Pattern Formation and Wave Propagation in the Belousov-Zhabotinskii Reaction

Similar documents
Oregonator model of the Belousov-Zhabotinsky reaction Richard K. Herz,

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

Chaos in the Belousov-Zhabotinsky Reaction

Reactions. John Vincent Department of Chemistry University of Alabama

CHEM-UA 652: Thermodynamics and Kinetics

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

REACTIONS THAT OSCILLATE

7. Well-Stirred Reactors IV

Answers to Unit 4 Review: Reaction Rates

k 1 HBrO 2 + HOBr (2) k 2 2HOBr (3) 3 2 BrO 2 + 2H 2 O (4) Ce 4+ + HBrO 2 (5)

An Analysis of the Belousov-Zhabotinskii Reaction

Chemiluminescence A Toast to Chemistry Chemiluminescence

Long Range Correlations in Chemical Oscillations

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

PhD Thesis. Investigation of the oxidation of iron(iii) complexes of porphyrin derivatives by bromate. Dénesné Rácz Krisztina

Development of Ce IV and Mn II Oscillometry: A New Estimation Method Based on Peak Potential and Oscillation Period of Chemical Oscillations

CHEM 515: Chemical Kinetics and Dynamics

Controlling Coupled Chemical Oscillators: Toward Synchronization Engineering and. Chemical Computation

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

Use this dramatic iodine clock reaction to demonstrate the effect of concentration, temperature, and a catalyst on the rate of a chemical reaction.

Dependence of Wave Speed on Acidity and Initial Bromate Concentration in the Belousov-Zhabotinsky Reaction-Diffusion System

Computational models: Reaction Diffusion. Matthias Vigelius Week 6

Pre-Lab Read the entire laboratory assignment. Answer all pre-lab questions before beginning the lab.

Kinetics; A Clock Reaction

Lectu re Notes in Biomathematics

Activation Energy of Aggregation-Disaggregation Self-Oscillation of Polymer Chain

Reaction Mechanism for Light Sensitivity of the Ru(bpy) 3. -Catalyzed Belousov-Zhabotinsky Reaction

Control of waves, patterns and turbulence in chemical systems

ACT Science Quick Guide

Numerical solution of stiff systems of differential equations arising from chemical reactions

How fast or slow will a reaction be? How can the reaction rate may be changed?

3 Mixtures. How do mixtures differ from elements and compounds? How can mixtures be separated? What are solutions, and how are they characterized?

Category V Physical Science Examples

Resonance in periodically inhibited reaction diffusion systems

Last Name: First Name: High School Name: Individual Exam 3 Solutions: Kinetics, Electrochemistry, and Thermodynamics

C1a The particulate nature of matter

2/22/2019 NEW UNIT! Chemical Interactions. Atomic Basics #19

Studies of the Catalyzed Bromate Ion-Benzoquinone- Sulfuric Acid Chemical Oscillator

Chapter 16. Rate Laws. The rate law describes the way in which reactant concentration affects reaction rate.

Module 3 - Thermodynamics. Thermodynamics. Measuring Temperatures. Temperature and Thermal Equilibrium

How do nerve cells behave? At

FLAME PHOTOMETRY AIM INTRODUCTION

Partner: Judy 6 October An Activity Series

Kinetics of an Iodine Clock Reaction Lab_Student Copy

Student Exploration: Chemical Changes

Advanced Higher Chemistry KINETICS. Learning Outcomes Questions & Answers

Name: Date: AP Chemistry. Titrations - Volumetric Analysis. Steps for Solving Titration Problems

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

Microfluidic Networks of the Compartmentalized Belousov-Zhabotinsky Reaction

#5 Chemical Kinetics: Iodine Clock Reaction

Examples of Excitable Media. Excitable Media. Characteristics of Excitable Media. Behavior of Excitable Media. Part 2: Cellular Automata 9/7/04

... [1] (ii) Draw a dot-and-cross diagram to show the bonding in NH 3

Directed Reading A. Section: Mixtures PROPERTIES OF MIXTURES. combined is a(n). of feldspar, mica, and quartz. SOLUTIONS

Photosynthesis: How do plants get engery? Teacher Version

Part A Answer all questions in this part.

