Price equation

Last updated

In the theory of evolution and natural selection, the Price equation (also known as Price's equation or Price's theorem) describes how a trait or allele changes in frequency over time. The equation uses a covariance between a trait and fitness, to give a mathematical description of evolution and natural selection. It provides a way to understand the effects that gene transmission and natural selection have on the frequency of alleles within each new generation of a population. The Price equation was derived by George R. Price, working in London to re-derive W.D. Hamilton's work on kin selection. Examples of the Price equation have been constructed for various evolutionary cases. The Price equation also has applications in economics. [1]

Contents

It is important to note that the Price equation is not a physical or biological law. It is not a concise or general expression of experimentally validated results. It is rather a purely mathematical relationship between various statistical descriptors of population dynamics. It is mathematically valid, and therefore not subject to experimental verification. In simple terms, it is a mathematical restatement of the expression "survival of the fittest" which is actually self-evident, given the mathematical definitions of "survival" and "fittest".

Statement

Example for a trait under positive selection Example of Price equation for a trait under positive selection.png
Example for a trait under positive selection

The Price equation shows that a change in the average amount of a trait in a population from one generation to the next () is determined by the covariance between the amounts of the trait for subpopulation and the fitnesses of the subpopulations, together with the expected change in the amount of the trait value due to fitness, namely :

Here is the average fitness over the population, and and represent the population mean and covariance respectively. 'Fitness' is the ratio of the average number of offspring for the whole population per the number of adult individuals in the population, and is that same ratio only for subpopulation .

If the covariance between fitness () and trait value () is positive, the trait value is expected to rise on average across population . If the covariance is negative, the characteristic is harmful, and its frequency is expected to drop.

The second term, , represents the portion of due to all factors other than direct selection which can affect trait evolution. This term can encompass genetic drift, mutation bias, or meiotic drive. Additionally, this term can encompass the effects of multi-level selection or group selection. Price (1972) referred to this as the "environment change" term, and denoted both terms using partial derivative notation (∂NS and ∂EC). This concept of environment includes interspecies and ecological effects. Price describes this as follows:

Fisher adopted the somewhat unusual point of view of regarding dominance and epistasis as being environment effects. For example, he writes (1941): ‘A change in the proportion of any pair of genes itself constitutes a change in the environment in which individuals of the species find themselves.’ Hence he regarded the natural selection effect on M as being limited to the additive or linear effects of changes in gene frequencies, while everything else – dominance, epistasis, population pressure, climate, and interactions with other species – he regarded as a matter of the environment.

G.R. Price (1972), Fisher's fundamental theorem made clear [2]

Proof

Suppose we are given four equal-length lists of real numbers [3] , , , from which we may define . and will be called the parent population numbers and characteristics associated with each index i. Likewise and will be called the child population numbers and characteristics, and will be called the fitness associated with index i. (Equivalently, we could have been given , , , with .) Define the parent and child population totals:

and the probabilities (or frequencies): [4]

Note that these are of the form of probability mass functions in that and are in fact the probabilities that a random individual drawn from the parent or child population has a characteristic . Define the fitnesses:

The average of any list is given by:

so the average characteristics are defined as:

and the average fitness is:

A simple theorem can be proved: so that:

and

The covariance of and is defined by:

Defining , the expectation value of is

The sum of the two terms is:

Using the above mentioned simple theorem, the sum becomes

where .

Derivation of the continuous-time Price equation

Consider a set of groups with that are characterized by a particular trait, denoted by . The number of individuals belonging to group experiences exponential growth:

where corresponds to the fitness of the group. We want to derive an equation describing the time-evolution of the expected value of the trait:

Based on the chain rule, we may derive an ordinary differential equation:

A further application of the chain rule for gives us:

Summing up the components gives us that:

which is also known as the replicator equation. Now, note that:

Therefore, putting all of these components together, we arrive at the continuous-time Price equation:

Simple Price equation

When the characteristic values do not change from the parent to the child generation, the second term in the Price equation becomes zero resulting in a simplified version of the Price equation:

which can be restated as:

where is the fractional fitness: .

This simple Price equation can be proven using the definition in Equation (2) above. It makes this fundamental statement about evolution: "If a certain inheritable characteristic is correlated with an increase in fractional fitness, the average value of that characteristic in the child population will be increased over that in the parent population."

