Chemical reaction network theory

Last updated

Chemical reaction network theory is an area of applied mathematics that attempts to model the behaviour of real-world chemical systems. Since its foundation in the 1960s, it has attracted a growing research community, mainly due to its applications in biochemistry and theoretical chemistry. It has also attracted interest from pure mathematicians due to the interesting problems that arise from the mathematical structures involved.

Contents

History

Dynamical properties of reaction networks were studied in chemistry and physics after the invention of the law of mass action. The essential steps in this study were introduction of detailed balance for the complex chemical reactions by Rudolf Wegscheider (1901), [1] development of the quantitative theory of chemical chain reactions by Nikolay Semyonov (1934), [2] development of kinetics of catalytic reactions by Cyril Norman Hinshelwood, [3] and many other results.

Three eras of chemical dynamics can be revealed in the flux of research and publications. [4] These eras may be associated with leaders: the first is the van 't Hoff era, the second may be called the SemenovHinshelwood era and the third is definitely the Aris era. The "eras" may be distinguished based on the main focuses of the scientific leaders:

The mathematical discipline "chemical reaction network theory" was originated by Rutherford Aris, a famous expert in chemical engineering, with the support of Clifford Truesdell, the founder and editor-in-chief of the journal Archive for Rational Mechanics and Analysis . The paper of R. Aris in this journal [5] was communicated to the journal by C. Truesdell. It opened the series of papers of other authors (which were communicated already by R. Aris). The well known papers of this series are the works of Frederick J. Krambeck, [6] Roy Jackson, Friedrich Josef Maria Horn, [7] Martin Feinberg [8] and others, published in the 1970s. In his second "prolegomena" paper, [9] R. Aris mentioned the work of N.Z. Shapiro, L.S. Shapley (1965), [10] where an important part of his scientific program was realized.

Since then, the chemical reaction network theory has been further developed by a large number of researchers internationally. [11] [12] [13] [14] [15] [16] [17] [18] [19] [20]

Overview

A chemical reaction network (often abbreviated to CRN) comprises a set of reactants, a set of products (often intersecting the set of reactants), and a set of reactions. For example, the pair of combustion reactions

 

 

 

 

(reaction 1)

form a reaction network. The reactions are represented by the arrows. The reactants appear to the left of the arrows, in this example they are (hydrogen), (oxygen) and C (carbon). The products appear to the right of the arrows, here they are (water) and (carbon dioxide). In this example, since the reactions are irreversible and neither of the products are used in the reactions, the set of reactants and the set of products are disjoint.

Mathematical modelling of chemical reaction networks usually focuses on what happens to the concentrations of the various chemicals involved as time passes. Following the example above, let a represent the concentration of in the surrounding air, b represent the concentration of , c represent the concentration of , and so on. Since all of these concentrations will not in general remain constant, they can be written as a function of time e.g. , etc.

These variables can then be combined into a vector

and their evolution with time can be written

This is an example of a continuous autonomous dynamical system, commonly written in the form . The number of molecules of each reactant used up each time a reaction occurs is constant, as is the number of molecules produced of each product. These numbers are referred to as the stoichiometry of the reaction, and the difference between the two (i.e. the overall number of molecules used up or produced) is the net stoichiometry. This means that the equation representing the chemical reaction network can be rewritten as

Here, each column of the constant matrix represents the net stoichiometry of a reaction, and so is called the stoichiometry matrix. is a vector-valued function where each output value represents a reaction rate, referred to as the kinetics.

Common assumptions

For physical reasons, it is usually assumed that reactant concentrations cannot be negative, and that each reaction only takes place if all its reactants are present, i.e. all have non-zero concentration. For mathematical reasons, it is usually assumed that is continuously differentiable.

It is also commonly assumed that no reaction features the same chemical as both a reactant and a product (i.e. no catalysis or autocatalysis), and that increasing the concentration of a reactant increases the rate of any reactions that use it up. This second assumption is compatible with all physically reasonable kinetics, including mass action, Michaelis–Menten and Hill kinetics. Sometimes further assumptions are made about reaction rates, e.g. that all reactions obey mass action kinetics.

