ArticlePDF Available

Abstract and Figures

Self-organized patterns are sensitive to microscopic external perturbations that modify the diffusion process. We find that Turing instability formed in a compartmented medium, a Belousov–Zhabotinski–aerosol-OT micelle reaction, responds sensitively to a change in the diffusion process. In order to modify the diffusion mechanism, we apply a centrifugal force that generates a perturbation with an anisotropic character. We find experimentally and numerically that the perturbation is able to modify the pattern and even force its disappearance. For different values of the perturbation significant changes can be seen in both the pattern wavelength and its morphology. Furthermore, for strong perturbations, the orientation of the patterns couples with the symmetry of the perturbation.
a shows experimental results in the absence of force. Turing patterns appear as a mixed state that is composed of stripes and spots with a labyrinthine conguration. However, when we increase the force to 4.7 m s À1 (Fig. 2b) two different phenomena can be distinguished: the dominance of stripes over spots and the so orientation of patterns along the direction of the centripetal force. Higher force values intensify these phenomena and stripe-like patterns completely orient along the direction of the force (Fig. 2c). The phase diagram can be divided into three representative regions, as shown in Fig. 2g. The rst region, from uR ¼ 0 m s À1 up to uR ¼ 2 m s À1 , is characterized by patterns that maintain a mixed conguration (labyrinth patterns). The second region corresponds to uR values ranging from 2 m s À1 up to 5 m s À1 (see discussion of Fig. 6). Here, the system exhibits an intermediate state where the orientation phenomenon begins to occur (transition stage). Finally above 5 m s À1 all patterns present are well-oriented stripes in the direction given by the centrifugal force (oriented patterns). Moreover, the pattern wavelength augments slightly according to the centrifugal force as will be discussed below (Fig. 5). It is interesting to note that the organization mechanisms studied here are similar to those studied for sh skin pattern formation by using the CDIMA reaction-diffusion model, 30 where the velocity of a growing boundary determines the nal structure of the pattern.
… 
Content may be subject to copyright.
Turing instability under centrifugal forces
Jacobo Guiu-Souto,
*
a
Lisa Michaels,
b
Alexandra von Kameke,
a
Jorge Carballido-
Landeira
a
and Alberto P. Mu
~
nuzuri
a
Self-organized patterns are sensitive to microscopic external perturbations that modify the diusion
process. We nd that Turing instability formed in a compartmented medium, a BelousovZhabotinski
aerosol-OT micelle reaction, responds sensitively to a change in the diusion process. In order to modify
the diusion mechanism, we apply a centrifugal force that generates a perturbation with an anisotropic
character. We nd experimentally and numerically that the perturbation is able to modify the pattern
and even force its disappearance. For dierent values of the perturbation signicant changes can be
seen in both the pattern wavelength and its morphology. Furthermore, for strong perturbations, the
orientation of the patterns couples with the symmetry of the perturbation.
Introduction
Self-organization is a characteristic of many living systems that
involves three fundamental properties: order, pattern and
form.
1
Thereby the involved physico-chemical mechanisms
allow these systems to evolve from an initially disordered stage
into a pattern-organized conguration with well-dened
macroscopic properties. In the last few decades, this phenom-
enon has been studied from a theoretical point of view
2,3
concluding that many nonlinear systems can lead to macro-
scopic self-organization.
One of the most important mechanisms capable of gener-
ating organization through pattern formation is the Turing
instability in reactiondiusion systems.
4,5
It is thought to be
responsible for morphogenesis
6,7
in living organisms, for
example, patterning on the sh skin
8
or seashells.
9
In Nature,
pattern formation does not occur in isolated systems but usually
in the presence of dierent external perturbations, as for
example, changes in the temperature, density gradients,
1012
and
even certain types of periodical forces such as modulated gravity
or dierential ows.
1315
Furthermore, pattern formation in the
presence of a centrifugal force has been studied in dierent
systems such as RayleighB
´
enard cells or active media.
1618
Here we considered the BelousovZhabotinsky reaction
19,20
encapsulated into AOT micelles
19
(BZAOT system). Experi-
mentally, this system is shown to exhibit Turing patterns for the
appropriate concentrations of the reactants. We impose a
centrifugal force in such a way that any uid ow is suppressed,
and the only transport process is the microscopic diusion of
micelles and reactants due to the physical constraints of the
system. The experimental observations are explained by theo-
retical analysis. Furthermore, numerical simulations conrm
and complement the experiments.
Methods
Experimental description
Two stock microemulsions (MEs), ME
1
and ME
2
, with the same
molar ratio, w ¼ [H
2
O]/[AOT] (¼18), and the same droplet
volume fraction, F
d
(¼0.72), were prepared at room temperature
by mixing aqueous solutions of BelosovZhabotinsky (BZ) reac-
tants and an oily 1.5 M solution of aerosol-OT (ref. 21) in octane.
ME
1
was composed of solutions of malonic acid (0.5 M) and
sulphuric acid (0.3 M), while for ME
2
we used bromate (0.32 M)
and ferroin (8.3 mM) as described in ref. 22. The reactive
microemulsion, BZAOT reaction system,
2325
was obtained by
mixing equal volumes of the two MEs and diluting the mixture
with octane to the desired droplet fraction (F
d
¼ 0.48).
The behaviour of the reaction is governed by the two main
species, also known as the activator (bromous acid) and the
inhibitor (bromine) due to their chemical competition
processes.
23
The activator molecules remain on the inside of the
AOT-micelles in the aqueous phase, due to their polar nature
and ionic dissociation, and only the non-polar molecules
(inhibitor) can go through the micelle membrane and move into
the oil phase. The polar compound, the activator, is conned to
the aqueous phase within the micelle and its diusion is linked
to that of the micelle. Thus, in the following, we will consider
the activator mass as that of the micelle containing it. The
micelle mass can be estimated
22
given the hydrodynamic radius
of our micelles
23
(R 1 nm) and we obtain a value of 2500 amu.
