fbpx
Wikipedia

Hardy–Weinberg principle

In population genetics, the Hardy–Weinberg principle, also known as the Hardy–Weinberg equilibrium, model, theorem, or law, states that allele and genotype frequencies in a population will remain constant from generation to generation in the absence of other evolutionary influences. These influences include genetic drift, mate choice, assortative mating, natural selection, sexual selection, mutation, gene flow, meiotic drive, genetic hitchhiking, population bottleneck, founder effect, inbreeding and outbreeding depression.

Hardy–Weinberg proportions for two alleles: the horizontal axis shows the two allele frequencies p and q and the vertical axis shows the expected genotype frequencies. Each line shows one of the three possible genotypes.

In the simplest case of a single locus with two alleles denoted A and a with frequencies f(A) = p and f(a) = q, respectively, the expected genotype frequencies under random mating are f(AA) = p2 for the AA homozygotes, f(aa) = q2 for the aa homozygotes, and f(Aa) = 2pq for the heterozygotes. In the absence of selection, mutation, genetic drift, or other forces, allele frequencies p and q are constant between generations, so equilibrium is reached.

The principle is named after G. H. Hardy and Wilhelm Weinberg, who first demonstrated it mathematically. Hardy's paper was focused on debunking the view that a dominant allele would automatically tend to increase in frequency (a view possibly based on a misinterpreted question at a lecture[1]). Today, tests for Hardy–Weinberg genotype frequencies are used primarily to test for population stratification and other forms of non-random mating.

Derivation edit

Consider a population of monoecious diploids, where each organism produces male and female gametes at equal frequency, and has two alleles at each gene locus. We assume that the population is so large that it can be treated as infinite. Organisms reproduce by random union of gametes (the "gene pool" population model). A locus in this population has two alleles, A and a, that occur with initial frequencies f0(A) = p and f0(a) = q, respectively.[note 1] The allele frequencies at each generation are obtained by pooling together the alleles from each genotype of the same generation according to the expected contribution from the homozygote and heterozygote genotypes, which are 1 and 1/2, respectively:

 

(1)
 

(2)
 
Length of p, q corresponds to allele frequencies (here p = 0.6, q = 0.4). Then area of rectangle represents genotype frequencies (thus AA : Aa : aa = 0.36 : 0.48 : 0.16).

The different ways to form genotypes for the next generation can be shown in a Punnett square, where the proportion of each genotype is equal to the product of the row and column allele frequencies from the current generation.

Table 1: Punnett square for Hardy–Weinberg
Females
A (p) a (q)
Males A (p) AA (p2) Aa (pq)
a (q) Aa (qp) aa (q2)

The sum of the entries is p2 + 2pq + q2 = 1, as the genotype frequencies must sum to one.

Note again that as p + q = 1, the binomial expansion of (p + q)2 = p2 + 2pq + q2 = 1 gives the same relationships.

Summing the elements of the Punnett square or the binomial expansion, we obtain the expected genotype proportions among the offspring after a single generation:

 

(3)
 

(4)
 

(5)

These frequencies define the Hardy–Weinberg equilibrium. It should be mentioned that the genotype frequencies after the first generation need not equal the genotype frequencies from the initial generation, e.g. f1(AA) ≠ f0(AA). However, the genotype frequencies for all future times will equal the Hardy–Weinberg frequencies, e.g. ft(AA) = f1(AA) for t > 1. This follows since the genotype frequencies of the next generation depend only on the allele frequencies of the current generation which, as calculated by equations (1) and (2), are preserved from the initial generation:

 

For the more general case of dioecious diploids [organisms are either male or female] that reproduce by random mating of individuals, it is necessary to calculate the genotype frequencies from the nine possible matings between each parental genotype (AA, Aa, and aa) in either sex, weighted by the expected genotype contributions of each such mating.[2] Equivalently, one considers the six unique diploid-diploid combinations:

 

and constructs a Punnett square for each, so as to calculate its contribution to the next generation's genotypes. These contributions are weighted according to the probability of each diploid-diploid combination, which follows a multinomial distribution with k = 3. For example, the probability of the mating combination (AA,aa) is 2 ft(AA)ft(aa) and it can only result in the Aa genotype: [0,1,0]. Overall, the resulting genotype frequencies are calculated as:

 

As before, one can show that the allele frequencies at time t + 1 equal those at time t, and so, are constant in time. Similarly, the genotype frequencies depend only on the allele frequencies, and so, after time t = 1 are also constant in time.

If in either monoecious or dioecious organisms, either the allele or genotype proportions are initially unequal in either sex, it can be shown that constant proportions are obtained after one generation of random mating. If dioecious organisms are heterogametic and the gene locus is located on the X chromosome, it can be shown that if the allele frequencies are initially unequal in the two sexes [e.g., XX females and XY males, as in humans], f′(a) in the heterogametic sex 'chases' f(a) in the homogametic sex of the previous generation, until an equilibrium is reached at the weighted average of the two initial frequencies.

Deviations from Hardy–Weinberg equilibrium edit

The seven assumptions underlying Hardy–Weinberg equilibrium are as follows:[3]

  • organisms are diploid
  • only sexual reproduction occurs
  • generations are nonoverlapping
  • mating is random
  • population size is infinitely large
  • allele frequencies are equal in the sexes
  • there is no migration, gene flow, admixture, mutation or selection

Violations of the Hardy–Weinberg assumptions can cause deviations from expectation. How this affects the population depends on the assumptions that are violated.

  • Random mating. The HWP states the population will have the given genotypic frequencies (called Hardy–Weinberg proportions) after a single generation of random mating within the population. When the random mating assumption is violated, the population will not have Hardy–Weinberg proportions. A common cause of non-random mating is inbreeding, which causes an increase in homozygosity for all genes.

If a population violates one of the following four assumptions, the population may continue to have Hardy–Weinberg proportions each generation, but the allele frequencies will change over time.

