Law of mass action

Last updated

In chemistry, the law of mass action is the proposition that the rate of a chemical reaction is directly proportional to the product of the activities or concentrations of the reactants. [1] 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. [2]

Contents

Two aspects are involved in the initial formulation of the law: 1) the equilibrium aspect, concerning the composition of a reaction mixture at equilibrium and 2) the kinetic aspect concerning the rate equations for elementary reactions. Both aspects stem from the research performed by Cato M. Guldberg and Peter Waage between 1864 and 1879 in which equilibrium constants were derived by using kinetic data and the rate equation which they had proposed. Guldberg and Waage also recognized that chemical equilibrium is a dynamic process in which rates of reaction for the forward and backward reactions must be equal at chemical equilibrium. In order to derive the expression of the equilibrium constant appealing to kinetics, the expression of the rate equation must be used. The expression of the rate equations was rediscovered independently by Jacobus Henricus van 't Hoff.

The law is a statement about equilibrium and gives an expression for the equilibrium constant, a quantity characterizing chemical equilibrium. In modern chemistry this is derived using equilibrium thermodynamics. It can also be derived with the concept of chemical potential. [3]

History

Two chemists generally expressed the composition of a mixture in terms of numerical values relating the amount of the product to describe the equilibrium state. Cato Maximilian Guldberg and Peter Waage, building on Claude Louis Berthollet's ideas [4] [5] about reversible chemical reactions, proposed the law of mass action in 1864. [6] [7] [8] These papers, in Danish, went largely unnoticed, as did the later publication (in French) of 1867 which contained a modified law and the experimental data on which that law was based. [9] [10]

In 1877 van 't Hoff independently came to similar conclusions, [11] [12] but was unaware of the earlier work, which prompted Guldberg and Waage to give a fuller and further developed account of their work, in German, in 1879. [13] Van 't Hoff then accepted their priority.

1864

The equilibrium state (composition)

In their first paper, [6] Guldberg and Waage suggested that in a reaction such as

the "chemical affinity" or "reaction force" between A and B did not just depend on the chemical nature of the reactants, as had previously been supposed, but also depended on the amount of each reactant in a reaction mixture. Thus the law of mass action was first stated as follows:

When two reactants, A and B, react together at a given temperature in a "substitution reaction," the affinity, or chemical force between them, is proportional to the active masses, [A] and [B], each raised to a particular power
.

In this context a substitution reaction was one such as . Active mass was defined in the 1879 paper as "the amount of substance in the sphere of action". [14] For species in solution active mass is equal to concentration. For solids, active mass is taken as a constant. , a and b were regarded as empirical constants, to be determined by experiment.

At equilibrium, the chemical force driving the forward reaction must be equal to the chemical force driving the reverse reaction. Writing the initial active masses of A,B, A' and B' as p, q, p' and q' and the dissociated active mass at equilibrium as , this equality is represented by

represents the amount of reagents A and B that has been converted into A' and B'. Calculations based on this equation are reported in the second paper. [7]

Dynamic approach to the equilibrium state

The third paper of 1864 [8] was concerned with the kinetics of the same equilibrium system. Writing the dissociated active mass at some point in time as x, the rate of reaction was given as

Likewise the reverse reaction of A' with B' proceeded at a rate given by

The overall rate of conversion is the difference between these rates, so at equilibrium (when the composition stops changing) the two rates of reaction must be equal. Hence

...

1867

The rate expressions given in Guldberg and Waage's 1864 paper could not be differentiated, so they were simplified as follows. [10] The chemical force was assumed to be directly proportional to the product of the active masses of the reactants.

This is equivalent to setting the exponents a and b of the earlier theory to one. The proportionality constant was called an affinity constant, k. The equilibrium condition for an "ideal" reaction was thus given the simplified form

[A]eq, [B]eq etc. are the active masses at equilibrium. In terms of the initial amounts reagents p,q etc. this becomes

The ratio of the affinity coefficients, k'/k, can be recognized as an equilibrium constant. Turning to the kinetic aspect, it was suggested that the velocity of reaction, v, is proportional to the sum of chemical affinities (forces). In its simplest form this results in the expression

where is the proportionality constant. Actually, Guldberg and Waage used a more complicated expression which allowed for interaction between A and A', etc. By making certain simplifying approximations to those more complicated expressions, the rate equation could be integrated and hence the equilibrium quantity could be calculated. The extensive calculations in the 1867 paper gave support to the simplified concept, namely,

The rate of a reaction is proportional to the product of the active masses of the reagents involved.

This is an alternative statement of the law of mass action.

