7
This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 12321 Cite this: Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 Low frequency temperature forcing of chemical oscillations Jan Novak, a Barnaby W. Thompson, b Mark C. T. Wilson, c Annette F. Taylor* b and Melanie M. Britton* a Received 7th April 2011, Accepted 12th May 2011 DOI: 10.1039/c1cp21096c The low frequency forcing of chemical oscillations by temperature is investigated experimentally in the Belousov-Zhabotinsky (BZ) reaction and in simulations of the Oregonator model with Arrhenius temperature dependence of the rate constants. Forcing with temperature leads to modulation of the chemical frequency. The number of response cycles per forcing cycle is given by the ratio of the natural frequency to the forcing frequency and phase locking is only observed in simulations when this ratio is a whole number and the forcing amplitude is small. The global temperature forcing of flow-distributed oscillations in a tubular reactor is also investigated and synchronisation is observed in the variation of band position with the external signal, reflecting the periodic modulation of chemical oscillations by temperature. 1. Introduction Oscillatory chemical processes take place in many biological systems, including NADH oscillations in yeast, cAMP oscilla- tions in the slime mold D. discoidium, and insulin pulses in pancreatic cells. 1,2 Oscillations in biological systems may be subject to periodic forcing by external parameters such as light. Entrained, quasiperiodic or chaotic responses are possible, depending on both the ratio of the forcing frequency, o f , to the natural frequency of the oscillator, o 0 , and the forcing ampli- tude A f . Modulation of the oscillations may also occur, for example the frequency of sensory neurons varies from cycle to cycle in response to variations in external stimuli. 3 In general, even with frequency modulation, if a fixed number of output cycles n is observed per m forcing cycles, then the system may be considered n:m synchronised. 4 Synchronisation of an oscillation with an external signal plays an important role in the functioning of natural systems, such as the circadian rythms governed by neurons of the suprachiasmatic nucleus (SCN) in mammals. Interestingly, while circadian rythms are insensitive to external changes in temperature in a phenomenon known as temperature compensation, 5 internal cycles in temperature driven by the SCN play an important role in synchronising other oscillatory cellular processes in the body to the day-night cycle. 6 Many chemical oscillators, including cAMP in slime mold 7 and the Belousov–Zhabotinsky (BZ) reaction, 8 exhibit an increase in the frequency of oscillations with increasing temperature. Changes in temperature can also influence the properties of oscillators, such as their precision. 9 The well- studied BZ reaction provides an ideal means for examination of the influence of temperature variations on chemical oscilla- tions and pattern formation. 10,11 Although the forcing of chemical oscillations with light has been extensively investigated using the light-sensitive ruthenium catalysed BZ reaction, 12 there are no investigations of temperature forcing of chemical oscillations. Additionally, previous studies with light concentrate on forcing with a similar or higher frequency than that of the natural frequency, rather than a low frequency. In the first part of this study we examine the influence of low-frequency temperature forcing on chemical oscillations exploiting the Belousov–Zhabotinsky reaction in a well-stirred closed reactor and the Oregonator model of the reaction with Arrhenius temperature dependence of the rate constants. We find that temperature forcing, even with a small amplitude of 1 K, results in significant modulation of the oscillatory frequency. Increasing the amplitude or period of forcing results in an increase in the range of frequencies observed. The number of chemical cycles per temperature cycle is given by the ratio of the natural frequency to the forcing frequency and phase locking is not observed in simulations unless this ratio is a whole number and the amplitude of the forcing is small. Thus it is not suprising that we did not obtain any phase locked states experimentally. The second part of this study examines the influence of temperature forcing on the BZ reaction with axial flow. A tubular packed bed reactor is fed from a continually-stirred a School of Chemistry, University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK. E-mail: [email protected]; Fax: +44 121 4144403; Tel: +44 121 4144391 b School of Chemistry, University of Leeds, Leeds, LS2 9JT, UK. E-mail: [email protected]; Fax: +44 113 3436565; Tel: +44 113 3436529 c School of Mechanical Engineering, University of Leeds, Leeds, LS2 9JT, UK. E-mail: [email protected] PCCP Dynamic Article Links www.rsc.org/pccp PAPER Published on 03 June 2011. Downloaded by Brown University on 29/10/2014 06:31:07. View Article Online / Journal Homepage / Table of Contents for this issue

Low frequency temperature forcing of chemical oscillations

Embed Size (px)

Citation preview

Page 1: Low frequency temperature forcing of chemical oscillations

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 12321

Cite this: Phys. Chem. Chem. Phys., 2011, 13, 12321–12327

Low frequency temperature forcing of chemical oscillations

Jan Novak,aBarnaby W. Thompson,

bMark C. T. Wilson,

cAnnette F. Taylor*

b

and Melanie M. Britton*a

Received 7th April 2011, Accepted 12th May 2011

DOI: 10.1039/c1cp21096c

The low frequency forcing of chemical oscillations by temperature is investigated experimentally

in the Belousov-Zhabotinsky (BZ) reaction and in simulations of the Oregonator model with