  • Selection, in general, causes allele frequencies to change, often quite rapidly. While directional selection eventually leads to the loss of all alleles except the favored one (unless one allele is dominant, in which case recessive alleles can survive at low frequencies), some forms of selection, such as balancing selection, lead to equilibrium without loss of alleles.
  • Mutation will have a very subtle effect on allele frequencies through the introduction of new allele into a population. Mutation rates are of the order 10−4 to 10−8, and the change in allele frequency will be, at most, the same order. Recurrent mutation will maintain alleles in the population, even if there is strong selection against them.
  • Migration genetically links two or more populations together. In general, allele frequencies will become more homogeneous among the populations. Some models for migration inherently include nonrandom mating (Wahlund effect, for example). For those models, the Hardy–Weinberg proportions will normally not be valid.
  • Small population size can cause a random change in allele frequencies. This is due to a sampling effect, and is called genetic drift. Sampling effects are most important when the allele is present in a small number of copies.

In real world genotype data, deviations from Hardy–Weinberg Equilibrium may be a sign of genotyping error.[4][5][6]

Sex linkage edit

Where the A gene is sex linked, the heterogametic sex (e.g., mammalian males; avian females) have only one copy of the gene (and are termed hemizygous), while the homogametic sex (e.g., human females) have two copies. The genotype frequencies at equilibrium are p and q for the heterogametic sex but p2, 2pq and q2 for the homogametic sex.

For example, in humans red–green colorblindness is an X-linked recessive trait. In western European males, the trait affects about 1 in 12, (q = 0.083) whereas it affects about 1 in 200 females (0.005, compared to q2 = 0.007), very close to Hardy–Weinberg proportions.

If a population is brought together with males and females with a different allele frequency in each subpopulation (males or females), the allele frequency of the male population in the next generation will follow that of the female population because each son receives its X chromosome from its mother. The population converges on equilibrium very quickly.

Generalizations edit

The simple derivation above can be generalized for more than two alleles and polyploidy.

Generalization for more than two alleles edit

 
Punnett square for three-allele case (left) and four-allele case (right). White areas are homozygotes. Colored areas are heterozygotes.

Consider an extra allele frequency, r. The two-allele case is the binomial expansion of (p + q)2, and thus the three-allele case is the trinomial expansion of (p + q + r)2.

 

More generally, consider the alleles A1, ..., An given by the allele frequencies p1 to pn;

 

giving for all homozygotes:

 

and for all heterozygotes:

 

Generalization for polyploidy edit

The Hardy–Weinberg principle may also be generalized to polyploid systems, that is, for organisms that have more than two copies of each chromosome. Consider again only two alleles. The diploid case is the binomial expansion of:

 

and therefore the polyploid case is the binomial expansion of:

 

where c is the ploidy, for example with tetraploid (c = 4):

Table 2: Expected genotype frequencies for tetraploidy
Genotype Frequency
AAAA  
AAAa  
AAaa  
Aaaa  
aaaa  

Whether the organism is a 'true' tetraploid or an amphidiploid will determine how long it will take for the population to reach Hardy–Weinberg equilibrium.

Complete generalization edit

For   distinct alleles in  -ploids, the genotype frequencies in the Hardy–Weinberg equilibrium are given by individual terms in the multinomial expansion of  :

 

Significance tests for deviation edit

Testing deviation from the HWP is generally performed using Pearson's chi-squared test, using the observed genotype frequencies obtained from the data and the expected genotype frequencies obtained using the HWP. For systems where there are large numbers of alleles, this may result in data with many empty possible genotypes and low genotype counts, because there are often not enough individuals present in the sample to adequately represent all genotype classes. If this is the case, then the asymptotic assumption of the chi-squared distribution, will no longer hold, and it may be necessary to use a form of Fisher's exact test, which requires a computer to solve. More recently a number of MCMC methods of testing for deviations from HWP have been proposed (Guo & Thompson, 1992; Wigginton et al. 2005)

Example chi-squared test for deviation edit

This data is from E. B. Ford (1971) on the scarlet tiger moth, for which the phenotypes of a sample of the population were recorded. Genotype–phenotype distinction is assumed to be negligibly small. The null hypothesis is that the population is in Hardy–Weinberg proportions, and the alternative hypothesis is that the population is not in Hardy–Weinberg proportions.

Table 3: Example Hardy–Weinberg principle calculation
Phenotype White-spotted (AA) Intermediate (Aa) Little spotting (aa) Total
Number 1469 138 5 1612

From this, allele frequencies can be calculated:

 

and

 

So the Hardy–Weinberg expectation is:

 

Pearson's chi-squared test states:

 

There is 1 degree of freedom (degrees of freedom for test for Hardy–Weinberg proportions are # genotypes − # alleles). The 5% significance level for 1 degree of freedom is 3.84, and since the χ2 value is less than this, the null hypothesis that the population is in Hardy–Weinberg frequencies is not rejected.

Fisher's exact test (probability test) edit

Fisher's exact test can be applied to testing for Hardy–Weinberg proportions. Since the test is conditional on the allele frequencies, p and q, the problem can be viewed as testing for the proper number of heterozygotes. In this way, the hypothesis of Hardy–Weinberg proportions is rejected if the number of heterozygotes is too large or too small. The conditional probabilities for the heterozygote, given the allele frequencies are given in Emigh (1980) as

 

where n11, n12, n22 are the observed numbers of the three genotypes, AA, Aa, and aa, respectively, and n1 is the number of A alleles, where  .

An example Using one of the examples from Emigh (1980),[7] we can consider the case where n = 100, and p = 0.34. The possible observed heterozygotes and their exact significance level is given in Table 4.

Table 4: Example of Fisher's exact test for n = 100, p = 0.34.[7]
Number of heterozygotes Significance level
0 0.000
2 0.000
4 0.000
6 0.000
8 0.000
10 0.000
12 0.000
14 0.000
16 0.000
18 0.001
20 0.007
22 0.034
34 0.067
24 0.151
32 0.291
26 0.474
30 0.730
28 1.000

Using this table, one must look up the significance level of the test based on the observed number of heterozygotes. For example, if one observed 20 heterozygotes, the significance level for the test is 0.007. As is typical for Fisher's exact test for small samples, the gradation of significance levels is quite coarse.

