Alouges_Num_Math_2014
Alouges_Num_Math_2014
Numerische Mathematik
ISSN 0029-599X
Numer. Math.
DOI 10.1007/s00211-014-0615-3
1 23
Your article is protected by copyright and
all rights are held exclusively by Springer-
Verlag Berlin Heidelberg. This e-offprint is
for personal use only and shall not be self-
archived in electronic repositories. If you wish
to self-archive your article, please use the
accepted manuscript version for posting on
your own website. You may further deposit
the accepted manuscript version in any
repository, provided it is only made publicly
available 12 months after official publication
or later and provided acknowledgement is
given to the original source of publication
and a link is inserted to the published article
on Springer's website. The link must be
accompanied by the following text: "The final
publication is available at link.springer.com”.
1 23
Author's personal copy
Numer. Math. Numerische
DOI 10.1007/s00211-014-0615-3 Mathematik
Abstract In this paper, we propose a new scheme for the numerical integration of
the Landau–Lifschitz–Gilbert (LLG) equations in their full complexity, in particular
including stray-field interactions. The scheme is consistent up to order 2 (in time), and
unconditionally stable. It combines a linear inner iteration with a non-linear renor-
malization stage for which a rigorous proof of convergence of the numerical solution
toward a weak solution is given, when both space and time stepsizes tend to 0. A numer-
ical implementation of the scheme shows its performance on physically relevant test
cases. We point out that to the knowledge of the authors this is the first finite element
scheme for the LLG equations which enjoys such theoretical and practical properties.
1 Introduction
In 1935 Landau and Lifschitz proposed a model for the description of the evolution of
the magnetization in a ferromagnetic material [15]: Suppose that the three dimensional
ferromagnetic sample under consideration occupies some domain Ω ⊂ R3 and let
F. Alouges (B)
CMAP-École Polytechnique, Route de Saclay, 91128 Palaiseau cedex, France
e-mail: [email protected]
J. Steiner
Institute for Applied Mathematics, Mathematik Zentrum, Endenicher Allee 60, 53115 Bonn, Germany
123
Author's personal copy
F. Alouges et al.
(As above, vectorial quantities will be denoted in bold letters in the sequel to avoid
ambiguity). The parameters in the equation are the damping parameter α, the gyro-
magnetic constant γ0 and the saturation magnetization Ms . The so-called effective
field Heff , measured in units of Ms , is given by the functional derivative of the micro-
magnetic (free) energy E, more precisely
∂E
Heff (m) = − = d 2 Δm + Hd (m) + Hext + Q (e · m) e (2)
∂m
The four contributions to the effective field in (2) and the energy in (3), respec-
tively, correspond to the so-called exchange, stray-field, applied and anisotropy field
or energy, respectively. The material constants in (2) and (3) are the exchange con-
stant d, the anisotropy constant Q and the anisotropy direction e (also called the easy
axis). Furthermore, the vector field Hext models an applied magnetic field. We will
also use the notation Haniso = Q (e · m) e. The stray field Hd (m) is the magnetic
field induced by the magnetization distribution m via the following (subset of) static
Maxwell equations
curl Hd (m) = 0 in R3
(4)
div (Hd (m) + m) = 0 in R3 .
Notice that in the equation above, all fields are measured in units of Ms . Below the Curie
temperature, the magnetization can be described by a directional field that we rescale
to unit length. It is straightforward to check that the magnitude of the magnetization
is formally conserved by the dynamics (1). Take note that the gyromagnetic term is a
conservative term while the damping term leads to the following energy dissipation
law
123
Author's personal copy
A convergent and precise finite element scheme
d α
E(m(t)) = − |∂t m|2 dx . (6)
dt γ0 M s
Ω
Let us recall the notion of a weak solution to (1) from [6] and [18].
