To see the other types of publications on this topic, follow the link: Autocatalytic reaction.

Journal articles on the topic 'Autocatalytic reaction'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Autocatalytic reaction.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Fedotov, Vladislav Kh, Nikolay I. Kol'tsov, and Petr M. Kosianov. "INFLUENCE OF THE AUTOCATALYTIC STAGES ON THE DYNAMICS OF CONJUGATED CHEMICAL REACTIONS." IZVESTIYA VYSSHIKH UCHEBNYKH ZAVEDENII KHIMIYA KHIMICHESKAYA TEKHNOLOGIYA 63, no. 2 (February 8, 2020): 14–20. http://dx.doi.org/10.6060/ivkkt.20206302.6053.

Full text
Abstract:
Chemical reactions occurring on nonlinear mechanisms, containing the stage of interaction of various reagents (feedback), can exhibit unusual kinetic properties - the multiplicity of equilibria (hysteresis of different shape dependency on the «velocity-parameter»), change the time of the motion to the equilibrium (slow or fast relaxation), sustained oscillations (regular, irregular), etc. All these critical phenomena are usually associated with the appearance of unstable equilibria in the reactions under study. From the kinetic point of view, one of the main causes of instability is the presence of autocatalytic stages in the reaction mechanism. Therefore, it is interesting to study the effect of autocatalytic stages on the kinetics of chemical reactions, especially far from equilibrium. In this regard, the dynamic characteristics of typical conjugate reactions occurring by non-autocatalytic and autocatalytic mechanisms in an isothermal reactor of ideal mixing under the same conditions are compared in this paper. It is shown that the kinetics of these reactions is different: autocatalysis can shift the equilibrium, change the relaxation time and the rate of reactions. In an irreversible consecutive reaction (far from equilibrium) autocatalysis shifts the equilibrium in the direction of increasing the proportion occupied by the surface of the catalyst and the reaction rate dominated by positive autocatalysis. As the reversible processes increase, the balance shifts to the other side, the reaction slows down and autoinhibition begins to prevail. In parallel conjugate reactions, negative autocatalysis is not observed. In both types of the considered conjugate reactions, the maximum positive change in concentrations and velocity due to autocatalysis observed when these reactions are irreversible.
APA, Harvard, Vancouver, ISO, and other styles
2

Qi, Yuanwei, and Yi Zhu. "Computational Study of Traveling Wave Solutions of Isothermal Chemical Systems." Communications in Computational Physics 19, no. 5 (May 2016): 1461–72. http://dx.doi.org/10.4208/cicp.scpde14.38s.

Full text
Abstract:
AbstractThis article studies propagating traveling waves in a class of reaction-diffusion systems which model isothermal autocatalytic chemical reactions as well as microbial growth and competition in a flow reactor. In the context of isothermal autocatalytic systems, two different cases will be studied. The first is autocatalytic chemical reaction of order m without decay. The second is chemical reaction of order m with a decay of order n, where m and n are positive integers and m>n≥1. A typical system in autocatalysis is A+2B→3B and B→C involving two chemical species, a reactant A and an auto-catalyst B and C an inert chemical species.The numerical computation gives more accurate estimates on minimum speed of traveling waves for autocatalytic reaction without decay, providing useful insight in the study of stability of traveling waves.For autocatalytic reaction of order m = 2 with linear decay n = 1, which has a particular important role in chemical waves, it is shown numerically that there exist multiple traveling waves with 1, 2 and 3 peaks with certain choices of parameters.
APA, Harvard, Vancouver, ISO, and other styles
3

Blokhuis, Alex, David Lacoste, and Philippe Nghe. "Universal motifs and the diversity of autocatalytic systems." Proceedings of the National Academy of Sciences 117, no. 41 (September 28, 2020): 25230–36. http://dx.doi.org/10.1073/pnas.2013527117.

Full text
Abstract:
Autocatalysis is essential for the origin of life and chemical evolution. However, the lack of a unified framework so far prevents a systematic study of autocatalysis. Here, we derive, from basic principles, general stoichiometric conditions for catalysis and autocatalysis in chemical reaction networks. This allows for a classification of minimal autocatalytic motifs called cores. While all known autocatalytic systems indeed contain minimal motifs, the classification also reveals hitherto unidentified motifs. We further examine conditions for kinetic viability of such networks, which depends on the autocatalytic motifs they contain and is notably increased by internal catalytic cycles. Finally, we show how this framework extends the range of conceivable autocatalytic systems, by applying our stoichiometric and kinetic analysis to autocatalysis emerging from coupled compartments. The unified approach to autocatalysis presented in this work lays a foundation toward the building of a systems-level theory of chemical evolution.
APA, Harvard, Vancouver, ISO, and other styles
4

Skorb, Ekaterina V., and Sergey N. Semenov. "Mathematical Analysis of a Prototypical Autocatalytic Reaction Network." Life 9, no. 2 (May 20, 2019): 42. http://dx.doi.org/10.3390/life9020042.