However, a table like this has to be created for every experiment, since the tables are dependent on both n and p.

Equivalence tests edit

The equivalence tests are developed in order to establish sufficiently good agreement of the observed genotype frequencies and Hardy Weinberg equilibrium. Let   denote the family of the genotype distributions under the assumption of Hardy Weinberg equilibrium. The distance between a genotype distribution   and Hardy Weinberg equilibrium is defined by  , where   is some distance. The equivalence test problem is given by   and  , where   is a tolerance parameter. If the hypothesis   can be rejected then the population is close to Hardy Weinberg equilibrium with a high probability. The equivalence tests for the biallelic case are developed among others in Wellek (2004).[8] The equivalence tests for the case of multiple alleles are proposed in Ostrovski (2020).[9]

Inbreeding coefficient edit

The inbreeding coefficient,   (see also F-statistics), is one minus the observed frequency of heterozygotes over that expected from Hardy–Weinberg equilibrium.

 

where the expected value from Hardy–Weinberg equilibrium is given by

 

For example, for Ford's data above:

 

For two alleles, the chi-squared goodness of fit test for Hardy–Weinberg proportions is equivalent to the test for inbreeding,  .

The inbreeding coefficient is unstable as the expected value approaches zero, and thus not useful for rare and very common alleles. For:  ;   is undefined.

History edit

Mendelian genetics were rediscovered in 1900. However, it remained somewhat controversial for several years as it was not then known how it could cause continuous characteristics. Udny Yule (1902) argued against Mendelism because he thought that dominant alleles would increase in the population.[10] The American William E. Castle (1903) showed that without selection, the genotype frequencies would remain stable.[11] Karl Pearson (1903) found one equilibrium position with values of p = q = 0.5.[12] Reginald Punnett, unable to counter Yule's point, introduced the problem to G. H. Hardy, a British mathematician, with whom he played cricket. Hardy was a pure mathematician and held applied mathematics in some contempt; his view of biologists' use of mathematics comes across in his 1908 paper where he describes this as "very simple":[13]

To the Editor of Science: I am reluctant to intrude in a discussion concerning matters of which I have no expert knowledge, and I should have expected the very simple point which I wish to make to have been familiar to biologists. However, some remarks of Mr. Udny Yule, to which Mr. R. C. Punnett has called my attention, suggest that it may still be worth making...
Suppose that Aa is a pair of Mendelian characters, A being dominant, and that in any given generation the number of pure dominants (AA), heterozygotes (Aa), and pure recessives (aa) are as p:2q:r. Finally, suppose that the numbers are fairly large, so that mating may be regarded as random, that the sexes are evenly distributed among the three varieties, and that all are equally fertile. A little mathematics of the multiplication-table type is enough to show that in the next generation the numbers will be as (p + q)2:2(p + q)(q + r):(q + r)2, or as p1:2q1:r1, say.
The interesting question is: in what circumstances will this distribution be the same as that in the generation before? It is easy to see that the condition for this is q2 = pr. And since q12 = p1r1, whatever the values of p, q, and r may be, the distribution will in any case continue unchanged after the second generation

The principle was thus known as Hardy's law in the English-speaking world until 1943, when Curt Stern pointed out that it had first been formulated independently in 1908 by the German physician Wilhelm Weinberg.[14][15] William Castle in 1903 also derived the ratios for the special case of equal allele frequencies, and it is sometimes (but rarely) called the Hardy–Weinberg–Castle Law.

Derivation of Hardy's equations edit

Hardy's statement begins with a recurrence relation for the frequencies p, 2q, and r. These recurrence relations follow from fundamental concepts in probability, specifically independence, and conditional probability. For example, consider the probability of an offspring from the generation   being homozygous dominant. Alleles are inherited independently from each parent. A dominant allele can be inherited from a homozygous dominant parent with probability 1, or from a heterozygous parent with probability 0.5. To represent this reasoning in an equation, let   represent inheritance of a dominant allele from a parent. Furthermore, let   and   represent potential parental genotypes in the preceding generation.

 

The same reasoning, applied to the other genotypes yields the two remaining recurrence relations. Equilibrium occurs when each proportion is constant between subsequent generations. More formally, a population is at equilibrium at generation   when

 ,  , and  

By solving these equations necessary and sufficient conditions for equilibrium to occur can be determined. Again, consider the frequency of homozygous dominant animals. Equilibrium implies

 

First consider the case, where  , and note that it implies that   and  . Now consider the remaining case, where  :

 

where the final equality holds because the allele proportions must sum to one. In both cases,  . It can be shown that the other two equilibrium conditions imply the same equation. Together, the solutions of the three equilibrium equations imply sufficiency of Hardy's condition for equilibrium. Since the condition always holds for the second generation, all succeeding generations have the same proportions.

Numerical example edit

Estimation of genotype distribution edit

An example computation of the genotype distribution given by Hardy's original equations is instructive. The phenotype distribution from Table 3 above will be used to compute Hardy's initial genotype distribution. Note that the p and q values used by Hardy are not the same as those used above.

 
 

As checks on the distribution, compute

 

and

 

For the next generation, Hardy's equations give

 

Again as checks on the distribution, compute

 

and

 

which are the expected values. The reader may demonstrate that subsequent use of the second-generation values for a third generation will yield identical results.

Estimation of carrier frequency edit

The Hardy–Weinberg principle can also be used to estimate the frequency of carriers of an autosomal recessive condition in a population based on the frequency of suffers.

Let us assume an estimated   babies are born with cystic fibrosis, this is about the frequency of homozygous individuals observed in Northern European populations. We can use the Hardy–Weinberg equations to estimate the carrier frequency, the frequency of heterozygous individuals,  .

 

As   is small we can take p,  , to be 1.

 

We therefore estimate the carrier rate to be  , which is about the frequency observed in Northern European populations.

