Co-reporter:Abhishek Saxena, Arpita Bhattacharya, Satish Kumar, Irving R. Epstein, Rachana Sahney
Journal of Colloid and Interface Science 2017 Volume 490(Volume 490) pp:
Publication Date(Web):15 March 2017
DOI:10.1016/j.jcis.2016.11.030
Alginate microparticles and nanoparticles crosslinked with Ca+2 ions are frequently employed in biomedical applications. Here we use microemulsion polymerization to prepare alginate nanoparticles (nanogels) using different crosslinking ions (Ca+2, Sr+2, Ba+2) to encapsulate a model protein, urease enzyme (jackbeans). With alginate concentrations of 0.2 wt% in the aqueous phase, emulsion droplets showed good stability and narrow, monomodal distributions with radii ∼65 ± 10 nm. The size of the nanogel varies with the crosslinking cation and its affinity for the mannuronate and guluronate units in the linear alginate chain. The nanogels were further characterized using dynamic light scattering, scanning electron microscopy, energy dispersive X-ray spectrometry and zeta potential. This work demonstrates the potential application of Ba-alginate as an alternative matrix for nano-encapsulation of proteins and its use for biomedical applications.Download high-res image (82KB)Download full-size image
Co-reporter:Viktor Horváth, Irving R. Epstein, and Kenneth Kustin
The Journal of Physical Chemistry A 2016 Volume 120(Issue 12) pp:1951-1960
Publication Date(Web):March 7, 2016
DOI:10.1021/acs.jpca.5b11152
Existing models of the ferrocyanide–iodate–sulfite (FIS) reaction seek to replicate the oscillatory pH behavior that occurs in open systems. These models exhibit significant differences in the amplitudes and waveforms of the concentration oscillations of such intermediates as I–, I3–, and Fe(CN)63– under identical conditions and do not include several experimentally found intermediates. Here we report measurements of sulfite concentrations during an oscillatory cycle. Knowing the correct concentration of sulfite over the course of a period is important because sulfite is the main component that determines the buffer capacity, the pH extrema, and the amount of oxidizer (iodate) required for the transition to low pH. On the basis of this new result and recent experimental findings on the rate laws and intermediates of component processes taken from the literature, we propose a mass action kinetics model that attempts to faithfully represent the chemistry of the FIS reaction. This new comprehensive mechanism reproduces the pH oscillations and the periodic behavior in [Fe(CN)63–], [I3–], [I–], and [SO32–]T with characteristics similar to those seen in experiments in both CSTR and semibatch arrangements. The parameter ranges at which stationary and oscillatory behavior is exhibited also show good agreement with those of the experiments.
Co-reporter:Miklós Orbán, Krisztina Kurin-Csörgei, and Irving R. Epstein
Accounts of Chemical Research 2015 Volume 48(Issue 3) pp:593
Publication Date(Web):February 23, 2015
DOI:10.1021/ar5004237
The hydrogen ion is arguably the most ubiquitous and important species in chemistry. It also plays a key role in nearly every biological process. In this Account, we discuss systems whose behavior is governed by oscillations in the concentration of hydrogen ion. The first chemical oscillators driven by changes in pH were developed a quarter century ago. Since then, about two dozen new pH oscillators, systems in which the periodic variation in pH is not just an indicator but an essential prerequisite of the oscillatory behavior, have been discovered. Mechanistic understanding of their behavior has grown, and new ideas for their practical application have been proposed and, in some cases, tested.Here we present a catalog of the known pH oscillators, divide them into mechanistically based categories based on whether they involve a single oxidant and reductant or an oxidant and a pair of reductants, and describe general mechanisms for these two major classes of systems. We also describe in detail the chemistry of one example from each class, hydrogen peroxide–sulfide and ferricyanide–iodate–sulfite. Finally, we consider actual and potential applications. These include using pH oscillators to induce oscillation in species that would otherwise be nonoscillatory, creating novel spatial patterns, generating periodic transitions between vesicle and micelle states, stimulating switching between folded and random coil states of DNA, building molecular motors, and designing pulsating drug delivery systems. We point out the importance for future applications of finding a batch pH oscillator, one that oscillates in a closed system for an extended period of time, and comment on the progress that has been made toward that goal.
Co-reporter:Viktor Horvath, Daniel J. Kutner, John T. Chavis III and Irving R. Epstein
Physical Chemistry Chemical Physics 2015 vol. 17(Issue 6) pp:4664-4676
Publication Date(Web):13 Jan 2015
DOI:10.1039/C4CP05416D
Coupled chemical oscillators are usually studied with symmetric coupling, either between identical oscillators or between oscillators whose frequencies differ. Asymmetric connectivity is important in neuroscience, where synaptic strength inequality in neural networks commonly occurs. While the properties of the individual oscillators in some coupled chemical systems may be readily changed, enforcing inequality between the connection strengths in a reciprocal coupling is more challenging. We recently demonstrated a novel way of coupling chemical oscillators, which allows for manipulation of individual connection strengths. Here we study two identical, pulse-coupled Belousov–Zhabotinsky (BZ) oscillators with unequal connection strengths. When the pulse perturbations contain KBr (inhibitor), this system exhibits simple out-of-phase and complex oscillations, oscillatory-suppressed states as well as temporally periodic patterns (N:M) in which the two oscillators exhibit different numbers of peaks per cycle. The N:M patterns emerge due to the long-term effect of the inhibitory pulse-perturbations, a feature that has not been considered in earlier works. Time delay was previously shown to have a profound effect on the system’s behaviour when pulse coupling was inhibitory and the coupling strengths were equal. When the coupling is asymmetric, however, delay produces no qualitative change in behaviour, though the 1:2 temporal pattern becomes more robust. Asymmetry in instantaneous excitatory coupling via AgNO3 injection produces a previously unseen temporal pattern (1:N patterns starting with a double peak) with time delay and high [AgNO3]. Numerical simulations of the behaviour agree well with theoretical predictions in asymmetrical pulse-coupled systems.
