You are on page 1of 12

Physica D 350 (2017) 1–12

Contents lists available at ScienceDirect

Physica D
journal homepage: www.elsevier.com/locate/physd

Pattern formation on the free surface of a ferrofluid: Spatial dynamics


and homoclinic bifurcation
M.D. Groves a,b,∗ , D.J.B. Lloyd c , A. Stylianou d
a
Fachrichtung Mathematik, Universität des Saarlandes, Postfach 151150, 66041 Saarbrücken, Germany
b
Department of Mathematical Sciences, Loughborough University, Loughborough, LE11 3TU, UK
c
Department of Mathematics, University of Surrey, Guildford, GU2 7XH, UK
d
Institut für Mathematik, Universität Kassel, 34132 Kassel, Germany

article info abstract


Article history: We establish the existence of spatially localised one-dimensional free surfaces of a ferrofluid near onset
Received 20 October 2016 of the Rosensweig instability, assuming a general (nonlinear) magnetisation law. It is shown that the
Accepted 11 March 2017 ferrohydrostatic equations can be derived from a variational principle that allows one to formulate them
Available online 23 March 2017
as an (infinite-dimensional) spatial Hamiltonian system in which the unbounded free-surface direction
Communicated by J. Dawes
plays the role of time. A centre-manifold reduction technique converts the problem for small solutions
near onset to an equivalent Hamiltonian system with finitely many degrees of freedom. Normal-form
Keywords:
Ferrofluids
theory yields the existence of homoclinic solutions to the reduced system, which correspond to spatially
Localised patterns localised solutions of the ferrohydrostatic equations.
Spatial dynamics © 2017 Elsevier B.V. All rights reserved.
Centre-manifold reduction
Hamiltonian-Hopf bifurcation

1. Introduction Ferrofluids theory

Ferrofluids The first attempt at a theoretical explanation for the nucleation


of static cellular patterns was by Gailitis [10,11]. Gailitis postulated
a free energy for the system (a ferrofluid of infinite depth with a
The Rosensweig instability – a surface instability of a ferrofluid
linear magnetisation law) into which he substituted an Ansatz for
– has been of interest since the 1960s (see Cowley & Rosensweig [1]
a cellular free surface. The resulting equations for the unknown
and Rosensweig [2]). In an experiment, a vertical magnetic field
coefficients were then solved to find regions of existence for the
is applied to a static ferrofluid layer, and regular cellular patterns
various cellular patterns (stripes, squares and hexagons). This work
(typically hexagons) emerge on the fluid surface as the field
was extended by Friedrichs & Engel [12] to finite-depth ferrofluids.
strength is increased through a critical value (see Gollwitzer
Zaitsev & Shliomis [13] considered one-dimensional static
et al. [3,4] for recent experimental results). Recently, experiments
spatially periodic solutions to the ferrohydrostatic equations
have shown that spatially localised free-surface structures occur
near onset of the Rosensweig instability, again assuming a linear
in the hysteresis region between the flat state and the cellular
magnetisation law and infinite fluid depth. These results were
spatially periodic patterns (see Fig. 1, Richter [5], Richter & subsequently extended by Twombly & Thomas [14] to two-
Barashenkov [6] and Lloyd et al. [7] for experimental results, and dimensional spatially periodic free-surface solutions and finite-
Lavrova et al. [8] and Cao & Ding [9] for finite-element simulations). depth fluids. These local bifurcation theories were complemented
Despite this wealth of experimental and numerical evidence little by a formal normal-form analysis by Silber & Knobloch [15] (for
is known theoretically on the existence of localised solutions to the two-dimensional patterns and infinite depth), and there have
ferrohydrostatic equations. recently been attempts at deriving time-dependent amplitude
equations for spatially periodic free-surface patterns near onset
(see Bohlius et al. [16,17]).
∗ Corresponding author at: Fachrichtung Mathematik, Universität des Saarlandes, All current theoretical studies, whether rigorous or heuristic,
Postfach 151150, 66041 Saarbrücken, Germany. assume a linear relationship between the magnetisation M of the
E-mail address: groves@math.uni-sb.de (M.D. Groves). ferrofluid and the strength of the magnetic field H. The linearity of
http://dx.doi.org/10.1016/j.physd.2017.03.004
0167-2789/© 2017 Elsevier B.V. All rights reserved.
2 M.D. Groves et al. / Physica D 350 (2017) 1–12

Fig. 1. Radially symmetric localised patterns have been observed in experiments


by Richter [5] (reproduced with permission).

the magnetisation law M = M(H) allows one to replace a nonlinear


Fig. 2. Static localised patterns appear on the interface between a non-
equation for the magnetic potential in the fluid by Laplace’s
magnetisable fluid (white) and a ferrofluid (shaded) as the strength of a vertically
equation and greatly simplifies computations of coefficients in aligned magnetic field is varied.
bifurcation and normal-form theory. However, experiments show
that realistic magnetisation laws are nonlinear (e.g. see Lloyd φ in respectively the upper and lower fluids) follow from the
et al. [7] for a discussion of this point), emphasising the need to variational principle
develop a theoretical framework which includes general nonlinear   ∞  η(x) 
magnetisation laws.
δ −µ0 T (|∇φ|) − T (H ) dy dx

−∞ −D
Localised travelling water waves  ∞  D
1
− µ0 |∇φ ′ |2 dy dx
The ferrohydrostatic problem has similarities with the govern- −∞ η(x) 2
ing equations for travelling gravity–capillary surface and interfa-
 ∞  
1 1
cial water waves. One-dimensional spatially localised solutions to + (ρ − ρ )g η + µ0 H µ(H )
′ 2 2
µ(H ) − 1 η
−∞ 2 2
those problems (‘gravity–capillary solitary waves’) with fluids of  
finite depth have been studied intensively in the last twenty-five +σ 1 + ηx2 − 1
years using a technique known as the Kirchgässner reduction (see
Iooss [18], Groves & Wahlén [19] (surface waves) and Nilsson [20]  
+ µ0 µ(H )H φ(−D) − φ (D) dx = 0, ′
 
(interfacial waves) and the references therein). In this approach
one formulates the governing equations as an evolutionary sys-
tem in which the horizontal spatial direction plays the role of time where µ0 , g , σ denote respectively the magnetic permeability of
(‘spatial dynamics’); a centre-manifold reduction principle is used free space, acceleration due to gravity and coefficient of surface
to show that all spatially localised solutions solve a system of or- tension, ρ ′ and ρ are the densities of the upper and lower fluids,
dinary differential equations, whose solution set can in principle H is the strength of the applied magnetic field and
be determined. A helpful feature of this water-wave problem is its
|M(s)| s

variational structure, which leads to a Hamiltonian spatial dynam- µ(s) = 1 + , T ( s) = t µ(t ) dt ;
ics formulation of the problem; the Hamiltonian structure is in- s 0
herited by the reduced system of ordinary differential equations, the variations are taken with respect to η, φ ′ and φ satisfying
which can be treated by well-developed methods for Hamiltonian φ ′ |y=η = φ|y=η . (Variational (Neumann) boundary conditions at
systems with finitely many degrees of freedom, e.g. the Birkhoff y = ±D are chosen so that the reference state (η, φ ′ , φ) =
normal-form theory. (0, µ(H )Hy, Hy) – corresponding to a uniform magnetic field and
a flat surface – is a solution to the governing equations.) We regard
The present contribution the above variational functional as an action functional of the form

We consider two static immiscible perfect fluids in the regions Jf (η, χ ′ , χ , ηx , χx′ , χx ) dx,

S ′ := {(x, y) : η(x) < y < D}, where χ ′ , χ are suitably chosen perturbations of the reference
S := {(x, y) : −D < y < η(x)} states of the magnetic potentials and Jf is an appropriately
transformed version of the ‘Lagrangian’ integrand. Performing a
separated by the free interface {y = η(x)} (see Fig. 2); the upper
classical Legendre transform yields the desired spatial Hamiltonian
fluid is non-magnetisable, while the lower is a ferrofluid with a
formulation
general nonlinear magnetisation law (see Section 2 for a complete
mathematical formulation of the ferrohydrostatic problem). Using δ Hf δ Hf δ Hf
ηx = , ωx = − , χx′ = ,
spatial dynamics and the Kirchgässner reduction we present a δω δη δξ ′
rigorous existence theory for small-amplitude localised surface (1.1)
δ Hf δ Hf δ Hf
patterns to the ferrohydrostatic problem near onset of the ξx′ = − ′ , χx = , ξx = −
Rosensweig instability.
δχ δξ δχ
Our starting point is the observation that the governing of the ferrohydrostatic problem, where ω, ξ , ξ ′ are the momenta
equations (formulated in terms of the magnetic potentials φ ′ and associated with the coordinates η, χ , χ ′ and Hf is the Hamiltonian
M.D. Groves et al. / Physica D 350 (2017) 1–12 3