Radius-Dependent Inhibition and Activation of Chemical Oscillations in Small Droplets

Kinetics of an Iodine Clock Reaction Lab_ Teacher s Key

Stresses Applied to Chemical Equilibrium

CIE Chemistry A-Level Practicals for Papers 3 and 5

KINETICS OF THE PERMANGANATE- ISOPROPYL ALCOHOL REACTION

Lab #14: Electrochemical Cells

UNIVERSITY OF CAMBRIDGE INTERNATIONAL EXAMINATIONS International General Certificate of Secondary Education CHEMISTRY

Describe in full the colour change at the end-point of this titration. ... (1)

THE EFFECT OF TEMPERATURE AND CONCENTRATION ON REACTION RATE

] 2+ and [Co(en) 3 CH 2. (30 g) ( en represents ethylenediamine, 1,2-diaminoethane or 1,2-ethanediamine) nickel(ii) chloride-6-water, NiCl 2 6H 2

Name Class Date. Read the words in the box. Read the sentences. Fill in each blank with the word or phrase that best completes the sentence.

Chemistry Final Exam Sample Items

Influence of Oxygen and Organic Substrate on Oscillations and Autocatalysis in the Belousov-Zhabotinsky Reaction

glossary 6 of. boxes with and units. question. the syllabus. to see if. nature. Day. Teacher. Pre readings. Topic. Mr Stocker.

Excellence International School Chemistry Academic Year Grade 9 Revision sheet 3 Topic: unit 8 speed of reaction

The rate of a chemical reaction is the speed at which the reactants are used up or the speed at which new products are formed.

"In Terms Of" 1. Explain, in terms of electron configuration, why arsenic and antimony are chemically similar.

EXPERIMENT 8 Reactions of Hydrocarbons

5.1. The Classification of Matter

NCEA COLLATED QUESTIONS ON RATES OF REACTION

6.3.4 Action potential

DYNAMICS OF A DISCRETE BRUSSELATOR MODEL: ESCAPE TO INFINITY AND JULIA SET

KINETIC STUDIES ON METAL-ION CATALYSED OSCILLATORY CHEMICAL REACTION WITH GALLIC ACID IN PRESENCE OF ADDITIVES

Reactions that Produce Heat

Use of Bifurcation Diagrams as Fingerprints of Chemical Mechanisms

Iodine Clock Challenge Rate Laws

Physical and Chemical Changes. 3. One of the new materials was a precipitate that settled out of solution.

Name Class Date. How do mixtures differ from elements and compounds? How can mixtures be separated? What are solutions?

Topic: APPLIED ELECTROCHEMISTRY. Q.1 What is polarization? Explain the various type of polarization.

Name Class Date. How do mixtures differ from elements and compounds? How can mixtures be separated? What are solutions?

Enzyme Catalysis. Objectives

A Clock Reaction: Determination of the Rate Law for a Reaction

EXPERIMENT 7- SAPONIFICATION RATE OF TERT- BUTYL CHLORIDE

AP Chemistry Unit 2 Test (Chapters 3 and 4)

Rate law Determination of the Crystal Violet Reaction Using the Isolation Method

Draw one line from each solution to the ph value of the solution. Solution ph value of the solution

Experiment 7 Buffer Capacity & Buffer Preparation

ph AND WATER Comparable substance

Experiment 7: SIMULTANEOUS EQUILIBRIA

SUPPLEMENTARY INFORMATION

The Design and Manipulation of Bromate-Based Chemical Oscillators

Chapter 5, Lesson 1: Water is a Polar Molecule

*AC134* Chemistry. Assessment Unit AS 3. [AC134] wednesday 27 MAY, MORNING. assessing Module 3: Practical Examination Practical Booklet B

