You are on page 1of 49

ARTICLE IN PRESS

J. Differential Equations 205 (2004) 1–49

Ostwald ripening in two dimensions—the


rigorous derivation of the equations from the
Mullins–Sekerka dynamics
Nicholas D. Alikakos,a,b, Giorgio Fusco,c and Georgia Karalid
a
Department of Mathematics, University of North Texas, USA
b
Department of Mathematics, University of Athens, Athens, Greece
c
Dipartimento di Matematica, Universita di L’Aquila, Italy
d
Department of Mathematics, University of Toronto, Toronto, Canada
Received January 10, 2002; revised March 20, 2003

Abstract

We consider a dilute mixture of a finite number of particles and we are interested in the
coarsening of the spatial distribution in two space dimensions under Mullins–Sekerka
dynamics. Under the appropriate scaling hypotheses we associate radii and centers to each
particle and derive equations for the whole evolution.
r 2004 Elsevier Inc. All rights reserved.

Keywords: Ostwald ripening; Mullins–Sekerka; Asymptotics; Stability

1. Introduction

Ostwald ripening or coarsening is a diffusion process occurring in the last stage of


a first-order phase transformation. Usually, any first-order phase transformation
process results in a two-phase mixture with a dispersed second phase in a matrix.
Initially, the average size of the dispersed particles is very small and therefore the
interfacial energy of the system is large and the mixture is not in thermodynamical
equilibrium. The force that drives the system towards equilibrium is the gradient of
the chemical potential that, according to the Gibbs–Thomson condition, on the


Corresponding author.
E-mail address: nalikako@earthlink.net (N.D. Alikakos).

0022-0396/$ - see front matter r 2004 Elsevier Inc. All rights reserved.
doi:10.1016/j.jde.2004.05.008
ARTICLE IN PRESS
2 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

interface, is proportional to its mean curvature. As a result matter diffuses from


regions of higher curvature to regions of lower curvature and large particles grow at
the expense of smaller particles that eventually shrink to nothing. The outcome of
this process, known as Ostwald ripening is an increasing of the average size of
particles and a reduction of the number of them that makes the mixture coarser. A
quantitative description of Ostwald ripening [34] has been developed by Lifschitz
and Slyozov and independently by Wagner under the assumption that the relative
fraction of the dispersed phase is very small. Their theory (LSW theory in the
following) is in three dimensions and assumes that there are many particles in the
system with the size of the particles small compared to the distance between them.
LSW is a mean field theory and is derived as follows: The point of departure is the
quasi-static Stefan problem with surface tension (otherwise known as Mullins–
Sekerka free boundary problem, see (1.7)). From this problem a differential equation
(1.2) describing the evolution of the size of a typical particle is formally derived. The
derivation is based on the assumption that particles are exact spheres and that their
centers stay fixed in time. The evolution is characterized by the particle distribution
nðR; tÞ dR which is defined as the number of particles which at time t have radius in
½R; R þ dR: More specifically the LSW theory provides the equation
 
@nðR; tÞ @ dR
þ nðR; tÞ ¼ 0 ð1:1Þ
@t @R dt

with
 
dR 1 1 1
¼  ; ð1:2Þ
dt %
RðtÞ RðtÞ RðtÞ

%
where RðtÞ is the average radius size
R
RnðR; tÞ dR
% ¼ R
RðtÞ : ð1:3Þ
nðR; tÞ dR

System (1.1)–(1.3) is analyzed in [27,41] and it is shown that there exist infinitely
many self-similar solutions, but only one is believed to describe the typical behavior
of the system for large times
 
1 RðtÞ
nðR; tÞD 4 g : ð1:4Þ
%
t3 RðtÞ
The theory predicts also the following temporal laws for the average radius and the
total number of particles:
 3 1
%
RðtÞD R% ð0Þ þ 49 t 3 ; ð1:5Þ

 1
NðtÞ ¼ R% 3 ð0Þ þ 49 t : ð1:6Þ
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 3

Niethammer [32] has rigorously derived Eqs. (1.1) and (1.2) through a homogeniza-
tion procedure starting by a suitable modification of the Mullins–Sekerka problem.
Alikakos and Fusco [3–5] obtained precise expressions for the equations of the
centers and the radii by taking also into account the geometry of the distribution
thus removing these restrictive hypotheses. These results will make it possible to pass
rigorously to the limit [7,33].
The quasistatic Stefan problem or Mullins–Sekerka model [31] in dimensionless
variables takes the form
8
>
> Du ¼ 0 off GðtÞ; in OCRm ; m ¼ 2; 3;
>
>
>
> u¼H on GðtÞ;
>
<
@u
¼0 on @O; ð1:7Þ
>
> @n  
>
>
>
> @u
>
:V ¼ on GðtÞ;
@n

where u is the chemical potential, H is the mean curvature of G; the sign convention
for H is that H is positive for a shrinking sphere; ~
n is the outward normal to @O; V is
@uþ @u
the normal velocity positive for a shrinking sphere; ½½@u @n ¼ @nþ þ @n is the jump of
the derivative of u in the normal direction to GðtÞ where uþ ; u are the restrictions of
u on the exterior Oþ ðtÞ and the interior O ðtÞ of GðtÞ in O and n ; nþ the unit
S
exterior normal to Oþ ðtÞ; O ðtÞ: Here GðtÞ ¼ N i¼1 Gi ðtÞ is the union of the
boundaries of the N particles and O is a bounded, smooth domain (the container of
the mixture). The Mullins–Sekerka model is a nonlocal evolution law in which the
normal velocity of a propagating interface depends on the jump across the interface
of the normal derivative of a function which is harmonic on either side and which
equals the mean curvature on the propagating interface. If H ¼ const: on GðtÞ then
uðxÞ ¼ H ¼ const:; 8xAO and V
0 solve (1.7). Therefore O is the union of NX1
equal balls with the same radius which are equilibria for (1.7).
The Mullins–Sekerka model arises as a singular limit for the Cahn–Hilliard
equation, a fourth-order parabolic equation which is used as a model for phase
separation and coarsening phenomena in a melted binary alloy [1,12,13,35]. In this
paper we are interested in the evolution in two space dimensions. This problem has
also some physical interest, for example, in the theory of thin films. We refer to
[30,36,38]. We denote by PerðGðtÞÞ; VolðO ðtÞÞ; the surface area and the enclosed
volume (perimeter, enclosed area for n ¼ 2), respectively, then (1.7) is a volume
preserving, perimeter shortening law. A standard computation (e.g. [14]) shows that
Z Z   Z
d @u
PerðGðtÞÞ ¼  HV ¼  u ¼ jruj2 p0; ð1:8Þ
dt G G @n O

Z Z   Z
d @u
VolðO ðtÞÞ ¼  V¼ ¼ Du ¼ 0: ð1:9Þ
dt G G @n O\G
ARTICLE IN PRESS
4 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

In this paper, we consider the case where O is the union of NX1 small ‘‘particles’’
which are initially very close to balls Bxi ;Ri ; i ¼ 1; y; N with center xi and radii Ri :
We assume that the radii are small with respect to interparticle distances. These
assumptions imply in particular that we work under the premise that the ‘‘volume
fraction’’, that is the ratio j between the volume of the dispersed phase
P
jO jD N 2
i¼1 pRi and the measure of the set O is a small quantity
PN
jO j pR2i
j¼ D i¼1 51: ð1:10Þ
jOj jOj

We show that under these restrictions for the initial condition the particles retain
their almost circular shape until one singularity occurs which always results in the
fact that one or more particles disappear shrinking to nothing. We derive corrected
equations for the radii which take into account the distance and the size of the
neighboring particles and also equations for the motion of the centers of the
particles. Moreover, the robustness of the circular shape is established. In the case
that the boundary @O is removed, the equations of the radii and the centers take the
form:
8
2 1 < 1 1


Ri ¼ 
jlog jj Ri : R% Ri
2 39
  X   =
2 4 Ri 1 1 1 1
þ log 1  þ log jxi  xh j   E 5 þ ?; ð1:11Þ
jlog jj % Ri R% Rh ;
j2 R hai

 
4 X 1 1 xh  xi
x’ i ¼   þ ?; ð1:12Þ
jlog jj hai R% Rh jxh  xi j2

where ‘‘dots’’ denote higher order terms, R% is the harmonic mean of Ri


defined by

1 1 X
N
1
¼ ð1:13Þ
%
R N j¼1 Rj

and E is determined by the conservation of volume and is given by


   
1 XN
Rk 1 1 1 XN X
1 1
E¼ log 1  þ log jxh  xk j  : ð1:14Þ
N k¼1 % R% Rh
j2 R Rk N k¼1 hak

We remark that the above products and sums are taken only over the particles which
have positive radius. This applies to all formulas throughout the paper. As it is well
known the Mullins–Sekerka problem has an invariance property. Indeed, if t-O ðtÞ
is a solution and m40 a positive constant then t-O *  ðtÞ ¼: mO ð t3 Þ is again a
m
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 5

solution. The equations


 
2 1 1 1
R’ i ¼  ; ð1:15Þ
jlog jj Ri R% Ri

 
4 X 1 1 xh  x i
x’ i ¼   ð1:16Þ
%
jlog jj hai R Rh jxh  xi j2

which are obtained by (1.11) and (1.12) by retaining the principal terms enjoy the
same invariance property: if t-ðRðtÞ; xðtÞÞ is a solution then t-ðR̃ðtÞ; xðtÞÞ * ¼:
ðmRðmt3 Þ; mxðmt3 ÞÞ is again a solution. The complete equations (1.11) do exhibit exactly
the invariance property shared by Eqs. (1.7) and (1.15).
Eqs. (1.13), (1.15) state that given t; the radius Ri ðtÞ of the ith particle, decreases or
increases according to whether at that time it is below or above the average value R: %
Moreover, Eqs. (1.15) preserve the total enclosed area and reduce the total
perimeter,
d XN
R2 ¼ 0; ð1:17Þ
dt i¼1 i

d X N
Ri p0 ð1:18Þ
dt i¼1

reflecting the enclosed area preserving and perimeter shortening properties of


Mullins–Sekerka problem. We assume
0oR1 ð0ÞoR2 ð0Þo?oRN ð0Þ;

and show (see Proposition 3.1) that the solution of (1.15) preserves the order:
R1 ðtÞo?oRN ðtÞ in its maximal interval of existence ½0; T1 Þ and that T1 is bounded
and characterized by the fact that
lim R1 ðtÞ ¼ 0:
t-T1

After t ¼ T1 ; the solution of (1.15) can be continued to an interval ½T1 ; T2 Þ by


removing the first equation; changing N to N  1; and taking Rj ðT1 Þ; j ¼ 2; y; N as
initial condition. Proceeding in this way one defines a sequence of times
T1 oT2 o?oTN1 characterized by the fact that limt-Tj Rj ðtÞ ¼ 0: We show that
S
under the above assumption that O ð0ÞD N i¼1 Bxi ð0Þ;Ri ð0Þ ; something similar holds
true also for the full Mullins–Sekerka free boundary problem. Indeed we prove (see
Theorem 2.1) that there are times T̂1 oT̂2 o; y; oT̂N1 near to T1 ; y; TN1 such
that at time T̂j a singularity occurs and the jth particle shrinks to nothing. Eq. (1.15)
ARTICLE IN PRESS
6 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

for the case of 2 particles where first derived in [43] by the method of
images Eq. (1.15) were also obtained in [33] starting from a variation of the
Mullins–Sekerka model which assumes the Gibbs–Thomson relation in an averaged
integral form:

Comparison of principal terms


3 Dimensions 2 Dimensions
   
’ 1 1 1 ’ 1 2 1 1
Ri ¼  Ri ¼ 
Ri R% Ri Ri jlog jj R% Ri
1 PN  1
R% ¼ 1 PN 1
N j¼1
Rj R% ¼
N j¼1 Rj
arithmetic mean harmonic mean
independent of distance independent of distance
1 1
scale invariance compatible with t3 law scale invariance compatible with t3 law
1 1
x-m3 x; t-mt x-m3 x; t-mt
1 1
No singularity for R% R%
becomes singular at
the extinction times
Singularity like that of Singularity like that of
1 1 2
R’ i ¼  2 R’ i ¼  2
Ri Ri jlog jj
milder than in 3D
No crossing of time lines No crossing of time lines
No effect of neighbors No effect of neighbors

We now present a formal derivation of the equations of the radii and the centers in
two dimensions. In the following, we will show that in the limit of small size the
distortion from sphericity measured by the function ri introduced in Section 5 does
not affect to principal order the evolution of the centers and the radii of the particles.
Therefore in this paragraph in doing the formal derivation we will assume that O is
the union of NX1 perfect balls of center xi and radius Ri 40; i ¼ 1; y; N that is
S
O ¼ N i¼1 Bxi ;Ri : We represent the boundary @Bxi ;Ri of Bxi ;Ri through the map
X i : S 1 -R2 defined by

S1 {u-x ¼ X i ðuÞ :¼ xi þ Ri u: ð1:19Þ

We work under the assumption that the radii Ri are small, that is we assume Ri 51;
i ¼ 1; y; N: For this reason in the analysis that follows we will work with the
rescaled radii ri defined by

Ri ¼ eri ;
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 7

where e40 is a small parameter e51: The parameter e can be identified with the
square root of the volume fraction j
P
jO j e2 N 2
i¼1 pri
j¼ D :
jOj jOj

By regarding xi and Ri ¼ eri in (1.19) as functions of time t and by differentiating


with respect to t we get x’ ¼ x’ i þ er’ i u: By projecting this equation on the unit vector u
which coincides with the exterior normal to Bxi ;Ri we get the following expression for
the normal velocity V of @Bxi ;Ri at X i ðuÞ:
X
2
V ðuÞ ¼ er’ i  x’ij /u; ej S; x’ij ¼ /x’ i ; ej S; ð1:20Þ
j¼1

where fe1 ; e2 g is the standard basis of R2 ; /; S the standard scalar product of R2 and
the ‘‘–’’ sign is due to the sign convention that we assume V positive for a shrinking
sphere. As we recall in Section 3, (1.7) can be reduced via potential theory (see also
[43]) to a problem that lives entirely on the interface:
Z
gðx; yÞV ðyÞ dy ¼ HðxÞ  E; xAG; ð1:21Þ
GðtÞ

1
where gðx; yÞ ¼  2p log jx  yj þ gðx; yÞ is the Green function of the Neumann
problem
8
> Du ¼ f ;
>
< @u
¼ 0;
>
> R@n R
:
O u¼ O f ¼ 0

G ¼ GðtÞ; E ¼ EðtÞ; V and H also depend on time and EðtÞ is to be chosen in order
to ensure that Z
V ¼0
G

which implies conservation of volume.