This can be simplified to the carrier frequency being about twice the square root of the birth frequency.

Graphical representation edit

 
A de Finetti diagram representing a distribution of genotype frequencies

It is possible to represent the distribution of genotype frequencies for a bi-allelic locus within a population graphically using a de Finetti diagram. This uses a triangular plot (also known as trilinear, triaxial or ternary plot) to represent the distribution of the three genotype frequencies in relation to each other. It differs from many other such plots in that the direction of one of the axes has been reversed.[16] The curved line in the diagram is the Hardy–Weinberg parabola and represents the state where alleles are in Hardy–Weinberg equilibrium. It is possible to represent the effects of natural selection and its effect on allele frequency on such graphs.[17] The de Finetti diagram was developed and used extensively by A. W. F. Edwards in his book Foundations of Mathematical Genetics.[18]

See also edit

Notes edit

  1. ^ The term frequency usually refers to a number or count, but in this context, it is synonymous with probability.

References edit

Citations edit

  1. ^ Edwards, A. W. F. (2008). "G. H. Hardy (1908) and Hardy–Weinberg Equilibrium". Genetics. 179 (3): 1143–1150. doi:10.1534/genetics.104.92940. ISSN 0016-6731. PMC 2475721. PMID 18645201.
  2. ^ Carr, Dr. Steven M. "Hardy–Weinberg in dioecious organisms". www.mun.ca.
  3. ^ Hartl DL, Clarke AG (2007) Principles of population genetics. Sunderland, MA: Sinauer
  4. ^ Hosking, Louise; Lumsden, Sheena; Lewis, Karen; Yeo, Astrid; McCarthy, Linda; Bansal, Aruna; Riley, John; Purvis, Ian; Xu, Chun-Fang (May 2004). "Detection of genotyping errors by Hardy–Weinberg equilibrium testing". European Journal of Human Genetics. 12 (5): 395–399. doi:10.1038/sj.ejhg.5201164. ISSN 1018-4813. PMID 14872201.
  5. ^ Pompanon, François; Bonin, Aurélie; Bellemain, Eva; Taberlet, Pierre (November 2005). "Genotyping errors: causes, consequences and solutions". Nature Reviews Genetics. 6 (11): 847–859. doi:10.1038/nrg1707. ISSN 1471-0064. PMID 16304600. S2CID 14031116.
  6. ^ Cox, David G.; Kraft, Peter (2006). "Quantification of the Power of Hardy–Weinberg Equilibrium Testing to Detect Genotyping Error". Human Heredity. 61 (1): 10–14. doi:10.1159/000091787. ISSN 0001-5652. PMID 16514241. S2CID 37599930.
  7. ^ a b Emigh, Ted H. (1980). "A Comparison of Tests for Hardy–Weinberg Equilibrium". Biometrics. 36 (4): 627–642. doi:10.2307/2556115. JSTOR 2556115. PMID 25856832.
  8. ^ Wellek, Stefan (September 2004). "Tests for establishing compatibility of an observed genotype distribution with Hardy–Weinberg equilibrium in the case of a biallelic locus". Biometrics. 60 (3): 694–703. doi:10.1111/j.0006-341X.2004.00219.x. PMID 15339292. S2CID 12028776.Official web link (subscription required)
  9. ^ Ostrovski, Vladimir (February 2020). "New equivalence tests for Hardy–Weinberg equilibrium and multiple alleles". Stats. 3: 34–39. doi:10.3390/stats3010004.Official web link
  10. ^ Yule, 1902
  11. ^ Castle, 1903
  12. ^ Pearson, 1903
  13. ^ Hardy, 1908
  14. ^ Crow, James F. (1999). "Hardy, Weinberg and language impediments". Genetics. 152 (3): 821–825. doi:10.1093/genetics/152.3.821. PMC 1460671. PMID 10388804.
  15. ^ Stern, Curt (1962). "Wilhelm Weinberg". Genetics. 47: 1–5.
  16. ^ Cannings, C.; Edwards, A.W.F. (1968). "Natural selection and the de Finetti diagram". Annals of Human Genetics. 31 (4): 421–428. doi:10.1111/j.1469-1809.1968.tb00575.x. PMID 5673165. S2CID 8863631.
  17. ^ See e.g. Ineichen & Batschelet 1975
  18. ^ Edwards, 1977

Sources edit

  • Castle, W. E. (1903). "The laws of Galton and Mendel and some laws governing race improvement by selection". Proceedings of the American Academy of Arts and Sciences. 35: 233–242.
  • Crow, Jf (July 1999). "Hardy, Weinberg and language impediments". Genetics. 152 (3): 821–5. doi:10.1093/genetics/152.3.821. ISSN 0016-6731. PMC 1460671. PMID 10388804.
  • Edwards, A.W.F. 1977. Foundations of Mathematical Genetics. Cambridge University Press, Cambridge (2nd ed., 2000). ISBN 0-521-77544-2
  • Emigh, T.H. (1980). "A comparison of tests for Hardy–Weinberg equilibrium". Biometrics. 36 (4): 627–642. doi:10.2307/2556115. JSTOR 2556115. PMID 25856832.
  • Ford, E.B. (1971). Ecological Genetics, London.
  • Guo, Sw; Thompson, Elizabeth A. (June 1992). "Performing the exact test of Hardy–Weinberg proportion for multiple alleles". Biometrics. 48 (2): 361–72. doi:10.2307/2532296. ISSN 0006-341X. JSTOR 2532296. PMID 1637966.
  • Hardy, G. H. (July 1908). "Mendelian Proportions in a Mixed Population" (PDF). Science. 28 (706): 49–50. Bibcode:1908Sci....28...49H. doi:10.1126/science.28.706.49. ISSN 0036-8075. PMC 2582692. PMID 17779291.
  • Ineichen, Robert; Batschelet, Eduard (1975). "Genetic selection and de Finetti diagrams". Journal of Mathematical Biology. 2: 33–39. doi:10.1007/BF00276014. S2CID 123415153.
  • Masel, Joanna (2012). "Rethinking Hardy–Weinberg and genetic drift in undergraduate biology". BioEssays. 34 (8): 701–10. doi:10.1002/bies.201100178. PMID 22576789. S2CID 28513167.
  • Pearson, K. (1903). "Mathematical contributions to the theory of evolution. XI. On the influence of natural selection on the variability and correlation of organs". Philosophical Transactions of the Royal Society A. 200 (321–330): 1–66. Bibcode:1903RSPTA.200....1P. doi:10.1098/rsta.1903.0001.
  • Stern, C. (1943). "The Hardy–Weinberg law". Science. 97 (2510): 137–138. Bibcode:1943Sci....97..137S. doi:10.1126/science.97.2510.137. JSTOR 1670409. PMID 17788516.
  • Weinberg, W. (1908). "Über den Nachweis der Vererbung beim Menschen". Jahreshefte des Vereins für vaterländische Naturkunde in Württemberg. 64: 368–382.
  • Wigginton, Je; Cutler, Dj; Abecasis, Gr (May 2005). "A Note on Exact Tests of Hardy–Weinberg Equilibrium". American Journal of Human Genetics. 76 (5): 887–93. doi:10.1086/429864. ISSN 0002-9297. PMC 1199378. PMID 15789306.
  • Yule, G. U. (1902). "Mendel's laws and their probable relation to intra-racial heredity". New Phytol. 1 (193–207): 222–238. doi:10.1111/j.1469-8137.1902.tb07336.x.