Other assumptions include mass balance, constant temperature, constant pressure, spatially uniform concentration of reactants, and so on.

Types of results

As chemical reaction network theory is a diverse and well-established area of research, there is a significant variety of results. Some key areas are outlined below.

Number of steady states

These results relate to whether a chemical reaction network can produce significantly different behaviour depending on the initial concentrations of its constituent reactants. This has applications in e.g. modelling biological switchesa high concentration of a key chemical at steady state could represent a biological process being "switched on" whereas a low concentration would represent being "switched off".

For example, the catalytic trigger is the simplest catalytic reaction without autocatalysis that allows multiplicity of steady states (1976): [21] [22]

 

 

 

 

(reaction 2)

 

 

 

 

(reaction 3)

 

 

 

 

(reaction 4)

This is the classical adsorption mechanism of catalytic oxidation.

Here, and are gases (for example, , and ), is the "adsorption place" on the surface of the solid catalyst (for example, ), and are the intermediates on the surface (adatoms, adsorbed molecules or radicals). This system may have two stable steady states of the surface for the same concentrations of the gaseous components.

Stability of steady states

Stability determines whether a given steady state solution is likely to be observed in reality. Since real systems (unlike deterministic models) tend to be subject to random background noise, an unstable steady state solution is unlikely to be observed in practice. Instead of them, stable oscillations or other types of attractors may appear.

Persistence

Persistence has its roots in population dynamics. A non-persistent species in population dynamics can go extinct for some (or all) initial conditions. Similar questions are of interests to chemists and biochemists, i.e. if a given reactant was present to start with, can it ever be completely used up?

Existence of stable periodic solutions

Results regarding stable periodic solutions attempt to rule out "unusual" behaviour. If a given chemical reaction network admits a stable periodic solution, then some initial conditions will converge to an infinite cycle of oscillating reactant concentrations. For some parameter values it may even exhibit quasiperiodic or chaotic behaviour. While stable periodic solutions are unusual in real-world chemical reaction networks, well-known examples exist, such as the Belousov–Zhabotinsky reactions. The simplest catalytic oscillator (nonlinear self-oscillations without autocatalysis) can be produced from the catalytic trigger by adding a "buffer" step. [23]

 

 

 

 

(reaction 5)

where (BZ) is an intermediate that does not participate in the main reaction.

Network structure and dynamical properties

One of the main problems of chemical reaction network theory is the connection between network structure and properties of dynamics. This connection is important even for linear systems, for example, the simple cycle with equal interaction weights has the slowest decay of the oscillations among all linear systems with the same number of states. [24]

For nonlinear systems, many connections between structure and dynamics have been discovered. First of all, these are results about stability. [25] For some classes of networks, explicit construction of Lyapunov functions is possible without apriori assumptions about special relations between rate constants. Two results of this type are well known: the deficiency zero theorem [26] and the theorem about systems without interactions between different components. [27]

The deficiency zero theorem gives sufficient conditions for the existence of the Lyapunov function in the classical free energy form , where is the concentration of the i-th component. The theorem about systems without interactions between different components states that if a network consists of reactions of the form (for , where r is the number of reactions, is the symbol of ith component, , and are non-negative integers) and allows the stoichiometric conservation law (where all ), then the weighted L1 distance between two solutions with the same M(c) monotonically decreases in time.

Model reduction

Modelling of large reaction networks meets various difficulties: the models include too many unknown parameters and high dimension makes the modelling computationally expensive. The model reduction methods were developed together with the first theories of complex chemical reactions. [28] Three simple basic ideas have been invented:

The quasi-equilibrium approximation and the quasi steady state methods were developed further into the methods of slow invariant manifolds and computational singular perturbation. The methods of limiting steps gave rise to many methods of the analysis of the reaction graph. [28]

Related Research Articles

In a chemical reaction, chemical equilibrium is the state in which both the reactants and products are present in concentrations which have no further tendency to change with time, so that there is no observable change in the properties of the system. This state results when the forward reaction proceeds at the same rate as the reverse reaction. The reaction rates of the forward and backward reactions are generally not zero, but they are equal. Thus, there are no net changes in the concentrations of the reactants and products. Such a state is known as dynamic equilibrium.

