Fact-checked by Grok 2 weeks ago

Neutral mutation

A neutral mutation is a genetic change in the DNA sequence that neither enhances nor impairs an organism's , meaning it has no significant on survival or . These mutations occur randomly and are typically fixed in populations through rather than . Examples include silent mutations that do not alter the sequence of proteins, as well as some synonymous codon changes or variations in non-coding regions that do not affect gene function. The concept of neutral mutations forms the cornerstone of the , first proposed by in 1968. This theory asserts that the majority of genetic variation observed at the molecular level arises from neutral mutations, which accumulate and become fixed in populations at a rate approximately equal to the mutation rate itself, independent of selective pressures. By emphasizing as the primary mechanism for these changes, the neutral theory explains the unexpectedly high rates of and intraspecific variability without invoking adaptive significance for most DNA differences. Neutral mutations play a crucial role in understanding evolutionary processes, as their fixation probability equals their initial frequency in the population, confirming their independence from selection. While not all mutations are neutral—many are deleterious and purged, or advantageous and favored—the neutral framework highlights that a substantial portion of genomic , particularly in non-functional regions, evolves .

Fundamentals

Definition

A neutral mutation refers to a change in the DNA sequence that has no discernible effect on an organism's , , or overall function within its environment. Such mutations neither confer an nor impose a , allowing their fate in a population to be governed primarily by rather than . In essence, the fixation or loss of a neutral mutation occurs independently of selective pressures, as it does not influence the organism's ability to survive or reproduce. The criteria for classifying a mutation as neutral center on the absence of any significant impact on key evolutionary parameters: survival rates, , or to environmental changes. At the molecular level, these mutations produce silent effects, meaning they do not disrupt , , or cellular processes in a way that alters organismal performance. , defined as the relative contribution to the next generation's , remains unchanged, distinguishing neutral mutations from those that could shift . Neutral mutations arise from various alterations in DNA sequences, such as point mutations (single nucleotide substitutions), insertions, or deletions, provided these changes result in no functional consequences. They commonly occur in non-coding regions of the genome, where sequence variations do not influence gene regulation or protein coding, or in redundant genetic elements like pseudogenes, which lack active transcriptional roles and accumulate changes without downstream effects. A typical example is a synonymous mutation in a coding region, which alters a codon but encodes the same amino acid, thereby preserving protein function.

Distinction from other mutations

Neutral mutations are characterized by their lack of significant effect on an organism's , distinguishing them from advantageous mutations that increase and are thus subject to positive selection, which accelerates their fixation in populations. In contrast, deleterious mutations decrease and are typically purged by negative or purifying selection, preventing their spread and maintaining functional genetic integrity. These distinctions highlight how selection shapes the trajectory of non-neutral mutations, either favoring or eliminating harm, whereas neutral mutations evade such pressures. Nearly neutral mutations represent a nuanced category at the boundary of neutrality, featuring subtle effects where the selection coefficient s is small enough that the product N_e s (with N_e as ) falls near or below 1, rendering their behavior effectively neutral under drift dominance in smaller populations. In larger populations, these mutations may experience weak selection, shifting them toward slightly deleterious or advantageous outcomes depending on demographic conditions like population bottlenecks. This conditional nature underscores how modulates the evolutionary relevance of near-neutral variants, bridging strict neutrality and selective influence. The evolutionary implications of neutrality emphasize random as the primary mechanism for fixation, in stark contrast to the propelled by selection on advantageous mutations or the constraint imposed by deleterious ones. mutations thus accumulate polymorphisms and substitutions without adaptive purpose, contributing to molecular clock-like rates of change across lineages. This framework, central to the neutral theory, posits that most proceeds via such stochastic processes rather than selection-driven .

