Uploaded by SUMIT KALE

Nonlinear Dynamical Behavior in the Photodecomposition of N-Bromo 1,4- Benzoquinoe-4-Imine

advertisement
Article
pubs.acs.org/JPCA
Nonlinear Dynamical Behavior in the Photodecomposition of
N‑Bromo-1,4-Benzoquinone-4-Imine
Jeffrey G. Bell, James R. Green,* and Jichang Wang*
Downloaded via INDIAN INST OF TECHNOLOGY GUWAHATI on March 26, 2019 at 15:20:15 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Department of Chemistry and Biochemistry, University of Windsor, Windsor, ON, N9B 3P4, Canada
ABSTRACT: Photodecomposition of N-bromo-1,4-benzoquinone-4-imine
in water with and without the presence of oxidants was investigated,
exhibiting nonlinear kinetic features such as autocatalytic excursion. An
acidic environment was found to be essential for the observed photodecomposition. Mechanistic studies through NMR and mass spectrometry
show the conversion of N-bromo-1,4-benzoquinone-4-imine into 1,4benzoquinone. When bromate was introduced to the studied system,
transient spontaneous oscillations were uncovered, forming a new
photocontrolled chemical oscillator. The system’s sensitivity to the intensity
of the illumination is great as decreases halt the oscillations from occurring.
Phase diagrams show that as the concentration of N-bromo-1,4benzoquinone-4-imine is increased, broader concentration ranges of both
bromate and sulfuric acid allow the system to exhibit spontaneous
oscillations. Electrospray-TOF mass spectroscopy and 13C NMR spectra
suggest that 3,4,4-tribromo-2-hydroxycyclohexa-2,5-dienone is a major product in the bromate system.
1. INTRODUCTION
The study of nonlinear dynamics in chemical systems has seen
substantial activity over the past few decades and remains a
topic of great attention.1−10 The autocatalytic properties of
bromate oscillators, first studied by Belousov, have allowed the
original parameters of the Belousov−Zhabotinsky reaction to
be expanded upon, in terms of catalyst, organic substrate, and
external forces.11−20 In 1978, Orbán and Körös published the
first evidence of oscillatory behavior in an uncatalyzed system
containing only H+, BrO3−, and gallic acid.21 In later work, they
published a list of 23 organic substrates which would be capable
of exhibiting nonlinear behavior, leading to a class of bromate
oscillators referred to as Uncatalyzed Bromate Oscillators
(UBOs).22 Recent exploration into the study of chemical
oscillators has led to much research on the effects of certain
external forces on nonlinear dynamics such as stirring rate,
temperature, and the influence of illumination on the oscillatory
phenomenon.23−36 This is partially driven by the need of
gaining insights into the behavior of less controllable nonlinear
systems in nature, which are frequently subjected to various
perturbations.
A variety of phenomena that do not exist in unperturbed
systems have been uncovered.26−28 Among various means to
affect reaction kinetics, light is arguably the most convenient
means to implement various forms of external forcing, such as
periodic or aperiodic spatiotemporal illumination.29−36 Photochemical oscillators can loosely be separated into two groups:
photosensitive or photocontrolled. A photocontrolled reaction
is dependent upon illumination in order for certain
intermediates to be created, whereas in photosensitive
reactions, the illumination simply gives an alternative pathway
for the formation of the intermediate. It has been shown that
© 2013 American Chemical Society
the bromate-4-aminophenol, bromate-benzoquinone, and
bromate-methyl-benzoquinone chemical reactions belong to
the subset of photocontrolled oscillators as evidenced by the
presence of spontaneous oscillations in an illuminated reaction
system and none in a nonilluminated system.36−38 The driving
force behind the nonlinear properties of the reaction lack as
complete an understanding as other more well-studied
oscillatory systems such as the BZ reaction.39,40
To further advance the development of photocontrolled
chemical oscillators, which are expected to provide better
model systems for understanding perturbed nonlinear dynamics, this study investigated the photoreaction kinetics of Nbromo-1,4-benzoquinone-4-imine (BBI) in aqueous solution
with and without the presence of other reactants. While there is
no reactivity in the absence of light, under illumination the
dissolution of BBI exhibits an autocatalytic feature, in which
there is a sharp Pt potential increase in the time series. When
sodium bromate is introduced to the BBI photodecomposition,
spontaneous oscillations occur, forming a new photocontrolled
oscillator. The results also provide new insights into the
reaction mechanism of the bromate-4-aminophenol oscillations.