Full text
Abstract:
Network autocatalysis, which is autocatalysis whereby a catalyst is not directly produced in a catalytic cycle, is likely to be more common in chemistry than direct autocatalysis is. Nevertheless, the kinetics of autocatalytic networks often does not exactly follow simple quadratic or cubic rate laws and largely depends on the structure of the network. In this article, we analyzed one of the simplest and most chemically plausible autocatalytic networks where a catalytic cycle is coupled to an ancillary reaction that produces the catalyst. We analytically analyzed deviations in the kinetics of this network from its exponential growth and numerically studied the competition between two networks for common substrates. Our results showed that when quasi-steady-state approximation is applicable for at least one of the components, the deviation from the exponential growth is small. Numerical simulations showed that competition between networks results in the mutual exclusion of autocatalysts; however, the presence of a substantial noncatalytic conversion of substrates will create broad regions where autocatalysts can coexist. Thus, we should avoid the accumulation of intermediates and the noncatalytic conversion of the substrate when designing experimental systems that need autocatalysis as a source of positive feedback or as a source of evolutionary pressure.
APA, Harvard, Vancouver, ISO, and other styles
5

Baier, Gerold, and Sven Sahle. "Spatio-temporal patterns with hyperchaotic dynamics in diffusively coupled biochemical oscillators." Discrete Dynamics in Nature and Society 1, no. 2 (1997): 161–67. http://dx.doi.org/10.1155/s1026022697000162.

Full text
Abstract:
We present three examples how complex spatio-temporal patterns can be linked to hyperchaotic attractors in dynamical systems consisting of nonlinear biochemical oscillators coupled linearly with diffusion terms. The systems involved are: (a) a two-variable oscillator with two consecutive autocatalytic reactions derived from the Lotka–Volterra scheme; (b) a minimal two-variable oscillator with one first-order autocatalytic reaction; (c) a three-variable oscillator with first-order feedback lacking autocatalysis. The dynamics of a finite number of coupled biochemical oscillators may account for complex patterns in compartmentalized living systems like cells or tissue, and may be tested experimentally in coupled microreactors.
APA, Harvard, Vancouver, ISO, and other styles
6

Simoyi, Reuben H. "Autocatalytic chlorite-bromide reaction." Journal of Physical Chemistry 89, no. 16 (August 1985): 3570–74. http://dx.doi.org/10.1021/j100262a029.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Gyevi-Nagy, László, Emese Lantos, Tünde Gehér-Herczegh, Ágota Tóth, Csaba Bagyinka, and Dezső Horváth. "Reaction fronts of the autocatalytic hydrogenase reaction." Journal of Chemical Physics 148, no. 16 (April 28, 2018): 165103. http://dx.doi.org/10.1063/1.5022359.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

El-Sayed, Saad A. "Explosion characteristics of autocatalytic reaction." Combustion and Flame 133, no. 3 (May 2003): 375–78. http://dx.doi.org/10.1016/s0010-2180(03)00014-2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

El-Sayed, Saad A. "Thermal explosion of autocatalytic reaction." Journal of Loss Prevention in the Process Industries 16, no. 4 (July 2003): 249–57. http://dx.doi.org/10.1016/s0950-4230(03)00039-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Stadler, Peter F., and Peter Schuster. "Mutation in autocatalytic reaction networks." Journal of Mathematical Biology 30, no. 6 (1992): 597–631. http://dx.doi.org/10.1007/bf00948894.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Canepa, Carlo. "The role of autocatalysis on the chemical diversity of the prebiotic ocean of early Earth." International Journal of Astrobiology 15, no. 1 (May 5, 2015): 57–64. http://dx.doi.org/10.1017/s1473550415000099.

Full text
Abstract:
AbstractThe spontaneous formation of catalytic polypeptides of various lengths in a primordial ocean endowed with a source of amino acids from micrometeorites was investigated and found to be sufficient to induce the transformation of potential substrates under the assumption of a high propensity of the environment to catalyse the formation of the peptide bond. This work aims to include in this picture the effect of autocatalysis, i.e. the ability of a polypeptide with a specific length to promote the formation of the peptide bond. Once the formation of an autocatalytic species is attained, the concentrations of the polypeptides, substrates and products of reaction exhibit a time-dependent rate of formation and undergo a catastrophic change. While in the absence of autocatalysis the concentrations of polypeptides are stationary and the formation of reaction products is limited by the proper frequency λ, autocatalysis induces a steady growth of the concentrations of polypeptides and a 100 − 105-fold increase of reaction products at t = ω−1<0.46 Gyr, with a subsequent linear growth in time according to the law u/z0 = 1+s(ω−1+t)/z0, provided the autocatalytic species be active with length fewer than 70 amino acid units. A relationship was found between the catalytic ability of the environment (expressed by the ratio η/ηh of the rate coefficient for peptide bond formation to the corresponding rate coefficient for hydrolysis) and the time of the sharp increase of the concentration of both the polypeptides and their products of transformation. Although the formation of autocatalytic polypeptides is able to rapidly induce a sharp increase in the concentration of both polypeptides and their products of transformation, the crucial formation of the first autocatalytic polypeptides relies on the ability of the environment to promote the formation of the peptide bond. The value of the ratio η/ηh, constrained by the available time for chemical evolution to values bordering the catalytic activity of present-day enzymes, suggests that the correlation between the presence of water and the formation of a complex chemistry should be taken with caution.
APA, Harvard, Vancouver, ISO, and other styles
12