Co-reporter:Ye Zhang ; Ning Zhou ; Ning Li ; Megan Sun ; Dongshin Kim ; Seth Fraden ; Irving R. Epstein ;Bing Xu
Journal of the American Chemical Society 2014 Volume 136(Issue 20) pp:7341-7347
Publication Date(Web):April 21, 2014
DOI:10.1021/ja503665t
While living systems have developed highly efficient ways to convert chemical energy (e.g., ATP hydrolysis) to mechanical motion (e.g., movement of muscle), it remains a challenge to build muscle-like biomimetic systems to generate mechanical force directly from chemical reactions. Here we show that a continuous flow of reactant solution leads to by far the largest volume change to date in autonomous active gels driven by the Belousov–Zhabotinsky reaction. These results demonstrate that microfluidics offers a useful and facile experimental approach to optimize the conditions (e.g., fabrication methods, counterions, flow rates, concentrations of reagents) for chemomechanical transduction in active materials. This work thus provides much needed insights and methods for the development of chemomechanically active systems based on combining soft materials and microfluidic systems.
Co-reporter:Mieczysław Sajewicz, Miloš Dolnik, Teresa Kowalska and Irving R. Epstein
RSC Advances 2014 vol. 4(Issue 14) pp:7330-7339
Publication Date(Web):09 Jan 2014
DOI:10.1039/C3RA46921B
We employ high-performance liquid chromatography with evaporative light scattering and mass spectrometric detection (HPLC/ELSD and LC/MS) to monitor the dynamic behavior of L-Pro, L-Hyp, and L-Pro–L-Hyp in 70% aqueous methanol. The individual amino acid solutions show evidence of oscillatory oligomerization. In the binary solution, the behavior is controlled by the dynamics of L-Pro oligomerization. A simple model involving oligomerization, formation of catalytic oligomer aggregates and cross-catalysis provides qualitative insight into the process.
Co-reporter:Xingjie Lu, Lin Ren, Qingyu Gao, Yuemin Zhao, Shaorong Wang, Jiaping Yang and Irving R. Epstein
Chemical Communications 2013 vol. 49(Issue 70) pp:7690-7692
Publication Date(Web):17 Jul 2013
DOI:10.1039/C3CC44480E
A photosensitive self-oscillating gel that incorporates the Belousov–Zhabotinsky reaction can undergo rhythmic mechanical oscillations. We exploit the dependence of the oscillation frequency on light intensity to generate both photophobic and phototropic movement of the gel under differential illumination. Our findings may be used in designing intelligent sensors that can execute biomimetic behaviours.
Co-reporter:Dr. Ye Zhang;Ning Zhou;Sathish Akella;Yi Kuang;Dr. Dongshin Kim;Alyssa Schwartz;Marc Bezpalko;Dr. Bruce M. Foxman;Dr. Seth Fraden;Dr. Irving R. Epstein;Dr. Bing Xu
Angewandte Chemie International Edition 2013 Volume 52( Issue 44) pp:11494-11498
Publication Date(Web):
DOI:10.1002/anie.201304437
Co-reporter:Dr. Ye Zhang;Ning Zhou;Sathish Akella;Yi Kuang;Dr. Dongshin Kim;Alyssa Schwartz;Marc Bezpalko;Dr. Bruce M. Foxman;Dr. Seth Fraden;Dr. Irving R. Epstein;Dr. Bing Xu
Angewandte Chemie 2013 Volume 125( Issue 44) pp:11708-11712
Publication Date(Web):
DOI:10.1002/ange.201304437
Co-reporter:Ye Zhang, Rong Zhou, Junfeng Shi, Ning Zhou, Irving R. Epstein, and Bing Xu
The Journal of Physical Chemistry B 2013 Volume 117(Issue 21) pp:6566-6573
Publication Date(Web):May 9, 2013
DOI:10.1021/jp401353e
We study the use of post-self-assembly cross-linking to combine molecular nanofibers of hydrogelators with copolymers to generate oscillatory materials using the Belousov–Zhabotinsky reaction. The formation of nanofibers from designed hydrogelators provides multiple polymerizable sites for copolymerizing with N-isopropylacrylamide and for attaching a catalytic ruthenium bipyridine complex on the copolymer. The combination of supramolecular self-assembly with copolymerization offers a versatile and facile approach for generating soft materials that have large pores in the gel network and robust mechanical integrity. These larger pores facilitate the diffusion of the reactants and accelerate the chemical oscillation by about a factor of 4 relative to a poly(NIPAAm-Ru) gel that contains no molecular nanofibers.
