Epistasis

Last updated
An example of epistasis is the interaction between hair colour and baldness. A gene for total baldness would be epistatic to one for blond hair or red hair. The hair-colour genes are hypostatic to the baldness gene. The baldness phenotype supersedes genes for hair colour, and so the effects are non-additive. Epistatic hair.png
An example of epistasis is the interaction between hair colour and baldness. A gene for total baldness would be epistatic to one for blond hair or red hair. The hair-colour genes are hypostatic to the baldness gene. The baldness phenotype supersedes genes for hair colour, and so the effects are non-additive.
Example of epistasis in coat colour genetics: If no pigments can be produced the other coat colour genes have no effect on the phenotype, no matter if they are dominant or if the individual is homozygous. Here the genotype "c c" for no pigmentation is epistatic over the other genes. Epistasis.png
Example of epistasis in coat colour genetics: If no pigments can be produced the other coat colour genes have no effect on the phenotype, no matter if they are dominant or if the individual is homozygous. Here the genotype "c c" for no pigmentation is epistatic over the 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. [2] 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. [3]

Contents

The concept of epistasis originated in genetics in 1907 [4] but is now used in biochemistry, computational biology and evolutionary biology. The phenomenon arises due to interactions, either between genes (such as mutations also being needed in regulators of gene expression) or within them (multiple mutations being needed before the gene loses function), leading to non-linear effects. Epistasis has a great influence on the shape of evolutionary landscapes, which leads to profound consequences for evolution and for the evolvability of phenotypic traits.

History

Understanding of epistasis has changed considerably through the history of genetics and so too has the use of the term. The term was first used by William Bateson and his collaborators Florence Durham and Muriel Wheldale Onslow. [4] In early models of natural selection devised in the early 20th century, each gene was considered to make its own characteristic contribution to fitness, against an average background of other genes. Some introductory courses still teach population genetics this way. Because of the way that the science of population genetics was developed, evolutionary geneticists have tended to think of epistasis as the exception. However, in general, the expression of any one allele depends in a complicated way on many other alleles.

In classical genetics, if genes A and B are mutated, and each mutation by itself produces a unique phenotype but the two mutations together show the same phenotype as the gene A mutation, then gene A is epistatic and gene B is hypostatic. For example, the gene for total baldness is epistatic to the gene for brown hair. In this sense, epistasis can be contrasted with genetic dominance, which is an interaction between alleles at the same gene locus. As the study of genetics developed, and with the advent of molecular biology, epistasis started to be studied in relation to quantitative trait loci (QTL) and polygenic inheritance.

The effects of genes are now commonly quantifiable by assaying the magnitude of a phenotype (e.g. height, pigmentation or growth rate) or by biochemically assaying protein activity (e.g. binding or catalysis). Increasingly sophisticated computational and evolutionary biology models aim to describe the effects of epistasis on a genome-wide scale and the consequences of this for evolution. [5] [6] [7] Since identification of epistatic pairs is challenging both computationally and statistically, some studies try to prioritize epistatic pairs. [8] [9]

Classification

Quantitative trait values after two mutations either alone (Ab and aB) or in combination (AB). Bars contained in the grey box indicate the combined trait value under different circumstances of epistasis. Upper panel indicates epistasis between beneficial mutations (blue). Lower panel indicates epistasis between deleterious mutations (red). Epistasis bars.png
Quantitative trait values after two mutations either alone (Ab and aB) or in combination (AB). Bars contained in the grey box indicate the combined trait value under different circumstances of epistasis. Upper panel indicates epistasis between beneficial mutations (blue). Lower panel indicates epistasis between deleterious mutations (red).
Since, on average, mutations are deleterious, random mutations to an organism cause a decline in fitness. If all mutations are additive, fitness will fall proportionally to mutation number (black line). When deleterious mutations display negative (synergistic) epistasis, they are more deleterious in combination than individually and so fitness falls with the number of mutations at an increasing rate (upper, red line). When mutations display positive (antagonistic) epistasis, effects of mutations are less severe in combination than individually and so fitness falls at a decreasing rate (lower, blue line). Synergistic versus antagonistic epistasis.svg
Since, on average, mutations are deleterious, random mutations to an organism cause a decline in fitness. If all mutations are additive, fitness will fall proportionally to mutation number (black line). When deleterious mutations display negative (synergistic) epistasis, they are more deleterious in combination than individually and so fitness falls with the number of mutations at an increasing rate (upper, red line). When mutations display positive (antagonistic) epistasis, effects of mutations are less severe in combination than individually and so fitness falls at a decreasing rate (lower, blue line).

Terminology about epistasis can vary between scientific fields. Geneticists often refer to wild type and mutant alleles where the mutation is implicitly deleterious and may talk in terms of genetic enhancement, synthetic lethality and genetic suppressors. Conversely, a biochemist may more frequently focus on beneficial mutations and so explicitly state the effect of a mutation and use terms such as reciprocal sign epistasis and compensatory mutation. [16] Additionally, there are differences when looking at epistasis within a single gene (biochemistry) and epistasis within a haploid or diploid genome (genetics). In general, epistasis is used to denote the departure from 'independence' of the effects of different genetic loci. Confusion often arises due to the varied interpretation of 'independence' among different branches of biology. [17] The classifications below attempt to cover the various terms and how they relate to one another.

Additivity

Two mutations are considered to be purely additive if the effect of the double mutation is the sum of the effects of the single mutations. This occurs when genes do not interact with each other, for example by acting through different metabolic pathways. Simply, additive traits were studied early on in the history of genetics, however they are relatively rare, with most genes exhibiting at least some level of epistatic interaction. [18] [19]

Magnitude epistasis

When the double mutation has a fitter phenotype than expected from the effects of the two single mutations, it is referred to as positive epistasis. Positive epistasis between beneficial mutations generates greater improvements in function than expected. [10] [11] Positive epistasis between deleterious mutations protects against the negative effects to cause a less severe fitness drop. [13]

Conversely, when two mutations together lead to a less fit phenotype than expected from their effects when alone, it is called negative epistasis. [20] [21] Negative epistasis between beneficial mutations causes smaller than expected fitness improvements, whereas negative epistasis between deleterious mutations causes greater-than-additive fitness drops. [12]

Independently, when the effect on fitness of two mutations is more radical than expected from their effects when alone, it is referred to as synergistic epistasis. The opposite situation, when the fitness difference of the double mutant from the wild type is smaller than expected from the effects of the two single mutations, it is called antagonistic epistasis. [15] Therefore, for deleterious mutations, negative epistasis is also synergistic, while positive epistasis is antagonistic; conversely, for advantageous mutations, positive epistasis is synergistic, while negative epistasis is antagonistic.