In the formal derivation that follows we do not consider the contribution of the
smooth part of the Green’s function. We will indicate at the end the extra terms that
one gets when the contribution of g is also accounted for.
S
If G ¼ N 1
i¼1 @Bxi ;eri then, at @Bxi ;eri {x ¼ xi þ eri u; uAS ; Eq. (1.21) can be written
in the form
Z X Z  
1 1
eri  log jeri ðu  vÞjVi ðvÞ dv þ erh  log jxi  xh þ eðri u  rh vÞj
S1 2p hai S1 2p
1
 Vh ðvÞ dv ¼  E; 8uAS 1 ; i ¼ 1; y; N: ð1:22Þ
eri
ARTICLE IN PRESS
8 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Substituting Eq. (1.20) into Eq. (1.22) for xAGi ðtÞ; we obtain
Z  " X 2
#
1 1 ’
eri  log jeri j  log ju  vj er’ i  xij /v; ej S dv
S1 2p 2p j¼1

X Z " #
1 e /xi  xh ; uS e /xi  xh ; vS 2
þ erh  log jxi  xh j  ri þ rh þ Oðe Þ
hai S1 2p 2p jxi  xh j2 2p jxi  xh j2
" #
X 2
1
 er’ h  x’ hj /v; ej S dv ¼  E; 8uAS 1 ; i ¼ 1; y; N: ð1:23Þ
j¼1
er i

We now observe that


Z Z
1 1 1
 log ju  vj dv ¼ 0;  log ju  vj/v; ej S dv ¼ /u; ej S;
S 1 2p S 1 2p 2
Z Z
du ¼ 2p; /u; ej S du ¼ 0;
S1 S1
Z Z
/u; ej S2 du ¼ p; /u; e1 S/u; e2 S du ¼ 0:
S1 S1

Moreover, we can write

X
2
/xi  xh ; uS ¼ /xi  xh ; ej S/u; ej S:
j¼1

Utilizing the above expressions, we get from Eq. (1.23)


" # "
1X 2

X
eri er’ i log jeri j  xij /u; ej S þ erh er’ h log jxi  xh j
2 j¼1 hai

1 X
2
e
þ e2 ri r’ h 2
/xi  xh ; ej S/u; ej S þ Oðe3 Þ  rh
jxi  xh j j¼1 2
#
X2
/x  x ; e S 1
x’ hj i h j
 2
¼  E: ð1:24Þ
j¼1 jx i  x h j er i

This equation implies


8 " #
>
> P e P2 /x  x ; e S 1
> x’ hj i h j
< eri er’ i log jeri j þ
> erh er’ h log jxi  xh j  rh
2 2
¼
er
 E;
hai j¼1 jx i  x h j i
>
> P 1
: eri ð 2 x’ ij Þ þ
>
> 1
erh e2 ri r’ h /xi  xh ; ej S ¼ 0; i ¼ 1; y; N; j ¼ 1; 2:
hai jxi  xh j2
ð1:25Þ
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 9

From Eq. ð1:25Þ2 we obtain

X /xi  xh ; ej S
x’ ij ¼ 2e2 rh r’ h ð1:26Þ
hai jxi  xh j2

while Eq. ð1:25Þ1 is equivalent to


X
eri er’ i log jeri j þ erh er’ h log jxi  xh j
hai
X
2 X /xh  xk ; ej S/xi  xh ; ej S 1
 erh e2 rk r’ k 2 2
¼  E: ð1:27Þ
j¼1 kah jxh  xk j jxi  xk j eri

If in Eq. (1.27) we disregard the third term on the left-hand side we get
X 1
eri er’ i log e þ eri er’ i log ri þ erh er’ h log jxi  xh j ¼  E: ð1:28Þ
hai
eri

PN
From this if we determine E by imposing k¼1 rk r’ k ¼ 0 we get
  " #
1 1 1 1 1 1 X
er’ i ¼   þ er’ i log ri þ er er’ h log jxi  xh j
jlog ej eri eri er% jlog ej eri hai h
" #
1 X
N X
 er er’ k log rk þ erh er’ h log jxk  xh j : ð1:29Þ
eri jlog ejN k¼1 k k;h hak

P
Since we have N ’ k ¼ 0 if t-ri ðtÞ is a solution of (1.29) then t-r# i ðtÞ ¼ mri ðmt3 Þ
k¼1 rk r
is again a solution. Indeed if LðrÞ; RðrÞ correspond to the left-hand side and right-
hand side terms of (1.29) we have

1
# ¼
LðrÞ LðrÞ;
m2

1 log m X N
log m XN
1
# ¼
RðrÞ RðrÞ þ er h e ’
r h  N erh er’ h ¼ 2 RðrÞ:
m2 eri jlog ej h¼1 eri jlog ejN h¼1
m

From this computation it appears that the extra terms in the expression of RðrÞ #
PN
cancel out even if r
k¼1 k ’
r k a0: Let Q i ðrÞ be any function which satisfies the
following conditions:
1
c1  Qi ðmrÞ ¼ Qi ðrÞ;
m2
ARTICLE IN PRESS
10 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

X
N
c2  ri Qi ðrÞ ¼ 0:
i¼1

From the above discussion it follows that if we replace in the right-hand side of
(1.29) er’ i with Qi ðerÞ then the ODE that we obtain has the desired scale invariance.
Clearly, the function
 
1 1 1 1
Qi ðrÞ ¼  
jlog ej ri ri r%

satisfies c1 and c2 : This leads to the ODE


( 
1 1 1 1
er’ i ¼ 
eri jlog ej er% eri
"     #)
1 1 1 X 1 1
þ log ri  þ log jxi  xh j  E ; ð1:30Þ
jlog ej er% eri hai
er% erh

where
   
1X N
1 1 1X N X
1 1
E¼ log rk  þ log jxh  xk j 
N k¼1 er% erk N k¼1 hak er% erh

and this equation satisfies the correct scaling law.


Finally, we obtain from (1.26)

X  
1 1 1 xh  xi
x’ i ¼ 2  : ð1:31Þ
hai
jlog ej er% erh jxh  xi j2

This concludes the derivation.

Remark. The above analysis can be extended to include also the effect of the
boundary. With similar computations one finds that the boundary contributes with
the following terms:
(  "  
1 1 1 1 1 1 1
er’ i ¼  þ ðlog ri þ gðxi ; xi ÞÞ 
jlog ej eri er% eri jlog ej er% eri
  #)
X 1 1
þ ðlog jxi  xh j þ gðxi ; xh ÞÞ  E ;
hai
er% erh
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 11

where
 
1X N
1 1
E¼ ðlog rk þ gðxk ; xk ÞÞ 
N k¼1 er% erk
 
1X N X
1 1
þ ðlog jxh  xk j þ gðxh ; xk ÞÞ 
N k¼1 hak er% erh

and
  !
2 X 1 1 x h  xi @gðxi ; xh Þ
x’ i ¼   þ :
jlog ej hai er% erh jxh  xi j2 @x

We now present some examples which illustrate the use of system (1.29), (1.31),
(1.32).

The two-particle case (See Fig. 1).


We calculate
8  
>
> ’ 2 1 1 1
>
< R1 ¼ jlog jj R R%  R ;
1 1
 
>
> R’ ¼ 2 1 1 1
>
: 2 
jlog jj R2 R% R2
) R’ 1 ðtÞo0; R’ 2 ðtÞ40
8  
>
> ’ 4 1 1 x2  x1
< x1 ¼ 
> 
jlog jj R% R2 jx2  x1 j2
;
 
>
> 4 1 1 x1  x2
> x’ 2 ¼ 
:  :
%
jlog jj R R1 jx2  x1 j2

Fig. 1. The two-particle case for R1 oR2 :


ARTICLE IN PRESS
12 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Fig. 2. The three-particle case for R1 ¼ R2 oR3 :

Notice that if ũ ¼ jxx22 x ’ ’


x1 j is the unit vector along x1 ; x2 then /x1 ; ũS40; /x2 ; ũS40:
1

Moreover, it holds that /x’ 1 ðtÞ; ũS ¼ /x’ 2 ðtÞ; ũS: In the two-particle system, we see
that the smaller particle becomes smaller and the bigger becomes bigger. The two
unequal particles move in the same direction, in the direction of the smaller particle
with equal speeds.
The three-particle case (Fig. 2).
We consider the above arrangement and we calculate
8 2
< 1   
R’ 1 ¼
2 1

1
þ
2 4log R1 1  1
jlog jj R1 : R% R1 jlog jj 1 % R1
j2 R
39
    =
1 1 1 1
þ log d  þ log 2d   E5 ;
R% R2 R% R3 ;

8 2
2 1 < 1 1

2

R2 1 1


R2 ¼  þ 4log 1 
jlog jj R2 : R% R2 jlog jj % R2
j2 R
39
    =
1 1 1 1
þ log d  þ log d   E5 :
R% R1 R% R3 ;

By subtracting the above equations we obtain


 
1 1
R’ 1  R’ 2 ¼ log 2  40:
R% R3

So, the proximity to particle 3 impedes the growth of particle 2 and as a result
particle 1 grows faster than particle 2 (see Fig. 3).
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 13

Fig. 3. Four-particle system with the particles numbered from left to right, the leftmost particle denoted as
particle 1, the center particle as particle 2 and the rightmost particles as particles 3 and 4. Particles 3 and 4
are close to particle 2. We observe that despite particle 2 being smaller than particle 1, particle 2 grows at
the expense of 3 and 4.

Four-particle case: In systems with more than two particles, the competition is
complicated. Although location of a particle between particles of smaller and larger
sizes is a necessary condition for migration, calculations show that it is not a
sufficient condition. Small changes in the locations of the particles relative to one
another can have large effects on the evolution. With the permission of the authors
[43] we reproduce below some of their numerical results on four-particle systems (see
Figs. 3–5). A similar problem concerning numerical studies can be found in [39]
where the Laplace equation is solved only inside the region of the interface.
The paper is organized as follows: In Section 2, we present the statement of the
main result. In Section 3, we analyze the system of ODEs and present the integral
formulation of the equation. In Section 4, we present the operators T; L and the
operator A ¼ TL: Moreover, we express the mean curvature H in x; r; r coordinates.
In Section 5, we study the coordinate system [2,5,11,42]. Given an interface G close to
circular, we would like to associate to it in a unique way a circle and view the
interface as a small perturbation of that circle such that

GðtÞ ¼ x þ erð1 þ erðuÞÞu; uAS1 ;

Z Z
rðuÞ du ¼ 0; rðuÞ /u; ei S du ¼ 0; i ¼ 1; 2:
S1 S1

% assuming that H is known


In Section 6, we solve the linear equation SðV Þ ¼ H  H;
and we obtain a system for x; r and r equations with estimates for the higher order
terms. In Section 7, we obtain bounds on r by analyzing the r-equation. The control
on r is accomplished by utilizing the maximal regularity theory of da Prato and
ARTICLE IN PRESS
14 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Fig. 4. A four-particle system with the same radius as in Fig. 3 with the difference that the two smaller
particles are away from particle 2.

Fig. 5. A four-particle system with the same radius as in Fig. 3 but now the distance between particle 2 and
particles 3 and 4 is larger than those in Fig. 3 or Fig. 4. As a result particle 2 disappears before particles 3
and 4.

Grisvard [18]. In the previous sections we assume that ejjrjjC 1þa ðS1 Þ od for d40; a
fixed small number. In Section 7, a uniform bound on jjrjj is established. The
uniform bound on jjrjj is one of the harder analytic results. We should mention that
the bound on r implies the robustness of the spherical shape. Chen [14] and
independently Constantin and Pugh [17] established the stability of a single circle
equilibrium in two space dimensions. Later [21] this was reestablished in a more
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 15

general way. In our case, we first note that a configuration of two or more circles is
unstable and we show for arbitrary initial data of unequal circles that the distortion
away from circularity is small globally in time. Although the general layout is quite
close to [5] there are differences between two and three dimensions also on the
technical level with the two dimensional case tending in general to be harder. These
differences can always be traced back to the fundamental solution of the Laplacian
in two dimensions.