2. EXPERIMENTAL SECTION
All reactions were carried out in a thermal-jacketed 50 mL glass
beaker (ChemGlass). The reaction temperature was held
constant at 25.0 ± 0.1 °C by a circulating water bath (Thermo
NesLab RTE 7). The volume for every reaction was held
Received: April 10, 2013
Revised: May 17, 2013
Published: May 20, 2013
4545
dx.doi.org/10.1021/jp403546g | J. Phys. Chem. A 2013, 117, 4545−4550
The Journal of Physical Chemistry A
Article
feature of autocatalytic reactions. The solution became a dark
orange color as the reaction proceeds and all of the N-bromo1,4-benzoquinone-4-imine was photodecomposed.
1
H NMR measurement, taken after all of the BBI had
dissolved, shows the presence of 1,4-benzoquinone at a spectral
resonance of δ = 6.80, indicating that the BBI- photoreaction
led to the production of 1,4-benzoquinone. This conclusion is
further supported by characterization with Electron Impact
(EI) mass spectrometry (m/e = 108). The formation of 1,4benzoquinone leads us to speculate that introducing bromate
into the photodecomposition of BBI may form a new
photocontrolled chemical oscillator. Figure 2a presents a time
constant at 30.0 mL. The reaction solution was stirred with a
magnetic stirring bar driven by a magnetic stirrer (Fisher
Isotemp), in order to ensure homogeneity. Reactions were
monitored with a platinum electrode coupled with a Hg|
Hg2SO4|K2SO4 reference electrode (Radiometer Analytical,
XR200 and M231 Pt-9). A Teflon cap was used on top of the
thermal-jacketed beaker in order to hold the electrodes. All
measurements were recorded through a pH/potential meter
(Radiometer PHM220) connected to a computer through a
PowerLab/4SP data logger. The source of illumination was a
150 W halogen light bulb (Fisher Scientific, model DLS100HD) and was placed at a distance of 55 mm away from the
reaction beaker.
Stock solutions of analytical grade sodium bromate (NaBrO3,
Aldrich, 99%), 0.1 M, sodium bromide (NaBr, Aldrich, 99%)
0.1 M, and sulfuric acid (Aldrich, 95−98%), 6.0 M, were
prepared with double-distilled water. N-bromo-1,4-benzoquinone-4-imine was prepared in our lab through the reaction
between sodium bromate and 4-aminophenol in 1.7 M H2SO4
solution, in which the N-bromo-1,4-benzoquinone-4-imine was
collected through filtration techniques and dried thoroughly.
The purity of the product was analyzed with 1H NMR
spectroscopy and thin layer chromatography. To avoid the
decomposition of N-bromo-1,4-benzoquinone-4-imine, it was
prepared immediately prior to its use in the reaction.
3. RESULTS AND DISCUSSION
BBI has very low solubility in water. For instance, when 3.0 ×
10−4 mole of BBI is mixed with 30 mL of water, there is no
visible decrease in the amount of solid after 24 h, while the
water shows a very light yellow color. When the BBI and water
mixture is exposed to light, the solution turns into yellow
rapidly, indicating photoassisted dissolution/reaction. Figure 1
Figure 2. (a) Time series of the uncatalyzed bromate−BBI reaction;
(b) time series illustrating the photocontrolled nature of the
uncatalyzed bromate−BBI reaction. Reaction concentrations are
[NaBrO3] = 0.04 M, [H2SO4] = 1.7 M and [BBI] = 9.41 × 10−3 M.
series under concentrations of bromate (0.040 M), BBI (9.41 ×
10−3 M), and sulfuric acid (1.7 M). The solution immediately
becomes yellow upon the addition of the bromobenzoquinone
imine and grows viscous after bromate is added and the
photoreaction proceeds. The sharp spike in potential marks the
complete dissolution of the BBI, at which point the solution
remains yellow but transparent. The potential reaches a
maximum and slowly decreases until spontaneous oscillations
occur. The oscillations continue for several cycles, which stop at
the low potential, at which point the color of the solution
remains yellow.