Arrhenius temperature dependence of the rate constants. Forcing with temperature leads to

modulation of the chemical frequency. The number of response cycles per forcing cycle is given

by the ratio of the natural frequency to the forcing frequency and phase locking is only observed

in simulations when this ratio is a whole number and the forcing amplitude is small. The global

temperature forcing of flow-distributed oscillations in a tubular reactor is also investigated and

synchronisation is observed in the variation of band position with the external signal, reflecting

the periodic modulation of chemical oscillations by temperature.

1. Introduction

Oscillatory chemical processes take place in many biological

systems, including NADH oscillations in yeast, cAMP oscilla-

tions in the slime mold D. discoidium, and insulin pulses in

pancreatic cells.1,2 Oscillations in biological systems may be

subject to periodic forcing by external parameters such as

light. Entrained, quasiperiodic or chaotic responses are possible,

depending on both the ratio of the forcing frequency, of, to the

natural frequency of the oscillator, o0, and the forcing ampli-

tude Af. Modulation of the oscillations may also occur, for

example the frequency of sensory neurons varies from cycle to

cycle in response to variations in external stimuli.3

In general, even with frequency modulation, if a fixed

number of output cycles n is observed per m forcing cycles,

then the system may be considered n:m synchronised.4

Synchronisation of an oscillation with an external signal plays

an important role in the functioning of natural systems,

such as the circadian rythms governed by neurons of the

suprachiasmatic nucleus (SCN) in mammals. Interestingly,

while circadian rythms are insensitive to external changes in

temperature in a phenomenon known as temperature

compensation,5 internal cycles in temperature driven by the

SCN play an important role in synchronising other oscillatory

cellular processes in the body to the day-night cycle.6

Many chemical oscillators, including cAMP in slime mold7

and the Belousov–Zhabotinsky (BZ) reaction,8 exhibit an

increase in the frequency of oscillations with increasing

temperature. Changes in temperature can also influence the

properties of oscillators, such as their precision.9 The well-

studied BZ reaction provides an ideal means for examination

of the influence of temperature variations on chemical oscilla-

tions and pattern formation.10,11

Although the forcing of chemical oscillations with light

has been extensively investigated using the light-sensitive

ruthenium catalysed BZ reaction,12 there are no investigations

of temperature forcing of chemical oscillations. Additionally,

previous studies with light concentrate on forcing with a

similar or higher frequency than that of the natural frequency,

rather than a low frequency.

In the first part of this study we examine the influence of

low-frequency temperature forcing on chemical oscillations

exploiting the Belousov–Zhabotinsky reaction in a well-stirred

closed reactor and the Oregonator model of the reaction with

Arrhenius temperature dependence of the rate constants. We

find that temperature forcing, even with a small amplitude of

1 K, results in significant modulation of the oscillatory

frequency. Increasing the amplitude or period of forcing

results in an increase in the range of frequencies observed.

The number of chemical cycles per temperature cycle is given

by the ratio of the natural frequency to the forcing frequency

and phase locking is not observed in simulations unless this

ratio is a whole number and the amplitude of the forcing is

small. Thus it is not suprising that we did not obtain any phase

locked states experimentally.

The second part of this study examines the influence of

temperature forcing on the BZ reaction with axial flow.

A tubular packed bed reactor is fed from a continually-stirred

a School of Chemistry, University of Birmingham, Edgbaston,Birmingham, B15 2TT, UK. E-mail: [email protected];Fax: +44 121 4144403; Tel: +44 121 4144391

b School of Chemistry, University of Leeds, Leeds, LS2 9JT, UK.E-mail: [email protected]; Fax: +44 113 3436565;Tel: +44 113 3436529

c School of Mechanical Engineering, University of Leeds, Leeds,LS2 9JT, UK. E-mail: [email protected]

PCCP Dynamic Article Links

www.rsc.org/pccp PAPER

Publ

ishe

d on

03

June

201

1. D

ownl

oade

d by

Bro

wn

Uni

vers

ity o

n 29

/10/

2014

06:

31:0

7.

View Article Online / Journal Homepage / Table of Contents for this issue

Page 2: Low frequency temperature forcing of chemical oscillations

12322 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 This journal is c the Owner Societies 2011

tank reactor (CSTR) containing the oscillatory BZ substrate.