HONMA, Hideo, Teruyoshi SAITO, and Takeshi KOBAYASHI. "Electroless Solder Plating by Autocatalytic Reaction." Journal of the Surface Finishing Society of Japan 45, no. 1 (1994): 108–9. http://dx.doi.org/10.4139/sfj.45.108.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Schwartz, Lowell M. "More on autocatalytic reaction data analysis." Journal of Chemical Education 66, no. 8 (August 1989): 677. http://dx.doi.org/10.1021/ed066p677.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Agliari, E., R. Burioni, D. Cassi, and F. M. Neri. "Autocatalytic reaction on low-dimensional substrates." Theoretical Chemistry Accounts 118, no. 5-6 (May 26, 2007): 855–62. http://dx.doi.org/10.1007/s00214-007-0323-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Zhou, Jun, and Junping Shi. "Qualitative analysis of an autocatalytic chemical reaction model with decay." Proceedings of the Royal Society of Edinburgh: Section A Mathematics 144, no. 2 (March 20, 2014): 427–46. http://dx.doi.org/10.1017/s0308210512001667.

Full text
Abstract:
In this paper, we revisit a reaction—diffusion autocatalytic chemical reaction model with decay. For higher-order reactions, we prove that the system possesses at least two positive steady-state solutions; hence, it has bistable dynamics similar to the system without decay. For the linear reaction, we determine the necessary and sufficient condition to ensure the existence of a solution. Moreover, in the one-dimensional case, we prove that the positive steady-state solution is unique. Our results demonstrate the drastic difference in dynamics caused by the order of chemical reactions.
APA, Harvard, Vancouver, ISO, and other styles
16

Priimägi, Linnar. "The problem of the autocatalytic origin of culture in Juri Lotman’s cultural philosophy." Sign Systems Studies 33, no. 1 (December 31, 2005): 191–204. http://dx.doi.org/10.12697/sss.2005.33.1.08.

Full text
Abstract:
The origin of culture remains in the sphere of hypotheses. Although the hypotheses derive from two presumptions: first, how the structure of culture is envisaged, and secondly, how culture is thought to function. Juri Lotman dealt with both aspects of culture, initially the structural and typological and later the dynamic aspects. Thereby, he arrived at the cultural-philosophical hypothesis of the autocatalytic origin of culture. A catalyst is a component of a chemical reaction which itself doesn’t transform during the reaction, but whose presence is needed to guarantee the reaction (or to stimulate it). Thus, autocatalysis is a paradoxical situation in which the genesis of something presumes the pre-existence of the final product. The paradox of the autocatalysis of culture lies in the fact that culture cannot emerge from anything other than from culture itself, from its own germination. In 1988, speaking about the autocatalysis of culture, Lotman refered to the cultural historicist Nikolai I. Konrad (1891–1970), who undoubtedly borrowed this idea from Jacob Christopher Burckhardt (1818–1897). This undiscovered connection reminds us of the fact, that a model for autocatalysis (or an autopoiesis) was basic to Naturphilosophie of the 19th century. In the 20th century, this was represented by Vladimir I. Vernadsky (1863–1945), from whom Lotman in 1982 received the impetus to formulate the concept of semiosphere as well as of the autocatalysis of culture. The autocatalysis model of culture is culturally diachronical, the semiosphere is, however, a synchronical one. In both cases, the natural philosophical cytology of the 19th century was Lotman’s semiotical meta-language.
APA, Harvard, Vancouver, ISO, and other styles
17

Kauffman, Stuart, and Mike Steel. "The Expected Number of Viable Autocatalytic Sets in Chemical Reaction Systems." Artificial Life 27, no. 1 (2021): 1–14. http://dx.doi.org/10.1162/artl_a_00333.