Figure 2b presents time series proving the photocontrolled
nature of this new bromate oscillator. It shows that when
illumination is removed from the reaction system while
oscillations are occurring, they stop abruptly and the potential
reaches a stable nonoscillatory state. However, when
illumination is reintroduced to the system, oscillations reemerge. Notably, removing the illumination at opposite phases
of the oscillation caused different responses of the system,
where turning off the illumination at the bottom of an
oscillation will cause the system to return to the high potential
immediately, followed by a decrease to the stable potential. The
magnitude of the oscillation is also influenced by the intensity
of the illumination involved; in fact, there is a threshold
intensity that must be reached in order for oscillations to occur.
We would like to note that both the oscillation waveform and
the long induction time are reproducible, despite their great
sensitivity to illumination.
Figure 1. Time series showing the photodecomposition of 9.41 × 10−3
M N-bromo-1,4-benzoquinone-4-imine in (a) neutral solution and (b)
1.7 M H2SO4.
presents two time series illustrating the critical influence of an
acidic environment on the photoreaction of BBI. In Figure 1a,
when the photodecomposition of BBI was studied in a neutral
solution, the Pt potential gradually increased over the first 90
min and then remained at an elevated level throughout the
reaction. The reaction solution became light orange with very
little of the BBI dissolving. In an acidic environment the
photodecomposition of BBI began to occur within 15 min
where the Pt potential can be seen to drastically decrease from
its initial high level to a very low level, exhibiting the kinetic
4546
dx.doi.org/10.1021/jp403546g | J. Phys. Chem. A 2013, 117, 4545−4550
The Journal of Physical Chemistry A
Article
Figure 3 illustrates the effect of changing the concentration of
bromate while holding the amount of N-bromo-1,4-benzoqu-
Figure 4. Time series of the uncatalyzed bromate−BBI reaction
carried out at different BBI concentrations: (a) 3.14 × 10−3 M, (b)
6.28 × 10−3 M, (c) 9.41 × 10−3 M, (d) 1.25 × 10−2 M, and (e) 1.88 ×
10−2 M. Other reaction conditions were [NaBrO3] = 0.04 M and
[H2SO4] = 1.7 M.
Figure 3. Time series of the uncatalyzed bromate−BBI reaction
carried out at different bromate concentrations: (a) 0.02 M, (b) 0.03
M, (c) 0.04 M, (d) 0.05M. Other reaction conditions were [H2SO4] =
1.7 M and [BBI] = 9.41 × 10−3 M.
ninone-4-imine (9.41 × 10−3 M), and sulfuric acid (1.7 M)
constant. It can be seen that the concentration of bromate is a
critical parameter influencing the nonlinear reaction kinetics. At
low concentrations of bromate, not enough oxidant is present
to oxidize the BBI resulting in a Pt potential which decreases
and is unable to spike, and the solution becomes a murky
yellow/orange color. An increased concentration of bromate
shows a spike in potential, and an induction time followed by a
drastic drop in potential with no oscillations. As the
concentration of bromate is increased further, to 0.030 M,
five oscillations are present. When the concentration is
increased to 0.040 M bromate, the number of oscillations
increased; however, the amplitude of the oscillations decreased.
This decrease in amplitude became obvious at 0.05 M bromate.
The number of oscillations can also be seen to decrease as the
concentration of bromate is increased and at 0.060 M, the
potential remains high after the potential spike, and the
solution becomes transparent with a yellow tint.
Figure 4 is a time series showing the effect of changing the
concentration of N-bromo-1,4-benzoquinone-4-imine while
holding the concentration of bromate (0.040 M) and sulfuric
acid (1.7 M) constant. As is apparent, there is no upper
boundary to the concentration of BBI, as adding 0.105 g of BBI
to the 30.0 mL reaction vessel (0.0188 M) gives full
dissolution; however adding more BBI effectively saturates
the solution leaving a film of undissolved powder in the system.