Co-reporter:Xing-Jie Lu, Lin Ren, Qing-Yu Gao, Ying-Ying Yang, Yue-Min Zhao, Jun Huang, Xiao-Li Lv, and Irving R. Epstein
The Journal of Physical Chemistry Letters 2013 Volume 4(Issue 22) pp:3891-3896
Publication Date(Web):November 1, 2013
DOI:10.1021/jz402117m
Patterns containing multiple length scales arise in a variety of natural systems such as lateral veins in leaves, fingerprints, wrinkled skin, and dendritic crystals. Here we observe period-doubling and bursting instabilities in the spatial extent of wave propagation in a gel-filled capillary tube open at one end and containing the Belousov–Zhabotinsky (BZ) reaction–diffusion system. We analyze the relationship between the multiple propagation distances of pulse waves and the local kinetics of the reaction–diffusion system. Simulations with a five-variable Oregonator model qualitatively mimic the multiple length scale patterns of pulse propagation observed in our experiments, suggesting that the study of these phenomena in reaction–diffusion systems may be helpful in understanding complex multiple length scale dynamical behaviors in nature.Keywords: Belousov−Zhabotinsky reaction; multiscale dissipative structures; Oregonator model; pulse waves; reaction-diffusion;
Co-reporter:Irving R. Epstein, Vladimir K. Vanag, Anna C. Balazs, Olga Kuksenok, Pratyush Dayal, and Amitabh Bhattacharya
Accounts of Chemical Research 2012 Volume 45(Issue 12) pp:2160
Publication Date(Web):December 28, 2011
DOI:10.1021/ar200251j
Evolution is a characteristic feature of living systems, and many fundamental processes in life, including the cell cycle, take place in a periodic fashion. From a chemistry perspective, these repeating phenomena suggest the question of whether reactions in which concentrations oscillate could provide a basis and/or useful models for the behavior of organisms, and perhaps even their ability to evolve.In this Account, we examine several aspects of the behavior of the prototype oscillating chemical reaction, the Belousov–Zhabotinsky (BZ) system, carried out in microemulsions, arrays of micrometer-sized aqueous droplets suspended in oil, or hydrogels. Each of these environments contains elements of the compartmentalization that likely played a role in the development of the first living cells, and within them we observe behaviors not found in the BZ reaction in simple aqueous solution. Several of these phenomena resemble traits displayed by living organisms. For example, the nanodroplets in a BZ microemulsion “communicate” with each other through a phenomenon analogous to quorum sensing in bacteria to produce a remarkable variety of patterns and waves on length scales 105 times the size of a single droplet. A photosensitive version can “remember” an imposed image. Larger, micrometer-sized droplets exhibit similarly rich behavior and allow for the observation and control of individual droplets. These droplets offer promise for building arrays capable of computation by varying the strength and sign of the coupling between drops. Gels that incorporate a BZ catalyst and are immersed in a solution containing the BZ reactants change their shape and volume in oscillations that follow the variation in the redox state of the catalyst. Using this phenomenon, we can construct phototactic gel “worms” or segments of gel that attract one another.Whether such systems will provide more realistic caricatures of life, and whether they can serve as useful materials will largely depend on the successful integration of various properties, including communication, motion, and memory, which we observed in separate experiments. Theoretical approaches that couple reaction and diffusion processes to mechanical and other material properties are likely to play a key role in this integration, and we describe one such approach. The evolution of systems of coupled chemical oscillators presents another challenge to the development of these systems, but one that we expect to be solved.
Co-reporter:Ye Zhang, Ning Li, Jorge Delgado, Ning Zhou, Ryo Yoshida, Seth Fraden, Irving R. Epstein and Bing Xu
Soft Matter 2012 vol. 8(Issue 26) pp:7056-7061
Publication Date(Web):31 May 2012
DOI:10.1039/C2SM25797A
We designed and synthesised two new polymerizable ruthenium complexes that catalyse the Belousov–Zhabotinsky (BZ) oscillating reaction and incorporated them into a copolymer to form hydrogels. The periodic oxidation and reduction of the attached ruthenium complex in the BZ reaction induces hydrating and dehydrating effects, respectively, that result in self-oscillatory volume changes of the hydrogel. We evaluated the correlation between the chemomechanical oscillation properties of the hydrogel and the proximity of the catalyst to the polymer backbone. Our results indicate that, like the change of such macroscopic parameters as temperature, reactant concentrations and pH, varying the microscopic distance between the catalyst and the polymeric chain provides a new way to tailor the chemomechanical behaviour, e.g., the initiation time, the frequency of oscillation, and the volume change of BZ hydrogels. Moreover, variation of the catalysts offers a new means to control the microstructure of the copolymer by expanding the range of monomer ratios. Modulation of molecular structure appears to be an effective way to alter the reaction–diffusion profile of species within heterogeneous chemoresponsive gels, thus contributing to the development of multifunctional, active soft materials capable of converting chemical energy into controllable mechanical forces.
Co-reporter: Viktor Horvath; Pier Luigi Gentili; Vladimir K. Vanag; Irving R. Epstein
Angewandte Chemie International Edition 2012 Volume 51( Issue 28) pp:6878-6881
Publication Date(Web):
DOI:10.1002/anie.201201962
Co-reporter:Ye Zhang, Ning Li, Jorge Delgado, Yuan Gao, Yi Kuang, Seth Fraden, Irving R. Epstein, and Bing Xu
Langmuir 2012 Volume 28(Issue 6) pp:3063-3066
Publication Date(Web):January 25, 2012
DOI:10.1021/la203923d
After a polymerizable hydrogelator self-assembles in water to form molecular nanofibers, post-self-assembly cross-linking allows the catalyst of the Belousov–Zhabotinsky (BZ) reaction to be attached to the nanofibers, resulting in a hydrogel that exhibits concentration oscillations, spiral waves, and concentric waves. In addition to the first report of the observation of BZ spiral waves in a hydrogel that covalently integrates the catalyst, we suggest a new approach to developing active soft materials as chemical oscillators and for exploring the correlation between molecular structure and far-from-equilibrium dynamics.
Co-reporter: Viktor Horvath; Pier Luigi Gentili; Vladimir K. Vanag; Irving R. Epstein
Angewandte Chemie 2012 Volume 124( Issue 28) pp:6984-6987
Publication Date(Web):
DOI:10.1002/ange.201201962
Co-reporter:Eszter Poros ; Viktor Horváth ; Krisztina Kurin-Csörgei ; Irving R. Epstein ;Miklós Orbán
Journal of the American Chemical Society 2011 Volume 133(Issue 18) pp:7174-7179
Publication Date(Web):April 15, 2011
DOI:10.1021/ja2010835
All pH-oscillators reported to date function only under open (flow reactor) conditions. We describe an approach to generating pH-oscillations in a closed system by starting from an open system pH-oscillator, finding semibatch conditions under which it oscillates with an inflow of a single reactant to an otherwise closed reactor containing the remaining components, and replacing this inflow with a layer of silica gel impregnated with the key reactant. We present data showing the successful application of this technique to the BrO3––Mn2+–SO32–, IO3––Fe(CN)64––SO32–, and BrO3––Fe(CN)64––SO32– systems. In all three cases, sulfite ion is the species that is replenished via dissolution from the gel layer.
