Neutral theory of molecular evolution

Last updated
The Neutral Theory of Molecular Evolution.jpg

The neutral theory of molecular evolution holds that most evolutionary changes occur at the molecular level, and most of the variation within and between species are due to random genetic drift of mutant alleles that are selectively neutral. The theory applies only for evolution at the molecular level, and is compatible with phenotypic evolution being shaped by natural selection as postulated by Charles Darwin.

Contents

The neutral theory allows for the possibility that most mutations are deleterious, but holds that because these are rapidly removed by natural selection, they do not make significant contributions to variation within and between species at the molecular level. A neutral mutation is one that does not affect an organism's ability to survive and reproduce.

The neutral theory assumes that most mutations that are not deleterious are neutral rather than beneficial. Because only a fraction of gametes are sampled in each generation of a species, the neutral theory suggests that a mutant allele can arise within a population and reach fixation by chance, rather than by selective advantage. [1]

The theory was introduced by the Japanese biologist Motoo Kimura in 1968, and independently by two American biologists Jack Lester King and Thomas Hughes Jukes in 1969, and described in detail by Kimura in his 1983 monograph The Neutral Theory of Molecular Evolution . The proposal of the neutral theory was followed by an extensive "neutralist–selectionist" controversy over the interpretation of patterns of molecular divergence and gene polymorphism, peaking in the 1970s and 1980s.

Neutral theory is frequently used as the null hypothesis, as opposed to adaptive explanations, for describing the emergence of morphological or genetic features in organisms and populations. This has been suggested in a number of areas, including in explaining genetic variation between populations of one nominal species, [2] the emergence of complex subcellular machinery, [3] and the convergent emergence of several typical microbial morphologies. [4]

Origins

While some scientists, such as Freese (1962) [5] and Freese and Yoshida (1965), [6] had suggested that neutral mutations were probably widespread, the original mathematical derivation of the theory had been published by R.A. Fisher in 1930. [7] Fisher, however, gave a reasoned argument for believing that, in practice, neutral gene substitutions would be very rare. [8] A coherent theory of neutral evolution was first proposed by Motoo Kimura in 1968 [9] and by King and Jukes independently in 1969. [10] Kimura initially focused on differences among species; King and Jukes focused on differences within species.

Many molecular biologists and population geneticists also contributed to the development of the neutral theory. [1] [11] [12] The principles of population genetics, established by J.B.S. Haldane, R.A. Fisher, and Sewall Wright, created a mathematical approach to analyzing gene frequencies that contributed to the development of Kimura's theory.

Haldane's dilemma regarding the cost of selection was used as motivation by Kimura. Haldane estimated that it takes about 300 generations for a beneficial mutation to become fixed in a mammalian lineage, meaning that the number of substitutions (1.5 per year) in the evolution between humans and chimpanzees was too high to be explained by beneficial mutations.

Functional constraint

The neutral theory holds that as functional constraint diminishes, the probability that a mutation is neutral rises, and so should the rate of sequence divergence.

When comparing various proteins, extremely high evolutionary rates were observed in proteins such as fibrinopeptides and the C chain of the proinsulin molecule, which both have little to no functionality compared to their active molecules. Kimura and Ohta also estimated that the alpha and beta chains on the surface of a hemoglobin protein evolve at a rate almost ten times faster than the inside pockets, which would imply that the overall molecular structure of hemoglobin is less significant than the inside where the iron-containing heme groups reside. [13]

There is evidence that rates of nucleotide substitution are particularly high in the third position of a codon, where there is little functional constraint. [14] This view is based in part on the degenerate genetic code, in which sequences of three nucleotides (codons) may differ and yet encode the same amino acid (GCC and GCA both encode alanine, for example). Consequently, many potential single-nucleotide changes are in effect "silent" or "unexpressed" (see synonymous or silent substitution). Such changes are presumed to have little or no biological effect. [15]

Quantitative theory

Kimura also developed the infinite sites model (ISM) to provide insight into evolutionary rates of mutant alleles. If were to represent the rate of mutation of gametes per generation of individuals, each with two sets of chromosomes, the total number of new mutants in each generation is . Now let represent the evolution rate in terms of a mutant allele becoming fixed in a population. [16]

According to ISM, selectively neutral mutations appear at rate in each of the copies of a gene, and fix with probability . Because any of the genes have the ability to become fixed in a population, is equal to , resulting in the rate of evolutionary rate equation:

This means that if all mutations were neutral, the rate at which fixed differences accumulate between divergent populations is predicted to be equal to the per-individual mutation rate, independent of population size. When the proportion of mutations that are neutral is constant, so is the divergence rate between populations. This provides a rationale for the molecular clock, which predated neutral theory. [17] The ISM also demonstrates a constancy that is observed in molecular lineages.