Applications

The Price equation can describe any system that changes over time, but is most often applied in evolutionary biology. The evolution of sight provides an example of simple directional selection. The evolution of sickle cell anemia shows how a heterozygote advantage can affect trait evolution. The Price equation can also be applied to population context dependent traits such as the evolution of sex ratios. Additionally, the Price equation is flexible enough to model second order traits such as the evolution of mutability. The Price equation also provides an extension to Founder effect which shows change in population traits in different settlements

Dynamical sufficiency and the simple Price equation

Sometimes the genetic model being used encodes enough information into the parameters used by the Price equation to allow the calculation of the parameters for all subsequent generations. This property is referred to as dynamical sufficiency. For simplicity, the following looks at dynamical sufficiency for the simple Price equation, but is also valid for the full Price equation.

Referring to the definition in Equation (2), the simple Price equation for the character can be written:

For the second generation:

The simple Price equation for only gives us the value of for the first generation, but does not give us the value of and , which are needed to calculate for the second generation. The variables and can both be thought of as characteristics of the first generation, so the Price equation can be used to calculate them as well:

The five 0-generation variables , , , , and must be known before proceeding to calculate the three first generation variables , , and , which are needed to calculate for the second generation. It can be seen that in general the Price equation cannot be used to propagate forward in time unless there is a way of calculating the higher moments and from the lower moments in a way that is independent of the generation. Dynamical sufficiency means that such equations can be found in the genetic model, allowing the Price equation to be used alone as a propagator of the dynamics of the model forward in time.

Full Price equation

The simple Price equation was based on the assumption that the characters do not change over one generation. If it is assumed that they do change, with being the value of the character in the child population, then the full Price equation must be used. A change in character can come about in a number of ways. The following two examples illustrate two such possibilities, each of which introduces new insight into the Price equation.

Genotype fitness

We focus on the idea of the fitness of the genotype. The index indicates the genotype and the number of type genotypes in the child population is:

which gives fitness:

Since the individual mutability does not change, the average mutabilities will be:

with these definitions, the simple Price equation now applies.

Lineage fitness

In this case we want to look at the idea that fitness is measured by the number of children an organism has, regardless of their genotype. Note that we now have two methods of grouping, by lineage, and by genotype. It is this complication that will introduce the need for the full Price equation. The number of children an -type organism has is:

which gives fitness:

We now have characters in the child population which are the average character of the -th parent.

with global characters:

with these definitions, the full Price equation now applies.

Criticism

The use of the change in average characteristic () per generation as a measure of evolutionary progress is not always appropriate. There may be cases where the average remains unchanged (and the covariance between fitness and characteristic is zero) while evolution is nevertheless in progress. For example, if we have , , and , then for the child population, showing that the peak fitness at is in fact fractionally increasing the population of individuals with . However, the average characteristics are z=2 and z'=2 so that . The covariance is also zero. The simple Price equation is required here, and it yields 0=0. In other words, it yields no information regarding the progress of evolution in this system.

A critical discussion of the use of the Price equation can be found in van Veelen (2005), [5] van Veelen et al. (2012), [6] and van Veelen (2020). [7] Frank (2012) discusses the criticism in van Veelen et al. (2012). [8]

Cultural references

Price's equation features in the plot and title of the 2008 thriller film WΔZ .

The Price equation also features in posters in the computer game BioShock 2 , in which a consumer of a "Brain Boost" tonic is seen deriving the Price equation while simultaneously reading a book. The game is set in the 1950s, substantially before Price's work.

See also

Related Research Articles

<span class="mw-page-title-main">Dirac delta function</span> Generalized function whose value is zero everywhere except at zero

In mathematical analysis, the Dirac delta function, also known as the unit impulse, is a generalized function on the real numbers, whose value is zero everywhere except at zero, and whose integral over the entire real line is equal to one. Since there is no function having this property, to model the delta "function" rigorously involves the use of limits or, as is common in mathematics, measure theory and the theory of distributions.

Covariance in probability theory and statistics is a measure of the joint variability of two random variables.