Systems Biology Across Scales: A Personal View XXVI. Waves in Biology: From cells & tissue to populations. Sitabhra Sinha IMSc Chennai

UNIVERSITY OF CAMBRIDGE INTERNATIONAL EXAMINATIONS International General Certificate of Secondary Education

Transcription:

Pattern Formation and Wave Propagation in the Belousov-Zhabotinskii Reaction Dan Hill and Timothy Morgan University of California, San Diego Physics Department

INTRODUCTION The Belousov-Zhabotinskii (BZ) reaction is the earliest well-understood example of a chemical system that results in oscillations and pattern formation. It was discovered in 1951 by the Soviet biophysicist Boris Belousov. At the time, chemists were skeptical of the possibility of chemical oscillators because of a misconception about the second law of thermodynamics. As a result, Belousov was unable to get his work published. A decade later, another Soviet scientist named Anatol Zhabotinskii reproduced Belousov s experiment and successfully persuaded more chemists to accept the idea of chemical oscillators. In 197, three researches at the University of Oregon, Field, Körös, and Noyes, published a complete mechanism describing the BZ reaction, known as the FKN mechanism. However, their equations were too complex for numerical analysis by the computers of the time. Consequently, Field and Noyes distilled their model to a set of ordinary differential equations with only three variables, dubbed the Oregonator. Belousov s original experiment studied only temporal oscillations in a well-stirred solution; however, much more interesting is the formation of spatial patterns in an unstirred solution. In these cases, if the reaction begins at a given point, the concentrations of intermediates will propagate outward via diffusion, initiating the reaction in the adjacent regions. This is known as a trigger wave. Periodically the reaction will reinitiate at the nucleation point resulting in successive bands (in a onedimensional test tube reaction) or concentric rings (in a two-dimensional Petri dish reaction). Systems of chemical oscillators are of great interest in biological systems. For example, the sinoatrial node, the heart s pacemaker, causes trigger waves which, though electrical in nature, propogate in much the same way as BZ trigger waves. The generation of cyclic adenosine monophosphate in aggregating Dictyostelium discoideum slime mold colonies creates spiral patterns nearly identical to those of the BZ reaction. RNA has been found to to self-replicate, with fronts of increased RNA concentration moving outward via diffusion. (All examples Epstein and Pojman.) THEORY As stated before, early skeptics of the BZ reaction held that chemical oscillators would violate the nd law of thermodynamics. They believed the second law meant that all chemical concentrations in a reaction must move directly towards equilibrium. Many compared the BZ reaction to a damped pendulum, which passes through its equilibrium position during each oscillation and eventually comes to rest. A chemical system that did this would indeed violate the second law of thermodynamics. Passing through the equilibrium point and then moving away from it would require an increase in Gibbs free energy, which must always decrease. However, the BZ reaction (like all other chemical oscillators) does not reach its equilibrium point until after the oscillations are finished oscillatory behavior is a process involving only intermediates. The BZ reaction is actually a system of several chemical reactions with dozens of elementary steps, but the overall process is the oxidation of malonic acid by bromate producing carbon dioxide. As the system slowly progresses towards equilibrium, the concentrations of several