This stochastic process is assumed to obey equations describing random genetic drift by means of accidents of sampling, rather than for example genetic hitchhiking of a neutral allele due to genetic linkage with non-neutral alleles. After appearing by mutation, a neutral allele may become more common within the population via genetic drift. Usually, it will be lost, or in rare cases it may become fixed, meaning that the new allele becomes standard in the population.

According to the neutral theory of molecular evolution, the amount of genetic variation within a species should be proportional to the effective population size.

The "neutralist–selectionist" debate

A heated debate arose when Kimura's theory was published, largely revolving around the relative percentages of polymorphic and fixed alleles that are "neutral" versus "non-neutral".

A genetic polymorphism means that different forms of particular genes, and hence of the proteins that they produce, are co-existing within a species. Selectionists claimed that such polymorphisms are maintained by balancing selection, while neutralists view the variation of a protein as a transient phase of molecular evolution. [1] Studies by Richard K. Koehn and W. F. Eanes demonstrated a correlation between polymorphism and molecular weight of their molecular subunits. [18] This is consistent with the neutral theory assumption that larger subunits should have higher rates of neutral mutation. Selectionists, on the other hand, contribute environmental conditions to be the major determinants of polymorphisms rather than structural and functional factors. [16]

According to the neutral theory of molecular evolution, the amount of genetic variation within a species should be proportional to the effective population size. Levels of genetic diversity vary much less than census population sizes, giving rise to the "paradox of variation" . [19] While high levels of genetic diversity were one of the original arguments in favor of neutral theory, the paradox of variation has been one of the strongest arguments against neutral theory.

There are a large number of statistical methods for testing whether neutral theory is a good description of evolution (e.g., McDonald-Kreitman test [20] ), and many authors claimed detection of selection. [21] [22] [23] [24] [25] [26] Some researchers have nevertheless argued that the neutral theory still stands, while expanding the definition of neutral theory to include background selection at linked sites. [27]

Nearly neutral theory

Tomoko Ohta also emphasized the importance of nearly neutral mutations, in particularly slightly deleterious mutations. [28] The Nearly neutral theory stems from the prediction of neutral theory that the balance between selection and genetic drift depends on effective population size. [29] Nearly neutral mutations are those that carry selection coefficients less than the inverse of twice the effective population size. [30] The population dynamics of nearly neutral mutations are only slightly different from those of neutral mutations unless the absolute magnitude of the selection coefficient is greater than 1/N, where N is the effective population size in respect of selection. [1] [11] [12] The effective population size affects whether slightly deleterious mutations can be treated as neutral or as deleterious. [31] In large populations, selection can decrease the frequency of slightly deleterious mutations, therefore acting as if they are deleterious. However, in small populations, genetic drift can more easily overcome selection, causing slightly deleterious mutations to act as if they are neutral and drift to fixation or loss. [31]

Constructive neutral evolution

The groundworks for the theory of constructive neutral evolution (CNE) was laid by two papers in the 1990s. [32] [33] [34] Constructive neutral evolution is a theory which suggests that complex structures and processes can emerge through neutral transitions. Although a separate theory altogether, the emphasis on neutrality as a process whereby neutral alleles are randomly fixed by genetic drift finds some inspiration from the earlier attempt by the neutral theory to invoke its importance in evolution. [34] Conceptually, there are two components A and B (which may represent two proteins) which interact with each other. A, which performs a function for the system, does not depend on its interaction with B for its functionality, and the interaction itself may have randomly arisen in an individual with the ability to disappear without an effect on the fitness of A. This present yet currently unnecessary interaction is therefore called an "excess capacity" of the system. However, a mutation may occur which compromises the ability of A to perform its function independently. However, the A:B interaction that has already emerged sustains the capacity of A to perform its initial function. Therefore, the emergence of the A:B interaction "presuppresses" the deleterious nature of the mutation, making it a neutral change in the genome that is capable of spreading through the population via random genetic drift. Hence, A has gained a dependency on its interaction with B. [35] In this case, the loss of B or the A:B interaction would have a negative effect on fitness and so purifying selection would eliminate individuals where this occurs. While each of these steps are individually reversible (for example, A may regain the capacity to function independently or the A:B interaction may be lost), a random sequence of mutations tends to further reduce the capacity of A to function independently and a random walk through the dependency space may very well result in a configuration in which a return to functional independence of A is far too unlikely to occur, which makes CNE a one-directional or "ratchet-like" process. [36] CNE, which does not invoke adaptationist mechanisms for the origins of more complex systems (which involve more parts and interactions contributing to the whole), has seen application in the understanding of the evolutionary origins of the spliceosomal eukaryotic complex, RNA editing, additional ribosomal proteins beyond the core, the emergence of long-noncoding RNA from junk DNA, and so forth. [37] [38] [39] [40] In some cases, ancestral sequence reconstruction techniques have afforded the ability for experimental demonstration of some proposed examples of CNE, as in heterooligomeric ring protein complexes in some fungal lineages. [41]