1879

In the 1879 paper [13] the assumption that reaction rate was proportional to the product of concentrations was justified microscopically in terms of the frequency of independent collisions, as had been developed for gas kinetics by Boltzmann in 1872 (Boltzmann equation). It was also proposed that the original theory of the equilibrium condition could be generalised to apply to any arbitrary chemical equilibrium.

The exponents α, β etc. are explicitly identified for the first time as the stoichiometric coefficients for the reaction.

Modern statement of the law

The affinity constants, k+ and k, of the 1879 paper can now be recognised as rate constants. The equilibrium constant, K, was derived by setting the rates of forward and backward reactions to be equal. This also meant that the chemical affinities for the forward and backward reactions are equal. The resultant expression

is correct [2] even from the modern perspective, apart from the use of concentrations instead of activities (the concept of chemical activity was developed by Josiah Willard Gibbs, in the 1870s, but was not widely known in Europe until the 1890s). The derivation from the reaction rate expressions is no longer considered to be valid. Nevertheless, Guldberg and Waage were on the right track when they suggested that the driving force for both forward and backward reactions is equal when the mixture is at equilibrium. The term they used for this force was chemical affinity. Today the expression for the equilibrium constant is derived by setting the chemical potential of forward and backward reactions to be equal. The generalisation of the law of mass action, in terms of affinity, to equilibria of arbitrary stoichiometry was a bold and correct conjecture.

The hypothesis that reaction rate is proportional to reactant concentrations is, strictly speaking, only true for elementary reactions (reactions with a single mechanistic step), but the empirical rate expression

is also applicable to second order reactions that may not be concerted reactions. Guldberg and Waage were fortunate in that reactions such as ester formation and hydrolysis, on which they originally based their theory, do indeed follow this rate expression.

In general many reactions occur with the formation of reactive intermediates, and/or through parallel reaction pathways. However, all reactions can be represented as a series of elementary reactions and, if the mechanism is known in detail, the rate equation for each individual step is given by the expression so that the overall rate equation can be derived from the individual steps. When this is done the equilibrium constant is obtained correctly from the rate equations for forward and backward reaction rates.

In biochemistry, there has been significant interest in the appropriate mathematical model for chemical reactions occurring in the intracellular medium. This is in contrast to the initial work done on chemical kinetics, which was in simplified systems where reactants were in a relatively dilute, pH-buffered, aqueous solution. In more complex environments, where bound particles may be prevented from disassociation by their surroundings, or diffusion is slow or anomalous, the model of mass action does not always describe the behavior of the reaction kinetics accurately. Several attempts have been made to modify the mass action model, but consensus has yet to be reached. Popular modifications replace the rate constants with functions of time and concentration. As an alternative to these mathematical constructs, one school of thought is that the mass action model can be valid in intracellular environments under certain conditions, but with different rates than would be found in a dilute, simple environment [ citation needed ].

The fact that Guldberg and Waage developed their concepts in steps from 1864 to 1867 and 1879 has resulted in much confusion in the literature as to which equation the law of mass action refers. It has been a source of some textbook errors. [15] Thus, today the "law of mass action" sometimes refers to the (correct) equilibrium constant formula, [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] and at other times to the (usually incorrect) rate formula. [26] [27]

Applications to other fields

In semiconductor physics

The law of mass action also has implications in semiconductor physics. Regardless of doping, the product of electron and hole densities is a constant at equilibrium. This constant depends on the thermal energy of the system (i.e. the product of the Boltzmann constant, , and temperature, ), as well as the band gap (the energy separation between conduction and valence bands, ) and effective density of states in the valence and conduction bands. When the equilibrium electron and hole densities are equal, their density is called the intrinsic carrier density as this would be the value of and in a perfect crystal. Note that the final product is independent of the Fermi level :

Diffusion in condensed matter

Yakov Frenkel represented diffusion process in condensed matter as an ensemble of elementary jumps and quasichemical interactions of particles and defects. Henry Eyring applied his theory of absolute reaction rates to this quasichemical representation of diffusion. Mass action law for diffusion leads to various nonlinear versions of Fick's law. [28]

In mathematical ecology

The Lotka–Volterra equations describe dynamics of the predator-prey systems. The rate of predation upon the prey is assumed to be proportional to the rate at which the predators and the prey meet; this rate is evaluated as xy, where x is the number of prey, y is the number of predator. This is a typical example of the law of mass action.

