 0. This
∆η̇ = SKm0 × η , can be seen by a gauge transformation of the effective
ṁ0 = sKη × mtot , Lagrangian. We use the constraint Eq. (50) explicitly to
ṁtot = sKη × m0 . (E12) rewrite the effective Lagrangian (D1). With

Further, Eq. (E10) yields m0 mtot = const., and for d


R = 2, in particular, we trivially have m0 mtot = 0. Gen- n0,i (η) = (η̇∇)n0,i (η) = zi η̇ (F2)
dt
erally, one cannot infer m0 = const. or mtot = const.,
opposed to the ASD and Eq. (43). We rather have we find
m20 + m2tot = const. and (m0 ± mtot )2 = const. only. We X X
conclude that the effective spin dynamics under tight- Leff (η, η̇, m, ṁ) = S A(mj )ṁj + s A(zi η)zi η̇
binding constraints differs from the naive adiabatic the- j i
ory as well as from ASD. − Heff (η, m) . (F3)
19

The curl of the vector potential A(r) ≡ Aez (r) = Appendix H: Alternative derivation of the ASD
−(1/r2 )(ez × r)/(1 + ez r/r) is invariant under a gauge
transformation that replaces ez by an arbitrary unit vec- Interestingly, one can indeed give an alternative deriva-
tor e. The Euler-Lagrange equations are in fact invariant tion of the ASD by starting from the effective spin dy-
under a local, i-dependent gauge transformation, speci- namics under tight-binding constraints discussed above
fied by ez 7→ zi ez , of the second term on the right-hand and by imposing the additional constraint
side:
η = η 0 (m) ≡ zK m0 /m0 , (H1)
1 ez × η
Aez (zi η) 7→ Azi e (zi η) = − . (F4)
η 2 1 + ez η/η with zK = −K/|K|, i.e., assuming that the mutually
bound host spins are, at any instant of time, in the
Note that this does no longer depend on the site index ground-state configuration corresponding to the respec-
i. The result of the transformation is that the second tive impurity-spin configuration.
term
P on the right-hand side of Eq. (F3) vanishes if ∆ = To prove our claim, we start from the effective Hamil-
i zi = 0 and, hence, the transformed but equivalent tonian Eq. (E8). Inserting the constraint, Eq. (H1), in
effective Lagrangian, the effective Hamiltonian Heff (η, m), yields an effective
X Hamiltonian depending on the impurity-spin degrees of
Leff (η, m, ṁ) = S A(mj )ṁj − Heff (η, m) , (F5) freedom only,
j
R
X
lacks the dependence on η̇. Therefore, it is not admissi- Heff (m) = E0 − |K|sSm0 − S mr B r , (H2)
ble. r=1

which, of course, equals the one derived earlier, see Eq.


(21). For the derivation of the effective equation of mo-
Appendix G: Single impurity spin coupled to a tion for m under the additional constraint, we use the
magnetic field action principle and follow the steps outlined in Ref.
[31]. The unconstrained dynamics is governed by the
For R = 1, i.e., for a single impurity spin S = Sm, Lagrangian
and for ∆ 6= 0, Eq. (E10) reads: X
Leff (η, η̇, m, ṁ) = A(η)s∆η̇ + A(mr )S ṁr
∆η̇ = SKm × η , ṁ = sKη × m − B × m , (G1) r
− Heff (η, m) . (H3)
where we have set, without loss of generality, zi1 = +1.
This implies Using the constraint Eq. (H1), we get an effective La-
grangian depending on m, ṁ only:
s
ṁ = − ∆η̇ − B × m . (G2)
X
S Leff (m, ṁ) = A(η 0 (m)) s∆ (ṁr ∇r )η 0 (m)
r
For given m, the ground state of the tightly bound host- +
X
A(mr )S ṁr − Heff (m) . (H4)
spin subsystem is given by ni = zi η with η = zK m = r
−signK m. Taking the time derivative of this additional
condition yields: This form of Leff differs only slightly from the one dis-
cussed in Ref. [31] such that the resulting Lagrange equa-
η̇ = zK ṁ . (G3) tions have exactly the same form as Eq. (7):

Inserting this relation into Eq. (G2), we find in case of ∂Heff (m)
S ṁr = × mr + T r × mr . (H5)
antiferromagnetic Kondo coupling (zK < 0), antiferro- ∂mr
magnetic host-spin structure and odd L (∆ = +1) Here T r is given via
1 X
ṁ = m×B. (G4) Trµ = Trµ (m, ṁ) = Ωrµ,sν (m)ṁsν , (H6)
1 − s/S sν

This describees precession with a renormalized frequency in terms of


as predicted by the ASD, see Ref. [31] and Eq. (37). It
seems that the ASD can be re-derived by imposing the Ωrµ,sν (m) = 4πs∆erµ,sν (m) , (H7)
above additional constraint. For the general case of arbi-
which differs from Eq. (9) by the missing sum over i and
trary R, however, we must carefully base the considera-
the additional factor ∆. We have:
tions on the action principle, as shown below, since using
Eq. (G3) to simplify the equation of motion lacks formal 1 ∂η 0 (m) ∂η 0 (m)
erµ,sν (m) = × · η 0 (m) . (H8)
justification. 4π ∂mrµ ∂msν
20