The term genetic enhancement is sometimes used when a double (deleterious) mutant has a more severe phenotype than the additive effects of the single mutants. Strong positive epistasis is sometimes referred to by creationists as irreducible complexity (although most examples are misidentified).

Sign epistasis

Sign epistasis [22] occurs when one mutation has the opposite effect when in the presence of another mutation. This occurs when a mutation that is deleterious on its own can enhance the effect of a particular beneficial mutation. [17] For example, a large and complex brain is a waste of energy without a range of sense organs, but sense organs are made more useful by a large and complex brain that can better process the information. If a fitness landscape has no sign epistasis then it is called smooth.

At its most extreme, reciprocal sign epistasis [23] occurs when two deleterious genes are beneficial when together. For example, producing a toxin alone can kill a bacterium, and producing a toxin exporter alone can waste energy, but producing both can improve fitness by killing competing organisms. If a fitness landscape has sign epistasis but no reciprocal sign epistasis then it is called semismooth. [24]

Reciprocal sign epistasis also leads to genetic suppression whereby two deleterious mutations are less harmful together than either one on its own, i.e. one compensates for the other. A clear example of genetic suppression was the demonstration that in the assembly of bacteriophage T4 two deleterious mutations, each causing a deficiency in the level of a different morphogenetic protein, could interact positively. [25] If a mutation causes a reduction in a particular structural component, this can bring about an imbalance in morphogenesis and loss of viable virus progeny, but production of viable progeny can be restored by a second (suppressor) mutation in another morphogenetic component that restores the balance of protein components.

The term genetic suppression can also apply to sign epistasis where the double mutant has a phenotype intermediate between those of the single mutants, in which case the more severe single mutant phenotype is suppressed by the other mutation or genetic condition. For example, in a diploid organism, a hypomorphic (or partial loss-of-function) mutant phenotype can be suppressed by knocking out one copy of a gene that acts oppositely in the same pathway. In this case, the second gene is described as a "dominant suppressor" of the hypomorphic mutant; "dominant" because the effect is seen when one wild-type copy of the suppressor gene is present (i.e. even in a heterozygote). For most genes, the phenotype of the heterozygous suppressor mutation by itself would be wild type (because most genes are not haplo-insufficient), so that the double mutant (suppressed) phenotype is intermediate between those of the single mutants.

In non reciprocal sign epistasis, fitness of the mutant lies in the middle of that of the extreme effects seen in reciprocal sign epistasis.

When two mutations are viable alone but lethal in combination, it is called Synthetic lethality or unlinked non-complementation. [26]

Haploid organisms

In a haploid organism with genotypes (at two loci) ab, Ab, aB or AB, we can think of different forms of epistasis as affecting the magnitude of a phenotype upon mutation individually (Ab and aB) or in combination (AB).

Interaction typeabAbaBAB
No epistasis (additive) 0112AB = Ab + aB + ab 
Positive (synergistic) epistasis0113 AB > Ab + aB + ab 
Negative (antagonistic) epistasis0111AB < Ab + aB + ab 
Sign epistasis01-12AB has opposite sign to AboraB
Reciprocal sign epistasis0-1-12AB has opposite sign to Ab andaB

Diploid organisms

Epistasis in diploid organisms is further complicated by the presence of two copies of each gene. Epistasis can occur between loci, but additionally, interactions can occur between the two copies of each locus in heterozygotes. For a two locus, two allele system, there are eight independent types of gene interaction. [27]

Additive A locusAdditive B locusDominance A locusDominance B locus
aaaAAAaaaAAAaaaAAAaaaAAA
bb10–1bb111bb–11–1bb–1–1–1
bB10–1bB000bB–11–1bB111
BB10–1BB–1–1–1BB–11–1BB–1–1–1
Additive by Additive EpistasisAdditive by Dominance EpistasisDominance by Additive EpistasisDominance by Dominance Epistasis
aaaAAAaaaAAAaaaAAAaaaAAA
bb10–1bb10–1bb1–11bb–11–1
bB000bB–101bB000bB1–11
BB–101BB10–1BB–11–1BB–11–1

Genetic and molecular causes

Additivity

This can be the case when multiple genes act in parallel to achieve the same effect. For example, when an organism is in need of phosphorus, multiple enzymes that break down different phosphorylated components from the environment may act additively to increase the amount of phosphorus available to the organism. However, there inevitably comes a point where phosphorus is no longer the limiting factor for growth and reproduction and so further improvements in phosphorus metabolism have smaller or no effect (negative epistasis). Some sets of mutations within genes have also been specifically found to be additive. [28] It is now considered that strict additivity is the exception, rather than the rule, since most genes interact with hundreds or thousands of other genes. [18] [19]

Epistasis between genes

Epistasis within the genomes of organisms occurs due to interactions between the genes within the genome. This interaction may be direct if the genes encode proteins that, for example, are separate components of a multi-component protein (such as the ribosome), inhibit each other's activity, or if the protein encoded by one gene modifies the other (such as by phosphorylation). Alternatively the interaction may be indirect, where the genes encode components of a metabolic pathway or network, developmental pathway, signalling pathway or transcription factor network. For example, the gene encoding the enzyme that synthesizes penicillin is of no use to a fungus without the enzymes that synthesize the necessary precursors in the metabolic pathway.

Epistasis within genes

Just as mutations in two separate genes can be non-additive if those genes interact, mutations in two codons within a gene can be non-additive. In genetics this is sometimes called intragenic suppression when one deleterious mutation can be compensated for by a second mutation within that gene. Analysis of bacteriophage T4 mutants that were altered in the rIIB cistron (gene) revealed that certain pairwise combinations of mutations could mutually suppress each other; that is the double mutants had a more nearly wild-type phenotype than either mutant alone. [29] The linear map order of the mutants was established using genetic recombination data, From these sources of information, the triplet nature of the genetic code was logically deduced for the first time in 1961, and other key features of the code were also inferred. [29]

Also intragenic suppression can occur when the amino acids within a protein interact. Due to the complexity of protein folding and activity, additive mutations are rare.

