Turing patterns in a simple gel reactor

Similar documents
Dynamic Mechanism of Photochemical Induction of Turing Superlattices in the Chlorine Dioxide-Iodine-Malonic Acid Reaction-Diffusion System

Spatial Periodic Perturbation of Turing Pattern Development Using a Striped Mask

Locking of Turing patterns in the chlorine dioxide iodine malonic acid reaction with one-dimensional spatial periodic forcing

Turing pattern formation induced by spatially correlated noise

7. Well-Stirred Reactors IV

Turing structures. Progress toward a room temperature, closed system. Irving R. Epstein, Istvfin Lengyel

Spatial variation of a short-lived intermediate chemical species in a Couette reactor

Resonance in periodically inhibited reaction diffusion systems

Resonant suppression of Turing patterns by periodic illumination

Krisztina Kurin-Cso1rgei, Anatol M. Zhabotinsky,*, Miklós Orbán,*, and Irving R. Epstein

TABLE OF CONTENT. Chapter 4 Multiple Reaction Systems 61 Parallel Reactions 61 Quantitative Treatment of Product Distribution 63 Series Reactions 65

arxiv:chao-dyn/ v1 12 Feb 1996

Lecture (9) Reactor Sizing. Figure (1). Information needed to predict what a reactor can do.

CHEMICAL ENGINEERING LABORATORY CHEG 4137W/4139W. Reaction Kinetics Saponification of Isopropyl Acetate with Sodium Hydroxide

On the nature of patterns arising during polymerization of acrylamide in the presence of the methylene blue-sulfide-oxygen oscillating reaction

Time-periodic forcing of Turing patterns in the Brusselator model

Chromatography Lab # 4

Complexity of Hydrodynamic Phenomena Induced by Spiral Waves in the Belousov-Zhabotinsky Reaction

CENG 501 Examination Problem: Estimation of Viscosity with a Falling - Cylinder Viscometer

Chemical Reaction Engineering Prof. Jayant Modak Department of Chemical Engineering Indian Institute of Science, Bangalore

CHAPTER 4 EXPERIMENTAL METHODS

Homogeneity of optical glass

EXPERIMENT 10 The Activity Series

The Blue Bottle Experiment and Pattern Formation in this System

B.Tech. First Semester Examination Physics-1 (PHY-101F)

Exercise 2-2. Titration of a Strong Acid EXERCISE OBJECTIVES

Well Stirred Reactor Stabilization of flames

Excitation of Waves in a Belousov-Zhabotinsky System in Emulsion Media

Part of the practical procedure is given below.

Shear instabilities in a tilting tube

How Silica Aerogels Are Made

Oscillatory Turing Patterns in a Simple Reaction-Diffusion System

1 Background: The Gray-Scott Reaction-Diffusion System

CFD STUDIES IN THE PREDICTION OF THERMAL STRIPING IN AN LMFBR

FLAME PHOTOMETRY AIM INTRODUCTION

Complex Patterns in a Simple System

Black spots in a surfactant-rich Belousov Zhabotinsky reaction dispersed in a water-in-oil microemulsion system

Recovery of Copper Renee Y. Becker Manatee Community College

Lecture 8. Mole balance: calculations of microreactors, membrane reactors and unsteady state in tank reactors

Hands-on table-top science

Contains ribosomes attached to the endoplasmic reticulum. Genetic material consists of linear chromosomes. Diameter of the cell is 1 m

M2 TP. Low-Energy Electron Diffraction (LEED)

Chem 2115 Experiment # 6 PERIODIC RELATIONSHIPS

Pattern Formation and Spatiotemporal Chaos in Systems Far from Equilibrium

Cardiac muscle waves (1), calcium waves in cells (2), glycolytic

Deep Desulfurization of Diesel using a Single-Phase Micro-reactor

Experiment 2: Reaction Stoichiometry by Thermometric Titration

Membrane-bound Turing patterns

Chemically Driven Convection in the Belousov-Zhabotinsky Reaction Evolutionary Pattern Dynamics

Asia Flow Chemistry System

PERIODIC RELATIONSHIPS

Investigation of Petasis and Ugi Reactions in Series in an Automated Microreactor System

Modulation Instability of Spatially-Incoherent Light Beams and Pattern Formation in Incoherent Wave Systems

Numerical exploration of diffusion-controlled solid-state reactions in cubic particles

Chapter Two (Chemistry of Life)

Technology offer: Environmentally friendly holographic recording material

Diffusion and Adsorption in porous media. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

EXPERIMENT C3: SOLUBILITY PRODUCT & COMMON ION EFFECT. Learning Outcomes. Introduction. Upon completion of this lab, the student will be able to:

Edexcel Chemistry Checklist

4.4.1 Reactivity of metals Metal oxides The reactivity series. Key opportunities for skills development.

Festival of the Mind Chaotic Chemical Waves: Oscillations and waves in chemical systems. Dr. Jonathan Howse Dr.