Co-reporter:Tamás Bánsági Jr.;Vladimir K. Vanag
Science 2011 Volume 331(Issue 6022) pp:1309-1312
Publication Date(Web):11 Mar 2011
DOI:10.1126/science.1200815
Tomography reveals three-dimensional Turing patterns created by the Belousov-Zhabotinsky reaction running in a microemulsion.
Co-reporter:Dr. Federico Rossi;Dr. Vladimir K. Vanag; Irving R. Epstein
Chemistry - A European Journal 2011 Volume 17( Issue 7) pp:2138-2145
Publication Date(Web):
DOI:10.1002/chem.201002069
Abstract
We measure cross-diffusion coefficients in a five-component system, an aerosol OT (AOT) water-in-oil microemulsion loaded with two constituents of the Belousov–Zhabotinsky (BZ) reaction (H2O/AOT/BZ1/BZ2/octane). The species BZ1 is either NaBr, an inhibitor of the BZ reaction, or ferroin, a catalyst for the reaction. As species BZ2, we choose Br2, an intermediate in the reaction. The cross-diffusion coefficients between BZ1 and BZ2 are found to be negative, which can be understood in terms of complexation between these species. Using a four-variable model for the BZ reaction, we find that the cross-diffusion coefficients measured here can lead to a noticeable shift in the onset of Turing instability in the BZ–AOT system.
Co-reporter:Yuemin Zhao ; Shasha Wang ; Hamilton Varela ; Qingyu Gao ; Xuefeng Hu ; Jiaping Yang
The Journal of Physical Chemistry C 2011 Volume 115(Issue 26) pp:12965-12971
Publication Date(Web):May 24, 2011
DOI:10.1021/jp202881h
Spatiotemporal pattern formation in the electrocatalytic oxidation of sulfide on a platinum disk is investigated using electrochemical methods and a charge-coupled device (CCD) camera simultaneously. The system is characterized by different oscillatory regions spread over a wide potential range. An additional series resistor and a large electrode area facilitate observation of multiple regions of kinetic instabilities along the current/potential curve. Spatiotemporal patterns on the working electrode, such as fronts, pulses, spirals, twinkling eyes, labyrinthine stripes, and alternating synchronized deposition and dissolution, are observed at different operating conditions of series resistance and sweep rate.
Co-reporter:Jorge Delgado, Ye Zhang, Bing Xu, and Irving R. Epstein
The Journal of Physical Chemistry A 2011 Volume 115(Issue 11) pp:2208-2215
Publication Date(Web):February 28, 2011
DOI:10.1021/jp111724t
We report the successful use of Ru(II)(terpy)2 (1, terpy = 2,2′:6′,2′′-terpyridine) as a catalyst in the Belousov−Zhabotinsky (BZ) oscillating chemical reaction. We also examine several additional Ru(II) complexes, Ru(II)(bipy)2(L′)2 (2, L′ = 4-pyridinecarboxylic acid; bipy = 2,2′-bipyridine) and Ru(II)(bipy)2(L′′) (3, L′′ = 4,4′-dicarboxy-2,2′-bipy; 4, L′′ = N-allyl-4′-methyl-[2,2′-bipy]-4-carboxamide; 5, L′′ = bipy), for catalyzing the BZ reaction. While 2 is unable to trigger BZ oscillations, probably because of the rapid loss of L′ in a BZ solution, the other bipyridine-based Ru(II)-complexes can catalyze the BZ reaction, although their catalytic activity is adversely affected by slow ligand substitution in a BZ solution. Nevertheless, the successfully tested Ru(II)(terpy)2 and Ru(II)(bipy)2(L′′) catalysts may provide useful building blocks for complex functional macromolecules.
Co-reporter:Mieczysław Sajewicz, Miloš Dolnik, Dorota Kronenbach, Monika Gontarska, Teresa Kowalska, and Irving R. Epstein
The Journal of Physical Chemistry A 2011 Volume 115(Issue 50) pp:14331-14339
Publication Date(Web):November 4, 2011
DOI:10.1021/jp2070216
We employ high-performance liquid chromatography with diode array, evaporative light scattering, and mass spectrometric detection to monitor the oligomerization of l-lactic acid in pure acetonitrile and in 70% aqueous ethanol. The production of higher oligomers appears to proceed in an oscillatory fashion. A model is presented that involves the formation of aggregates (micelles), which catalyze the oligomerization.
Co-reporter:Yongchao Lu ; Qingyu Gao ; Li Xu ; Yuemin Zhao
Inorganic Chemistry 2010 Volume 49(Issue 13) pp:6026-6034
Publication Date(Web):June 1, 2010
DOI:10.1021/ic100573a
We have carried out species determination and kinetic analysis in the hydrogen peroxide−thiosulfate reaction by means of capillary electrophoresis and high performance liquid chromatography. In addition to thiosulfate, dithionate, trithionate, and tetrathionate, other polythionates such as pentathionate and hexathionate were detected during the oxidation process. The polythionates found are sensitive to the pH, with the average length of the sulfur chain decreasing with increasing pH. By varying the pH and the concentrations of the reactants, we find that the reaction is first order with respect to each of the reactants with rate constant k = 0.025 M−1 s−1. With HOS2O3−, HSO3−/SO32−, S3O62−, S4O62−, and S5O62− as key intermediates that are eventually oxidized to sulfate, a proposed 14-step kinetic model simulates the reaction process, including the evolution of the thiosulfate and tetrathionate concentrations at various pHs.