Proteins are held in their tertiary structure by a distributed, internal network of cooperative interactions (hydrophobic, polar and covalent). [30] Epistatic interactions occur whenever one mutation alters the local environment of another residue (either by directly contacting it, or by inducing changes in the protein structure). [31] For example, in a disulphide bridge, a single cysteine has no effect on protein stability until a second is present at the correct location at which point the two cysteines form a chemical bond which enhances the stability of the protein. [32] This would be observed as positive epistasis where the double-cysteine variant had a much higher stability than either of the single-cysteine variants. Conversely, when deleterious mutations are introduced, proteins often exhibit mutational robustness whereby as stabilising interactions are destroyed the protein still functions until it reaches some stability threshold at which point further destabilising mutations have large, detrimental effects as the protein can no longer fold. This leads to negative epistasis whereby mutations that have little effect alone have a large, deleterious effect together. [33] [34]

In enzymes, the protein structure orients a few, key amino acids into precise geometries to form an active site to perform chemistry. [35] Since these active site networks frequently require the cooperation of multiple components, mutating any one of these components massively compromises activity, and so mutating a second component has a relatively minor effect on the already inactivated enzyme. For example, removing any member of the catalytic triad of many enzymes will reduce activity to levels low enough that the organism is no longer viable. [36] [37] [38]

Heterozygotic epistasis

Diploid organisms contain two copies of each gene. If these are different (heterozygous / heteroallelic), the two different copies of the allele may interact with each other to cause epistasis. This is sometimes called allelic complementation, or interallelic complementation. It may be caused by several mechanisms, for example transvection, where an enhancer from one allele acts in trans to activate transcription from the promoter of the second allele. Alternately, trans-splicing of two non-functional RNA molecules may produce a single, functional RNA.

Similarly, at the protein level, proteins that function as dimers may form a heterodimer composed of one protein from each alternate gene and may display different properties to the homodimer of one or both variants. Two bacteriophage T4 mutants defective at different locations in the same gene can undergo allelic complementation during a mixed infection. [39] That is, each mutant alone upon infection cannot produce viable progeny, but upon mixed infection with two complementing mutants, viable phage are formed. Intragenic complementation was demonstrated for several genes that encode structural proteins of the bacteriophage [39] indicating that such proteins function as dimers or even higher order multimers. [40]

Evolutionary consequences

Fitness landscapes and evolvability

The top row indicates interactions between two genes that show either (a) additive effects, (b) positive epistasis or (c) reciprocal sign epistasis. Below are fitness landscapes which display greater and greater levels of global epistasis between large numbers of genes. Purely additive interactions lead to a single smooth peak (d); as increasing numbers of genes exhibit epistasis, the landscape becomes more rugged (e), and when all genes interact epistatically the landscape becomes so rugged that mutations have seemingly random effects (f). Epistasis and landscapes.png
The top row indicates interactions between two genes that show either (a) additive effects, (b) positive epistasis or (c) reciprocal sign epistasis. Below are fitness landscapes which display greater and greater levels of global epistasis between large numbers of genes. Purely additive interactions lead to a single smooth peak (d); as increasing numbers of genes exhibit epistasis, the landscape becomes more rugged (e), and when all genes interact epistatically the landscape becomes so rugged that mutations have seemingly random effects (f).

In evolutionary genetics, the sign of epistasis is usually more significant than the magnitude of epistasis. This is because magnitude epistasis (positive and negative) simply affects how beneficial mutations are together, however sign epistasis affects whether mutation combinations are beneficial or deleterious. [10]

A fitness landscape is a representation of the fitness where all genotypes are arranged in 2D space and the fitness of each genotype is represented by height on a surface. It is frequently used as a visual metaphor for understanding evolution as the process of moving uphill from one genotype to the next, nearby, fitter genotype. [18]

If all mutations are additive, they can be acquired in any order and still give a continuous uphill trajectory. The landscape is perfectly smooth, with only one peak (global maximum) and all sequences can evolve uphill to it by the accumulation of beneficial mutations in any order. Conversely, if mutations interact with one another by epistasis, the fitness landscape becomes rugged as the effect of a mutation depends on the genetic background of other mutations. [41] At its most extreme, interactions are so complex that the fitness is 'uncorrelated' with gene sequence and the topology of the landscape is random. This is referred to as a rugged fitness landscape and has profound implications for the evolutionary optimisation of organisms. If mutations are deleterious in one combination but beneficial in another, the fittest genotypes can only be accessed by accumulating mutations in one specific order. This makes it more likely that organisms will get stuck at local maxima in the fitness landscape having acquired mutations in the 'wrong' order. [34] [42] For example, a variant of TEM1 β-lactamase with 5 mutations is able to cleave cefotaxime (a third generation antibiotic). [43] However, of the 120 possible pathways to this 5-mutant variant, only 7% are accessible to evolution as the remainder passed through fitness valleys where the combination of mutations reduces activity. In contrast, changes in environment (and therefore the shape of the fitness landscape) have been shown to provide escape from local maxima. [34] In this example, selection in changing antibiotic environments resulted in a "gateway mutation" which epistatically interacted in a positive manner with other mutations along an evolutionary pathway, effectively crossing a fitness valley. This gateway mutation alleviated the negative epistatic interactions of other individually beneficial mutations, allowing them to better function in concert. Complex environments or selections may therefore bypass local maxima found in models assuming simple positive selection.

High epistasis is usually considered a constraining factor on evolution, and improvements in a highly epistatic trait are considered to have lower evolvability. This is because, in any given genetic background, very few mutations will be beneficial, even though many mutations may need to occur to eventually improve the trait. The lack of a smooth landscape makes it harder for evolution to access fitness peaks. In highly rugged landscapes, fitness valleys block access to some genes, and even if ridges exist that allow access, these may be rare or prohibitively long. [44] Moreover, adaptation can move proteins into more precarious or rugged regions of the fitness landscape. [45] These shifting "fitness territories" may act to decelerate evolution and could represent tradeoffs for adaptive traits.