123
Author's personal copy
F. Alouges et al.
∂t m · dx dt − α (m × ∂t m) · dx dt
ΩT ΩT
d
= d2 m × ∂xi m · ∂xi dx dt (7)
i=1Ω
T
3. The magnetization initially satisfies m(x, 0) = m0 (x) in the trace sense, and
4. The energy decreases according to
E(m(T )) + α |∂t m|2 dx dt ≤ E(m(0)). (8)
ΩT
As in [5], our discretization relies on piecewise linear finite elements in space com-
bined with a linear interpolation in time. The domain Ω is discretized by a conformal
triangulation Th of mesh size h with vertices (xih )1≤i≤Nh . Let us denote by (φih )1≤i≤Nh
the set of associated piecewise linear basis functions that satisfy φih (xhj ) = δi, j at the
vertices xhj for 1 ≤ i, j ≤ Nh , where δi, j denotes the Kronecker symbol. This amounts
to a standard P 1 (Th )-discretization. Based on the scalar basis (φih )1≤i≤Nh we construct
the vector-valued finite element space as
Vh = uh = ui φih , s.t. ∀i, ui ∈ R3 .
i
Due to the unit-length constraint (5), the solution to (7) is sought in the subset
Mh = uh ∈ Vh , s.t. ∀i, ui ∈ S2 ⊂ Vh .
Let us also introduce the tangent space to mh = i mi φih ∈ Mh , denoted by
K m = vh = vi φih , s.t. ∀i, vi · mi = 0 .
i
Ih : C 0 (Ω, R3 ) → Vh
u → u(xih )φih . (9)
i
123
Author's personal copy
A convergent and precise finite element scheme
To simplify notations, the index h of the ansatz functions will be neglected from now
on most of the times, i.e., we write u, v, etc. instead of uh , vh , respectively, in case
this does not lead to any ambiguities.
The finite element scheme proposed in [5] relies on the observation that the LLG Eq.
(1)—with the notation v = ∂t m—can be rewritten in the following weak form
α v · dx + α m × v · dx
Ω Ω
= −d 2
∇m · ∇ dx + (Hd (m) + Hext + Haniso (m)) · dx. (10)
Ω Ω
Equation (10) holds for every test function ∈ H 1 (Ω, R3 ) that satisfies (x)·m(x) =
0 for a.e. x in Ω. The reformulation of (1) in the form of (10) motivated the following
θ -scheme introduced in [1]:
It is noteworthy that this procedure requires the solution of a linear equation in each
time step only. Moreover, due to the fact that the symmetric part of the underlying
matrix is positive definite, existence and uniqueness of a solution to (11) is guaranteed.
T
The time discrete solution constructed via algorithm (11) at time-steps N =
k
is interpolated as follows:
t − nk n+1 (n + 1)k − t n
mh,k (x, t) = m (x) + m (x),
k k
−
mh,k = mn (x), vh,k = vn (x).
123
Author's personal copy
F. Alouges et al.
−
Our notational convention is thus that mh,k , mh,k and vh,k refer to suitable time
interpolants of the time discrete approximation m and vn . Notice that mh,k is piece-
n
−
wise linear in time whereas mh,k and vh,k are piecewise constant. (The introduction
of the piecewise constant magnetization will be useful in the convergence proof).
Based on this discretization, weak convergence of the constructed approximation was
established in [1]. Both the proof of this result and the proof in case of our new
scheme consist of the following two main “classical” steps: as a first step establish-
ing an energy estimate which guarantees the convergence (sufficiently strong) of the
sequence constructed and then in a second step verifying that the limit indeed satisfies
the equation. As far as the first step is concerned, the following section addresses
the fact that the energy behaves well under renormalization—in principle a strongly
nonlinear modification of the flow.
The influence of the renormalization on the exchange energy was for instance inves-
tigated in [2] in the continuous case. More precisely, it was shown that for maps
w ∈ H 1 (Ω, R3 ) with |w(x)| ≥ 1 a.e. x ∈ Ω one has
w 2
∇ dx ≤ |∇w|2 dx. (12)
|w|
Ω Ω
In 3d, the condition (13)—and hence (14)—is for instance satisfied provided all
dihedral angles of the tetrahedra of the mesh are smaller than π/2, see [17].
Consider an exact solution m(nk) at time nk. As remarked in [1], it is a priori not suf-
ficient to choose θ = 21 in (11) to achieve quadratic order. Indeed, the renormalization
stage inherently introduces an order 2 term due to the expansion
123
Author's personal copy
A convergent and precise finite element scheme
m(nk) + kv k2
mn+1 := = m(nk) + kv − |v|2 m(nk) + O(k 3 ) . (15)
|m(nk) + kv| 2
Hence, in order to have mn+1 = m((n + 1)k) + O(k 3 ), it is necessary to take this
term into account in the variational formulation that defines v. We hence modify the
time-discrete approximation of the magnetization m which we heuristically derive in
the following. It is well known that the midpoint rule is exact up to cubic error, i.e.,
1
m((n + 1) k) = m(nk) + kmt (n + ) k + O(k 3 )
2
k
= m(nk) + k mt (nk) + mtt (nk) + O(k 3 ) .