External links edit

  • EvolutionSolution (at bottom of page)
  • Hardy–Weinberg Equilibrium Calculator
  • genetics Population Genetics Simulator[permanent dead link]
  • HARDY C implementation of Guo & Thompson 1992
  • Source code (C/C++/Fortran/R) for Wigginton et al. 2005
  • Online de Finetti Diagram Generator and Hardy–Weinberg equilibrium tests
  • Online Hardy–Weinberg equilibrium tests and drawing of de Finetti diagrams 26 May 2015 at the Wayback Machine

hardy, weinberg, principle, this, article, includes, list, general, references, lacks, sufficient, corresponding, inline, citations, please, help, improve, this, article, introducing, more, precise, citations, april, 2020, learn, when, remove, this, template, . This article includes a list of general references but it lacks sufficient corresponding inline citations Please help to improve this article by introducing more precise citations April 2020 Learn how and when to remove this template message In population genetics the Hardy Weinberg principle also known as the Hardy Weinberg equilibrium model theorem or law states that allele and genotype frequencies in a population will remain constant from generation to generation in the absence of other evolutionary influences These influences include genetic drift mate choice assortative mating natural selection sexual selection mutation gene flow meiotic drive genetic hitchhiking population bottleneck founder effect inbreeding and outbreeding depression Hardy Weinberg proportions for two alleles the horizontal axis shows the two allele frequencies p and q and the vertical axis shows the expected genotype frequencies Each line shows one of the three possible genotypes In the simplest case of a single locus with two alleles denoted A and a with frequencies f A p and f a q respectively the expected genotype frequencies under random mating are f AA p2 for the AA homozygotes f aa q2 for the aa homozygotes and f Aa 2pq for the heterozygotes In the absence of selection mutation genetic drift or other forces allele frequencies p and q are constant between generations so equilibrium is reached The principle is named after G H Hardy and Wilhelm Weinberg who first demonstrated it mathematically Hardy s paper was focused on debunking the view that a dominant allele would automatically tend to increase in frequency a view possibly based on a misinterpreted question at a lecture 1 Today tests for Hardy Weinberg genotype frequencies are used primarily to test for population stratification and other forms of non random mating Contents 1 Derivation 2 Deviations from Hardy Weinberg equilibrium 3 Sex linkage 4 Generalizations 4 1 Generalization for more than two alleles 4 2 Generalization for polyploidy 4 3 Complete generalization 5 Significance tests for deviation 5 1 Example chi squared test for deviation 5 2 Fisher s exact test probability test 6 Equivalence tests 7 Inbreeding coefficient 8 History 8 1 Derivation of Hardy s equations 8 2 Numerical example 8 2 1 Estimation of genotype distribution 8 2 2 Estimation of carrier frequency 9 Graphical representation 10 See also 11 Notes 12 References 12 1 Citations 12 2 Sources 13 External linksDerivation editConsider a population of monoecious diploids where each organism produces male and female gametes at equal frequency and has two alleles at each gene locus We assume that the population is so large that it can be treated as infinite Organisms reproduce by random union of gametes the gene pool population model A locus in this population has two alleles A and a that occur with initial frequencies f0 A p and f0 a q respectively note 1 The allele frequencies at each generation are obtained by pooling together the alleles from each genotype of the same generation according to the expected contribution from the homozygote and heterozygote genotypes which are 1 and 1 2 respectively f t A f t AA 1 2 f t Aa displaystyle f t text A f t text AA tfrac 1 2 f t text Aa nbsp 1 f t a f t aa 1 2 f t Aa displaystyle f t text a f t text aa tfrac 1 2 f t text Aa nbsp 2 nbsp Length of p q corresponds to allele frequencies here p 0 6 q 0 4 Then area of rectangle represents genotype frequencies thus AA Aa aa 0 36 0 48 0 16 The different ways to form genotypes for the next generation can be shown in a Punnett square where the proportion of each genotype is equal to the product of the row and column allele frequencies from the current generation Table 1 Punnett square for Hardy Weinberg Females A p a q Males A p AA p2 Aa pq a q Aa qp aa q2 The sum of the entries is p2 2pq q2 1 as the genotype frequencies must sum to one Note again that as p q 1 the binomial expansion of p q 2 p2 2pq q2 1 gives the same relationships Summing the elements of the Punnett square or the binomial expansion we obtain the expected genotype proportions among the offspring after a single generation f 1 AA p 2 f 0 A 2 displaystyle f 1 text AA p 2 f 0 text A 2 nbsp 3 f 1 Aa p q q p 2 p q 2 f 0 A f 0 a displaystyle f 1 text Aa pq qp 2pq 2f 0 text A f 0 text a nbsp 4 f 1 aa q 2 f 0 a 2 displaystyle f 1 text aa q 2 f 0 text a 2 nbsp 5 These frequencies define the Hardy Weinberg equilibrium It should be mentioned that the genotype frequencies after the first generation need not equal the genotype frequencies from the initial generation e g f1 AA f0 AA However the genotype frequencies for all future times will equal the Hardy Weinberg frequencies e g ft AA f1 AA for t gt 1 This follows since the genotype frequencies of the next generation depend only on the allele frequencies of the current generation which as calculated by equations 1 and 2 are preserved from the initial generation f 1 A f 1 AA 1 2 f 1 Aa p 2 p q p p q p f 0 A f 1 a f 1 aa 1 2 f 1 Aa q 2 p q q p q q f 0 a displaystyle begin aligned f 1 text A amp f 1 text AA tfrac 1 2 f 1 text Aa p 2 pq p p q p f 0 text A f 1 text a amp