<span class="mw-page-title-main">Half-life</span> Time for exponential decay to remove half of a quantity

Half-life is the time required for a quantity to reduce to half of its initial value. The term is commonly used in nuclear physics to describe how quickly unstable atoms undergo radioactive decay or how long stable atoms survive. The term is also used more generally to characterize any type of exponential decay. For example, the medical sciences refer to the biological half-life of drugs and other chemicals in the human body. The converse of half-life is doubling time.

<span class="mw-page-title-main">Activation energy</span> Minimum amount of energy that must be provided for a system to undergo a reaction or process

In the Arrhenius model of reaction rates, activation energy is the minimum amount of energy that must be available to reactants for a chemical reaction to occur. The activation energy (Ea) of a reaction is measured in kilojoules per mole (kJ/mol) or kilocalories per mole (kcal/mol). Activation energy can be thought of as the magnitude of the potential barrier (sometimes called the energy barrier) separating minima of the potential energy surface pertaining to the initial and final thermodynamic state. For a chemical reaction to proceed at a reasonable rate, the temperature of the system should be high enough such that there exists an appreciable number of molecules with translational energy equal to or greater than the activation energy. The term "activation energy" was introduced in 1889 by the Swedish scientist Svante Arrhenius.

<span class="mw-page-title-main">Reaction rate</span> Speed at which a chemical reaction takes place

The reaction rate or rate of reaction is the speed at which a chemical reaction takes place, defined as proportional to the increase in the concentration of a product per unit time and to the decrease in the concentration of a reactant per unit time. Reaction rates can vary dramatically. For example, the oxidative rusting of iron under Earth's atmosphere is a slow reaction that can take many years, but the combustion of cellulose in a fire is a reaction that takes place in fractions of a second. For most reactions, the rate decreases as the reaction proceeds. A reaction's rate can be determined by measuring the changes in concentration over time.

<span class="mw-page-title-main">Michaelis–Menten kinetics</span> Model of enzyme kinetics

In biochemistry, Michaelis–Menten kinetics, named after Leonor Michaelis and Maud Menten, is the simplest case of enzyme kinetics, applied to enzyme-catalysed reactions of one substrate and one product. It takes the form of an equation describing the reaction rate to , the concentration of the substrate A. Its formula is given by the Michaelis–Menten equation:

In chemistry, the law of mass action is the proposition that the rate of the chemical reaction is directly proportional to the product of the activities or concentrations of the reactants. It explains and predicts behaviors of solutions in dynamic equilibrium. Specifically, it implies that for a chemical reaction mixture that is in equilibrium, the ratio between the concentration of reactants and products is constant.

Chemical kinetics, also known as reaction kinetics, is the branch of physical chemistry that is concerned with understanding the rates of chemical reactions. It is different from chemical thermodynamics, which deals with the direction in which a reaction occurs but in itself tells nothing about its rate. Chemical kinetics includes investigations of how experimental conditions influence the speed of a chemical reaction and yield information about the reaction's mechanism and transition states, as well as the construction of mathematical models that also can describe the characteristics of a chemical reaction.

In chemical kinetics, the overall rate of a reaction is often approximately determined by the slowest step, known as the rate-determining step or rate-limiting step. For a given reaction mechanism, the prediction of the corresponding rate equation is often simplified by using this approximation of the rate-determining step.

In chemical kinetics, a reaction rate constant or reaction rate coefficient is a proportionality constant which quantifies the rate and direction of a chemical reaction by relating it with the concentration of reactants.

In chemistry, the rate equation is an empirical differential mathematical expression for the reaction rate of a given reaction in terms of concentrations of chemical species and constant parameters only. For many reactions, the initial rate is given by a power law such as

In chemistry, molecularity is the number of molecules that come together to react in an elementary (single-step) reaction and is equal to the sum of stoichiometric coefficients of reactants in the elementary reaction with effective collision and correct orientation. Depending on how many molecules come together, a reaction can be unimolecular, bimolecular or even trimolecular.