Spontaneous oscillations are still present in the saturated
system, but the exact concentration of BBI is unknown. For this
reason the phase diagrams in Figure 5 show no upper limit in
the concentration of the N-bromo-1,4-benzoquinone-4-imine
plane. Below the concentration of BBI required for oscillations
the solution does not become murky at the onset of the
reaction and as the reactant dissolves the solution becomes
transparent with a slight yellow color. The potential undergoes
a gradual decrease until the potential stabilizes. As the
Figure 5. Phase diagram illustrating regions of oscillatory phenomenon in the concentration of (a) BBI and bromate concentration
plane, and (b) BBI and sulfuric acid concentration plane.
concentration of BBI is increased, the amount of time for the
potential spike to occur increases, which is reasonable due to an
increased amount of reactant; the induction time also increases.
Figure 5a is a phase diagram in the concentration of BBI and
bromate concentration plane, where the filled triangle (▲)
indicates the conditions under which the system displays
spontaneous oscillations. Figure 6 illustrates the effect of
changing the concentration of sulfuric acid while holding the
amount of BBI (9.41 × 10−3 M), and bromate (0.04 M)
4547
dx.doi.org/10.1021/jp403546g | J. Phys. Chem. A 2013, 117, 4545−4550
The Journal of Physical Chemistry A
Article
Figure 6. Time series of the uncatalyzed bromate−BBI reaction
carried out at different sulfuric acid concentrations: (a) 0.4 M, (b) 1.0
M, (c) 1.7 M, and (d) 2.2 M. Other reaction conditions were
[NaBrO3] = 0.04 M and [BBI] = 9.41 × 10−3 M.
constant. Immediately it is obvious that the range of H+ which
allows for oscillations to occur is much larger than the range for
BrO3−; this is because the H2SO4 serves simply as a source of
H+ needed to ensure that the pH is kept low enough for the
autocatalytic cycle to continue. Below the concentration of
sulfuric acid which produces oscillations a sharp potential
decrease is observed, but unlike being below the lower limit in
bromate concentration (where the sharp decline occurs
quickly), the potential remains stable for over 90 min. The
most notable difference in the time series in Figures 4 and 6 is
apparent at the extreme cases. High acid concentrations have
their spike early, whereas high bromate concentrations undergo
the potential spike after remaining at a stable potential for a
prolonged period. Figure 5b is a phase diagram in the amount
of BBI and sulfuric acid concentration plane, where the filled
triangle (▲) indicates the conditions under which the system
exhibits spontaneous oscillations.
A study of the compounds involved in the nonlinear behavior
of the oscillatory system was also undertaken using NMR
spectroscopy. Figure 7 presents three 1H NMR spectra
recorded in CDCl3 solvent taken at key intervals as the
reaction progressed. Figure 7a spectrum was taken after all of
the BBI had dissolved, but prior to the occurrence of
spontaneous oscillations. The spectral resonance of δ = 6.80
indicates the presence of 1,4-benzoquinone and was confirmed
by the δ = 136.6 and 187.3 resonances in the 13C NMR
spectrum of a more concentrated sample. Figure 7b shows the
1
H NMR spectrum of the composition of the reaction solution
during the oscillatory window. As can be seen the component
giving resonances at δ = 6.97 and 7.02 with a distinct AB
coupling pattern (AB quartet, J = 10.5 Hz), which was only
present in minute quantities before oscillations occurred has
increased at the expense of 1,4-benzoquinone. Finally in Figure
7c, after the oscillatory window has ceased, the component at δ
= 6.97 and 7.02 has grown substantially in comparison to the
benzoquinone.
Because of the absence of bromobenzoquinone, a known
product of other bromate-benzoquinone oscillators, which
Figure 7. 1H NMR spectra at three unique points in the reaction: (a)
after potential spike and before oscillations; (b) during the oscillatory
window; (c) after the oscillations have occurred.
develops as the reaction proceeds, further spectroscopic
experiments were conducted to establish a reasonable candidate
for the identity of this new compound. This compound is of
limited thermal stability, which has precluded chromatographic
purification or crystallization. Nevertheless, the 13C NMR
spectrum of more highly concentrated samples displayed
resonances at δ = 69.0, 94.6, 135.3, 137.7, 182.3, and 189.2.