2
Now, setting
k
v = Pm⊥ mt (nk) + mtt (nk)
2
k
= mt (nk) + Pm⊥ mtt (nk) , (16)
2
where Pm⊥ denotes the projection onto the orthogonal component of m(nk), and
noticing that v · m(nk) = 0, a Taylor expansion up to cubic order reveals
m(nk) + kv k2
= m(nk) + kv − |v|2 m(nk) + O(k 3 )
|m(nk) + kv| 2
k2
= m(nk) + kmt (nk) + Pm⊥ (mtt (nk))
2
k 2
− |mt (nk)|2 m(nk) + O(k 3 )
2
k2
= m(nk) + kmt (nk) + mtt (nk) + O(k 3 )
2
= m((n + 1) k) + O(k 3 ) .
Notice that we have made use in the above computation that, due to the unit length
constraint, one has
We therefore propose to modify the original first order scheme by replacing the
tangential update with the higher order approximation (16).
We will use the simplifying shorthand notation m = m(nk) and mt = mt (nk).
Let us proceed with the derivation of the equation that is satisfied by v = mt (nk) +
2 Pm⊥ mtt (nk), i.e. the counterpart to (10). The equation will be inferred from the
k
123
Author's personal copy
F. Alouges et al.
Remark 1 Although the midpoint rule is of order 2, our scheme is only almost of
order 2—as the section’s title suggests and as we will see in the sequel. This is due
to a regularizing term that is introduced in order to obtain the necessary estimates
in the convergence proof. This term prevents the scheme from being of order 2, in
the sense that the consistency error is not of order O(k 3 ) but only O(k 3− ) for any
> 0. On the other hand, this regularization approach allows for unconditional
convergence of the scheme. The approximation in space is still of order 1. Indeed,
only piecewise linear functions (P 1 Lagrange finite elements) are used in order to
ensure condition (13). It is well known that this condition, which for the time being is
mandatory to prove the convergence of the scheme, does not hold for higher order finite
elements.
where
∂Heff
(mt ) = d 2 mt + Hd (mt ) + Q (e · mt ) e
∂m
and where we once again used the unit length constraint (5). Now, we compute from
(17) and (18).
k
αv + m × v = αmt + m × mt + α Pm⊥ mtt + m × Pm⊥ mtt
2
k
= Pm⊥ αmt + m × mt + (αmtt + m × mtt )
2
k ∂Heff k
= Pm⊥ Heff (m) + (mt ) − (Heff (m) · m)mt
2 ∂m 2
for any test function with · m = 0. Observe that mt (nk) = v + O(k), cf. (16).
Therefore up to higher order terms
123
Author's personal copy
A convergent and precise finite element scheme
k k ∂Heff
α + (Heff (m) · m ) v · + m × v · dx − (v) · dx
2 2 ∂m
Ω Ω
= Heff (m) · dx + O(k 2 ), (19)
Ω
where we remind that m = m(nk) and mt = mt (nk). Observe that, if we neglect the
residual term, the latter equation is (at first sight surprisingly) linear in v. However, its
well-posedness is non-obvious since both the first and the last contributions on the l.h.s.
of (19) potentially affect the definiteness of the symmetric part of the operator. In order
to guarantee solvability and uniqueness we proceed with higher order modifications
that will finally lead to a well posed formulation. We address the first contribution and
define (see Fig. 1)
⎧
⎪ k
⎪
⎪α + min(x, M) for x ≥ 0,
⎪
⎪ 2
⎨
ϕ̃ M (x) = α (20)
⎪
⎪ for x < 0.