Historical Context

Early observations

In the pre-molecular era, early population geneticists began recognizing the role of random processes in altering gene frequencies, independent of natural selection. Sewall Wright's work in the 1930s highlighted genetic drift as a mechanism driving random changes in allele frequencies within finite populations, where neutral genetic variations could fluctuate and potentially fix without adaptive advantage. This concept, formalized in Wright's seminal 1931 paper, provided a foundational insight into non-adaptive evolutionary changes, emphasizing that mutation and drift could maintain genetic diversity even in the absence of selective pressures. By the 1960s, the advent of molecular techniques revealed unexpectedly high levels of at the protein level, challenging the prevailing view that most polymorphisms were under strict selective control. noted that rates of , inferred from early protein sequence comparisons, appeared too rapid to be explained solely by adaptive substitutions, suggesting instead that many changes were selectively and governed by and drift. This observation was bolstered by Émile Zuckerkandl and Pauling's 1965 analysis of protein sequences, which demonstrated a roughly constant rate of substitutions across , implying the accumulation of neutral mutations over time akin to a . Empirical support came from protein electrophoresis studies, which uncovered widespread protein polymorphisms, often presumed to be functionally , in natural populations. In 1966, John L. Hubby and Richard C. Lewontin applied to pseudoobscura, finding that approximately 30% of surveyed loci were polymorphic, with heterozygosity levels indicating substantial neutral variation persisting without evident selective cost. Concurrently, Harry Harris's electrophoresis surveys of enzymes revealed similar high polymorphism rates, further evidencing that much molecular diversity was consistent with neutral processes rather than adaptive . These accumulating observations culminated in 1968 with foundational publications on rates, including Kimura's paper explicitly proposing that neutral mutations dominate evolutionary change at the molecular level, setting the stage for a in understanding .

Formulation of neutral theory

formulated the in 1968, proposing that the majority of evolutionary changes at the molecular level arise from neutral mutations that are fixed in populations primarily through rather than . This theory emerged as a response to observations of unexpectedly high rates of molecular evolution, suggesting that most substitutions and replacements do not significantly affect and thus accumulate stochastically. At its core, the neutral theory posits that the rate at which mutations occur equals the itself, denoted as \mu, and that the probability of fixation for a neutral mutation in a diploid is $1/(2N), where N is the . Consequently, the rate of k under neutrality simplifies to k = \mu, independent of or selection pressures, in stark contrast to adaptive evolution where fixation rates depend on selective advantages. This mathematical foundation provided a null model for molecular change, emphasizing as the predominant mechanism for neutral alleles in sufficiently large populations. The theory quickly elicited responses and refinements from the scientific community. In 1969, Jack Lester King and Thomas H. Jukes independently advanced similar ideas in their paper on non-Darwinian evolution, reinforcing the role of mutations and in protein evolution and sparking broader debates with selectionists who argued for the prevalence of adaptive changes. Ohta extended the framework in 1973 with the nearly neutral theory, incorporating slightly deleterious mutations whose fixation probability varies with , thus bridging and selective processes in . These developments highlighted ongoing controversies, particularly regarding the proportion of versus adaptive substitutions, but solidified the neutral theory's influence on .

Classification

Synonymous mutations

Synonymous mutations are alterations in the sequence of a codon that do not change the it encodes, a phenomenon enabled by the degeneracy of the whereby most of the 20 standard are specified by two to six synonymous codons. This redundancy allows multiple DNA triplets to direct the incorporation of the same during protein , preserving the protein's primary despite the . The primary mechanism underlying synonymous mutations involves variability, or "wobble," particularly at the third of the codon, where base changes frequently yield synonymous outcomes due to relaxed base-pairing rules in the tRNA anticodon. For example, the codons CUU and CUC both code for , differing only by a U-to-C in the third , illustrating how such changes maintain identity without altering . This positional bias contributes to the prevalence of synonymous mutations in coding sequences, as the third base often tolerates substitutions more readily than the first or second. While synonymous mutations do not alter the sequence, they can influence through effects on efficiency, mRNA stability, splicing, and ; historically considered , recent studies indicate that approximately 76% are significantly deleterious. Consequently, synonymous mutations accumulate at high frequencies within protein-coding regions, reflecting minimal purifying selection in many cases and serving as a for baseline evolutionary rates, though with caveats due to potential non-neutral effects. The synonymous substitution rate (), which quantifies these changes, is widely used as a proxy for ; in mammals, averages approximately $2.2 \times 10^{-9} substitutions per site per year.