CNE has also been put forwards as the null hypothesis for explaining complex structures, and thus adaptationist explanations for the emergence of complexity must be rigorously tested on a case-by-case basis against this null hypothesis prior to acceptance. Grounds for invoking CNE as a null include that it does not presume that changes offered an adaptive benefit to the host or that they were directionally selected for, while maintaining the importance of more rigorous demonstrations of adaptation when invoked so as to avoid the excessive flaws of adaptationism criticized by Gould and Lewontin. [42] [3] [43]

Empirical evidence for the neutral theory

Predictions derived from the neutral theory are generally supported in studies of molecular evolution. [44] One of corollaries of the neutral theory is that the efficiency of positive selection is higher in populations or species with higher effective population sizes. [45] This relationship between the effective population size and selection efficiency was evidenced by genomic studies of species including chimpanzee and human [45] and domesticated species. [46] In small populations (e.g., a population bottleneck during a speciation event), slightly deleterious mutations should accumulate. Data from various species supports this prediction in that the ratio of nonsynonymous to synonymous nucleotide substitutions between species generally exceeds that within species. [31] In addition, nucleotide and amino acid substitutions generally accumulate over time in a linear fashion, which is consistent with neutral theory. [44] Arguments against the neutral theory cite evidence of widespread positive selection and selective sweeps in genomic data. [47] Empirical support for the neutral theory may vary depending on the type of genomic data studied and the statistical tools used to detect positive selection. [44] For example, Bayesian methods for the detection of selected codon sites and McDonald-Kreitman tests have been criticized for their rate of erroneous identification of positive selection. [31] [44]

See also

Related Research Articles

Microevolution is the change in allele frequencies that occurs over time within a population. This change is due to four different processes: mutation, selection, gene flow and genetic drift. This change happens over a relatively short amount of time compared to the changes termed macroevolution.

<span class="mw-page-title-main">Mutation</span> Alteration in the nucleotide sequence of a genome

In biology, a mutation is an alteration in the nucleic acid sequence of the genome of an organism, virus, or extrachromosomal DNA. Viral genomes contain either DNA or RNA. Mutations result from errors during DNA or viral replication, mitosis, or meiosis or other types of damage to DNA, which then may undergo error-prone repair, cause an error during other forms of repair, or cause an error during replication. Mutations may also result from insertion or deletion of segments of DNA due to mobile genetic elements.

Genetic drift, also known as random genetic drift, allelic drift or the Wright effect, is the change in the frequency of an existing gene variant (allele) in a population due to random chance.

Molecular evolution is the process of change in the sequence composition of cellular molecules such as DNA, RNA, and proteins across generations. The field of molecular evolution uses principles of evolutionary biology and population genetics to explain patterns in these changes. Major topics in molecular evolution concern the rates and impacts of single nucleotide changes, neutral evolution vs. natural selection, origins of new genes, the genetic nature of complex traits, the genetic basis of speciation, the evolution of development, and ways that evolutionary forces influence genomic and phenotypic changes.

Population genetics is a subfield of genetics that deals with genetic differences within and among populations, and is a part of evolutionary biology. Studies in this branch of biology examine such phenomena as adaptation, speciation, and population structure.