Formation of white-eye pattern with microdischarge in an air. dielectric barrier discharge system

CIE Chemistry A-Level Practicals for Papers 3 and 5

In-Plane Liquid Distribution In Nonwoven Fabrics: Part 2 Simulation

IDEAL REACTORS FOR HOMOGENOUS REACTION AND THEIR PERFORMANCE EQUATIONS

Use of Bifurcation Diagrams as Fingerprints of Chemical Mechanisms

Liquid-Rocket Transverse Triggered Combustion Instability: Deterministic and Stochastic Analyses

hydrideicp Hydride Generation System

CHAPTER 3 MODELLING AND ANALYSIS OF THE PACKED COLUMN

On the effects of Non-Newtonian fluids above the ribbing instability

Hysteresis-free reactive high power impulse magnetron sputtering

AN014e. Non-standard geomtries for rheological characterization of complex fluids. A. Franck, TA Instruments Germany

5.4 Chemical changes Reactivity of metals Metal oxides The reactivity series. Key opportunities for skills development

CHM 130 Acid-Base Titration Molarity of Acetic Acid in Vinegar

Drift velocity of rotating spiral waves in the weak deformation approximation

Semi-analytical solutions for cubic autocatalytic reaction-diffusion equations; the effect of a precursor chemical

Reactors. Reaction Classifications

White Paper FINAL REPORT AN EVALUATION OF THE HYDRODYNAMICS MECHANISMS WHICH DRIVE THE PERFORMANCE OF THE WESTFALL STATIC MIXER.

Chemical Reaction between Solids x 2 = Kt

Heat and Mass Transfer Prof. S.P. Sukhatme Department of Mechanical Engineering Indian Institute of Technology, Bombay

METHOD 9012 TOTAL AND AMENABLE CYANIDE (COLORIMETRIC, AUTOMATED UV)

Research Article Chlorine Dioxide-Iodide-Methyl Acetoacetate Oscillation Reaction Investigated by UV-Vis and Online FTIR Spectrophotometric Method

On the Early Development of Dispersion in Flow through a Tube with Wall Reactions

Chemical Kinetics and Reaction Engineering Midterm 1

Quiz 5 Introduction to Polymers

Chemical Kinetics and Reaction Engineering

Outline. Definition and mechanism Theory of diffusion Molecular diffusion in gases Molecular diffusion in liquid Mass transfer

CFD STUDY OF MASS TRANSFER IN SPACER FILLED MEMBRANE MODULE

Nucleophilic displacement - Formation of an ether by an S N 2 reaction The Williamson- Ether Synthesis

PAPER AND THIN LAYER CHROMATOGRAPHY (TLC)

LAB #6 Chromatography Techniques

Kinetics of an Iodine Clock Reaction

Introduction to the course ``Theory and Development of Reactive Systems'' (Chemical Reaction Engineering - I)

Atoms, Molecules, and Life

Part I.

Kinetics of an Iodine Clock Reaction

e - Galvanic Cell 1. Voltage Sources 1.1 Polymer Electrolyte Membrane (PEM) Fuel Cell

EXPERIMENT NINE Part I - The Standardization of Thiosulfate Solutions

Chemistry 283g- Experiment 4

Page 2. The tripeptide shown is formed from the amino acids alanine, threonine and lysine.