Existence theories for homoclinic solutions to (1.2) have been


given by Iooss & Pérouème [21] and Buffoni & Groves [22] under
the assumption that the coefficients c1 and c3 in the expansion
HNF = ε c1 |A|2 + ε ic2 (AB − AB) + c3 |A|4
+ ic4 |A|2 (AB − AB) − c5 (AB − AB)2
+ ε 2 c6 |A|2 + ε 2 ic7 (AB − AB) + · · ·
are respectively negative and positive. Iooss [18] and Iooss &
Fig. 3. A homoclinic solution to the spatial dynamics formulation (left, shown as a
trajectory in phase space) corresponds to a localised solution of the ferrohydrostatic Pérouème [21] establish the existence of two distinct symmetric
problem (right, sketch of the free surface). homoclinic solutions with asymptotic expansions
1/2
c3 ε

A(x) = ± − sech (−c1 ε)1/2 x eiβ0 qx + O(ε),
 
(see Section 3). Homoclinic solutions of (1.1) (solutions with
(η, ω, χ ′ , ξ ′ , χ , ξ ) → 0 as x → ±∞) are of particular interest c1
since they correspond to localised solutions of the ferrohydrostatic B(x) = O(ε)
problem (see Fig. 3). as ε → 0, while Buffoni & Groves [22] show that (1.2) has
The Rosensweig instability (the trivial solution becoming unsta- an infinite number of geometrically distinct homoclinic solu-
ble to Fourier modes ±eisx ) is associated with a Hamiltonian–Hopf tions which generically resemble multiple copies of one of the
bifurcation for (1.1): two pairs of simple, purely imaginary eigen- Iooss–Pérouème solutions. The free surface is given by the formula
values of our linearised Hamiltonian system become complex by η(x) ∼ 2 Re A(x) + O(ε), where the constant of proportionality is
colliding on the imaginary axis at the points ±is. Working in di- the first component of the eigenvector e, and Fig. 4 shows sketches
mensionless variables (see Section 2), one finds that ±is are double of the free surface corresponding to the various homoclinic solu-
eigenvalues if and only if s is a double root of the equation tions. Explicit formulae for the coefficients c1 and c3 are given in
some special cases in Section 5 (such formulae are unwieldy, and
s2 µ(1)(µ(1) − 1)2
     it appears in general more appropriate to calculate them numeri-
s̃ s cally for a specific choice of µ).
= (γ + s2 ) s̃ coth + sµ(1) coth ,
β β The results presented in this article are mathematically
rigorous. In order to make them accessible to as wide a readership
(with the assumption µ(1) + µ̇(1) > 0), where as possible we begin each of Sections 3 (spatial dynamics), 4
(centre-manifold reduction) and 5 (homoclinic bifurcation) with

µ(1) σ (ρ − ρ ′ )σ g
s̃ = s , β= , γ = . an informal exposition of the theory and a discussion of its
µ(1) + µ̇(1) µ0 H 2D µ20 H 4 computational aspects; the rigorous mathematics is presented in
the second part of each of these sections, specifically Section 3.2,
We therefore choose a value (β0 , γ0 ) of (β, γ ) for which this
Section 4.2 and Remark 5.2. The article is intended to be self-
equation has a pair of double roots and set (β, γ ) = (β0 , γ0 + ε), contained, although we omit the details of several lengthy
so that a Hamiltonian–Hopf bifurcation takes place as ε increases calculations with analogues in other papers as well as the proofs
through zero. In Section 4 we use a centre-manifold reduction of specific theorems available elsewhere.
principle to show that our Hamiltonian system near onset admits
a locally invariant manifold of the form
2. The ferrohydrostatic problem
{Ae + Bf + Ae + Bf + r̃ (A, B, A, B, ε)},
We consider two static immiscible perfect fluids in the regions
where e and f are suitably normalised generalised eigenvectors
at ε = 0 and r̃ (A, B, A, B, ε) satisfies r̃ = O(|(A, B, A, B)| S ′ := {(x, y) : η(x) < y < D},
|(ε, A, B, A, B)|). The flow on this manifold is described by the two- S := {(x, y) : −D < y < η(x)}
degree-of-freedom, reversible Hamiltonian system separated by the free interface {y = η(x)} (see Fig. 2). The upper,
ε ε non-magnetisable fluid has unit relative permeability and density
∂ H̃f ∂ H̃f
Ax = , Bx = − , (1.2) ρ ′ , while the lower is a ferrofluid with density ρ . We denote the
∂B ∂A magnetic and induction fields in the fluids by respectively H′ , H
where and B′ , B, and suppose that the relationships between them are
given by the identities
H̃fε (A, B, A, B) = Hfε (Ae + Bf + Ae + Bf + r̃ (A, B, Ã, B, ε))
B′ = µ0 H′ , B = µ0 (H + M(H)),
(the superscript is added to emphasise the dependence of the
where µ0 is the magnetic permeability of free space and M is the
Hamiltonian upon ε ). Homoclinic solutions (A, B) to (1.2) (solu-
(prescribed) magnetic intensity of the ferrofluid. We suppose that
tions with A(x), B(x) → 0 as x → ±∞) generate homoclinic solu-
tions to (1.1) and hence localised solutions to the ferrohydrostatic H
M(H) = m(|H|)
problem. |H|
According to the Birkhoff normal-form theory (Section 5) we
where m is a (prescribed) nonnegative function, so that in
can select the coordinates A, B, A, B so that the reduced Hamiltonian
particular M and H are collinear. According to Maxwell’s equations
takes the form
the magnetic and induction fields are respectively irrotational
H̃fε (A, B) = iβ0 q(AB − AB) + |B|2 and solenoidal, and introducing magnetic potential functions φ ′ , φ
with H′ = −∇φ ′ , H = −∇φ , we therefore find that
+ HNF (|A|2 , i(AB − AB), ε)
+ O(|(A, B)|2 |(ε, A, B)|n0 ), 1φ ′ = 0, ∇ · (µ(|∇φ|)∇φ) = 0, (2.1)
in which
where HNF is a real polynomial of order n0 + 1 satisfying
m(s)
HNF (|A|2 , i(AB − AB), ε) = O(|(A, B)|2 |(ε, A, B)|). µ(s) = 1 +
s
4 M.D. Groves et al. / Physica D 350 (2017) 1–12

Fig. 4. Symmetric unipulse patterns (left and centre) co-exist with an infinite family of multipulse patterns (right).

is the magnetic permeability of the ferrofluid relative to that of The constant b is selected so that (η, φ ′ , φ) = (0, µ(H )Hy, Hy)
free space. (Here, and in the remainder of this paper, equations is a solution to (2.1), (2.3) and (2.4) (corresponding to a uniform
for ‘primed’ and ‘non-primed’ quantities are supposed to hold in magnetic field and a flat surface); we therefore set b = −µ0 T (H )−
respectively S ′ and S.) We assume that µ is a smooth function of s µ0 H 2 µ(H )( 12 µ(H ) − 1). Finally, choosing compatible Neumann
near unity which satisfies µ(1) + µ̇(1) > 0, where the dot denotes boundary conditions
differentiation with respect to s.
The ferrohydrostatic Euler equations are given by φy′ |y=D = µ(H )H , µ(|∇φ|)φy |y=−D = µ(H )H (2.5)
ensures that the system of Eqs. (2.1), (2.3)–(2.5) has a variational
−∇(p′ + ρ ′ gy) = 0,
structure: they follow from the formal variational principle
µ0 (M · ∇) H − ∇(p⋆ + ρ gy) = 0   ∞  η(x) 
δ −µ0 T (|∇φ|) − T (H ) dy dx

(Rosensweig [2, Section 5.1]), where g is the acceleration due to
gravity, p′ is the hydrodynamic pressure in the upper fluid and p⋆ −∞ −D
∞  D
is the composite pressure in the lower fluid. The calculation 1
− µ0 |∇φ ′ |2 dy dx
 |H| 
2
−∞ η(x)
(M · ∇) H = |M|∇(|H|) = ∇ |M(t )| dt  ∞  
0 1 1
+ (ρ − ρ ′ )g η2 + µ0 H 2 µ(H ) µ(H ) − 1 η
shows that these equations are equivalent to −∞ 2 2
 
 |H|
− (p′ + ρ ′ gy) = b′0 , µ0 m(t ) dt − (p⋆ + ρ gy) = b0 , +σ 1 + ηx2 − 1
0
 
(2.2)
+ µ0 µ(H )H φ(−D) − φ ′ (D) dx = 0,
 
where b′0 , b0 are constants.
The magnetic boundary conditions at {y = η(x)} are
where the variations are taken with respect to η, φ ′ and φ satisfying
H · t = H · t,

