Population dynamics

Last updated

Population dynamics is the type of mathematics used to model and study the size and age composition of populations as dynamical systems.

Contents

History

Population dynamics has traditionally been the dominant branch of mathematical biology, which has a history of more than 220 years, [1] although over the last century the scope of mathematical biology has greatly expanded.[ citation needed ]

The beginning of population dynamics is widely regarded as the work of Malthus, formulated as the Malthusian growth model. According to Malthus, assuming that the conditions (the environment) remain constant ( ceteris paribus ), a population will grow (or decline) exponentially. [2] :18 This principle provided the basis for the subsequent predictive theories, such as the demographic studies such as the work of Benjamin Gompertz [3] and Pierre François Verhulst in the early 19th century, who refined and adjusted the Malthusian demographic model. [4]

A more general model formulation was proposed by F. J. Richards in 1959, [5] further expanded by Simon Hopkins, in which the models of Gompertz, Verhulst and also Ludwig von Bertalanffy are covered as special cases of the general formulation. The Lotka–Volterra predator-prey equations are another famous example, [6] [7] [8] [9] [10] [11] [12] [13] as well as the alternative Arditi–Ginzburg equations. [14] [15]

Logistic function

Simplified population models usually start with four key variables (four demographic processes) including death, birth, immigration, and emigration. Mathematical models used to calculate changes in population demographics and evolution hold the assumption of no external influence. Models can be more mathematically complex where "...several competing hypotheses are simultaneously confronted with the data." [16] For example, in a closed system where immigration and emigration does not take place, the rate of change in the number of individuals in a population can be described as:

where N is the total number of individuals in the specific experimental population being studied, B is the number of births and D is the number of deaths per individual in a particular experiment or model. The algebraic symbols b, d and r stand for the rates of birth, death, and the rate of change per individual in the general population, the intrinsic rate of increase. This formula can be read as the rate of change in the population (dN/dT) is equal to births minus deaths (BD). [2] [13] [17]

Using these techniques, Malthus' population principle of growth was later transformed into a mathematical model known as the logistic equation:

where N is the biomass density, a is the maximum per-capita rate of change, and K is the carrying capacity of the population. The formula can be read as follows: the rate of change in the population (dN/dT) is equal to growth (aN) that is limited by carrying capacity (1 − N/K). From these basic mathematical principles the discipline of population ecology expands into a field of investigation that queries the demographics of real populations and tests these results against the statistical models. The field of population ecology often uses data on life history and matrix algebra to develop projection matrices on fecundity and survivorship. This information is used for managing wildlife stocks and setting harvest quotas. [13] [17]

Intrinsic rate of increase

The rate at which a population increases in size if there are no density-dependent forces regulating the population is known as the intrinsic rate of increase. It is

where the derivative is the rate of increase of the population, N is the population size, and r is the intrinsic rate of increase. Thus r is the maximum theoretical rate of increase of a population per individual – that is, the maximum population growth rate. The concept is commonly used in insect population ecology or management to determine how environmental factors affect the rate at which pest populations increase. See also exponential population growth and logistic population growth. [18]

Epidemiology

Population dynamics overlap with another active area of research in mathematical biology: mathematical epidemiology, the study of infectious disease affecting populations. Various models of viral spread have been proposed and analysed, and provide important results that may be applied to health policy decisions.[ citation needed ]

Geometric populations

Operophtera brumata populations are geometric. Operophtera.brumata.6961.jpg
Operophtera brumata populations are geometric.

The mathematical formula below can used to model geometric populations. Geometric populations grow in discrete reproductive periods between intervals of abstinence, as opposed to populations which grow without designated periods for reproduction. Say that N denotes the number of individuals in each generation of a population that will reproduce. [20]

where Nt is the population size in generation t, and Nt+1 is the population size in the generation directly after Nt; Bt is the sum of births in the population between generations t and t + 1 (i.e. the birth rate); It is the sum of immigrants added to the population between generations; Dt is the sum of deaths between generations (death rate); and Et is the sum of emigrants moving out of the population between generations.

When there is no migration to or from the population,

Assuming in this case that the birth and death rates are constants, then the birth rate minus the death rate equals R, the geometric rate of increase.

where λ = 1 + R is the finite rate of increase.

At t + 1Nt+1 = λNt
At t + 2Nt+2 = λNt+1 = λλNt = λ2Nt
At t + 3Nt+3 = λNt+2 = λλ2Nt = λ3Nt

Therefore:

where λt is the Finite rate of increase raised to the power of the number of generations (e.g. for t + 2 [two generations] → λ2, for t + 1 [one generation] → λ1 = λ, and for t [before any generations - at time zero] → λ0 = 1

Doubling time

G. stearothermophilus has a shorter doubling time (td) than E. coli and N. meningitidis. Growth rates of 2 bacterial species will differ by unexpected orders of magnitude if the doubling times of the 2 species differ by even as little as 10 minutes. In eukaryotes such as animals, fungi, plants, and protists, doubling times are much longer than in bacteria. This reduces the growth rates of eukaryotes in comparison to Bacteria. G. stearothermophilus, E. coli, and N. meningitidis have 20 minute, 30 minute, and 40 minute doubling times under optimal conditions respectively. If bacterial populations could grow indefinitely (which they do not) then the number of bacteria in each species would approach infinity ([?]). However, the percentage of G. stearothermophilus bacteria out of all the bacteria would approach 100% whilst the percentage of E. coli and N. meningitidis combined out of all the bacteria would approach 0%. This graph is a simulation of this hypothetical scenario. In reality, bacterial populations do not grow indefinitely in size and the 3 species require different optimal conditions to bring their doubling times to minima.
Time in minutes
% that is G. stearothermophilus
30
44.4%
60
53.3%
90
64.9%
120
72.7%
-[?]
100%
Time in minutes
% that is E. coli
30
29.6%
60
26.7%
90
21.6%
120
18.2%
-[?]
0.00%
Time in minutes
% that is N. meningitidis
30
25.9%
60
20.0%
90
13.5%
120
9.10%
-[?]
0.00%
Disclaimer: Bacterial populations are logistic instead of geometric. Nevertheless, doubling times are applicable to both types of populations. G. stearothermophilus has a shorter doubling time (td) than E. coli and N. meningitidis.png
G. stearothermophilus has a shorter doubling time (td) than E. coli and N. meningitidis. Growth rates of 2 bacterial species will differ by unexpected orders of magnitude if the doubling times of the 2 species differ by even as little as 10 minutes. In eukaryotes such as animals, fungi, plants, and protists, doubling times are much longer than in bacteria. This reduces the growth rates of eukaryotes in comparison to Bacteria. G. stearothermophilus , E. coli , and N. meningitidis have 20 minute, 30 minute, and 40 minute doubling times under optimal conditions respectively. If bacterial populations could grow indefinitely (which they do not) then the number of bacteria in each species would approach infinity (∞). However, the percentage of G. stearothermophilus bacteria out of all the bacteria would approach 100% whilst the percentage of E. coli and N. meningitidis combined out of all the bacteria would approach 0%. This graph is a simulation of this hypothetical scenario. In reality, bacterial populations do not grow indefinitely in size and the 3 species require different optimal conditions to bring their doubling times to minima.
Time in minutes % that is G. stearothermophilus
3044.4%
6053.3%
9064.9%
12072.7%
→∞100%
Time in minutes % that is E. coli
3029.6%
6026.7%
9021.6%
12018.2%
→∞0.00%
Time in minutes % that is N. meningitidis
3025.9%
6020.0%
9013.5%
1209.10%
→∞0.00%
Disclaimer: Bacterial populations are logistic instead of geometric. Nevertheless, doubling times are applicable to both types of populations.

The doubling time (td) of a population is the time required for the population to grow to twice its size. [24] We can calculate the doubling time of a geometric population using the equation: Nt = λtN0 by exploiting our knowledge of the fact that the population (N) is twice its size (2N) after the doubling time. [20]

The doubling time can be found by taking logarithms. For instance:

Or:

Therefore:

Half-life of geometric populations

The half-life of a population is the time taken for the population to decline to half its size. We can calculate the half-life of a geometric population using the equation: Nt = λtN0 by exploiting our knowledge of the fact that the population (N) is half its size (0.5N) after a half-life. [20]

where t1/2 is the half-life.

The half-life can be calculated by taking logarithms (see above).

Geometric (R) growth constant

where ΔN is the change in population size between two generations (between generation t + 1 and t).

Finite (λ) growth constant

Mathematical relationship between geometric and logistic populations

In geometric populations, R and λ represent growth constants (see 2 and 2.3). In logistic populations however, the intrinsic growth rate, also known as intrinsic rate of increase (r) is the relevant growth constant. Since generations of reproduction in a geometric population do not overlap (e.g. reproduce once a year) but do in an exponential population, geometric and exponential populations are usually considered to be mutually exclusive. [25] However, both sets of constants share the mathematical relationship below. [20]

The growth equation for exponential populations is

where e is Euler's number, a universal constant often applicable in logistic equations, and r is the intrinsic growth rate.

To find the relationship between a geometric population and a logistic population, we assume the Nt is the same for both models, and we expand to the following equality:

Giving us

and

Evolutionary game theory

Evolutionary game theory was first developed by Ronald Fisher in his 1930 article The Genetic Theory of Natural Selection . [26] In 1973 John Maynard Smith formalised a central concept, the evolutionarily stable strategy. [27]

Population dynamics have been used in several control theory applications. Evolutionary game theory can be used in different industrial or other contexts. Industrially, it is mostly used in multiple-input-multiple-output (MIMO) systems, although it can be adapted for use in single-input-single-output (SISO) systems. Some other examples of applications are military campaigns, water distribution, dispatch of distributed generators, lab experiments, transport problems, communication problems, among others.

Oscillatory

Population size in plants experiences significant oscillation due to the annual environmental oscillation. [28] Plant dynamics experience a higher degree of this seasonality than do mammals, birds, or bivoltine insects. [28] When combined with perturbations due to disease, this often results in chaotic oscillations. [28]

The computer game SimCity , Sim Earth and the MMORPG Ultima Online, among others, tried to simulate some of these population dynamics.

See also

Related Research Articles

<span class="mw-page-title-main">Exponential distribution</span> Probability distribution

In probability theory and statistics, the exponential distribution or negative exponential distribution is the probability distribution of the distance between events in a Poisson point process, i.e., a process in which events occur continuously and independently at a constant average rate; the distance parameter could be any meaningful mono-dimensional measure of the process, such as time between production errors, or length along a roll of fabric in the weaving manufacturing process. It is a particular case of the gamma distribution. It is the continuous analogue of the geometric distribution, and it has the key property of being memoryless. In addition to being used for the analysis of Poisson point processes it is found in various other contexts.

<span class="mw-page-title-main">Logistic function</span> S-shaped curve

A logistic function or logistic curve is a common S-shaped curve with the equation

The Black–Scholes or Black–Scholes–Merton model is a mathematical model for the dynamics of a financial market containing derivative investment instruments, using various underlying assumptions. From the parabolic partial differential equation in the model, known as the Black–Scholes equation, one can deduce the Black–Scholes formula, which gives a theoretical estimate of the price of European-style options and shows that the option has a unique price given the risk of the security and its expected return. The equation and model are named after economists Fischer Black and Myron Scholes; Robert C. Merton, who first wrote an academic paper on the subject, is sometimes also credited.

<span class="mw-page-title-main">Exponential growth</span> Growth of quantities at rate proportional to the current amount

Exponential growth is a process that increases quantity over time at an ever-increasing rate. It occurs when the instantaneous rate of change of a quantity with respect to time is proportional to the quantity itself. Described as a function, a quantity undergoing exponential growth is an exponential function of time, that is, the variable representing time is the exponent. Exponential growth is the inverse of logarithmic growth.

<span class="mw-page-title-main">Weibull distribution</span> Continuous probability distribution

In probability theory and statistics, the Weibull distribution is a continuous probability distribution. It models a broad range of random variables, largely in the nature of a time to failure or time between events. Examples are maximum one-day rainfalls and the time a user spends on a web page.

<span class="mw-page-title-main">Exponential decay</span> Decrease in value at a rate proportional to the current value

A quantity is subject to exponential decay if it decreases at a rate proportional to its current value. Symbolically, this process can be expressed by the following differential equation, where N is the quantity and λ (lambda) is a positive rate called the exponential decay constant, disintegration constant, rate constant, or transformation constant:

The Lotka–Volterra equations, also known as the Lotka–Volterra predator–prey model, are a pair of first-order nonlinear differential equations, frequently used to describe the dynamics of biological systems in which two species interact, one as a predator and the other as prey. The populations change through time according to the pair of equations:

<span class="mw-page-title-main">Haldane's dilemma</span> Limit on the speed of beneficial evolution

Haldane's dilemma, also known as the waiting time problem, is a limit on the speed of beneficial evolution, calculated by J. B. S. Haldane in 1957. Before the invention of DNA sequencing technologies, it was not known how much polymorphism DNA harbored, although alloenzymes were beginning to make it clear that substantial polymorphism existed. This was puzzling because the amount of polymorphism known to exist seemed to exceed the theoretical limits that Haldane calculated, that is, the limits imposed if polymorphisms present in the population generally influence an organism's fitness. Motoo Kimura's landmark paper on neutral theory in 1968 built on Haldane's work to suggest that most molecular evolution is neutral, resolving the dilemma. Although neutral evolution remains the consensus theory among modern biologists, and thus Kimura's resolution of Haldane's dilemma is widely regarded as correct, some biologists argue that adaptive evolution explains a large fraction of substitutions in protein coding sequence, and they propose alternative solutions to Haldane's dilemma.

<span class="mw-page-title-main">Basic reproduction number</span> Metric in epidemiology

In epidemiology, the basic reproduction number, or basic reproductive number, denoted , of an infection is the expected number of cases directly generated by one case in a population where all individuals are susceptible to infection. The definition assumes that no other individuals are infected or immunized. Some definitions, such as that of the Australian Department of Health, add the absence of "any deliberate intervention in disease transmission". The basic reproduction number is not necessarily the same as the effective reproduction number , which is the number of cases generated in the current state of a population, which does not have to be the uninfected state. is a dimensionless number and not a time rate, which would have units of time−1, or units of time like doubling time.

Compartmental models are a very general modelling technique. They are often applied to the mathematical modelling of infectious diseases. The population is assigned to compartments with labels – for example, S, I, or R,. People may progress between compartments. The order of the labels usually shows the flow patterns between the compartments; for example SEIS means susceptible, exposed, infectious, then susceptible again.

Variational Bayesian methods are a family of techniques for approximating intractable integrals arising in Bayesian inference and machine learning. They are typically used in complex statistical models consisting of observed variables as well as unknown parameters and latent variables, with various sorts of relationships among the three types of random variables, as might be described by a graphical model. As typical in Bayesian inference, the parameters and latent variables are grouped together as "unobserved variables". Variational Bayesian methods are primarily used for two purposes:

  1. To provide an analytical approximation to the posterior probability of the unobserved variables, in order to do statistical inference over these variables.
  2. To derive a lower bound for the marginal likelihood of the observed data. This is typically used for performing model selection, the general idea being that a higher marginal likelihood for a given model indicates a better fit of the data by that model and hence a greater probability that the model in question was the one that generated the data.

The birth–death process is a special case of continuous-time Markov process where the state transitions are of only two types: "births", which increase the state variable by one and "deaths", which decrease the state by one. It was introduced by William Feller. The model's name comes from a common application, the use of such models to represent the current size of a population where the transitions are literal births and deaths. Birth–death processes have many applications in demography, queueing theory, performance engineering, epidemiology, biology and other areas. They may be used, for example, to study the evolution of bacteria, the number of people with a disease within a population, or the number of customers in line at the supermarket.

The Gompertz curve or Gompertz function is a type of mathematical model for a time series, named after Benjamin Gompertz (1779–1865). It is a sigmoid function which describes growth as being slowest at the start and end of a given time period. The right-side or future value asymptote of the function is approached much more gradually by the curve than the left-side or lower valued asymptote. This is in contrast to the simple logistic function in which both asymptotes are approached by the curve symmetrically. It is a special case of the generalised logistic function. The function was originally designed to describe human mortality, but since has been modified to be applied in biology, with regard to detailing populations.

In population biology and demography, generation time is the average time between two consecutive generations in the lineages of a population. In human populations, generation time typically has ranged from 20 to 30 years, with wide variation based on gender and society. Historians sometimes use this to date events, by converting generations into years to obtain rough estimates of time.

In statistics, a power transform is a family of functions applied to create a monotonic transformation of data using power functions. It is a data transformation technique used to stabilize variance, make the data more normal distribution-like, improve the validity of measures of association, and for other data stabilization procedures.

A local volatility model, in mathematical finance and financial engineering, is an option pricing model that treats volatility as a function of both the current asset level and of time . As such, it is a generalisation of the Black–Scholes model, where the volatility is a constant. Local volatility models are often compared with stochastic volatility models, where the instantaneous volatility is not just a function of the asset level but depends also on a new "global" randomness coming from an additional random component.

<span class="mw-page-title-main">Conway–Maxwell–Poisson distribution</span> Probability distribution

In probability theory and statistics, the Conway–Maxwell–Poisson distribution is a discrete probability distribution named after Richard W. Conway, William L. Maxwell, and Siméon Denis Poisson that generalizes the Poisson distribution by adding a parameter to model overdispersion and underdispersion. It is a member of the exponential family, has the Poisson distribution and geometric distribution as special cases and the Bernoulli distribution as a limiting case.

In the study of age-structured population growth, probably one of the most important equations is the Euler–Lotka equation. Based on the age demographic of females in the population and female births, this equation allows for an estimation of how a population is growing.

Kermack–McKendrick theory is a hypothesis that predicts the number and distribution of cases of an infectious disease as it is transmitted through a population over time. Building on the research of Ronald Ross and Hilda Hudson, A. G. McKendrick and W. O. Kermack published their theory in a set of three articles from 1927, 1932, and 1933. While Kermack–McKendrick theory was indeed the source of SIR models and their relatives, Kermack and McKendrick were thinking of a more subtle and empirically useful problem than the simple compartmental models discussed here. The text is somewhat difficult to read, compared to modern papers, but the important feature is it was a model where the age-of-infection affected the transmission and removal rates.

<span class="mw-page-title-main">Pest insect population dynamics</span>

The population dynamics of pest insects is a subject of interest to farmers, agricultural economists, ecologists, and those concerned with animal welfare.

References

  1. Malthus, Thomas Robert. An Essay on the Principle of Population: Library of Economics
  2. 1 2 Turchin, P. (2001). "Does Population Ecology Have General Laws?". Oikos . John Wiley & Sons Ltd. (Nordic Society Oikos). 94 (1): 17–26. doi:10.1034/j.1600-0706.2001.11310.x. S2CID   27090414.
  3. Gompertz, Benjamin (1825). "On the Nature of the Function Expressive of the Law of Human Mortality, and on a New Mode of Determining the Value of Life Contingencies". Philosophical Transactions of the Royal Society of London . 115: 513–585. doi:10.1098/rstl.1825.0026. S2CID   145157003.
  4. Verhulst, P. H. (1838). "Notice sur la loi que la population poursuit dans son accroissement". Corresp. Mathématique et Physique . 10: 113–121.
  5. Richards, F. J. (June 1959). "A Flexible Growth Function for Empirical Use". Journal of Experimental Botany . 10 (29): 290–300. doi:10.1093/jxb/10.2.290. JSTOR   23686557 . Retrieved 16 November 2020.
  6. Hoppensteadt, F. (2006). "Predator-prey model". Scholarpedia . 1 (10): 1563. Bibcode:2006SchpJ...1.1563H. doi: 10.4249/scholarpedia.1563 .
  7. Lotka, A. J. (1910). "Contribution to the Theory of Periodic Reaction". J. Phys. Chem. 14 (3): 271–274. doi:10.1021/j150111a004.
  8. Goel, N. S.; et al. (1971). On the Volterra and Other Non-Linear Models of Interacting Populations. Academic Press.
  9. Lotka, A. J. (1925). Elements of Physical Biology. Williams and Wilkins.
  10. Volterra, V. (1926). "Variazioni e fluttuazioni del numero d'individui in specie animali conviventi". Mem. Acad. Lincei Roma . 2: 31–113.
  11. Volterra, V. (1931). "Variations and fluctuations of the number of individuals in animal species living together". In Chapman, R. N. (ed.). Animal Ecology. McGraw–Hill.
  12. Kingsland, S. (1995). Modeling Nature: Episodes in the History of Population Ecology. University of Chicago Press. ISBN   978-0-226-43728-6.
  13. 1 2 3 Berryman, A. A. (1992). "The Origins and Evolution of Predator-Prey Theory" (PDF). Ecology . 73 (5): 1530–1535. doi:10.2307/1940005. JSTOR   1940005. Archived from the original (PDF) on 2010-05-31.
  14. Arditi, R.; Ginzburg, L. R. (1989). "Coupling in predator-prey dynamics: ratio dependence" (PDF). Journal of Theoretical Biology . 139 (3): 311–326. Bibcode:1989JThBi.139..311A. doi:10.1016/s0022-5193(89)80211-5. Archived from the original (PDF) on 2016-03-04. Retrieved 2020-11-17.
  15. Abrams, P. A.; Ginzburg, L. R. (2000). "The nature of predation: prey dependent, ratio dependent or neither?". Trends in Ecology & Evolution . 15 (8): 337–341. doi:10.1016/s0169-5347(00)01908-x. PMID   10884706.
  16. Johnson, J. B.; Omland, K. S. (2004). "Model selection in ecology and evolution" (PDF). Trends in Ecology and Evolution . 19 (2): 101–108. CiteSeerX   10.1.1.401.777 . doi:10.1016/j.tree.2003.10.013. PMID   16701236. Archived from the original (PDF) on 2011-06-11. Retrieved 2010-01-25.
  17. 1 2 Vandermeer, J. H.; Goldberg, D. E. (2003). Population ecology: First principles. Woodstock, Oxfordshire: Princeton University Press. ISBN   978-0-691-11440-8.
  18. Jahn, Gary C.; Almazan, Liberty P.; Pacia, Jocelyn B. (2005). "Effect of Nitrogen Fertilizer on the Intrinsic Rate of Increase of Hysteroneura setariae (Thomas) (Homoptera: Aphididae) on Rice (Oryza sativa L.)". Environmental Entomology . 34 (4): 938–43. doi: 10.1603/0046-225X-34.4.938 .
  19. Hassell, Michael P. (June 1980). "Foraging Strategies, Population Models and Biological Control: A Case Study". The Journal of Animal Ecology. 49 (2): 603–628. doi:10.2307/4267. JSTOR   4267.
  20. 1 2 3 4 "Geometric and Exponential Population Models" (PDF). Archived from the original (PDF) on 2015-04-21. Retrieved 2015-08-17.
  21. "Bacillus stearothermophilus NEUF2011". Microbe wiki.
  22. Chandler, M.; Bird, R.E.; Caro, L. (May 1975). "The replication time of the Escherichia coli K12 chromosome as a function of cell doubling time". Journal of Molecular Biology . 94 (1): 127–132. doi:10.1016/0022-2836(75)90410-6. PMID   1095767.
  23. Tobiason, D. M.; Seifert, H. S. (19 February 2010). "Genomic Content of Neisseria Species". Journal of Bacteriology . 192 (8): 2160–2168. doi:10.1128/JB.01593-09. PMC   2849444 . PMID   20172999.
  24. Boucher, Lauren (24 March 2015). "What is Doubling Time and How is it Calculated?". Population Education.
  25. "Population Growth" (PDF). University of Alberta . Archived from the original (PDF) on 2018-02-18. Retrieved 2020-11-16.
  26. "Evolutionary Game Theory". Stanford Encyclopedia of Philosophy. The Metaphysics Research Lab, Center for the Study of Language and Information (CSLI), Stanford University. 19 July 2009. ISSN   1095-5054 . Retrieved 16 November 2020.
  27. Nanjundiah, V. (2005). "John Maynard Smith (1920–2004)" (PDF). Resonance . 10 (11): 70–78. doi:10.1007/BF02837646. S2CID   82303195.
  28. 1 2 3 Altizer, Sonia; Dobson, Andrew; Hosseini, Parviez; Hudson, Peter; Pascual, Mercedes; Rohani, Pejman (2006). "Seasonality and the dynamics of infectious diseases". Reviews and Syntheses. Ecology Letters . Blackwell Publishing Ltd (French National Centre for Scientific Research (CNRS)). 9 (4): 467–84. doi: 10.1111/J.1461-0248.2005.00879.X . hdl: 2027.42/73860 . PMID   16623732. S2CID   12918683.

Further reading