<span class="mw-page-title-main">Covariance matrix</span> Measure of covariance of components of a random vector

In probability theory and statistics, a covariance matrix is a square matrix giving the covariance between each pair of elements of a given random vector.

<span class="mw-page-title-main">Helmholtz free energy</span> Thermodynamic potential

In thermodynamics, the Helmholtz free energy is a thermodynamic potential that measures the useful work obtainable from a closed thermodynamic system at a constant temperature (isothermal). The change in the Helmholtz energy during a process is equal to the maximum amount of work that the system can perform in a thermodynamic process in which temperature is held constant. At constant temperature, the Helmholtz free energy is minimized at equilibrium.

In physics, an operator is a function over a space of physical states onto another space of physical states. The simplest example of the utility of operators is the study of symmetry. Because of this, they are useful tools in classical mechanics. Operators are even more important in quantum mechanics, where they form an intrinsic part of the formulation of the theory.

In mathematics, the Hodge star operator or Hodge star is a linear map defined on the exterior algebra of a finite-dimensional oriented vector space endowed with a nondegenerate symmetric bilinear form. Applying the operator to an element of the algebra produces the Hodge dual of the element. This map was introduced by W. V. D. Hodge.

<span class="mw-page-title-main">Path integral formulation</span> Formulation of quantum mechanics

The path integral formulation is a description in quantum mechanics that generalizes the stationary action principle of classical mechanics. It replaces the classical notion of a single, unique classical trajectory for a system with a sum, or functional integral, over an infinity of quantum-mechanically possible trajectories to compute a quantum amplitude.

In mathematics, spectral theory is an inclusive term for theories extending the eigenvector and eigenvalue theory of a single square matrix to a much broader theory of the structure of operators in a variety of mathematical spaces. It is a result of studies of linear algebra and the solutions of systems of linear equations and their generalizations. The theory is connected to that of analytic functions because the spectral properties of an operator are related to analytic functions of the spectral parameter.

<span class="mw-page-title-main">Equipartition theorem</span> Theorem in classical statistical mechanics

In classical statistical mechanics, the equipartition theorem relates the temperature of a system to its average energies. The equipartition theorem is also known as the law of equipartition, equipartition of energy, or simply equipartition. The original idea of equipartition was that, in thermal equilibrium, energy is shared equally among all of its various forms; for example, the average kinetic energy per degree of freedom in translational motion of a molecule should equal that in rotational motion.

<span class="mw-page-title-main">Correlation function (quantum field theory)</span> Expectation value of time-ordered quantum operators

In quantum field theory, correlation functions, often referred to as correlators or Green's functions, are vacuum expectation values of time-ordered products of field operators. They are a key object of study in quantum field theory where they can be used to calculate various observables such as S-matrix elements. They are closely related to correlation functions between random variables, although they are nonetheless different objects, being defined in Minkowski spacetime and on quantum operators.

Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator lowers the number of particles in a given state by one. A creation operator increases the number of particles in a given state by one, and it is the adjoint of the annihilation operator. In many subfields of physics and chemistry, the use of these operators instead of wavefunctions is known as second quantization. They were introduced by Paul Dirac.

In quantum mechanics, the canonical commutation relation is the fundamental relation between canonical conjugate quantities. For example,

In the theory of stochastic processes, the Karhunen–Loève theorem, also known as the Kosambi–Karhunen–Loève theorem states that a stochastic process can be represented as an infinite linear combination of orthogonal functions, analogous to a Fourier series representation of a function on a bounded interval. The transformation is also known as Hotelling transform and eigenvector transform, and is closely related to principal component analysis (PCA) technique widely used in image processing and in data analysis in many fields.

<span class="mw-page-title-main">Schwinger–Dyson equation</span> Equations for correlation functions in QFT

The Schwinger–Dyson equations (SDEs) or Dyson–Schwinger equations, named after Julian Schwinger and Freeman Dyson, are general relations between correlation functions in quantum field theories (QFTs). They are also referred to as the Euler–Lagrange equations of quantum field theories, since they are the equations of motion corresponding to the Green's function. They form a set of infinitely many functional differential equations, all coupled to each other, sometimes referred to as the infinite tower of SDEs.