The frustration of adaptive evolution by rugged fitness landscapes was recognized as a potential force for the evolution of evolvability. Michael Conrad in 1972 was the first to propose a mechanism for the evolution of evolvability by noting that a mutation which smoothed the fitness landscape at other loci could facilitate the production of advantageous mutations and hitchhike along with them. [46] [47] Rupert Riedl in 1975 proposed that new genes which produced the same phenotypic effects with a single mutation as other loci with reciprocal sign epistasis would be a new means to attain a phenotype otherwise too unlikely to occur by mutation. [48] [49]

Rugged, epistatic fitness landscapes also affect the trajectories of evolution. When a mutation has a large number of epistatic effects, each accumulated mutation drastically changes the set of available beneficial mutations. Therefore, the evolutionary trajectory followed depends highly on which early mutations were accepted. Thus, repeats of evolution from the same starting point tend to diverge to different local maxima rather than converge on a single global maximum as they would in a smooth, additive landscape. [50] [51]

Evolution of sex

Negative epistasis and sex are thought to be intimately correlated. Experimentally, this idea has been tested in using digital simulations of asexual and sexual populations. Over time, sexual populations move towards more negative epistasis, or the lowering of fitness by two interacting alleles. It is thought that negative epistasis allows individuals carrying the interacting deleterious mutations to be removed from the populations efficiently. This removes those alleles from the population, resulting in an overall more fit population. This hypothesis was proposed by Alexey Kondrashov, and is sometimes known as the deterministic mutation hypothesis [52] and has also been tested using artificial gene networks. [20]

However, the evidence for this hypothesis has not always been straightforward and the model proposed by Kondrashov has been criticized for assuming mutation parameters far from real world observations. [53] In addition, in those tests which used artificial gene networks, negative epistasis is only found in more densely connected networks, [20] whereas empirical evidence indicates that natural gene networks are sparsely connected, [54] and theory shows that selection for robustness will favor more sparsely connected and minimally complex networks. [54]

Methods and model systems

Regression analysis

Quantitative genetics focuses on genetic variance due to genetic interactions. Any two locus interactions at a particular gene frequency can be decomposed into eight independent genetic effects using a weighted regression. In this regression, the observed two locus genetic effects are treated as dependent variables and the "pure" genetic effects are used as the independent variables. Because the regression is weighted, the partitioning among the variance components will change as a function of gene frequency. By analogy it is possible to expand this system to three or more loci, or to cytonuclear interactions [55]

Double mutant cycles

When assaying epistasis within a gene, site-directed mutagenesis can be used to generate the different genes, and their protein products can be assayed (e.g. for stability or catalytic activity). This is sometimes called a double mutant cycle and involves producing and assaying the wild type protein, the two single mutants and the double mutant. Epistasis is measured as the difference between the effects of the mutations together versus the sum of their individual effects. [56] This can be expressed as a free energy of interaction. The same methodology can be used to investigate the interactions between larger sets of mutations but all combinations have to be produced and assayed. For example, there are 120 different combinations of 5 mutations, some or all of which may show epistasis...

Computational prediction

Numerous computational methods have been developed for the detection and characterization of epistasis. Many of these rely on machine learning to detect non-additive effects that might be missed by statistical approaches such as linear regression. [57] For example, multifactor dimensionality reduction (MDR) was designed specifically for nonparametric and model-free detection of combinations of genetic variants that are predictive of a phenotype such as disease status in human populations. [58] [59] Several of these approaches have been broadly reviewed in the literature. [60] Even more recently, methods that utilize insights from theoretical computer science (the Hadamard transform [61] and compressed sensing [62] [63] ) or maximum-likelihood inference [64] were shown to distinguish epistatic effects from overall non-linearity in genotype–phenotype map structure, [65] while others used patient survival analysis to identify non-linearity. [66]

See also

Related Research Articles

<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.

<span class="mw-page-title-main">Dominance (genetics)</span> One gene variant masking the effect of another in the other copy of the gene

In genetics, dominance is the phenomenon of one variant (allele) of a gene on a chromosome masking or overriding the effect of a different variant of the same gene on the other copy of the chromosome. The first variant is termed dominant and the second is called recessive. This state of having two different variants of the same gene on each chromosome is originally caused by a mutation in one of the genes, either new or inherited. The terms autosomal dominant or autosomal recessive are used to describe gene variants on non-sex chromosomes (autosomes) and their associated traits, while those on sex chromosomes (allosomes) are termed X-linked dominant, X-linked recessive or Y-linked; these have an inheritance and presentation pattern that depends on the sex of both the parent and the child. Since there is only one copy of the Y chromosome, Y-linked traits cannot be dominant or recessive. Additionally, there are other forms of dominance, such as incomplete dominance, in which a gene variant has a partial effect compared to when it is present on both chromosomes, and co-dominance, in which different variants on each chromosome both show their associated traits.

<span class="mw-page-title-main">Neutral theory of molecular evolution</span> Theory of evolution by changes at the molecular level

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.

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.

In evolutionary biology, fitness landscapes or adaptive landscapes are used to visualize the relationship between genotypes and reproductive success. It is assumed that every genotype has a well-defined replication rate. This fitness is the "height" of the landscape. Genotypes which are similar are said to be "close" to each other, while those that are very different are "far" from each other. The set of all possible genotypes, their degree of similarity, and their related fitness values is then called a fitness landscape. The idea of a fitness landscape is a metaphor to help explain flawed forms in evolution by natural selection, including exploits and glitches in animals like their reactions to supernormal stimuli.

<span class="mw-page-title-main">Polymorphism (biology)</span> Occurrence of two or more clearly different morphs or forms in the population of a species

In biology, polymorphism is the occurrence of two or more clearly different morphs or forms, also referred to as alternative phenotypes, in the population of a species. To be classified as such, morphs must occupy the same habitat at the same time and belong to a panmictic population.

<span class="mw-page-title-main">Evolution of sexual reproduction</span> How sexually reproducing multicellular organisms could have evolved from a common ancestor species

Sexual reproduction is an adaptive feature which is common to almost all multicellular organisms and various unicellular organisms. Currently, the adaptive advantage of sexual reproduction is widely regarded as a major unsolved problem in biology. As discussed below, one prominent theory is that sex evolved as an efficient mechanism for producing variation, and this had the advantage of enabling organisms to adapt to changing environments. Another prominent theory, also discussed below, is that a primary advantage of outcrossing sex is the masking of the expression of deleterious mutations. Additional theories concerning the adaptive advantage of sex are also discussed below. Sex does, however, come with a cost. In reproducing asexually, no time nor energy needs to be expended in choosing a mate and, if the environment has not changed, then there may be little reason for variation, as the organism may already be well-adapted. However, very few environments have not changed over the millions of years that reproduction has existed. Hence it is easy to imagine that being able to adapt to changing environment imparts a benefit. Sex also halves the amount of offspring a given population is able to produce. Sex, however, has evolved as the most prolific means of species branching into the tree of life. Diversification into the phylogenetic tree happens much more rapidly via sexual reproduction than it does by way of asexual reproduction.