Full text
Abstract:
Abstract The emergence of self-sustaining autocatalytic networks in chemical reaction systems has been studied as a possible mechanism for modeling how living systems first arose. It has been known for several decades that such networks will form within systems of polymers (under cleavage and ligation reactions) under a simple process of random catalysis, and this process has since been mathematically analyzed. In this paper, we provide an exact expression for the expected number of self-sustaining autocatalytic networks that will form in a general chemical reaction system, and the expected number of these networks that will also be uninhibited (by some molecule produced by the system). Using these equations, we are able to describe the patterns of catalysis and inhibition that maximize or minimize the expected number of such networks. We apply our results to derive a general theorem concerning the trade-off between catalysis and inhibition, and to provide some insight into the extent to which the expected number of self-sustaining autocatalytic networks coincides with the probability that at least one such system is present.
APA, Harvard, Vancouver, ISO, and other styles
18

KOLB, M. "SURFACE EXPLOSION IN AUTOCATALYTIC MONOMER-MONOMER REACTIONS." Fractals 04, no. 02 (June 1996): 187–92. http://dx.doi.org/10.1142/s0218348x9600025x.

Full text
Abstract:
A lattice model of the A-B type has been introduced recently to describe temporal oscillations and pattern formation for the NO-CO reaction catalyzed by a Pt surface. Under certain experimental conditions one observes the rapid reactional removal of an adsorbed layer upon nucleation — a direct consequence of the autocatalytic nature of the reaction. Here this surface explosion is described by the same model, but now only depending on a single parameter: the initial fraction of A sites. Simulations show that a reactive "cleaning" of the surface is possible for a wide range of A coverages. For low (high) values of the parameter, a transition towards a B (A) poisoned state is observed. Fractal reaction patterns appear in the vicinity of the transition points. Both transitions have large scale characteristics of standard percolation.
APA, Harvard, Vancouver, ISO, and other styles
19

Kol'tsov, Nikolay I. "CHAOTIC OSCILLATIONS IN SIMPLEST CHEMICAL REACTION." IZVESTIYA VYSSHIKH UCHEBNYKH ZAVEDENIY KHIMIYA KHIMICHESKAYA TEKHNOLOGIYA 61, no. 4-5 (April 17, 2018): 133. http://dx.doi.org/10.6060/tcct.20186104-05.5654.

Full text
Abstract:
It is known that chaotic oscillations for chemical reactions can be described by non-stationary kinetic models consisting of three ordinary differential equations. Rossler established the first examples of chemical reactions, including the two-route five-stage reaction of the Villamovski-Rossler, with three intermediate substances, containing three autocatalytic on intermediates stages, the dynamic model of which describes chaotic oscillations. In given article presents a simple one-route four-stages reaction A+E=D involving two autocatalytic and one linear on intermediate stage, the non-stationary kinetic model of which describes chaotic oscillations. The non-stationary kinetic model under the assumption of quasistationarity with respect to the main substances within the framework of the law of acting masses is a system of three ordinary differential equations. The presence of chaos is confirmed by numerical calculations of the kinetic model and Lyapunov exponentials. The Lyapunov exponents satisfy the condition L1+L2+L3<0, which proves the existence of chaotic oscillations.Forcitation:Kol'tsov N.I. Chaotic oscillations in simplest chemical reaction. Izv. Vyssh. Uchebn. Zaved. Khim. Khim. Tekhnol. 2018. V. 61. N 4-5. P. 133-135
APA, Harvard, Vancouver, ISO, and other styles
20

Arcadia, Christopher E., Amanda Dombroski, Kady Oakley, Shui Ling Chen, Hokchhay Tann, Christopher Rose, Eunsuk Kim, Sherief Reda, Brenda M. Rubenstein, and Jacob K. Rosenstein. "Leveraging autocatalytic reactions for chemical domain image classification." Chemical Science 12, no. 15 (2021): 5464–72. http://dx.doi.org/10.1039/d0sc05860b.

Full text
Abstract:
Kinetic models of autocatalytic reactions have mathematical forms similar to activation functions used in artificial neural networks. Inspired by these similarities, we use a copper-catalyzed reaction to perform digital image recognition tasks.
APA, Harvard, Vancouver, ISO, and other styles
21

Mrákavová, Marta, Milan Melicherčík, Anna Olexová, and Ľudovít Treindl. "The Autocatalytic Reduction of Ferriin by Malonic Acid with Regard to the Ferroin-Catalyzed Belousov-Zhabotinsky Reaction." Collection of Czechoslovak Chemical Communications 68, no. 1 (2003): 23–34. http://dx.doi.org/10.1135/cccc20030023.