2. Statement of the main result

In Section 5, following [5], we will show that given an interface G close to circular,
we associate to it in a unique way a circle and view the interface as a small
perturbation of this circle. So, each interface will have unique xAR2 ; r40 and
rAC 1 ðS 1 Þ satisfying

GðtÞ ¼ fx=x ¼ x þ erð1 þ erðuÞÞu; uAS 1 g; ð2:1Þ

Z Z
rðuÞ du ¼ 0; rðuÞ/u; ei S du ¼ 0; i ¼ 1; 2: ð2:2Þ
S1 S1

Theorem 2.1. Let OCR2 be a bounded, connected and smooth domain. Assume that
S
Gð0Þ ¼ N i¼1 Gi ð0Þ; NX2; with Gi ð0Þ of the form (2.1), xi ð0ÞAO; Ri ð0Þ40 and
ri ð0ÞAC 3þa ðS 1 Þ satisfying (2.2). Assume that

xi ð0Þaxj ð0Þ for iaj;

R1 ð0ÞoR2 ð0Þo?oRN ð0Þ

then there is e%40 such that


PN
pR2i ð0Þ 2
j¼ i¼1
oe% ;
jOj

jjri ð0ÞjjC 3þa ðS1 Þ oe%;

imply that the solution t-GðtÞ of the Mullins–Sekerka problem (1.7) satisfies GðtÞ ¼
SN 3þa
i¼1 Gi ðtÞ with Gi ðtÞ of the form (2.1) with xi ðtÞ; Ri ðtÞ40; ri ðtÞAC ðS1 Þ satisfying
(2.2) and exists globally as a weak solution in the sense of [15,37]. Moreover,
(i) There exist times T̂1 o?oT̂N1 such that limt-T̂i Ri ðtÞ ¼ 0; i ¼ 1; y; N  1:
The solution is classical except at that times.
ARTICLE IN PRESS
16 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Ri ð0Þ
(ii) There are constants Cr ; CR 40 depending on R1 ð0Þ i ¼ 2; y; N such that
 
2 1 1 1
R’ i ¼  þ CR gi ; ð2:3Þ
jlog jj Ri R% Ri

where R% is the harmonic mean of Ri defined by

1 1XN
1
¼
R% N j¼1 Rj

and gi ðR; e; rÞ a smooth function satisfying


1
jgi jo :
jlog jj2

The above expression holds for T̂i1 otoT̂i where for t4T̂i i ¼ 1; y; N we have that
Ri ðtÞ ¼ 0:
In addition,

jjri jjC 3þa ðS1 Þ oCr :

3. The integral equation formulation and the equations of Ostwald Ripening

We would like to formulate the Mullins–Sekerka problem as an integral equation


in the class of C 3þa interfaces. The immediate advantage of this approach is that the
space dimension of the problem will be reduced by one. We will only need to solve
the integral equations along the boundaries of the evolving domains instead of
solving the two space dimension PDE problem. This approach has been used in
many problems such as Ostwald ripening [39], water waves [10], etc. It is known [43]
that we have the following integral formulation for the Mullins–Sekerka problem
(1.7):
Z Z Z
1
gðx; yÞV ðyÞ dSy  gðx; yÞV ðyÞ dSy dSx ¼ HðxÞ  H% ð3:1Þ
GðtÞ jGðtÞj GðtÞ GðtÞ

with
Z
1
H% ¼ H dSy :
jGðtÞj GðtÞ

Then problem (1.7) takes the form SðV Þ ¼ H  H% where S is a linear operator, V
the normal velocity, H the mean curvature, H% the average value of H over GðtÞ and
jGðtÞj the surface area of GðtÞ: Moreover, we have that the Green’s function in two
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 17

dimensions has the form

1
gðx; yÞ ¼  log jx  yj þ gðx; yÞ; ð3:2Þ
2p

where g is the function associated to the problem 5


8
> Du ¼ f ; in O;
>
< @u
¼ 0; on @O; ð3:3Þ
>
> @n
:R R
O u dx ¼ O f dx ¼ 0

and satisfies
8 1
> Dy gðx; yÞ ¼ dx ðyÞ  jOj in O;
>
>
<
@g
¼0 on @O; ð3:4Þ
> @ny
>
>
:R
O gðx; yÞ dy ¼ 0:

On the other hand, g is the smooth part of the Green’s function and satisfies
8
>
> 1
>
> Dy gðx; yÞ ¼  ; xAO; yAO;
>
> jOj
>
<  
@gðx; yÞ @ 1
¼ log jx  yj ; xAO; yA@O; ð3:5Þ
>
> @n y @n y 2p
>
>
>
> R R 1
>
:
O gðx; yÞ dy ¼ O 2p log jx  yjdy;

where O is an open, bounded, connected, smooth set in R2 (the container of the


mixture) and dx ðyÞ is the Dirac d supported at xAO:
From classical elliptic theory [24] one has the estimates
 
@gðx; yÞ
  2
jgðx; yÞjpC logðdistðx; @OÞÞ;  @y pC log ðdistðx; @OÞÞ: ð3:6Þ

Proposition 3.1. We consider the system of ODEs


8  
>
> d Ri 2 1 1 1
< dt ¼ jlog jj R R
>  ; Ti1 otoTi ; i ¼ 1; y; N;
i % Ri
  ð3:7Þ
>
> dxi 4 P 1 1 xh  xi
>
: dt ¼ jlog jj  ;
hai
%
R Rh jxh  xi j2
ARTICLE IN PRESS
18 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

where

1 1 XN
1
¼
% N
R R j
j¼1

and we assume that R1 ð0ÞoR2 ð0Þo?oRN ð0Þ:


For system (3.7) the following properties hold true:
(i) If Ri ð0ÞoRj ð0Þ then Ri ðtÞoRj ðtÞ on their common domain of existence.
d PN 2 
(ii) R ðtÞ ¼ 0:
dt  i¼1 i 
d PN
(iii) i¼1 Ri ðtÞ p0:
dt
(iv) R1 ðtÞ is nonincreasing in time and RN ðtÞ is nondecreasing.
(v) If we assume RN1 ð0ÞoRN ð0Þ then all except the Nth particle get extinct in finite
times T1 o?oTN1 and we have the estimate

jlog jj 3 jlog jj 3 N RN ð0Þ


R1 ð0ÞpT1 p R1 ð0Þ
6 6 RN ð0Þ  R1 ð0Þ

1
(vi) System (3.7) has a scale invariance compatible with the t3 law, that is if RðtÞ; xðtÞ
is a solution then so is mRðmt3 Þ; mxðmt3 Þ; m40:

Proof. (i) Suppose that there exists t 40 such that Ri ðt Þ ¼ Rj ðt Þ40: Then we
observe that Ri ðtÞ and Rj ðtÞ satisfy the same equation. By uniqueness, we can
conclude that Ri ð0Þ ¼ Rj ð0Þ; contradicting the assumption.
(ii) Between extinction times we have

!  
d X N
2
XN
’i ¼
X N
1 2 1 1
Ri ðtÞ ¼ 2Ri R 2Ri 
dt i¼1 Ri jlog jj R% Ri
i¼1 i¼1
!
N  
4 X 1 1 4 N X N
1
¼  ¼  ¼ 0:
jlog jj i¼1 R % Ri jlog jj R % Ri
i¼1

(iii) Between extinction times we have

 
d X N XN
’ i ðtÞ ¼
XN
1 2 1 1
Ri ðtÞ ¼ R 
dt i¼1 Ri jlog jj R % Ri
i¼1 i¼1
! !
2 1X N
1 X N
1 2 N X N
1
¼  ¼  p0:
jlog jj R % Ri 2 %2
jlog jj R 2
i¼1 i¼1 Ri i¼1 Ri
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 19

(iv) We have that

1 1X
N
1 1 N 1
¼ p ¼ :
%
R N j¼1 Rj N R1 R1

Analogously,

1 1
X :
% RN
R

So,

1 1 1
% ðtÞpRN ðtÞ:
p p 3R1 ðtÞpR
% R1
RN R

The result follows immediately from the above expression and the use of system
(3.7).
(v) We have that
   
d R1 ðtÞ 2 1 1 1 2 1 R1 2 1
¼  ¼ 1 X :
dt %
jlog jj R1 R R1 2
jlog jj R1 R % jlog jj R21

By integration, we obtain the inequality on the left

6
R31 ðtÞX  t þ R31 ð0Þ:
jlog jj

For the inequality on the right, we have the estimate


     
1 1 1 1 1 N 1 1 1 1 1
 ¼ þ?þ  ¼  þ?þ 
% R1 N R1
R RN N R1 N R2 R1 RN R1
   
1 1 1 1 R 1  RN
p  ¼ :
N RN R 1 N R 1 RN

By making use of this estimate, we compute


   
d R1 ðtÞ 2 1 1 1 2 1 1 R1  R N
¼  p
dt jlog jj R1 R % R1 jlog jj R1 N R1 RN
   
2 1 1 R 1  RN 2 1 1 R1 ð0Þ  RN ð0Þ
p p :
jlog jj N R21 RN jlog jj N R21 RN ð0Þ

Integrating we obtain the right-hand side in (v).


(vi) Straightforward verification.
The proof of Proposition 3.1 is complete. &
ARTICLE IN PRESS
20 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

4. The operators T; L and A and the mean curvature in n; q; r coordinates

The operator T: Consider OCR2 a bounded, smooth, connected set in R2 and G a


C 1þa closed, orientable surface in O; O represents the part of O enclosed by G and
Oþ ¼ O\O %  : Set
 
@u @uþ @u
Tf ¼  þ þ ¼: ; xAG: ð4:1Þ
@n @n @n G
R
For GAC 1þa ; f : G-R a sufficiently regular function satisfying G f ¼ 0 and u ; uþ
are the solutions to the Dirichlet problems

Du ¼ 0; xAO ;
u ¼ f; xAG;

8
> Duþ ¼ 0; xAOþ ;
>
< þ
u ¼ f; xAG;
>
> @u 
: ¼ 0; xA@O;
@n
-
where n ; nþ are the outward normals to @O ; @Oþ and n is the outward normal to
@O: T is the Dirichlet–Neumann operator
 
@u
Tf ¼ ¼c ð4:2Þ
@n G

and is invertible in the class


Z Z
c¼ f ¼ 0;
G G

where the inverse is given by


Z Z Z
1
ðScÞðxÞ ¼ gðx; yÞcðyÞ dy  gðx; yÞcðyÞ dy dx: ð4:3Þ
G jGj G G

The operator S can Rbe interpreted as the inverse of the restriction T to the set of
functions satisfying G f ¼ 0: In fact if
Z Z Z
1
uðxÞ ¼ gðx; yÞcðyÞ dy  gðx; yÞcðyÞ dy dx; xAO; ð4:4Þ
G jGj G G
!!@u""
then u is harmonic in O ; Oþ ; @n G
¼ c and so S is the Neumann–Dirichlet
operator.
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 21

R 1 1
Let X ¼ ffAL2 ðGÞ; G f ¼ 0g: We denote by X 2 ; X 2 the closures of
R
ffAC 1 ðGÞ; G f ¼ 0g under the norms
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
jj  jj 1 ¼ /S; SL2 ðGÞ ; jj  jj1 ¼ /T; SL2 ðGÞ ; ð4:5Þ
2 2

1 1
where /; SL2 ðGÞ denotes the standard inner product in L2 ðGÞ; X 2 ; X 2 are Hilbert
spaces and Eq. (4.5) implies that

jjfjj1 ¼ jjcjj 1 ð4:6Þ


2 2

and so T; S are isometries,


1 1
T : X 2 -X 2 ;
1 1
S : X 2 -X 2

with

ST ¼ id 1; TS ¼ id 1: ð4:7Þ

X2 X 2

The operator L: L is the classical Jacobi operator

L ¼ D G0 þ k 2 ; ð4:8Þ

where G0 ¼ S1 ; DG0 is the Laplace–Beltrami operator on G0 and k2 is the principal


curvature of G0 :
The operator A: The linearized Mullins–Sekerka operator A at G0 ¼ S1 is given by

A ¼ TL ð4:9Þ

considering conservative perturbations along the normal direction and where T; L


are defined as above.

Remarks.
* In what follows, we use the operator T0 defined by

@u @uþ
T0 X ¼ þ in S 1 ; X AC 1þa ðS1 Þ;
@n @nþ
where u ; uþ are the harmonic functions determined by
(
Du ¼ 0 on B1 ¼ fxAR2 =jxjo1g;
u ¼ X on S 1 ;
ARTICLE IN PRESS
22 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

8
> þ
< Du ¼ 0 on R2 \B% 1 ;
uþ ¼ X on S 1 ;
>
:
limx-N uþ ¼ 0:

T0 is the analog of T for G ¼ S 1 and O replaced by R2 and it holds that

@u @uþ n
T0 Yn ¼ þ ¼ Yn ;
@n @nþ 2

where Yn are the spherical harmonics of degree n [23].