This topological charge density can be computed as in This is exactly the result found earlier, see Eq. (32), and
Sec. VI, and we find thus yields the same expression for the topological spin
torque, Eq. (35), and the same equations of motion, Eq.
1 X 1 (36), for mr .
erµ,sν (m) = zK zir zis εµντ m0τ 3 . (H9)
4π τ
m 0

Therewith, we have
X 1
Ωrµ,sν (m) = zir zis zK s∆ εµντ m0τ . (H10)
τ
m30

[1] D. C. Mattis, The Theory of Magnetism (Springer, [23] T. Liu, J. Ren, and P. Tong, Phys. Rev. B 98, 184426
Berlin, 1981). (2018).
[2] A. Auerbach, Interacting electrons and quantum mag- [24] M. Onoda and N. Nagaosa, Phys. Rev. Lett. 96, 066603
netism (Springer, New York, 1994). (2006).
[3] W. Nolting and A. Ramakanth, Quantum Theory of Mag- [25] N. Umetsu, D. Miura, and A. Sakuma, J. Appl. Phys.
netism (Springer, Berlin, 2009). 111, 07D117 (2012).
[4] P. W. Anderson, Phys. Rev. 79, 350 (1950). [26] M. Sayad, R. Rausch, and M. Potthoff, Phys. Rev. Lett.
[5] C. Zener, Phys. Rev. 82, 403 (1951). 117, 127201 (2016).
[6] P.-G. de Gennes, Phys. Rev. 118, 141 (1960). [27] M. Sayad, R. Rausch, and M. Potthoff, Europhys. Lett.
[7] M. A. Ruderman and C. Kittel, Phys. Rev. 96, 99 (1954). 116, 17001 (2016).
[8] T. Kasuya, Prog. Theor. Phys. 16, 45 (1956). [28] D. Marx and J. Hutter, Ab initio molecular dynamics:
[9] K. Yosida, Phys. Rev. 106, 893 (1957). Theory and Implementation, In: Modern Methods and
[10] A. Schwabe, I. Titvinidze, and M. Potthoff, Phys. Rev. Algorithms of Quantum Chemistry, NIC Series, Vol. 1,
B 88, 121107(R) (2013). Ed. by J. Grotendorst, p. 301 (John von Neumann Insti-
[11] A. Secchi, S. Brener, A. I. Lichtenstein, and M. I. Kat- tute for Computing, Jülich, 2000).
snelson, Ann. Phys. (N.Y.) 333, 221 (2013). [29] M. Nakahara, Geometry, Topology, and Physics (Inst. of
[12] A. C. Hewson, The Kondo Problem to Heavy Fermions Physics Publ., Bristol, 1998).
(Cambridge University Press, Cambridge, 1993). [30] A. Bohm, A. Mostafazadeh, H. Koizumi, Q. Niu, and
[13] S. Doniach, Physica B+C 91, 231 (1977). J. Zwanziger, The Geometric Phase in Quantum Systems
[14] C. Jayaprakash, H. R. Krishnamurthy, and J. W. (Springer, Berlin, 2003).
Wilkins, Phys. Rev. Lett. 47, 737 (1981). [31] M. Elbracht, S. Michel, and M. Potthoff, Phys. Rev.
[15] Y. Luo, C. Verdozzi, and N. Kioussis, Phys. Rev. B 71, Lett. 124, 197202 (2020).
033304 (2005). [32] H.-B. Braun, Adv. Phys. 61, 1 (2012).
[16] A. Schwabe, D. Gütersloh, and M. Potthoff, Phys. Rev. [33] G. Finocchio, F. Büttner, R. Tomasello, M. Carpentieri,
Lett. 109, 257202 (2012). and M. Kläui, J. Phys. D 49, 423001 (2016).
[17] A. Schwabe, M. Hänsel, M. Potthoff, and A. K. Mitchell, [34] M. Stier, W. Häusler, T. Posske, G. Gurski, and
Phys. Rev. B 92, 155104 (2015). M. Thorwart, Phys. Rev. Lett. 118, 267203 (2017).
[18] B. Skubic, J. Hellsvik, L. Nordström, and O. Eriksson, [35] T. Gilbert, Phys. Rev. 100, 1243 (1955).
J. Phys.: Condens. Matter 20, 315203 (2008). [36] T. Gilbert, Magnetics, IEEE Transactions on 40, 3443
[19] A. Sakuma, J. Phys. Soc. Jpn. 81, 084701 (2012). (2004).
[20] S. Bhattacharjee, L. Nordström, and J. Fransson, Phys. [37] N. R. Cooper, Phys. Rev. Lett. 82, 1554 (1999).
Rev. Lett. 108, 057204 (2012). [38] P. A. M. Dirac, Proc. R. Soc. London A 133, 60 (1931).
[21] M. Sayad and M. Potthoff, New J. Phys. 17, 113058 [39] P. A. M. Dirac, Can. J. Math. 2, 129 (1951).
(2015). [40] J. L. Anderson and P. G. Bergmann, Phys. Rev. 83, 1018
[22] J. Fransson, J. Ren, and J.-X. Zhu, Phys. Rev. Lett. (1951).
113, 257201 (2014). [41] P. G. Bergmann, I. Goldberg, A. Janis, and E. Newman,
Phys. Rev. 103, 807 (1956).

You might also like