Full text
Abstract:
The reduction of ferriin ([Fe(phen)3]3+, phen = 1,10-phenanthroline) by malonic acid (MA) differs from the reduction of Ce(IV) or Mn(III) ions by MA in its autocatalytic character and in a pregnant influence of oxygen, which behaves obviously as a catalyst. The time dependence of the ferroin-ferriin redox potential at the last stage of this reaction has a sigmoidal shape, which indicates autocatalysis. Under anaerobic conditions, the inflection time is of the order of several tens of minutes, since autocatalysis cannot proceed unless a sufficient amount of oxygen is produced via oxidation of water (OH- ions) with Fe(IV) formed by the ferriin dismutation. Under aerobic conditions, the inflection time decreases to a value of a few seconds. The probable reaction mechanism is discussed in detail.
APA, Harvard, Vancouver, ISO, and other styles
22

LIU, Jiang-Hong, and Zhi-Xin WANG. "Kinetic analysis of ligand-induced autocatalytic reactions." Biochemical Journal 379, no. 3 (May 1, 2004): 697–702. http://dx.doi.org/10.1042/bj20031365.

Full text
Abstract:
Protein phosphorylation and limited proteolysis are two most common regulatory mechanisms involving the energy-dependent covalent modification of regulatory enzymes. In addition to modifying other proteins, many protein kinases and proteases catalyse automodification reactions (i.e. reactions in which the kinase or zymogen serves as its own substrate), and their activities are frequently regulated by other regulatory ligands. In the present study, a kinetic analysis of autocatalytic reaction modulated by regulatory ligands is presented. On the basis of the kinetic equation, a novel procedure is developed to evaluate the kinetic parameters of the reaction. As an example of an application of this method, the effects of calcium ions on the autoacatalytic activation of trypsinogen by trypsin is re-examined. The results indicate that the binding affinity for Ca2+-bound trypsinogen to trypsin is at least two orders of magnitude higher than that for Ca2+-free trypsinogen, and therefore that the effect of Ca2+ ions on Km* values for trypsinogen is very much greater than that for the model peptides. Based on the experimental results, one possible molecular mechanism has been proposed.
APA, Harvard, Vancouver, ISO, and other styles
23

Bahrami, Homayoon, and Mansour Zahedi. "Kinetics and mechanism of the oxidation of L-α-amino-n-butyric acid in moderately concentrated sulfuric acid medium." Canadian Journal of Chemistry 82, no. 3 (March 1, 2004): 430–36. http://dx.doi.org/10.1139/v03-181.

Full text
Abstract:
The reaction kinetics of L-α-amino-n-butyric acid oxidation by permanganate ion have been investigated in moderately strong acid medium using a spectrophotometric technique. In all cases studied an autocatalytic effect, due to Mn+2 ions formed as a reaction product, was observed. Both catalytic and noncatalytic processes were determined to be first order with respect to the permanganate ion and the amino acid. The overall rate equation for this process may be written as: [Formula: see text] where k′1 and k′2 are pseudo-order rate constants for the noncatalytic and catalytic reactions, respectively. The influence of some factors such as temperature and reactant concentration on the rate constants has been studied, and the activation parameters have been calculated. Reaction mechanisms satisfying observations for both catalytic and noncatalytic routes have been presented.Key words: L-α-amino-n-butyric acid, permanganate oxidation, concentrated acidic medium, autocatalysis, free radical intermediates.
APA, Harvard, Vancouver, ISO, and other styles
24

Kay, Steven R. "Global bifurcations in a simple, autocatalytic reaction." Dynamics and Stability of Systems 5, no. 3 (January 1990): 137–48. http://dx.doi.org/10.1080/02681119008806092.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Šnita, D., H. Ševčíková, M. Marek, and J. H. Merkin. "Ionic Autocatalytic Reaction Fronts in Electric Fields." Journal of Physical Chemistry 100, no. 48 (January 1996): 18740–48. http://dx.doi.org/10.1021/jp9613302.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Carey, Michael R., Stephen W. Morris, and Paul Kolodner. "Convective fingering of an autocatalytic reaction front." Physical Review E 53, no. 6 (June 1, 1996): 6012–15. http://dx.doi.org/10.1103/physreve.53.6012.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Suresh, A. K., T. Sridhar, and O. E. Potter. "Autocatalytic oxidation of cyclohexane—modeling reaction kinetics." AIChE Journal 34, no. 1 (January 1988): 69–80. http://dx.doi.org/10.1002/aic.690340109.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Aukrust, T., D. A. Browne, and I. Webman. "Critical behavior of an autocatalytic reaction model." Physical Review A 41, no. 10 (May 1, 1990): 5294–301. http://dx.doi.org/10.1103/physreva.41.5294.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Sashuk, Volodymyr, Helena Butkiewicz, Marcin Fiałkowski, and Oksana Danylyuk. "Triggering autocatalytic reaction by host–guest interactions." Chemical Communications 52, no. 22 (2016): 4191–94. http://dx.doi.org/10.1039/c5cc10063a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Das, Debojyoti, Pushpita Ghosh, and Deb Shankar Ray. "Marginal states in a cubic autocatalytic reaction." Journal of Chemical Physics 135, no. 12 (September 28, 2011): 124104. http://dx.doi.org/10.1063/1.3640012.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Filisetti, Alessandro, Alex Graudenzi, Roberto Serra, Marco Villani, Rudolf M. Füchslin, Norman Packard, Stuart A. Kauffman, and Irene Poli. "A stochastic model of autocatalytic reaction networks." Theory in Biosciences 131, no. 2 (October 7, 2011): 85–93. http://dx.doi.org/10.1007/s12064-011-0136-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Hordijk, Wim, Jonathan Naylor, Natalio Krasnogor, and Harold Fellermann. "Population Dynamics of Autocatalytic Sets in a Compartmentalized Spatial World." Life 8, no. 3 (August 18, 2018): 33. http://dx.doi.org/10.3390/life8030033.