Neutral amino acid substitutions

Neutral substitutions, also known as conservative substitutions, involve the replacement of one with another that shares similar physicochemical properties, such as hydrophobicity, size, charge, or polarity, thereby preserving the overall structure and function of the protein. These changes typically occur without significantly altering , stability, or activity, distinguishing them from more disruptive mutations. A classic example is the substitution of for , both of which are non-polar, branched-chain aliphatic with comparable volumes and hydrophobic characteristics, allowing the protein to maintain its native conformation. In structural proteins like , the replacement of with in the alpha-globin chain represents another neutral conservative substitution, as these small, neutral minimally impact the protein's and oxygen-binding capacity. Similarly, in enzymes such as , conservative substitutions in regions interacting with the template-primer, such as replacing one hydrophobic residue with another of similar size, are frequently tolerated without compromising catalytic efficiency. Evidence for the neutrality of these substitutions comes from functional assays and studies, which show no measurable cost or phenotypic alteration despite the amino acid change. For instance, naturally occurring variants in mouse hemoglobin with alanine-for-glycine substitutions exhibit no detectable impairment in oxygen transport or survival rates, as confirmed by high-resolution electrophoretic techniques that resolve these proteins without functional deficits. In conserved protein regions, such substitutions accumulate over evolutionary time without driving adaptive selection, supporting their neutral status through comparisons of polymorphism and divergence data. In contrast to radical substitutions, which involve amino acids differing markedly in polarity, charge, or size—often leading to disrupted interactions, altered folding, or loss of function—neutral conservative changes maintain biochemical compatibility, as evidenced by higher substitution rates for similar amino acid pairs in phylogenetic analyses. This conservation of properties, such as replacing a positively charged with , ensures minimal perturbation to electrostatic or hydrophobic networks critical for protein activity.

Assessment Methods

Identification techniques

Sequence comparison is a foundational computational for identifying potential neutral mutations by aligning genomic sequences across individuals or to detect polymorphisms, such as single nucleotide polymorphisms (SNPs), that do not alter protein function. Tools like software (e.g., MAFFT or Clustal Omega) facilitate this process, enabling the pinpointing of synonymous substitutions or conservative nonsynonymous changes as candidates for neutrality. Synonymous mutations, which do not change the sequence, are common initial targets for such analysis due to their presumed lack of functional impact. A key metric in this approach is the dN/dS ratio, which compares the rate of nonsynonymous substitutions (dN, altering amino acids) to synonymous substitutions (dS, preserving amino acids); a ratio approximately equal to 1 suggests neutral evolution, as both types of changes accumulate at similar rates without selective pressure. This ratio was first formalized to quantify substitution types in protein evolution, providing evidence for neutrality when dN ≈ dS across aligned sequences. Software such as PAML implements likelihood-based models to compute dN/dS from alignments, flagging sites or genes with values near 1 as potentially neutral. Functional assays directly test the impact of candidate mutations on protein activity through in vitro expression and biochemical evaluation. In these experiments, site-directed mutagenesis introduces specific variants into a gene, followed by recombinant protein production in systems like E. coli or yeast, and assessment of enzymatic activity, stability, or binding affinity using techniques such as fluorescence spectroscopy or enzyme kinetics. Mutations resulting in no measurable loss of function—comparable to wild-type performance—are classified as neutral, as demonstrated in studies of enzymes like beta-lactamase where variants retained full catalytic efficiency. High-throughput variants of these assays, including deep mutational scanning, systematically evaluate thousands of mutations via coupled expression and selection, identifying neutral ones as those maintaining wild-type-like fitness in reporter gene contexts. Population genetics methods scan distributions in population samples to detect deviations from expectations, inferring neutrality from the absence of selection-driven distortions. Under neutrality, polymorphisms should follow predictions from models like the infinite-sites model, with site frequency spectra showing an excess of low-frequency variants as expected under neutrality (approximately proportional to 1/i); tools such as statistic quantify this by comparing observed polymorphism patterns to neutral simulations, where values near zero indicate no selection and thus potential neutrality. This approach analyzes genomic from cohorts, flagging variants with balanced frequencies as neutral candidates. Modern tools leverage -based to empirically test effects on organismal , introducing precise variants into model like , , or and measuring growth rates or survival. In competitive assays, edited s compete against wild-type, with sequencing tracking frequencies; neutral mutations show no frequency shift. High-throughput sequencing complements this by enabling variant calling from edited populations to accurately identify and quantify polymorphisms introduced via CRISPR, thus validating neutrality at scale.