B · n = B · n,

φ ′ |y=η = φ|y=η .
The next step is to introduce dimensionless variables
where
(1, ηx )T (−ηx , 1)T µ0 H 2 µ0 H
t=  , n=  (x̂, ŷ) = (x, y), φ̂ := φ,
σ σ
1 + ηx2 1 + ηx2
µ0 H ′ µ0 H 2
are the tangent and normal vectors to the interface; it follows that φ̂ ′ := φ, η̂ := η
σ σ
φ ′ − φ = 0, φn′ − µ(|∇φ|)φn = 0 (2.3) and functions
 s
for y = η(x). The ferrohydrostatic boundary condition is given by µ̂(s) := µ(Hs), T̂ (s) :=
1
T (Hs) = t µ̂(t ) dt .
⋆ µ0 H 2
0
p + (M · n) = p + 2σ κ 2 ′
2 Writing (η̂, φ̂ ′ , φ̂) = (η̂, ψ̂ ′ + µ(1)y, ψ̂ + y) (so that (η̂, ψ̂ ′ , ψ̂) =
(Rosensweig [2, Section 5.2]), in which σ > 0 is the coefficient of (0, 0, 0) is the ‘trivial’ solution), we find that
surface tension and
ηxx 1ψ ′ = 0, ∇ · (µ (|∇(ψ + y)|) ∇(ψ + y)) = 0 (2.6)
2κ = −
(1 + ηx2 )3/2 with boundary conditions
ψy′  = 0, µ (|∇(ψ + y)|) (ψy + 1) − µ(1) = 0
 
is the mean curvature of the interface. Using (2.2), we find that y=
1
y=−
1
β β
|H|
µ0

(2.7)
µ0 m(t ) dt + (M · n)2 − (ρ − ρ ′ )g η
0 2 and
σ ηxx
+ + b = 0, ψ − ψ ′ − (µ(1) − 1) η = 0, (2.8)
(1 + ηx2 )3/2
µ (|∇(ψ + y)|) (ψ + y)n − (ψ + µ(1)y)n = 0, ′
(2.9)
where b = b′0 − b0 , or equivalently
1 
µ0
2
T (|∇(ψ + y)|) − ∇ ψ ′ + µ(1)y 
µ0 T (|∇φ|) − |∇φ ′ |2 − µ0 µ(|∇φ|) − 1 φn (φn − φn′ )
 
2
2 
σ ηxx + 1 + ηx ψy + µ(1) ψ ′ + µ(1)y n
 ′  
2
−(ρ − ρ )g η + ′
+ b = 0, (2.4)
(1 + ηx2 )3/2 
− 1 + ηx2 µ (|∇(ψ + y)|) ψy + 1 (ψ + y)n − γ η
 
where
ηxx
s
 
1

T (s) = t µ(t ) dt . − T (1) − µ(1) µ(1) − 1 + =0 (2.10)
0
2 (1 + ηx2 )3/2
M.D. Groves et al. / Physica D 350 (2017) 1–12 5

for y = η(x), where on y = 0; here


(ρ − ρ ′ )gD σ
 
α= , β= , γ = αβ µ⋆ = µ (χx − ηx K1 χy )2 + (K2 χy + 1)2 ,
µ0 H 2 µ0 H 2 D
and the hats have been dropped for notational simplicity. We use  
β and γ as parameters, noting that the limit β → 0 corresponds T⋆ = T (χx − ηx K1 χy )2 + (K2 χy + 1)2
to fluids of infinite depth.
Finally, note that Eqs. (2.6)–(2.10) follow from the formal and
variational principle δ I = 0, where
1 − βy 1 + βy
 ∞  η(x)  K1′ = , K1 = ,
I (η, ψ ′ , ψ) := − T (|∇ (ψ + y) |) − T (1) dy dx 1 − βη 1 + βη

1
−∞ − 1 1
β
K2′ = , K2 = .
 ∞  1
β 1 1 − βη 1 + βη
|∇ ψ ′ + µ(1)y |2 dy dx
 

−∞ η(x) 2
Observe that Eqs. (3.1)–(3.6) follow from the new variational
γ
 ∞    
1 principle δ J = 0, where
+ η + µ(1)
2
µ(1) − 1 η + 1+η −1
2
x
−∞ 2 2  ∞  0
1 ⋆
J (η, χ , χ ) := −

T − T (1) dy dx

    
+ µ(1) ψ − β1 − ψ ′ 1
β
dx −∞ −
1
β
K2

1
and the variations are taken with respect to η, ψ ′ and ψ with ∞
 
β 1  ′ 2  2 
ψ ′ (0) = ψ(0) − (µ(1) − 1)η. − ′ χx − K1′ ηx χy′ + K2′ χy′ + µ(1) dy dx
−∞ 0 2K2
∞
γ
    
3. Spatial dynamics 1
+ η2 + µ(1) µ(1) − 1 η + 1 + ηx2 − 1
−∞ 2 2
3.1. Formulation as a spatial Hamiltonian system     
+ µ(1) ψ − β1 − ψ ′ β1 dx (3.7)
The first step is to use the ‘flattening’ transformation
 y − η(x)
 , y ≥ η(x), and the variations are taken in η, χ ′ and χ satisfying the side
1 − βη(x)

Y = constraint
y − η(x)
 , y < η(x), χ ′ (0) = χ (0) − (µ(1) − 1)η (3.8)
1 + βη(x)

to map the variable fluid domains S ′ and S into fixed strips R × (the functional J is obtained from I by ‘flattening’). We exploit this
(0, β1 ) and R × (− β1 , 0) and the free interface {y = η(x)} into variational principle by regarding J as an action functional of the
{y = 0}. Replacing the symbol Y by y for notational simplicity, we form
find that the corresponding ‘flattened’ variables

J = Jf (η, χ ′ , χ , ηx , χx′ , χx ) dx,
χ ′ (x, Y ) = ψ ′ (x, y), χ (x, Y ) = ψ(x, y)
satisfy the equations in which J is the integrand on the right-hand side of Eq. (3.7),
and deriving a canonical Hamiltonian formulation of (3.1)–(3.6) by
χxx − K1 ηxx χy − 2K1 ηx χxy + K1 ηx χyy
′ ′ ′ ′ ′ ′2 2 ′
means of the Legendre transform. To this end, let us introduce new
+K2′2 χyy

− 2β K1′ K2′ ηx2 χy′ = 0, (3.1) variables ω, ξ ′ and ξ by the formulae

µ χxx − K1 ηxx χy − 2K1 ηx χxy + K1 ηx χyy + K2 χyy + 2β K1 K2 ηx χy
2 2 2 2
δ Jf
   0
K1 ⋆
ω= = µ (χx − K1 ηx χy )χy dy
+ χx − ηx K1 χy , K2 χy + 1 · (µ⋆x − ηx K1 µ⋆y , K2 µ⋆y ) = 0 δηx 1 K2
 
(3.2) −
β
with boundary conditions 1
K1′ ηx

β

µ (K2 χy + 1) − µ(1) = 0, K2 χ y + (χx′ − K1′ ηx χy′ )χy′ dy +  ,
 ′

′
=0 (3.3) K2′
y=−
1
β
y=
1
β 0 1 + ηx2

and δ Jf 1
ξ′ = = − ′ (χx′ − K1′ ηx χy′ ),
χ − χ − (µ(1) − 1) η = 0,

(3.4) δχx′ K2
δ Jf
 ⋆ ⋆ ⋆
1 + ηx (µ K2 χy − K2 χy ) − ηx (µ χx − χx ) + µ − µ(1) = 0,
2 ′ ′ ′ 1

ξ= = − µ⋆ (χx − K1 ηx χy )
(3.5) δχx K2
1 and define the Hamiltonian function by
T ⋆ − χx′2 + 1 + ηx2 µ(1)K2′ χy′ − µ⋆ K2 χy
  
2
  Hf (η, ω, χ ′ , ξ ′ , χ , ξ )
 1  ′ ′ 2
K 2 χ y − µ⋆ K 2 χ y
2
+ 1 + ηx2
 
1
2  0 
β
= ξ χy dy + ξ ′ χx′ dy + ωηx
− ηx (µ(1)χx′ − µ⋆ χx ) + µ K2 χy ηx χx − K2 χy − 1 ⋆
 
1
− 0
β
ηxx
− γ η − T (1) + µ(1) + =0 (3.6)
− J (η, χ ′ , χ , ηx , χx′ , χx )
(1 + ηx2 )3/2
6 M.D. Groves et al. / Physica D 350 (2017) 1–12