Full text
Abstract:
Autocatalytic sets are self-sustaining and collectively catalytic chemical reaction networks which are believed to have played an important role in the origin of life. Simulation studies have shown that autocatalytic sets are, in principle, evolvable if multiple autocatalytic subsets can exist in different combinations within compartments, i.e., so-called protocells. However, these previous studies have so far not explicitly modeled the emergence and dynamics of autocatalytic sets in populations of compartments in a spatial environment. Here, we use a recently developed software tool to simulate exactly this scenario, as an important first step towards more realistic simulations and experiments on autocatalytic sets in protocells.
APA, Harvard, Vancouver, ISO, and other styles
33

Boman, Bruce M., Arthur Guetter, Ryan M. Boman, and Olaf A. Runquist. "Autocatalytic Tissue Polymerization Reaction Mechanism in Colorectal Cancer Development and Growth." Cancers 12, no. 2 (February 17, 2020): 460. http://dx.doi.org/10.3390/cancers12020460.

Full text
Abstract:
The goal of our study was to measure the kinetics of human colorectal cancer (CRC) development in order to identify aberrant mechanisms in tissue dynamics and processes that contribute to colon tumorigenesis. The kinetics of tumor development were investigated using age-at-tumor diagnosis (adenomas and CRCs) of familial adenomatous coli (FAP) patients and sporadic CRC patients. Plots of age-at-tumor diagnosis data as a function of age showed a distinct sigmoidal-shaped curve that is characteristic of an autocatalytic reaction. Consequently, we performed logistics function analysis and found an excellent fit (p < 0.05) of the logistic equation to the curves for age-at-tumor diagnoses. These findings indicate that the tissue mechanism that becomes altered in CRC development and growth involves an autocatalytic reaction. We conjecture that colonic epithelium normally functions as a polymer of cells which dynamically maintains itself in a steady state through an autocatalytic polymerization mechanism. Further, in FAP and sporadic CRC patients, mutation in the adenomatous polyposis coli (APC) gene increases autocatalytic tissue polymerization and induces tumor tissues to autocatalyze their own progressive growth, which drives tumor development in the colon.
APA, Harvard, Vancouver, ISO, and other styles
34

Jiang, Jifa, and Junping Shi. "Dynamics of a reaction-diffusion system of autocatalytic chemical reaction." Discrete & Continuous Dynamical Systems - A 21, no. 1 (2008): 245–58. http://dx.doi.org/10.3934/dcds.2008.21.245.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Tóth-Szeles, Eszter, Ágota Tóth, and Dezső Horváth. "Diffusive fingering in a precipitation reaction driven by autocatalysis." Chem. Commun. 50, no. 42 (2014): 5580–82. http://dx.doi.org/10.1039/c4cc00142g.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

López, Mar, Valentín Santos, and Juan Carlos Parajó. "Autocatalytic Fractionation of Wood Hemicelluloses: Modeling of Multistage Operation." Catalysts 10, no. 3 (March 17, 2020): 337. http://dx.doi.org/10.3390/catal10030337.

Full text
Abstract:
Eucalyptus globulus wood samples were treated with hot, compressed water (autohydrolysis) in consecutive stages under non-isothermal conditions in order to convert the hemicellulose fraction into soluble compounds through reactions catalyzed by in situ generated acids. The first stage was a conventional autohydrolysis, and liquid phase obtained under conditions leading to an optimal recovery of soluble saccharides was employed in a new reaction (second crossflow stage) using a fresh wood lot, in order to increase the concentrations of soluble saccharides. In the third crossflow stage, the best liquid phase from the second stage was employed to solubilize the hemicelluloses from a fresh wood lot. The concentration profiles determined for the soluble saccharides, acids, and furans present in the liquid phases from the diverse crossflow stages were employed for kinetic modeling, based on pseudohomogeneous reactions and Arrhenius-type dependence of the kinetic coefficients on temperature. Additional characterization of the reaction products by High Pressure Size Exclusion Chromatography, High Performance Anion Exchange Chromatography with Pulsed Amperometric Detection, and Matrix Assisted Laser Desorption/Ionization Time of Flight Mass Spectrometry provided further insight on the properties of the soluble saccharides present in the various reaction media.
APA, Harvard, Vancouver, ISO, and other styles
37

Vasil’eva, Olga. "Computer Research of Building Materials." E3S Web of Conferences 97 (2019): 02011. http://dx.doi.org/10.1051/e3sconf/20199702011.