intermediate species oscillate while the concentrations of products move monotonically towards equilibrium. The following is a quick summary of the development of oscillatory behavior: Normally bromide reduces the bromate. This reaction is fast, quickly using up the available bromide. Once bromide drops below a critical level, bromous acid takes over the reduction of bromate in a reaction that autocatalytically produces more bromous acid. This leads to exponential growth in [HBrO ]. This is eventually checked by a reaction that converts HBrO to HOBr and bromate. Meanwhile, the decomposition of malonic acid results in the reduction of bromine to bromide, nearly restoring the initial concentration and allowing the whole process to begin again. Much of our understanding of the BZ reaction stems from the FKN mechanism. Below is an overview of the FKN mechanism. (For a more thorough treatment, see Tyson.) The key steps are: (1) BrO 3 + 5Br + 6H + 3Br + 3H 0 () HOBr + Br + H + Br + H 0 (3) BrO 3 + Br + H + HBrO + HOBr (4) Br + MA BrMA + Br + H + (5) Ce +3 + BrO 3 + HBrO + 3H + Ce +4 + HBrO + H 0 (6) HBrO HOBr + BrO 3 + H + (7) 6Ce +4 + MA + H 0 6Ce +3 + HCOOH + CO + 6H + (8) 4Ce +4 + BrMA + H 0 4Ce +3 + Br + HCOOH + CO 5H + (9) Br + HCOOH Br + CO + H + Important Points: When [Br ] is high, (1) dominates the reduction of bromate When [Br ] is low, (5) dominates the reduction of bromate (5) includes the autocatalytic production of HBrO ; when (5) is dominant, HBrO increases exponentially until limited by (6) (7), (8), and (9) are the final processes, oxidizing malonic and bromomalonic acid The rates of (7), (8), and (9) are much slower than those of the other reactions Double- and triple-brominated malonic acid also occur, but in much smaller concentrations than BrMA, so the FKN mechanism neglects these. They would be oxidized in a manner similar to BrMA. Oscillations occur in the following manner: Process (1) lowers the bromide concentration and increases bromine, allowing for the bromination of malonic acid by

(3). When [Br ] has been significantly lowered, (5) causes an exponential increase in bromous acid and the oxidized form of the metal ion catalyst and indicator, cerium. Bromous acid is subsequently converted to bromate and HOBr. Meanwhile, the ratelimiting steps (7), (8), and (9) reduce the cerium to Ce +3 and simultaneously increase bromide concentration. Once the bromide concentration is high enough, it reacts with bromate and HOBr in (1) and () to form bromine, and the process begins again. The different absorption spectra of the two ionization states of cerium causes the solution to change from yellow to clear and back as their relative concentrations change, allowing the oscillations to be observed visually. Two-dimensional spatial patterns can either form spontaneously or by stimulation depending on the set-up of the experiment. When the reaction nucleates at a point, diffusion brings bromide from the surrounding region into the nucleation point. This facilitates the resetting mechanism at the nucleation point while simultaneously allowing process (5) to dominate in the surrounding region, oxidizing the metal ion catalyst and perpetuating the front of lowered bromide concentration. In this manner, a circular wave will progress outward from the nucleation point. If the conditions are such that oscillations will continue, a second circular wave will begin at the nucleation point, and the end result will be many concentric rings in a target pattern. If the nucleation point does not continue to oscillate, a single wavefront will expand until it is annihilated at the edge of the container. Field and Noyes discovered experimentally that the wavefront velocity is given by: v = 0.04 cm sec -1 M -1 ([H + ][ BrO 3 ]) 1/ (Tyson, p85). Furthermore, there is a curvature dependence on the velocity. For very small target patterns, the curvature is relatively high and this slows the reaction s progress because diffusion from a point is less than diffusion from a line. This dependence is given by the eikonal equation: v* = v + DK (Epstein, p13), where D is the diffusion constant (on the order of 10-5 cm sec -1 ), K = ±r - 1 is the curvature (positive for a contracting circle and negative for an expanding circle), and v is the propagation velocity for the reaction at zero curvature. Using the above equations and typical numbers for hydrogen and bromate concentrations and D, we can estimate v and v*. Using the following values: [BrO 3 ]= 0.3 M [H + ] = 0.5 M D = 10-5 cm / sec results in: v = 6.577 mm / min v* = 5.977 mm / min (00 ìm expanding circle) v* = 6.5697 mm / min (4 cm expanding circle) v* = 6.5757 mm / min (4 cm contracting circle) v* = 7.177 mm / min (00 ìm contracting circle)