1 K2 W ξ (0)
+ µĎ |y=0 K2 χy (0) + 1 − µ(1)
  
β 1 ′ ′2 √
= K2 (χy − ξ ′2 ) dy 1− W2
0 2
K2′ W ξ ′ (0)
 0

1 Ď  K2
 − √ − K2′ χy′ (0) = 0, (3.18)
+ T − T (1) − Ď ξ 2 dy 1 − W2

1 K2 µ 1 K2 W χy (0)
β
K2 ξ (0) − √ − K2′ ξ ′ (0)
  γ
  µĎ |y=0 1 − W2
− µ(1) χ (0) + χ − β1 − η2 + 1 − 1 − W 2 , (3.9)
2 K2′ W χy′ (0) W
+ √ + (µ(1) − 1) √ = 0. (3.19)
in which 1− W2 1 − W2
1
 0 
β The first three of these equations arise from the integration by parts
W =ω+ K1 ξ χy dy + K1′ ξ ′ χy′ dy, necessary to compute (3.13) and (3.15) subject to the constraint
1
− 0 (3.8), while the fourth is the compatibility condition which ensures
β
  that χx′ (0) = χx (0) − (µ(1) − 1)ηx . Note further that our
µĎ = µ ν −1 (K2 ξ ; K2 χy + 1)2 + (K2 χy + 1)2 , equations are reversible, that is invariant under the transformation
  (η, ω, χ ′ , ξ ′ , χ , ξ )(x) → R(η, ω, χ ′ , ξ ′ , χ , ξ )(−x), where the
Ď reverser is defined by
T =T ν (K2 ξ ; K2 χy + 1) + (K2 χy + 1)
− 1 2 2

R(η, ω, χ ′ , ξ ′ , χ , ξ ) = (η, −ω, χ ′ , −ξ ′ , χ , −ξ ).


and√ν(·; t ) : R → R is given by the formula ν(s; t ) =
Our equations are also invariant under the transformation
sµ( s2 + t 2 ), where t is a parameter whose value is near unity. χ ′ → χ ′ + c, χ → χ + c for any constant c. To eliminate
(Using the calculations ν(0, 1) = 0 and ∂1 ν(0, 1) = µ(1) > 1, this symmetry it is convenient to replace (ξ ′ , χ ′ , ξ , χ ) with new
one finds from the inverse-function theorem that ν is invertible ′
variables (χ ′ , ξ , χ , ξ , χ̂ , ξ̂ ), where
for (s, t ) near (0, 1) (see Section 3.2).)

Hamilton’s equations are given explicitly by χ ′ := χ ′ − χ̂ , χ = χ − χ̂ , ξ := ξ ′ − ξ̂ , ξ = ξ − ξ̂ ,
δ Hf W and
ηx = = √ , (3.10)
δω 1 − W2

1

  0
1 β
δ Hf χ̂ := χ ′ dy + χ dy ,
 0 
ωx = − = −β 1 T Ď − T (1) − µĎ K2 χy (K2 χy + 1) dy
 
2 1
0 −
δη −
β
β  
1
  0
1 1 β
ξ̂ := ξ dy +

ξ dy .

β 1 ′2 ′2 
+β K2 (ξ − χy′2 ) dy 2 0 −
1
2 β
0

βW 0 This transformation leads to a new canonical Hamiltonian system
+ √ K1 K2 ξ χy dy with Hamiltonian
1
1− W2 −
β ′
 H f (η, ω, χ ′ , ξ , χ̂ ′ , ξ̂ ′ , χ , ξ , χ̂ , ξ̂ )
1

= Hf (η, ω, χ ′ + χ̂ , ξ + ξ̂ , χ + χ̂ , ξ + ξ̂ )

β
− K1′ K2′ ξ ′ χy′ dy

0
= Hf (η, ω, χ ′ , ξ + ξ̂ , χ , ξ + ξ̂ )
W and additional constraints
− √ (µ(1)K2′ ξ ′ (0) − K2 ξ (0))
1 − W2  1  0  1  0
β β ′
Ď
+ µ |y=0 (K2 χy (0) + 1) χ ′ dy + 1
χ dy = 0, ξ dy + 1
ξ dy = 0.
0 − 0 −
β β
− µ(1)(K2′ χy′ (0) + 1) + γ η, (3.11)
(3.20)
δ Hf WK1′ χy′
χx′ = = −K2′ ξ ′ + √ , (3.12) Observe that χ̂ is a cyclic variable whose conjugate ξ̂ is a conserved
δξ ′ 1 − W2 quantity; we proceed in standard fashion by setting ξ̂ = 0,

δ Hf W (K ′ ξ ′ )y considering the equations for (η, ω, χ ′ , ξ , χ , ξ ), and recovering
ξx′ = − ′ = K2′ χyy

+√ 1 , (3.13)
χ̂ by quadrature. Dropping the bars for notational simplicity, one
δχ 1 − W2
finds that Hamilton’s equations for the reduced system are
δ Hf K2 WK1 χy
χx = = − Ďξ + √ , (3.14) W
δξ µ 1 − W2 ηx = √ , (3.21)
1 − W2
δ Hf W (K1 ξ )y
ξx = − = (µĎ (K2 χy + 1))y + √ 0

(3.15)
δχ ωx = −β T Ď − T (1) − µĎ K2 χy (K2 χy + 1) dy
 
1 − W2 1

β
and are accompanied by the side constraint (3.8) and further  1
boundary conditions β 1 ′2 ′2
+β K2 (ξ − χy′2 ) dy
 
0 2
χy′ 1
β
= 0, (3.16) 
1

βW 0
 
β
− √ K1 K2 ξ χy dy + K1 K2 ξ χy dy
′ ′ ′ ′
   
µĎ | K2 χy − β1 + 1 − µ(1) = 0,

1 (3.17) 1 − W2 −
1
0
y=−
β β
M.D. Groves et al. / Physica D 350 (2017) 1–12 7

W To apply this definition to the Hamiltonian system derived in


+ √ (µ(1)K2′ ξ ′ (0) − K2 ξ (0))
1 − W2 Section 3.1, we introduce the Hilbert spaces
+ µĎ |y=0 (K2 χy (0) + 1) − µ(1)(K2′ χy′ (0) + 1) + γ η, (3.22)     
WK1 χy ′ ′ Ms = (η, ω, χ ′ , ξ ′ , χ , ξ ) ∈ R × R × H s+1 0, β1 × H s 0, β1
χx′ = −K2′ ξ ′ + √
1 − W2    

 1 × H s+1 − β1 , 0 × H s − β1 , 0 :
WK1′ χy′

1 β
−  −K2′ ξ ′ + √ dy χ ′ (0) = χ (0) − (µ(1) − 1)η,
2 0 1 − W2
 1  0
  β
WK1 χy
 0 
K2 χ ′ dy + 1 χ dy = 0,
+ − Ďξ + √ dy , (3.23)

1 µ 1 − W2
0 −
β
β
1
W (K1 ξ )y 0
  
β
ξx′ = K2′ χyy

+√ , (3.24) ξ ′ dy + ξ dy = 0 , s ≥ 0,
1 − W2 0 −
1
β
K2 WK1 χy
χx = − Ď ξ + √ and let N1 be a neighbourhood of the origin in M1 such that |W | <
µ 1 − W2

 1 1, |η| < β1 and K2 χy (y) + 1 ∈ Λ2 , K2 ξ (y) ∈ Λ3 for each y ∈ [0, 1]
WK1′ χy′

1 β
−  −K2′ ξ ′ + √ dy (recall that H 1 (− β1 , 0) is continuously embedded in C [− β1 , 0]).
2 0 1 − W2 Observe that N1 is a manifold domain of M0 , while the formula
Ωf ((η1 , ω1 , χ1′ , ξ1′ , χ1 , ξ1 ), (η2 , ω2 , χ2′ , ξ2′ , χ2 , ξ2 ))
 
WK1 χy
 0 
K2
+ − Ďξ + √ dy , (3.25)

1 µ 1 − W2  1
β
β
W (K1 ξ )y
= ω2 η1 − η2 ω1 + (ξ2′ χ1′ − χ2′ ξ1′ ) dy
ξx = (µĎ (K2 χy + 1))y + √ , (3.26) 0
1 − W2 0

+ 1
(ξ2 χ1 − χ2 ξ1 ) dy,
with constraints (3.8), (3.20) and boundary conditions −
β
(3.16)–(3.19); the quantity χ̂ is recovered by quadrature from the
equation defines a weakly nondegenerate bilinear form M0 × M0 → R
 and hence a constant symplectic 2-form TM0 × TM0 → R (its
1 closure follows from the fact that it is constant). Furthermore, the
WK1′ χy′
  
1 β
χ̂x =  −K 2 ξ + √
′ ′
dy function Hf given by (3.9) belongs to C ∞ (N1 , R), so that the triple
2 0 1 − W2 (M0 , Ωf , Hf ) is a Hamiltonian system.
  Definition 3.3. Consider a Hamiltonian system (M , Ω , H ), where
0
WK1 χy
 
K2
+ − Ďξ + √ dy . (3.27) H ∈ C ∞ (N , R) and N is a manifold domain of M. Its Hamiltonian

1 µ 1 − W2 vector field vH with domain D(vH ) ⊆ N is defined as follows. The
β
point n ∈ N belongs to D(vH ) with vH |n := v̂ ∈ TM |n if and only if
Notice that these equations are also reversible (with respect to the
same reverser R). Ω |n (v̂, v) = dH |n (v)
for all tangent vectors v ∈ TM |n (by construction dH |n ∈ T ∗ N |n
admits a unique extension dH |n ∈ T ⋆ M |n ). Hamilton’s equations
3.2. Functional-analytic basis
for (M , Ω , H ) are the differential equations