Depending on the flow rate and chemical concentrations,

the system can display stationary patterns (flow-distributed

oscillations), downstream travelling chemical waves (a convective

instability) or bulk oscillations (an absolute instability).13,14 The

open flow with a fixed upstream boundary in the laboratory

system is equivalent, by a Galilean transformation, to a statio-

nary mediumwith a moving boundary.15 Thus these mechanisms

are applicable to biological morphogenesis processes,16 such as

somitogenesis in vertebrae development.17

Here we show that low frequency (global) temperature

forcing of flow-distributed oscillations in a tubular flow reactor

results in a periodic variation in band position, reflecting the

modulation of the chemical oscillatory frequency. Thus syn-

chronisation with the external signal is observed in space-time

plots of the flow-distributed oscillations.

2. Modelling section

2.1 Well-stirred reactor model

For the zero-dimensional well-stirred system, we utilised the

three variable five-step Oregonator model for the BZ reaction,18

derived from the FKN mechanism.19 The key reaction steps

are given in Table 1. Reactions (1) and (2) represent ‘‘Process A’’,

the removal of the inhibitor; reactions (3) and (4) represent

‘‘Process B’’, the autocatalytic oxidation of the catalyst limited

by disproportionation; and reaction (5) represents ‘‘Process C’’,

the process that ‘‘resets the clock’’ through generation of the

inhibitor. Generally these equations are scaled to give non-

dimensional values of the key intermediates: the autocatalyst

bromous acid, HBrO2 (X), the inhibitor bromide ion, Br� (Y),

and the oxidised metal catalyst, Mox (Z); resulting in the

dimensionless version of Oregonator model.20 However, these

scalings involve rate constants which vary with temperature;

hence it is not appropriate to scale these equations when the

temperature is time-dependent. The rate equations for the key

intermediates are given by:

dX

dt¼ r1 � r2 þ r3 � 2r4 ð1Þ

dY

dt¼ �r1 � r2 þ 1

2frc ð2Þ

dZ

dt¼ 2r3 � rc ð3Þ

where r1 to r5 represent the rate of reactions (1)–(5) in Table 1.

The rate constants were taken from the literature. The ODEs

(1)–(3) were solved using the XPPAUT package21 with integ-

ration method ‘‘stiff’’ and the parameter values were:

A = 0.16 M, B = 0.26 M, H = 0.17 M, f = 0.7 in this

work.

2.2 Reaction-diffusion-advection (RDA) model

To describe the BZ reaction in a 1D ‘‘plug-flow’’ reactor,

we transformed the ODE’s (1)–(3) to equations of the

form:22

@X

@t¼ f ðX;Y ;ZÞ þD

@2X

@r2�U

@X

@rð4Þ

Here f(X, Y, Z) represents the reaction terms, r is the spatial

coordinate where R is the total domain length = 4200 mm,

D is a typical diffusion coefficient = 2 � 10�3 mm2 s�1 and U

is the flow velocity = 1.4 mm s�1 in this work.

The equations were solved using an explicit finite difference

method for space and time, with spatial step size dr = 1 mm

and time step size dt = 0.001 s. Diffusion was approximated

using a central difference term while advection exploited a

backward difference term. The boundary condition at the

entry point of the reactor, X, Y, Z [r = 1, t] was set to the

steady state values in a CSTR with inverse residence time of

10�1 s�1 and inflow concentrations of Yi = 1 � 10�5 M and

Zi = 1 � 10�3 M. No flux boundary conditions were set at the

other boundary, i.e., X, Y, Z [r=R, t] =X, Y, Z [r=R� 1, t].

At all other grid points the initial concentrations, X, Y, Z

[r = 2. . .R � 1, t = 0], were set to (unstable) steady state

values for f(X,Y,Z) = 0. The parameter values were the same

as in section 2.1.

2.3 Temperature effects

In order to examine temperature effects in the model, rate

constant values at temperature T, kn,T, were expressed in

Arrhenius form, relative to rate constant values at T0 =

293 K (assuming the pre-exponential factor does not vary

with temperature):

kn;T ¼ kn;T0e�EnR

1T�

1T0

� �ð5Þ

where n designates the number of the reaction step, En is the

associated activation energy and R is the gas constant =

8.314 J K�1 mol�1. The rate constants and activation energies

(Table 1) were taken from earlier estimates,23,24 except for

E3 which was adjusted to match the experimental results

(see discussion).

Temperature forcing was achieved via sinusoidal varia-

tions about a mean of T1 with amplitude Af and forcing

frequency of:

T = Afsin(oft) + T1 (6)

Table 1 Oregonator model reactions where X = [HBrO2], Y = [Br�], Z = [Mox], A = [BrO3�], B = [MA + BrMA] (total organic species),

H = [H2SO4] and P = [HOBr], with respective rate constants and activation energies used in this work

Reaction Rate Rate constants (293 K) Activation energy/kJ mol�1

A + Y + 2H - X + P k1H2AY k1 = 2 M�1 s�1 E1 = 54

X + Y + H - 2P k2HXY k2 = 3 � 106 M�1 s�1 E2 = 25A + X + H - 2X + 2Z k3HAX k3 = 42 M�1 s�1 E3 = 402X - A + P + H k4X

2 k4 = 3000 M�1 s�1 E4 = 23B+Z - 1

2fY kcBZ kc = 1 M�1 s�1 E5 = 70

Publ

ishe

d on

03

June

201

1. D

ownl

oade

d by

Bro

wn

Uni

vers

ity o

n 29

/10/

2014

06:

31:0

7.