f 1 text aa tfrac 1 2 f 1 text Aa q 2 pq q p q q f 0 text a end aligned nbsp For the more general case of dioecious diploids organisms are either male or female that reproduce by random mating of individuals it is necessary to calculate the genotype frequencies from the nine possible matings between each parental genotype AA Aa and aa in either sex weighted by the expected genotype contributions of each such mating 2 Equivalently one considers the six unique diploid diploid combinations AA AA AA Aa AA aa Aa Aa Aa aa aa aa displaystyle left text AA text AA text AA text Aa text AA text aa text Aa text Aa text Aa text aa text aa text aa right nbsp and constructs a Punnett square for each so as to calculate its contribution to the next generation s genotypes These contributions are weighted according to the probability of each diploid diploid combination which follows a multinomial distribution with k 3 For example the probability of the mating combination AA aa is 2 ft AA ft aa and it can only result in the Aa genotype 0 1 0 Overall the resulting genotype frequencies are calculated as f t 1 AA f t 1 Aa f t 1 aa f t AA f t AA 1 0 0 2 f t AA f t Aa 1 2 1 2 0 2 f t AA f t aa 0 1 0 f t Aa f t Aa 1 4 1 2 1 4 2 f t Aa f t aa 0 1 2 1 2 f t aa f t aa 0 0 1 f t AA 1 2 f t Aa 2 2 f t AA 1 2 f t Aa f t aa 1 2 f t Aa f t aa 1 2 f t Aa 2 f t A 2 2 f t A f t a f t a 2 displaystyle begin aligned amp left f t 1 text AA f t 1 text Aa f t 1 text aa right amp qquad f t text AA f t text AA left 1 0 0 right 2f t text AA f t text Aa left tfrac 1 2 tfrac 1 2 0 right 2f t text AA f t text aa left 0 1 0 right amp qquad qquad f t text Aa f t text Aa left tfrac 1 4 tfrac 1 2 tfrac 1 4 right 2f t text Aa f t text aa left 0 tfrac 1 2 tfrac 1 2 right f t text aa f t text aa left 0 0 1 right amp qquad left left f t text AA tfrac 1 2 f t text Aa right 2 2 left f t text AA tfrac 1 2 f t text Aa right left f t text aa tfrac 1 2 f t text Aa right left f t text aa tfrac 1 2 f t text Aa right 2 right amp qquad left f t text A 2 2f t text A f t text a f t text a 2 right end aligned nbsp As before one can show that the allele frequencies at time t 1 equal those at time t and so are constant in time Similarly the genotype frequencies depend only on the allele frequencies and so after time t 1 are also constant in time If in either monoecious or dioecious organisms either the allele or genotype proportions are initially unequal in either sex it can be shown that constant proportions are obtained after one generation of random mating If dioecious organisms are heterogametic and the gene locus is located on the X chromosome it can be shown that if the allele frequencies are initially unequal in the two sexes e g XX females and XY males as in humans f a in the heterogametic sex chases f a in the homogametic sex of the previous generation until an equilibrium is reached at the weighted average of the two initial frequencies Deviations from Hardy Weinberg equilibrium editThe seven assumptions underlying Hardy Weinberg equilibrium are as follows 3 organisms are diploid only sexual reproduction occurs generations are nonoverlapping mating is random population size is infinitely large allele frequencies are equal in the sexes there is no migration gene flow admixture mutation or selection Violations of the Hardy Weinberg assumptions can cause deviations from expectation How this affects the population depends on the assumptions that are violated Random mating The HWP states the population will have the given genotypic frequencies called Hardy Weinberg proportions after a single generation of random mating within the population When the random mating assumption is violated the population will not have Hardy Weinberg proportions A common cause of non random mating is inbreeding which causes an increase in homozygosity for all genes If a population violates one of the following four assumptions the population may continue to have Hardy Weinberg proportions each generation but the allele frequencies will change over time Selection in general causes allele frequencies to change often quite rapidly While directional selection eventually leads to the loss of all alleles except the favored one unless one allele is dominant in which case recessive alleles can survive at low frequencies some forms of selection such as balancing selection lead to equilibrium without loss of alleles Mutation will have a very subtle effect on allele frequencies through the introduction of new allele into a population Mutation rates are of the order 10 4 to 10 8 and the change in allele frequency will be at most the same order Recurrent mutation will maintain alleles in the population even if there is strong selection against them Migration genetically links two or more populations together In general allele frequencies will become more homogeneous among the populations Some models for migration inherently include nonrandom mating Wahlund effect for example For those models the Hardy Weinberg proportions will normally not be valid Small population size can cause a random change in allele frequencies This is due to a sampling effect and is called genetic drift Sampling effects are most important when the allele is present in a small number of copies In real world genotype data deviations from Hardy Weinberg Equilibrium may be a sign of genotyping error 4 5 6 Sex linkage editWhere the A gene is sex linked the heterogametic sex e g mammalian males avian females have only one copy of the gene and are termed hemizygous while the homogametic sex e g human females have two copies The genotype frequencies at equilibrium are p and q for the heterogametic sex but p2 2pq and q2 for the homogametic sex For example in humans red green colorblindness is an X linked recessive trait In western European males the trait affects about 1 in 12 q 0 083 whereas it affects