EXPERIMENT & METHODS Reagents Rather than use cerium as a catalyst and indicator, as Field, Körös, and Noyes did, we used ferroin, which has a more conspicuous color change between its two ionization states, from dark red to bright blue. The reactants were stock solutions of sodium bromate, potassium bromide, and malonic acid dissolved in sulfuric acid. The ferroin used was a 0.05 M stock solution in water. The initial concentrations in the experiment were: [BrO 3 ] = 0.3 M [Br ] = 0.05 M [MA] = 0.3 M [Ferroin] = 0.004 M [H + ] = 0.5 M The specific recipe used was ml of 0.9375 M NaBrO 3, 1.5 ml of 0.5 M KBr, ml of 0.9375 M malonic acid, and 1mL of 0.05 M ferroin. All stock solutions except for ferroin were made in 0.3 M sulfuric acid, in order to achieve the final acid concentration of 0.5 M. A 0.1% solution of Triton X-100 was also used in our experiment. Apparatus To capture video images of our reaction, our set-up involved a camera overlooking a light-box. The light-box contained a blue sheet of plastic that served as a filter which closely matched the transmission wavelength (blue) of ferriin, the reduced form of ferroin. The use of the light filter provided the greatest possible contrast between the two states in the video output. In order to initiate target patterns we used a 9-volt battery with silver leads which we taped to the light-box. A CCD video camera mounted above the dish outputted a black-and-white video signal to a CRT and to a computer IMAQ card. A framegrabber written in LabVIew captured images from the camera with a user-controlled framerate (usually 1 fps). A timestamp was also saved for each frame. Experiment The amount of solution used in our experiments was 6.5 ml in a 10 cm diameter polystyrene Petri dish, resulting in a depth of 0.8 mm. This was shallow enough that the system was essentially two-dimensional. In other words, the reaction progressed at the same rate on the top surface and bottom surface; there was no variation in the z-direction. A similar experimental setup has been described for carrying out lecture demonstrations of chemical pattern formation (Shakashiri). The reactants were combined in a beaker, first by adding bromide to bromate (producing bromine, turning the solution yellow), then adding malonic acid, which slowly eliminated the bromine. When the solution was clear again, a drop of 0.1% Triton

X-100 detergent was added to facilitate spreading of the solution in the dish by reducing surface tension. Finally, ferroin was added. The beaker was then swirled several times. With the initial concentrations listed above, the solution will typically oscillate once while being swirled, turning blue, then back to red. After this, the solution was poured into the Petri dish and swirled carefully so that it spread evenly throughout the dish. The positive lead (silver wire) of a 9V battery was touched to the center of the dish for 1- seconds and the negative lead to the outer edge. Although several past experimenters used this method, none of the literature explains why this works. However, it seems likely that the the silver is ionized and thus reduces bromate in the vicinity, starting off the reaction. Simply using a piece of silver wire to seed the reaction (with no battery) is almost as effective, reducing the bromate in the same manner, but without the added push of a voltage source. After the wires are removed, a small blue dot will appear in the center and propogate outward. With the initial concentrations listed above, oscillations will continue at the nucleation point, creating target patterns. One problem with the above recipe is that, because the solution is so close to the oscillatory regime, the reaction can spontaneously nucleate, especially in regions of inhomogeneity, for example along the edge of the dish. For this reason, circular trigger waves often develop along the edge and progress inward, annihilating the planned target patterns when the two wavefronts meet. One way to reduce the excitability of the solution is to reduce the malonic acid concentration. Diluting the malonic acid threefold (in 0.3 M sulfuric acid) resulted in a solution in which the outer edge was unlikely to spontaneously nucleate trigger waves. In this solution, target patterns were unlikely to develop as well, as the reaction will not continue without stimulus by the battery or wire. In this case, one sees only single trigger waves resulting from each stimulus. However, this mixture is ideal for studying spirals, which are self-propagating, and can continue even if the solution is in the exicitable (but not oscillatory) regime. Spiral patterns can be made by breaking a wavefront, either physically (with a pipette tip or a blast of air from a micropipette) or chemically (with a small drop (10 ìl or less) of dilute (0.1 M or less) KCl (chloride ion limits the reaction). When a wave front is broken, the two ends will begin to spiral in on themselves. RESULTS AND ANALYSIS Our main experimental goal was to determine whether the rate of expansion of a circle pattern was dependent on curvature. That is, is velocity a function of radius? In order to calculate velocity, we stimulated individual circle patterns and recorded them for analysis with a computer. For each of six circle patterns, we captured approximately 300 frames taken every 500 ms where each image was 640 by 480 pixels or about.4 by 1.6 cm. Each video was filtered for noise and had an average background image subtracted from it. The videos were then analyzed with an algorithm that found a best-fit circle for each image as determined by a maximum correlation between the circle and the image. The magnitude of the radius over time was then plotted to compare it to a straight line (the constant velocity hypothesis). Here are these plots.