The HMBC spectrum (Figure 8) of the same sample showed
correlations of both the δ = 69.0 and 189.2 13C resonances with
the δ = 7.02 1H resonance, as well as the δ = 94.6 and 182.3 13C
resonances with the δ = 6.97 1H resonance. Finally,
electrospray-TOF mass spectroscopy (negative ion mode)
gives an intense set of ions at m/e = 345, 347, 349, and 351,
with a characteristic isotope pattern of a tribrominated
compound. Given the propensity for negative ion electrospray
MS to give prominent (M + H)− ions,41 this is strongly
suggestive of a compound with a C6H3Br3O2 molecular
formula. In light of this evidence, it is our working hypothesis
that this new compound is 3,4,4-tribromo-2-hydroxycyclohexa2,5-dienone (Figure 9).
■
CONCLUSION
This study presents a new photochemical oscillator in which
the presence of illumination was imperative in order for the
oscillatory phenomenon to be observed. The extreme
sensitivity of the system to the intensity of the illumination
was apparent through the manipulation of the oscillatory
window by both abruptly removing and reintroducing the
4548
dx.doi.org/10.1021/jp403546g | J. Phys. Chem. A 2013, 117, 4545−4550
The Journal of Physical Chemistry A
Article
Figure 8. Heteronuclear multiple bond correlation (HMBC) spectrum of sample taken after oscillations had occurred.
Figure 9. Electrospray-TOF mass spectroscopy spectrum and proposed reaction intermediate 3,4,4-tribromo-2-hydroxycyclohexa-2,5-dienone.
illumination, or by decreasing and then increasing the intensity
of the illumination. The system was also found to be sensitive
to the concentrations of bromate, sulfuric acid, and BBI.
Changing the concentrations of the reactants not only had an
effect on the number of oscillations but also the amplitude of
oscillations as well as the induction time before the
spontaneous oscillations occurred. Spectroscopic measurements indicate that N-bromo-1,4-benzoquinone-4-imine is
transformed into 1,4-benzoquinone preceding the onset of
spontaneous oscillations, leading to a major product assigned as
3,4,4-tribromo-2-hydroxycyclohexa-2,5-dienone.
■
AUTHOR INFORMATION
Corresponding Author
*Email: jgreen@uwindsor.ca; jwang@uwindsor.ca. Fax: 1-5199737089.
Notes
The authors declare no competing financial interest.
4549
dx.doi.org/10.1021/jp403546g | J. Phys. Chem. A 2013, 117, 4545−4550
The Journal of Physical Chemistry A
■
Article
(23) Wang, J. Light-Induced Pattern Formation in the Excitable
Belousov−Zhabotinsky Medium. Chem. Phys. Lett. 2001, 339, 357−
361.
(24) Steinbock, O.; Zykov, V. S.; Müller, S. C. Control of SpiralWave Dynamics in Active Media by Periodic Modulation of
Excitability. Nature 1993, 366, 322−324.
(25) Hemming, C.; Kapral, R. Resonantly Forced Inhomogeneous
Reaction-Diffusion Systems. Chaos 2000, 10, 720.
(26) Pelle, K.; Wittmann, M.; Nosziticzius, Z. Perturbation of the
Oscillatory BZ Reaction with Methanol and Ethylene Glycol:
Experiments and Model Calculations. J. Phys. Chem. A 2003, 107,
2039−2047.
(27) Sriram, K.; Gopinathan, M. S. Effects of Delayed Linear
Electrical Perturbation of the Belousov−Zhabotinsky Reaction: A Case
of Complex Mixed Mode Oscillations in a Batch Reactor. React. Kinet.
Catal. Lett. 2003, 79, 341−349.
(28) Rossi, F.; Lombardo, R.; Sciascia, L.; Sbriziolo, C.; Liveri, M. L.
T. Spatio-Temporal Perturbation of the Dynamics of the Ferroin
Catalyzed Belousov−Zhabotinsky Reaction in a Batch Reactor Caused
by Sodium Dodecyl Sulfate Micelles. J. Phys. Chem. B 2008, 112,
7244−7250.