The molecular clock is a figurative term for a technique that uses the mutation rate of biomolecules to deduce the time in prehistory when two or more life forms diverged. The biomolecular data used for such calculations are usually nucleotide sequences for DNA, RNA, or amino acid sequences for proteins. The benchmarks for determining the mutation rate are often fossil or archaeological dates. The molecular clock was first tested in 1962 on the hemoglobin protein variants of various animals, and is commonly used in molecular evolution to estimate times of speciation or radiation. It is sometimes called a gene clock or an evolutionary clock.

<span class="mw-page-title-main">Motoo Kimura</span> Japanese biologist

Motoo Kimura was a Japanese biologist best known for introducing the neutral theory of molecular evolution in 1968. He became one of the most influential theoretical population geneticists. He is remembered in genetics for his innovative use of diffusion equations to calculate the probability of fixation of beneficial, deleterious, or neutral alleles. Combining theoretical population genetics with molecular evolution data, he also developed the neutral theory of molecular evolution in which genetic drift is the main force changing allele frequencies. James F. Crow, himself a renowned population geneticist, considered Kimura to be one of the two greatest evolutionary geneticists, along with Gustave Malécot, after the great trio of the modern synthesis, Ronald Fisher, J. B. S. Haldane, and Sewall Wright.

Genetic load is the difference between the fitness of an average genotype in a population and the fitness of some reference genotype, which may be either the best present in a population, or may be the theoretically optimal genotype. The average individual taken from a population with a low genetic load will generally, when grown in the same conditions, have more surviving offspring than the average individual from a population with a high genetic load. Genetic load can also be seen as reduced fitness at the population level compared to what the population would have if all individuals had the reference high-fitness genotype. High genetic load may put a population in danger of extinction.

<span class="mw-page-title-main">Tomoko Ohta</span> Japanese biologist

Tomoko Ohta is a Japanese scientist and Professor Emeritus of the National Institute of Genetics. Ohta works on population genetics/molecular evolution and is known for developing the nearly neutral theory of evolution.

Genetic hitchhiking, also called genetic draft or the hitchhiking effect, is when an allele changes frequency not because it itself is under natural selection, but because it is near another gene that is undergoing a selective sweep and that is on the same DNA chain. When one gene goes through a selective sweep, any other nearby polymorphisms that are in linkage disequilibrium will tend to change their allele frequencies too. Selective sweeps happen when newly appeared mutations are advantageous and increase in frequency. Neutral or even slightly deleterious alleles that happen to be close by on the chromosome 'hitchhike' along with the sweep. In contrast, effects on a neutral locus due to linkage disequilibrium with newly appeared deleterious mutations are called background selection. Both genetic hitchhiking and background selection are stochastic (random) evolutionary forces, like genetic drift.

<i>The Neutral Theory of Molecular Evolution</i>

The Neutral Theory of Molecular Evolution is an influential monograph written in 1983 by Japanese evolutionary biologist Motoo Kimura. While the neutral theory of molecular evolution existed since his article in 1968, Kimura felt the need to write a monograph with up-to-date information and evidences showing the importance of his theory in evolution.

Neutral mutations are changes in DNA sequence that are neither beneficial nor detrimental to the ability of an organism to survive and reproduce. In population genetics, mutations in which natural selection does not affect the spread of the mutation in a species are termed neutral mutations. Neutral mutations that are inheritable and not linked to any genes under selection will be lost or will replace all other alleles of the gene. That loss or fixation of the gene proceeds based on random sampling known as genetic drift. A neutral mutation that is in linkage disequilibrium with other alleles that are under selection may proceed to loss or fixation via genetic hitchhiking and/or background selection.

In population genetics, fixation is the change in a gene pool from a situation where there exists at least two variants of a particular gene (allele) in a given population to a situation where only one of the alleles remains. That is, the allele becomes fixed. In the absence of mutation or heterozygote advantage, any allele must eventually either be lost completely from the population, or fixed, i.e. permanently established at 100% frequency in the population. Whether a gene will ultimately be lost or fixed is dependent on selection coefficients and chance fluctuations in allelic proportions. Fixation can refer to a gene in general or particular nucleotide position in the DNA chain (locus).

The nearly neutral theory of molecular evolution is a modification of the neutral theory of molecular evolution that accounts for the fact that not all mutations are either so deleterious such that they can be ignored, or else neutral. Slightly deleterious mutations are reliably purged only when their selection coefficient are greater than one divided by the effective population size. In larger populations, a higher proportion of mutations exceed this threshold for which genetic drift cannot overpower selection, leading to fewer fixation events and so slower molecular evolution.