Evolvability is defined as the capacity of a system for adaptive evolution. Evolvability is the ability of a population of organisms to not merely generate genetic diversity, but to generate adaptive genetic diversity, and thereby evolve through natural selection.

Genetic architecture is the underlying genetic basis of a phenotypic trait and its variational properties. Phenotypic variation for quantitative traits is, at the most basic level, the result of the segregation of alleles at quantitative trait loci (QTL). Environmental factors and other external influences can also play a role in phenotypic variation. Genetic architecture is a broad term that can be described for any given individual based on information regarding gene and allele number, the distribution of allelic and mutational effects, and patterns of pleiotropy, dominance, and epistasis.

<span class="mw-page-title-main">Pleiotropy</span> Influence of a single gene on multiple phenotypic traits

Pleiotropy occurs when one gene influences two or more seemingly unrelated phenotypic traits. Such a gene that exhibits multiple phenotypic expression is called a pleiotropic gene. Mutation in a pleiotropic gene may have an effect on several traits simultaneously, due to the gene coding for a product used by a myriad of cells or different targets that have the same signaling function.

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.

Complementation refers to a genetic process when two strains of an organism with different homozygous recessive mutations that produce the same mutant phenotype have offspring that express the wild-type phenotype when mated or crossed. Complementation will ordinarily occur if the mutations are in different genes. Complementation may also occur if the two mutations are at different sites within the same gene, but this effect is usually weaker than that of intergenic complementation. When the mutations are in different genes, each strain's genome supplies the wild-type allele to "complement" the mutated allele of the other strain's genome. Since the mutations are recessive, the offspring will display the wild-type phenotype. A complementation test can test whether the mutations in two strains are in different genes. Complementation is usually weaker or absent if the mutations are in the same gene. The convenience and essence of this test is that the mutations that produce a phenotype can be assigned to different genes without the exact knowledge of what the gene product is doing on a molecular level. American geneticist Edward B. Lewis developed the complementation test.

Evolutionary capacitance is the storage and release of variation, just as electric capacitors store and release charge. Living systems are robust to mutations. This means that living systems accumulate genetic variation without the variation having a phenotypic effect. But when the system is disturbed, robustness breaks down, and the variation has phenotypic effects and is subject to the full force of natural selection. An evolutionary capacitor is a molecular switch mechanism that can "toggle" genetic variation between hidden and revealed states. If some subset of newly revealed variation is adaptive, it becomes fixed by genetic assimilation. After that, the rest of variation, most of which is presumably deleterious, can be switched off, leaving the population with a newly evolved advantageous trait, but no long-term handicap. For evolutionary capacitance to increase evolvability in this way, the switching rate should not be faster than the timescale of genetic assimilation.

Inbreeding depression is the reduced biological fitness that has the potential to result from inbreeding. The loss of genetic diversity that is seen due to inbreeding, results from small population size. Biological fitness refers to an organism's ability to survive and perpetuate its genetic material. Inbreeding depression is often the result of a population bottleneck. In general, the higher the genetic variation or gene pool within a breeding population, the less likely it is to suffer from inbreeding depression, though inbreeding and outbreeding depression can simultaneously occur.

<span class="mw-page-title-main">Bateson–Dobzhansky–Muller model</span> Model of the evolution of genetic incompatibility

The Bateson–Dobzhansky–Muller model, also known as Dobzhansky–Muller model, is a model of the evolution of genetic incompatibility, important in understanding the evolution of reproductive isolation during speciation and the role of natural selection in bringing it about. The theory was first described by William Bateson in 1909, then independently described by Theodosius Dobzhansky in 1934, and later elaborated in different forms by Herman Muller, H. Allen Orr and Sergey Gavrilets.

Epistasis refers to genetic interactions in which the mutation of one gene masks the phenotypic effects of a mutation at another locus. Systematic analysis of these epistatic interactions can provide insight into the structure and function of genetic pathways. Examining the phenotypes resulting from pairs of mutations helps in understanding how the function of these genes intersects. Genetic interactions are generally classified as either Positive/Alleviating or Negative/Aggravating. Fitness epistasis is positive when a loss of function mutation of two given genes results in exceeding the fitness predicted from individual effects of deleterious mutations, and it is negative when it decreases fitness. Ryszard Korona and Lukas Jasnos showed that the epistatic effect is usually positive in Saccharomyces cerevisiae. Usually, even in case of positive interactions double mutant has smaller fitness than single mutants. The positive interactions occur often when both genes lie within the same pathway Conversely, negative interactions are characterized by an even stronger defect than would be expected in the case of two single mutations, and in the most extreme cases the double mutation is lethal. This aggravated phenotype arises when genes in compensatory pathways are both knocked out.

<span class="mw-page-title-main">Robustness (evolution)</span> Persistence of a biological trait under uncertain conditions

In evolutionary biology, robustness of a biological system is the persistence of a certain characteristic or trait in a system under perturbations or conditions of uncertainty. Robustness in development is known as canalization. According to the kind of perturbation involved, robustness can be classified as mutational, environmental, recombinational, or behavioral robustness etc. Robustness is achieved through the combination of many genetic and molecular mechanisms and can evolve by either direct or indirect selection. Several model systems have been developed to experimentally study robustness and its evolutionary consequences.

Weak selection in evolutionary biology is when individuals with different phenotypes possess similar fitness, i.e. one phenotype is weakly preferred over the other. Weak selection, therefore, is an evolutionary theory to explain the maintenance of multiple phenotypes in a stable population.

<span class="mw-page-title-main">Genetic variance</span> Biological concept

Genetic variance is a concept outlined by the English biologist and statistician Ronald Fisher in his fundamental theorem of natural selection. In his 1930 book The Genetical Theory of Natural Selection, Fisher postulates that the rate of change of biological fitness can be calculated by the genetic variance of the fitness itself. Fisher tried to give a statistical formula about how the change of fitness in a population can be attributed to changes in the allele frequency. Fisher made no restrictive assumptions in his formula concerning fitness parameters, mate choices or the number of alleles and loci involved.

Genetic interaction networks represent the functional interactions between pairs of genes in an organism and are useful for understanding the relation between genotype and phenotype. The majority of genes do not code for particular phenotypes. Instead, phenotypes often result from the interaction between several genes. In humans, "Each individual carries ~4 million genetic variants and polymorphisms, the overwhelming majority of which cannot be pinpointed as the single cause for a given phenotype. Instead, the effects of genetic variants may combine with one another both additively and synergistically, and each variant's contribution to a quantitative trait or disease risk could depend on the genotypes of dozens of other variants. Interactions between genetic variants, along with the environmental conditions, are likely to play a major role in determining the phenotype that arises from a given genotype." Genetic interaction networks help to understand genetic interactions by identifying such interactions between pairs of genes.

References

  1. Neil A. Campbell, Jane B. Reece: Biologie. Spektrum-Verlag Heidelberg-Berlin 2003, ISBN   3-8274-1352-4, page 306
  2. Gros PA, Le Nagard H, Tenaillon O (May 2009). "The Evolution of Epistasis and its Links with Genetic Robustness, Complexity and Drift in a Phenotypic Model of Adaptation". Genetics. 182 (1): 277–93. doi:10.1534/genetics.108.099127. PMC   2674823 . PMID   19279327.
  3. Rieger R, Michaelis A, Green MM (1968), A Glossary of Genetics and Cytogenetics: Classical and Molecular, New York: Springer-Verlag, ISBN   978-0-387-07668-3
  4. 1 2 Richmond, Marsha L. (2001). "Women in the Early History of Genetics: William Bateson and the Newnham College Mendelians, 1900–1910". Isis. 92 (1). The History of Science Society: 55–90. doi:10.1086/385040. JSTOR   237327. PMID   11441497. S2CID   29790111.
  5. Szendro IG, Martijn F S, Franke J, Krug J, de Visser J, Arjan GM (16 January 2013). "Quantitative analyses of empirical fitness landscapes". Journal of Statistical Mechanics: Theory and Experiment. 2013 (1): P01005. arXiv: 1202.4378 . Bibcode:2013JSMTE..01..005S. doi:10.1088/1742-5468/2013/01/P01005. S2CID   14356766.
  6. Edlund JA, Adami C (Spring 2004). "Evolution of robustness in digital organisms". Artificial Life. 10 (2): 167–79. CiteSeerX   10.1.1.556.2318 . doi:10.1162/106454604773563595. PMID   15107229. S2CID   13371337.
  7. Chattopadhyay S (Spring 2019). "Genome-wide interaction and pathway-based identification of key regulators in multiple myeloma". Communications Biology. 4 (2): 89–96. doi:10.1038/s42003-019-0329-2. PMC   6399257 . PMID   30854481.
  8. Ayati, Marzieh; Koyutürk, Mehmet (2014). "Prioritization of genomic locus pairs for testing epistasis". Proceedings of the 5th ACM Conference on Bioinformatics, Computational Biology, and Health Informatics. pp. 240–248. CiteSeerX   10.1.1.715.1549 . doi:10.1145/2649387.2649449. ISBN   978-1-4503-2894-4. S2CID   17343957.
  9. Piriyapongsa J, Ngamphiw C, Intarapanich A, Kulawonganunchai S, Assawamakin A, Bootchai C, Shaw PJ, Tongsima S (2012-12-13). "iLOCi: a SNP interaction prioritization technique for detecting epistasis in genome-wide association studies". BMC Genomics. 13 (Suppl 7): S2. doi: 10.1186/1471-2164-13-S7-S2 . PMC   3521387 . PMID   23281813.
  10. 1 2 3 Phillips PC (November 2008). "Epistasis--the essential role of gene interactions in the structure and evolution of genetic systems". Nature Reviews. Genetics. 9 (11): 855–67. doi:10.1038/nrg2452. PMC   2689140 . PMID   18852697.
  11. 1 2 Domingo E, Sheldon J, Perales C (June 2012). "Viral quasispecies evolution". Microbiology and Molecular Biology Reviews. 76 (2): 159–216. doi:10.1128/mmbr.05023-11. PMC   3372249 . PMID   22688811.
  12. 1 2 3 Tokuriki N, Tawfik DS (October 2009). "Stability effects of mutations and protein evolvability". Current Opinion in Structural Biology. 19 (5): 596–604. doi:10.1016/j.sbi.2009.08.003. PMID   19765975.
  13. 1 2 3 He X, Qian W, Wang Z, Li Y, Zhang J (March 2010). "Prevalent positive epistasis in Escherichia coli and Saccharomyces cerevisiae metabolic networks". Nature Genetics. 42 (3): 272–6. doi:10.1038/ng.524. PMC   2837480 . PMID   20101242.
  14. Ridley M (2004). Evolution (3rd ed.). Blackwell Publishing.
  15. 1 2 Charlesworth B, Charlesworth D (2010). Elements of Evolutionary Genetics. Roberts and Company Publishers.
  16. Ortlund EA, Bridgham JT, Redinbo MR, Thornton JW (September 2007). "Crystal structure of an ancient protein: evolution by conformational epistasis". Science. 317 (5844): 1544–8. Bibcode:2007Sci...317.1544O. doi:10.1126/science.1142819. PMC   2519897 . PMID   17702911.
  17. 1 2 Cordell HJ (October 2002). "Epistasis: what it means, what it doesn't mean, and statistical methods to detect it in humans". Human Molecular Genetics. 11 (20): 2463–8. CiteSeerX   10.1.1.719.4634 . doi:10.1093/hmg/11.20.2463. PMID   12351582.
  18. 1 2 3 Kauffman, Stuart A. (1993). The Origins of Order: Self-organization and Selection in Evolution. Oxford University Press. ISBN   978-0-19-507951-7.[ page needed ]
  19. 1 2 Bornscheuer UT, Huisman GW, Kazlauskas RJ, Lutz S, Moore JC, Robins K (May 2012). "Engineering the third wave of biocatalysis". Nature. 485 (7397): 185–94. Bibcode:2012Natur.485..185B. doi:10.1038/nature11117. PMID   22575958. S2CID   4379415.
  20. 1 2 3 Azevedo RB, Lohaus R, Srinivasan S, Dang KK, Burch CL (March 2006). "Sexual reproduction selects for robustness and negative epistasis in artificial gene networks". Nature. 440 (7080): 87–90. Bibcode:2006Natur.440...87A. doi:10.1038/nature04488. PMID   16511495. S2CID   4415072.
  21. Bonhoeffer S, Chappey C, Parkin NT, Whitcomb JM, Petropoulos CJ (November 2004). "Evidence for positive epistasis in HIV-1". Science. 306 (5701): 1547–50. Bibcode:2004Sci...306.1547B. doi:10.1126/science.1101786. PMID   15567861. S2CID   45784964.
  22. Weinreich DM, Watson RA, Chao L (June 2005). "Perspective: Sign epistasis and genetic constraint on evolutionary trajectories". Evolution; International Journal of Organic Evolution. 59 (6): 1165–74. doi: 10.1111/j.0014-3820.2005.tb01768.x . JSTOR   3448895. PMID   16050094.
  23. Poelwijk FJ, Kiviet DJ, Weinreich DM, Tans SJ (January 2007). "Empirical fitness landscapes reveal accessible evolutionary paths". Nature. 445 (7126): 383–6. Bibcode:2007Natur.445..383P. doi:10.1038/nature05451. PMID   17251971. S2CID   4415468.
  24. Kaznatcheev, Artem (1 May 2019). "Computational Complexity as an Ultimate Constraint on Evolution". Genetics. 212 (1): 245–265. doi:10.1534/genetics.119.302000. PMC   6499524 . PMID   30833289.
  25. Floor, E. (1970). "Interaction of morphogenetic genes of bacteriophage T4". Journal of Molecular Biology. 47 (3): 293–306. doi:10.1016/0022-2836(70)90303-7. PMID   4907266.
  26. French-Mischo S (July 2002). "Synthetic Lethal Mutations". Department of Microbiology, University of Illinois, Urbana. Archived from the original on 2016-08-23. Retrieved 2017-08-03.
  27. Kempthorne O (1969). An introduction to genetic statistics. Iowa State University Press. ISBN   978-0-8138-2375-1.
  28. Lunzer M, Miller SP, Felsheim R, Dean AM (October 2005). "The biochemical architecture of an ancient adaptive landscape". Science. 310 (5747): 499–501. Bibcode:2005Sci...310..499L. doi:10.1126/science.1115649. PMID   16239478. S2CID   28379541.
  29. 1 2 Crick, F. H.; Barnett, L-; Brenner, S.; Watts-Tobin, R. J. (1961). "General nature of the genetic code for proteins". Nature. 192 (4809): 1227–1232. Bibcode:1961Natur.192.1227C. doi:10.1038/1921227a0. PMID   13882203. S2CID   4276146.
  30. Shakhnovich BE, Deeds E, Delisi C, Shakhnovich E (March 2005). "Protein structure and evolutionary history determine sequence space topology". Genome Research. 15 (3): 385–92. arXiv: q-bio/0404040 . doi:10.1101/gr.3133605. PMC   551565 . PMID   15741509.
  31. Harms MJ, Thornton JW (August 2013). "Evolutionary biochemistry: revealing the historical and physical causes of protein properties". Nature Reviews. Genetics. 14 (8): 559–71. doi:10.1038/nrg3540. PMC   4418793 . PMID   23864121.
  32. Witt D (2008). "Recent developments in disulfide bond formation". Synthesis . 2008 (16): 2491–2509. doi:10.1055/s-2008-1067188.
  33. Bershtein S, Segal M, Bekerman R, Tokuriki N, Tawfik DS (December 2006). "Robustness-epistasis link shapes the fitness landscape of a randomly drifting protein". Nature. 444 (7121): 929–32. Bibcode:2006Natur.444..929B. doi:10.1038/nature05385. PMID   17122770. S2CID   4416275.
  34. 1 2 3 Steinberg B, Ostermeier M (January 2016). "Environmental changes bridge evolutionary valleys". Science Advances. 2 (1): e1500921. Bibcode:2016SciA....2E0921S. doi:10.1126/sciadv.1500921. PMC   4737206 . PMID   26844293.
  35. Halabi N, Rivoire O, Leibler S, Ranganathan R (August 2009). "Protein sectors: evolutionary units of three-dimensional structure". Cell. 138 (4): 774–86. doi:10.1016/j.cell.2009.07.038. PMC   3210731 . PMID   19703402.
  36. Neet KE, Koshland DE (November 1966). "The conversion of serine at the active site of subtilisin to cysteine: a "chemical mutation"". Proceedings of the National Academy of Sciences of the United States of America. 56 (5): 1606–11. Bibcode:1966PNAS...56.1606N. doi: 10.1073/pnas.56.5.1606 . PMC   220044 . PMID   5230319.
  37. Beveridge AJ (July 1996). "A theoretical study of the active sites of papain and S195C rat trypsin: implications for the low reactivity of mutant serine proteinases". Protein Science. 5 (7): 1355–65. doi:10.1002/pro.5560050714. PMC   2143470 . PMID   8819168.
  38. Sigal IS, Harwood BG, Arentzen R (December 1982). "Thiol-beta-lactamase: replacement of the active-site serine of RTEM beta-lactamase by a cysteine residue". Proceedings of the National Academy of Sciences of the United States of America. 79 (23): 7157–60. Bibcode:1982PNAS...79.7157S. doi: 10.1073/pnas.79.23.7157 . PMC   347297 . PMID   6818541.
  39. 1 2 Bernstein, H.; Edgar, R. S.; Denhardt, G. H. (1965). "Intragenic Complementation Among Temperature Sensitive Mutants of Bacteriophage T4D". Genetics. 51 (6): 987–1002. doi:10.1093/genetics/51.6.987. PMC   1210828 . PMID   14337770.
  40. CRICK FH; ORGEL LE (1964). "The Theory of Inter-Allelic Complementation". Journal of Molecular Biology. 8: 161–165. doi:10.1016/s0022-2836(64)80156-x. PMID   14149958.
  41. Poelwijk FJ, Tănase-Nicola S, Kiviet DJ, Tans SJ (March 2011). "Reciprocal sign epistasis is a necessary condition for multi-peaked fitness landscapes" (PDF). Journal of Theoretical Biology. 272 (1): 141–4. Bibcode:2011JThBi.272..141P. doi:10.1016/j.jtbi.2010.12.015. PMID   21167837.
  42. Reetz MT, Sanchis J (September 2008). "Constructing and analyzing the fitness landscape of an experimental evolutionary process". ChemBioChem. 9 (14): 2260–7. doi:10.1002/cbic.200800371. PMID   18712749. S2CID   23771783.
  43. Weinreich DM, Delaney NF, Depristo MA, Hartl DL (April 2006). "Darwinian evolution can follow only very few mutational paths to fitter proteins". Science. 312 (5770): 111–114. Bibcode:2006Sci...312..111W. doi:10.1126/science.1123539. PMID   16601193. S2CID   21186834.
  44. Gong LI, Suchard MA, Bloom JD (May 2013). "Stability-mediated epistasis constrains the evolution of an influenza protein". eLife. 2: e00631. doi: 10.7554/eLife.00631 . PMC   3654441 . PMID   23682315.
  45. Steinberg B, Ostermeier M (July 2016). "Shifting Fitness and Epistatic Landscapes Reflect Trade-offs along an Evolutionary Pathway". Journal of Molecular Biology. 428 (13): 2730–43. doi: 10.1016/j.jmb.2016.04.033 . PMID   27173379.
  46. Conrad M (1972). "The importance of molecular hierarchy in information processing". Towards a Theoretical Biology (4). Edinburgh University Press Edinburgh: 222.
  47. Conrad, Michael (1979). "Bootstrapping on the adaptive landscape". BioSystems. 11 (2–3): 167–182. Bibcode:1979BiSys..11..167C. doi:10.1016/0303-2647(79)90009-1. hdl: 2027.42/23514 . PMID   497367.
  48. Riedl, Rupert J. (1975). Die Ordnung des Lebendigen: Systembedingungen der Evolution. Hamburg and Berlin: Parey.
  49. Riedl, Rupert J. (1977). "A systems-analytical approach to macroevolutionary phenomena". Quarterly Review of Biology. 52 (4): 351–370. doi:10.1086/410123. PMID   343152. S2CID   25465466.
  50. Lobkovsky AE, Wolf YI, Koonin EV (December 2011). "Predictability of evolutionary trajectories in fitness landscapes". PLOS Computational Biology. 7 (12): e1002302. arXiv: 1108.3590 . Bibcode:2011PLSCB...7E2302L. doi: 10.1371/journal.pcbi.1002302 . PMC   3240586 . PMID   22194675.
  51. Bridgham JT, Ortlund EA, Thornton JW (September 2009). "An epistatic ratchet constrains the direction of glucocorticoid receptor evolution". Nature. 461 (7263): 515–9. Bibcode:2009Natur.461..515B. doi:10.1038/nature08249. PMC   6141187 . PMID   19779450.
  52. Kondrashov AS (December 1988). "Deleterious mutations and the evolution of sexual reproduction". Nature. 336 (6198): 435–40. Bibcode:1988Natur.336..435K. doi:10.1038/336435a0. PMID   3057385. S2CID   4233528.
  53. MacCarthy T, Bergman A (July 2007). "Coevolution of robustness, epistasis, and recombination favors asexual reproduction". Proceedings of the National Academy of Sciences of the United States of America. 104 (31): 12801–6. Bibcode:2007PNAS..10412801M. doi: 10.1073/pnas.0705455104 . PMC   1931480 . PMID   17646644.
  54. 1 2 Leclerc RD (August 2008). "Survival of the sparsest: robust gene networks are parsimonious". Molecular Systems Biology. 4 (213): 213. doi:10.1038/msb.2008.52. PMC   2538912 . PMID   18682703.
  55. Wade MJ, Goodnight CJ (April 2006). "Cyto-nuclear epistasis: two-locus random genetic drift in hermaphroditic and dioecious species". Evolution; International Journal of Organic Evolution. 60 (4): 643–59. doi:10.1554/05-019.1. PMID   16739448. S2CID   41900960.
  56. Horovitz A (1996). "Double-mutant cycles: a powerful tool for analyzing protein structure and function". Folding & Design. 1 (6): R121–6. doi:10.1016/s1359-0278(96)00056-9. PMID   9080186.
  57. Chicco D, Faultless T (2021-03-18). "Brief Survey on Machine Learning in Epistasis". Epistasis. Methods in Molecular Biology. Vol. 2212. pp. 169–179. doi:10.1007/978-1-0716-0947-7_11. ISBN   978-1-0716-0947-7. PMID   33733356. S2CID   232303194.
  58. Moore JH, Andrews PC (2015-01-01). "Epistasis Analysis Using Multifactor Dimensionality Reduction". Epistasis. Methods in Molecular Biology. Vol. 1253. pp. 301–14. doi:10.1007/978-1-4939-2155-3_16. ISBN   978-1-4939-2154-6. PMID   25403539.
  59. Moore JH, Williams SM, eds. (2015). Epistasis: Methods and Protocols. Methods in Molecular Biology. Vol. 1253. Springer. doi:10.1007/978-1-4939-2155-3. ISBN   978-1-4939-2154-6. S2CID   241378477.
  60. Cordell HJ (June 2009). "Detecting gene-gene interactions that underlie human diseases". Nature Reviews. Genetics. 10 (6): 392–404. doi:10.1038/nrg2579. PMC   2872761 . PMID   19434077.
  61. Weinreich DM, Lan Y, Wylie CS, Heckendorn RB (December 2013). "Should evolutionary geneticists worry about higher-order epistasis?". Current Opinion in Genetics & Development. Genetics of system biology. 23 (6): 700–7. doi:10.1016/j.gde.2013.10.007. PMC   4313208 . PMID   24290990.
  62. Poelwijk FJ, Krishna V, Ranganathan R (June 2016). "The Context-Dependence of Mutations: A Linkage of Formalisms". PLOS Computational Biology. 12 (6): e1004771. arXiv: 1502.00726 . Bibcode:2016PLSCB..12E4771P. doi: 10.1371/journal.pcbi.1004771 . PMC   4919011 . PMID   27337695.
  63. Poelwijk FJ, Socolich, M, Ranganathan R (September 2019). "Learning the pattern of epistasis linking genotype and phenotype in a protein". Nature Communications. 10 (1): 4213. Bibcode:2019NatCo..10.4213P. doi:10.1038/s41467-019-12130-8. PMC   6746860 . PMID   31527666.
  64. Otwinowski J, McCandlish DM, Plotkin JB (August 2018). "Inferring the shape of global epistasis". Proceedings of the National Academy of Sciences of the United States of America. 115 (32): E7550–E7558. Bibcode:2018PNAS..115E7550O. doi: 10.1073/pnas.1804015115 . PMC   6094095 . PMID   30037990.
  65. Sailer ZR, Harms MJ (March 2017). "Detecting High-Order Epistasis in Nonlinear Genotype-Phenotype Maps". Genetics. 205 (3): 1079–1088. doi:10.1534/genetics.116.195214. PMC   5340324 . PMID   28100592.
  66. Magen, A (2019). "Beyond Synthetic Lethality: Charting the Landscape of Pairwise Gene Expression States Associated with Survival in Cancer". Cell Reports. 28 (4): P938–948.E6. doi: 10.1016/j.celrep.2019.06.067 . PMC   8261641 . PMID   31340155.