By normalizing and shifting each set of data so that its best fit line was x=y, we can compare all plots to a constant velocity on the same graph.

There appears a definite trend of velocity increasing with radius and/or time. This agrees with the eikonal equation described earlier. We did not attempt to fit the eikonal equation to our data, but here is an example of the eikonal equation predicted velocity for parameters consistent with our experiments. We attempted two other experiments during the quarter. One was to model a BZ reaction wavefront using the Oregonator model. The equations we used were: where X t Y t = D X X + k3aby k r AXY + k Y = DY k3aby kaxy r A = [H+] B = [BrO 3 -] X = [HBrO ] Y=[Br-] D is a diffusion constant k is a reaction rate 5 ABX k 4 X

A and B were assumed to be constants. In order to solve these as ordinary differential equations we assumed constant velocity and were thus able to remove the partial derivative with respect to time. We used initial conditions given by (Field and Noyes) and numerically solved these equations in MATLAB. We found the following plots of X and Y. Although the direction of change and timescale are approximately correct for these reactants, the wavefront is not physically realistic since it involves negative concentrations. Finally, we attempted to fit a spiral to an image of one of our spiral patterns. The particular spiral equation we used was an Archimedes spiral which assumes that the increase in radius of curvature per turn is constant. Since the increase in the radius of curvature for circle patterns is approximately constant, we hypothesized this would be true for spirals as well. We fit one spiral pattern satisfactorily with the Archimedes spiral r = (.97mm)θ.

As far as future experiments, we would have liked to continue the modeling effort to obtain reasonable simulated behavior with an ODE or PDE model. This should be a realizable goal since there are multiple MATLAB implementations of BZ models available through a Google search. One property of the BZ reaction we never verified experimentally is that although the wavelength of individual target patterns may vary in a single solution, the velocity of all patterns should be a constant of the solution. This could be verified by stimulating multiple target patterns in a dish simultaneously. Similarly, spiral patterns have constant periods and velocities. Finally, it should be possible to observed 3-dimensionsal patterns such as scroll waves in thicker solutions. It was left unexplored what changes to the experimental set-up are necessary to form these patterns.

REFERENCES I.R. Epstein and J.A. Pojman. Introduction to Nonlinear Chemical Dynamics. Oscillations, Waves, Patterns and Chaos. Oxford UP. New York: 1998. Richard J. Field and Richard M. Noyes. Oscillations in Chemical Systems. V. Quantitative Explanation of Band Migration in the Belousov-Zhabotinskii Reaction. Journal of the American Chemical Society, 96:7:001-006, 1974. Stephen K. Scott. Oscillations, Waves and Chaos in Chemical Kinetics. Oxford UP. New York: 1994. Bassam Z. Shakhashiri. Chemical Demonstrations: A Handbook for Teachers of Chemistry. University of Wisconson Press. Madison, Wisconson: 1983. John J. Tyson. The Belousov-Zhabotinskii Reaction. Springer-Verlag. New York: 1976.