(29) Borderie, B.; Lavabre, D.; Micheau, J. C.; Laplante, J. P.
Nonlinear Dynamics, Multiple Steady-States, and Oscillations in
Photochemistry. J. Phys. Chem. 1992, 96, 2953−2961.
(30) Kaminaga, A.; Hanazaki, I. Photo-Induced Excitability in the
Tris-(bipyridyl)Ruthenium (II)-Catalyzed Minimal Bromate Oscillator. Chem. Phys. Lett. 1997, 278, 16−20.
(31) Huh, D. S.; Kim, H. S.; Kang, J. K.; Kim, Y. J.; Kim, D. H.; Park,
S. H.; Yadav, K.; Wang, J. Photo-Controlled Bifurcations in the 1,4Cyclohexanedione−Hydroquinone−Bromate−Ferroin Reaction.
Chem. Phys. Lett. 2003, 378, 78−84.
(32) Vanag, V. K.; Epstein, I. R. Design and Control of Patterns in
Reaction-Diffusion Systems. Chaos 2008, 18, 026107.
(33) Sakurai, T.; Mihaliuk, E.; Chirila, F.; Showalter, K. Design and
Control of Wave Propagation Patterns in Excitable Media. Science
2002, 296, 2009−2012.
(34) Wang, J.; Yadav, K.; Zhao, B.; Gao, Q. Y.; Huh, D. S.
Photocontrolled Oscillatory Dynamics in the Bromate−1,4-Cyclohexanedione Reaction. J. Chem. Phys. 2004, 121, 10138−10144.
(35) Li, J.; Wang, J. Complex Dynamical Behavior in the Highly
Photosensitive Cerium-Bromate−1,4-Benzoquinone Reaction. J. Phys.
Chem. A 2012, 116, 8130−8137.
(36) Zhao, B.; Wang, J. Photo-Mediated Bromate−1,4-Benzoquinone
Reaction: A Novel Photochemical Oscillator. Chem. Phys. Lett. 2006,
430, 41−44.
(37) Amemiya, T.; Wang, J. A Chemical Oscillator Based on the
Photoreduction of 2-Methyl-1,4-Benzoquinone. J. Phys. Chem. A 2010,
114, 13347−13352.
(38) Harati, M.; Amiralaei, S.; Green, J. R.; Wang, J. Chemical
Oscillations in the 4-Aminophenol−Bromate-Photoreaction. Chem.
Phys. Lett. 2007, 439, 337−341.
(39) Field, R. J.; Körös, E.; Noyes, R. M. Oscillations in Chemical
Systems. 2. Thorough Analysis of Temporal Oscillation in Bromate−
Cerium−Malonic Acid System. J. Am. Chem. Soc. 1972, 94, 8649−
8664.
(40) Gyorgyi, L.; Field, R. J. A 3-Variable Model of Deterministic
Chaos in the Belousov−Zhabotinsky Reaction. Nature 1992, 355,
808−810.
(41) Zhao, Y.; Qin, F.; Boyd, J. M.; Anichina, J.; Li, X. F.
Characterization and Determination of Chloro- and BromoBenzoquinones as New Chlorination Disinfections Byproducts in
Drinking Water. Anal. Chem. 2010, 82, 4599−4605.
ACKNOWLEDGMENTS
This research was supported by Natural Sciences and
Engineering Research Council of Canada (NSERC) and
Canada Foundation for Innovation (CFI).
■
REFERENCES
(1) Field, R. J., Burger, M., Eds. Oscillations and Travelling Waves in
Chemical Systems; John Wiley & Sons: New York, 1985.
(2) Epstein, I. R., Pojman, J. A. An Introduction to Nonlinear Chemical
Dynamics: Oscillations, Waves, Patterns and Chaos; Oxford University
Press: Oxford, 1998.
(3) Gray, P., Scott, S. K., Chemical Oscillations and Instabilities: Nonlinear Chemical Kinetics; Clarendon Press: Oxford, 1994.
(4) Scott, S. K. Oscillations, Waves, and Chaos in Chemical Kinetics;
Oxford University Press: Oxford, 1994.
(5) Sagués, F.; Epstein, I. R. Nonlinear Chemical Dynamics. Dalton
Trans. 2003, 1201−1217.