about 1 in 200 females 0 005 compared to q2 0 007 very close to Hardy Weinberg proportions If a population is brought together with males and females with a different allele frequency in each subpopulation males or females the allele frequency of the male population in the next generation will follow that of the female population because each son receives its X chromosome from its mother The population converges on equilibrium very quickly Generalizations editThe simple derivation above can be generalized for more than two alleles and polyploidy Generalization for more than two alleles edit nbsp Punnett square for three allele case left and four allele case right White areas are homozygotes Colored areas are heterozygotes Consider an extra allele frequency r The two allele case is the binomial expansion of p q 2 and thus the three allele case is the trinomial expansion of p q r 2 p q r 2 p 2 q 2 r 2 2 p q 2 p r 2 q r displaystyle p q r 2 p 2 q 2 r 2 2pq 2pr 2qr nbsp More generally consider the alleles A1 An given by the allele frequencies p1 to pn p 1 p n 2 displaystyle p 1 cdots p n 2 nbsp giving for all homozygotes f A i A i p i 2 displaystyle f A i A i p i 2 nbsp and for all heterozygotes f A i A j 2 p i p j displaystyle f A i A j 2p i p j nbsp Generalization for polyploidy edit The Hardy Weinberg principle may also be generalized to polyploid systems that is for organisms that have more than two copies of each chromosome Consider again only two alleles The diploid case is the binomial expansion of p q 2 displaystyle p q 2 nbsp and therefore the polyploid case is the binomial expansion of p q c displaystyle p q c nbsp where c is the ploidy for example with tetraploid c 4 Table 2 Expected genotype frequencies for tetraploidy Genotype Frequency AAAA p 4 displaystyle p 4 nbsp AAAa 4 p 3 q displaystyle 4p 3 q nbsp AAaa 6 p 2 q 2 displaystyle 6p 2 q 2 nbsp Aaaa 4 p q 3 displaystyle 4pq 3 nbsp aaaa q 4 displaystyle q 4 nbsp Whether the organism is a true tetraploid or an amphidiploid will determine how long it will take for the population to reach Hardy Weinberg equilibrium Complete generalization edit For n displaystyle n nbsp distinct alleles in c displaystyle c nbsp ploids the genotype frequencies in the Hardy Weinberg equilibrium are given by individual terms in the multinomial expansion of p 1 p n c displaystyle p 1 cdots p n c nbsp p 1 p n c k 1 k n N k 1 k n c c k 1 k n p 1 k 1 p n k n displaystyle p 1 cdots p n c sum k 1 ldots k n in mathbb N k 1 cdots k n c c choose k 1 ldots k n p 1 k 1 cdots p n k n nbsp Significance tests for deviation editTesting deviation from the HWP is generally performed using Pearson s chi squared test using the observed genotype frequencies obtained from the data and the expected genotype frequencies obtained using the HWP For systems where there are large numbers of alleles this may result in data with many empty possible genotypes and low genotype counts because there are often not enough individuals present in the sample to adequately represent all genotype classes If this is the case then the asymptotic assumption of the chi squared distribution will no longer hold and it may be necessary to use a form of Fisher s exact test which requires a computer to solve More recently a number of MCMC methods of testing for deviations from HWP have been proposed Guo amp Thompson 1992 Wigginton et al 2005 Example chi squared test for deviation edit This data is from E B Ford 1971 on the scarlet tiger moth for which the phenotypes of a sample of the population were recorded Genotype phenotype distinction is assumed to be negligibly small The null hypothesis is that the population is in Hardy Weinberg proportions and the alternative hypothesis is that the population is not in Hardy Weinberg proportions Table 3 Example Hardy Weinberg principle calculation Phenotype White spotted AA Intermediate Aa Little spotting aa Total Number 1469 138 5 1612 From this allele frequencies can be calculated p 2 o b s AA o b s Aa 2 o b s AA o b s Aa o b s aa 2 1469 138 2 1469 138 5 3076 3224 0 954 displaystyle begin aligned p amp 2 times mathrm obs text AA mathrm obs text Aa over 2 times mathrm obs text AA mathrm obs text Aa mathrm obs text aa amp 2 times 1469 138 over 2 times 1469 138 5 amp 3076 over 3224 amp 0 954 end aligned nbsp and q 1 p 1 0 954 0 046 displaystyle begin aligned q amp 1 p amp 1 0 954 amp 0 046 end aligned nbsp So the Hardy Weinberg expectation is E x p AA p 2 n 0 954 2 1612 1467 4 E x p Aa 2 p q n 2 0 954 0 046 1612 141 2 E x p aa q 2 n 0 046 2 1612 3 4 displaystyle begin aligned mathrm Exp text AA amp p 2 n 0 954 2 times 1612 1467 4 mathrm Exp text Aa amp 2pqn 2 times 0 954 times 0 046 times 1612 141 2 mathrm Exp text aa amp q 2 n 0 046 2 times 1612 3 4 end aligned nbsp Pearson s chi squared test states x 2 O E 2 E 1469 1467 4 2 1467 4 138 141 2 2 141 2 5 3 4 2 3 4 0 001 0 073 0 756 0 83 displaystyle begin aligned chi 2 amp sum O E 2 over E amp 1469 1467 4 2 over 1467 4 138 141 2 2 over 141 2 5 3 4 2 over 3 4 amp 0 001 0 073 0 756 amp 0 83 end aligned nbsp There is 1 degree of freedom degrees of freedom for test for Hardy Weinberg proportions are genotypes alleles The 5 significance level for 1 degree of freedom is 3 84 and since the x2 value is less than this the null hypothesis that the population is in Hardy Weinberg frequencies is not rejected Fisher s exact test probability test edit Fisher s exact test can be applied to testing for Hardy Weinberg proportions Since the test is conditional on the allele frequencies p and q the problem can be viewed as testing for the proper number of heterozygotes In this way the hypothesis of Hardy Weinberg proportions is rejected if the number of heterozygotes is too large or too small The conditional probabilities for the heterozygote given the allele frequencies are given in Emigh 1980 as prob n 12 n 1 n