View Article Online

Page 3: Low frequency temperature forcing of chemical oscillations

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 12323

3. Experimental section

3.1 Well-stirred reactor

Malonic acid (Sigma, 99%), sodium bromate (Sigma, 99%),

sulfuric acid (Fisher, 98%), iron (II) sulfate heptahydrate

(Sigma, 98%) and 1, 10 phenanthroline (Sigma, 99%) were

all used without further purification. Ferroin was prepared by

mixing iron (II) sulfate heptahydrate and 1, 10 phenanthroline

in a 1 : 3 ratio. The initial concentrations were: [MA]0 = 0.26 M,

[NaBrO3]0 = 0.16 M, [H2SO4]0 = 0.13 M and [ferroin]0 =

1.86 � 10�3 M.

The reaction mixture was placed inside a water-jacketed

reaction vessel, which was stirred by a magnetic follower

driven by a magnetic stirrer (IKA). The temperature of the

vessel was controlled by a water bath (polystat cc3, Huber).

The temperature of the water bath was periodically changed

via a sawtooth waveform at various frequencies. Temporal

oscillations in the BZ reaction were followed by measuring the

potential difference of the reacting solution over time with a

Pt-combination electrode employing a Picoscope Oscilloscope

(Picotech) and the Picotech software.

3.2 Tubular flow reactor

Stationary chemical waves were produced in a water-jacketed

tube containing chemicals for the BZ reaction in a 0.6% w/v

agar medium (Fig. 4(a)). Stock solutions of bromate/malonic

acid (A) and acid/ferroin (B) were mixed with warm agar (C),

which formed a gel on cooling. Agar was used to produce

plug-flow in an alternative flow environment to a packed bed

reactor. This avoided the need for packing material, which can

create inhomogeneities in the flow.25 The gelation point was

controlled to be 5–10 cm from the tube inlet. Reactants/agar

were mixed in a continuously-fed stirred tank reactor (CSTR)

and pumped into a 16 mm internal diameter tube using

capillary tubing. By rapidly pumping the reaction mixture

from the CSTR into the reaction tube, the concentrations

at the boundary of the tube remained constant. Reactant

concentrations were as per the batch system in 3.1.

The temperature of the tube was periodically varied through

a sawtooth waveform and simultaneously measured using a

thermocouple. In the ferroin-catalysed BZ reaction, red

ferroin is oxidised to blue ferriin and hence the formation of

chemical waves was observed optically. A sequence of images

of the tube was taken at 20 s intervals, using a digital camera

(Canon A520). Space-time plots of the reaction were constructed

using Photoshop (Adobe).

4. Results

4.1 Temperature forcing in the well-stirred system

Experimentally, the natural period of oscillations at 288 K

between 10 000 and 20 000 s was on average 370 s. We note

that under the batch conditions given in section 3.1 there is a

25% decrease in period between 0 and 13 000 s and a further

decrease of less than 2% from 13 000 s to 23 000 s, thus the

latter part of the experiment was deemed to be suitable for

studying the effects of forcing with temperature. An increase

in temperature from 287 to 289 gave a 15% decrease in

period with a small decrease in amplitude. The Oregonator

model reproduces a decrease in the amplitude and period of

oscillations with an increase in temperature with the parameter

values given in section 2.1 (Fig. 1). A value of f was chosen to

reasonably match the experimental results: an increase in

temperature from 287 to 289 K resulted in a 17% decrease

in period from 422 to 353 s and the natural period was 386 s

at 288 K.

In Fig. 1(c) an oscillation is plotted with a linear increase in

temperature from 270 to 310 K. The peak-to-peak period was

determined with an increase in temperature in each of the rate

constants individually. The dominant step in determining the

overall decrease in period with increasing temperature was

reaction 1 (Table 1). Reactions 4 and 5 also resulted in a, less

marked, decrease in period, while reactions 2 and 3 resulted in

a slight increase in period with increasing temperature.

Low-frequency forcing with temperature about T1 = 288 K

results in periodic modulation of the chemical oscillations.

Fig. 2 shows experimental (a) and simulated data (b) and the

change in peak-to-peak period in time is shown in (c) and (d).