Co-reporter:Jorge Carballido-Landeira, Vladimir K. Vanag and Irving R. Epstein
Physical Chemistry Chemical Physics 2010 vol. 12(Issue 15) pp:3656-3665
Publication Date(Web):23 Feb 2010
DOI:10.1039/B919278F
We investigate the effect of changing temperature in the ferroin-catalysed Belousov–Zhabotinsky (BZ) reaction dispersed in the water nanodroplets of a water-in-oil aerosol OT (AOT) microemulsion, which undergoes a temperature-induced percolation transition at about 38 °C. We observe stationary Turing patterns at temperatures in the range 15–35 °C and bulk oscillations at T = 40–55 °C. When a temperature gradient ΔT is applied normal to a thin layer of the BZ–AOT reaction mixture, the range of patterns observed is dramatically expanded. Anti-phase oscillatory Turing patterns, leaping waves, and chaotic waves emerge, depending on the temperature gradient and the average temperature. These new patterns originate from the coupling between a low temperature Turing mode and a high temperature Hopf mode. Simulations with a simple model of the BZ–AOT system give good agreement with our experimental results.
Co-reporter:Viktor Horváth, Krisztina Kurin-Csörgei, Irving R. Epstein and Miklós Orbán
Physical Chemistry Chemical Physics 2010 vol. 12(Issue 6) pp:1248-1252
Publication Date(Web):23 Dec 2009
DOI:10.1039/B919924A
Oscillations in the concentration of divalent ions Cd2+, Ca2+, Zn2+, Co2+ and Ni2+ are induced by adding these species to the BrO3−–SO32− chemical oscillator in a flow reactor. Producing periodic pulses in the concentrations of these non-redox ions extends our earlier approach to generating forced periodic behavior. Instead of driving pH-dependent equilibrium reactions of the target ion by a pH oscillator backward and forward, we now couple a redox core oscillating reaction to two consecutive reactions taking place between the components of the oscillator and the target element. In the systems examined here, the oscillatory reductant SO32− binds the free metal ion in a MSO3 precipitate, reducing its level to a minimal value when [SO32−] is high, followed by release of the metal ion when the sulfite is oxidized by BrO3−.
Co-reporter:Mieczys&x142;aw Sajewicz;Marek Matlengiewicz;Marcin Leda;Monika Gontarska;Dorota Kronenbach;Teresa Kowalska
Journal of Physical Organic Chemistry 2010 Volume 23( Issue 11) pp:1066-1073
Publication Date(Web):
DOI:10.1002/poc.1739
Abstract
In earlier studies, we have collected experimental evidence (mostly from thin-layer chromatography and polarimetry) on the spontaneous oscillatory in vitro chiral conversion of simple carboxylic acids dissolved in 70% aqueous ethanol. To elucidate this phenomenon, we developed a simple theoretical model of two linked Templators. Recently, we have obtained additional experimental evidence of the spontaneous condensation of chiral carboxylic acids, based on the biuret test (amino acids), high performance liquid chromatography, and 13C NMR spectroscopy (profens and hydroxy acids). We briefly describe our experimental results in the context of the existing literature and outline an improved theoretical model for these phenomena. Our system resembles in some respects the reported oscillatory condensation of organic silanols. Here, the key reaction is the formation of carboxylic acid-derived enols. Finally, we discuss the importance of the oscillatory chiral conversion of simple carboxylic acids for biochemistry, pharmacology, and related life sciences. Copyright © 2010 John Wiley & Sons, Ltd.
Co-reporter:Mieczysław Sajewicz;Monika Gontarska;Dorota Kronenbach
Journal of Systems Chemistry 2010 Volume 1( Issue 1) pp:
Publication Date(Web):2010 December
DOI:10.1186/1759-2208-1-7
In earlier studies, we showed that certain low-molecular-weight carboxylic acids (profens, amino acids, hydroxy acids) can undergo spontaneous in vitro chiral conversion accompanied by condensation to from oligomers, and we proposed two simple models to describe these processes. Here, we present the results of investigations using non-chiral high-performance liquid chromatography with diode array detector (HPLC-DAD) and mass spectrometry (MS) on the dynamics of peptidization of S-, R-, and rac-phenylglycine dissolved in 70% aqueous ethanol and stored for times up to one year. The experimental results demonstrate that peptidization of phenylglycine can occur in an oscillatory fashion. We also describe, and carry out simulations with, three models that capture key aspects of the oscillatory condensation and chiral conversion processes.
Co-reporter:Masahiro Toiya, Hector O. González-Ochoa, Vladimir K. Vanag, Seth Fraden and Irving R. Epstein
The Journal of Physical Chemistry Letters 2010 Volume 1(Issue 8) pp:1241-1246
Publication Date(Web):March 26, 2010
DOI:10.1021/jz100238u
Many phenomena of biological, physical, and chemical importance involve synchronization of oscillatory elements. We explore here, in several geometries, the behavior of diffusively coupled, nanoliter volume, aqueous drops separated by oil gaps and containing the reactants of the oscillatory Belousov−Zhabotinsky (BZ) reaction. A variety of synchronous regimes are found, including in- and antiphase oscillations, stationary Turing patterns, and more complex combinations of stationary and oscillatory BZ drops, including three-phase patterns. A model consisting of ordinary differential equations based on a simplified description of the BZ chemistry and diffusion of messenger (primarily inhibitory) species qualitatively reproduces most of the experimental results.Keywords (keywords): antiphase oscillation; Belousov-Zhabotinsky reaction; in-phase oscillation; synchronized oscillators; Turing patterns;
Co-reporter:Ling Yuan, Qingyu Gao, Yuemin Zhao, Xiaodong Tang and Irving R. Epstein
The Journal of Physical Chemistry A 2010 Volume 114(Issue 26) pp:7014-7020
Publication Date(Web):June 10, 2010
DOI:10.1021/jp102010k
We study the oxidation dynamics of thiosulfate ions by hydrogen peroxide in the presence of trace amounts of copper(II) using the reaction temperature as a control parameter in a continuous flow stirred tank reactor. The system displays period-doubling, aperodic, and mixed-mode oscillations at different temperatures. We are able to simulate these complex dynamics with a model proposed by Kurin-Csörgei et al. The model suggests that the Cu2+-containing term is not essential for the observed oscillations. We find small-amplitude and high-frequency oscillations in the catalyst-free experimental system. The reaction between H2O2 and S2O32− contains the core mechanism of the H2O2−S2O32−−Cu2+ and H2O2−S2O32−−SO32- oscillatory systems, while the Cu2+ and SO32− modulate the feedback loops so as to strengthen the oscillatory dynamics.