Measurement approaches

One primary quantitative approach to measuring the neutrality of involves the dN/dS ratio, which compares the rate of nonsynonymous substitutions (dN, those altering the sequence) to the rate of synonymous substitutions (, those preserving the ). This ratio is calculated by first estimating the number of nonsynonymous (S_N) and synonymous (S_S) substitutions per site between two sequences, accounting for multiple substitutions at the same site using models like the Jukes-Cantor correction, and then deriving dN = - (3/4) ln(1 - (4/3) p_N) and = - (3/4) ln(1 - (4/3) p_S), where p_N and p_S are the proportions of nonsynonymous and synonymous differences, respectively; under neutrality, dN/dS ≈ 1, indicating equal fixation rates driven by . The McDonald-Kreitman (MK) test provides another framework for assessing neutrality by contrasting within-species polymorphism to between-species divergence, using a 2x2 of synonymous and nonsynonymous sites: polymorphisms (P_S and P_N) versus fixed differences (D_S and D_N). Neutrality is tested via a statistic, where deviation from equality (P_N/P_S ≈ D_N/D_S) suggests selection; the proportion of adaptive substitutions is estimated as α = 1 - (P_N D_S)/(P_S D_N), with values near zero supporting neutrality. Site-frequency (SFS) analysis evaluates by examining the distribution of frequencies in a sample, where under , an excess of variants (low-frequency alleles) is expected due to recent mutations fixed by drift rather than selection. The expected SFS under the model folds the unfolded , showing the number of sites with i proportional to 1/i for i = 1 to 2n-1 in a sample of n diploids, with variants comprising most of the as a signature of drift-dominated processes. Statistical models based on neutral coalescent theory further quantify neutrality through likelihood-based inference, comparing observed data to expectations under drift and mutation. For instance, the expected heterozygosity H at a locus is given by H = 4N_e \mu where N_e is the and \mu is the per ; deviations from this equilibrium value, tested via maximum likelihood methods, indicate departures from neutrality.