Transcription:

Physica A 188 (1992) 17-25 North-Holland Turing patterns in a simple gel reactor R. Dennis Vigil, Q. Ouyang and Harry L. Swinney Center for Nonlinear Dynamics, Department of Physics, The University of Texas, Austin, TX 78712, USA We introduce a very simple open two-dimensional gel reactor for studying chemical patterns that arise in reaction-diffusion systems. The reactor consists of a thin disk-shaped layer of polyvinyl alcohol gel with one face in direct contact with a well-stirred flow reactor containing reactants of the chlorite-iodide-malonic acid system; the other face of the gel is in contact with a piece of transparent plexiglass. We have used this reactor to study the transition from a uniform state to a hexagonal pattern; for other concentrations, striped patterns were observed. The wavelength of these structures is greater than the thickness of the gel layer, which indicates that the patterns are two-dimensional (a single layer). This reactor provides a promising new tool for studying chemical patterns since the diffusion time of reactants into and out of the gel can be made small compared to the total residence time in the stirred flow reactor. This feature facilitates the comparison between theory and experiment since the chemical concentrations leading to pattern formation are close to those in the stirred flow reactor and hence can be directly measured. 1. Introduction Pattern formation in reaction-diffusion systems has attracted the interest of experimentalists and theorists alike during the last few decades. Until recently experimental work centered on spatio-temporal patterns in closed systems. Typically, a thin layer of reactive solution was placed in a petri dish and patterns such as spiral waves and target patterns were observed. However, since these appeared in a closed system, they were only short-lived transients. Several open chemical reactor designs have been introduced in order to allow the study of chemical patterns at conditions maintained far from equilibrium. Most of these designs employ an inert gel reaction medium in which convection is suppressed; such a reactor has been called a CFUR (continuous flow unstirred reactor), in analogy with the acronym CSTR used for continuous flow stirred tank reactors. In 1987 Noszticzius et al. [1] used a CFUR with an annular gel that was fed at the inner and outer rims to study traveling waves in the Belousov-Zhabotinskii (BZ) reaction. Another CFUR design consists of a thin gel layer that is sandwiched between two thin porous glass plates [2]. The face of each porous glass plate opposite to the gel is in contact with a stirred 0378-4371/92/$05.00 (~) 1992- Elsevier Science Publishers B.V. All rights reserved

18 R.D. Vigil et al. / Turing patterns in a simple gel reactor chamber that is continuously fed with chemicals; the chemicals in each chamber are chosen to be unreactive. The chemicals diffuse through the porous glass plates into the gel, forming crossed gradients in the gel. The concentrations in planes parallel to the glass plates would be uniform in the absence of (a) (b) Fig. 1. Stationary chemical patterns formed in the CSTR-membrane reactor with a gel layer 0.27 mm thick. (a) Hexagonal structures with wavelength 0.32 mm ([CH2(COOH)2] = 1.0 ram). (b) Striped patterns with wavelength 0.52mm ([CHz(COOH)2] =2.2mM). The parameters common to both patterns were [ClOz] = 6.0mM, [H2SO~] - 5.0mM, [Nal] = 3.5 ram, and temperature 7.0 C. The chlorite reservoir also contains sodium hydroxide (1 ram) to stabilize the chlorite. The bar beside each picture represents 1 cm. Wavelengths were calculated from twodimensional Fourier transforms.

R.D. Vigil et al. / Turing patterns in a simple gel reactor 19 reaction, but the reaction-diffusion process can lead to the formation of spatial patterns. This membrane reactor has been used by Ouyang and Swinney [3] in the first study of the transitions from uniform concentration to stationary chemical patterns (Turing patterns). Turing had predicted stationary spatial structures four decades ago [4], but the first observation in laboratory experiments was made only in 1990 by the Bordeaux group [5-7]. The Bordeaux experiments and the subsequent experiments of Ouyang and Swinney were conducted on the chlorite-iodide-malonic acid (CIMA) reaction. Further experiments by Ouyang and Swinney [8], using the membrane reactor, led to the discovery of a transition from Turing patterns to time-dependent patterns ("chemical turbulence"). A third design also consists of a thin gel layer, one side of which is in contact with a capillary array, which in turn is in contact with a CSTR; the other side of the gel is in contract with a glass plate. This reactor was used by Tam et al. [9] in studies of the formation of spirals in the BZ reaction and by Skinner and Swinney [10] in studies of the transition from simple to compound rotation of the spirals. The reactor developed for the present experiments is similar to the disk reactor of Tam et al. [9], except that the capillary array has been eliminated. The capillary array was originally thought to be necessary to prevent the stirring in the CSTR from disrupting the pattern formation. Unfortunately, because of the long diffusion time through the 1 mm thick capillary array, the concentrations imposed on the gel were not accurately known, which makes it difficult to make direct connection between the experimental observations and reaction-diffusion models. The present reactor is the ultimate in simplicity: a thin gel layer lies on one wall of a CSTR. Very high contrast Turing patterns have been observed with a gel thickness of only 0.27 mm, as fig. 1 illustrates. The diffusion time into the gel is very short; hence the concentrations of the feed chemicals for the reaction-diffusion process can be determined fairly accurately from measurements in the CSTR, and consequently it should be possible to make direct contact between the observations and reaction-diffusion models. We call this new CFUR a CSTR-membrane reactor. We will describe it and then present some results obtained using it. 2. Experimental system A schematic diagram of the chemical reactor and data aquisition system used in this study is shown in fig. 2. Three Pharmacia P-500 precision piston pumps