(6) Szabo, E.; Adamčiková, L.; Ševčik, P. Are Uncatalyzed Bromate
Oscillators Truly Gas-Free? J. Phys. Chem. A 2011, 115, 6518−6524.
(7) Yamaguchi, T.; Kuhnert, L.; Nagy-Ungrarai, Zs.; Müller, S. C.;
Hess, B. Gel Systems for the Belousov−Zhabotinskii Reaction. J. Phys.
Chem. 1991, 95, 5831−5837.
(8) Nosziticzius, Z. Non-Br−-Controlled Oscillations in the
Belousov−Zhabotinskii Reaction of Malonic Acid. J. Am. Chem. Soc.
1979, 101, 3660−3663.
(9) Gao, Q.; Wang, J. pH Oscillations and Complex Reaction
Dynamics in the Non-buffered Chlorite−Thiourea Reaction. Chem.
Phys. Lett. 2004, 391, 349−353.
(10) Noyes, R. M.; Furrow, S. D. The Oscillatory Brigg−Rauscher
Reaction. 3. A Skeleton Mechanism for Oscillations. J. Am. Chem. Soc.
1982, 104, 45−48.
(11) Szalai, I.; Kurin-Csörgei, K.; Epstein, I. R.; Orbán, M. Dynamics
and Mechanism of Bromate Oscillators with 1,4-Cyclohexanedione. J.
Phys. Chem. A 2003, 107, 10074−10081.
(12) Dutt, A. K.; Müller, S. C. Bistability in an Uncatalyzed Bromate
Oscillator in a Continuously Fed Stirred Tank Reactor. J. Chem. Phys.
1996, 104, 583−586.
(13) Burger, M.; Körös, E. Conditions for the Onset of Chemical
Oscillation. J. Phys. Chem. 1980, 84, 496−500.
(14) Zhabotinsky, A. M.; Buchholtz, F.; Kiyatkin, A. B.; Epstein, I. R.
Oscillations and Waves in Metal-Ion-Catalyzed Bromate Oscillating
Reactions in Highly Oxidized States. J. Phys. Chem. 1993, 97, 7578−
7584.
(15) Ganaie, N. B.; Nath, M. A.; Peerzada, G. M. Kinetic Studies on
Resorcinol−Bromate−Manganese (II) Ion Oscillator with and without
Additives. Kinet. Catal. 2010, 51, 25−30.
(16) Adamčiková, L.; Bielik, R.; Ševčik, P. Pattern Formation in
Uncatalyzed Bromate Oscillator with Pyrogallol. Collect. Czech. Chem.
Commun. 2004, 69, 1553−1565.
(17) Kádár, S.; Amemiya, T.; Showalter, K. Reaction Mechanism for
Light Sensitivity of the Ru(bpy)32+-Catalyzed Belousov−Zhabotinsky
Reaction. J. Phys. Chem. A 1997, 101, 8200−8206.
(18) Ramaswamy, R.; Lalitha, P. V. New Substrates in the FerroinCatalyzed Bromate System in Aqueous-Organic Mixed-Media. Int. J.
Chem. Kinet. 1993, 25, 457−468.
(19) Srivastava, P. K.; Mori, Y.; Hanazaki, I. Photoinhibition of
Chemical Oscillation in the Ru(bpy)32+-Catalyzed Belousov−Zhabotinskii Reaction. Chem. Phys. Lett. 1992, 190, 279−284.
(20) Harati, M.; Green, J. R.; Cooper, B. F. T.; Wang, J. Gas
Production in the Bromate-Pyrocatechol Oscillator. Chem. Phys. Lett.
2009, 113, 6548−6551.
(21) Orbán, M.; Körös, E. Uncatalyzed Oscillatory ChemicalReactions. Nature 1978, 273, 371−372.
(22) Sridevi, V.; Ramaswamy, R. Potential Response in the
Uncatalyzed Bromate Oscillatory System. React. Kinet. Catal. Lett.
1998, 65, 93−100.
4550
dx.doi.org/10.1021/jp403546g | J. Phys. Chem. A 2013, 117, 4545−4550
Download