* The linearization of the one phase Mullins–Sekerka operator for general inter-
faces G has been determined by Kimura [26], for the sphere it can be found in [21].
More information about T; L; A operators and their spectrum can be found in [5].

In what follows, we prefer to rescale R and r according to R ¼ er; r ¼ er and


introduce the quantities r; r which are of order 1.
Eq. (2.1) reads

GðtÞ ¼ fx=x ¼ x þ erð1 þ erðuÞÞu; uAS1 g: ðÞ

Instead of (), we could more precisely scale R ¼ er; r ¼ er with two independent
small parameters e1 ; e2 and write

GðtÞ ¼ fx=x ¼ x þ e1 rð1 þ e2 rðuÞÞu; uAS 1 g:

Going through the various steps of the proof of Proposition 6.1, it appears that the
whole argument applies to this more general situation. We limit ourselves to case ()
where e1 ¼ e2 ¼ e because all our estimates were originally done under this
assumption and also to keep the notation of already complicated computations
into reasonable limits.
The mean curvature in x; r; r coordinates: We would like to derive an expression for
the mean curvature HðX ðuÞÞ of G at a point X ðuÞ assuming that G ¼ fx=x ¼
x þ erð1 þ erðuÞÞu; uAS 1 g with rAC 1þa ðS 1 Þ:

Proposition 4.1. The mean curvature HðX ðuÞÞ of G has the form

1
HðX ðuÞÞ ¼ ð1  eLr þ BÞ; ð4:10Þ
er

where L is the classical Jacobi operator on S 1 as described above and B is an operator


 1
2
of the form B ¼ bðer; eGr; eG 2 rÞ with G ¼ ð1 þ erÞ2 þ e2 r2W and bðz; p; PÞ is a linear
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 23

function in P that satisfies the estimate


 
jbðz; p; PÞjpC jzj2 þ jpj2 þ ðjzj þ jpjÞjPj for jzjod:

Proof. H as a function of r; W (r; W polar coordinates) has the following expression:

1 cos W
H¼ 1
1 þ er  erWW  erW
erð1 þ erÞG 2 sin W
%
1
þ ferW ½eð1 þ erÞrW þ e2 rW rWW g ; ð4:11Þ
G

where

 1
2
G ¼ ð1 þ erÞ2 þ e2 r2W :

The desired result is obtained by (4.11) and also by using the expression of the
Laplace–Beltrami operator Ds r on S 1 which takes the form sin1 W ððrW sin WÞW Þ
[19]. &

5. The coordinate system

The definition of coordinate system is very important. Given an interface close to


circular, we associate to it in a unique way a circle and view the interface as a small
perturbation of that circle. Specifically, if the interface is already circular, the
procedure associates the same circle. So, to each interface in a certain class we will
give a center and a radius. There are many different coordinate systems that can be
used to accomplish this. The important fact about the coordinate system comes from
the way we intend to utilize it which is for studying the global stability properties of
the circular shape for a class of geometric operators related to the mean curvature.
By Sx;r CR2 we denote the circle of center x and radius r:

Proposition 5.1. Given an interface G in a sufficiently small C 1 neighborhood of a circle


Sx;* r* ; there are unique xAR2 ; r40; rAC 1 ðS1 Þ such that

G ¼ fx=x ¼ x þ erð1 þ erðuÞÞu; uAS1 g; ð5:1Þ

Z
rðuÞ du ¼ 0; ð5:2Þ
S1
ARTICLE IN PRESS
24 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Z
rðuÞ/u; ei S du ¼ 0; i ¼ 1; 2; ð5:3Þ
S1

where e1 ¼ ð1; 0Þ; e2 ¼ ð0; 1Þ and /; S is the euclidian inner product in R2 :

Proof. The representation of all G’s in a C 1 neighborhood of Sx;* r* is given by the


form

G ¼ fx=x ¼ x* þ erð1
* þ er̃ðũÞÞũ; ũAS1 g:

Choosing the new origin at x and for all x’s in the neighborhood of x* we have an
alternative expression

G ¼ fx=x ¼ x þ erð1 þ erðuÞÞu; uAS1 g;

* r;
where x; r; r and x; * r̃ are related through

x* þ erð1
* þ er̃ðũÞÞũ ¼ x þ erð1 þ erðuÞÞu: ð5:4Þ

From the above equation we have

x*  x þ erð1
* þ er̃ðũÞÞũ
u¼ ð5:5Þ
* * þ er̃ðũÞÞũj
jx  x þ erð1

and

erð1 þ erðuÞÞ ¼ jx*  x þ erð1


* þ er̃ðũÞÞũj: ð5:6Þ

Condition (5.2) is equivalent to taking


Z
1
er ¼ jx*  x þ erð1
* þ er̃ðũðuÞÞÞũðuÞ du ð5:7Þ
2p S1

and x is the remaining free variable while the map ũ-u for jx*  xj small is a C 1
diffeomorphism. What we would like to show next is to choose x such that Eq. (5.3)
is satisfied. In other words,
Z
jx*  x þ erð1
* þ er̃ðũÞÞũj/u; ei Sdu ¼ 0; i ¼ 1; 2: ð5:8Þ
S1

Hence, (5.8) is equivalent to solving the system


Z
* er̃Þ :¼
0 ¼ Fi ðx; er; jx*  x þ erð1
* þ er̃ðũÞÞũj/u; ei Sdu ¼ 0; i ¼ 1; 2; ð5:9Þ
S1
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 25

* er̃Þ is implicitly defined by (5.5). So, we seek to solve


where ũ ¼ ũðu; x; er;

* er̃Þ ¼ 0;
Fi ðx; er; i ¼ 1; 2

and we observe that


* er;
Fi ðx; * 0Þ ¼ 0:

We would like to employ the implicit function theorem. For this purpose we need to
calculate
Z
De Fi ðx; * 0Þx# ¼ Dx
* er; jx* þ erð1 #
* þ er̃ðũÞÞũj/u; ei SduðxÞ
S1
Z
¼ Dx /x* þ erð1 #
* þ er̃ðũÞÞũ; uS /u; ei SduðxÞ
S1
Z
¼ /  x# þ erD # uS /u; ei Sdu; xAR
* e ũx; # 2
; ð5:10Þ
S1

where Dx is the gradient of Fi with respect to the first entry and we have set
* er;
Dx ũ ¼ Dx ũðu; x; * 0Þ:

Moreover, by differentiation of (5.5) with r̃ ¼ 0; for x ¼ x* we get

* x ũx#  /  x# þ erD
x# þ erD # uSu ¼ 0:
* x ũx; ð5:11Þ

# uS ¼ 0 we obtain from (5.11) that


By using the fact that /Dx ũx;

* x ũx# ¼ x#  /x;
erD # uSu: ð5:12Þ

From (5.12) and (5.10) we conclude


Z
* 0Þx# ¼ 
* er;
Dx Fi ðx; #
/u; xS/u; ei Sdu ¼ px# i : ð5:13Þ
S1

So, the implicit function theorem applies for jjer̃jjC 1 ðS1 Þ od; for d40 and claims that
* erÞ
Eq. (5.9) has a unique solution x ¼ xðer̃; x; * such that x ¼ xð0; x; * erÞ * &
* ¼ x:

Remarks.
* It is important to note that rðuÞ is the distortion from circularity and we would
like it to say under control during the evolution. By imposing condition (5.3) we
remove from it the element corresponding to translation and so we have r under
control. The spectrum of the restricted operator is also stable, something which
reflects the stability of the circular shape and suggests that the coordinate system
will be preserved along evolution. In order to see the meaning of condition (5.3),
we observe that the translate Sxþde* i ;r% of Sx;
* *
% r% by d; where d40 is a small number, in
ARTICLE IN PRESS
26 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

*
% þ ðr% þ d/u;
% i ;r% ¼ fx=x ¼ x
the direction ei is given by Sxþde ei S þ Oðd* 2 ÞÞu; uAS 1 g:
Therefore if Q is an operator which has Sx;% r% and all its translates as equilibria then
/u; ei S has to be a zero eigenfunction of the linearization DQ0 of Q at Sx;% r% : The
Mullins–Sekerka operator has this property and so (5.3) are orthogonality
conditions. Condition (5.2) implies that, to principal order, r is the average radius
of G:
* Conditions (5.2) and (5.3) introduce 3 constraints but we also have 3 parameters
xAR2 and r40:
* We can find similar coordinate systems in [2,11] concerning three and two
dimensions, respectively.

% for given H
6. Solving the linear equation SðV Þ ¼ H  H

As it was mentioned in Section 3, instead of solving the two-space dimension


Mullins–Sekerka problem, we will solve the integral equations along the boundaries
of the evolving domains. We have the integral formulation
Z Z Z
1
gðx; yÞV ðyÞ dSy  %
gðx; yÞV ðyÞ dSy dSx ¼ HðxÞ  H: ð6:1Þ
GðtÞ jGðtÞj GðtÞ GðtÞ
S
Throughout this section, we write G instead of GðtÞ and we take G ¼ N i¼1 Gi with
Gi ¼ fx=x ¼ X ðuÞ :¼ xi þ eri ð1 þ eri ðuÞÞu; uAS g: For e40 small, the map X i :
i 1

S 1 -Gi is a diffeomorphism with the same regularity as ri : We let ui : Gi -S 1 be the


inverse of X i : Eq. (6.1) can be written in the form
N Z
X
gðx; yÞVh ðuh ðyÞÞ dy ¼ HðxÞ  E; xAGi ; i ¼ 1; y; N;
h¼1 Gh

where Z Z
1
E ¼ H%  gðx; yÞV ðyÞ dy dx
jGj G G

and Vh ðuh ðÞÞ is the restriction of V to Gh : We can see Eq. (6.1) as an equation in V
which is nonlinear in G due to the nonlinear dependence of H on the representation
of the interface. We are going to solve the above linear equation for H given and
then we are going to derive an expression for V in terms of x; r and r:

Proposition 6.1. Let xi AO; ri 40; ri AC 1þa ðS 1 Þ; i ¼ 1; y; N be given and assume


xi axj for iaj: Then the system
N Z
X
gðx; yÞVh ðuh ðyÞÞ dy ¼ HðxÞ  E; xAGi ; i ¼ 1; y; N ð6:2Þ
h¼1 Gh
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 27

has a unique solution Vi AC a ðS 1 Þ and

jjeri Vi  eri V% i þ KjjC a ðS1 Þ pjjF jj; ð6:3Þ

where
  !
1 1 1 jjVh jj
V% i ¼   E þ O sup 2
eri jlog ej eri h jlog ej

and

X
N
1 2 2
K¼ e rh log jxi  xh j
hai;h¼1
2p
Z X
N Z
1
 Vh ðvÞ dv þ e2 r2h gðxi ; xh Þ Vh ðvÞ dv  T0 Lri þ e2 T0
S1 h¼1 S1 eri
Z
1
 log jeri u  eri vjð3ri ðuÞ  ri ðvÞÞV% i dv
S1 2p

and F includes precise estimates for higher order terms


 
2 2 2 @gðxi ; xi Þ 2
jjF jj ¼ OC 1þa ðS1 Þ e jjri jjC 1þa ðS1 Þ rh gðxi ; xi Þ þ e ri þ e jjri jjC 1þa 2ðS1 Þ erh jjVi jjC a ðS1 Þ
@x
 
eri þ erh
þ erh OC 1þa ðS1 Þ ejjrh jjC 1þa ðS1 Þ log jxi  xh j þ jjVh jjC a ðS1 Þ
jxi  xh j
 
@gðxi ; xh Þ
þ erh OC 1þa ðS1 Þ erh þ ejjrh jjC 1þa ðS1 Þ gðxi ; xh Þ jjVh jjC a ðS1 Þ :
@y

Proof. We have mentioned that the Green’s function has the form

1
gðx; yÞ ¼  log jx  yj þ gðx; yÞ
2p
and we have
Z Z
h 1
gðx; yÞVh ðu ðyÞÞ dy ¼  log jx  yjVh ðuh ðyÞÞ dy
Gh Gh 2p
Z
þ gðx; yÞVh ðuh ðyÞÞ dy: ð6:4Þ
Gh

In what follows, we employ the notation f ¼ OC a ðS1 Þ ðjjrjjC 3þa ðS1 Þ Þ meaning that the
C a ðS1 Þ norm of f is OC a ðS1 Þ ðjjrjjC 3þa ðS1 Þ Þ:
ARTICLE IN PRESS
28 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Step 1: We consider the case h ¼ i; xAGi and we are interested in the analysis of
the two integrals on the right-hand side of Eq. (6.4).
(a) We start our analysis with the study of the first integral on the right-hand side
of Eq. (6.4).
Consider the case h ¼ i; xAGi : Let Oi ¼ fz=z ¼ lu; 0plo1 þ erðuÞ; uAS 1 g and
consider the function U i : Oi -R defined by
Z
1
U i ðzÞ :¼  log jxi þ eri z  yjVi ðui ðyÞÞ dy
2p
Gi
Z
1
¼ eri  log jeri z  eri z0 jVi ðui ðxi þ eri z0 ÞÞ dz0 : ð6:5Þ
@Oi 2p