20 R.D. Vigil et al. / Turing patterns in a simple gel reactor CAMERA VIDEO ;UT TO [~ PC WITH FRAME I I 3RABBER QUARTZ WINDOW TO WASTE - ~-~'\\\\\\\\\~- 1 LIGftT (490 rim) SULFURIC~~ -:' ~ ~ MA ONICACID ( SODIUM IODIDE SODIUM tlydroxide SODIUM CItLORITE Fig. 2. Schematic of the CSTR-membrane reactor showing the configuration of the gel and reactant streams. The reactor is shown in cross section and is not to scale. The PVA gel plus Anapore thickness is 0.31 mm and the diameter is 25.4 mm (see text). are used to feed solutions from three reservoirs containing sulfuric acid, a malonic acid-sodium iodide mixture, and a sodium chlorite-sodium hydroxide solution (the NaOH is needed to stabilize the chlorite) into a 0.5 ml stirredtank reactor. The three reservoirs are separately unreactive and thus the reaction does not occur until the reagents are mixed in the CSTR. The reaction mixture then enters a small plexiglass chamber (1.6 ml volume) that is in direct contact with the gel (fig. 2). The disk-shaped gel is made from polyvinyl alcohol (PVA) instead of polyacrylamide, which has been used in most previous studies. The properties of PVA differ in several important respects from those of polyacrylamide. In previous work using polyacrylamide, the gel had to be loaded with a soluble starch indicator (usually Thiod~ne, Prolabo) in order to make spatial variations in triiodide concentration visible. In contrast, PVA acts both as a convectionfree reaction medium and is itself a triiodide indicator. This gel is also less brittle than polyacrylamide; hence the gel layer can be very thin. The thickness of the gel used in the present study was 0.27 mm; further studies could use even thinner gels. PVA gel was prepared by mixing 5 ml of 12 percent by weight PVA solution with 2 drops of 25 percent aqueous solution of glutaral-

R.D. Vigil et al. / Turing patterns in a simple gel reactor 21 dehyde at room temperature. Polymerization was initiated using 0.2ml of concentrated hydrochloric acid. After one-half hour the gel set and upon subsequent immersion in water it shrunk approximately 10 percent after several hours. Recently, Noszticzius et al. [11] have studied the role of triiodide indicators, including Thiodbne and polyvinyl alcohol, in the formation of Turing patterns. They have shown that in both cases triiodide reacts with the indicator to form an optically detectable complex. It is believed that these complexes are immobile and result in a rescaling of the effective diffusivity of some of the species [12], which is a necessary condition for a Turing instability to occur. However, Thiod6ne contains a large amount of urea, which may interfere with other reactions [11]. The possibility of this happening can be circumvented by using PVA gel instead of polyacrylamide loaded with Thiod~ne. The 0.27 mm thick gel disk is 2.54 cm in diameter and it rests on top of a transparent piece of plexiglass (see fig. 2). It is held in place around its edges by an overlaying piece of plexiglass with a circular 1.78 cm diameter opening, which exposes the gel to the small 1.6ml chamber containing the mixed reactants that are fed from the CSTR. High contrast patterns were obtained with this arrangement, but there was some difficulty at times with buckling of the gel. To prevent this buckling, experiments were also conducted with a thin inert rigid membrane (Anapore from Whatman) placed over the PVA gel; the Anapore disc is 0.04 mm thick and has an average pore size of 0.02 wm. The Anapore membrane keeps the gel flush with the plexiglass but is so thin that no difference could be observed in the patterns formed with and without this membrane. The membrane is brittle, and we therefore impregnated it with one drop of polyacrylamide gel (no starch) in order to make it stronger and more flexible. A peristalic pump operating at 1000 ml/hr is used to recirculate the reactants that exit from the chamber in order to maintain well-stirred conditions in the fluid above the gel. The entire reactor assembly is immersed in a constant temperature water bath. The PVA gel reacts with triiodide to form an immobile complex that absorbs light with a broad range of wavelengths that peaks near 490 nm, giving a color change from clear to a reddish-brown during the redox reaction. Therefore, patterns can be observed by using standard optical techniques. The gel is illuminated from below with a uniform light source with a peak intensity of 470-520 rim. A Sanyo CCD video camera fitted with a macro lens is placed above the gel to detect the intensity of transmitted light as a function of position in the plane of the gel. The video output from the camera is sent to an IBM PC-AT equipped with a 480 512 frame grabber. The acquired images are then analyzed on a Silicon Graphics workstation.