Evolutionary Significance

Integration with evolutionary theory

The , proposed by in 1968, marked a significant shift from the prevailing view that was the primary driver of all change. Traditional Darwinian perspectives emphasized adaptive evolution through selection acting on phenotypic traits, but observations of nearly constant rates of molecular change across diverse taxa—despite long periods of phenotypic —challenged this selection-only paradigm. Neutral theory posits that the majority of molecular-level changes arise from random of selectively neutral mutations, which neither enhance nor impair , thereby explaining the observed uniformity in independent of adaptive pressures. This framework complements by delineating distinct roles for neutral processes and selection: neutral mutations account for " sequences and silent synonymous substitutions that accumulate without phenotypic consequences, while remains the dominant force shaping adaptive traits and organismal fitness. Kimura argued that neutral evolution operates primarily at the molecular level, coexisting with selection-driven changes at higher phenotypic levels, thus integrating drift as a complementary within the modern synthesis rather than supplanting it. This compatibility allows to incorporate molecular data, recognizing that much genomic evolution proceeds neutrally while selection acts on a subset of functionally significant variants. The integration sparked enduring debates between neutralists and selectionists. Selectionist critiques, exemplified by Stephen Jay Gould's advocacy for , contended that overemphasized drift at the expense of selection's role in generating , arguing that rapid adaptive bursts better explain evolutionary patterns than gradual neutral accumulation. Neutralists countered that drift of neutral variants provides a foundation for , enabling and through random fixation without requiring constant selection, and that molecular supports neutrality as the baseline for most genomic changes. These exchanges highlighted tensions but ultimately enriched evolutionary by clarifying the interplay between drift and selection in shaping . Recent challenges, such as the 2025 adaptive tracking , suggest higher prevalence of beneficial in microbial and , potentially revising the balance toward more adaptive processes. In contemporary evolutionary synthesis, neutral mutations serve as a null model for identifying genuine adaptations, with genome-wide studies revealing that a substantial proportion of aligns with expectations, underscoring their role as a key source of molecular diversity. This baseline facilitates detection of selective signals amid pervasive neutrality, as evidenced by analyses showing most polymorphisms fixed by drift rather than selection. Such findings reinforce theory's foundational status, providing a rigorous framework for interpreting genomic data and resolving long-standing debates on evolutionary mechanisms.

Application to molecular clocks

Neutral mutations form the basis of the molecular clock hypothesis, which posits that these changes accumulate at a relatively constant rate over time due to , independent of . Under neutral theory, the rate of for neutral alleles equals the neutral mutation rate μ, leading to a predictable between lineages. The time since divergence t can thus be estimated using the formula t = \frac{d}{2\mu}, where d represents the observed (typically the number of substitutions per site) between two species, and the factor of 2 accounts for mutations accumulating independently in each . This , first articulated in the context of , allows for the inference of evolutionary timelines from sequence data. To apply the , substitution rates must be calibrated using independent evidence, such as records or well-documented events, to determine μ for specific lineages. For instance, -calibrated clocks often anchor rates to known times, revealing variations in clock-like behavior influenced by factors like , which affects the number of reproductive cycles and thus mutation opportunities per unit calendar time. Shorter s generally accelerate the clock, as seen in comparisons across mammals where exhibit faster rates than due to more rapid generations. These calibrations enable reliable dating when neutral sites, such as synonymous substitutions, are prioritized to minimize selective biases. In practice, neutral mutation-based clocks have dated key speciation events, including the divergence of humans and chimpanzees at approximately 6–7 million years ago, derived from calibrated genomic comparisons showing consistent neutral substitution rates. Genome-scale analyses further enhance precision in phylogenetics by aggregating neutral variants across thousands of loci, constructing robust timetrees for diverse clades like or birds, and revealing fine-scale evolutionary histories. These applications have revolutionized fields like and conservation by providing temporal frameworks for lineage splits without relying solely on paleontological data. Despite these strengths, molecular clocks exhibit rate heterogeneity across lineages, arising from differences in mutation processes, population sizes, or environmental factors, which can violate strict clock assumptions. Such variations lead to in divergence estimates, particularly over deep timescales. To address this, relaxed clock models have been developed, including Bayesian approaches that allow rates to evolve autocorrelated along branches or vary stochastically while integrating priors and sequence data for probabilistic inference. These models, implemented in software like , improve accuracy by accommodating non-clock-like behavior without abandoning the neutral foundation.