As soon as @Oi is a surface of class C 1þa ðS 1 Þ; ri AC 1þa ðS1 Þ; we have by Theorem 2.I,
p. 307 in [29] applied to the derivatives of U i ; that U i can be extended as a C 1þa
function to the closure O% i of Oi and we have the estimate

jjU i ðÞjjC 1þa ðO% i Þ peri CjjVi ðui ðxi þ eri ÞÞjjC a ð@Oi Þ ; ð6:6Þ

where C is a constant independent of r under the assumption jjri jjode: Moreover, the
map zA@Oi -ui ðxi þ eri zÞAS1 is a C 1þa diffeomorphism and

jjui ðxi þ eri ÞjjC 1þa ð@Oi Þ oC̃ð1 þ ejjri jjC 1þa ðS1 Þ ÞoC

while we have a similar estimate for the inverse map u-z:

jjVi ðui ðxi þ eri ÞÞjjC 1þa ð@Oi Þ pCjjVi jjC a ðS1 Þ : ð6:7Þ

From (6.5)–(6.7), we obtain that we have a map Vi AC a ðS 1 Þ-U i j@Oi AC 1þa ðS 1 Þ and
together with the above properties of the diffeomorphism u-zðuÞ :¼ X i ðuÞ  xi ; we
can define a map I1i : C a ðS1 Þ-C 1þa ðS 1 Þ as follows:

eri ðI1i Vi ÞðuÞ ¼ U i ðX i ðuÞ  xi Þ: ð6:8Þ

Moreover, from Eqs. (6.6) and (6.8) we obtain the estimate

jjI1i Vi jjC 1þa ðS1 Þ pCjjVi jjC a ðS1 Þ ð6:9Þ

Our purpose is to compute the main term in I1i Vi : From (6.5) and

dz0 ¼ ð1 þ 2eri þ OC a ðS1 Þ ðe2 jjri jj2C 1þa ðS1 Þ ÞÞ du ð6:10Þ


ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 29

we have that
Z
1
ðI1i Vi ÞðuÞ ¼  log jX i ðuÞ  X i ðvÞjð1 þ 2eri ðvÞ þ OC a ðS1 Þ ðe2 jjri jj2C 1þa ðS1 Þ ÞÞðvÞVi ðvÞ dv
S1 2p
Z Z
1 1
¼  log jeri u  eri vjVi ðvÞ dv  e log jeri u  eri vj2ri ðvÞVi ðvÞ dv
S 1 2p S 1 2p
Z
1
 e2 log jeri u  eri vjOC a ðS1 Þ ðjjri jj2C 1þa ðS1 Þ ÞðvÞVi ðvÞ dv
S 1 2p
Z  
1 log jeri u  eri vj  log jX i ðuÞ  X i ðvÞj
e log jeri u  eri vj
S 1 2p log jX i ðuÞ  X i ðvÞj
 ð1 þ 2eri þ OC a ðS1 Þ ðe2 jjri jj2C 1þa ðS1 Þ ÞÞVi ðvÞ dv: ð6:11Þ

By Mirandd [29], I1i Vi and the first 3 integrals on the right-hand side of (6.11) are
C 1þa ðS 1 Þ functions. Also the last integral on the right-hand side of (6.11) belongs to
C 1þa ðS 1 Þ:
Let I Vi be the last integral. We have
Z
1
ðI Vi ÞðuÞ ¼  log jeri u  eri vj
S 1 2p
 
log jeri u  eri vj  log jeri u  eri v þ e2 ðri ri ðuÞu  ri ri ðvÞvj

log jeri u  eri v þ e2 ðri ri ðuÞu  ri ri ðvÞvj
 ð1 þ 2eri þ OC a ðS1 Þ ðe2 jjri jj2C 1þa ðS1 Þ ÞÞðvÞVi ðvÞ dv

¼: ðI 1 Vi ÞðuÞ þ ðI 2 Vi ÞðuÞ þ ðI 3 Vi ÞðuÞ: ð6:12Þ

After estimating ðI 1 Vi Þ; ðI 2 Vi Þ; ðI 3 Vi Þ; we conclude that


Z
1
ðI Vi ÞðuÞ ¼  log jeri u  eri vjðri ðuÞ þ ri ðvÞÞVi ðvÞ dv
S 1 2p
þ eOC 1þa ðS1 Þ ðjjri jj2C 1þa ðS1 Þ ÞjjVi jjca ðS1 Þ : ð6:13Þ

(b) We now study the second integral on the right-hand side of Eq. (6.4)
Z
gðX i ðuÞ; yÞVi ðui ðyÞÞ dy
Gi
Z
¼ eri gðX i ðuÞ; xi þ eri zÞVi ðui ðxi þ eri zÞÞ dz ¼: eri ðI2i Vi ÞðuÞ: ð6:14Þ
@Oi

The definition of X i ðuÞ implies that


 
@gðxi ; xi Þ
gðX i ðuÞ; eri zÞ ¼ gðxi ; xi Þ þ OC 1þa ðS1 Þ eri : ð6:15Þ
@x
ARTICLE IN PRESS
30 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

From (6.14) and (6.15) it follows


Z  
@gðxi ; xi Þ
I2i Vi ¼ gðxi ; xi Þ Vi ðvÞdv þ OC 1þa ðS1 Þ ejjri jjC 1þa ðS1 Þ gðxi ; xi Þ þ eri
S1 @x
 jjVi jjC a ðS1 Þ : ð6:16Þ

Eqs. (6.16) and (6.7) imply the estimate

jjI2i Vi jjC 1þa ðS1 Þ pCjjVi jjC a ðS1 Þ : ð6:17Þ

So, from Eqs. (6.4), (6.11), (6.13), (6.16) we conclude


Z Z  
i i 1
gðX ðuÞ; yÞVi ðu ðyÞÞ dy ¼ eri  logjeri u  eri vj þ gðxi ; xi Þ Vi ðvÞ dv
Gi S1 2p
þ eri ðI ii Vi ÞðuÞ; ð6:18Þ

where I ii satisfies

jjI ii Vi jjC 1þa ðS1 Þ pCjjVi jjC a ðS1 Þ ð6:19Þ

and
 
@gðxi ; xi Þ
C ¼ OC 1þa ðS1 Þ ejjri jjC 1þa ðS1 Þ gðxi ; xi Þ þ eri : ð6:20Þ
@x

Step 2: We now consider the case hai; where xAGi and yAGh : Again our aim is to
analyze the two integrals on the right-hand side of Eq. (6.4).
For hai and X ¼ X i ðuÞ both integrals on the right-hand side of Eq. (6.4) have as
functions of uAS1 the same smoothness as X i : The first integral on the right-hand
side of Eq. (6.4) gives
Z
1
 log jX i ðuÞ  yjVh ðuh ðyÞÞdy
Gh 2p
Z
1
¼ erh  log jX i ðuÞ  X h ðvÞjð1 þ 2 erh ðvÞ þ OC a ðS1 Þ ðe2 jjrh jj2C 1þa ðS1 Þ ÞÞVh ðvÞ dv
S 1 2p
Z
1
¼ erh  logjxi  xh jVh ðvÞ dv
S 1 2p
 
xr þ erh
þ erh OC 1þa ðS1 Þ ejjrh jjC 1þa ðS1 Þ log jxi  xh j þ i jjVh jjC a ðS1 Þ : ð6:21Þ
jxi  xh j
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 31

The second integral on the right-hand side of Eq. (6.4) implies


Z
gðX i ðuÞ; yÞVh ðuh ðyÞÞ dy
Gh
Z
¼ erh gðX i ðuÞ; X h ðvÞÞð1 þ 2erh ðvÞ þ OC a ðS1 Þ ðe2 jjrh jj2C 1þa ðS1 Þ ÞÞVh ðvÞ dv
S1
Z 
@gðxi ; xh Þ @gðxi ; xh Þ
¼ erh gðxi ; xh Þ Vh ðvÞdv þ erh OC 1þa ðS1 Þ eri þ erh
S1 @x @y

þ ejjrh jjC 1þa ðS1 Þ gðxi ; xh Þ jjVh jjC a ðS1 Þ ; ð6:22Þ

where the following expansion has been used:


 
@gðxi ; xh Þ @gðxi ; xh Þ
gðX i ðuÞ; X h ðvÞÞ ¼ gðxi ; xh Þ þ OC a ðS1 Þ eri þ erh : ð6:23Þ
@x @y

From (6.21), (6.22) we conclude that (6.4) takes the form


Z
gðX i ðuÞ; yÞVh ðuh ðyÞÞ dy
Gh
Z Z
1
 log jX i ðuÞ  yjVh ðuh ðyÞÞ dy þ gðX i ðuÞ; yÞVh ðuh ðyÞÞ dy
Gh 2p Gh
Z  
1
¼ erh  log jxi  xh j þ gðxi ; xh Þ Vh ðvÞ dv þ erh ðI ih Vh ÞðvÞ; ð6:24Þ
S1 2p

where
 
ih eri þ erh
I Vh ¼ OC 1þa ðS1 Þ ejjrh jjC 1þa ðS1 Þ log jxi  xh j þ jjVh jjC a ðS1 Þ
jxi  xh j
 
@gðxi ; xh Þ
þ OC 1þa ðS1 Þ erh þ ejjrh jjC 1þa ðS1 Þ gðxi ; xh Þ
@y

and I ih is a linear operator that satisfies

jjI ih Vh jjpCjjVh jjC a ðS1 Þ ; hai: ð6:25Þ

Step 3: We compute V% i : the average of Vi on the sphere S1 ; by utilizing the integral


representation (6.2) and the expressions obtained in Steps 1 and 2.
We saw that Eq. (6.1) can be written in the form

N Z
X
gðx; yÞVh ðuh ðyÞÞ dy ¼ HðxÞ  E; xAGi ; i ¼ 1; y; N
h¼1 Gh
ARTICLE IN PRESS
32 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

equivalently

N Z
X
gðX i ðuÞ; yÞVh ðuh ðyÞÞ dy ¼ Hi ðxÞ  E; i ¼ 1; y; N:
h¼1 Gh

From Eqs. (6.18) and (6.24), we have


Z
1
 ðlog jeri u  eri vj þ 2pgðxi ; xi ÞÞ eri Vi ðvÞ dv
2p S1
Z
1 X
 ðlog jxi  xh j þ 2pgðxi ; xh ÞÞerh Vh ðvÞ dv
2p hai S1
X
þ eri ðI ii Vi ÞðuÞ þ erh ðI ih Vh ÞðuÞ ¼ Hi  E: ð6:26Þ
hai

From (6.26) we obtain


Z Z
1
log jeri u  eri vjeri Vi ðvÞ dv ¼  gðxi ; xi Þ eri Vi ðvÞ dv
2p S1 S1
Z
1 X
 ðlog jxi  xh j þ 2pgðxi ; xh ÞÞ erh Vh ðvÞ dv
2p hai S1
X
N
þ erh ðI ih Vh ÞðuÞ  Hi þ E: ð6:27Þ
h¼1

R
Let V% i denotes the average of Vi on the circle S 1 then by utilizing V% i and S 1 ðVi 
V% i Þ ¼ 0; Eq. (6.27) takes the form
Z Z
1 1
log jeri u  eri vjeri ðVi  V% i Þ dv ¼  log jeri u  eri vjeri V% i dv
2p S1 2p S1
 gðxi ; xi Þeri V% i

1 X
 log jxi  xh jerh V% h
2p hai
X
 gðxi ; xh Þerh V% h  Hi þ E
hai
X
N
þ erh ðI ih Vh ÞðuÞ: ð6:28Þ
h¼1

Then by the theory of single layer potential, we know that the left-hand side defines
an harmonic function which is bounded at infinity. We extend the right-hand side of
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 33

Eq. (6.28) in order to have also an harmonic bounded function at infinity


Z
1
log jeri u  eri vjeri ðVi  V% i Þ dv
2p S1
1 X
¼ eri V% i log r  gðxi ; xi Þeri V% i  log jxi  xh jerh V% h
2p hai
X X
N
 gðxi ; xh Þerh V% h þ c1 log r þ c2 þ erh ðI ih Vh ÞðuÞ: ð6:29Þ
hai h¼1

Our aim is to determine V% i : We will accomplish it by the following procedure: On the


boundary of the ith particle we have the leading order to condition
1
c1 log eri þ c2 ¼ 2 þ E: ð6:30Þ
eri

Moreover, in order to obtain that the right-hand side is bounded, we need to impose
the following conditions:

eri V% i ¼ c1 ; ð6:31Þ

1 X X
eri V% i gðxi ; xi Þ  log jxi  xh jerh V% h  gðxi ; xh Þerh V% h ¼ c2 : ð6:32Þ
2p hai hai

The above equations (6.30)–(6.32) form a system of 3N equations with 3N


unknowns. By substituting (6.31), (6.32) into (6.30) we obtain that
!
1 X X
%
ðlog jeri j þ gðxi ; xi ÞÞeri Vi   %
logjxi  xh jerh Vh  %
gðxi ; xh Þerh Vh
2p hai hai
1
¼E : ð6:33Þ
eri

Eq. (6.33) forms a diagonal dominant system for 0oe51 and it has a unique
solution which is given to principal order by the formula
  !
1 1 1 1
V% i ¼   E þ OC 1þa ðS1 Þ : ð6:34Þ
jlog ej eri eri jlog ej2