The history of molecular evolution starts in the early 20th century with "comparative biochemistry", but the field of molecular evolution came into its own in the 1960s and 1970s, following the rise of molecular biology. The advent of protein sequencing allowed molecular biologists to create phylogenies based on sequence comparison, and to use the differences between homologous sequences as a molecular clock to estimate the time since the last common ancestor. In the late 1960s, the neutral theory of molecular evolution provided a theoretical basis for the molecular clock, though both the clock and the neutral theory were controversial, since most evolutionary biologists held strongly to panselectionism, with natural selection as the only important cause of evolutionary change. After the 1970s, nucleic acid sequencing allowed molecular evolution to reach beyond proteins to highly conserved ribosomal RNA sequences, the foundation of a reconceptualization of the early history of life.

"Non-Darwinian Evolution" is a scientific paper written by Jack Lester King and Thomas H. Jukes and published in 1969. It is credited, along with Motoo Kimura's 1968 paper "Evolutionary Rate at the Molecular Level", with proposing what became known as the neutral theory of molecular evolution. The paper brings together a wide variety of evidence, ranging from protein sequence comparisons to studies of the Treffers mutator gene in E. coli to analysis of the genetic code to comparative immunology, to argue that most protein evolution is due to neutral mutations and genetic drift. It was published in the journal Science on May 16, 1969.

A nonsynonymous substitution is a nucleotide mutation that alters the amino acid sequence of a protein. Nonsynonymous substitutions differ from synonymous substitutions, which do not alter amino acid sequences and are (sometimes) silent mutations. As nonsynonymous substitutions result in a biological change in the organism, they are subject to natural selection.

A neutral network is a set of genes all related by point mutations that have equivalent function or fitness. Each node represents a gene sequence and each line represents the mutation connecting two sequences. Neutral networks can be thought of as high, flat plateaus in a fitness landscape. During neutral evolution, genes can randomly move through neutral networks and traverse regions of sequence space which may have consequences for robustness and evolvability.

The rate of evolution is quantified as the speed of genetic or morphological change in a lineage over a period of time. The speed at which a molecular entity evolves is of considerable interest in evolutionary biology since determining the evolutionary rate is the first step in characterizing its evolution. Calculating rates of evolutionary change is also useful when studying phenotypic changes in phylogenetic comparative biology. In either case, it can be beneficial to consider and compare both genomic data and paleontological data, especially in regards to estimating the timing of divergence events and establishing geological time scales.

<span class="mw-page-title-main">Epistasis</span> Dependence of a gene mutations phenotype on mutations in other genes

Epistasis is a phenomenon in genetics in which the effect of a gene mutation is dependent on the presence or absence of mutations in one or more other genes, respectively termed modifier genes. In other words, the effect of the mutation is dependent on the genetic background in which it appears. Epistatic mutations therefore have different effects on their own than when they occur together. Originally, the term epistasis specifically meant that the effect of a gene variant is masked by that of different gene.

References

  1. 1 2 3 4 Kimura, Motoo (1983). The neutral theory of molecular evolution. Cambridge University Press. ISBN   978-0-521-31793-1.
  2. Fenchel, Tom (2005-11-11). "Cosmopolitan microbes and their 'cryptic' species". Aquatic Microbial Ecology. 41 (1): 49–54. doi: 10.3354/ame041049 . ISSN   0948-3055.
  3. 1 2 Koonin, Eugene V. (2016). "Splendor and misery of adaptation, or the importance of neutral null for understanding evolution". BMC Biology. 14 (1): 114. doi: 10.1186/s12915-016-0338-2 . ISSN   1741-7007. PMC   5180405 . PMID   28010725.
  4. Lahr, Daniel J. G.; Laughinghouse, Haywood Dail; Oliverio, Angela M.; Gao, Feng; Katz, Laura A. (2014). "How discordant morphological and molecular evolution among microorganisms can revise our notions of biodiversity on Earth: Prospects & Overviews". BioEssays. 36 (10): 950–959. doi:10.1002/bies.201400056. PMC   4288574 . PMID   25156897.
  5. Freese, E. (July 1962). "On the evolution of the base composition of DNA". Journal of Theoretical Biology. 3 (1): 82–101. Bibcode:1962JThBi...3...82F. doi:10.1016/S0022-5193(62)80005-8.
  6. Freese, E.; Yoshida, A. (1965). "The role of mutations in evolution.". In Bryson, V.; Vogel, H. J. (eds.). Evolving Genes and Proteins. New York: Academic. pp. 341–355.
  7. Fisher R.A. 1930. The distribution of gene ratios for rare mutations. Proceedings of the Royal Society of Edinburgh volume 50, pages 205-230.
  8. R.J. Berry, T.J. Crawford, G.M. Hewitt 1992. Genes in Ecology. Blackwell Scientific Publications, Oxford. pp.29-54 J.R.G.Turner: Stochastic processes in populations: the horse behind the cart?.
  9. Kimura, Motoo (February 1968). "Evolutionary rate at the molecular level". Nature. 217 (5129): 624–6. Bibcode:1968Natur.217..624K. doi:10.1038/217624a0. PMID   5637732. S2CID   4161261.
  10. King, J. L.; Jukes, T. H. (May 1969). "Non-Darwinian evolution". Science. 164 (3881): 788–98. Bibcode:1969Sci...164..788L. doi:10.1126/science.164.3881.788. PMID   5767777.
  11. 1 2 Nei, Masatoshi (December 2005). "Selectionism and neutralism in molecular evolution". Molecular Biology and Evolution. 22 (12): 2318–2342. doi:10.1093/molbev/msi242. PMC   1513187 . PMID   16120807.
  12. 1 2 Nei, Masatoshi (2013). Mutation-driven evolution. Oxford University Press.
  13. Kimura, M. (1969-08-01). "The Rate of Molecular Evolution Considered from the Standpoint of Population Genetics". Proceedings of the National Academy of Sciences. 63 (4): 1181–1188. Bibcode:1969PNAS...63.1181K. doi: 10.1073/pnas.63.4.1181 . ISSN   0027-8424. PMC   223447 . PMID   5260917.
  14. Bofkin, L.; Goldman, N. (2006-11-13). "Variation in Evolutionary Processes at Different Codon Positions". Molecular Biology and Evolution. 24 (2): 513–521. doi: 10.1093/molbev/msl178 . ISSN   0737-4038. PMID   17119011.
  15. Crick, F.H.C. (1989), "Codon—Anticodon Pairing: The Wobble Hypothesis", Molecular Biology, Elsevier, pp. 370–377, doi:10.1016/b978-0-12-131200-8.50026-5, ISBN   978-0-12-131200-8 , retrieved 2021-04-03
  16. 1 2 Kimura, Motoo (November 1979). "The neutral theory of molecular evolution". Scientific American . 241 (5): 98–100, 102, 108 passim. Bibcode:1979SciAm.241e..98K. doi:10.1038/scientificamerican1179-98. JSTOR   24965339. PMID   504979. S2CID   5119551.
  17. Zuckerkandl, Emile; Pauling, Linus B. (1962). "Molecular disease, evolution, and genetic heterogeneity" . In Kasha, M.; Pullman, B. (eds.). Horizons in Biochemistry. Academic Press. pp.  189–225.
  18. Eanes, Walter F. (November 1999). "Analysis of Selection on Enzyme Polymorphisms". Annual Review of Ecology and Systematics. 30 (1): 301–326. doi:10.1146/annurev.ecolsys.30.1.301.
  19. Lewontin, Richard C. (1973). The genetic basis of evolutionary change (4th printing ed.). Columbia University Press. ISBN   978-0231033923.
  20. Kreitman, M. (2000). "Methods to detect selection in populations with applications to the human". Annual Review of Genomics and Human Genetics. 1 (1): 539–59. doi: 10.1146/annurev.genom.1.1.539 . PMID   11701640.
  21. Fay, J. C.; Wyckoff, G. J.; Wu, C. I. (February 2002). "Testing the neutral theory of molecular evolution with genomic data from Drosophila". Nature. 415 (6875): 1024–6. Bibcode:2002Natur.415.1024F. doi:10.1038/4151024a. PMID   11875569. S2CID   4420010.
  22. Begun, D. J.; Holloway, A. K.; Stevens, K.; Hillier, L. W.; Poh, Y. P.; Hahn, M. W.; Nista, P. M.; Jones, C. D.; Kern, A. D.; Dewey, C. N.; Pachter, L.; Myers, E.; Langley, C. H. (November 2007). "Population genomics: whole-genome analysis of polymorphism and divergence in Drosophila simulans". PLOS Biology. 5 (11): e310. doi: 10.1371/journal.pbio.0050310 . PMC   2062478 . PMID   17988176.
  23. Shapiro, J. A.; Huang, W.; Zhang, C.; Hubisz, M. J.; Lu, J.; Turissini, D. A.; Fang, S.; Wang, H. Y.; Hudson, RR; Nielsen, R.; Chen, Z.; Wu, C. I. (February 2007). "Adaptive genic evolution in the Drosophila genomes". PNAS. 104 (7): 2271–6. Bibcode:2007PNAS..104.2271S. doi: 10.1073/pnas.0610385104 . PMC   1892965 . PMID   17284599.
  24. Hahn, M. W. (February 2008). "Toward a selection theory of molecular evolution". Evolution; International Journal of Organic Evolution. 62 (2): 255–65. doi: 10.1111/j.1558-5646.2007.00308.x . PMID   18302709.
  25. Akey, J. M. (May 2009). "Constructing genomic maps of positive selection in humans: where do we go from here?". Genome Research. 19 (5): 711–22. doi:10.1101/gr.086652.108. PMC   3647533 . PMID   19411596.
  26. Kern, A. D.; Hahn, M. W. (June 2018). "The Neutral Theory in Light of Natural Selection". Molecular Biology and Evolution. 35 (6): 1366–1371. doi:10.1093/molbev/msy092. PMC   5967545 . PMID   29722831.
  27. Jensen, J.D.; Payseur, B. A.; Stephan, W.; Aquadro C. F.; Lynch, M. Charlesworth, D.; Charlesworth, B. (January 2019). "The importance of the Neutral Theory in 1968 and 50 years on: A response to Kern and Hahn 2018". Evolution; International Journal of Organic Evolution. 73 (1): 111–114. doi:10.1111/evo.13650. PMC   6496948 . PMID   30460993.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  28. Ohta, T. (December 2002). "Near-neutrality in evolution of genes and gene regulation". Proceedings of the National Academy of Sciences of the United States of America. 99 (25): 16134–7. Bibcode:2002PNAS...9916134O. doi: 10.1073/pnas.252626899 . PMC   138577 . PMID   12461171.
  29. Ohta, Tomoko (1973). "Slightly Deleterious Mutant Substitutions in Evolution". Nature. 246 (5428): 96–98. Bibcode:1973Natur.246...96O. doi:10.1038/246096a0. ISSN   1476-4687. PMID   4585855. S2CID   4226804.
  30. Kimura, Motoo (1983). The neutral theory of molecular evolution. Cambridge University Press. ISBN   978-0-521-31793-1.
  31. 1 2 3 4 Hughes, Austin L. (2008). "Near Neutrality: the leading edge of the Neutral Theory of Molecular Evolution". Annals of the New York Academy of Sciences. 1133 (1): 162–179. Bibcode:2008NYASA1133..162H. doi:10.1196/annals.1438.001. ISSN   0077-8923. PMC   2707937 . PMID   18559820.
  32. Covello, Patrick S.; Gray, MichaelW. (1993). "On the evolution of RNA editing". Trends in Genetics. 9 (8): 265–268. doi:10.1016/0168-9525(93)90011-6. PMID   8379005.
  33. Stoltzfus, Arlin (1999). "On the Possibility of Constructive Neutral Evolution". Journal of Molecular Evolution. 49 (2): 169–181. Bibcode:1999JMolE..49..169S. doi:10.1007/PL00006540. ISSN   0022-2844. PMID   10441669. S2CID   1743092.
  34. 1 2 Muñoz-Gómez, Sergio A.; Bilolikar, Gaurav; Wideman, Jeremy G.; Geiler-Samerotte, Kerry (2021-04-01). "Constructive Neutral Evolution 20 Years Later". Journal of Molecular Evolution. 89 (3): 172–182. Bibcode:2021JMolE..89..172M. doi:10.1007/s00239-021-09996-y. ISSN   1432-1432. PMC   7982386 . PMID   33604782.
  35. Speijer, Dave (2011). "Does constructive neutral evolution play an important role in the origin of cellular complexity?: Making sense of the origins and uses of biological complexity". BioEssays. 33 (5): 344–349. doi:10.1002/bies.201100010. PMID   21381061. S2CID   205470421.
  36. Stoltzfus, Arlin (2012-10-13). "Constructive neutral evolution: exploring evolutionary theory's curious disconnect". Biology Direct. 7 (1): 35. doi: 10.1186/1745-6150-7-35 . ISSN   1745-6150. PMC   3534586 . PMID   23062217.
  37. Gray, Michael W.; Lukeš, Julius; Archibald, John M.; Keeling, Patrick J.; Doolittle, W. Ford (2010-11-12). "Irremediable Complexity?". Science. 330 (6006): 920–921. Bibcode:2010Sci...330..920G. doi:10.1126/science.1198594. ISSN   0036-8075. PMID   21071654. S2CID   206530279.
  38. Lukeš, Julius; Archibald, John M.; Keeling, Patrick J.; Doolittle, W. Ford; Gray, Michael W. (2011). "How a neutral evolutionary ratchet can build cellular complexity". IUBMB Life. 63 (7): 528–537. doi:10.1002/iub.489. PMID   21698757. S2CID   7306575.
  39. Lamech, Lilian T.; Mallam, Anna L.; Lambowitz, Alan M. (2014-12-23). Herschlag, Daniel (ed.). "Evolution of RNA-Protein Interactions: Non-Specific Binding Led to RNA Splicing Activity of Fungal Mitochondrial Tyrosyl-tRNA Synthetases". PLOS Biology. 12 (12): e1002028. doi: 10.1371/journal.pbio.1002028 . ISSN   1545-7885. PMC   4275181 . PMID   25536042.
  40. Palazzo, Alexander F.; Koonin, Eugene V. (2020-11-25). "Functional Long Non-coding RNAs Evolve from Junk Transcripts". Cell. 183 (5): 1151–1161. doi: 10.1016/j.cell.2020.09.047 . PMID   33068526. S2CID   222815635.
  41. Finnigan, Gregory C.; Hanson-Smith, Victor; Stevens, Tom H.; Thornton, Joseph W. (2012-01-09). "Evolution of increased complexity in a molecular machine". Nature. 481 (7381): 360–364. Bibcode:2012Natur.481..360F. doi:10.1038/nature10724. ISSN   0028-0836. PMC   3979732 . PMID   22230956.
  42. Gould, S. J.; Lewontin, R. C.; Maynard Smith, J.; Holliday, Robin (1979-09-21). "The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme". Proceedings of the Royal Society of London. Series B. Biological Sciences. 205 (1161): 581–598. Bibcode:1979RSPSB.205..581G. doi:10.1098/rspb.1979.0086. PMID   42062. S2CID   2129408.
  43. Brunet, T. D. P.; Doolittle, W. Ford (2018-03-19). "The generality of Constructive Neutral Evolution". Biology & Philosophy. 33 (1): 2. doi:10.1007/s10539-018-9614-6. ISSN   1572-8404. S2CID   90290787.
  44. 1 2 3 4 Nei, Masatoshi; Suzuki, Yoshiyuki; Nozawa, Masafumi (2010-09-01). "The Neutral Theory of Molecular Evolution in the Genomic Era". Annual Review of Genomics and Human Genetics. 11 (1): 265–289. doi:10.1146/annurev-genom-082908-150129. ISSN   1527-8204. PMID   20565254.
  45. 1 2 Bakewell, Margaret A.; Shi, Peng; Zhang, Jianzhi (May 2007). "More genes underwent positive selection in chimpanzee evolution than in human evolution". Proceedings of the National Academy of Sciences. 104 (18): 7489–7494. Bibcode:2007PNAS..104.7489B. doi: 10.1073/pnas.0701705104 . ISSN   0027-8424. PMC   1863478 . PMID   17449636.
  46. Chen, Jianhai; Ni, Pan; Li, Xinyun; Han, Jianlin; Jakovlić, Ivan; Zhang, Chengjun; Zhao, Shuhong (2018-01-19). "Population size may shape the accumulation of functional mutations following domestication". BMC Evolutionary Biology. 18 (1): 4. Bibcode:2018BMCEE..18....4C. doi: 10.1186/s12862-018-1120-6 . ISSN   1471-2148. PMC   5775542 . PMID   29351740.
  47. Kern, Andrew D; Hahn, Matthew W (2018-06-01). Kumar, Sudhir (ed.). "The Neutral Theory in Light of Natural Selection". Molecular Biology and Evolution. 35 (6): 1366–1371. doi: 10.1093/molbev/msy092 . ISSN   0737-4038. PMC   5967545 .