We begin with a precise statement of the invertibility of the ux = vH (u) := vH |u

√ ν(·; t ) : R → R defined by the formula ν(s; t ) =


function which determine the trajectories u ∈ C 1 (R, M0 ) ∩ C (R, N1 ) of its
sµ( s2 + t 2 ). Hamiltonian vector field.
Applying the criterion in this definition to the ferrofluid
Proposition 3.1. There exist open neighbourhoods Λ1 , Λ3 of the Hamiltonian system (M0 , Ωf , Hf ), one finds that
origin and Λ2 of unity in R such that ν(·, t ) : Λ1 → Λ3 is a D(vHf ) = {(η, ω, χ ′ , ξ ′ , χ , ξ ) ∈ N1 : B(η, ω, χ ′ , ξ ′ , χ , ξ ) = 0},
bijection for each t ∈ Λ2 . Furthermore ν ∈ C ∞ (Λ1 × Λ2 ) and
where B ∈ C ∞ (N1 , R4 ) is defined by the left-hand sides of
ν −1 ∈ C ∞ (Λ1 × Λ3 ). Eqs. (3.16)–(3.19), and that Hamilton’s equations are given ex-
plicitly by (3.10)–(3.15). Observe that these equations constitute
Next we recall the differential-geometric definitions of a a quasilinear evolutionary system in the phase space M0 with non-
Hamiltonian system and Hamilton’s equations for its associated linear boundary conditions (3.16)–(3.19); its right-hand side is a
vector field. smooth mapping N1 → M0 .
It remains to confirm the relationship between a solution
to Hamilton’s equations for (M0 , Ωf , Hf ) and a solution to the
Definition 3.2. A Hamiltonian system consists of a triple (M , Ω ,
‘flattened’ ferrohydrostatic problem (3.1)–(3.6). Suppose that
H ), where M is a manifold, Ω : TM × TM → R is a closed, (η, ω, χ ′ , ξ ′ , χ , ξ ) is a smooth solution of Hamilton’s equations
weakly nondegenerate bilinear form (the symplectic 2-form) and and compute χ̂ from (3.27) by quadrature. An explicit calculation
the Hamiltonian H : N → R is a smooth function on a manifold shows that the variables η̃, χ̃ ′ , χ̃ given by η̃(x) = η(x), χ̃ ′ (x, z ) =
domain N of M (that is, a manifold N which is smoothly embedded χ ′ (x)(z ) + χ̂ (x)(z ), χ̃ (x, z ) = χ (x)(z ) + χ̂ (x)(z ) solve (3.1)–(3.6)
in M and has the property that TN |n is densely embedded in TM |n (see Buffoni, Groves & Toland [23, Theorem 2.1] for a discussion of
for each n ∈ N). this procedure in the context of water waves).
8 M.D. Groves et al. / Physica D 350 (2017) 1–12

4. Centre-manifold reduction

4.1. Reduction to a two-degree of freedom Hamiltonian system

We now set (β, γ ) = (β0 , γ0 + ε), where the values of β0 ,


γ0 are appropriately chosen and fixed (see below) and ε plays the
role of a bifurcation parameter. The Hamiltonian formulation of our
ferrohydrostatic problem is accordingly written as

ux = f ε (u), (4.1)
ε
where u = (η, ω, χ , ξ , χ , ξ ) and f is given by the right-
′ ′

hand side of (3.21)–(3.26), with linear constraints (3.8), (3.20) and


nonlinear boundary conditions
B(u) = 0, (4.2)
where B is given by the left-hand sides of Eqs. (3.16)–(3.19).
Similarly, we denote the Hamiltonian (3.9) with this parameter
choice by Hfε .
Fig. 5. Purely imaginary eigenvalues as functions of the parameters β0 and α0 . A
The corresponding linearised system is Hamiltonian–Hopf bifurcation takes place at points of the curve C , and homoclinic
solutions exist in the shaded region.
ux = Lu,
where (0 is not an eigenvalue of L). This equation has either zero, one
or two pairs ±iq of solutions (see Fig. 5). A Hamiltonian–Hopf
 ω 
bifurcation takes place at points of the curve
 (µ(1) + µ̇(1)) τy (0) − µ(1)τy (0) + γ0 η 

η   0  1  C = {(βHH (q), αHH (q)) : q ∈ (0, ∞)}
 −ζ ′ + 1
 1 β0 ′ 
ω ζ dy + ζ dy 
in the (β0 , α0 )-plane: at the point (βHH (q), αHH (q)) two pairs of
τ   2µ(1) − 1 2 0
 ′  
β0  , (4.3) simple, purely imaginary eigenvalues become complex by colliding

L =
ζ ′   τyy

 
on the imaginary axis at ±iβ0 q and forming two Jordan chains of

τ
   
 0  1 length 2. Here
β0 ′
 
1 1 1
ζ ζ+ ζ ζ
 
− dy + dy
 µ(1) 2µ(1) − 1 2 0  µ(1)(µ(1) − 1)2
β0 βHH (q) =
(µ(1) + µ̇(1))τyy 2(q̃ coth q̃ + qµ(1) coth q)
with linear constraints (3.8), (3.20) and boundary conditions µ(1)(µ(1) − 1)2 (q̃2 cosech2 q̃ + q2 µ(1)cosech2 q)
+ ,
Bl (u) = 0, where 2(q̃ coth q̃ + qµ(1) coth q)2
η
  q2 µ(1)(µ(1) − 1)2
τy′ αHH (q) =
 1

β0 2(q̃ coth q̃ + qµ(1) coth q)
 ω′     
τ   (µ(1) + µ̇(1))τy − β10 
q2 µ(1)(µ(1) − 1)2 (q̃2 cosech2 q̃ + q2 µ(1)cosech2 q)
Bl  ′  =  . (4.4) .
 

ζ

(µ( 1 ) + µ̇(1 ))τ (0) − τ ′
( 0) −
 y y  2(q̃ coth q̃ + qµ(1) coth q)2
τ
  
1 
ζ ζ (0) − ζ (0) + (µ(1) − 1)ω

This eigenvalue collision is associated with the bifurcation of
µ(1) a branch of homoclinic solutions into the region with complex
An explicit calculation shows that a non-zero complex number λ eigenvalues (the shaded region in Fig. 5); we therefore set
is an eigenvalue of the linear problem (that is, the linear problem (β0 , α0 ) = (βHH (q), αHH (q)) (and γ0 = α0 β0 ) for some q ∈ (0, ∞).
admits a solution of the form u(x) = exp(λx)w with w ̸= 0) if and Choose e, f such that
only if λ = β0 σ , where
Le = iβ0 qe, Le = −iβ0 qe,
(µ(1) − 1)2 (µ(1) + µ̇(1))σ σ̃ sin σ sin σ̃ (L − iβ0 qI )f = e, (L + iβ0 qI )f = e,
= (σ 2 β0 − α0 ) σ sin σ cos σ̃ + (µ(1) + µ̇(1))σ̃ sin σ̃ cos σ
 
where Re = e, Rf = −f , the ‘symplectic products’ Ω (e, f ) and
and Ω (f , e) are respectively 1 and −1, and the symplectic products of
 all other combinations are zero (note that Ω acts bilinearly on pairs
µ(1) of complex vectors).
σ̃ = σ ;
µ(1) + µ̇(1)
Theorem 4.1. The Hamiltonian formulation of our ferrohydrostatic
furthermore 0 is not an eigenvalue of L. In particular, L has a finite problem admits a locally invariant manifold of the form {Ae + Bf +
number of purely imaginary eigenvalues. Ae + Bf + r̃ (A, B, A, B, ε)}, where r̃ (A, B, A, B, ε) satisfies r̃ =
According to this calculation, a purely imaginary number σ = is O(|(A, B, A, B)||(ε, A, B, A, B)|). The flow on this manifold is described
is an eigenvalue of L if and only if s = β0 q, where by the two-degree-of-freedom Hamiltonian system
q2 µ(1)(µ(1) − 1)2 = (α0 + q2 β0 ) q̃ coth q̃ + qµ(1) coth q
 