Co-reporter:Changwei Pan, Qingyu Gao, Jingxuan Xie, Yu Xia and Irving R. Epstein
Physical Chemistry Chemical Physics 2009 vol. 11(Issue 46) pp:11033-11039
Publication Date(Web):13 Oct 2009
DOI:10.1039/B904445K
Two-dimensional Liesegang patterns formed when the boundary between electrolytes is polygonal display a variety of patterns, such as dislocations (radial alleys of gaps), branches (anastomoses) and spirals, many of which can be found in nature. Each vertex of the polygon can produce a pair of dislocation lines or branch lines. The effect caused by a vertex decreases with the number of vertices. Double-armed spirals are observed in experiments with a pentagonal boundary. Hexagons, which begin to approach smooth circular boundaries, do not give rise to dislocations, but instead yield concentric precipitation rings. A simple model of nucleation growth enables us to simulate dislocations and spirals consistent with those seen in our experiments.
Co-reporter:Vladimir K. Vanag and Irving R. Epstein
Physical Chemistry Chemical Physics 2009 vol. 11(Issue 6) pp:897-912
Publication Date(Web):11 Dec 2008
DOI:10.1039/B813825G
Cross-diffusion, the phenomenon in which a gradient in the concentration of one species induces a flux of another chemical species, has generally been neglected in the study of reaction–diffusion systems. We summarize experiments that demonstrate that cross-diffusion coefficients can be quite significant, even exceeding “normal,” diagonal diffusion coefficients in magnitude in systems that involve ions, micelles, complex formation, excluded volume effects (e.g., surface or polymer reactions) and other phenomena commonly encountered in situations of interest to chemists. We then demonstrate with a series of model calculations that cross-diffusion can lead to spatial and spatiotemporal pattern formation, even in relatively simple systems. We also show that, in the absence of cross-diffusion among the reacting species, introduction of a nonreactive species that induces appropriate cross-diffusive fluxes with reactive species can lead to pattern formation.
Co-reporter:Klara Kovacs, Marcin Leda, Vladimir K. Vanag and Irving R. Epstein
The Journal of Physical Chemistry A 2009 Volume 113(Issue 1) pp:146-156
Publication Date(Web):December 16, 2008
DOI:10.1021/jp807840g
The BrO3−−SO32−−Fe(CN)64− (BSF) pH-oscillatory system is coupled to the Al(OH)3 precipitation equilibrium (BSFA system) and studied in a stirred flow reactor. The dynamic behavior of the BSFA system differs significantly from that of the BSF system. In addition to the large-amplitude pH oscillations found in the BSF system, new small-amplitude and mixed-mode oscillations occur. A detailed mechanism of the BSFA system is developed and investigated.
Co-reporter:Jiamin Feng, Qingyu Gao, Xiaoli Lv and Irving R. Epstein
The Journal of Physical Chemistry A 2008 Volume 112(Issue 29) pp:6578-6585
Publication Date(Web):June 27, 2008
DOI:10.1021/jp802002k
We explored the temperature-dependent dynamics of the electrochemical oxidation of thiourea under potential-control mode and found complex oscillations with one large peak and one small peak per period. Adjusting the temperature caused the relative amplitudes and positions of the two peaks to vary. Experiments showed that there were two distinct oscillatory regimes as a function of the external current, and at some temperatures, three-frequency quasiperiodic oscillations occurred for current densities between the two oscillatory windows. We examined the species present and the component reactions by employing cyclic voltammetry and electrolysis−HPLC−MS in the two oscillatory regimes. Adsorption and desorption of multiple species, including water and chloride, contribute to the rich dynamical phenomena that were observed.
Co-reporter:Vladimir K. Vanag, Federico Rossi, Alexander Cherkashin and Irving R. Epstein
The Journal of Physical Chemistry B 2008 Volume 112(Issue 30) pp:9058-9070
Publication Date(Web):July 9, 2008
DOI:10.1021/jp800525w
We describe an improved Taylor dispersion method for four-component systems, which we apply to measure the main- and cross-diffusion coefficients in an Aerosol OT water-in-oil microemulsion loaded with one of the reactants of the Belousov−Zhabotinsky (BZ) reaction, water(1)/AOT(2)/R(3)/octane(4) system, where R is malonic acid or ferroin. With [H2O]/[AOT] = 11.8 and volume droplet fraction φd = 0.18, when the microemulsion is below the percolation transition, the cross-diffusion coefficients D13 and D23 are large and positive (D13/D33 ≅ 14, D23/D33 ≅ 3) for malonic acid and large and negative for ferroin (D13/D33 ≅ −112, D23/D33 ≅ −30) while coefficients D31 and D32 are small and negative for malonic acid (D31/D33 ≅ −0.01, D32/D33 ≅ −0.14) and small and positive for ferroin (D31/D33 ≅ 5 × 10−4, D32/D33 ≅ 8 × 10−3). These data represent the first direct determination of cross-diffusion effects in a pattern-forming system and of the full matrix of diffusion coefficients for a four-component system. The results should provide a basis for modeling pattern formation in the BZ−AOT system.