References

  1. [1]
    The neutral theory - Understanding Evolution
    The neutral theory of molecular evolution suggests that most of the genetic variation in populations is the result of mutation and genetic drift and not ...
  2. [2]
    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.
  3. [3]
    Mutations (article) | Khan Academy
    Neutral mutations have no observable effect on an organism's traits. For example, some gene mutations do not lead to amino acid changes, and so do not affect ...<|control11|><|separator|>
  4. [4]
    The importance of the Neutral Theory in 1968 and 50 years on - NIH
    Furthermore, the rate of substitution of neutral mutations between species is equal to the mutation rate (Kimura 1968). A critical first extension of this ...
  5. [5]
    Derivation of the relationship between neutral mutation and fixation ...
    A mutation whose fixation is independent of natural selection is termed a neutral mutation. Therefore selective neutrality of a mutation can be defined by ...
  6. [6]
    None
    No readable text found in the HTML.<|control11|><|separator|>
  7. [7]
    Definition of mutation - NCI Dictionary of Genetics Terms
    Although the term often has a negative connotation, mutations (including polymorphisms) can be harmful, beneficial, or neutral in their effect on cell function.
  8. [8]
    Mutation Rates and Gene Location: Some Like It Hot - PMC - NIH
    Neutral mutations can also occur within gene-coding regions. For example, there are many instances where more than one codon—say, CUU, CUC, CUA, CUG—specify the ...Missing: non- | Show results with:non-
  9. [9]
    The Extinction Dynamics of Bacterial Pseudogenes - PMC - NIH
    Aug 5, 2010 · Pseudogenes are usually considered to be completely neutral sequences whose evolution is shaped by random mutations and chance events.<|control11|><|separator|>
  10. [10]
    The distribution of fitness effects of new mutations - Nature Reviews Genetics
    ### Summary of Mutation Classification from https://www.nature.com/articles/nrg2146
  11. [11]
    Near-neutrality in evolution of genes and gene regulation - PNAS
    The most abundant mutations are nucleotide substitutions in binding sites for transcription factors. Under mutation and selection for maintaining expression ...
  12. [12]
    The Neutral Theory of Molecular Evolution
    Motoo Kimura. Publisher: Cambridge University Press. Online ... He first proposed the theory in 1968 to explain the unexpectedly high rate of evolutionary ...
  13. [13]
    Evolution in Mendelian Populations - PMC - NIH
    Evolution in Mendelian Populations ... Sewall Wright. 1University of Chicago, Chicago, Illinois. Find articles by Sewall Wright. 1. Author information
  14. [14]
    Non-Darwinian Evolution - Science
    Non-Darwinian Evolution: Most evolutionary change in proteins may be due to neutral mutations and genetic drift. Jack Lester King and Thomas H. JukesAuthors ...
  15. [15]
    Slightly Deleterious Mutant Substitutions in Evolution - Nature
    Nov 9, 1973 · Slightly Deleterious Mutant Substitutions in Evolution. TOMOKO OHTA. Nature volume 246, pages 96–98 (1973)Cite this article.
  16. [16]
    Molecular Mechanisms and the Significance of Synonymous Mutations
    Jan 20, 2024 · Synonymous mutations result from the degeneracy of the genetic code. Most amino acids are encoded by two or more codons.
  17. [17]
    Synonymous mutations break their silence | Nature Reviews Genetics
    The degeneracy of the genetic code means that many mutations in coding sequences, especially at the third base of codons, do not affect protein sequence and ...
  18. [18]
    Silent Mutation - an overview | ScienceDirect Topics
    For example, the codon CUU codes for the amino acid leucine. Any mutation in the third position of the nucleic acid sequence, that is, to CUC, CUA, or CUG ...
  19. [19]
    Mutation rates in mammalian genomes - PMC - NIH
    Our results suggest that the average mammalian genome mutation rate is 2.2 × 10 −9 per base pair per year.
  20. [20]
    Are radical and conservative substitution rates useful statistics in ...
    A DNA mutation in a protein coding gene which causes an amino acid change can be classified as "conservative" or "radical" depending on the magnitude of the ...
  21. [21]
    Amino Acid Properties, Substitution Rates, and the Nearly Neutral ...