The range of periods observed in simulations was 354 to

420 s. The number of chemical oscillations per temperature

oscillation is around 8 and the plot of chemical period versus

temperature in the inset shows the synchronisation between

the oscillation in period and temperature oscillation. However

the periods are not the same during each loop around the

period-T cycle i.e. phase locking is not observed.

In Fig. 3(a)i, the time series is shown in simulations with

Af = 1 K and o0/of = 3.1. The range of periods observed

was 357–410 s. There are now 3 chemical oscillations per

temperature cycle. The period of the n + 1 cycle versus the

Fig. 1 Bifurcation diagrams in the Oregonator model showing

change in (a) amplitude of Z, (b) oscillation period with temperature.

(c) Oscillation in Z with a linear increase in temperature (upper plot)

and period (calculated as the time between the nth and n + 1 peak)

with a linear increase in temperature in each of the rate constants

individually (with the others fixed at T = 290 K).

Publ

ishe

d on

03

June

201

1. D

ownl

oade

d by

Bro

wn

Uni

vers

ity o

n 29

/10/

2014

06:

31:0

7.

View Article Online

Page 4: Low frequency temperature forcing of chemical oscillations

12324 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 This journal is c the Owner Societies 2011

period of the nth cycle is plotted in Fig. 3(a)ii illustrating that

there is no phase locking. When the ratio of natural to forcing

frequency is exactly 3, phase locking is observed and there

are 3 points in the corresponding period map (Fig. 3(b)).

However, as the amplitude of the forcing is increased, the

phase locking is lost (Fig. 3(c)). A larger range of periods is

observed (284–486 s) but there are still only 3 chemical

oscillations per temperature cycle.

In general we found that the number of chemical oscillations

per temperature oscillation was given by the ratio of the

natural to the forcing frequency. Phase locking was not

observed in any of the experimental runs with o0/of between

2 and 8 (increased in 0.1 increments), and only in simulations

when the ratio of the natural to the forcing frequency was a

whole number and the forcing amplitude was small.

4.2 Temperature forcing of flow distributed oscillations

When the BZ reaction is performed in a tubular flow reactor at

constant temperature, stationary chemical waves form. In the

kinematic limit, the chemical temporal dynamics are mapped

onto the flow axis.26 In other words, the flow carries a time

oscillating element, behaving as an individual batch reactor

down the tube (packets p1 etc. in Fig. 4(b)) while the fixed

inflow boundary condition locks the phase of oscillation. The

stationary chemical waves are essentially flow-distributed

oscillations with the wavelength given by l = Ut0 where U

is the flow velocity and t0 is the natural oscillatory period

(the relationship is in fact more complex as a result of

axial dispersive mixing, resulting in a shorter wavelength than

that predicted by the relation above).27,28 The formation of

stationary waves with constant temperature is shown in

Fig. 4(c). The first band of oxidised catalyst formed close to

the inlet, with bands forming subsequently along the length of

the tube via a wave-splitting mechanism described in detail

elsewhere.27

The effect of temperature forcing on the flow-distributed

oscillations is shown in Fig. 4(d). In the example, the

temperature is varied between 289 and 293 K with a forcing

frequency of 1 � 10�3 s�1, resulting in the ‘‘wavy’’ patterns

shown in the space-time plot.

The patterns are further analysed in Fig. 5. Four bands of

oxidised catalyst are observed in the space-time plot

(Fig. 5(a)), along with three diagonal lines that have slope

equal to the flow velocity of the fluid. The intensity profile

along each line corresponds to the concentration of oxidised

catalyst in time, following a packet of oscillating fluid as it

moves up the tube. If a packet enters the reactor on a

temperature low, the frequency increases with the subsequent

increase in global temperature (Fig. 5(b)i, iii); on the other

hand, if a packet enters on a temperature high, the frequency

decreases with the subsequent decrease in global temperature

Fig. 2 Temperature forcing of the BZ reaction where Af = 1 K,

o0/of =8.1 (see sections 2.1 and 3.1). Oscillations (baselined) and

period (calculated as the time between succesive peaks) in experiments:

(a) and (c), and simulations: (b) and (d). The inset shows the period of

the nth peak versus the temperature at the time of nth peak.

Fig. 3 Oscillations (i) and corresponding period maps (ii) in

simulations of the Oregonator model with temperature forcing where

o0/of = (a) 3.1 (b) 3.0 (c) 3.0 and Af = (a) 1, (b) 1, (c) 4.

Fig. 4 (a) Experimental set-up for the BZ reaction with axial flow

(section 3.2). (b) Illustration of consecutive images of the reactor in

time showing packets of fluid p1 etc. travelling with the flow. Space-

time plots (white = oxidized catalyst) from experimental images

showing formation of flow distributed oscillations with (c) constant

temperature, U = 0.12 cm s�1 (d) oscillations in temperature, U =

0.19 cm s�1.