Full text
Abstract:
A computer study of the technological processes for the production of building materials based on the use of reversible autocatalysis reactions is carried out. For the mathematical description of the process of autocatalytic reactions, mathematical models are used, described by a initial-boundary problems for the Carleman system. The considered mathematical model describes the autocatalysis reaction under the following assumptions: the rate of decrease in the density of positive ions is proportional to the square of the deviation of the density of positive ions from the equilibrium state, the rate of decrease in the density of negative ions is proportional to the square of the deviation of the density of negative ions from the equilibrium position and the law of conservation of the total charge of the system including liquid solution. This mathematical model is non-linear, therefore numerical methods are used for its study. For a numerical solution of the initial-boundary value problem for the Carleman system, a finite-difference method of the second order of accuracy is used. For the numerical investigation of the solution obtained, linear interpolation and spline interpolation are used. A numerical study is made on the dependence of the chemical reaction time from the parameters of the model and various initial conditions.
APA, Harvard, Vancouver, ISO, and other styles
38

Ortiz de la Plata, Guadalupe B., Orlando M. Alfano, and Alberto E. Cassano. "The heterogeneous photo-Fenton reaction using goethite as catalyst." Water Science and Technology 61, no. 12 (June 1, 2010): 3109–16. http://dx.doi.org/10.2166/wst.2010.258.

Full text
Abstract:
In the present work the degradation of 2-chlorophenol (2-CP) used as model compound, applying the Heterogeneous photo-Fenton reaction, was studied. Small particles of goethite or iron oxyhydroxide were used as a source of iron. The influence of catalyst loading, radiation intensity and the molar ratio between hydrogen peroxide and contaminant were examined. Improvement by illumination is highly significant. During the progress of 2-CP degradation, the reaction shows an unusual acceleration. This autocatalytic comportment, with stronger tendencies at higher temperatures, implies a completely different behaviour from the one typically expected. The autocatalytic performance is successfully explained by the joint action of two factors: (i) the evolution of the available iron in the homogeneous phase during the course of the reaction and (ii) the autocatalytic contribution of some of the reaction intermediates in the iron cycle. The small iron concentration leaching into the solution is produced by two typical liquid medium – solid goethite surface dissolution processes. A reaction mechanism has been proposed and, in a first stage, parameters have been obtained for the dark reaction. In a second step, the complete data for the irradiated operation were obtained.
APA, Harvard, Vancouver, ISO, and other styles
39

Schmitz, Guy, and Henri Rooze. "Mécanisme des réactions du chlorite et du dioxyde de chlore. 5. Cinétique de la réaction chlorite–bromure." Canadian Journal of Chemistry 65, no. 3 (March 1, 1987): 497–501. http://dx.doi.org/10.1139/v87-086.

Full text
Abstract:
In acidic solutions of chlorite and bromide two processes occur simultaneously, the disproportionation of chlorite and an autocatalytic reaction leading to a rapid production of ClO2. The unusual features of this reaction are described. In this complex system the rate of HClO2 + Br− + H+ → HClO + HBrO cannot be measured. It can be, however, with added ortho-tolidine to remove the HBrO and HClO. We obtained r = k[HClO2][Br−][H+] with k = 1.48 × 10−2 M−2 s−1 at 25 °C. Without added ortho-tolidine this reaction initiates the autocatalytic reaction for which we suggest a kinetic scheme.
APA, Harvard, Vancouver, ISO, and other styles
40

Ke, Wei, Guang-Jin Chen, and Daoyi Chen. "Methane–propane hydrate formation and memory effect study with a reaction kinetics model." Progress in Reaction Kinetics and Mechanism 45 (January 2020): 146867832090162. http://dx.doi.org/10.1177/1468678320901622.

Full text
Abstract:
Although natural gas hydrates and hydrate exploration have been extensively studied for decades, the reaction kinetics and nucleation mechanism of hydrate formation is not fully understood. In its early stage, gas hydrate formation can be assumed to be an autocatalytic kinetic reaction with nucleation and initial growth. In this work, a reaction kinetics model has been established to form structure II methane–propane hydrate in an isochoric reactor. The computational model consists of six pseudo-elementary reactions for three dynamic processes: (1) gas dissolution into the bulk liquid, (2) a slow buildup of hydrate precursors for nucleation onset, and (3) rapid and autocatalytic hydrate growth after onset. The model was programmed using FORTRAN, with initiating parameters and rate constants that were derived or obtained from data fitted using experimental results. The simulations indicate that the length of nucleation induction is determined largely by an accumulation of oligomeric hydrate precursors up to a threshold value. The slow accumulation of precursors is the rate-limiting step for the overall hydrate formation, and its conversion into hydrate particles is critical for the rapid, autocatalytic reaction. By applying this model, the memory effect for hydrate nucleation was studied by assigning varied initial amounts of precursor or hydrate species in the simulations. The presence of pre-existing precursors or hydrate particles could facilitate the nucleation stage with a reduced induction time, and without affecting hydrate growth. The computational model with the performed simulations provides insight into the reaction kinetics and nucleation mechanism of hydrate formation.
APA, Harvard, Vancouver, ISO, and other styles
41