Co-reporter:Masahiro Toiya Dr.;VladimirK. Vanag ;IrvingR. Epstein
Angewandte Chemie 2008 Volume 120( Issue 40) pp:7867-7869
Publication Date(Web):
DOI:10.1002/ange.200802339
Co-reporter:Masahiro Toiya Dr.;VladimirK. Vanag ;IrvingR. Epstein
Angewandte Chemie International Edition 2008 Volume 47( Issue 40) pp:7753-7755
Publication Date(Web):
DOI:10.1002/anie.200802339
Co-reporter:Qingyu Gao, Guangping Wang, Yanyan Sun and Irving R. Epstein
The Journal of Physical Chemistry A 2008 Volume 112(Issue 26) pp:5771-5773
Publication Date(Web):June 5, 2008
DOI:10.1021/jp8003932
We employ reversed-phase ion-pair high-performance liquid chromatography to quantitatively characterize the oxidation kinetics of thiourea oxidation by hydrogen peroxide. The HPLC technique makes it possible to monitor the concentrations of a variety of sulfur-containing species with different oxidation states and to elucidate the relative phase relations among them. The experimental results are in good agreement with simulations from an 8-step reaction mechanism.
Co-reporter:Vladimir K. Vanag;David G. Míguez
PNAS 2007 Volume 104 (Issue 17 ) pp:6992-6997
Publication Date(Web):2007-04-24
DOI:10.1073/pnas.0611438104
Waves and patterns in living systems are often driven by biochemical reactions with enzymes as catalysts and regulators. We
present a reaction–diffusion system catalyzed by the enzyme glucose oxidase that exhibits traveling wave patterns in a spatially
extended medium. Fronts and pulses propagate as a result of the coupling between the enzyme-catalyzed autocatalytic production
and diffusion of hydrogen ions. A mathematical model qualitatively explains the experimental observations.
Co-reporter:Lingfa Yang, Anatol M. Zhabotinsky and Irving R. Epstein
Physical Chemistry Chemical Physics 2006 vol. 8(Issue 40) pp:4647-4651
Publication Date(Web):11 Sep 2006
DOI:10.1039/B609214D
We describe a new type of solitary waves, which propagate in such a manner that the pulse periodically disappears from its original position and reemerges at a fixed distance. We find such jumping waves as solutions to a reaction–diffusion system with a subcritical short-wavelength instability. We demonstrate closely related solitary wave solutions in the quintic complex Ginzburg–Landau equation. We study the characteristics of and interactions between these solitary waves and the dynamics of related wave trains and standing waves.
Co-reporter:Akiko Kaminaga Dr.;Vladimir K. Vanag Dr. Professor
Angewandte Chemie 2006 Volume 118(Issue 19) pp:
Publication Date(Web):29 MAR 2006
DOI:10.1002/ange.200600400
Ein Gesicht, das man nicht vergisst: Ein Bild, das durch Licht in eine reaktive Mikroemulsion eingeprägt wurde, bleibt mehr als eine Stunde erhalten. Solche lokalisierten Muster in Reaktions-Diffusions-Systemen könnten in Speichereinheiten eingesetzt werden. Die Abbildung zeigt das Bild eines Gesichts unmittelbar nach der Bestrahlung (a) und nach 1 h (b). Im rechten Bereich ist die Entwicklung eines spontanen Turing-Musters zu erkennen.
Co-reporter:Akiko Kaminaga, Vladimir K. Vanag,Irving R. Epstein
Angewandte Chemie International Edition 2006 45(19) pp:3087-3089
Publication Date(Web):
DOI:10.1002/anie.200600400
Co-reporter:Krisztina Kurin-Csörgei,
Irving R. Epstein
and
Miklós Orbán
Nature 2005 433(7022) pp:139
Publication Date(Web):
DOI:10.1038/nature03214
Co-reporter:Yaneer Bar-Yam
PNAS 2004 Volume 101 (Issue 13 ) pp:4341-4345
Publication Date(Web):2004-03-30
DOI:10.1073/pnas.0400673101
We consider the response of complex systems to stimuli and argue for the importance of both sensitivity, the possibility of
large response to small stimuli, and robustness, the possibility of small response to large stimuli. Using a dynamic attractor
network model for switching of patterns of behavior, we show that the scale-free topologies often found in nature enable more
sensitive response to specific changes than do random networks. This property may be essential in networks where appropriate
response to environmental change is critical and may, in such systems, be more important than features, such as connectivity,
often used to characterize network topologies. Phenomenologically observed exponents for functional scale-free networks fall
in a range corresponding to the onset of particularly high sensitivities, while still retaining robustness.
Co-reporter:Layla Badr, Irving R. Epstein
Chemical Physics Letters (February 2017) Volume 669() pp:
Publication Date(Web):February 2017
DOI:10.1016/j.cplett.2016.11.050
•Copper(I) oxide nanoparticles can be synthesized by a simple reaction-diffusion process.•The mean diameter and the size dispersion of the particles are controlled by two experimental parameters.•The synthetic approach is a modification of that used to produce Liesegang rings.Copper (I) oxide nanoparticles are synthesized by a simple reaction-diffusion process involving Cu+ ions and sodium hydroxide in gelatin. The mean diameter and the size dispersion of the nanoparticles can be controlled by two experimental parameters, the percent of gelatin in the medium and the hydroxide ion concentration. UV–visible spectroscopy, transmission electron microscopy and X-ray diffraction are used to analyze the size, morphology, and chemical composition of the nanoparticles generated.
Co-reporter:Xingjie Lu, Lin Ren, Qingyu Gao, Yuemin Zhao, Shaorong Wang, Jiaping Yang and Irving R. Epstein
Chemical Communications 2013 - vol. 49(Issue 70) pp:NaN7692-7692
Publication Date(Web):2013/07/17
DOI:10.1039/C3CC44480E
A photosensitive self-oscillating gel that incorporates the Belousov–Zhabotinsky reaction can undergo rhythmic mechanical oscillations. We exploit the dependence of the oscillation frequency on light intensity to generate both photophobic and phototropic movement of the gel under differential illumination. Our findings may be used in designing intelligent sensors that can execute biomimetic behaviours.