In mathematical epidemiology

The law of mass action forms the basis of the compartmental model of disease spread in mathematical epidemiology, in which a population of humans, animals or other individuals is divided into categories of susceptible, infected, and recovered (immune). The principle of mass action is at the heart of the transmission term of compartmental models in epidemiology, which provide a useful abstraction of disease dynamics. [29] The law of mass action formulation of the SIR model corresponds to the following "quasichemical" system of elementary reactions:

The list of components is S (susceptible individuals), I (infected individuals), and R (removed individuals, or just recovered ones if we neglect lethality);
The list of elementary reactions is
.
If the immunity is unstable then the transition from R to S should be added that closes the cycle (SIRS model):
.

A rich system of law of mass action models was developed in mathematical epidemiology by adding components and elementary reactions.

Individuals in human or animal populations  unlike molecules in an ideal solution  do not mix homogeneously. There are some disease examples in which this non-homogeneity is great enough such that the outputs of the classical SIR model and their simple generalizations like SIS or SEIR, are invalid. For these situations, more sophisticated compartmental models or distributed reaction-diffusion models may be useful.

In sociophysics

Sociophysics [30] uses tools and concepts from physics and physical chemistry to describe some aspects of social and political behavior. It attempts to explain why and how humans behave much like atoms, at least in some aspects of their collective lives. The law of mass action (generalized if it is necessary) is the main tool to produce the equation of interactions of humans in sociophysics.

See also

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.

Chemical thermodynamics is the study of the interrelation of heat and work with chemical reactions or with physical changes of state within the confines of the laws of thermodynamics. Chemical thermodynamics involves not only laboratory measurements of various thermodynamic properties, but also the application of mathematical methods to the study of chemical questions and the spontaneity of processes.

In physical chemistry, the Arrhenius equation is a formula for the temperature dependence of reaction rates. The equation was proposed by Svante Arrhenius in 1889, based on the work of Dutch chemist Jacobus Henricus van 't Hoff who had noted in 1884 that the van 't Hoff equation for the temperature dependence of equilibrium constants suggests such a formula for the rates of both forward and reverse reactions. This equation has a vast and important application in determining the rate of chemical reactions and for calculation of energy of activation. Arrhenius provided a physical justification and interpretation for the formula. Currently, it is best seen as an empirical relationship. It can be used to model the temperature variation of diffusion coefficients, population of crystal vacancies, creep rates, and many other thermally induced processes and reactions. The Eyring equation, developed in 1935, also expresses the relationship between rate and energy.

In chemistry, a dynamic equilibrium exists once a reversible reaction occurs. Substances transition between the reactants and products at equal rates, meaning there is no net change. Reactants and products are formed at such a rate that the concentration of neither changes. It is a particular example of a system in a steady state.

<span class="mw-page-title-main">Gibbs free energy</span> Type of thermodynamic potential

In thermodynamics, the Gibbs free energy is a thermodynamic potential that can be used to calculate the maximum amount of work, other than pressure-volume work, that may be performed by a thermodynamically closed system at constant temperature and pressure. It also provides a necessary condition for processes such as chemical reactions that may occur under these conditions. The Gibbs free energy is expressed as

<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.

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.

A reversible reaction is a reaction in which the conversion of reactants to products and the conversion of products to reactants occur simultaneously.

The equilibrium constant of a chemical reaction is the value of its reaction quotient at chemical equilibrium, a state approached by a dynamic chemical system after sufficient time has elapsed at which its composition has no measurable tendency towards further change. For a given set of reaction conditions, the equilibrium constant is independent of the initial analytical concentrations of the reactant and product species in the mixture. Thus, given the initial composition of a system, known equilibrium constant values can be used to determine the composition of the system at equilibrium. However, reaction parameters like temperature, solvent, and ionic strength may all influence the value of the equilibrium constant.

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

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.

In physics, a mass balance, also called a material balance, is an application of conservation of mass to the analysis of physical systems. By accounting for material entering and leaving a system, mass flows can be identified which might have been unknown, or difficult to measure without this technique. The exact conservation law used in the analysis of the system depends on the context of the problem, but all revolve around mass conservation, i.e., that matter cannot disappear or be created spontaneously.

The Van 't Hoff equation relates the change in the equilibrium constant, Keq, of a chemical reaction to the change in temperature, T, given the standard enthalpy change, ΔrH, for the process. The subscript means "reaction" and the superscript means "standard". It was proposed by Dutch chemist Jacobus Henricus van 't Hoff in 1884 in his book Études de Dynamique chimique.

An elementary reaction is a chemical reaction in which one or more chemical species react directly to form products in a single reaction step and with a single transition state. In practice, a reaction is assumed to be elementary if no reaction intermediates have been detected or need to be postulated to describe the reaction on a molecular scale. An apparently elementary reaction may be in fact a stepwise reaction, i.e. a complicated sequence of chemical reactions, with reaction intermediates of variable lifetimes.