22 R.D. Vigil et al. / Turing patterns in a simple gel reactor 3. Results The system can be controlled by changing one or more of several control parameters including the concentration of the chemical reactants, the residence time, and the temperature. Beyond critical values of the control parameters patterns emerge spontaneously from an initially homogeneous state. For example, fig. 1 shows steady-state patterns observed at two different sets of conditions At low concentrations of malonic acid, hexagonal patterns separated by grain boundaries are observed (fig. la). These patterns evolve over a period of several hours before becoming stationary with only slight changes in the grain boundary positions. We have studied the transition from the homogeneous state to hexagons by using temperature as a control parameter while keeping all other parameters held fixed. As the temperature is decreased we find that the pattern emerges near 17 C, as shown in fig. 3. Within the experimental resolution we do not observe any hysteresis upon subsequent increase of temperature. We do not detect any critical slowing down near the transition, in contrast to previous work with a membrane reactor [3]. The wavelength of the patterns, 0.32 mm, also appears to be fairly insensitive to the temperature. At higher concentrations of malonic acid, striped patterns are observed, as illustrated in fig. lb. Again, these structures evolve toward a stationary state from an initially homogeneous state over a period of several hours. The 4" [] rn < 2 [3 [3 0. 4 6 8 10 12 14 16 18 TEMPERATURE ( C) Fig. 3. The transition from a uniform state to a hexagonal Turing pattern. The amplitude of the patterns was obtained from the modulus of a two-dimensional Fourier transform. The points (U]) were measured for temperature decreasing (increasing) in steps. The concentrations of the reactants were [CIO2] = 6.0 ram, [H2SO4] = 5.0 ram, [Nail = 3.5 mm and [CH2(COOH)2 ] = 2.0 mm.

R.D. Vigil et al. / Turing patterns in a simple gel reactor 23 appearance of stripes at larger concentrations of malonic acid is also accompanied by an increase in wavelength from 0.32mm (at [CH2(COOH)2 ] = 1.0mM, hexagons) to 0.56mm (at [CHz(COOH)2 ] =2.5mM, stripes). The transition from hexagons to stripes as the concentration of malonic acid is increased was also found in the membrane reactor, but the strong dependence of the wavelength on malonic acid was not observed [3]. We have also found that the appearance of Turing structures is sensitive to the residence time in the CSTR-membrane reactor. We define the residence time, r, to be the volume of mixed reactants divided by the total volumetric flow rate of the reactants. The volume containing the reaction mixture includes the CSTR, the tubing leading from the CSTR to the gel reactor, the small volume of reactants above the gel, and the recirculation loop. We have found patterns only when 60s<r<180s. The wavelength and character of the patterns changes very little as the residence time is varied in this range. 4. Discussion A major advantage of the CSTR-membrane reactor over previous reactors designed for the study of reaction-diffusion patterns is the simplicity of the present design. Another advantage is the small thickness of the gel. If we neglect chemical reaction and reflections off the boundary, we can obtain an upper bound estimate on the time scale for diffusion into and out of the gel from the relation t z -z2/2d, where D is the diffusivity and z is the gel thickness. In our system z = 0.031 cm (gel + Anapore), D ~ 10-5 cm2/s, and therefore we find t z-50s. We emphasize that this is only a largest-value estimate, and that the actual value could be much less. By comparison, typical residence times in the CSTR-membrane reactor at which we observe patterns are four times greater. Hence, the chemical concentrations probably do not vary greatly across the thickness of the gel and are near those in the CSTR. A second benefit of the short time-scale for diffusion into and out of the gel is that the spatial structures in the plane of the gel can evolve on a much shorter time-scale than they would in a membrane reactor. This could be especially important if one were interested in studying the dynamics of spatio-temporal structures such as those recently reported by Ouyang and Swinney [8]. Thirdly, it may be possible to draw a correspondence between experimental observations of temporal structures in a CSTR and spatial or spatio-temporal structures in the CSTR-membrane reactor. For example, Lengyel and Epstein [12] have argued, using certain assumptions, that the chemical conditions in a gel medium in which Turing structures are present should correspond to