Co-reporter:Viktor Horvath, Daniel J. Kutner, John T. Chavis III and Irving R. Epstein
Physical Chemistry Chemical Physics 2015 - vol. 17(Issue 6) pp:NaN4676-4676
Publication Date(Web):2015/01/13
DOI:10.1039/C4CP05416D
Coupled chemical oscillators are usually studied with symmetric coupling, either between identical oscillators or between oscillators whose frequencies differ. Asymmetric connectivity is important in neuroscience, where synaptic strength inequality in neural networks commonly occurs. While the properties of the individual oscillators in some coupled chemical systems may be readily changed, enforcing inequality between the connection strengths in a reciprocal coupling is more challenging. We recently demonstrated a novel way of coupling chemical oscillators, which allows for manipulation of individual connection strengths. Here we study two identical, pulse-coupled Belousov–Zhabotinsky (BZ) oscillators with unequal connection strengths. When the pulse perturbations contain KBr (inhibitor), this system exhibits simple out-of-phase and complex oscillations, oscillatory-suppressed states as well as temporally periodic patterns (N:M) in which the two oscillators exhibit different numbers of peaks per cycle. The N:M patterns emerge due to the long-term effect of the inhibitory pulse-perturbations, a feature that has not been considered in earlier works. Time delay was previously shown to have a profound effect on the system’s behaviour when pulse coupling was inhibitory and the coupling strengths were equal. When the coupling is asymmetric, however, delay produces no qualitative change in behaviour, though the 1:2 temporal pattern becomes more robust. Asymmetry in instantaneous excitatory coupling via AgNO3 injection produces a previously unseen temporal pattern (1:N patterns starting with a double peak) with time delay and high [AgNO3]. Numerical simulations of the behaviour agree well with theoretical predictions in asymmetrical pulse-coupled systems.
Co-reporter:Jorge Carballido-Landeira, Vladimir K. Vanag and Irving R. Epstein
Physical Chemistry Chemical Physics 2010 - vol. 12(Issue 15) pp:NaN3665-3665
Publication Date(Web):2010/02/23
DOI:10.1039/B919278F
We investigate the effect of changing temperature in the ferroin-catalysed Belousov–Zhabotinsky (BZ) reaction dispersed in the water nanodroplets of a water-in-oil aerosol OT (AOT) microemulsion, which undergoes a temperature-induced percolation transition at about 38 °C. We observe stationary Turing patterns at temperatures in the range 15–35 °C and bulk oscillations at T = 40–55 °C. When a temperature gradient ΔT is applied normal to a thin layer of the BZ–AOT reaction mixture, the range of patterns observed is dramatically expanded. Anti-phase oscillatory Turing patterns, leaping waves, and chaotic waves emerge, depending on the temperature gradient and the average temperature. These new patterns originate from the coupling between a low temperature Turing mode and a high temperature Hopf mode. Simulations with a simple model of the BZ–AOT system give good agreement with our experimental results.
Co-reporter:Changwei Pan, Qingyu Gao, Jingxuan Xie, Yu Xia and Irving R. Epstein
Physical Chemistry Chemical Physics 2009 - vol. 11(Issue 46) pp:NaN11039-11039
Publication Date(Web):2009/10/13
DOI:10.1039/B904445K
Two-dimensional Liesegang patterns formed when the boundary between electrolytes is polygonal display a variety of patterns, such as dislocations (radial alleys of gaps), branches (anastomoses) and spirals, many of which can be found in nature. Each vertex of the polygon can produce a pair of dislocation lines or branch lines. The effect caused by a vertex decreases with the number of vertices. Double-armed spirals are observed in experiments with a pentagonal boundary. Hexagons, which begin to approach smooth circular boundaries, do not give rise to dislocations, but instead yield concentric precipitation rings. A simple model of nucleation growth enables us to simulate dislocations and spirals consistent with those seen in our experiments.
Co-reporter:Vladimir K. Vanag and Irving R. Epstein
Physical Chemistry Chemical Physics 2009 - vol. 11(Issue 6) pp:NaN912-912
Publication Date(Web):2008/12/11
DOI:10.1039/B813825G
Cross-diffusion, the phenomenon in which a gradient in the concentration of one species induces a flux of another chemical species, has generally been neglected in the study of reaction–diffusion systems. We summarize experiments that demonstrate that cross-diffusion coefficients can be quite significant, even exceeding “normal,” diagonal diffusion coefficients in magnitude in systems that involve ions, micelles, complex formation, excluded volume effects (e.g., surface or polymer reactions) and other phenomena commonly encountered in situations of interest to chemists. We then demonstrate with a series of model calculations that cross-diffusion can lead to spatial and spatiotemporal pattern formation, even in relatively simple systems. We also show that, in the absence of cross-diffusion among the reacting species, introduction of a nonreactive species that induces appropriate cross-diffusive fluxes with reactive species can lead to pattern formation.
Co-reporter:Viktor Horváth, Krisztina Kurin-Csörgei, Irving R. Epstein and Miklós Orbán
Physical Chemistry Chemical Physics 2010 - vol. 12(Issue 6) pp:NaN1252-1252
Publication Date(Web):2009/12/23
DOI:10.1039/B919924A
Oscillations in the concentration of divalent ions Cd2+, Ca2+, Zn2+, Co2+ and Ni2+ are induced by adding these species to the BrO3−–SO32− chemical oscillator in a flow reactor. Producing periodic pulses in the concentrations of these non-redox ions extends our earlier approach to generating forced periodic behavior. Instead of driving pH-dependent equilibrium reactions of the target ion by a pH oscillator backward and forward, we now couple a redox core oscillating reaction to two consecutive reactions taking place between the components of the oscillator and the target element. In the systems examined here, the oscillatory reductant SO32− binds the free metal ion in a MSO3 precipitate, reducing its level to a minimal value when [SO32−] is high, followed by release of the metal ion when the sulfite is oxidized by BrO3−.