<span class="mw-page-title-main">Transition state theory</span> Theory describing the reaction rates of elementary chemical reactions

In chemistry, transition state theory (TST) explains the reaction rates of elementary chemical reactions. The theory assumes a special type of chemical equilibrium (quasi-equilibrium) between reactants and activated transition state complexes.

Equilibrium constants are determined in order to quantify chemical equilibria. When an equilibrium constant K is expressed as a concentration quotient,

In electrochemistry, the Butler–Volmer equation, also known as Erdey-Grúz–Volmer equation, is one of the most fundamental relationships in electrochemical kinetics. It describes how the electrical current through an electrode depends on the voltage difference between the electrode and the bulk electrolyte for a simple, unimolecular redox reaction, considering that both a cathodic and an anodic reaction occur on the same electrode:

The rate of a chemical reaction is influenced by many different factors, such as temperature, pH, reactant, and product concentrations and other effectors. The degree to which these factors change the reaction rate is described by the elasticity coefficient. This coefficient is defined as follows:

<span class="mw-page-title-main">Langmuir adsorption model</span> Model describing the adsorption of a mono-layer of gas molecules on an ideal flat surface

The Langmuir adsorption model explains adsorption by assuming an adsorbate behaves as an ideal gas at isothermal conditions. According to the model, adsorption and desorption are reversible processes. This model even explains the effect of pressure i.e. at these conditions the adsorbate's partial pressure, , is related to the volume of it, V, adsorbed onto a solid adsorbent. The adsorbent, as indicated in the figure, is assumed to be an ideal solid surface composed of a series of distinct sites capable of binding the adsorbate. The adsorbate binding is treated as a chemical reaction between the adsorbate gaseous molecule and an empty sorption site, S. This reaction yields an adsorbed species with an associated equilibrium constant :

Equilibrium chemistry is concerned with systems in chemical equilibrium. The unifying principle is that the free energy of a system at equilibrium is the minimum possible, so that the slope of the free energy with respect to the reaction coordinate is zero. This principle, applied to mixtures at equilibrium provides a definition of an equilibrium constant. Applications include acid–base, host–guest, metal–complex, solubility, partition, chromatography and redox equilibria.