In theoretical physics, the Weinberg–Witten (WW) theorem, proved by Steven Weinberg and Edward Witten, states that massless particles (either composite or elementary) with spin j > 1/2 cannot carry a Lorentz-covariant current, while massless particles with spin j > 1 cannot carry a Lorentz-covariant stress-energy. The theorem is usually interpreted to mean that the graviton (j = 2) cannot be a composite particle in a relativistic quantum field theory.

<span class="mw-page-title-main">Fundamental thermodynamic relation</span>

In thermodynamics, the fundamental thermodynamic relation are four fundamental equations which demonstrate how four important thermodynamic quantities depend on variables that can be controlled and measured experimentally. Thus, they are essentially equations of state, and using the fundamental equations, experimental data can be used to determine sought-after quantities like G or H (enthalpy). The relation is generally expressed as a microscopic change in internal energy in terms of microscopic changes in entropy, and volume for a closed system in thermal equilibrium in the following way.

In mathematics, the Skorokhod integral, also named Hitsuda–Skorokhod integral, often denoted , is an operator of great importance in the theory of stochastic processes. It is named after the Ukrainian mathematician Anatoliy Skorokhod and Japanese mathematician Masuyuki Hitsuda. Part of its importance is that it unifies several concepts:

This article relates the Schrödinger equation with the path integral formulation of quantum mechanics using a simple nonrelativistic one-dimensional single-particle Hamiltonian composed of kinetic and potential energy.

An electric dipole transition is the dominant effect of an interaction of an electron in an atom with the electromagnetic field.

In statistical mechanics, the mean squared displacement is a measure of the deviation of the position of a particle with respect to a reference position over time. It is the most common measure of the spatial extent of random motion, and can be thought of as measuring the portion of the system "explored" by the random walker. In the realm of biophysics and environmental engineering, the Mean Squared Displacement is measured over time to determine if a particle is spreading slowly due to diffusion, or if an advective force is also contributing. Another relevant concept, the variance-related diameter, is also used in studying the transportation and mixing phenomena in the realm of environmental engineering. It prominently appears in the Debye–Waller factor and in the Langevin equation.

References

  1. Knudsen, Thorbjørn (2004). "General selection theory and economic evolution: The Price equation and the replicator/interactor distinction". Journal of Economic Methodology. 11 (2): 147–173. doi:10.1080/13501780410001694109. S2CID   154197796 . Retrieved 2011-10-22.
  2. Price, G.R. (1972). "Fisher's "fundamental theorem" made clear". Annals of Human Genetics. 36 (2): 129–140. doi:10.1111/j.1469-1809.1972.tb00764.x. PMID   4656569. S2CID   20757537.
  3. The lists may in fact be members of any field (i.e. a set on which addition, subtraction, multiplication, and division are defined and behave as the corresponding operations on rational and real numbers do
  4. Frank, Steven A. (1995). "George Price's Contributions to Evolutionary Genetics". J. Theor. Biol. 175 (3): 373–388. Bibcode:1995JThBi.175..373F. doi:10.1006/jtbi.1995.0148. PMID   7475081 . Retrieved Mar 19, 2023.
  5. van Veelen, M. (December 2005). "On the use of the Price equation". Journal of Theoretical Biology. 237 (4): 412–426. Bibcode:2005JThBi.237..412V. doi:10.1016/j.jtbi.2005.04.026. PMID   15953618.
  6. van Veelen, M.; García, J.; Sabelis, M.W.; Egas, M. (April 2012). "Group selection and inclusive fitness are not equivalent; the Price equation vs. models and statistics". Journal of Theoretical Biology. 299: 64–80. Bibcode:2012JThBi.299...64V. doi:10.1016/j.jtbi.2011.07.025. PMID   21839750.
  7. van Veelen, M. (March 2020). "The problem with the Price equation". Philosophical Transactions of the Royal Society B. 375 (1797): 1–13. doi: 10.1098/rstb.2019.0355 . PMC   7133513 . PMID   32146887.
  8. Frank, S.A. (2012). "Natural Selection IV: The Price equation". Journal of Evolutionary Biology. 25 (6): 1002–1019. arXiv: 1204.1515 . doi:10.1111/j.1420-9101.2012.02498.x. PMC   3354028 . PMID   22487312.

Further reading