Step 4: As soon as the right-hand side of Eq. (6.4) has been analyzed and V% i has
been determined, we utilize the integral representation (6.2) in order to conclude on
to the desired result which is (6.3).
Using again theorem 2.I in [29], we observe that the function I ih Vh ; h ¼ 1; y; N
has a harmonic extension both to the interior Bi ¼ fx=jxjo1g and to the exterior
R2 \B% 1 of S 1 : These harmonic functions can be extended as a C 1þa functions to B1
ARTICLE IN PRESS
34 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

and R2 \B% 1 with the estimate

jjðI ih Vh Þ jjC 1þa ðB% 1 Þ p CjjI ih Vh jjC 1þa ðS1 Þ ;

jjðI ih Vh Þþ jjC 1þa ðR2 \B1 Þ p CjjI ih Vh jjC 1þa ðS1 Þ ;

where the subscripts 7 denote the extensions, and C is a universal constant. From
these inequalities, it follows that

jjT0 I ih Vh jjC a ðS1 Þ pCjjI ih Vh jjC 1þa ðS1 Þ :

After applying T0 to Eq. (6.28) and utilizing the expression for H given in
Proposition 4.1, Eq. (6.28) takes the form

X
N
1
eri Vi ¼ eri V% i  erh ðT0 I ih V% h Þ þ O 1þa 1 ðejjri jjC 1þa ðS1 Þ Þ  T0 Wi ; ð6:35Þ
h¼1
eri C ðS Þ

where Wi ¼ HjGi  E ¼ er1 ð1  eLri þ Bi Þ  E; i ¼ 1; 2; y; N and T0 Wi ¼


i
1 e
eri T0 Lri eri OC a ðS 1 Þ ðjjri jjC 1þa ðS 1 Þ jjri jjC 3þa ðS 1 Þ Þ:
For 0oe51; system (6.35) has a unique solution that can be computed by
ðnÞ ðnÞ
iteration. The approximate solution ðrV ÞðnÞ ¼ ðr1 V1 ; y; rN VN Þ at step n is
computed by solving Eq. (6.35) with rV ¼ ðrV Þðn1Þ in the right-hand side starting
with ðrV Þð0Þ ¼ 0; Moreover, Eqs. (6.11), (6.13), (6.14), (6.18), (6.19) imply
Z
1
jjI ii Vi  logjeri u  eri vjð3ri ðvÞ  ri ðÞÞVi ðvÞ dv  eri gðxi ; xi ÞV% i jjC 1þa ðS1 Þ
S 1 2p
 
@gðxi ; xi Þ 2
¼ OC 1þa ðS1 Þ ejjri jjC 1þa ðS1 Þ gðxi ; xi Þ þ eri þ e jjri jjC 1þa ðS1 Þ jjVi jjC a ðS1 Þ ð6:36Þ
@x

and from Eqs. (6.21), (6.22), (6.24), (6.25)


  Z 
 ih 
I Vh þ er 1 log jxi  xh j  gðxi ; xh Þ V ðvÞ dv 
 h
2p
h  1þa 1
S1 C ðS Þ
 
eri þ erh
¼ OC 1þa ðS1 Þ ejjrh jjC 1þa ðS1 Þ log jxi  xh j þ jjVh jjC a ðS1 Þ
jxi  xh j
 
@gðxi ; xh Þ
þ OC 1þa ðS1 Þ erh þ ejjrh jjC 1þa ðS1 Þ gðxi ; xh Þ
@y
 jjVh jjC a ðS1 Þ : ð6:37Þ

Inserting these expressions of I ii Vi ; I ih Vh into Eq. (6.29) we have the desired result
which is Eq. (6.3). &
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 35

As soon as V% i is known, we notice that the previous proposition reduces the


determination of Vi to a system that, to principle order is coupled only through E:
The coupling constant E can be determined by the conservation requirement
Z N Z
X
V¼ Vh ¼ 0: ð6:38Þ
G h¼1 Gh

Proposition 6.2. Condition (6.39) uniquely determines the constant E: Moreover,


 
1 XN
1 jjrjjC 1þa ðS1 Þ
E¼ þ OC 1þa ðS1 Þ 1 þ ; ð6:39Þ
N j¼1 erj r%

where

1 1 X
N
1
¼
er% N j¼1 erj

and jjrjjC 1þa ðS1 Þ is the norm of r ¼ ðr1 ; y; rN Þ:

Proof. By inserting the expression of V% i obtained from (6.34) into (6.3) under the
standing assumption that for e jjrh jjC 1þa ðS1 Þ od for d40 and by [29] we obtain

1
eri Vi ¼ eri V% i þ O a 1 ðjjri jjC 3þa ðS1 Þ Þ þ OC a ðS1 Þ ðe2 jjri jj2C 1þa ðS1 Þ Þ
ri C ðS Þ
1
þ Oðejjri jjC 1þa ðS1 Þ Þ: ð6:40Þ
eri

By utilizing the conservation requirement (6.38) we have


N Z
X X
N Z
0¼ Vi ¼ eri Vi ð1 þ 2eri þ OC 1þa ðS1 Þ ðe2 jjri jj2C 1þa ðS1 Þ ÞÞ du: ð6:41Þ
i¼1 Gi i¼1 S1

Substituting Eq. (6.41) into (6.40) and after some calculations we conclude
 
1 XN
1 jjrjjC 1þa ðS1 Þ
E¼ þ OC 1þa ðS1 Þ %
þ 1 þ eðjjrjjC 1þa ðS1 Þ þ rÞE :
N j¼1 erj r%

If e40 is smaller than some e%40 this equation can be solved for E yielding estimate
(6.39). &

We now give a decomposition result for a general V in terms of p; ’ rrt for


’ x;
interfaces with representation (5.1). We let V ¼ V ðu; tÞ to be the speed of GðtÞ in the
orthogonal direction to GðtÞ at the point xAGðtÞ and we study the relationship
between V and r; rrt and x: ’
ARTICLE IN PRESS
36 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Proposition 6.3. Assume that ejjrjjC 1þa ðS1 Þ od for d40 a small fixed number, so that
Proposition 5.1 holds. Then V is a linear combination of er; ’ e2 rrt ; x’ and the equation
a 2
V ¼ Z with ZAC ðS Þ a given function, determines uniquely er; ’ Moreover,
’ e2 rrt ; x:
the following estimates hold true:
8 pffiffiffi 
> j2 ’ þ /Z; S 2 ðS 2 Þ jpCe ejjrjj 1þa
2
>
>
>
p e r w 0 L C ðS 2 Þ jjZjjC a ðS 2 Þ
>
> 
>
> P 2
> þjjrjjC 1þa ðS2 Þ h¼1 j/Z; wh SL2 ðS2 Þ j ;
>
>
>
> 
>
> pffiffi
< j2 px’ j þ /Z; wj SL2 ðS2 Þ jpCe ejjrjj2 1þa 2 jjZjj a 2
3 C ðS Þ C ðS Þ
 ð6:42Þ
>
> P 2
>
> þjjrjj 1þa
C ðS Þ2 j/Z; w h S 2 2
L ðS Þ j ;
>
>
h¼1
>
> P2
>
> jje rrt þ Z  j¼0 /Z; wj SL2 ðS2 Þ wj  e/Z; w0 SL2 ðS2 Þ w0 rjjC a ðS2 Þ
2
>
>
>
>  P 
>
: pCe ejjrjj2 1þa 2 jjZjj a 2 þ jjrjj2 a 2 2 j/Z; wh SL2 ðS2 Þ j
C ðS Þ C ðS Þ C ðS Þ h¼1

for some constant C40 and wj ; j ¼ 0; 1; 2 defined as follows: w0 ¼ p1ffiffi2p; w1 ¼


p1ffiffi /u; e1 S ¼ p1ffiffi cos W; w2 ¼ p1ffiffi /u; e2 S ¼ p1ffiffi sin W:
2p 2p 2p 2p

Proof. We refer the reader to [5] for a detailed proof which can be easily adapted to
two dimensions. &

Proposition 6.4. There is e% 40 such that for joe% 2 Eq. (6.1) is equivalent to the
following system of evolution equations:
8  
>
> dri 2 1 1 1
>
> e ¼  þ fir ðr; x; rÞ;
>
> dt jlog ej er e %
r er
>
<  i i

dri 2 1 1
¼ Ãr i  ð3r i  T r
0 i Þ þ fir ðr; x; rÞ; ð6:43Þ
>
> dt jlog ej e 3 r3 e3 r2 r
%
>
> i i
>
> dx
>
: i x
¼ fi ðr; x; rÞ;
dt
where

à ¼ T0 ðDs  IÞ þ 3I

and fir ðr; x; rÞ; fir ðr; x; rÞ; fix ðr; x; rÞ are smooth, uniformly bounded in e functions of
r ¼ ðr1 ; y; rN Þ; x ¼ ðx1 ; y; xN Þ and r ¼ ðr1 ; y; rN Þ and er% is the harmonic mean of
eri defined by

1 1 X
N
1
¼ :
er% N j¼1 erj
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 37

Proof. Eqs. (6.43) follow from Eqs. (6.42) and Proposition 6.1. More details can be
found in [5]. &

7. The R; r estimates

Now, we turn back to the old notation R ¼ er; r ¼ er and our aim is to provide
estimates for the R; r equations. Especially, the control on r is a very important
factor. In the previous sections we assume that jjrjjC 1þa ðS1 Þ od for d40 a fixed small
number. In this section, we are going to establish a uniform bound on jjrjjC 3þa ðS1 Þ : By
PN
pR2i
recalling that j ¼ i¼1jOj ; we have from Proposition 6.5 the following system of
evolution equations:
8  
>
> dRi 2 1 1 1
>
> ¼  þ fiR ðR; x; rÞ;
>
> dt jlogjj Ri R% Ri
>
<  
d ri 2 1 1
¼ Ãri  2 ð3ri  T0 ri Þ þ fir ðR; x; rÞ; ð7:1Þ
>
> dt jlogjj R3i
> Ri R%
>
>
>
: dxi ¼ f x ðR; x; rÞ:
>
i
dt

In [16,22] the local (in time) existence of classical solutions is established for the
Mullins–Sekerka model for arbitrary space dimensions. From Eqs. (7.1) with initial
conditions Ri ð0Þ; ri ð0Þ; xi ð0Þ we have a classical solution Ri ¼ Ri ðtÞ; ri ¼ ri ðtÞ; xi ¼
xi ðtÞ in some maximal interval of existence ½0; T̂Þ: We start the analysis of this section
by providing information which will be used in the proof of subsequent theorems.

Lemma 7.1. Let à be the operator defined in Eq. (6.43). Then à is a self-adjoint
1
operator on X 2 and the eigenvalues of à are given by

mn ¼ 2nð1  n2 Þ þ 3; n ¼ 1; 2; y; ð7:2Þ

where mn has multiplicity 2n and the corresponding eigenspace is spanned by the 2n


spherical harmonics Yn of degree n: Additionally, the eigenvalues of à restricted to the
subspace

1
X ¼ frAX 2 ; /r; wi SL2 ¼ 0; i ¼ 0; 1; 2; 3g

satisfy

mp  9: ð7:3Þ
ARTICLE IN PRESS
38 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Proof. The spherical harmonics of degree n are eigenfunctions of the operator T0 :

@ui @ue
T0 Yn ¼ þ ¼ 2nYn : ð7:4Þ
@ni @ne

Moreover, we have

Ds Yn ¼ n2 Yn ; ð7:5Þ

where the eigenvalue n2 has multiplicity n2 : From the definition of à and the
completeness of the set of spherical harmonics we have the desired result. &

Remarks. We will need to obtain estimates on r for

rðtÞ ¼ T0 Lr þ f ¼ Ãr þ f ðrðtÞÞ: ð7:6Þ

If r is a solution of rt ¼ Ãr þ f ðrðtÞÞ then r satisfies the ‘‘variation of constants


formula’’
Z t
Ãt
rðtÞ ¼ e rð0Þ þ eÃðtsÞ f ðrðsÞÞ ds: ð7:7Þ
0

Moreover, from well known properties of analytic semigroups [25] we have: If à is a


self-adjoint densely defined operator and if à is bounded below, then à is a sectorial
and if à is a sectorial operator then it is the infinitesimal generator of an analytic
semigroup, while the following estimates hold:

jjeÃs jjjC 3þa ðS1 Þ pMems jjjjjC 3þa ðS1 Þ ; jAE1 ; ð7:8Þ

M
jjeÃs jjjC 3þa ðS1 Þ p b Mems jjjC 2þa ðS1 Þ ; jAC 2þa ðS 1 Þ-E0 ; b ¼ 13; ð7:9Þ
s
where m; M40 constants. We can assume that the constant M satisfies M41: We
will be utilizing the semigroup setting of maximal regularity due to da Prato and
Grisvard, [8,9,28]. The result of this theory is:
AAM1 ðE0 ; E1 Þ if the following estimate holds:
Z s 
 

sup  e ÃðssÞ
jðsÞ ds pC% sup jjjðsÞjjC a ðS1 Þ ; ð7:10Þ
0psp%s 0 C 3þa ðS 1 Þ 0psp%s

where C40% is a constant independent of s%: In general, C% is dependent of s but here


from Lemma 7.1 the spectrum of à is bounded above by a negative number. Let
hkþa ðS 1 Þ be the ‘‘little’’ Hölder space and let

E0 ¼ ha ðS 1 Þ-X ; E1 ¼ haþ3 ðS 1 Þ-X : ð7:11Þ


ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 39

The ‘‘little’’ Hölder spaces ha ; 0oao1 are obtained by completing the C N functions
in the C a norm. More generally, the spaces of the little Hölder continuous functions
are defined by
( )
jjf ðtÞ  f ðsÞjj
ha ðI; X Þ ¼ f AC a ðI; X Þ : lim sup ¼0 ;
d-0 jtsjod jt  sja

hkþa ðI; X Þ ¼ ff ACbk ðI; X Þ : f ðkÞ Aha ðI; X Þg:

One checks immediately that C y ðI; X ÞCha ðI; X Þ for y4a: Moreover, if 0oao1 and
y4a then ha ðI; X Þ is the closure of C y ðI; X Þ in C a ðI; X Þ and the following
interpolation fact is true:

ðha ; h1þa Þy ¼ hð1yÞaþyð1aÞ :

Moreover, if AAM1 ðE0 ; E1 Þ we have a relationship between the fractional spaces and
the interpolation spaces Ey :

jjAy xjj0 pjjxjjEy ;

where

jjxjjEy ¼ sup m1y jjAeA xjj0


m40

and Ey is defined as
%
Ey ¼ xAE0 : lim m1y jjAemA xjj0 ¼ 0 ; 0oyo1:
m-0

The following proposition justifies Eqs. (7.1) by showing that Ri can be


approximated well by Ri ; the solutions of (3.7). We focus on the first extinction
interval ½0; T̂1 ). By repeating the argument, we obtain analogous results in
ðT̂1 ; T̂2 Þ; y; ðT̂N2 ; T̂N1 Þ:

Proposition 7.2. Assume NX2: Assume there exists z40 such that

jjri ðtÞjjC 3þa ðS1 Þ oz; tA½0; T̂Þ: ð7:12Þ

Assume also

R1 ð0ÞoR2 ð0Þo?oRN1 ð0Þ; Ri ð0Þ ¼ Ri ð0Þ; i ¼ 1; y; N

and let T1 be the extinction time of R1 characterised by R1 ðT1 Þ ¼ 0 and T̂1 ¼ T̂ be the
extinction time of R1 characterized by R1 ðT̂1 Þ ¼ 0: Then there exists e%40 such that
ARTICLE IN PRESS
40 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

joe% 2 and positive constants CT ; c1 ; C1 ; CR ; CR ; a depending on Ri ð0Þ; i ¼ 1; y; N


such that the following estimates hold true:
(i) jT̂  T1 joCT 1
;
jlog jj2
(ii) 1 1
c1 ðT̂  tÞ3 oR1 ðtÞoC1 ðT̂  tÞ3 ;
(iii) cR oRi ðtÞoCR ; tA½0; T̂Þ; i41;
(iv) jRi ðtÞ  Ri ðtÞjoCR 1 4 ; tA½0; T1  CT 1
;
jlog jj2
jlog jj3
(v) 1  R1 ðtÞ4a; tA½0; T̂Þ:
%
RðtÞ

Proof. (A) Eq. ð7:1Þ1 can be written in the form


   
dRi 2 1 Ri 1
¼  1 þ R O ; i ¼ 1; y; N: ð7:13Þ
R%
i
dt jlog jj R2i jlog jj

(B) The following estimates hold:


PN
(a) 2 2 1
i¼1 ðRi ðtÞ  Ri ð0ÞÞ ¼ Oðjlog jj2 Þ;
PN
(b) 2 2 1
i¼1 ðRi ðtÞ  Ri ðtÞÞ ¼ Oðjlog jj2 Þ;
(c) Ri : uniformly bounded in j; equicontinuous in j and uniformly continuous on
½0; T̂Þ:

 
d 1 PN P

PN
Verification: (a) R ðtÞ ¼ N
2
i¼1 Ri ðtÞRi ðtÞ ¼
2 1 Ri
i¼1 jlog jj Ri ð R%  1Þ þ
dt 2 i¼1 i
Oðjlog1jj2 Þ ¼ Oðjlog1jj2 Þ; by recalling Proposition 3.1(ii).
P
Integrating we obtain N 2 2 2
i¼1 ðRi ðtÞ  Ri ð0ÞÞpOðjlog jj2 Þ:
P
From this by Schwartz and Gronwall we obtain a bound on N 2
i¼1 Ri ðtÞ hence
PN
i¼1 Ri and as a result (a).
P
Condition (b) follows from (a) by utilizing the conservation of N 2
i¼1 Ri ðtÞ:
Condition (c) follows from (a).
(C) Let T̂ 4T1 arbitrary otherwise. Then

lim min R1 ¼ 0: ð7:14Þ


j-0 ½T1 ;T̂ 

We argue by contradiction. Assume that limj-0 inf ½T1 ;T̂  R1 40: Then
limj-0 inf ½T1 ;T̂  Rj Xc40; j ¼ 1; y; N: By continuous dependence as long as there
is no singularity in ½T1 ; T̂  we can pass to the limit in (7.13) and deduce that
R1 ðtÞXc40 in ½T1 ; T̂  contradicting that T1 oT̂ : Finally the lim can be replaced by
lim by the equicontinuity of R:
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 41

(D) We show that T̂1 oN; R1 40 on ½0; T̂1 Þ; T̂1 oT̂2 ; y; T̂N1 : We argue by
contradiction. Assume that T̂1 ¼ N; so T̂ ¼ N: By (7.13) for i ¼ 1; we have
    !
d 1 3 2 R1 1
R p 1 þO : ð7:15Þ
dt 3 1 jlog jj R% jlog jj2

1 ðt Þ
We fix a T̂ 4T1 : By (C) there exists t A½T1 ; T̂  such that RRðt
%  Þ o2 and hence over an
1

interval ½T̂ ; T̂ þ d which can be chosen uniformly in j (by the equicontinuity of
R1 ). It follows that R1 is decreasing on ½T̂ ; T̂ þ d by a fixed amount. By
equicontinuity we can repeat the argument over ½T̂ þ d; T̂ þ 2d; y to conclude
that T̂1 oN: Next, we consider Eq. (7.13) for i ¼ 1; 2: By subtracting them we have

1d 2 1 1 1
ðR3  R31 ÞX ðR2  R1 Þ  C XC on ½0; T̂1 Þ;
3 dt 2 jlog jj R% jlog jj2 jlog jj2

where we used that R2 4R1 : Integrating we obtain

1
R32 ðtÞXR32 ð0Þ  R31 ð0Þ  C t
jlog jj2

from which it follows that

1
R32 ðT̂1 ÞXR32 ð0Þ  R31 ð0Þ  C T̂1
jlog jj2

and as a result T̂2 4T̂1 for j: small. Similarly we show that


T̂N1 4T̂N2 4?4T̂2 4T̂1 : From which (iii) and (v) follow.
(E) We argue near T̂1 : Given d40; there exists d% such that for T̂1  d% otoT̂1 )
j % jod: Let tA½T̂1  d% ; T̂1 Þ; from (7.13) with i ¼ 1; we obtain
R1
R

    !  
d 1 3 2 R1 1 2 R1 1
R1 ¼ 1 þO p  1 þC
dt 3 jlog jj R% jlog jj 2 jlog jj R% jlog jj2
2 C 1
p ð1 þ dÞ þ o :
jlog jj jlog jj2 2

Integrating this inequality from t to T̂1 we obtain the inequality on the left-hand side
of (ii). To obtain the inequality on the right-hand side we utilize an upper bound on
jRR%1 j on ½T̂1  d% ; T̂1 Þ and we have
   
d 1 3 2 R1 1
R1 X 1 C X  C:
dt 3 jlog jj R% jlog jj2
ARTICLE IN PRESS
42 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

Integrating from t to T̂1 we obtain the right-hand side of estimate (ii).


(F) We have
  !
1d 3 3 2 Ri Ri 1
ðR  Ri Þ ¼  þO :
3 dt i jlog jj R% R% jlog jj2

By integration
Z  
2 t Ri Ri  1
jR3i ðtÞ  R3i ðtÞjp   dt þ C :
R% R% 
jlog jj 0 jlog jj2

Utilizing R1 pRi we obtain

X
N Z t X
N
2 1
R21 jRi  Ri jpC jRi  Ri jdt þ C :
i¼1
jlog jj 0 i¼1 jlog jj2

Set

X
N
yðtÞ ¼ R21 ðtÞ jRi  Ri j:
i¼1

Then
Z t
2 yðsÞ 1
yðtÞp 2
ds þ C :
jlog jj 0 R1 ðsÞ jlog jj2

Utilizing that
1 1
C1 ðT1  tÞ3 XR1 ðtÞXc1 ðT1  tÞ2 ;

we obtain that yðtÞpC jlog1jj2 hence

X
N
1
R21 ðtÞ jRi  Ri jpC :
i¼1 jlog jj2

So,

X
N
1
jRi  Ri jpC 2
:
i¼1 R21 ðtÞjlog jj

Now,
 2
1 3 1
R1 ðtÞXC if tpT1  CT
jlog jj jlog jj2
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 43

and so
0 1 " #
X
N
1 1
jRi  Ri jpC @ 4
A; tA 0; T1  CT :
i¼1 jlog jj3 jlog jj2

(G) We now prove (i). We first note that (ii) implies a lower bound on T̂1 :
Indeed, if T̂1 pT1  C jlog1jj2 ; CXCt then

1
jR1 ðT̂1 Þ  R1 ðT̂1 ÞjpCR 4
jlog jj3

1
hence jR1 ðT̂1 ÞjpCR 4:
jlog jj3
On the other hand by the lower bound on R1
4 1
cðT1  T̂1 Þ3 pCR 4
jlog jj3

4
which if cðcÞ3 4cR we arrive at a contradiction. Thus,

1
T̂1 XT1  c ; c : appropriate:
jlog jj2

To obtain an upper bound we argue as follows: From


 ! !
 C C  1
 
R1 T1   R 1 T 1  oCR ;
 jlog jj2 jlog jj2  4
jlog jj3

we obtain
!
C 1
R1 T1  2
oC  4
:
jlog jj jlog jj3

dt pð1
dR1
For tXT1  jlogCjj2 we have þ dÞR12 : Integrating we obtain
1

!!
1 1 C
0p R31 ðtÞpC 3  t T1 
3 jlog jj2 jlog jj2

from which we obtain an upper bound for t of the form T1 þ C jlog1jj2 : The proof of
the proposition is complete. &
ARTICLE IN PRESS
44 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

For convenience, we utilize the ‘‘pseudo-times’’ si in estimating ri in order to


handle the factor jlog2 jj R13 infront of Ãri in Eq. ð7:1Þ2 : The ‘‘pseudo-times’’ are defined
i
by
Z t
2 1
si ¼: 3
dt; i ¼ 1; y; N; tA½0; T̂Þ: ð7:16Þ
0 jlog jj Ri ðtÞ

Proposition 7.3. Assume NX2: Then there exists e%40 such that joe%2 ;
jjri ð0ÞjjC 3þa ðS1 Þ oe% and z independent of e% such that the following inequality holds true:

jjri ðtÞjjC 3þa ðS1 Þ oz; tA½0; T̂Þ i ¼ 1; y; N: ð7:17Þ

Proof. If we introduce the ‘‘pseudo-times’’ si defined in Eq. (7.16) then system (7.1)
takes the form
8   
>
> dRi Ri R
>
> ¼ R  1 þ g ð R; x; r Þ ;
R%
i i
>
> dsi
>
<
drit Ri 2
¼ Ãri  ð3ri  T0 ri Þ þ gri ðR; x; rÞ; ð7:18Þ
>
> ds i R% jlog jj
>
>
>
> dx
>
: i ¼ gxi ðR; x; rÞ;
dsi
where

gR R
i ¼ R i fi ; gri ¼ R3i fir ; gxi ¼ Ri fix :

1. Let z40 any number that satisfies

jjri ð0ÞjjC 3þa ðS2 Þ oz; i ¼ 1; y; N: ð7:19Þ

Then by continuity there exists T̂0 pT̂ such that the inequality jjri ðtÞjjC 3þa ðS1 Þ oz holds
in ½0; T̂0 Þ: Then estimates (iii), (v) in Proposition 7.2 hold in ½0; T̂0 Þ and we can also
assume jgR i joa: Therefore Eq. ð7:18Þ1 implies

R1 ðtðs1 ÞÞpR1 ð0Þeas1 : ð7:20Þ

Let ŝ1 be chosen large enough so that,


Z s1
R1 ð0Þ aŝ1 emðs1 sÞ 1
Mð3 þ jjT0 jjÞN e sup dso ð7:21Þ
RN ð0Þ s1 Xŝ1 ŝ1 ðs1  sÞb 4

for b fixed as in (7.9).


ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 45

2. In the interval ½0; T̂0 Þ we have from (7.16), (7.20) and estimate (iii) in Proposition
7.2
Z   Z
s1
R31 0 R1 ð0Þ 3 s1 3as1 0 0
si ¼ ds 1 p e ds1
0 R3i cR 0
 
1 R1 ð0Þ 3
p ¼ ŝ; i41: ð7:22Þ
3a cR

Let s%1 ¼ maxfŝ; ŝ1 g:


3. The map si -½RR%i ðri  T0 ri Þðtðsi ÞÞ is a continuous map from si ð½0; T̂0 ÞÞ1 into
C 2þa ðS 2 Þ and we have
 
R i  NCR
 ð3ri  T0 ri Þ
 R%  3þa 1 pR ð0Þ ð3 þ jjT0 jjÞjjri jjC 3þa ðS1 Þ : ð7:23Þ
C ðS Þ N

From this and (7.9) it follows that


Z si   
 Aðsi sÞ Ri

 e ð3ri  T0 ri Þ ðtðsÞÞ ds
 %
R
0 C 3þa ðS 1 Þ
NCR
pM ð3 þ jjT0 jjÞs1b
i sup jjri ðtðsÞÞjjC 3þa ðS1 Þ : ð7:24Þ
RN ð0Þ 0ospsi

We fix s40 small, so that

CR 1
MN ð3 þ jjT0 jjÞs1b o : ð7:25Þ
RN ð0Þ 4

4. Let k% ¼ ½s%s1  þ 1 where s%n ; s are defined in 3 and 4 and where [ ] stands for the
integer part.
5. Utilizing estimate (iii) in Proposition 7.2 we have for tA½0; T̂0 Þ

jjgri jjC a ðS2 Þ pC0 ðCR z þ zjjri jjC 3þa ðS1 Þ Þ:

Therefore from (7.10) it follows that


Z s 
 2 
sup   e ÃðssÞ r 
gi  
0ospsi 0 jlog jj C 3þa ðS 1 Þ
2 2
¼ % 0 CR ðCR þ zÞ þ
CC % 0 z sup jjri ðtðsÞÞjjC 3þa ðS1 Þ
CC
jlog jj jlog jj 0ospsi
2 2
p C1 ð1 þ zÞ þ C1 z sup jjri ðtðsÞÞjjC 3þa ðS1 Þ ; ð7:26Þ
jlog jj jlog jj 0ospsi

where C1 is a suitably chosen constant.


ARTICLE IN PRESS
46 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

6. Assume e%40 so small that for joe%2


2 1
2 C1 zo ; ð7:27Þ
jlog jj 4

%
2 X
k1
1
2 C1 ð2MÞk o : ð7:28Þ
jlog jj k¼0
8M

We now make a definite choice for z (cf. 7.19)


%
z ¼ 8Mð2MÞk r0 þ 1; ð7:29Þ

where k% is defined in 4 and r0 ¼ maxi jjri ð0ÞjjC 3þa ðS1 Þ :


7. From the variation of constants formula applied to Eq. ð7:18Þ2 it follows via
(7.24), (7.26), that
2 2
zi pMzi ð0Þ þ C2 s1b
i zi þ C1 ð1 þ zÞ þ C 1 zi ; ð7:30Þ
jlog jj jlog jj

where we have set

zi ¼ sup jjri ðtðsÞÞjjC 3þa ðS1 Þ ; zi0 ¼ jjri ð0ÞjjC 3þa ðS1 Þ ð7:31Þ
0ospsi

and
CR
C2 ¼ MN ð3 þ jjT0 jjÞ:
RN ð0Þ

Eq. (7.30) is valid in the interval Ii ¼ ½0; si ðT̂0 ÞÞ: From (7.25) and (7.27) it follows that
for si AIi -½0; s we have
2
zi o2Mzi ð0Þ þ 2 C1 ð1 þ zÞ; si AIi -½0; s: ð7:32Þ
jlog jj

If we replace in (7.30) zi ð0Þ and zi ðsÞ and si by ðsi  sÞ and use Eq. (7.34) to estimate
zi ðsÞ we get
2
zi oð2MÞ2 zi0 þ 2 C1 ð1 þ zÞð2M þ 1Þ; si AIi -½0; 2s: ð7:33Þ
jlog jj

By iterating this procedure we get

2 X
k1
zi oð2MÞk zi ð0Þ þ 2 C1 ð1 þ zÞ ð2MÞh ; si AIi -½0; ks: ð7:34Þ
jlog jj h¼0

Eq. (7.34) is one of the basic estimates needed to complete the proof.
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 47

8. We now prove that ŝi AI1 : This follows from Eq. (7.34). In fact for kpk% we have

2 X
k1
ð2MÞk zi ð0Þ þ 2 C1 ð1 þ zÞ ð2MÞh
jlog jj h¼0
% %
% 2 X
k1
2 X
k1
pð2MÞk r0 þ 2 C1 ð2MÞh þ 2 C1 ð2MÞz
jlog jj h¼0
jlog jj h¼0
% 1 z z
pð2MÞk r0 þ þ o ; ð7:35Þ
8M 8M 4M

where we have utilized definition (7.29) of z and (7.34). This inequality, recalling the
definition of k% in 4, shows that Eq. (7.34) implies

z
zi ðsi Þo oz; i ¼ 1; y; N ð7:36Þ
4M

for si A½0; s%i ; where s%1 is defined in 2 and

Z s%i
R31
s%i ¼ ds1 if i41: ð7:37Þ
0 R3i

Since, by definition s%i Xŝi ; the claim is proved.


9. From Eq. (7.36) it follows that the condition

%
jjri ðtÞjjC 3þa ðS1 Þ oz ¼ 8Mð2MÞk r0 þ 1 ð7:38Þ

is satisfied in the time interval ½0; T̂0 Þ: Notice that T̂0 cannot be infinity because, as we
have seen from Proposition 7.2(ii), condition (7.36) implies R1 ðtÞ-0 in finite time.
Thus the maximal interval of existence of the solution of system (7.1) and therefore
T̂0 ¼ T̂: The proof of the proposition is complete. &

Acknowledgments

The authors thank the anonymous referee for his substantial suggestions that
resulted to an improved paper.
N.A. and G.K. were partially supported by a IIENED 99/527 interdisciplinary
grant in Materials. N.A. and G.K. also acknowledge partial support by the RTN
European network: Fronts-Singularities, HPRN-CT-2002-00274. N.A. also grate-
fully acknowledges the support from the office of special accounts at the University
of Athens.
ARTICLE IN PRESS
48 N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49

References

[1] N. Alikakos, P. Bates, X. Chen, The convergence of solutions of the Cahn–Hilliard equation to the
solution of the Hele–Shaw model, Arch. Rational Mech. Anal. 128 (1994) 165–205.
[2] N. Alikakos, P.W. Bates, X. Chen, G. Fusco, Mullins–Sekerka motion of small droplet on a fixed
boundary, J. Geom. Anal. 4 (2000) 575–596.
[3] N.D. Alikakos, G. Fusco, The equations of Ostwald ripening for dilute systems, J. Statist. Phys. 95
(1999) 851–866.
[4] N.D. Alikakos, G. Fusco, The effect of distribution in space in Ostwald ripening, Centre de
Recherches Mathematiques CRM Proceedings and Lecture Notes, Vol. 27, 2001.
[5] N.D. Alikakos, G. Fusco, Ostwald ripening for dilute systems under quasistationary dynamics,
Comm. Math. Phys., 238 (2003) 429–479.
[7] N.D. Alikakos, G. Fusco, G. Karali, Continuum limits of particles interacting via diffusion, Abstract
Appl. Anal., 2004 (3) (2004) 215–237.
[8] H. Amann, Linear and Quasilinear Parabolic. I. Abstract Linear Theory, Monographs in
Mathematics, Vol. 89, Birkhäuser, Basel, 1995.
[9] S.B. Angenent, Nonlinear analytic semigroups, Proc. Roy. Soc. Edinburgh A 115 (1990) 91–107.
[10] G. Baker, O. Meiron, S. Orszag, J. Fluid Mech. 123 (1982) 477.
[11] G. Bellettini, G. Fusco, Some aspects of the dynamics of SðVÞ ¼ H  H; % J. Differential Equations
157 (1999) 206–246.
[12] J.W. Cahn, On spinodal decomposition, Acta Metall. 9 (1961) 795–801.
[13] J.W. Cahn, J.E. Hilliard, Free energy of a nonuniform system I. interfacial free energy, J. Phys.
Chem. 28 (1958) 258–267.
[14] X. Chen, The Hele–Shaw problem and area-preserving curve shortening motion, Arch. Rational
Mech. Anal. 123 (1993) 117–151.
[15] X. Chen, Global asymptotic limit of solutions of the Cahn–Hilliard equation, J. Differential Geom.
44 (2) (1996) 262–311.
[16] X. Chen, X. Hong, F. Yi, Existence, uniqueness and regularity of solutions of the Mullins–Sekerka
problem, Comm. Partial Differential Equations 21 (1996) 1705–1727.
[17] P. Constantin, M. Pugh, Global solutions for small data to the Hele–Shaw problem, Nonlinearity 6
(1993) 393–415.
[18] G. da Prato, P. Grisvard, Equations d’evolution abstraites non lineaires de type parabolique, Ann.
Mat. Pura Appl. 4 (120) (1979) 329–396.
[19] M.P. Do Carmo, Differential Geometry of Curves and Surfaces, Prentice-Hall, New York, 1976.
[21] J. Escher, G. Simonett, Classical solutions for Hele Shaw models with surface tension, Adv.
Differential Equations 2 (1997) 619–642.
[22] J. Escher, G. Simonett, A center manifold analysis for the Mullins–Sekerka model, J. Differential
Equations 143 (1998) 267–292.
[23] G. Folland, Introduction to Partial Differential Equations, Princeton University Press and University
of Tokyo Press, Princeton, NJ, 1976.
[24] D. Gilbarg, N. Trudinger, Elliptic Partial Differential Equations of Second Order, Springer, Berlin.
[25] D. Henry, Geometric Theory of Semilinear Parabolic Equations, Lecture Notes in Mathematics,
Springer, Berlin, 1981.
[26] M. Kimura, An application of the shape derivative to the quasistationary Stefan problem, preprint.
[27] I.M. Lifschitz, V.V. Slyozov, The kinetics of precipitation from supersaturated solid solutions, J.
Phys. Chem. Solids 19 (1961) 35–50.
[28] A. Lunardi, Analytic semigroups and optimal regularity for parabolic problems, Progress of
Nonlinear Differential Equations and its Applications, Vol. 16, Birkhäuser, Basel, 1995.
[29] C. Miranda, Sulle proprieta di regolarita di certe trasformazioni integrali, Atti Accad. Naz. Lincei
Mem. Cl. Sci. Fis. Mat. Natur. Sez. I 8 (7) (1965) 303–336.
[30] W.W. Mullins, Grain growth of uniform boundaries with scaling, Acta Mater. 46 (1998) 6219–6226.
[31] W.W. Mullins, R.F. Sekerka, Morphological stability of a particle growing by diffusion and heat
flow, J. Appl. Phys. 34 (1963) 323–329.
ARTICLE IN PRESS
N.D. Alikakos et al. / J. Differential Equations 205 (2004) 1–49 49

[32] B. Niethammer, Derivation of the LSW—theory for Ostwald ripening by homogenization methods,
Arch. Rational Mech. Anal. 147 (2) (1999) 119–178.
[33] B. Niethammer, F. Otto, Domain coarsening in thin films, Comm. Pure Appl. Math. 54 (3) (2001)
361–384.
[34] W. Ostwald, Z. Phys. Chem. 37 (1901) 385.
[35] R.L. Pego, Front migration in the nonlinear Cahn–Hilliard equation, Proc. Roy. Soc. London Ser. A
422 (1989) 261–278.
[36] G.E. Poirier, W.P. Fitts, J.M. White, Two-dimensional phase diagram of decanethiol on Au(111),
Langmuir 17 (2001) 1176–1183.
[37] H.M. Soner, Convergence of the phase-field equations to the Mullins–Sekerka problem with kinetic
undercooling, Arch. Rational Mech. Anal. 131 (1995) 139–197.
[38] J. Vinals, W.W. Mullins, Self-similarity and coarsening of three dimensional particles on a two
dimensional matrix, J. Appl. Phys. 83 (2) (1998).
[39] P.W. Voorhees, G.B. McFadden, D. Boisvert, D.I. Meiron, Numerical simulation of morphological
development during Ostwald ripening, Acta Metall. 36 (1) (1988) 207–222.
[41] C. Wagner, Theorie der Alterung von Niederschlagen dursch Umlosen, Z. Elektrochem. 65 (1961).
[42] R. Ye, Foliation by constant mean curvature spheres, Pacific J. Math. 147 (2) (1991) 381–396.
[43] J. Zhu, X. Chen, T.Y. Hou, An efficient boundary integral method for the Mullins–Sekerka problem,
J. comput. Phys. 126 (2) (1996) 246–267.

Further reading

N.D. Alikakos, G. Fusco, G. Karali, The effect of the geometry of the particle distribution in Ostwald
ripening, Comm. Math. Phys., 238 (2003) 481–488.
J. Duchon, R. Robert, Evolution d’une interface par capillarite et diffusion de volume I, existence locale en
temps, Ann. Inst. H.Poincaré Anal. Non Linéaire 1 (1984) 361–378.
P.W. Voorhees, Schaffer, In situ observation of particle motion and diffusion interactions during
coarsening, Acta Metall. 33 (1987) 327–339.

You might also like