As a conclusion, the activator diusion coecient is much
smaller than the inhibitor diusion coecient (approximately,
10
9
m
2
s
1
and 10
7
m
2
s
1
, respectively
23
).
a
Group of Nonlinear Physics, Universidade de Santiago de Compostela, E-15782
Santiago de Compostela, Spain. E-mail: jacobo.guiu@usc.es
b
Max-Planck-Institut fuer Kernphysik, Saupfercheckweg 1, 69117 Heidelberg, Germany
Present address: Planetary Sciences, Harvard University, 100 Edwin H. Land
Boulevard Cambridge, MA 02142-1204
Cite this: DOI: 10.1039/c3sm27624d
Received 14th November 2012
Accepted 22nd February 2013
DOI: 10.1039/c3sm27624d
www.rsc.org/softmatter
This journal is ª The Royal Society of Chemistry 2013 Soft Matter
Soft Matter
PAPER
Downloaded by Harvard University on 22 March 2013
Published on 21 March 2013 on http://pubs.rsc.org | doi:10.1039/C3SM27624D
View Article Online
View Journal
A small amount of the reactive ME was sandwiched between
two optical glasses separated by a Teon gasket (thickness
80 mm, inner diameter 25 mm, and outer diameter 50 mm, see
Fig. 1). This reactor was placed at dierent distances, R, from
0 up to 20 cm from the rotation center of the rotor (Geared
Brushless 24V DC), performing an uniform circular motion on
the horizontal plane, with a maximum peak angular velocity of
u ¼ 40 rad s
1
. The system is such that no convective ow may
occur and this will be discussed below.
The rotating BZAOT system was analyzed with the following
acquisition image set. In order to achieve optimal contrast, the
sample was illuminated with a blue high-intensity diode array
LED. The light passes through a series of light diusers until
reaching the reactor in order to guarantee uniform background
illumination. The image of the reactor was magnied with an
achromatic objective (DIN 4 Edmund Optics), recorded by a
camera (Guppy AVT 64 fps) and was nally processed in the
computer. In order to quantify the eect of the rotational force,
we compare the dierences between two identical reaction
samples. One of them is subjected to a rotational motion and
the other is le at rest, it is subsequently referred to as the
standard pattern. This procedure is repeated for dierent
angular velocities and radii, obtaining a range of force from
uR ¼ 0ms
1
to uR ¼ 8ms
1
. The required time for the pattern
to appear and its evolution into the steady state is 40 min. The
room temperature was kept constant at 20 2
C. All experi-
ments lying on the same iso-velocity line (v ¼ uR) exhibit the
same behaviour and thus can be averaged for statistical
purposes.
BZAOT numerical model
The kinetics of the BZAOT reaction can be modelled by the
Oregonator equations revised in ref. 26:
F
1
¼
dc
1
dt
¼
1
3
fc
2
q c
1
q þ c
1
þ c
1
1 mc
2
1 mc
2
þ 3
1
c
1
2
(1)
F
2
¼
dc
2
dt
¼ c
1
1 mc
2
1 mc
2
þ 3
1
c
1
(2)
where the subindex i ¼ 1, 2 labels the activator and inhibitor
species, and c
i
are their respective dimensionless concentra-
tions. The parameters f, m, q, 3
1
and 3 depend on the reaction
rates and they appear as a result of applying the mass action law
on the chemical equations.
27
The suitable values to model the
BZAOT system
28
are f ¼ 1.2, m ¼ 190, q ¼ 0.001, 3
1
¼ 0.01 and
3 ¼ 0.8, as used throughout this study. Numerical simulations
of eqn (7) are performed by applying an explicit three-level Du
FortFrankel scheme
29
with a spatial step of 0.2 s.u. and zero
ux boundary conditions. To trigger instability we use random
noise as an initial condition.
Experimental results
Fig. 2a shows exper imental results in the absence of force.
Turing patterns app ear as a mixed state that is composed of
stripes and spots with a labyrinthine congurat ion. However,
when we increase the force to 4.7 m s
1
(Fig. 2b) two dierent
phenomena can be distin guished: the dominance of stripes
over spots and the so orientation of patterns along the
direction of the centripetal force. Higher force values intensify
these phenomena an d stripe-lik e patterns completely orient
along the direction of the for ce (Fig. 2c). The phase diagram
can be divided into three represe ntative regions, as shown in
Fig. 2g.
The rst region, from uR ¼ 0ms
1
up to uR ¼ 2ms
1
,is
characterized by patterns that maintain a mixed conguration
(labyrinth patterns). The second region corresponds to uR
values ranging from 2 m s
1
up to 5 m s
1
(see discussion of
Fig. 6). Here, the system exhibits an intermediate state where
the orientation phenomenon begins to occur (transition stage).
Finally above 5 m s
1
all patterns present are well-oriented
stripes in the direction given by the centrifugal force (oriented
patterns).
Moreover, the pattern wavelength augments slightly
according to the centrifugal force as will be discussed below
(Fig. 5). It is interesting to note that the organization mecha-
nisms studied here are similar to those studied for sh skin
pattern formation by using the CDIMA reactiondiusion
model,
30
where the velocity of a growing boundary determines
the nal structure of the pattern.
Theoretical approach
In order to understand the phenomena induced by a centrifugal
force, we propose a theoretical model based on a statistical
mechanics approach. The characteristics of our encapsulated
reactor allow us to neglect any macroscopic ow,
22
and there-
fore, the only transport mechanism is due to molecular diu-
sion.
31,32
This approximation is possible because the thickness
(80 mm) of our system is much smaller than the estimated
boundary layer thickness
33
(3 cm). Given that no convective
transport may exist in this system, the only eect of the external
force is an alteration of the diusion mechanism. According to
Fig. 1 Experimental setup. The rotor is controlled by a voltage source (VS) that
generates a uniform angular velocity u. BL is the blue LED array used for illumi-
nation and DF is the diuser. The CCD camera on top of the reactor center is
connected to a PC. The reactor (RE), where the BZAOT reaction takes place, is
composed of a Teon separator (TS) and two optical glasses (GL). R is the distance
from the rotation axis to the reactor center.
Soft Matter This journal is ª The Royal Society of Chemistry 2013
Soft Matter Paper
Downloaded by Harvard University on 22 March 2013
Published on 21 March 2013 on http://pubs.rsc.org | doi:10.1039/C3SM27624D
View Article Online
this, canonical ensemble considerations can be applied in order
to evaluate the inuence of the force on the microscopic diu-
sion processes.