∂ H̃fε ∂ H̃fε
Ax = , Bx = − , (4.5)
and ∂B ∂A

µ(1) where
q̃ = q
µ(1) + µ̇(1) H̃fε (A, B, A, B) = Hfε (Ae + Bf + Ae + Bf + r̃ (A, B, Ã, B, ε)),
M.D. Groves et al. / Physica D 350 (2017) 1–12 9

which is reversible with reverser R(A, B) = (A, −B). In particular, (iv) M̃ λ is a symplectic submanifold of M and the flow determined by
(i) all small bounded solutions of (4.1), (4.2) (with constraints (3.8), the Hamiltonian system (M̃ λ , Ω̃ λ , H̃ λ ), where the tilde denotes
(3.20)) take the form restriction to M̃ λ , coincides with the flow on M̃ λ determined by
(M , Ω λ , H λ ). The reduced Eq. (4.8) is reversible and represents
u(x) = A(x)e + B(x)f + A(x)e + B(x)f Hamilton’s equations for (M̃ λ , Ω̃ λ , H̃ λ ).
+ r̃ (A(x), B(x), Ã(x), B̃(x), ε), (4.6)
Mielke’s theorem cannot be applied directly to the Hamiltonian
where (A(x), B(x)), x ∈ R solves (4.5); system (M0 , Ωf , Hfε ) because of the nonlinear boundary conditions
(ii) any small, bounded solution (A(x), B(x)), x ∈ R of (4.5) defines (4.2) in the domain of the Hamiltonian vector field vHfε (the right-
a small, bounded solution (4.1), (4.2) (with constraints (3.8), hand sides of (3.21)–(3.26) define a smooth mapping f ε : N1 → M0
(3.20)) via formula (4.6). with vHfε |u = f ε (u) for any u ∈ D(vHfε )). We overcome this
difficulty by using the change of variable (η, ω, τ ′ , ζ ′ , τ , ζ ) =
4.2. Proof of the reduction theorem G(η, ω, χ ′ , ξ ′ , χ , ξ ), where
K1′ W ξ ′
 y 
The key result is the following theorem, which is a parametrised, τ =χ +
′ ′
√ + K2 χy − χy ′ ′ ′
dt − β0 τ̂ ,
Hamiltonian version of a reduction principle for quasilinear evo- 0 1 − W2
lutionary equations presented by Mielke [24, Theorem 4.1] (see WK1′ χy′
Buffoni, Groves & Toland [23, Theorem 4.1]). ζ′ = ξ′ − √ + K2′ ξ ′
1 − W2
Theorem 4.2. Consider the differential equation W
− (µ(1) − 1) √ − ξ ′ + (µ(1) − 1)ω − 3y2 β03 ζ̂ ,
ux = Lu + N (u; λ), (4.7) 1 − W2
0
K1 W ξ
 
1
which represents Hamilton’s equations for the reversible Hamiltonian τ =χ+ −√
system (M , Ω λ , H λ ). Here u belongs to a Hilbert space X , λ ∈ Rℓ is a µ(1) + µ̇(1) y 1 − W2
parameter and L : D(L) ⊂ X → X is a densely defined, closed linear − µĎ (K2 χy + 1) + µ(1)
operator. Regarding D(L) as a Hilbert space equipped with the graph 
norm, suppose that 0 is an equilibrium point of (4.7) when λ = 0 and + (µ(1) + µ̇(1))χy dt − β0 τ̂ ,
that
WK1 χy
 
(H1) the part of the spectrum σ (L) of L which lies on the K2 1
ζ = ξ + µ(1) − √ + Ďξ − ξ − 3y2 β03 ζ̂ ,
imaginary axis consists of a finite number of eigenvalues of finite 1 − W2 µ µ(1)
multiplicity and is separated from the rest of σ (L) in the sense of
and
Kato, so that X admits the decomposition X = X1 ⊕ X2 , where
X1 = P (X ), X2 = (I − P )(X ) and P is the spectral projection
0 0
K1 W ξ
  
1
corresponding the purely imaginary part of σ (L).
τ̂ = −√
2(µ(1) + µ̇(1)) − 1
y 1 − W2
(H2) the operator L2 = L|X2 satisfies the estimate β0

− µĎ (K2 χy + 1) + µ(1)
C
∥(L2 − isI ) ∥X2 →X2 ≤
−1
, s ∈ R, 
1 + | s| + (µ(1) + µ̇(1))χy dt dy
for some constant C that is independent of s.
(H3) there exist a natural number k and neighbourhoods Λ ⊂ Rℓ
1
K1′ W ξ ′
  y 
1 β0
of 0 and U ⊂ D(L) of 0 such that N is (k + 1) times + √ + K2 χy − χy
′ ′ ′
dt dy,
2 1 − W2
continuously differentiable on U ×Λ, its derivatives are bounded 0 0

and uniformly continuous on U × Λ and N (0, 0) = 0, µ(1)



β0
1 
WK1 χy K2 1

d1 N [0, 0] = 0. ζ̂ = −√ + ξ− ξ dy
2 0 1 − W2 µĎ µ(1)
Under these hypotheses there exist neighbourhoods Λ̃ ⊂ Λ of 0 and 1
WK1′ χy′

Ũ1 ⊂ U ∩X1 , Ũ2 ⊂ U ∩X2 of 0 and a reduction function r : Ũ1 ×Λ̃ →

1 β0
+ −√ + K2′ ξ ′
Ũ2 with the following properties. The reduction function r is k times 2 0 1 − W2
continuously differentiable on Ũ1 × Λ̃, its derivatives are bounded and W
uniformly continuous on Ũ1 × Λ̃ and r (0; 0) = 0, d1 r [0; 0] = 0. The − (µ(1) − 1) √
1 − W2
graph M̃ λ = {u1 + r (u1 ; λ) ∈ X1 ⊕ X2 : u1 ∈ Ũ1 } is a Hamiltonian 
centre manifold for (4.7), so that − ξ + (µ(1) − 1)ω dy,

λ
(i) M̃ is a locally invariant manifold of (4.7): through every point
which transforms the nonlinear boundary conditions in D(vHfε )
in M̃ λ there passes a unique solution of (4.7) that remains on M̃ λ
into their linearisations
as long as it remains in Ũ1 × Ũ2 .
(ii) every small bounded solution u(x), x ∈ R of (4.7) that satisfies Bl (η, ω, τ ′ , ζ ′ , τ , ζ ) = 0
(u1 (x), u2 (x)) ∈ Ũ1 × Ũ2 lies completely in M̃ λ .
where Bl = dB[0] (see formula (4.4)).
(iii) every solution u1 : (x1 , x2 ) → Ũ1 of the reduced equation
u1x = Lu1 + PN (u1 + r (u1 ; λ); λ) (4.8) Lemma 4.3. (i) There exists an open neighbourhood N̂1 of the origin
generates a solution such that G : N1 → N̂1 is a diffeomorphism.
(ii) For each u ∈ N1 the operator dG[u] : M1 → M1 extends
u(x) = u1 (x) + r (u1 (x); λ) (4.9) [u] : M0 → M0 . The operators dG  [ u] ,
to an isomorphism dG
of the full Eq. (4.7). [u]−1 ∈ L(M0 , M0 ) depend smoothly upon u ∈ N1 .
dG
10 M.D. Groves et al. / Physica D 350 (2017) 1–12

Proof. (i) We observe that G is a smooth, near identity mapping (ii) There exist real constants C , s0 > 0 such that
N1 → M1 and apply the inverse-function theorem. (ii) The C
result for dG[u] follows from a direct calculation, while the ∥(L − isI )−1 ∥L(M0 ,M0 ) ≤
second assertion is proved using the argument given by Groves & |s|
Mielke [25, Lemma 3.3].  for each real number s with |s| > s0 .

A simple calculation shows that the diffeomorphism G trans- The centre manifold M̃ ε is equipped with the single coordinate
forms chart Ũ1 ⊂ X1 and coordinate map π : M̃ ε → Ũ1 defined
by π −1 (u1 ) = u1 + r (u1 ; ε). It is however more convenient to
ux = f ε ( u)
use an alternative coordinate map for calculations. According to
into the parameter-dependent version of Darboux’s theorem (e.g. see
Buffoni & Groves [22, Theorem 4]) there exists a near-identity
vx = fˆ ε (v), (4.10) change of variable
where fˆ ε : N̂1 → M0 is the smooth vector field defined by u1 = û1 + D(û1 ; ε)
ˆε
f (v) = dG
 [G −1 ε
(v)] (f (G (v))).
−1 of class C k−1 which transforms Γ̃ ε into Υ , where

Formula (4.10) represents Hamilton’s equations for the Hamilto- Υ (v1 , v2 ) = Γ̃ 0 |0 (v1 , v2 ) = Γ |0 (v1 , v2 ).
nian system (M0 , Γ , Ĥfε ), where Define the function r̃ : U1 ×Λ̃ → Ũ1 ×Ũ2 with U1 = P G−1 (Ũ1 ×Ũ2 )
Γ n (v1 , v2 ) = Ω (dG
 [G−1 (v)]−1 (v1 ), dG
 [G−1 (n)]−1 (v2 )), (which in general has components in X1 and X2 ) by the formula