Chandler, W. David, Zhao Wang, and Donald G. Lee. "Kinetics and mechanism of the oxidation of alcohols by tetrapropylammonium perruthenate." Canadian Journal of Chemistry 83, no. 9 (September 1, 2005): 1212–21. http://dx.doi.org/10.1139/v05-114.

Full text
Abstract:
2-Propanol is oxidized by tetrapropylammonium perruthenate (TPAP) in a reaction that is second order in TPAP and first order in 2-propanol. One of the products, believed to be ruthenium dioxide, is an effective catalyst for the reaction, making it an autocatalytic process. The rate of oxidation is relatively insensitive to the presence of substituents. Primary kinetic deuterium isotope effects are observed when either the hydroxyl or the α hydrogen is replaced by deuterium. The only product obtained from the oxidation of cyclobutanol is cyclobutanone, indicating that the reaction is a two-electron process. Tetrahydrofuran is oxidized at a rate that is several orders of magnitude slower than that observed for 2-propanol, suggesting that the reaction of TPAP with alcohols may be initiated by formation of perruthenate esters. A tentative mechanism consistent with these observations is proposed.Key words: oxidation, alcohols, tetrapropylammonium perruthenate, reaction mechanism, autocatalysis.
APA, Harvard, Vancouver, ISO, and other styles
42

Rusheng, Li. "AN AUTOCATALYTIC REACTION MODEL FOR BIRHYTHUM AND TRISTABILITY." Acta Physico-Chimica Sinica 4, no. 06 (1988): 657–59. http://dx.doi.org/10.3866/pku.whxb19880618.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

STADLER, B., and P. STADLER. "Small autocatalytic reaction networks—III. Monotone growth functions." Bulletin of Mathematical Biology 53, no. 3 (1991): 469–85. http://dx.doi.org/10.1016/s0092-8240(05)80399-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Niu, Hong, Qingling Zhang, Chunyu Yang, and Fenglan Bai. "Variable structure control for three-variable autocatalytic reaction." Journal of Control Theory and Applications 11, no. 3 (July 4, 2013): 393–400. http://dx.doi.org/10.1007/s11768-013-2044-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Alhumaizi, Khalid. "Chaotic behavior of an autocatalytic reaction with mutation." Chaos, Solitons & Fractals 11, no. 8 (June 2000): 1279–86. http://dx.doi.org/10.1016/s0960-0779(99)00037-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Malham, I. Bou, N. Jarrige, J. Martin, N. Rakotomalala, L. Talon, and D. Salin. "Lock-exchange experiments with an autocatalytic reaction front." Journal of Chemical Physics 133, no. 24 (December 28, 2010): 244505. http://dx.doi.org/10.1063/1.3507899.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Kato, Jun, Toshio Kubota, and Shukuro Igarashi. "Length Detection‐Flow Analytical System Using Autocatalytic Reaction." Analytical Letters 38, no. 14 (November 2005): 2431–37. http://dx.doi.org/10.1080/00032710500318122.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Chan, Tzu-Hsin, Po-Tuan Chen, Hsuan-Hau Chang, Ming-Yu Lai, Michitoshi Hayashi, Juen-Kai Wang, and Yuh-Lin Wang. "Autocatalytic Reaction in Hydrolysis of Difructose Anhydride III." Journal of Physical Chemistry A 115, no. 37 (September 22, 2011): 10309–14. http://dx.doi.org/10.1021/jp206494r.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Asakura, K., K. Kurihara, A. Ikumo, A. Tanaka, T. Miura, T. Ozawa, Y. Kushibe, S. Osanai, and D. K. Kondepudi. "Chirally autocatalytic reaction performed in highly supersaturated conditions." Macromolecular Symposia 160, no. 1 (October 2000): 7–14. http://dx.doi.org/10.1002/1521-3900(200010)160:1<7::aid-masy7>3.0.co;2-h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Liao, Xiaoli, Jing Su, and Milan Mrksich. "An Adaptor Domain-Mediated Autocatalytic Interfacial Kinase Reaction." Chemistry - A European Journal 15, no. 45 (November 16, 2009): 12303–9. http://dx.doi.org/10.1002/chem.200901345.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography