Fact-checked by Grok 2 weeks ago

Moran process

The Moran process is a continuous-time model in that simulates the evolution of frequencies in a finite of constant size N, where each individual reproduces and dies at a uniform rate, leading to overlapping generations and a birth-death dynamics that maintains population constancy. In this process, at each step, one individual is selected for reproduction proportional to its , and its replaces a randomly chosen individual for death, modeling and selection in haploid or effectively haploid populations. Introduced by mathematician Patrick A. P. Moran in 1958 as an analytically tractable alternative to earlier discrete-generation models, the process was originally formulated to analyze random genetic changes under and selection in well-mixed populations. Moran's work built on foundational ideas from and but emphasized continuous-time Markov chains, with states representing the number of copies of a particular (from 0 to N), where 0 and N are absorbing states corresponding to fixation or loss of the . The model assumes a structure for interactions in its classic form, allowing exact computations of fixation probabilities and times, which were detailed in Moran's subsequent book The Statistical Processes of Evolutionary Theory (1962). In the neutral case, where all individuals have equal , the probability of a fixing in the starting from i copies is simply i/N, and the expected time to fixation or loss scales with squared N^2, approximately -N^2 [p \ln p + (1-p) \ln (1-p)] for initial frequency p = i/N. Under selection, with relative r > 1 for , the birth rate for the mutant type increases to (N - i) i r / N, while the death rate remains i (N - i) / N, yielding a fixation probability of [1 - (1/r)^i] / [1 - (1/r)^N], which amplifies the role of selection compared to drift. This contrasts with the Wright-Fisher model, which uses non-overlapping generations and sampling for offspring distribution, making the Moran process more suitable for deriving approximations and handling weak selection limits. Beyond classical population genetics, the Moran process has been extended to structured populations on graphs, where reproduction is limited to neighbors, influencing fixation probabilities based on graph topology—such as amplifiers (e.g., star graphs) that enhance selection or suppressors that hinder it. These generalizations, pioneered in works like Lieberman, Hauert, and Nowak (2005), apply to evolutionary game theory, epidemiology, and cancer modeling, where spatial structure affects the spread of advantageous mutants. Recent developments continue to explore multiallelic selection and time-inhomogeneous variants, underscoring the model's enduring utility in theoretical biology.

Fundamentals

Definition and Overview

The Moran process is a stochastic model in population genetics that simulates the evolution of allele frequencies in a finite population of constant size N, representing genetic drift and potential selection through a birth-death mechanism in overlapping generations. Introduced by Patrick Moran in 1958, it treats the population as a continuous-time Markov chain where the state is defined by the number i of individuals carrying a focal allele A (with the remainder N - i carrying the alternative allele a), ranging from i = 0 (no A alleles) to i = N (complete fixation of A). This setup assumes asexual reproduction, where offspring inherit the parent's allele exactly, and the population maintains strict size constancy via immediate replacement after each event. In , at each step, one is selected for proportional to its relative , and its replaces a uniformly at random chosen in the population (which may include the itself); the baseline assumes equal fitness for neutrality. This birth-death dynamic models evolutionary , where a single (starting at i = 1) may spread to fixation (i = N) or be lost (i = 0), capturing the nature of trajectories in finite populations. Unlike non-overlapping generation models, the Moran process allows continuous generational overlap, enabling finer-grained analysis of short-term fluctuations and long-term times. Since the , the Moran process has served as a foundational tool in , providing insights into neutral drift—where alleles evolve by random chance without fitness differences—and extensions incorporating selection to study adaptive evolution. Its simplicity and tractability have made it influential for theoretical developments and simulations in .

Historical Background

The Moran process was introduced by the Australian statistician Patrick Alfred Pierce Moran (1917–1988) in his seminal paper "Random processes in genetics," where he developed a model to describe changes in gene frequencies within finite populations, serving as a continuous-time alternative to the discrete-generation Wright-Fisher model. Moran formulated the process to capture birth-death dynamics in overlapping generations, addressing key aspects of evolutionary change in bounded populations. He expanded on these ideas in his 1962 book The Statistical Processes of Evolutionary Theory, which provided a comprehensive treatment of processes in , including detailed analyses of the model's implications for dynamics. This development occurred amid mid-20th-century advances in , where researchers recognized the shortcomings of deterministic frameworks like the Hardy-Weinberg equilibrium, which idealized infinite population sizes and overlooked random in finite settings. Moran's model emerged as a response to these limitations, offering a tractable way to incorporate effects in small populations, such as those encountered in experimental or natural systems with constrained sizes. Early applications by Moran focused on neutral evolution scenarios, particularly the fixation of alleles in haploid populations, which proved especially relevant to asexually reproducing organisms like , where continuous reproduction aligns with the model's overlapping generations. In the 1960s, the model gained further traction through extensions by Samuel Karlin and James McGregor, who provided detailed analyses of transition probabilities and stationary distributions in their 1962 paper "On a Genetics Model of ." By the and , contributions from Karlin and collaborators, alongside integrations into influential texts such as and Kimura's An Introduction to Theory (1970), helped standardize the Moran process as a foundational tool in for modeling drift and selection in finite populations.

Mathematical Formulation

Basic Model Setup

The Moran process is formulated as a on the discrete state space \{i = 0, 1, \dots, N\}, where i denotes the number of individuals of type A in a of constant size N. The consists of two types of individuals, type A and type a, with overlapping generations maintained through birth and death events that preserve the fixed total size N. Key parameters of the model include the b and rate d for individuals, along with values assigned to each type; in the case, these values are equal for types A and a. The states i=0 and i=N are absorbing states, corresponding to fixation of type a (all individuals are type a) and fixation of type A (all individuals are type A), respectively. In the process, an is selected for with probability proportional to its relative to the average, while an is selected for death uniformly at random across all N individuals. The offspring of the selected reproducing individual then replaces the deceased individual, ensuring the population size remains constant at N. In the case, where fitness values are equal, the selection probability for reproduction reduces to uniform across all individuals.

Transition Dynamics

The transition dynamics of the Moran process are described by a on the state space {0, 1, \dots, N}, where the state i denotes the number of type A individuals in a fixed N. The process allows transitions only between adjacent states i \to i \pm 1, reflecting single birth-death events that change the type composition by one individual. These transitions capture the replacement mechanism: an individual is selected to reproduce, and another is selected to die, with the offspring replacing the deceased. In the discrete-time formulation, the probability of increasing the state from i to i+1 (birth of a type A offspring replacing a type a ) is given by the product of the probability that a type A is selected for and the probability that a type a is selected for death. With relative f_A for type A and f_a for type a, the total fitness is T = i f_A + (N - i) f_a, so the probability for type A is i f_A / T and the uniform death probability for type a is (N - i)/N. Thus, P_{i,i+1} = \frac{i f_A}{T} \cdot \frac{N - i}{N}. Similarly, the probability of decreasing the state from i to i-1 (birth of a type a offspring replacing a type A ) is P_{i,i-1} = \frac{(N - i) f_a}{T} \cdot \frac{i}{N}. The adjustment arises from the fitness-proportional selection of the reproducer (f_A / T or f_a / T), while death selection remains uniform. The death rate adjustment is absent in the standard model, as deaths occur equally across types. The probability of no change is $1 - P_{i,i+1} - P_{i,i-1}. The continuous-time version approximates the process as a birth-death chain, where events occur according to an exponential waiting time. In the standard normalization, each individual has a death rate of 1, yielding a total update rate of N per unit time (one full population turnover on average). The transition rates then follow from scaling the discrete probabilities by this update rate, but only counting type-changing events. The generator matrix Q has entries Q_{i,i+1} = \frac{i (N - i) f_A}{T}, \quad Q_{i,i-1} = \frac{i (N - i) f_a}{T}, with Q_{i,i} = - (Q_{i,i+1} + Q_{i,i-1}) and zero elsewhere. This yields the general form of the for the embedded continuous-time process. The total rate of type-changing events is \frac{i (N - i) (f_A + f_a)}{T} from state i, which in the case (f_A = f_a = 1, T = N) simplifies to \frac{2 i (N - i)}{N} and reaches a maximum of \frac{N}{2} per unit time at i = \frac{N}{2}. In the neutral case with equal fitnesses, the transitions become symmetric, P_{i,i+1} = P_{i,i-1} = i (N - i) / N^2 in discrete time and Q_{i,i+1} = Q_{i,i-1} = i (N - i) / N in continuous time, emphasizing drift without directional bias.

Neutral Case

Fixation Probability

In the neutral process, where all individuals possess equal reproductive , the probability that a specific achieves fixation, starting from an initial count of i copies in a finite of constant size N, is given by \pi_i = \frac{i}{N}. This linear relationship reflects the inherent symmetry of the process under ity, where birth and death rates for the allele are balanced at each state i. This result follows from the martingale property of the count X_t, which maintains a constant E[X_t] = i due to equal transition rates upward and downward from state i. Upon at the boundaries (either fixation at N or loss at $0), the becomes N \cdot \pi_i + 0 \cdot (1 - \pi_i) = i, yielding \pi_i = \frac{i}{N}. The boundary conditions are \pi_0 = 0 (certain loss if absent) and \pi_N = 1 (certain fixation if already fixed). Alternatively, for large N, a diffusion approximation models the allele frequency p = i/N as a continuous process satisfying the backward Kolmogorov equation \mu(p) \frac{d\pi}{dp} + \frac{1}{2} \sigma^2(p) \frac{d^2\pi}{dp^2} = 0, where the neutral drift \mu(p) = 0 and variance \sigma^2(p) = \frac{p(1-p)}{N}. The general solution simplifies to \pi(p) = p under the boundary conditions \pi(0) = 0 and \pi(1) = 1, confirming the exact discrete result in the limit. This fixation probability underscores the role of genetic drift in neutral evolution: each allele has a chance of ultimate dominance strictly proportional to its initial frequency, with outcomes determined randomly regardless of the trajectory taken to absorption.

Time to Absorption

In the neutral Moran process, the time to absorption refers to the expected duration until one allele reaches fixation or loss, starting from an initial count of i individuals of the mutant allele in a population of size N. This quantity quantifies the pace of genetic drift under neutrality and is derived from the fundamental equations of Markov chain absorption times. The mean time to absorption T_i, measured in generations (where each birth-death event contributes $1/N to a generation), satisfies the recursive relation N T_i = 1 + \sum_{j=0}^N p_{ij} (N T_j), with boundary conditions T_0 = T_N = 0, where p_{ij} are the transition probabilities from state i to j. For the neutral case, p_{i,i+1} = p_{i,i-1} = \frac{i (N-i)}{N^2} and p_{i,i} = 1 - \frac{2 i (N-i)}{N^2}. For large N, T_i \approx -2 \left[ \frac{i}{N} \ln \left( \frac{i}{N} \right) + \left(1 - \frac{i}{N} \right) \ln \left(1 - \frac{i}{N} \right) \right] N, emphasizing the logarithmic scaling near the boundaries. A simpler quadratic approximation, valid for intermediate frequencies away from 0 or 1, is T_i \approx N \frac{i}{N} \left(1 - \frac{i}{N}\right). The variance of the absorption time, \mathrm{Var}(T_i), accounts for fluctuations in drift trajectories and is computed using the fundamental matrix of the transient states. Let Q be the (N-1) \times (N-1) submatrix of transitions among states $1toN-1; then the matrix V = (I - Q)^{-1}$ gives the expected number of visits to each transient state, and \mathrm{Var}(T_i) = 2 \sum_{j=1}^{N-1} \sum_{k=1}^{N-1} v_{jk} - \sum_{j=1}^{N-1} v_{ij}^2, scaled appropriately by the time unit $1/N per step. This yields \mathrm{Var}(T_i) = O(N^2) for fixed i/N, reflecting high variability due to the diffusive nature of neutral evolution. Conditional absorption times condition on the outcome of drift, with the neutral fixation probability i/N serving as the conditioning factor. The expected time to fixation given that the mutant fixes, T_i^\text{fix}, is T_i^\text{fix} \approx -\frac{N}{\frac{i}{N}} \left(1 - \frac{i}{N}\right) \ln \left(1 - \frac{i}{N}\right), which for small i/N scales as O(N), consistent with the overall coalescence timescale of approximately N generations under neutrality. Symmetrically, the time to loss given loss follows a similar form but is shorter on average near the boundaries.

Selection Case

Fixation under Weak Selection

In the Moran process with selection, the of type A individuals is 1 + s and that of type a individuals is 1, where s is the selection coefficient denoting the relative fitness advantage (or disadvantage if s < 0) of type A. This formulation modifies the transition probabilities, introducing a bias toward increase when s > 0 and i > 0. The weak selection regime is characterized by |s| << 1/N, under which genetic drift continues to dominate evolutionary dynamics while selection exerts only a minor influence, allowing approximations that perturb from neutrality. The fixation probability π_i, starting from i copies of type A, can be approximated via perturbation expansion from the neutral case π_i = i/N. By assuming π_i = i/N + s ψ_i and substituting into the —derived from the birth-death transition rates T_{i,i+1} = \frac{i(1+s)(N-i)}{N[i(1+s) + (N-i)]} and T_{i,i-1} = \frac{i(N-i)}{N[i(1+s) + (N-i)]}—the zeroth-order equation is satisfied by the neutral solution, and the first-order equation yields a linear system for ψ_i. Solving this system gives the approximation \pi_i \approx \frac{i}{N} \left[1 + s \frac{i (N-i)}{N (N-1)}\right]. This reveals a positive correction for s > 0 that is quadratic in i, peaking at intermediate frequencies where both types coexist. An alternative approximation arises from the diffusion limit of the Moran process, treating the allele frequency x = i/N as a continuous variable for large N with Ns = O(1). The corresponding Kolmogorov backward equation becomes the ordinary differential equation s x (1-x) \frac{d \pi}{dx} + \frac{x (1-x)}{N} \frac{d^2 \pi}{dx^2} = 0, with boundary conditions π(0) = 0 and π(1) = 1. Solving this boundary value problem yields \pi_i \approx \frac{1 - e^{-s i}}{1 - e^{-s N}}. This formula integrates the selective drift over the stochastic trajectory, providing a smooth transition from neutrality as s → 0. For |s| << 1/N, the perturbation approach is more appropriate as it directly quantifies the small bias without requiring the diffusion scaling, while both methods confirm that positive selection elevates π_i above i/N, with the neutral case recovered exactly at s = 0.

Fixation under Strong Selection

In the Moran process under selection, the fixation probability \pi_i of a mutant allele starting from i individuals in a population of constant size N, where the mutant has relative fitness r = 1 + s (with s the selection coefficient) and the wild-type has fitness 1, admits an exact closed-form solution. This probability is \pi_i = \frac{1 - (1/r)^i}{1 - (1/r)^N}, with boundary conditions \pi_0 = 0 and \pi_N = 1. This expression is derived by solving the backward for the absorption probabilities in the birth-death process underlying the Moran model. Specifically, the fixation probabilities satisfy the recursion \pi_i = \frac{p_{+}^{(i)} \pi_{i+1} + p_{-}^{(i)} \pi_{i-1}}{p_{+}^{(i)} + p_{-}^{(i)}}, \quad i = 1, \dots, N-1, where p_{+}^{(i)} = \frac{i (N - i) r}{N (N - 1)} is the transition probability from state i to i+1, and p_{-}^{(i)} = \frac{i (N - i)}{N (N - 1)} is the probability from i to i-1. For general fitness values, this linear system can be solved numerically via iterative methods or matrix inversion, though the closed form above provides an efficient evaluation for the constant-fitness case. Under strong selection, defined by large |Ns| (where stochastic drift is negligible compared to deterministic forces), the behavior of \pi_i simplifies markedly. For an advantageous (s > 0 and large r = 1 + s), \pi_i \approx 1 for any i \geq 1, as the denominator approaches 1 and the numerator nears 1, reflecting near-certain fixation independent of initial frequency except at the boundary i = 0. Conversely, for a deleterious (s < 0 and small r = 1 + s), \pi_i \approx 0 for i < N, with appreciable probability only when i is close to N, as the process is strongly biased toward loss. These deterministic outcomes dominate when selection overwhelms genetic drift. A key threshold for selective sweeps in the Moran process occurs when s > 1/N, or equivalently Ns > 1, marking the regime where selection overcomes drift such that the fixation probability of a single advantageous mutant (\pi_1) substantially exceeds the neutral value $1/N. Below this threshold (s < 1/N), the process behaves nearly neutrally.

Comparisons and Extensions

Relation to Wright-Fisher Model

The Wright-Fisher model describes genetic drift in a population of fixed size N (or $2N alleles in the diploid case) through discrete, non-overlapping generations, where the next generation is formed by sampling N alleles with replacement from the current generation according to a binomial distribution, incorporating selection via fitness-biased sampling probabilities. In contrast, the Moran process models overlapping generations in continuous time (or discrete steps), where at each event, one individual is chosen to reproduce proportional to its fitness, and its offspring replaces a randomly selected individual in the population, leading to a birth-death dynamics that maintains constant population size. In the diffusion approximation for large population size N, both models converge to the same forward and backward Kolmogorov equations governing the evolution of allele frequencies under drift, mutation, and selection, establishing their mathematical equivalence in the continuous limit where demographic noise is scaled appropriately. This equivalence arises because the infinitesimal mean and variance of allele frequency changes match after rescaling, allowing results derived from one model to be applied to the other in asymptotic analyses of fixation probabilities and site frequency spectra. A key difference lies in time scaling: in the Moran process, coalescence times scale on the order of N generations, whereas in the , they scale as $2N generations, reflecting the faster rate of genetic drift in the Moran model (twice that of Wright-Fisher when measured per generation) due to frequent overlapping updates. To align the models, time in the Moran process is typically rescaled by a factor of 2, ensuring comparable evolutionary dynamics. Recent efforts have developed hybrid models to bridge the two frameworks for finite N, such as updating a fixed fraction of the population per discrete time step, which interpolates between Wright-Fisher's full generational replacement and Moran's incremental changes, while preserving tractability for simulations and analytical fixation times. These hybrids facilitate exact mappings and effective population size calculations, enabling more precise finite-N approximations without relying solely on diffusion limits. The Moran process offers advantages for exact computations in small populations, as its Markov chain structure allows straightforward calculation of transition probabilities and absorption times without the binomial sampling complexity of Wright-Fisher, and its continuous-time formulation simplifies modeling variable rates.

Structured and Multiallelic Extensions

The Moran process has been extended to structured populations where individuals interact on graphs rather than in a well-mixed setting, allowing the study of spatial effects on evolutionary dynamics. In the spatial Moran process, typically implemented as a birth-death update rule, an individual is selected proportional to fitness for reproduction, and its offspring replaces a randomly chosen neighbor on the graph. Fixation probabilities in these models depend strongly on the graph's topology; for instance, star graphs act as strong amplifiers of selection, increasing the fixation probability of advantageous mutants far beyond the well-mixed case, while cycle graphs suppress selection, reducing it below neutral levels. Multiallelic extensions generalize the biallelic to k alleles with arbitrary fitness interactions, often represented via a fitness matrix that captures frequency-dependent selection. These models embed the dynamics in a higher-dimensional state space, where the process tracks the counts of each allele, and absorption occurs when one allele reaches fixation. Analytical solutions for fixation probabilities and times can be derived using embedded Markov chains, facilitating the study of complex interactions like overdominance or directional selection among multiple types. Recent work has focused on colonization times and invasion processes in heterogeneous graphs under the Moran process, where a single mutant invades an empty or resident-occupied network. For example, on directed graphs like total orders, colonization times scale differently from fixation times in well-mixed populations, highlighting how structure influences the speed of spatial spread. These analyses reveal that heterogeneous graphs can accelerate or delay invasions depending on connectivity patterns, with exact formulas available for common structures like complete or cycle graphs. The spatial Moran process is closely related to the voter model, providing a duality for ancestry tracing: forward in time, the voter model describes opinion spread (analogous to allele invasion), while backward in time, it reconstructs genealogies under the Moran dynamics, aiding coalescent approximations for structured populations. However, extensions to large graphs pose significant challenges, including high computational complexity for exact fixation calculations, which scale exponentially with population size and often require matrix exponentiation or simulation-based approximations like effective population sizes to handle realistic networks.

Applications

Rate of Molecular Evolution

The Moran process connects directly to Kimura's neutral theory of molecular evolution, introduced in 1968, by modeling genetic drift in finite populations as the primary driver of neutral substitutions, offering corrections to infinite-population assumptions where fixation occurs deterministically. In the neutral Moran process, a newly arising mutant allele has a fixation probability of \frac{1}{N}, where N is the constant haploid population size, reflecting the equal chance of any allele reaching fixation under drift alone. This fixation probability ensures that the rate of molecular evolution, measured as the neutral substitution rate per site, equals the underlying neutral mutation rate u per site per generation and is independent of population size. Specifically, new mutations arise at a total rate k = N u across the population per unit time (in the standard scaling where the total birth rate is N), and each fixes with probability \frac{1}{N}, yielding a substitution rate \mu = \frac{k}{N} = u. Moran process simulations for small populations highlight elevated variance in substitution rate estimates, stemming from intensified stochastic fluctuations in mutant trajectories and fixation events, which widen confidence intervals and complicate precise inference of evolutionary rates compared to larger populations.

Evolutionary Game Theory

The Moran process provides a foundational framework for modeling the evolution of strategies in finite populations within evolutionary game theory, where individual fitness is determined by payoffs from strategic interactions rather than constant values. In this setup, the population consists of individuals adopting different strategies, and at each step, an individual is selected for reproduction proportional to its payoff-derived fitness, while another is chosen uniformly for death. This frequency-dependent selection captures how the success of a strategy, such as cooperation or defection, varies with the composition of the population. Seminal work on structured populations demonstrated that such dynamics can lead to the emergence of cooperation in social dilemmas when spatial structure amplifies selection, contrasting with infinite-population replicator dynamics where defectors typically dominate. Variants of the Moran process, such as imitation or pairwise comparison processes, adapt the reproduction mechanism to reflect social learning, where strategies spread based on observed payoffs. In the pairwise comparison process, an individual compares its payoff to that of another randomly selected individual and imitates the higher-payoff strategy with a probability depending on the payoff difference and an intensity parameter, effectively making reproduction proportional to relative success. This formulation bridges biological evolution and cultural transmission, allowing analysis across a spectrum from weak selection (neutral drift) to strong selection (deterministic imitation). These processes maintain the core stochastic structure of the Moran model but emphasize payoff-driven imitation, enabling the study of strategy evolution in complete graphs where interactions are well-mixed. In symmetric two-strategy games like the or played on complete graphs, the Moran process yields explicit fixation probabilities for a mutant strategy invading a resident population. For the , where cooperation yields mutual benefits but is vulnerable to exploitation, a single cooperator's fixation probability is less than the neutral value of $1/N under standard parameters, as defectors exploit rare cooperators. Finite populations introduce stochastic effects that can allow cooperation to fix by drift, but selection favors defection. In the game, which models conflicts over resources with risks of injury for aggressive "Hawks" versus passive "Doves," fixation dynamics reveal how mixed equilibria in infinite populations translate to stochastic absorption in finite ones, with aggressive strategies fixating more readily when costs of conflict are low. These examples illustrate how frequency dependence alters fixation compared to constant-fitness selection, with analytical solutions derived from Markov chain absorption probabilities. A key condition for strategy replacement under weak selection in these games involves the structure coefficient \sigma, which corrects infinite-population criteria like risk dominance for finite-size effects. For the Moran birth-death process in well-mixed populations, \sigma \approx 1 for large N, meaning the condition for a mutant strategy to invade approximates the infinite-population case, such as payoff advantage >0 for risk dominance. Risk dominance, traditionally requiring payoff advantages like benefit/cost >1 in coordination games, is thus minimally adjusted by finite-size effects in large well-mixed populations. Post-2000 developments have extended the Moran process to and incorporated finite-population corrections in , enhancing its applicability to behavioral . The pairwise comparison rule, for instance, models where imitation probability follows a Fermi of payoff differences, linking to decision-making experiments. Finite-size corrections, formalized via structure coefficients, generalize and payoff dominance conditions across rules, revealing that well-mixed Moran yield \sigma \approx 1 for large N, but deviations in small populations amplify effects on selection. Recent studies (as of 2025) have applied these to model evolution in AI agents and cultural in societies. These advances have influenced models of and policy design, emphasizing how population size modulates game-theoretic outcomes without relying on spatial structure.

References

  1. [1]
    [PDF] Moran model 1 Neutral case - Duke Math
    The Moran model describes the evolution of a collection of cells that is maintained at a constant population size of N. Thinking of the population genetics ...
  2. [2]
    Random processes in genetics | Mathematical Proceedings of the ...
    Oct 24, 2008 · Random processes in genetics. Published online by Cambridge University Press: 24 October 2008. P. A. P. Moran.
  3. [3]
    [PDF] A survey of the modified Moran process and evolutionary graph theory
    Wright-Fisher model, Patrick Moran19 defined a new model for genetic drift, which today is known as the Moran Process [54, 55]. At the beginning of his.
  4. [4]
    Modeling Multiallelic Selection Using a Moran Model - PMC
    We present a Moran-model approach to modeling general multiallelic selection in a finite population and show how it may be used to develop theoretical models.
  5. [5]
    [PDF] The Moran model of population genetics - CORE
    The offspring inherits the parent's type and replaces another individual, randomly chosen from the entire population (again including the parent individual). In ...
  6. [6]
    Bridging Wright–Fisher and Moran models - ScienceDirect.com
    Feb 21, 2025 · The Wright–Fisher model and the Moran model are both widely used in population genetics. They describe the time evolution of the frequency ...Bridging Wright--Fisher And... · 2. Model With Fixed Updated... · 3. Generalizations Of The...
  7. [7]
    Patrick Alfred Pierce Moran 1917-1988
    The paper 'Random processes in genetics' ( Proc. Camb. Phil. Soc. 54, 60-71), published in 1958, was Pat's first attempt at a genetic problem and was ...
  8. [8]
    The statistical processes of evolutionary theory - Internet Archive
    Sep 4, 2019 · Moran, P. A. P. (Patrick Alfred Pierce), 1917-. Publication date: 1962. Topics: Genetics -- Mathematical models. Publisher: Oxford, Clarendon ...Missing: population | Show results with:population
  9. [9]
    Population Genetics - Stanford Encyclopedia of Philosophy
    Sep 22, 2006 · Another benefit of the Hardy-Weinberg principle is that it greatly simplifies the task of modelling evolutionary change. When a population is in ...
  10. [10]
    [PDF] Mathematical Population Genetics: Lecture Notes Cornell University ...
    As stated in the above, the Hardy-Weinberg law assumes that the population considered is infinite in size, so that that random, or stochastic, changes in ...
  11. [11]
    On a Genetics Model of Moran† | Mathematical Proceedings of the ...
    Oct 24, 2008 · Each of the gametes of the next generation is independently formed by making a random selection from the gametes of the present generation. The ...Missing: Patrick | Show results with:Patrick
  12. [12]
    An introduction to population genetics theory : Crow, James F ...
    Sep 30, 2022 · Publication date: 1970. Topics: Population genetics. Publisher: New York, Harper & Row. Collection: internetarchivebooks; inlibrary ...
  13. [13]
    The duality of spatial death–birth and birth–death processes and ...
    Apr 1, 2015 · ... birth rate (and not by their death rate). In this paper, we discuss ... Moran process. To this end, we consider a regular spatial ...
  14. [14]
  15. [15]
    [PDF] CHAPTER 15 - UBC Zoology
    This equation is referred to as the backward Kolmogorov equation: . (15.11) ... (c) Use equation (15.35b) to write the probability of fixation for an.
  16. [16]
  17. [17]
    [PDF] Theoretical Population Biology Absorption and fixation times for ...
    Sep 18, 2008 · Since it is a one dimensional stochastic process, the absorption and fixation times can be written down explicitly and efficiently calculated.
  18. [18]
    [PDF] Fixation probabilities in evolutionary dynamics under weak selection
    Feb 1, 2021 · Fixation probabilities in evolutionary dynamics under weak selection ... This formula depends on the first-order effects of selection.
  19. [19]
  20. [20]
    [PDF] Diffusion Process Models in Mathematical Genetics
    The Moran model is a continuous time chain - which just means that the times between events are determined by a sequence of independent exponentially.
  21. [21]
    (PDF) Quantitative approximation of the discrete Moran process by a ...
    Aug 7, 2025 · The Wright–Fisher diffusion is commonly used as an approximation in order to understand the dynamics of population genetics models.
  22. [22]
    Extensions of the Coalescent Effective Population Size - PMC
    For the Wright–Fisher model, 1 time step constitutes 1 generation. For the Moran model, it takes N time steps to make 1 generation.
  23. [23]
    [PDF] The Coalescent - Santa Fe Institute Events Wiki
    Thus the rate of genetic drift per time unit in the Moran model is equal to 2/N2. To make this comparable to drift in the Wright-Fisher model, we can define a ...
  24. [24]
  25. [25]
  26. [26]
    Moran model as a dynamical process on networks and its ...
    Sep 1, 2011 · In population genetics, the Moran model describes the neutral evolution of a biallelic gene in a population of haploid individuals subjected ...
  27. [27]
    Colonization times in Moran process on graphs - Research journals
    Moran Birth-death process is a standard stochastic process that is used to model natural selection in spatially structured populations.
  28. [28]
    Genealogical constructions of population models - Project Euclid
    The time-evolution of the levels captures the genealogies of the particles in the population. Key forces of ecology and genetics can be captured within this com ...Missing: ancestry | Show results with:ancestry
  29. [29]
    Evolutionary Rate at the Molecular Level - Nature
    Calculating the rate of evolution in terms of nucleotide substitutions seems to give a value so high that many of the mutations involved must be neutral ones.
  30. [30]
    [PDF] The Moran model with selection: Fixation probabilities, ancestral ...
    Dec 6, 2013 · Abstract. We reconsider the Moran model in continuous time with pop- ulation size N, two allelic types, and selection.
  31. [31]
    [PDF] Population size and the rate of evolution - Rob Lanfear
    As a result, the neutral substitution rate equals the mutation rate, no matter what the value of Ne. Thus, if the mutation rate does not change, then the NeRR ...
  32. [32]
    A study on a nearly neutral mutation model in finite populations.
    The substitution rate is reduced as 4N sigma is increased. Three tests of neutrality, one using the relationship between the average and the variance of ...
  33. [33]
  34. [34]