In chemistry, a steady state is a situation in which all state variables are constant in spite of ongoing processes that strive to change them. For an entire system to be at steady state, i.e. for all state variables of a system to be constant, there must be a flow through the system. A simple example of such a system is the case of a bathtub with the tap running but with the drain unplugged: after a certain time, the water flows in and out at the same rate, so the water level stabilizes and the system is in a steady state.

The principle of detailed balance can be used in kinetic systems which are decomposed into elementary processes. It states that at equilibrium, each elementary process is in equilibrium with its reverse process.

<span class="mw-page-title-main">Continuous stirred-tank reactor</span> Type of chemical reactor

The continuous stirred-tank reactor (CSTR), also known as vat- or backmix reactor, mixed flow reactor (MFR), or a continuous-flow stirred-tank reactor (CFSTR), is a common model for a chemical reactor in chemical engineering and environmental engineering. A CSTR often refers to a model used to estimate the key unit operation variables when using a continuous agitated-tank reactor to reach a specified output. The mathematical model works for all fluids: liquids, gases, and slurries.

<span class="mw-page-title-main">Enzyme kinetics</span> Study of biochemical reaction rates catalysed by an enzyme

Enzyme kinetics is the study of the rates of enzyme-catalysed chemical reactions. In enzyme kinetics, the reaction rate is measured and the effects of varying the conditions of the reaction are investigated. Studying an enzyme's kinetics in this way can reveal the catalytic mechanism of this enzyme, its role in metabolism, how its activity is controlled, and how a drug or a modifier might affect the rate.

In chemical kinetics, the Lindemann mechanism is a schematic reaction mechanism for unimolecular reactions. Frederick Lindemann and J. A. Christiansen proposed the concept almost simultaneously in 1921, and Cyril Hinshelwood developed it to take into account the energy distributed among vibrational degrees of freedom for some reaction steps.

This bibliography of Rutherford Aris contains a comprehensive listing of the scientific publications of Aris, including books, journal articles, and contributions to other published material.

Transient kinetic isotope effects occur when the reaction leading to isotope fractionation does not follow pure first-order kinetics and therefore isotopic effects cannot be described with the classical equilibrium fractionation equations or with steady-state kinetic fractionation equations. In these instances, the general equations for biochemical isotope kinetics (GEBIK) and the general equations for biochemical isotope fractionation (GEBIF) can be used.

<span class="mw-page-title-main">Grigoriy Yablonsky</span>

Grigoriy Yablonsky is an expert in the area of chemical kinetics and chemical engineering, particularly in catalytic technology of complete and selective oxidation, which is one of the main driving forces of sustainable development.

<span class="mw-page-title-main">Alexander Gorban</span> Russian-British scientist (born 1952)

Alexander Nikolaevich Gorban is a scientist of Russian origin, working in the United Kingdom. He is a professor at the University of Leicester, and director of its Mathematical Modeling Centre. Gorban has contributed to many areas of fundamental and applied science, including statistical physics, non-equilibrium thermodynamics, machine learning and mathematical biology.