34
We establish an expression for the position of the activator
and the inhibitor inside of the reactor by coupling the thermal
uctuations with the rotational motion:
r
i
¼ x
i
^
x + y
i
ˆ
y + z
i
^
z ¼ (x
0
i
+ x
rot
)
^
x +(y
0
i
+ y
rot
)
ˆ
y + z
0
i
^
z (3)
where x
rot
¼ Rcos(ut), y
rot
¼ Rsin(ut) and the subindex i ¼ 1, 2
represents the activator and inhibitor species, respectively. The
coordinates (x
0
i
, y
0
i
, z
0
i
) indicate the position of the particles due to
thermal motion. Thus the resulting Hamiltonian associated
with the kinetic contribution is:
H
i
¼
p
0
i
2
2m
i
þ
1
2
m
i
u
2
R
2
þ uðp
0
y;i
x
rot
p
0
x;i
y
rot
Þ (4)
where p
0
i
¼ (p
0
x,i
, p
0
y,i
, p
0
z,i
) and m
i
denote the momentum and the
masses of the particles. The rst and the second term of the
Hamiltonian represent the kinetic and rotational energy,
respectively. The last term represents the angular moment of
the particles and it accounts for the coupling between the
rotational motion and thermal uctuations.
In order to estimate the changes in the diusion processes
due to the applied force, we consider a diluted system approx-
imation
34
and assume that the microscopic collisions occur only
between identical particles, i.e. 11or22. The 12 collisions are
considered negligible because in the BZAOT reaction the
inhibitor molecules can diuse through the micelle membrane
due their non-polar character.
23
One can also estimate easily the
frequency of collisions and the results indicate that collisions of
type 22 occur 100 times more frequently that those of type 11.
All of these arguments allow us to consider each species
independently.
Under these considerations we obtain the following expres-
sion for the diusion coecient of each species:
31,34
D
1;2
ðb; m
1;2
; u; RÞ¼
D
0
1;2
ðb; m
1;2
Þ
3
ffiffi
3
p
3 þ bm
1;2
v
2
3=2
(5)
D
0
1;2
ðb; m
1;2
Þ¼
1
8Ns
tot
ffiffiffiffiffiffiffiffiffiffi
3
bm
1;2
s
(6)
where v ¼ uR is the characteristic force velocity, N is the volu-
metric density of the particles, s
tot
is the total collision cross-
section, D
0
1,2
(b,m
1,2
) is the diusion coecient in the absence of
force and b ¼ 1/k
B
T. This analytical expression, eqn (5), was
obtained by considering that bm
1,2
u
2
R
2
1 which is in agree-
ment with the experimental conditions. For example, for the
activator bm
1
u
2
R
2
0.095 at the experimental force threshold.
It is also important to note that in the absence of force, i.e.
uR ¼ 0ms
1
, we recover the expressions for free diusion.
34
In accordance with eqn (5) the diusion process is aniso-
tropic, since it depends on the direction and magnitude of the
centrifugal force. Moreover, as x
rot
, y
rot
[ x
0
, y
0
, we consider the
following approximations x x
rot
and y y
rot
. Taking into
account all these considerations the resulting reactiondiu-
sion equation for our rotating system is:
vc
i
vt
¼ F
i
þ 2D
0
i
bm
i
u
2
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ
1
3
F
i
ðx; yÞ
r
x
vc
i
vx
þ y
vc
i
vy
þ
D
0
i
3
ffiffi
3
p
ð3 þ F
i
ðx; yÞÞ
3=2
v
2
c
i
vx
2
þ
v
2
c
i
vy
2
(7)
where c
i
with i ¼ 1, 2 represents the activator and inhibitor
species, respectively. F
i
(x,y) ¼ bm
i
u
2
(x
2
+ y
2
) is related to the
magnitude of the centrifugal forces. F
i
is the reaction term from
eqn (1) and (2). The second term on the right corresponds to a
microscopic ow obtained from VD
i
Vc
i
. Finally, the last term
represents the diusive transport given by D
i
V
2
c
i
.
By performing a linear stability analysis
7
of eqn (7), we nd
the following eigenvalues L
:
Fig. 2 Turing pattern under centrifugal force (direction represented by the red arrow) far from the rotation center. The patterns orient in the direction of the force.
Experimental (top row) and numerical (bottom row) Turing patterns obtained at dierent force values: (a and d) uR ¼ 0ms
1
, (b and e) uR ¼ 4.7 m s
1
and (c and f)
uR ¼ 7ms
1
. The experimental images have a radius of 1.53 mm and the computational domain consists of a two-dimensional mesh of 300 300 grid points, with a
spatial step of 0.2 s.u.; (g) is the experimental (u, R)-phase diagram with the corresponding characteristic velocities uR. The iso-velocity curves uR ¼ 2ms
1
and uR ¼
5ms
1
divide the diagram into three regions; labyrinth patterns, transition stage and oriented patterns.
This journal is ª The Royal Society of Chemistry 2013 Soft Matter
Paper Soft Matter
Downloaded by Harvard University on 22 March 2013
Published on 21 March 2013 on http://pubs.rsc.org | doi:10.1039/C3SM27624D
View Article Online
2L
¼
k
2
trðDÞjktrðZ ÞtrðJÞ
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
½k
2
trðDÞjktrðZÞtrðJÞ
2
4Pðk; D; J; ZÞ
q
(8)
with,
P(k,D,J,Z) ¼ k
4
det(D) jk
3
tr(D*Z) k
2
[tr(D*J) + det(Z)]
+ jktr(JZ*) + det(J) (9)
where k is the wavenumber, j the imaginary unit and J the
Jacobian matrix corresponding to the linearization of the reac-
tion terms, F
i
. D is the diagonal diusion matrix (d
ii
¼ D
i
), and Z
is an auxiliary diagonal matrix (z
ii
¼ 3D
0
i
bm
i
u(x + y)) that
accounts for the anisotropy of the diusion processes. The
matrices D* and Z* are dened by D
1
det(D) and Z
1
det(Z),
respectively. Once again, in the absence of force we recover the
standard expressions for the reactiondiusion system.
35
Fig. 3 shows the real eigenvalues, Re(L
+
) versus the wave-
number for dierent forces. Positive values of Re(L
+
) are
responsible for generating the Turing instability. Negative
values of the eigenvalue indicate the region where the Turing
structures are unstable. Thus for each uR-isovelocity-curve of
Fig. 3 there are a range of wavenumbers denoted by (k
min
, k
max
),
where Turing patterns can be generated. Moreover, the most
unstable wavenumber
7
(k
C
) inside that range determines the
nal observed stable conguration of the pattern, i.e. its critical
wavelength given by the relationship l
C
¼ 2p/k
C
. As the force is
increased k
max
and k
C
decrease and k
min
remains almost
invariable. It points to the fact that the pattern wavelength, l
C
,
increases with the force. Note that for values of the force above
uR ¼ 30 m s
1
Turing instability is inhibited.
Numerical results
Eqn (7) was numerically integrated and the results are shown in
Fig. 2df. Note that they show a good agreement with the
experimental observations; the radial force gradually orients the
previously labyrinthine patterns in the direction of the centrif-
ugal force. Additionally, we again obtain a structure that evolves
from a mixed state composed of stripes and spots to a well-
dened stripes con guration. Analogous to the experimental
results, the numerical wavelength undergoes an increment that
will be discussed in Fig. 5.
Fig. 4 shows the eect of a strong force close to the rotation
center. In the absence of any force (Fig. 4a) we obtain a laby-
rinth-like pattern with a uniform wavelength and without a
Fig. 3 The real part of the eigenvalues obtained by a linear stability analysis of
eqn (7). The curves were calculated for the following values of force uR ¼ 0, 5, 15,
25 and 40 m s
1
. The most probable Turing wavenumber k
C
, is the maximum of
each Re(L
+
) curve and the Turing region (k
min
, k
max
) is given by the zero-crossing,
i.e. Re(L
+
(k
min,max
)) ¼0. For forces above uR ¼30 m s
1
Turing patterns disappear.
Fig. 4 Numerical study of Turing patterns under rotation. The center of rotation lies in the middle of the numerical domain. The radius scale is around the sizeof
patterns (R ranges from 0 to 2 mm). (a) Labyrinth pattern in the absence of force. (b) Labyrinth pattern and transition stage at u ¼ 6 10
3
rad s
1
. (c) Developed radial
modications (oriented stage and black spots), at u ¼ 35 10
3
rad s
1
. (d) Pattern inhibition at u ¼ 57 10
3
rad s
1
. A summary of all these results is plotted in the
(u, R)-phase diagram in (e). The iso-velocity curves uR ¼ 2ms
1
, uR ¼ 5ms
1
, uR ¼ 20 m s
1
and uR ¼ 30 m s
1
delimit ve regions: I, II, III, IV and V related to the
eects of force over the patterns. Lines (a) to (d) indicate the force situations of panel (a) to (d) on the left.
Soft Matter This journal is ª The Royal Society of Chemistry 2013
Soft Matter Paper
Downloaded by Harvard University on 22 March 2013
Published on 21 March 2013 on http://pubs.rsc.org | doi:10.1039/C3SM27624D
View Article Online
preferred direction. Similar to the experimental results, when
we increase the angular velocity up to u ¼ 6 10
3
rad s
1
, the
pattern shows an increment in the wavelength with the radius,
and also a growing dominance of stripes (Fig. 4b). At higher
angular velocities, the patterns re-align along the direction of
the force (see Fig. 4c and d) but also new phenomena are
observed. On the one hand, black spots appear at the boundary
(see for example Fig. 4c), and on the other hand, for values
higher than uR ¼ 30 m s
1
, Turing patterns are strongly
inhibited (see Fig. 4d).
These dierent states are summarized in the numerical
phase diagram shown in Fig. 4e. The iso-velocity curves (uR ¼
2ms
1
, uR ¼5ms
1
, uR ¼20 m s
1
and uR ¼30 m s
1
) divide
the phase diagram into ve regions denoted as I, II, III, IV and V.
Region I includes the labyrinth patterns and region II the
transition state shown in Fig. 2g. Region III corresponds to the
experimental oriented regime. The black spots appear in region
IV and the inhibition phenomenon occurs in region V. When
this last region is reached the conditions for the Turing insta-
bility are not satised
7
as we also observed from the analytical
considerations. Lines marked with (a) to (d) in Fig. 4e show the
values of R and u in the corresponding Fig. 4ad. Each one of
these lines determines the transitions along the dierent
regions (related to the type of pattern organization) in the phase
diagram as seen in the corresponding Fig. 4ad.
Discussion
Fig. 5 compares the dierences in the wavelength of the Turing
pattern for the experimental results, the numerical results and
the theoretical predictions from Fig. 2 and 3. This comparison
is feasible due to the dimensionless force term (bm
1,2
u
2
R
2
) and
the use of the normalized wavelength increment, Dl. This
quantity is dened as Dl ¼ (l l
0
)/l
0
where l
0
stands for the
wavelength of the pattern in the absence of external force, i.e.,
the standard pattern. Note that the normalized wavelength
Fig. 5 Normalized wavelength (Dl ¼ (l l
0
)/l
0
, with l
0
as the standard pattern
wavelength) versus the force applied. The theoretical critical wavelength incre-
ment, Dl
C
(red continuous line) from stability analysis, numerical estimations,
Dl
num
(blue circles), and experimental measures, Dl
exp
(black squares)reect a
monotonous growth with the force. The theoretical region for Turing patterns lies
between the maximum and the minimum wavelength increments, i.e. Dl
max
and
Dl
min
(dashed lines). The uncertainties in the Dl were estimat ed from the width
of the wavenumber ring obtained by a Fourier transform of the Turing patterns
images. Uncertainties associated with uR are estimated by the sensitivity of the
experimental devices.
Fig. 6 Representation of the L and T-length mode increment depending on the force. The increments Dl
L,T
are calculated with respect to the wavelength of the
standard pattern (l
0
), analogous to Fig. 5. Circles (squares) and red (blue) color stand for transversal (longitudinal) mode. Filled (empty) markers correspond to
experimental (numerical) results. (a) Characterization of dierent orientation regimes versus applied force. Point P
1
marks the rst divergence between the L-mode and
T-mode and corresponds to the transition state (transition between region I and II). This can be more clearly seen in the inset of the gure. Point P
2
marks the transition
between regions II and III. (b) Semilog plot of the same results (log(Dl
L,T
) ¼ a
L,T
uR with a
L
¼ 1.02 0.16 and a
T
¼ 0.42 0.09).
This journal is ª The Royal Society of Chemistry 2013 Soft Matter
Paper Soft Matter
Downloaded by Harvard University on 22 March 2013
Published on 21 March 2013 on http://pubs.rsc.org | doi:10.1039/C3SM27624D
View Article Online
augments with the force. Experimentally, increments of up to
18% were reported while theoretically and numerically any
increment can be achieved up to the inhibition of the Turing
pattern (Fig. 3). In this sense, the experimentally encapsulated
BZAOT reactions are quite robust to a rotational force, and
therefore Turing patterns are only moderately modied. Similar
eects on the Turing patterns were observed in systems where
temperature or concentrations modify the diusivity.
36,37
In addition to the changes in the wavelength, we quantify the
orientation of the pattern with respect to the centrifugal forces.
From the numerical and experimental data shown in Fig. 2 we
calculate the mean separation between stripes in the direction
of the force (Dl
L
), and in the transversal direction (Dl
T
), deno-
ted as the longitudinal (L) and transversal (T) wavelength
modes, respectively (the denition of these normalized quan-
tities is equivalent to that in Fig. 5). The T-modes are related to
pattern wavelength and L-modes take into account the pattern
orientation. According to this, patterns in the absence of force
(Fig. 2a and d) have similar values for both L and T-modes since
there is no preferential direction. The relative wavelength
increments, Dl
L,T
, for the two modes are shown in Fig. 6. Note
that both experimental data and numerically data t reasonably
well onto the same curve.
Fig. 6a shows that above uR ¼5ms
1
(P
2
in the gure) the L-
modes experience an important increment because stripes
become largely oriented. At the bifurcation point P
1
(uR ¼ 2m
s
1
), the L and T-modes rst separate from each other. Thus,
the interval of force between P
1
and P
2
(from 2 to 5 m s
1
)
denes a transition zone where the orientation phenomenon
begins to evolve. P
1
and P
2
were calculated by intersecting the
best correlated linear ts of both modes (see Fig. 6). Note that
the points P
1
and P
2
coincide with the transition from stage I to
II and from II to III, respectively (see Fig. 2 and 4).
The growth of the L and T-modes can be characterized by a
semilog function (log( Dl
L,T
) ¼ a
L,T
uR) (see Fig. 6b). We nd that
the growth of Dl
L,T
follows exponential laws whose exponents
for the L and T modes are a
L
¼ 1.02 0.16 and a
T
¼ 0.42 0.09.
According to this, the longitudinal modes have a rate of growth
twice as high as the translational mode. In order to maximize
the correlation coecient of the t, low force points were
neglected due to their logarithmic dispersion.
Conclusions
We studied the eect of an external rotational force on the
Turing instability. The force modies the diusion dynamics
at the microscopic scale due to coupling of the pattern forma-
tion and the centrifugal force. An estimation of the external
force states that it is about a 10% of the thermal energy
(m
1
u
2
R
2
0.1 k
B
T).
Nevertheless, their eects have a crucial role in the pattern
formation mechanism. This may seem counterintuitive at rst
sight. Nevertheless, we have to consider that this is a highly
non-linear phenomenon where small disturbances may be
reected in a dramatic change in the macroscopic properties of
the system. In our present case, this small perturbation destroys
the isotropy of the system that it is translated into a non-
isotropic diusion coecient that induces the reported
changes. A theoretical model for the anisotropic diusion
process was developed, neglecting convective ows due to
boundary constraints. We show, both experimentally and
numerically, that the anisotropic (inhomogeneous) force has
various important eects on the pattern organization. The force
changes the pattern from a labyrinth con guration to stripes
oriented in the direction of the centrifugal forces. Moreover, a
numerical study reveals that for increasing angular velocity, the
patterns exhibit a clear radial orientation before passing
through an intermediate black-spots conguration and, nally,
are inhibited. In order to clarify these transitions, phase
diagrams for the experimental and the numerical case were
calculated as shown in Fig. 2e and 4e, respectively. A posterior
stability analysis of the reactiondiusion equations indicates
that the force is also able to augment the pattern wavelength
and completely inhibit the formation of Turing structures. We
also observed an increment in the wavelength with the force
experimentally and numerically, see Fig. 5. The numerical and
the experimental results are in good agreement. Furthermore,
the eect of the centrifugal force on the directionality of the
patterns has been quantied by comparing the transversal and
longitudinal wavelength modes (see Fig. 6).
Previous work
22
showed that it was possible to modify the
diusive regime homogenously; here we present a mechanism
capable of introducing anisotropies in the system. The coupling
between the perturbations and the pattern development allows
us to control the type of nal structure. In this regard, the
authors want to remark that the present work contributes to
obtaining a better understanding of the eects that microscopic
ows produce on Turing pattern formation. These mechanisms
might be extended to other systems that involve self-organiza-
tion phenomena. Further, this study also constitutes a
straightforward method to introduce anisotropy in a system by
simply applying an external eld without actually altering the
nature of the system.
Acknowledgements
The authors thank Dr. P. Taboada for helpful discussions. This
work was supported by the Ministerio de Educaci
´
on y Ciencia
and Xunta de Galicia under Research Grants no. FIS2010-21023
and no. CN2012/315. J.C.-L. and A.v.K. were supported by the
MICINN under a FPI and FPU (AP-2009-0713), respectively and
J.G.-S. by the Xunta de Galicia under a Predoctoral Grant.
Notes and references
1 J. Tabony, Biol. Cell, 2006, 98, 589.
2 N. Rashevsky, Bull. Math. Biophys., 1940, 2, 15.
3 I. Prigogine and G. Nicolis, Q. Rev. Biophys., 1971, 4, 107.
4 A. M. Turing, Philos. Trans. R. Soc., B, 1952, 237, 37.
5 S. Setayeshgar and M. C. Cross, Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top., 1998, 58(4), 4485.
6 J. M. W. Slack, From Egg to Embryo. Determinative Events in
Early Development, Cambridge University Press, 1983.
Soft Matter This journal is ª The Royal Society of Chemistry 2013
Soft Matter Paper
Downloaded by Harvard University on 22 March 2013
Published on 21 March 2013 on http://pubs.rsc.org | doi:10.1039/C3SM27624D
View Article Online
7 J. D. Murray, Mathematical Biology, Springer, Berling, 2nd
edn, 1993; P. A. Lawrence, The Making of a Fly: The Genetics
of Animal Design, Blackwell Scientic Publications, Boston,
1992.
8 S. Kondo and R. Asai, Nature, 1995, 376, 765.
9 H. Meinhardt, The Algorithmic Beauty of Sea Shells, Springer,
Berlin, 1995.
10 D. A. Vasquez and C. Lengacher, Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top., 1998, 58, 6865.
11 I. P. Nagy and J. A. Pojman, J. Phys. Chem., 1993, 97, 3443.
12 J. Carballido-Landeira, V. K. Vanag and I. R. Epstein, Phys.
Chem. Chem. Phys., 2010, 12, 3656.
13 S. R
¨
udiger, D. G. M
´
ıguez, A. P. Mu
~
nuzuri, F. Sagu
´
es and
J. Casademunt, Phys. Rev. Lett., 2003, 90, 128301.
14 G. Fern
´
andez-Garc
´
ıa, D. I. Roncaglia, V. P
´
erez-Villar,
A. P. Mu
~
nuzuri and V. P
´
erez Mu
~
nuzuri, Phys. Rev. E: Stat.
Phys., Plasmas, Fluids, Relat. Interdiscip. Top., 2008, 77,
026204.
15 A. von Kameke, F. Huhn, G. Fern
´
andez-Garc
´
ıa,
A. P. Mu
~
nuzuri and V. P
´
erez-Mu
~
nuzuri, Phys. Rev. E: Stat.
Phys., Plasmas, Fluids, Relat. Interdiscip. Top., 2010, 81,
066211.
16 V. P
´
erez-Villar, J. L. F. Porteiro and A. P. Mu
~
nuzuri, Phys. Rev.
E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top., 2006,
74, 046203.
17 M. Fantz, R. Friedrich, M. Bestehorn and H. Haken, Phys. D,
1992, 61, 147.
18 A. P. J. Breu, C. A. Kruelle and I. Rehberg, Europhys. Lett.,
2003, 62, 491.
19 V. K. Vanag and D. V. Boulanov, J. Phys. Chem., 1994, 98,
1449.
20 A. N. Zaikin and A. M. Zhabotinsky, Nature, 1970, 225, 535.
21 T. K. De and A. Maitra, Adv. Colloid Interface Sci., 1995,
59, 95.
22 J. Guiu-Souto, J. Carballido-Landeira, V. P
´
erez-Villar and
A. P. Mu
~
nuzuri, Phys. Rev. E: Stat. Phys., Plasmas, Fluids,
Relat. Interdiscip. Top., 2010, 82, 066209.
23 V. K. Vanag, Phys.-Usp., 2004, 47, 923; V. K. Vanag and
I. R. Epstein, Phys. Rev. Lett., 2001, 87, 228301.
24 V. K. Vanag and I. R. Epstein, Phys. Rev. Lett., 2003, 90,
098301.
25 I. R. Epstein and V. K. Vanag, Chaos, 2005, 15, 047510.
26 A. Kaminaga, V. K. Vanag and I. R. Epstein, J. Chem. Phys.,
2005, 122, 174706.
27 L. Gyorgyi, T. Turanyi and R. J. Field, J. Phys. Chem., 1990, 94,
7162.
28 A. Kaminaga, V. K. Vanag and I. R. Epstein, Angew. Chem.,
Int. Ed., 2006, 118, 3159.
29 F. A. William, Numerical Methods for Partial Dierential
Equations, Academic Press, New York, 1977.
30 D. G. M
´
ıguez and A. P. Mu
~
nuzuri, Biophys. Chem., 2006,
124, 161.
31 J. Guiu-Souto, J. Carballido-Landeira and A. P. Mu
~
nuzuri,
J. Phys.: Conf. Ser., 2011, 296, 012016.
32 J. Guiu-Souto, J. Carballido-Landeira and A. P. Mu
~
nuzuri,
Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip.
Top., 2012, 85, 056205.
33 I. M. Cohen and P. K. Kundu, Fluid Mechanics, Academic
Press, San Diego, 1990.
34 K. Huang, Statistical Mechanics, John Wiley & Sons,
Massachusetts, 1987.
35 S. H. Strogatz, Nonlinear Dynamics and Chaos: With
Applications to Physics, Biology, Chemistry, and Engineering,
Westview Press, 1994.
36 O. Ouyang, R. Li, G. Li and H. L. Swinney, J. Chem. Phys.,
1995, 102, 2551.
37 R. McIlwaine, V. K. Vanag and I. R. Epstein, Phys. Chem.
Chem. Phys., 2009, 11, 1581.
This journal is ª The Royal Society of Chemistry 2013 Soft Matter
Paper Soft Matter
Downloaded by Harvard University on 22 March 2013
Published on 21 March 2013 on http://pubs.rsc.org | doi:10.1039/C3SM27624D
View Article Online
... A previous work [32] has already shown, both numerically and experimentally, that Turing patterns in BZ-AOT medium under the effect of a constant centrifugal force may experience important changes in the diffusion mechanisms. In particular, the intensity of the external forcing directly affects the diffusion coefficients of both species, the effect being modulated by the molecular weight of each species. ...
... The effect of a centrifugal force on this system has been previously modeled [32,36] by introducing a radial-dependent diffusion coefficient as follows: ...
... Note that the effect of the forcing on the diffusion coefficients is multiplied by the molecular weight [Eq. (4)]; thus for large forcings the heaviest species is expected to diffuse faster than the lightest and the Turing mechanism conditions might be violated [32]. As the diffusion coefficient depends explicitly on R, this mathematically introduces some convective terms in the equations that are going to be responsible for some of the dynamics described below. ...
Article
Full-text available
Constantly acting centrifugal forces on Turing pattern forming systems have been observed to induce orientation and wavelength changes on Turing structures. Here, we will consider a periodic modulation of such centrifugal forces and their effects on pattern formation. Depending on the oscillation period the system exhibits a wide variety of stationary (stripes, H0, etc.) or nonstationary patterns (black eyes, etc.), as well as transitions and instabilities such as Eckhaus, zigzag, etc. In this paper, a detailed description of the different patterns and patterning mechanisms will be described and understood within the previous context. The system considered is the Belousov-Zhabotinsky reaction encapsulated in AOT micelles modeled by the adapted version of the Oregonator model.
... Physical and chemical mechanisms enable these living systems to evolve from an initially disordered stage to a patternorganized configuration with well-defined macroscopic properties [1]. As it is known, in animals, drawings patterns on the hide are strictly individual and the configuration of the latter depends on self-organization processes at the stage of embryo development [22][23][24]. As believed by a famous English scientist A. Turing, biological systems are initially characterized by the chemical mechanism of proceeding reactions [1]. ...
... However, Turing's pioneering ideas have found applications lately to explain many phenomena in physics, biology, chemistry, engineering, and even social sciences [25]. For instance, they are used to analyse physical and chemical processes in reaction-diffusion systems modeling the growth of metals during electrodeposition (as exemplified by Au-Cu, Au-Cu-Cd [26], and Co-In Co-In alloys [27]); to simulate processes that are the basis of structure formation on anisotropic and heterogeneous catalytic surfaces [28], in emulsions; and to describe oscillatory reactions in solutions of inorganic and organic substances including surfactants [23,29,30]. Nevertheless, we found no information regarding the use of the theory of morphogenesis to describe formation mechanisms of Turing's patterns in sol-gel technology-derived systems. ...
Article
Full-text available
The paper briefly reports on a contemporary view of the fractal structure of tetraethoxysilane (TEOS)- derived sol–gel synthesis products including in the presence of inorganic dopants and organic modifiers (H3PO4, polyionene, i.e., a high molecular weight quaternary ammonium compound, and epoxy resin), relying on the author’s and literature data. Morphological features of thin film surfaces and the effect of additives of inorganic salts and acids (Co(NO3)2, Al(NO3)3, H3BO3, HNO3, and HCl) thereon were analysed. The influence of spin-coating conditions (temperature and relative humidity) was also investigated, as exemplified by borosilicate “spin-on glass” films. The major components of film-forming TEOS-derived sols (TEOS and H3BO3, TEOS and epoxy-resin) were considered as morphogens. Moreover, their contribution to the formation of 2D and 3D stable spatial structures that were close to Turing patterns according to their configuration was assayed. An opportunity of formation of triply periodic minimal surfaces in thin-layer structures in the result of self-organization of viscous epoxy-siloxane sols has been discussed. Self-organization phenomena that are inherent in sol–gel systems based on tetraethoxysilane (TEOS) including those modified with inorganic dopants and organic modifiers are illustrated in TEM-, optical-, and AFM-images of the thin films surface. Two main types of fractal organization (mass and surface fractals) of silicophosphate sols both unmodified and modified with an organic modifier (polyionene) are exemplified in TEM-images. The possibility of the formation of stable spatial structures similar in surface morphology to Turing patterns, including those of the biological objects (animal prints), is illustrated by optical images of the surface morphology of thin epoxy-siloxane films, both unmodified and modified with the additive of detonation nanodiamonds. The possibility of the formation of sol–gel composites close in topology to triply periodic minimal surfaces is illustrated in the AFM-image of an epoxy-siloxane film in comparison with the Spit-P surface model. Self-organization phenomena that are inherent in sol–gel systems based on tetraethoxysilane (TEOS) including those modified with inorganic dopants and organic modifiers are illustrated in TEM-, optical-, and AFM-images of the thin films surface. Two main types of fractal organization (mass and surface fractals) of silicophosphate sols both unmodified and modified with an organic modifier (polyionene) are exemplified in TEM-images. The possibility of the formation of stable spatial structures similar in surface morphology to Turing patterns, including those of the biological objects (animal prints), is illustrated by optical images of the surface morphology of thin epoxy-siloxane films, both unmodified and modified with the additive of detonation nanodiamonds. The possibility of the formation of sol–gel composites close in topology to triply periodic minimal surfaces is illustrated in the AFM-image of an epoxy-siloxane film in comparison with the Spit-P surface model. The TEOS- and dopants-derived sols are usually multilevel fractal objects of mass or surface type.Stable spatial structures of various configurations can form in thin TEOS-derived thin films.Turing patterns can form in thin films obtained by spin-coating from TEOS-derived sols including doped ones.Precursors of sol–gel synthesis (i.e., TEOS–H3BO3, TEOS–epoxy resin) can be considered as morphogens.Viscous epoxy-siloxane sols can form films with a structure close to triply periodic minimal surfaces. The TEOS- and dopants-derived sols are usually multilevel fractal objects of mass or surface type. Stable spatial structures of various configurations can form in thin TEOS-derived thin films. Turing patterns can form in thin films obtained by spin-coating from TEOS-derived sols including doped ones. Precursors of sol–gel synthesis (i.e., TEOS–H3BO3, TEOS–epoxy resin) can be considered as morphogens. Viscous epoxy-siloxane sols can form films with a structure close to triply periodic minimal surfaces.
... 11,12 When a constant force is applied on Turing structures, the medium becomes anisotropic, and its morphology and type of pattern are affected. 11,12 The effect on chemical waves has been analyzed although not on spiral waves. 12 More complicated cases can be envisioned when the amplitude of the centrifugal force is periodically modulated. ...
... III. STATIONARY PATTERNS: TURING STRUCTURESIn Ref.11, experiments were reported describing the effect of a constant centrifugal force on Turing structures. Experiments were conducted in the Belousov-Zhabotinsky reaction 6,15 encapsulated into AOT micelle 15 (BZ-AOT system). ...
Article
Full-text available
The effect of centrifugal forces is analyzed in a pattern-forming reaction-diffusion system. Numerical simulations conducted on the appropriate extension of the Oregonator model for the Belousov-Zhabotinsky reaction show a great variety of dynamical behaviors in such a system. In general, the system exhibits an anisotropy that results in new types of patterns or in a global displacement of the previous one. We consider the effect of both constant and periodically modulated centrifugal forces on the different types of patterns that the system may exhibit. A detailed analysis of the patterns and behaviors observed for the different parameter values considered is presented here.
Article
Full-text available
Turing structures appear naturally and they are demonstrated under different spatial configurations such as stripes and spots as well as mixed states. The traditional tool to characterize these patterns is the Fourier transformation, which accounts for the spatial wavelength, but it fails to discriminate among different spatial configurations or mixed states. In this paper, we propose a parameter that clearly differentiates the different spatial configurations. As an application, we considered the transitions induced by an external forcing in a reaction-diffusion system although the results are straightforwardly extended to different problems with similar topologies. The method was also successfully tested on a temporally evolving pattern.
Article
Full-text available
Advances in nonequilibrium pattern formation in reaction-diffusion systems are reviewed. Special emphasis is placed on patterns found in the spatially extended Belousov-Zhabotinsky reaction dispersed in aerosol OT water-in-oil microemulsions (BZ-AOT system): Turing patterns, packet and standing waves, antispirals and segmented spirals, and accelerating waves and oscillons. All experimental results are explained theoretically and reproduced in computer simulations.
Book
Introduction. Computer Program Packages. Typical Problems. Classification of Equations. Discrete Methods. Finite Differences and Computational Molecules. Finite Difference Operators. Method of Weighted Residuals. Finite Elements. Method of Lines. Errors. Stability and Convergence. Irregular Boundaries. Choice of Discrete Network. Dimensionless Forms. Parabolic Equations: Introduction. Properties of a Simple Explicit Method. Fourier Stability Method. Implicit Methods. Additional Stability Considerations. Matrix Stability Analysis. Extension of Matrix Stability Analysis. Consistency, Stability, And Convergence. Pure Initial Value Problems. Variable Coefficients. Examples of Equations with Variable Coefficients. General Concepts of Error Reduction. Methods of Lines (MOL) for Parabolic Equations. Weighted Residuals and the Method of Lines. Bubnov-Galerkin Scheme for Parabolic Equations. Finite Elements and Parabolic Equations. Hermite Basis. Finite Elements and Parabolic Equations. General Basis Fucntion. Finite Elements and Parabolic Equations. Special Basis Functions. Explicity Finite Difference Methods for Nonlinear Problems. Further Applications on One Dimentions. Asymmetric Approximations. Elliptic Equations: Introduction. Simple Finite Difference Schemes. Direct Methods. Iterative Methods. Linear Elliptic Equations. Some Poit Iterative Methods. Convergence of Point Iterative Methods. Rates of Convergence. Accelerations. Conjugate Gradient Method. Extensions of SOR. Auliative Examples of Over-Relaxation. Other Point Iterative Methods. Block Iterative Methods. Alternating Direction Methods. Summary of ADI Results. Triangular Elements. Boundary Element Method (BEM). Spectral Methods. Some Nonlinear Examples. Hyperbolic Equations: Introduction. The Quasilinear System. Introductory Examples. Method of Characteristics. Constant States and Simple Waves. Typical Application of Characteristics. Finite Differences for First-Order Equations. Lax-Wendroff Methods and Other Algorithms Dissipation and Dispersion. Explicity Finite Difference Methods. Attenuation. Implicit Methods for Second-Order Equations. Time Quasilinear Examples. Simultaneous First-Order Equations. Explicit Methods. An Implicity Method for First-Order Equations. Hybrid Methods for First-Order Equations. Finite Elements and the Wave Equation. Spectral Methods and Periodic Systems. Gas Dyunamics in One Space Variable. Eulerian Difference Equations. Lagrangian Difference Equations. Hopscotch Methods for Conservation Laws. Explicity-Implicity Schemes for Conservation Laws. Special Topics: Introduction. Singularities. Shocks. Eigenvale Problems. Parabolic Equations in Segeral Space Variables. Additional Comments on Elliptic Equations. Hyperbolic Equations in Higher Dimensions. Mixed Systems. Higher Order Equations in Elasticity and Vibrations. Computational Fluid Mechanics. Stream Function. Vorticity Method for Fluid Mechanics. Primitive Variable Methods for Fluid Mechanics. Vector Potential Methods for Fluid Mechanics. Introduction to Monte Carlo Mehtods. Fast Fourier Transform and Applications. Method of Fractional Steps. Applications of Group Theory in Computation. Computational Ocean Acoustics. Enclosure Methods. Chapter References. Author Index. Subject Index.
Article
It is suggested that a system of chemical substances, called morphogens, reacting together and diffusing through a tissue, is adequate to account for the main phenomena of morphogenesis. Such a system, although it may originally be quite homogeneous, may later develop a pattern or structure due to an instability of the homogeneous equilibrium, which is triggered off by random disturbances. Such reaction-diffusion systems are considered in some detail in the case of an isolated ring of cells, a mathematically convenient, though biologically unusual system. The investigation is chiefly concerned with the onset of instability. It is found that there are six essentially different forms which this may take. In the most interesting form stationary waves appear on the ring. It is suggested that this might account, for instance, for the tentacle patterns on Hydra and for whorled leaves. A system of reactions and diffusion on a sphere is also considered. Such a system appears to account for gastrulation. Another reaction system in two dimensions gives rise to patterns reminiscent of dappling. It is also suggested that stationary waves in two dimensions could account for the phenomena of phyllotaxis. The purpose of this paper is to discuss a possible mechanism by which the genes of a zygote may determine the anatomical structure of the resulting organism. The theory does not make any new hypotheses; it merely suggests that certain well-known physical laws are sufficient to account for many of the facts. The full understanding of the paper requires a good knowledge of mathematics, some biology, and some elementary chemistry. Since readers cannot be expected to be experts in all of these subjects, a number of elementary facts are explained, which can be found in text-books, but whose omission would make the paper difficult reading.
Article
IN 1952, Turing proposed a hypothetical molecular mechanism, called the reaction–diffusion system, which can develop periodic patterns from an initially homogeneous state. Many theoretical models based on reaction–diffusion have been proposed to account for patterning phenomena in morphogenesis, but, as yet, there is no conclusive experimental evidence for the existence of such a system in the field of biology. The marine angel fish, Pomacanthus has stripe patterns which are not fixed in their skin. Unlike mammal skin patterns, which simply enlarge proportionally during their body growth, the stripes of Pomacanthus maintain the spaces between the lines by the continuous rearrangement of the patterns. Although the pattern alteration varies depending on the conformation of the stripes, a simulation program based on a Turing system can correctly predict future patterns. The striking similarity between the actual and simulated pattern rearrangement strongly suggests that a reaction–diffusion wave is a viable mechanism for the stripe pattern of Pomacanthus.
Article
Using an extension of the Swift-Hohenberg equation we study pattern formation in the Bénard experiment close to the onset of convection in the case of rotating cylindrical fluid containers. For small Taylor numbers we emphasize the existence of slowly rotating patterns and describe behaviour exhibiting defect motion. Finally, we study pattern formation close to the Küppers-Lortz instability. The instability is nucleated at defects and proceeds through front propagation into the bulk patterns.
Article
The relation between the diffusion coefficient of reactants and the wavelength of Turing patterns is examined in experiments on the chlorite-iodide-malonic acid (CIMA) reaction in gel media. The diffusion coefficients in polyacrylamide and agarose gels are varied by varying the gel densities. The diffusion coefficient D of NaCl is found to vary from 0.5×10-5 to 1.8×10-5 cm2/s for the gel conditions considered. The CIMA reactants are assumed to have diffusion coefficients that are directly proportional to that of NaCl. The wavelength λ of the observed hexagonal patterns (0.13-0.28 mm) varies in accord with the predicted relation for Turing patterns, λ∼D1/2. Moreover, the predicted relationship to a characteristic period of oscillation τ, λ = (2πτD)1/2, is supported by measurements of τ just beyond a Hopf bifurcation in a stirred flow reactor.
Article
The formation of localized structures in the chlorine dioxide-idodine-malonic acid (CDIMA) reaction-diffusion system is investigated numerically using a realistic model of this system. We analyze the one-dimensional patterns formed along the gradients imposed by boundary feeds, and study their linear stability to symmetry-breaking perturbations (Turing instability) in the plane transverse to these gradients. We establish that an often-invoked simple local linear analysis that neglects longitudinal diffusion is inappropriate for predicting the linear stability of these patterns. Using a fully nonuniform analysis, we investigate the structure of the patterns formed along the gradients and their stability to transverse Turing pattern formation as a function of the values of two control parameters: the malonic acid feed concentration and the size of the reactor in the dimension along the gradients. The results from this investigation are compared with existing experiments.