Publ

ishe

d on

03

June

201

1. D

ownl

oade

d by

Bro

wn

Uni

vers

ity o

n 29

/10/

2014

06:

31:0

7.

View Article Online

Page 5: Low frequency temperature forcing of chemical oscillations

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 12325

(Fig. 5(b)ii). Thus modulation of the chemical oscillation is

observed in time.

The variation in band position is, however, synchronised

with the temperature oscillation (Fig. 5(c)). The linear plot

with negative slope for the first band (nearest to the inlet) is

indicative of frequency synchronisation with oscillations that

are anti-phase with the external temperature change. The

second and third bands are frequency synchronised with an

increasing phase delay relative to the temperature signal.

The experimental results are reproduced in RDA simulations

with temperature dependent rate constants. When Af = 0,

stationary concentration patterns form (Fig. 6(a)). When low

frequency temperature forcing is introduced, the bands that

form the stationary pattern adjust their position periodically

(Fig. 6(b)).

In agreement with the experimental findings, synchronisation

is observed in space-time. The variation in band position

versus temperature in Fig. 7(b) are closed loops. Each band

up the tube becomes progressively more delayed in phase

relative to the temperature forcing as distance increases from

the source.

The profiles obtained along the flow axis in the space-time

plot demonstrate the modulation of the oscillations in time —

an increase in frequency with increasing temperature and

vice versa (Fig. 7(c)). There are 4 chemical oscillations per

temperature cycle and the response is not phase locked with

the temperature signal, as illustrated in the period map

(Fig. 7(d)i); however for low amplitude forcing the first

9 peaks are clustered in 4 groups. This correlation is rapidly

lost as the amplitude of the forcing increases(Fig. 7(d)ii). We

did not find phase locking in time for any values of the forcing

frequency or amplitude explored.

Discussion

In the BZ reaction, the combination of activation energies

of the individual reaction steps governs the dependence of

oscillation period and amplitude on temperature. These

activation energies have been estimated experimentally for

the ferroin-catalysed BZ reaction,29 but there is still some

ambiguity with regard to the values, particularly E3.30 We

take a value of E3 = 40 kJ mol�1 to better reproduce the

generally observed experimental results of an amplitude and

frequency decrease with increasing temperature; the amplitude

increases with temperature if the value of E3 is greater than

40 kJ mol�1 for the temperature range and concentrations

used here.

As shown in earlier studies,31 the dominating step in deter-

mining the period of the oscillations is that of the reaction of

bromide with bromate to make bromous acid (reaction 1).

Increasing temperature in reactions 2 or 3 leads to a decrease

Fig. 5 (a) Measured variations in forcing temperature and band

position extracted from a grey-scale image of FDO patterns where

lines (i–iii) have slope = the flow velocity. (b) Profiles of the blue

intensity (i.e. relative value of Z) along flow lines (i–iii) in the grey-

scale image. (c) Position of the first three bands versus temperature

(about the mean s1, T1).

Fig. 6 Flow distributed oscillations in the 1D Oregonator RDA

model (section 2.2) where T1 = 291 K. (a) T = T1 and (b) T is

oscillatory with Af = 2 K; and o0/of = 4.

Fig. 7 Analysis of FDO patterns in Fig. 6(b). (a) Temperature and

space-time plot, with diagonal lines (i) and (ii) of slope = the flow

velocity. (b) Variation of band position with T (about the mean), for

the first 3 bands in (a). (c) Z profiles along the flow lines (i) and (ii) in

the grey-scale image, with the corresponding variation of temperature.

(d) the period of the n + 1 peak versus the period of the nth peak

(obtained from profile b(i) over 3000 s) for forcing amplitudes Af/K =

(i) 2, (ii) 5.

Publ

ishe

d on

03

June

201

1. D

ownl

oade

d by

Bro

wn

Uni

vers

ity o

n 29

/10/

2014

06:

31:0

7.

View Article Online

Page 6: Low frequency temperature forcing of chemical oscillations

12326 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 This journal is c the Owner Societies 2011

in period. Such opposing influences can result in temperature

compensation in some systems (i.e. a constant period over a

limited range in temperatures); not so in the BZ reaction which

displays a great temperature senstivity. Only a 2 K increase in

temperature from 287 K can result in a 16% decrease in

period.

Low frequency forcing with temperature results in modulation

of the chemical oscillations as the system is driven between two

limit cycles: one at low and one at high temperature. The lower

the frequency of the forcing, the more time the system has to

approach these limit cycles, thus the range of periods observed

increases with decreasing frequency and increasing amplitude

of the forcing.

It is possible to obtain phase locking even with modulation:

then the same peak-to-peak periods would be observed during

each forcing cycle. However we do not observe phase locking

in experiments on low frequency temperature forcing of the

BZ reaction. Preliminary investigations with high frequency

temperature forcing in the Oregonator model also suggest a

lack of phase locking. This is in stark contrast to the large

body of work on forcing of the oscillatory BZ reaction, where

entrainment is seen for a number of n:m regimes in purely

temporal32 as well as spatio-temporal systems.33–36 In all these

systems, the forcing typically involves periodic variations in

only one reaction rate within the rate equation of only one

variable. It is likely that the opposing forces generated by

temperature-dependent rate constants play a role in the

absense of phase locking in the temperature forced system.

Despite the lack of phase locking, synchronisation manifests

in the oscillation in the period with the oscillation in the

temperature. The period versus the next period forms a closed

loop and the number of points per transition around the loop

is given by the ratio of the natural to the forcing frequency.

Increasing the amplitude of the forcing simply results in a

greater degree of modulation and a larger range of periods

observed but the number of peaks per forcing cycle remains

the same.

Part of the interest in forcing with temperature comes from

the fact that this may provide a more general means of

controlling pattern formation in chemical/biological systems.12

Thus systems that are opaque could be manipulated. Tempera-

ture forcing of the stationary patterns in the flow reactor

indicates that while phase locking is not observed in time,

synchronisation does occur in space-time, manifested as the

‘‘wavy’’ change in band position for all values of the forcing

period and amplitude. This can be explained by the fact that

the band position is dictated by the oscillatory period. The

synchronisation reflects the periodic switch from a lengthening

in the oscillatory period to a decrease in the oscillatory period

with the changing temperature.

Menzinger and co-workers have shown that a periodic

change in the flow velocity in the FDO system also gives rise

to the ‘‘wavy’’ patterns.26 They have also simulated the

modulation of Fitz-Hugh-Nagumo FDO dynamics by a

sinusoidal variation in flow velocity. This results in periodic

longitudinal displacement of the FDO bands in the kinematic

limit. Low frequency periodic excitation by light of the

CDIMA reaction under flow and of the corresponding

Lengyel-Epstein model generates both FDO (for of/o0 o 1)

and Turing (for of/o0 4 1) patterns.37 Away from the

kinematic limit, the wave pattern can be disrupted, leading

to travelling waves with frequencies equal to a rational multi-

ple of the velocity perturbation frequency — a form of

resonance forcing — with resonances as high as 10 : 1,

i.e. for low frequency forcing.38 Further investigation of

temperature forcing of FDO patterns may also yield such

responses.

It remains to be seen what the influence of temperature

forcing on dynamic patterns (such as travelling waves) will be.

It is evident that the forcing amplitude and frequency are key

parameters in determining pattern formation. Exploring

the effect of varying these and chemical parameters in the

temperature forced BZ reaction may shed further light on the

dynamics and the stability of chemical patterns subject to

external parameter variations. This in turn may provide insights

into the equivalent biological processes.16,17,39–41

Conclusions

In summary, we have shown that low frequency forcing

with temperature modulates chemical oscillations in the BZ

reaction. The number of cycles per forcing is given by the ratio

of natural to forcing frequency and increasing the amplitude

of forcing results in a greater range of observed periods. Phase

locking is only observed in simulations when the ratio of the

forcing frequency to the natural is a whole number and the

amplitude is small. Temperature forcing of flow distributed

oscillations in the BZ reactions demonstrates that an entrained

periodic signal may be generated in space-time from a quasi-

periodic signal in time. We expect that the phenomena

observed may also be observed in biological systems with

temperature-dependent oscillatory frequencies, such as cAMP

in slime mold.

Acknowledgements

The authors thank the EPSRC Grants EP/F048777/1 and

EP/F050410/1 and the Royal Society for funding.

Notes and references

1 A. Goldbeter, Biochemical Oscillations and Cellular Rhythms,Cambridge University Press, Cambridge, 1996.

2 A. T. Winfree, The Geometry of Biological Time, Springer,New York, 1980.

3 C. E. Schroeder and P. Lakatos, Trends Neurosci., 2009, 32, 9–18.4 A. Pikovsky, M. Rosenblum and J. Kurths, Synchronisation:a universal concept in nonlinear sciences, Cambridge UniversityPress, Cambridge, 2001.

5 S. Sen, S. S. Riaz and D. S. Ray, J. Theor. Biol., 2008, 250,103–112.

6 E. D. Buhr, S. H. Yoo and J. S. Takahashi, Science, 2010, 330,379–385.

7 B. Wurster, Nature, 1976, 260, 703–704.8 E. Koros, Nature, 1974, 251, 703–704.9 M. Wickramasinghe and I. Z. Kiss, Chaos, 2010, 20, 23125.10 T. Bansagi, M. Leda, M. Toiya, A. M. Zhabotinsky and

I. R. Epstein, J. Phys. Chem. A, 2009, 113, 5644–5648.11 R. McIlwaine, V. K. Vanag and I. R. Epstein, Phys. Chem. Chem.

Phys., 2009, 11, 1581–1587.12 A. S. Mikhailov and K. Showalter, Phys. Rep., 2006, 425, 79–194.13 M. Kaern and M. Menzinger, Phys. Rev. E: Stat. Phys., Plasmas,

Fluids, Relat. Interdiscip. Top., 1999, 60, R3471–R3474.

Publ

ishe

d on

03

June

201

1. D

ownl

oade

d by

Bro

wn

Uni

vers

ity o

n 29

/10/

2014

06:

31:0

7.

View Article Online

Page 7: Low frequency temperature forcing of chemical oscillations

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 12321–12327 12327

14 S. P. Kuznetsov, E. Mosekilde, G. Dewel and P. Borckmans,J. Chem. Phys., 1997, 106, 7609–7616.

15 P. N. McGraw, M. Menzinger and A. P. Munuzuri, Phys. Rev. E:Stat., Nonlinear, Soft Matter Phys., 2009, 80, 26209.

16 M. Kaern, M. Menzinger, R. Satnoianu and A. Hunding, FaradayDiscuss., 2002, 120, 295–312.

17 M. Kaern, M. Menzinger and A. Hunding, Biophys. Chem., 2000,87, 121–126.

18 R. J. Field and R. M. Noyes, J. Am. Chem. Soc., 1974, 96,2001–2006.

19 R. J. Field, E. Koros and R. M. Noyes, J. Am. Chem. Soc., 1972,94, 8649.

20 J. J. Tyson and P. C. Fife, J. Chem. Phys., 1980, 73,2224–2237.

21 B. Ermentrout, Simulating, Analyzing and Animating DynamicalSystems, SIAM, Philadelphia, 2002.

22 J. R. Bamforth, S. Kalliadasis, J. H. Merkin and S. K. Scott, Phys.Chem. Chem. Phys., 2000, 2, 4013–4021.

23 J. A. Agreda and R. J. Field, J. Phys. Chem. A, 2006, 110,7867–7873.

24 G. Kshirsagar and R. J. Field, J. Phys. Chem., 1988, 92,7074–7079.

25 A. F. Taylor and M. M. Britton, Chaos, 2006, 16, 37103.26 P. N. McGraw and M. Menzinger, Phys. Rev. E: Stat., Nonlinear,

Soft Matter Phys., 2005, 72, 27202.27 J. R. Bamforth, R. Toth, V. Gaspar and S. K. Scott, Phys. Chem.

Chem. Phys., 2002, 4, 1299–1306.

28 M. M. Britton, A. J. Sederman, A. F. Taylor, S. K. Scott andL. F. Gladden, J. Phys. Chem. A, 2005, 109, 8306–8313.

29 J. Z. Zhang, L. Q. Zhou and O. Y. Qi, J. Phys. Chem. A, 2007, 111,1052–1056.

30 S. R. Pullela, D. Cristancho, P. He, D. W. Luo, K. R. Hall andZ. D. Cheng, Phys. Chem. Chem. Phys., 2009, 11, 4236–4243.

31 M. J. Blandamer and D. L. Roberts, J. Chem. Soc., Faraday Trans. 1,1977, 73, 1056–1065.

32 M. Dolnik, J. Finkeova, I. Schreiber and M. Marek, J. Phys.Chem., 1989, 93, 2764–2774.

33 A. Toth, V. Gaspar and K. Showalter, J. Phys. Chem., 1994, 98,522–531.

34 V. Petrov, Q. Ouyang and H. L. Swinney, Nature, 1997, 388,655–657.

35 A. L. Lin, M. Bertram, K. Martinez, H. L. Swinney, A. Ardeleaand G. F. Carey, Phys. Rev. Lett., 2000, 84, 4240–4243.

36 O. Steinbock, V. Zykov and S. C. Muller, Nature, 1993, 366,322–324.

37 D. G. Miguez, P. McGraw, A. P. Munuzuri and M. Menzinger,Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2009, 80, 26208.

38 P. N. McGraw, M. Menzinger and A. P. Munuzuri, Phys. Rev. E:Stat., Nonlinear, Soft Matter Phys., 2009, 80, 13.

39 L. Glass, Nature, 2001, 410, 277–284.40 L. G. a. M. C. Mackey, From Clock to Chaos: the Rhythms of Life,

Princeton University Press, Princeton, 1988.41 A. T. Winfree, The geometry of biological time, Springer,

New York, 2001.

Publ

ishe

d on

03

June

201

1. D

ownl

oade

d by

Bro

wn

Uni

vers

ity o

n 29

/10/

2014

06:

31:0

7.

View Article Online