References

  1. Wegscheider, R. (1901) Über simultane Gleichgewichte und die Beziehungen zwischen Thermodynamik und Reactionskinetik homogener Systeme, Monatshefte für Chemie / Chemical Monthly 32(8), 849--906.
  2. Semyonov's Nobel Lecture Some Problems Relating to Chain Reactions and to the Theory of Combustion
  3. Hinshelwood's Nobel Lecture Chemical Kinetics in the Past Few Decades
  4. A.N. Gorban, G.S. Yablonsky Three Waves of Chemical Dynamics, Mathematical Modelling of Natural Phenomena 10(5) (2015), 1–5.
  5. R. Aris, Prolegomena to the rational analysis of systems of chemical reactions, Archive for Rational Mechanics and Analysis, 1965, Volume 19, Issue 2, pp 81-99.
  6. F.J. Krambeck, The mathematical structure of chemical kinetics in homogeneous single-phase systems, Archive for Rational Mechanics and Analysis, 1970, Volume 38, Issue 5, pp 317-347,
  7. F. J. M. Horn and R. Jackson, "General Mass Action Kinetics", Archive Rational Mech., 47:81, 1972.
  8. M. Feinberg, "Complex balancing in general kinetic systems", Arch. Rational Mech. Anal. , 49:187–194, 1972.
  9. R. Aris, Prolegomena to the rational analysis of systems of chemical reactions II. Some addenda, Archive for Rational Mechanics and Analysis, 1968, Volume 27, Issue 5, pp 356-364
  10. N.Z. Shapiro, L.S. Shapley, Mass action law and the Gibbs free energy function, SIAM J. Appl. Math. 16 (1965) 353–375.
  11. P. Érdi and J. Tóth, "Mathematical models of chemical reactions", Manchester University Press , 1989.
  12. H. Kunze and D. Siegel, "Monotonicity properties of chemical reactions with a single initial bimolecular step", J. Math. Chem., 31(4):339–344, 2002.
  13. M. Mincheva and D. Siegel, "Nonnegativity and positiveness of solutions to mass action reaction–diffusion systems", J. Math. Chem., 42:1135–1145, 2007.
  14. P. De Leenheer, D. Angeli and E. D. Sontag, "Monotone chemical reaction networks" Archived 2014-08-12 at the Wayback Machine , J. Math. Chem.', 41(3):295–314, 2007.
  15. M. Banaji, P. Donnell and S. Baigent, "P matrix properties, injectivity and stability in chemical reaction systems", SIAM J. Appl. Math., 67(6):1523–1547, 2007.
  16. G. Craciun and C. Pantea, "Identifiability of chemical reaction networks", J. Math. Chem., 44:1, 2008.
  17. M. Domijan and M. Kirkilionis, "Bistability and oscillations in chemical reaction networks", J. Math. Biol. , 59(4):467–501, 2009.
  18. A. N. Gorban and G. S. Yablonsky, "Extended detailed balance for systems with irreversible reactions", Chemical Engineering Science, 66:5388–5399, 2011.
  19. E. Feliu, M. Knudsen and C. Wiuf., "Signaling cascades: Consequences of varying substrate and phosphatase levels", Adv. Exp. Med. Biol. (Adv Syst Biol), 736:81–94, 2012.
  20. I. Otero-Muras, J. R. Banga and A. A. Alonso, "Characterizing multistationarity regimes in biochemical reaction networks", PLoS ONE,7(7):e39194,2012.
  21. M.G. Slin'ko, V.I. Bykov, G.S. Yablonskii, T.A. Akramov, "Multiplicity of the Steady State in Heterogeneous Catalytic Reactions", Dokl. Akad. Nauk SSSR226 (4) (1976), 876.
  22. V.I. Bykov, V.I. Elokhin, G.S. Yablonskii, "The simplest catalytic mechanism permitting several steady states of the surface", React. Kinet. Catal. Lett.4 (2) (1976), 191–198.
  23. V.I. Bykov, G.S. Yablonskii, V.F. Kim, "On the simple model of kinetic self-oscillations in catalytic reaction of CO oxidation", Doklady AN USSR (Chemistry) 242 (3) (1978), 637–639.
  24. A.N. Gorban, N. Jarman, E. Steur, C. van Leeuwen, I.Yu. Tyukin, Leaders do not Look Back, or do They? Math. Model. Nat. Phenom. Vol. 10, No. 3, 2015, pp. 212–231.
  25. B.L. Clarke, Theorems on chemical network stability. The Journal of Chemical Physics. 1975, 62(3), 773-775.
  26. M. Feinberg, Chemical reaction network structure and the stability of complex isothermal reactors—I. The deficiency zero and deficiency one theorems. Chemical Engineering Science. 1987 31, 42(10), 2229-2268.
  27. A.N. Gorban, V.I. Bykov, G.S. Yablonskii, Thermodynamic function analogue for reactions proceeding without interaction of various substances, Chemical Engineering Science, 1986 41(11), 2739-2745.
  28. 1 2 A.N.Gorban, Model reduction in chemical dynamics: slow invariant manifolds, singular perturbations, thermodynamic estimates, and analysis of reaction graph. Current Opinion in Chemical Engineering 2018 21C, 48-59.