û1 + r̃ (û1 ; ε) = G−1 û1 + D(û1 ; ε) + r (û1 + D(û1 ; ε); ε) , (4.11)


 
n ∈ N̂1 , v1 , v2 ∈ TX0 |v ,

and where r̃ (0; 0) = 0, d1 r̃ [0; 0] = 0, and equip M̃ ε with the


ε ε coordinate map π̂ : M̃ ε → U1 given by π̂ −1 (û1 ) = û1 + r̃ (û1 ; ε).
Ĥf (n) = Hf (G −1
(n)), n ∈ N̂1 .
We use a basis for X1 with respect to which Υ is the canonical
The domain of the Hamiltonian vector field vĤ ε is symplectic 2-form (a ‘symplectic basis’). Choose e, f ∈ D(L) such
f
that
D(vĤ ε ) = {(η, ω, τ , ζ , τ , ζ ) ∈ N̂1 : Bl (η, ω, τ ′ , ζ ′ , τ , ζ ) = 0}
′ ′
f Le = iβ0 qe, Le = −iβ0 qe,

and vĤ ε |n = fˆ ε (n) for any n ∈ D(vĤ ε ). (L − iβ0 qI )f = e, (L + iβ0 qI )f = e,


f f
The next step is to verify that (4.10) satisfies the hypotheses of where Re = e, Rf = −f and Ω (e, f ) = 1, Ω (f , e) = −1 and
Theorem 4.2 (with X = M0 ), so that we obtain a finite-dimensional the symplectic products of all other combinations are zero (note
reduced Hamiltonian system (M̃ ε , Γ̃ ε , H̃fε ). We write (4.10) as that Ω acts bilinearly on pairs of complex vectors). It follows that
{e, f , e, f } is a symplectic basis for the central subspace of L (so
ux = Lu + N ε (u),
that the coordinates A, B, A and B in the e, f , e and f directions are
in which the linear operator L : D(L) ⊂ M0 → M0 with L = dg 0 [0] canonical coordinates) and the action of the reverser R on this space
and is given by

D(L) = {(η, ω, χ ′ , ξ ′ , χ , ξ ) ∈ M1 : Bl (η, ω, χ ′ , ξ ′ , χ , ξ ) = 0} R(A, B) = (A, −B).

is given explicitly by the formula (4.3). (Observe that dg 0 [0] = We can now identify (M̃ ε , Γ̃ ε , H̃fε ) with the four-dimensional
df 0 [0] and Bl = dB[0], so that L is the formal linearisation of vHfε .) canonical Hamiltonian system (U1 , Υ , H̃fε ) using the coordinate
It follows from Lemma 4.4 below that L satisfies hypotheses (H1) map π̂ , where
and (H2); hypothesis (H3) is clearly satisfied for an arbitrary value
of k. Part (i) of Lemma 4.4 is proved using the elementary theory Υ ((A1 , B1 , A1 , B1 ), (A2 , B2 , A2 , B2 ))
of ordinary differential equations, while part (ii) established using = A1 B2 − A2 B1 + A1 B2 − A2 B1
arguments similar to those employed for other problems treated
using centre-manifold reduction (e.g. see Buffoni, Groves & Toland and
[23, Proposition 3.2] or Groves & Wahlén [19, Lemma 3.4]). H̃fε (A, B) = Ĥfε (û1 + r̃ (û1 ; ε)), û1 = Ae + Bf + Ae + Bf ;

Lemma 4.4. (i) The spectrum σ (L) of L consists entirely of isolated Hamilton’s equations for (U1 , Υ , H̃fε ) are given by (4.5).
eigenvalues of finite algebraic multiplicity. Zero is not an
eigenvalue of L, and a non-zero complex number λ is an 5. Homoclinic bifurcation
eigenvalue if and only if λ = β0 σ , where

(µ(1) − 1)2 (µ(1) + µ̇(1))σ σ̃ sin σ sin σ̃ The coordinates A, B, A, B for the reduced Hamiltonian system
may be chosen so that the reduced Hamiltonian takes the normal
= (σ 2 β0 − α0 ) σ sin σ cos σ̃

form
+ (µ(1) + µ̇(1))σ̃ sin σ̃ cos σ

H̃fε (A, B) = iβ0 q(AB − AB) + |B|2
and + HNF (|A|2 , i(AB − AB), ε)

µ(1) + O(|(A, B)|2 |(ε, A, B)|n0 ), (5.1)
σ̃ = σ .
µ(1) + µ̇(1) where HNF is a real polynomial of order n0 + 1 satisfying
(In particular, σ (L) ∩ iR is a finite set.) HNF (|A|2 , i(AB − AB), ε) = O(|(A, B)|2 |(ε, A, B)|)
M.D. Groves et al. / Physica D 350 (2017) 1–12 11

and n0 ≥ 2 is a fixed integer; Hamilton’s equations for the reduced specific magnetisation law). Here we confine ourselves to stating
system are given by the values of the coefficients for two particular special cases.

Ax = iβ0 qA + B + iA∂2 HNF (|A|2 , i(AB − AB), ε)


(i) Constant relative permeability µ (corresponding to a linear
+ O(|(A, B)||(ε, A, B)|n0 ), (5.2) magnetisation law): We find that
Bx = iβ0 qB + iB∂2 HNF (|A|2 , i(AB − AB), ε)  −1
µ (µ − 1)2

− A∂1 HNF (|A|2 , i(AB − AB), ε) c1 = − 1 + (q tanh q − 1)sech2 q ,
β0 µ + 1
+ O(|(A, B)||(ε, A, B)|n0 ). (5.3) 1

µ (µ − 1)2
−1
c3 = 1+ (q tanh q − 1)sech2 q
2 β0 µ + 1
Existence theories for homoclinic solutions to (5.2), (5.3) have
q β0 µ (µ − 1)6 (−3 − 4 cosh 2q + cosh 4q)(−2 + sech2 q + 4sech2q)
 4 4 2
been given by Iooss & Pérouème [21] and Buffoni & Groves [22] ×
(µ + 1)4 2qβ (µ−1)2 µ tanh 2q − γ − 4q2 β 2 cosh2 q cosh 2q
 
under the assumption that the coefficients c1 and c3 in the 4
0 µ+1 0 0
expansion
q4 β04 µ2 (µ − 1)6
+ sech4 q
HNF = ε c1 |A|2 + ε ic2 (AB − AB) + c3 |A|4 γ0 (µ + 1)4
+ ic4 |A|2 (AB − AB) − c5 (AB − AB)2 q3 β03 cosech2q 
+ −16µ(µ2 − 1)2 cosh2 q
4 (µ + 1)3
+ ε 2 c6 |A|2 + ε 2 ic7 (AB − AB) + · · · 
+ 16(µ − 1)4 µsech2q + 6qβ0 (µ + 1)3 sinh 2q ,
are respectively negative and positive.

where β0 q is the critical wavenumber associated with the


Theorem 5.1. Suppose that c1 < 0 and c3 > 0. Rosensweig instability. The sign of c3 clearly depends upon µ and
q (see Fig. 6). For large fluid depths one requires µ ≥ µc ≈ 3.5 for
(i) (Iooss & Pérouème) For each sufficiently small, positive value of localised one-dimensional free surfaces to bifurcate from the trivial
ε the two-degree-of-freedom Hamiltonian system (5.2), (5.3) has state, while for experimentally relevant ferrofluids (2 < µ < 6)
two distinct symmetric homoclinic solutions. one requires q > 2. For sufficiently shallow depths, we find that
(ii) (Buffoni & Groves) For each sufficiently small, positive value bifurcation occurs for all values of µ. However, due to the choice
of ε the two-degree-of-freedom Hamiltonian system (5.2), of the boundary conditions, the model described in Section 2 may
(5.3) has an infinite number of geometrically distinct homoclinic become invalid at these shallow depths.
solutions which generically resemble multiple copies of one of the (ii) Small values of... β0 (corresponding to deep fluids):... Abbreviating
homoclinic solutions in part (i). µ(1), µ̇(1), µ̈(1), µ(1) to respectively µ1 , µ̇1 , µ̈1 , µ1 , one finds that
The homoclinic solutions identified above correspond to envelope pat-
terns whose amplitude is O((−c1 ε)1/2 ) and which decay exponen- c1 = −1 + o(1),
tially as x → ±∞; they are sketched in Fig. 4.
3s4 4s5
 
m
c3 = − 1+
The coefficients c1 and c3 are given by the formulae 4 (µ1 − 1)2 µ1 + µ̇1
s6
c1 = 2H21 [e, e] −
(µ1 − 1)4 µ1 (µ1 + µ̇1 )3
and 
× −6µ21 µ̈1 + 6µ1 µ̈1 + 24µ31 µ̇1 + 24µ21 µ̇21 − 50µ21 µ̇1 + 8µ1 µ̇31
c3 = 6H40 [e, e, e, e] + 3H30 [r̃2000
0
, e, e] + 3H30 [r̃1010
0
, e, e], 
− 22µ1 µ̇21 + 26µ1 µ̇1 + 6µ̇21 + 8µ41 − 16µ31 + 8µ21
j k3 k4
where r̃k1 k2 k3 k4 is the coefficient of ε j Ak1 Bk2 A B in the Taylor ms7
expansion of r̃ and +
4(µ1 − 1) µ31 (µ1 + µ̇1 )4
4

ε
∂H  ...
   
1 1 × 4µ31 µ1 − 9µ21 µ̈21 + 20µ31 µ̈1 − 5µ̇41 − 18µ1 µ̇31 − 37µ21 µ̇21
Hk0 = dk H 0 [0], Hk1 = dk [0];
k! k! ∂ε ε=0 ... 
+ 12µ31 µ̇1 + 4µ21 µ̇1 µ1 − 2µ1 µ̇21 µ̈1 − 18µ21 µ̇1 µ̈1
0 0
the quantities r̃2000 and r̃1010 are found from the linear boundary-
4s4 (µ̇1 + µ1 − 1) a1
value problems −
(µ1 − 1)3 (µ1 + µ̇1 )
(f10 − 2iβ0 qI )r̃2000
0
= −f20 [e, e], s4 a2
+
0
Bl r̃2000 = −B2 [e, e], (µ1 − 1) µ1 (µ1 + µ̇1 )2
3

0,0
 
f10 r̃1010 = −2f20 [e, e], × µ21 µ̈1 − µ1 µ̈1 + 11µ21 µ̇1 + 5µ1 µ̇21 − 7µ1 µ̇1 − µ̇21 + 4µ31 − 4µ21
0
Bl r̃1010 = −2B2 [e, e], + o(1),

where as β0 → 0, where
1 1 (µ1 − 1)2 µ1 µ1

fk0 = dk f 0 [0], Bk = dk B[0]. s= , m=
k! k! 2(m + µ1 ) µ1 + µ̇1
These formulae are derived using the method explained by Groves
and
& Mielke [25, Appendix B]. −1
2ms(µ1 + µ̇1 )

Attempting to compute explicit general expressions for c1 and   2s
a1 2
c2 leads to unwieldy formulae (it appears more appropriate to = 5s 5s2
2 µ1 s − 2ms(µ1 + µ̇1 ) +

a2
calculate them numerically for a specific choice of µ, that is a µ1 − 1 µ1 − 1
12 M.D. Groves et al. / Physica D 350 (2017) 1–12

Fig. 6. The sign of the coefficient c3 as a function of µ and q for a linear magnetisation law (left) and as a function of M and γ for the Langevin magnetisation law (5.4) in
a ferrofluid of great depth (right). The coefficient c1 is negative in both panels. The shaded areas show the regions in which localised one-dimensional interfaces bifurcate
from the trivial state.

4(m + 1)s3 s4 µ̇21 + µ1 (µ̈1 + 3µ̇1 )


   
[3] C. Gollwitzer, G. Matthies, R. Richter, I. Rehberg, L. Tobiska, The surface topog-
− − raphy of a magnetic fluid: a quantitative comparison between experiment and
 µ1 − 1 (µ1 − 1)2 µ1 (µ1 + µ̇1 )2  numerical simulation, J. Fluid Mech. 571 (2007) 455–474.
s4
 
[4] C. Gollwitzer, I. Rehberg, R. Richter, From phase space representation to
−2s2 (µ1 − 1)µ1 −
 
amplitude equations in a pattern-forming experiment, New J. Phys. 12 (2010)
(µ1 − 1) µ1 (µ1 + µ̇1 ) 

2 2
× . 093037.

  [5] R. Richter, Magnetic liquid mountains, Europhys. News 42 (2011) 17–19.
[6] R. Richter, I.V. Barashenkov, Two-dimensional solitons on the surface of
 
 × µ1 µ̈1 − 28µ2 µ̇1 − 14µ1 µ̇2

magnetic fluids, Phys. Rev. Lett. 94 (2005) 184503.

 1 1 
 [7] D.J.B. Lloyd, C. Gollwitzer, I. Rehberg, R. Richter, Homoclinc snaking near the
+17µ1 µ̇1 + µ̇21 − 14µ31 + 14µ21 surface instability of a polarisable fluid, J. Fluid Mech. 783 (2015) 283–305.
[8] O. Lavrova, G. Matthies, L. Tobiska, Numerical study of soliton-like surface
configurations on a magnetic fluid layer in the Rosensweig instability,
(Note that the critical wavenumber associated with the Rosensweig
Commun. Nonlinear Sci. Numer. Simul. 13 (2008) 1302–1310.
instability is s + o(1) as β0 → 0.) Fig. 6 shows the sign of c3 for the [9] Y. Cao, Z.J. Ding, Formation of hexagonal pattern of ferrofluid in magnetic field,
Langevin magnetisation law J. Magn. Magn. Mater. 355 (2014) 93–99.
  [10] A. Gailitis, Form of surface instability of a ferromagnetic fluid, Magnetohydro-
M 1 dynamics 5 (1969) 44–45.
µ(s) = 1 + coth(γ s) − (5.4) [11] A. Gailitis, Formation of the hexagonal pattern on the surface of a
s γs ferromagneticfluid in an applied magnetic field, J. Fluid Mech. 82 (1977)
401–413.
in the limit β0 → 0, where M and χ0 are respectively the magnetic [12] R. Friedrichs, A. Engel, Pattern and wave number selection in magnetic fluids,
saturation and initial susceptibility of the ferrofluid and γ = Phys. Rev. E 64 (2001) 021406.
[13] V.M. Zaitsev, M.I. Shliomis, Nature of the instability of the interface between
3χ0 /M. Fig. 6 shows that there is a critical magnetic saturation two liquids in a constant field, Dokl. Akad. Nauk SSSR 188 (1969) 1261.
Mc ≈ 10 below which no localised one-dimensional interfaces [14] E.E. Twombly, J.W. Thomas, Bifurcating instability of the free surface of a
bifurcate from the trivial state. ferrofluid, SIAM J. Math. Anal. 14 (1983) 736–766.
[15] M. Silber, E. Knobloch, Pattern selection in ferrofluids, Physica D 30 (1988)
83–98.
Remark 5.2 (Mathematics of the Normal-Form Theory). The above [16] S. Bohlius, H. Pleiner, H.R. Brand, Solution of the adjoint problem for
choice of coordinates (A, B, Ā, B̄) is accomplished by the Birkhoff instabilities with a deformable surface: Rosensweig and Marangoni instability,
normal-form theory for (U1 , Υ , Hfε ), which states that for each Phys. Fluids 19 (2007) 094103.
[17] S. Bohlius, H. Pleiner, H.R. Brand, The amplitude equation for the Rosensweig
n0 ≥ 2 there is a near-identity, analytic, symplectic change of
instability in magnetic fluids and gels, Progr. Theoret. Phys. 125 (2011)
coordinates with the property that Hfε takes the form (5.1) in the 1–46.
new coordinates (see Buffoni & Groves [22, pp. 196–197]). We [18] G. Iooss, Capillary–gravity water-waves problem as a dynamical system,
in: A. Mielke, K. Kirchgässner (Eds.), Structure and Dynamics of Nonlinear
incorporate this feature into the construction of r̃ by replacing the Waves in Fluids, World Scientific, Singapore, 1995, pp. 42–57.
Darboux transformation used in formula (4.11) by its composition [19] M.D. Groves, E. Wahlén, Spatial dynamics methods for solitary gravity-
with the normal-form transformation. capillary water waves with an arbitrary distribution of vorticity, SIAM J. Math.
Anal. 39 (2007) 932–964.
[20] D.V. Nilsson, Internal gravity-capillary solitary waves in finite depth, Math.
Acknowledgements Methods Appl. Sci. 40 (2017) 1053–1080.
[21] G. Iooss, M.C. Pérouème, Perturbed homoclinic solutions in reversible 1:1
A.S. was supported by the Deutsche Forschungsgemeinschaft resonance vector fields, J. Differential Equations 102 (1993) 62–88.
[22] B. Buffoni, M.D. Groves, A multiplicity result for solitary gravity-capillary
under grant GR 3348/1-1. D.J.B.L. acknowledges support from waves in deep water via critical-point theory, Arch. Ration. Mech. Anal. 146
an EPSRC grant (Nucleation of Ferrosolitons and Ferropatterns, (1999) 183–220.
EP/H05040X/1). No new data were created during this study. [23] B. Buffoni, M.D. Groves, J.F. Toland, A plethora of solitary gravity-capillary
water waves with nearly critical Bond and Froude numbers, Philos. Trans. R.
Soc. Lond. Ser. A 354 (1996) 575–607.
References [24] A. Mielke, Reduction of quasilinear elliptic equations in cylindrical domains
with applications, Math. Methods Appl. Sci. 10 (1988) 51–66.
[1] M.D. Cowley, R.E. Rosensweig, The interfacial stability of a ferromagnetic fluid, [25] M.D. Groves, A. Mielke, A spatial dynamics approach to three-dimensional
J. Fluid Mech. 30 (1967) 671–688. gravity-capillary steady water waves, Proc. Roy. Soc. Edinburgh Sect. A 131
[2] R.E. Rosensweig, Ferrohydrodynamics, Dover, New York, 1997. (2001) 83–136.

You might also like