⎪
⎪
⎪
⎩1 + k
min(−x, M)
2α
123
Author's personal copy
F. Alouges et al.
We introduce only one further, final modification which implements the strategy delin-
eated in Remark 1: In order to maintain unconditional convergence we additionally
modify the second highest order term on the r.h.s. in the following way
k k
d ∇v · ∇ dx
2
(1 + ρ(k)) d 2 ∇v · ∇ dx,
2 2
Ω Ω
where ρ(k) → 0 as k → 0. Take note that for ρ decreasing at least linearly, quadratic
order is conserved. However, only in case that ρ is slightly sublinear, for example
ρ(k) = k| ln(k)|, do we in fact achieve unconditional convergence.
Adopting Algorithm 1, we arrive at the following scheme:
The appropriate choice of ρ and M can be inferred from our convergence result, see
Theorem 2.
Let us sum up: the new scheme replaces the search of v as solution to (11) by the
search of v as a solution to (24). Besides this substitution, the algorithm outlined in
Sect. 4 remains as before in the sense that the renormalization and the interpolation
w.r.t. time are left unchanged. Since Eq. (24) is linear in v, our algorithm is very
favorable in practice.
123
Author's personal copy
A convergent and precise finite element scheme
Before we state our theorem about the convergence let us explicitly make a statement
about its order.
Proposition 1 Consider a smooth (in space and time) solution m to (1) at time t = nk
and a semi-discrete time-approximation to m at time t + k on the basis of (24).
More precisely, given m at time t = nk determine v as a solution to the variational
formulation (24) with ρ(k) = 0 and M(k) sufficiently large and set
m(x, t) + kv(x, t)
m̃(x, t + k) = for all x ∈ Ω.
|m(x, t) + kv(x, t)|
Argument for Proposition 1 The proof is a direct consequence of the Taylor expansion
performed in (16).
Remark 2 The smoothness of solutions to (1) has been widely studied during the past
years. In general, the formation of singularities cannot be ruled out and we can usually
not assume that a solution to the initial value problem will be regular. Our statement
about the order of the approximation is thus only a first little step on the way to a proof
of the order of convergence, which is way beyond the scope of this paper. Let us insist
on the fact that, to the knowledge of the authors, there does not yet exist any numerical
scheme for LLG equations for which an order of convergence can be proven.
Proof of Theorem 2 As stated before, the proof consists of two main steps: estab-
lishing estimates which guarantee the existence of a sufficiently strong converging
subsequence, then proving that the latter converges indeed to a solution (which sat-
isfies the energy estimate). Several arguments in the proof have already been used in
e.g. [1,3,8], but we restate them for the self-consistency of the paper. We will need
the following classical estimate from elliptic regularity theory, namely
for all p ∈ (1, +∞) and a positive constant C which depends only on p.
123
Author's personal copy
F. Alouges et al.
Since we assume that the triangulation Th satisfies the angle condition (13) we have
that
2
2
∇mn+1 dx ≤ ∇(mn + kvn ) dx
Ω Ω
n 2
≤ ∇mn 2 dx + 2k ∇m · ∇v dx + k
n n 2 ∇v dx.
Ω Ω Ω
Ω
+2k (Hd (mn ) + Haniso (vn ) + Hext ) · vn dx
Ω
n 2
−k ρ(k) d 2
2 ∇v dx. (27)
Ω
We partially neglect the negative contributions on the r.h.s. of (28)—those which are
quadratic in vn —and use (25) to obtain
123
Author's personal copy
A convergent and precise finite element scheme
2
d2 ∇mn+1 dx ≤ d 2 ∇mn 2 dx − 2k ϕ M (m) vn 2 dx
Ω Ω Ω
n 2
+2k||H̄eff (m )|| L 2 ||v || L 2 − k 2 ρ(k)d 2
n n ∇v dx
Ω
2
≤ d2 ∇mn 2 dx − 2k ϕ M (m) vn dx
Ω Ω
n 2
+Ck||vn || L 2 − k ρ(k) d 2
2 ∇v dx, (29)
Ω
where the generic constant C depends on Q and |Ω|. Due to Young’s inequality, we
2
have that Ck||vn || L 2 (Ω) ≤ kβ||vn ||2L 2 (Ω) + kC
4β for β > 0. Using the uniform bound
α
ϕ M (m) ≥ β =
1 + k2 M
2 −1
N
N
n 2 n 2
d2
∇m dx + βk v dx + k ρ(k) d
2 2 ∇v dx
Ω n=0 Ω Ω
⎛ ⎞
≤ C ⎝T, d 2 |∇m0 |2 dx, β, Q, Haniso ⎠ (31)
Ω
From now on, most of the arguments follow the same line as in [1]. It holds that
mn+1 − mn
i i
≤ |vin |, for all n ≤ N , and i ∈ {1, · · · , Nh }.