24 R.D. Vigil et al. / Turing patterns in a simple gel reactor oscillations in a CSTR. In order to monitor the iodide concentration, we have placed a platinum probe in the recirculation loop of the CSTR-membrane reactor, and no oscillations have been observed when Turing patterns are present. This seems to suggest that at least some of the assumptions in the analysis of Lengyel and Epstein are not fulfilled in our system. Another important finding of our studies using the CSTR-membrane reactor is that the wavelength of the observed patterns (0.32-0.56 mm) is greater than the thickness of the gel. Thus, only a single-layer structure is present, and patterns do not depend upon the presence of structure orthogonal to the gel. In contrast to a CSTR-membrane reactor, in the membrane with reservoirs on both sides of the reactor [2, 3] the concentration gradients are deliberately created and maintained across the thickness. The gel rests between two membranes, which further complicates the understanding of the chemical concentrations at which the spatial structures occur. Furthermore, these membranes do not simply serve as supports for the gel. As the reactants diffuse towards each other across the membranes and into the gel, a reaction front is formed at some intermediate location between the boundaries. On one side of the front the system will be in the reduced state and on the other side it will be in the oxidized state. The presence of starch indicator in the gel will render a dark color in the reduced region. Therefore patterns cannot be optically detected unless the position of the reaction front is moved to the gel-membrane boundary so that the reduced region resides in the indicator-free membrane. Consequently, there are restrictive conditions on the feed concentrations and flow rates, and it is difficult to determine if there is a single or multi-layer structure. However, the membrane reactor has a significant advantage for some studies in that the reagents in each chamber can be separately unreactive; reaction occurs only in the gel layer. Hence this reactor is complementary to the CSTR-membrane reactor we have described. In summary, we have introduced a simple new gel reactor that consists of a thin layer of PVA that is fed uniformly over one face and which has no-flux boundary conditions on the other face. We have observed Turing patterns and transitions similar to those previously found in a membrane reactor and we have shown that these patterns are quasi two-dimensional. The CSTRmembrane reactor allows accurate characterization of the chemical conditions that lead to spatial pattern formation and thus it facilitates the comparison of reaction-diffusion models with experimental observations. Acknowledgements We would like to acknowledge fruitful discussions with Z. Noszticzius, W.D. McCormick, and K. Lee. This work is supported by the Department of Energy

R.D. Vigil et al. / Turing patterns in a simple gel reactor 25 Office of Basic Energy Sciences. R. Dennis Vigil acknowledges the support provided by a Ford Foundation post-doctoral fellowship, administered by the National Research Council. References [1] Z. Noszticzius, W. Horsthemke, W.D. McCormick and H.L. Swinney, Nature 329 (1987) 619. [2] G. Kshirsagar, Z. Noszticzius, W.D. McCormick and H.L. Swinney, Physica D 49 (1991) 5. [3] Q. Ouyang and H.L. Swinney, Nature 352 (1991) 610. [4] A.M. Turing, Philos. Trans. R. Soc. Lond. B 327 (1952) 37. [5] V. Castets, E. Dulos, J. Boissonade and P. De Kepper, Phys. Rev. Lett. 64 (1990) 2953. [6] J. Boissonade, V. Castets, E. Dulos and P. De Kepper, Int. Ser. Num. Math. 97 (1991) 67. [7] P. De Kepper, V. Castets, E. Dulos and J. Boissonade, Physica D 49 (1991) 161. [8] Q. Ouyang and H.L. Swinney, Chaos 1 (1991) 411. [9] W.Y. Tam, W. Horsthemke, Z. Noszticzius and H.L. Swinney, J. Chem. Phys. 88 (1988) 3395. [10] G.S. Skinner and H.L. Swinney, Physica D 48 (1991) 1. [11] Z. Noszticzius, Q. Ouyang, W.D. McCormick and H.L. Swinney, J. Phys. Chem. (June 1992). [12] I. Lengyel and I.R. Epstein, Proc. Natl. Acad. Sci. 89 (1992) 3977.