    For example, Smith (2003) defined amino acid substitutions as either “radical” (physicochemically different) or “conservative” (physicochemically similar) ...Introduction · Results · Discussion · Materials and Methods
  22. [22]
    The tangled bank of amino acids - PMC - NIH
    The use of amino acid substitution matrices to model protein evolution has yielded important insights into both the evolutionary process and the properties ...
  23. [23]
    Detection of neutral amino acid substitutions in proteins. | PNAS
    ### Summary of Neutral Amino Acid Substitutions
  24. [24]
    DNA polymerase active site is highly mutable - PubMed Central - NIH
    A wide variety of amino acid substitutions were obtained at sites that are evolutionarily maintained, and conservative substitutions predominate at regions that ...
  25. [25]
    Neutrality tests of conservative-radical amino acid changes in nuclear
    Dec 30, 2000 · Neutrality tests of conservative-radical amino acid changes in nuclear- and mitochondrially-encoded proteins · Results. For initial comparison, ...
  26. [26]
    Detection of neutral amino acid substitutions in proteins - PMC - NIH
    Mouse hemoglobin tetramers that differ only by the substitution of alanine for glycine in the alpha-globin chains are resolved by several millimeters with the ...Missing: examples | Show results with:examples<|separator|>
  27. [27]
    dN/dS-Based Methods Detect Positive Selection Linked to Trade ...
    The dN/dS ratio between nonsynonymous and synonymous substitution rates has been used extensively to identify codon positions involved in adaptive processes.
  28. [28]
    Statistical Method for Testing the Neutral Mutation Hypothesis by ...
    A statistical method for testing the neutral mutation hypothesis is developed. This method needs only the data of DNA polymorphism.
  29. [29]
    Why Time Matters: Codon Evolution and the Temporal Dynamics of ...
    Oct 14, 2013 · Originally the dN/dS ratio was developed in a phylogenetics context, and its estimation was based on codon sequences of distantly related ...
  30. [30]
    A competitive precision CRISPR method to identify the fitness effects ...
    Sep 26, 2022 · Here we describe a competitive genome editing method that measures the effect of mutations on molecular functions, based on precision CRISPR editing.
  31. [31]
    Adaptive protein evolution at the Adh locus in Drosophila - Nature
    Jun 20, 1991 · Cite this article. McDonald, J., Kreitman, M. Adaptive protein evolution at the Adh locus in Drosophila. Nature 351, 652–654 (1991). https ...
  32. [32]
    Error | Proceedings of the National Academy of Sciences
    Insufficient relevant content. The provided text from https://www.pnas.org/doi/10.1073/pnas.61.4.1252 does not contain the full article or specific details about Kimura's argument on neutral mutations and molecular evolution. It only includes a 404 error page, metadata, and unrelated chemical element data. No substantive information about the topic is available.
  33. [33]
  34. [34]
    Theorists Debate How 'Neutral' Evolution Really Is | Quanta Magazine
    Nov 8, 2018 · Motoo Kimura proposed in 1968 that most mutations might be neutral in effect rather than beneficial or harmful, and that shifts in the frequency ...
  35. [35]
  36. [36]
    The Molecular Clock and Estimating Species Divergence - Nature
    The molecular clock hypothesis states that DNA and protein sequences evolve at a rate that is relatively constant over time and among different organisms.Missing: formula | Show results with:formula
  37. [37]
    Neutrality and Molecular Clocks | Learn Science at Scitable - Nature
    The term 'molecular clock' was initially coined to describe changes in amino acids occurring in proportional to time since species divergence.<|control11|><|separator|>
  38. [38]
    Placing confidence limits on the molecular age of ... - PubMed - NIH
    For a given the ape-OWM divergence time, the 95% confidence interval of the human-chimpanzee divergence ranges from -12% to 19% of the estimated time.
  39. [39]
    Modeling Substitution Rate Evolution across Lineages and Relaxing ...
    A summary of the most widely used Bayesian dating programs that relax the molecular clock assumption, detailing their available models of rate evolution and ...Early Models Of Rate... · Non-Bayesian Approaches · Literature Cited