References

  1. Péter Érdi; János Tóth (1989). Mathematical Models of Chemical Reactions: Theory and Applications of Deterministic and Stochastic Models. Manchester University Press. p. 3. ISBN   978-0-7190-2208-1.
  2. 1 2 Chieh, Chung. "Chemical Equilibria - The Law of Mass Action". Chemical reactions, chemical equilibria, and electrochemistry. Archived from the original on 3 October 2018. Retrieved 21 July 2019. The law of mass action is universal, applicable under any circumstance... The mass action law states that if the system is at equilibrium at a given temperature, then the following ratio is a constant.
  3. Föll, Helmut. "Mass Action Law". Defects in Crystals.
  4. Berthollet, C.L. (1803). Essai de statique chimique [Essay on chemical statics [i.e., equilibrium]] (in French). Paris, France: Firmin Didot. On pp. 404–407, Berthollet mentions that when he accompanied Napoleon on his expedition to Egypt, he (Berthollet) visited Lake Natron and found sodium carbonate along its shores. He realized that this was a product of the reverse of the usual reaction Na2CO3 + CaCl2 → 2NaCl + CaCO3↓ and therefore that the final state of a reaction was a state of equilibrium between two opposing processes. From p. 405: " … la décomposition du muriate de soude continue donc jusqu'à ce qu'il se soit formé assez de muriate de chaux, parce que l'acide muriatique devant se partager entre les deux bases en raison de leur action, il arrive un terme où leurs forces se balancent." ( … the decomposition of the sodium chloride thus continues until enough calcium chloride is formed, because the hydrochloric acid must be shared between the two bases in the ratio of their action [i.e., capacity to react]; it reaches an end [point] at which their forces are balanced.)
  5. Levere, Trevor, H. (1971). Affinity and Matter – Elements of Chemical Philosophy 1800–1865. Gordon and Breach Science Publishers. ISBN   2-88124-583-8.{{cite book}}: CS1 maint: multiple names: authors list (link)
  6. 1 2 Waage, P.; Guldberg, C.M. (1864). "Studier over Affiniteten" [Studies of affinities]. Forhandlinger I Videnskabs-selskabet I Christiania (Transactions of the Scientific Society in Christiania) (in Danish): 35–45.
  7. 1 2 Waage, P. (1864). "Forsøg til Bestemmelse af Lovene for Affiniteten" [Experiment for the determination of the laws of affinity]. Forhandlinger I Videnskabs-selskabet I Christiania (Transactions of the Scientific Society in Christiania) (in Danish): 92–94.
  8. 1 2 Guldberg, C.M. (1864). "Foredrag om Lovene for Affiniteten, specielt Tidens Indflydelse paa de kemiske Processer" [Lecture on the laws of affinity, especially the influence of time on chemical processes]. Forhandlinger I Videnskabs-selskabet I Christiania (Transactions of the Scientific Society in Christiania) (in Danish): 111–120.
  9. Guldberg, C.M.; Waage, P. (1867). Études sur les affinités chimiques [Studies of chemical affinities] (in French). Christiania [Oslo], Norway: Brøgger & Christie.
  10. 1 2 C.M. Guldberg and P. Waage, "Experiments concerning Chemical Affinity"; German translation by Abegg in Ostwalds Klassiker der Exacten Wissenschaften, no. 104, Wilhelm Engleman, Leipzig, 1899, pp 10-125
  11. McLean, Franklin C. (1938). "Application of the Law of Chemical Equilibrium (Law of Mass Action) to Biological Problems" (PDF). Physiological Reviews. 18 (4): 495–523. doi:10.1152/physrev.1938.18.4.495.
  12. van 't Hoff, J.H. (1877). "Die Grenzebene, ein Beitrag zur Kenntniss der Esterbildung" [The limit plane: a contribution to our knowledge of ester formation]. Berichte der Deutschen Chemischen Gesellschaft zu Berlin (in German). 10: 669–678. doi:10.1002/cber.187701001185.
  13. 1 2 Guldberg, C.M.; Waage, P. (1879). "Ueber die chemische Affinität" [On chemical affinity]. Journal für praktische Chemie. 2nd series (in German). 19: 69–114. doi:10.1002/prac.18790190111. Reprinted, with comments by Abegg in Ostwalds Klassiker der Exacten Wissenschaften, no. 104, Wilhelm Engleman, Leipzig, 1899, pp 126-171
  14. (Guldberg & Waage, 1879), p. 71: "Eigentlich verstehen wir unter der activen Masse nur die Menge des Stoffes innerhalb der Actionsphäre; unter sonst gleichen Umständen kann aber die Actionsphäre durch die Volumeneinheit repräsentirt werden." (Actually we understand by "active mass" only the quantity of substance inside the sphere of action; under otherwise identical conditions, however, the sphere of action can be represented by the unit volume.)
  15. Guggenheim, E.A. (1956). "Textbook errors IX: More About the Laws of Reaction Rates and of Equilibrium". Journal of Chemical Education. 33 (11): 544–545. doi:10.1021/ed033p544.
  16. Law of Mass Action
  17. SC.edu
  18. The Law of Mass Action
  19. Leech, Mary. "Acids and Bases" (PDF). Geochemistry — GEOL 480. San Francisco State University. Archived from the original (PDF) on 2006-09-21.
  20. "Recap of Fundamental Acid-Base Concepts". Chemistry 152. Washington University in St. Louis. Archived from the original on 2012-02-06.
  21. Chemical Equilibria: Basic Concepts
  22. "Chemical equilibrium - The Law of Mass Action". Archived from the original on 2010-02-01. Retrieved 2007-07-12.
  23. "Indiana.edu" (PDF). Archived from the original (PDF) on 2007-06-15. Retrieved 2007-07-12.
  24. Berkeley.edu
  25. General Chemistry Online: FAQ: Acids and bases: What is the pH at the equivalence point an HF/NaOH titration?
  26. law of mass action definition
  27. Lab 4 – Slow Manifolds Archived 2007-11-17 at the Wayback Machine
  28. A.N. Gorban, H.P. Sargsyan and H.A. Wahab (2011). "Quasichemical Models of Multicomponent Nonlinear Diffusion". Mathematical Modelling of Natural Phenomena. 6 (5): 184–262. arXiv: 1012.2908 . doi: 10.1051/mmnp/20116509 . S2CID   18961678.
  29. von Csefalvay, Chris (2023), "Simple compartmental models", Computational Modeling of Infectious Disease, Elsevier, pp. 19–91, doi:10.1016/b978-0-32-395389-4.00011-6, ISBN   978-0-323-95389-4 , retrieved 2023-03-02
  30. S. Galam, Sociophysics. A Physicist's Modeling of Psycho-political Phenomena, Springer, 2012, DOI: 10.1007/978-1-4614-2032-3, ISBN   978-1-4614-2032-3

Further reading