k
Moreover, there exists c > 0 such that for all 1 ≤ p < +∞ and all φh ∈ Vh there
holds
1 p
p
||φh || L p (Ω) ≤ h d |φh (xih )| p ≤ c||φh || L p (Ω) , (32)
c
i
123
Author's personal copy
F. Alouges et al.
which implies
n+1
m − mn
≤ c2 ||vn || L 2 . (33)
k
L2
Hence we obtain from the energy estimate (31) using (33) the following bounds
Due to (34) and (35), there exist m̄ ∈ H 1 (ΩT ) and v ∈ L 2 (ΩT ) such that up to the
extraction of subsequences
−1 T
N
n 2
k ρ(k)
2 ∇v dx = kρ(k) ∇vh,k 2 dx ≤ C < +∞.
n=0 Ω 0 Ω
If ρ decreases linearly or faster we have to resort to the inverse estimate ||∇v|| L 2 (ΩT )
−1 → 0 when both (h, k) → 0.
h ||v|| L 2 (ΩT ) so that (39) holds true if kh
1
Preliminary estimates We want to prove that m̄ satisfies (7) and follow the strategy
of [1]. To begin with, we restate some further estimates from [1] and derive some
necessary statements about convergence. Observe that for all n = 0, · · · , J and all
t ∈ [nk, (n + 1)k)
n+1
m (x) − mn (x)
−
|mh,k (x, t) − mh,k (x, t)|
= (t − nk) ≤ k ∂t mh,k (x, t) .
k
Therefore
−
||mh,k − mh,k || L 2 (ΩT ) ≤ k ∂t mh,k L 2 (Ω →(h,k)→0 0,
T)
123
Author's personal copy
A convergent and precise finite element scheme
Moreover, on any tetrahedron K of Th , and for any u ∈ Mh , and any vertex xih of K ,
one has
2
|u(x)| − |u(xih )| ≤ Ch 2 |∇u|2 ,
−
(recall that ∇u is constant on K ), from which one deduces (since |mh,k (xih )| = 1)
2
− − 2
1 − |mh,k | dx ≤ Ch 2 ||∇mh,k || L 2 (Ω ) .
T
ΩT
1 2 n2
|min+1 − min − kvin | = |min + kvin | − 1 ≤ k |vi | , (40)
2
we derive
mn+1 − mn 1
i i n
− vi ≤ k|vin |2 .
k 2
General properties of interpolation operator Before we start with the penultimate step
of proving convergence, let us state some general properties of the nodal interpolation
operator which we repeatedly use in the sequel. Up to dimension three, there holds
for any function ϕ ∈ H 2 (Ω) ⊂ C 0 ()¯
Since the basis functions are linear on each triangle one can deduce from (41) that
− ˜ − Ih (m− × )||
˜ L 2 ([0,T ],H 1 ) ≤ Ch||m− || H 1 (Ω ) ||||W 2,∞ ,
||mh,k × h,k h,k T
(42)
123
Author's personal copy
F. Alouges et al.
− ˜ where
that Ih is the nodal interpolation, cf. (9), and test (24) with = Ih (mh,k × )
˜ ∞
∈ C0 (ΩT ) . After suitable integration in time we get
3
− − ˜ dx dt + − − ˜ dx dt
ϕ M (mh,k ) vh,k · Ih (mh,k × ) mh,k × vh,k · Ih (mh,k × )
ΩT T
k −
+ (1 + ρ(k)) d 2 ∇vh,k · ∇Ih (mh,k ˜ − Hd (vh,k ) · Ih (m− × )
× ) ˜
h,k
2
T
− ˜ dx dt
−Q(e · vh,k )(e · Ih (mh,k × ))
− − ˜ + Hd (mh,k ) · Ih (m− × )
˜
= − d 2 ∇mh,k · ∇Ih (mh,k × ) h,k
T
− ˜ + Hext · Ih (m− × )
˜ dx dt.
+Q(e · mh,k )(e · Ih (mh,k × )) h,k (43)
Our goal is to pass to the limit (k, h) → 0 in the latter equation (43) to recover the
LLG equation (10). As we shall see, the first and the third term on the l.h.s. and the
first term on the r.h.s. are a little bit subtle and have to be treated with caution. The
remaining contributions behave well under the established convergence; this is par-
ticularly due to the fact that Hd is L 2 -continuous. For the second contribution on the
l.h.s. one further uses that the L ∞ bound on m− improves (37) to strong convergence
in any L p with 1 < p < +∞.
Observing that |ϕ M | is uniformly bounded, and that |ϕ M − α| ≤ k 2M , as long as
−
k M → 0 for both (h, k) → 0 the strong convergence of mh,k is sufficient to conclude
that
− − ˜ dx dt →(h,k)→0 α ˜ dx dt. (44)
ϕ M (mh,k ) vh,k · Ih (mh,k × ) v · (m̄ × )
ΩT T
− − ˜ dx dt
˜ dx dt − α v · (m̄ × )
ϕ M (mh,k ) vh,k · Ih (mh,k × )
n
ΩT T
− − ˜ dx dt
˜ dx dt − α v · (m̄ × )
≤ ϕ M (mh,k ) vh,k · (mh,k × )
ΩT T
− − −
˜ − Ih (m × ))˜ dx dt .
+ ϕ M (mh,k ) vh,k · ((mh,k × ) (45)
h,k
ΩT
123
Author's personal copy
A convergent and precise finite element scheme
it suffices to show
− ˜ dx →(h,k)→0 0,
k
2 d2 (1 + ρ(k))∇vh,k · ∇(mh,k × ) (47)
T
123
Author's personal copy
F. Alouges et al.
Energy estimate We finally establish the energy estimate. From (27) we deduce that
∀n ∈ {0, · · · , Nh }
2
E(mn+1 ) − E(mn ) ≤ − k ϕ(mn ) vn dx + k H̄eff mn · vn dx
Ω Ω
k2 ∂ H̄eff n k2 n 2
+ (v ) · vn dx − ρ(k) d 2 ∇v dx
2 ∂m 2
Ω Ω
1
− H̄eff mn+1 + mn · (mn+1 − mn ) dx, (49)
2
Ω
n = H̄
eff (m ) for the remaining effective
cf. (3). Let us introduce the shorthand H̄eff n
field. We consider the contributions in (49) separately and start with the observation
that
1 n+1
k H̄eff · v dx −
n n
(H̄eff + H̄eff
n
) · (mn+1 − mn ) dx
2
Ω Ω
1 n+1
= H̄eff · (m
n n+1
− mn − kvn ) dx + (H̄eff − H̄eff
n
) · (mn+1 − mn ) dx.
2
Ω Ω
The stray-field contribution can be bounded with the help of (25) with p = 4. The
contributions in the second line on the r.h.s. of (49) are of higher order in k. The first
term can be easily bounded using Young’s inequality:
∂ H̄eff n
(v ) · v dx ≤ C||vn ||2L 2 .
n
(51)
∂m
Ω
123
Author's personal copy
A convergent and precise finite element scheme
where C denotes a generic constant. Here we made use of the classical Sobolev
embedding
||vn || L 4 ≤ C||∇vn || L 2 .
We are now ready to pass to the limit. Noticing once again that k ||∇vn || L 2 (ΩT ) is
uniformly bounded from (39) we derive that
n 2
E(m(N k)) − E(m(0)) + α v dx dt ≤ 0. (52)
ΩT
7 Numerical results
In this section, we report on two numerical experiments designed to test the scheme.
The first one is intended to confirm the order of the scheme (expected to be nearly
2); the second one investigates the effect of overdamping i.e. taking ρ as suggested in
Theorem 2. In both cases, we refer to numerical test cases that are well-known in the
physics literature of ferromagnetic materials (see for instance [12]).
In submicron Permalloy dots, the two main contributions to the effective field Heff (m)
are the exchange field Hex (m) and the demagnetizing field Hd (m), solution to (4). In
our finite element code “feeLLGood”, Hd (m) is computed by the FMM [14].
The first study mentioned above addresses the dynamical relaxation of the mag-
netization towards equilibrium in a flat cylindrical dot which is 200 nm wide and
20 nm thick. Typical material parameters are A = 1.0 × 10−11 Jm−1 and Ms =
8.0 × 105 Am−1 corresponding to an exchange length
d = (2 A μ−1 −2 1/2
0 Ms ) = 5 nm.
The damping factor α in the Landau–Lifschitz equation is set to 0.05, and we have
taken γ0 = 2.21 × 105 mA−1 s−1 [(cf. (1)].
The initial state is a vortex whose core is displaced in the x-direction. Writing in
complex variables and renormalizing the radius of the dot to 1, the initial magnetization
123
Author's personal copy
F. Alouges et al.
z−a
with the complex function f (z) = ci 1−az̄ and z = x + i y. Parameters a ∈ [0, 1)
and c ∈ R respectively specify the displacement of the vortex core from the sample’s
center and the characteristic size of the vortex. Our simulation uses a = c = 0.5. The
corresponding initial magnetization distribution is depicted in Fig. 2.
Notice that our construction ensures a distribution locally tangent to the circular
boundary. It is well known that the stray-field energy strongly penalizes non-tangential
configurations in thin films and thus, at least over short times, the nucleation of further
vortices. The simulations were performed on a tetrahedral mesh of 46,016 elements
(8,355 nodes) with an in-plane mesh-size h = 4nm < d. The final simulation time is
T = 10 ns, while the bound M [(cf. (20)] is set s.t. k M (2α)−1 = 0.1.
Figure 3 shows the time evolution of the average magnetization components and
monitors the convergence of the scheme as the time step tends to 0. At large time the
vortex turns as expected around the dot’s center at a characteristic frequency of f =
0.62 GHz in agreement with the analytical value given by Guslienko’s ferromagnetic
resonance model [12] within a physically acceptable error of about 10 %.
During its first rotation, the vortex core dilates, dissipating excess energy—
predominantly exchange energy. The rotation and core expansion combined altogether
cause fast magnetization variations, as observed in Fig. 3 at short times.
We performed simulations with a range of time steps to study the order of the
scheme. A log–log plot of the energy gap of the solution at the intermediate time
t = 0.7 ns and an estimation corresponding to k → 0 is presented in Fig. 4. The gap
follows a quadratic power law, confirming the presumed quadratic order.
123
Author's personal copy
A convergent and precise finite element scheme
Fig. 4 Energy gap versus time step following almost a quadratic law
We now consider a Permalloy dot with an elliptic 120 nm × 75 nm cross section and
2.5 nm thickness. It is discretized into 23,862 tetrahedral elements (4,732 nodes). The
material parameters are the same as in the preceding experiment.
Starting from a saturated magnetization state along the [110] direction, the system
relaxes to its equilibrium configuration: a symmetric distribution roughly aligned with
the long axis of the ellipse. Again, the magnetization tends to be tangent to the ellipse
boundary, in order to minimize the demagnetizing energy.
In order to test the effect of the (slight) overdamping induced by the parameter ρ,
we performed two simulations with the same time step k = 1 ps but with different
values for ρ:
1. ρ = 0, and
2. ρ(k) = τk | ln τk | ≈ 0.022, corresponding to the characteristic time τ = 250 ps.
123
Author's personal copy
F. Alouges et al.
Fig. 5 Time evolution of the average magnetization component along Oy with k = 1 ps obtained with
ρ = 0 and ρ ≈ 0.022
Fig. 6 Distribution of the y-component of the magnetization along the y-axis at t = 1ns for ρ = 0 (le f t)
and ρ ≈ 0.022 (right)
In Fig. 5, the evolution of the average magnetization seems almost identical (2 ns),
independent of ρ. A look at Fig. 6, however, reveals that the magnetization distributions
differ significantly, with an instability occurring in the case ρ = 0.
In Fig. 7, which shows the evolution of the energy, one clearly observes the stabiliz-
ing effect of ρ in the numerical scheme. This effect was observed in practice whatever
the time step used. For ρ = 0 it is necessary to drop down the time step (typically
below kc ≈ 0.5 ps in our case) to recover stability.
8 Summary
We started out with an introduction and discussion of the newly proposed tempo-
ral scheme. We proved its order 2 consistency to the equation, its uniform stability
and (weak) convergence. Numerical simulations on physical test cases show strong
evidence of good performance and confirm in practice the theoretical properties of
the algorithm stated in Theorem 2, namely the quadratic convergence and stabilizing
effect of ρ. A rigorous proof of the order 2 convergence is yet missing, though, due
123
Author's personal copy
A convergent and precise finite element scheme
Fig. 7 Total energy versus time for ρ = 0 (r ed) and ρ ≈ 0.022 (blue)
to the very strong nonlinearities in the equations such a result is clearly out of reach
for the time being.
Acknowledgments FA and JS both acknowledge support from the chair “Mathematical Modelling and
Numerical Simulation, F-EADS-Ecole Polytechnique-INRIA-F-X” and the ANR project ANR-08-BLAN-
0199. JS acknowledges support from the German Science Foundation through the Collaborative Research
Center 611 “Singular Phenomena and Scaling in Mathematical Models”.
References
1. Alouges, F.: A new finite element scheme for Landau-Lifschitz equations. Disc. Cont. Dyn. Syst. Ser.
S 1(2), 187–196 (2008)
2. Alouges, F.: A new algorithm for computing liquid crystal stable configurations: the harmonic mapping
case. SIAM J. Numer. Anal. 34(5), 1708–1726 (1997)
3. Alouges, F., Kritsikis, E., Toussaint, J.-C.: A convergent finite element approximation for the Landau–
Lifschitz–Gilbert equation. Physica B 407 (2012)
4. Alouges, F., Conti, S., DeSimone, A., Pokern, Y.: Energetics and switching of quasi-uniform states in
small ferromagnetic particles. M2AN Math. Model. Numer. Anal. 38(2), 235–248 (2004)
5. Alouges, F., Jaisson, P.: Convergence of a finite element discretization for the Landau–Lifshitz equa-
tions in micromagnetism. Math. Model. Method. Appl. Sci. 16(2), 299–316 (2006)
6. Alouges, F., Soyeur, A.: On global weak solutions for Landau–Lifchitz equations: existence and
nonuniqueness. In: Nonlinear analysis, theory, methods and applications 18(11), 1071–1084 (1992)
7. Bartels, S.: Stability and convergence of finite-element approximation schemes for harmonic maps.
SIAM J. Numer. Anal. 43(1), 220–238 (2005)
8. Bartels, S., Ko, J., Prohl, A.: Numerical analysis of an explicit approximation for the Landau–Lifshitz–
Gilbert equation. Math. Comp. 77, 773–788 (2008)
9. Bartels, S., Prohl, A.: Convergence of an implicit finite element method for the Landau–Lifshitz–Gilbert
equation. SIAM J. Numer. Anal. 44(4), 1405–1419 (2006)
10. Brown, W.F.: Micromagnetics. Wiley, New York (1963)
11. Cimràk, I.: A survey on the numerics and computations for the Landau–Lifshitz equation of micro-
magnetism, pp. 277–309. ACME, Netherlands (2007)
12. Guslienko et al: Eigenfrequencies of vortex state excitations in magnetic submicron-size disks. J. Appl.
Phys. 91, 8037–8040 (2002)
13. Hubert, A., Schäfer, R.: Magnetic domains. Springer, Berlin (1998)
123
Author's personal copy
F. Alouges et al.
14. Kritsikis, E., Toussaint, J.-C., Fruchart, O., Szambolics, H., Buda-Prejbeanu, L.: Fast computations of
magnetostatic fields by non-uniform fast Fourier transforms. Appl. Phys. Lett. 93, 132508 (2008)
15. Landau, L., Lifschitz, I.: On the theory of the dispersion of magnetic permeability in ferromagnetic
bodies. Phys. Zeitsch. der Sow. 8, 153–169 (1935)
16. Szambolics, H., Toussaint, J.-C., Buda-Prejbeanu, L., Alouges, F., Kritsikis, E., Fruchart, O.: Innovative
weak formulation for the LLG equation. IEEE Trans. Magn. 44(11), 3153–3156 (2008)
17. Vanselow, R.: About Delaunay triangulations and discrete maximum principles for the linear conform-
ing FEM applied to the Poisson equation. Appl. Math. 46(1), 13–28 (2001)
18. Visintin, A.: On Landau–Lishitz equations for ferromagnetism. Jpn. J. Appl. Math. 2, 69–84 (1985)
123