n 11 n 12 n 22 2 n n 1 n 2 2 n 12 displaystyle operatorname prob n 12 mid n 1 frac binom n n 11 n 12 n 22 binom 2n n 1 n 2 2 n 12 nbsp where n11 n12 n22 are the observed numbers of the three genotypes AA Aa and aa respectively and n1 is the number of A alleles where n 1 2 n 11 n 12 displaystyle n 1 2n 11 n 12 nbsp An example Using one of the examples from Emigh 1980 7 we can consider the case where n 100 and p 0 34 The possible observed heterozygotes and their exact significance level is given in Table 4 Table 4 Example of Fisher s exact test for n 100 p 0 34 7 Number of heterozygotes Significance level 0 0 000 2 0 000 4 0 000 6 0 000 8 0 000 10 0 000 12 0 000 14 0 000 16 0 000 18 0 001 20 0 007 22 0 034 34 0 067 24 0 151 32 0 291 26 0 474 30 0 730 28 1 000 Using this table one must look up the significance level of the test based on the observed number of heterozygotes For example if one observed 20 heterozygotes the significance level for the test is 0 007 As is typical for Fisher s exact test for small samples the gradation of significance levels is quite coarse However a table like this has to be created for every experiment since the tables are dependent on both n and p Equivalence tests editThe equivalence tests are developed in order to establish sufficiently good agreement of the observed genotype frequencies and Hardy Weinberg equilibrium Let M displaystyle mathcal M nbsp denote the family of the genotype distributions under the assumption of Hardy Weinberg equilibrium The distance between a genotype distribution p displaystyle p nbsp and Hardy Weinberg equilibrium is defined by d p M min q M d p q displaystyle d p mathcal M min q in mathcal M d p q nbsp where d displaystyle d nbsp is some distance The equivalence test problem is given by H 0 d p M e displaystyle H 0 d p mathcal M geq varepsilon nbsp and H 1 d p M lt e displaystyle H 1 d p mathcal M lt varepsilon nbsp where e gt 0 displaystyle varepsilon gt 0 nbsp is a tolerance parameter If the hypothesis H 0 displaystyle H 0 nbsp can be rejected then the population is close to Hardy Weinberg equilibrium with a high probability The equivalence tests for the biallelic case are developed among others in Wellek 2004 8 The equivalence tests for the case of multiple alleles are proposed in Ostrovski 2020 9 Inbreeding coefficient editThe inbreeding coefficient F displaystyle F nbsp see also F statistics is one minus the observed frequency of heterozygotes over that expected from Hardy Weinberg equilibrium F E f Aa O f Aa E f Aa 1 O f Aa E f Aa displaystyle F frac operatorname E f text Aa operatorname O f text Aa operatorname E f text Aa 1 frac operatorname O f text Aa operatorname E f text Aa nbsp where the expected value from Hardy Weinberg equilibrium is given by E f Aa 2 p q displaystyle operatorname E f text Aa 2pq nbsp For example for Ford s data above F 1 138 141 2 0 023 displaystyle F 1 138 over 141 2 0 023 nbsp For two alleles the chi squared goodness of fit test for Hardy Weinberg proportions is equivalent to the test for inbreeding F 0 displaystyle F 0 nbsp The inbreeding coefficient is unstable as the expected value approaches zero and thus not useful for rare and very common alleles For F E 0 O 0 displaystyle F big E 0 O 0 infty nbsp F E 0 O gt 0 displaystyle F big E 0 O gt 0 nbsp is undefined History editMendelian genetics were rediscovered in 1900 However it remained somewhat controversial for several years as it was not then known how it could cause continuous characteristics Udny Yule 1902 argued against Mendelism because he thought that dominant alleles would increase in the population 10 The American William E Castle 1903 showed that without selection the genotype frequencies would remain stable 11 Karl Pearson 1903 found one equilibrium position with values of p q 0 5 12 Reginald Punnett unable to counter Yule s point introduced the problem to G H Hardy a British mathematician with whom he played cricket Hardy was a pure mathematician and held applied mathematics in some contempt his view of biologists use of mathematics comes across in his 1908 paper where he describes this as very simple 13 To the Editor of Science I am reluctant to intrude in a discussion concerning matters of which I have no expert knowledge and I should have expected the very simple point which I wish to make to have been familiar to biologists However some remarks of Mr Udny Yule to which Mr R C Punnett has called my attention suggest that it may still be worth making Suppose that Aa is a pair of Mendelian characters A being dominant and that in any given generation the number of pure dominants AA heterozygotes Aa and pure recessives aa are as p 2q r Finally suppose that the numbers are fairly large so that mating may be regarded as random that the sexes are evenly distributed among the three varieties and that all are equally fertile A little mathematics of the multiplication table type is enough to show that in the next generation the numbers will be as p q 2 2 p q q r q r 2 or as p1 2q1 r1 say The interesting question is in what circumstances will this distribution be the same as that in the generation before It is easy to see that the condition for this is q2 pr And since q12 p1r1 whatever the values of p q and r may be the distribution will in any case continue unchanged after the second generation The principle was thus known as Hardy s law in the English speaking world until 1943 when Curt Stern pointed out that it had first been formulated independently in 1908 by the German physician Wilhelm Weinberg 14 15 William Castle in 1903 also derived the ratios for the special case of equal allele frequencies and it is sometimes but rarely called the Hardy Weinberg Castle Law Derivation of Hardy s equations edit Hardy s statement begins with a recurrence relation for the frequencies p 2q and r These recurrence relations follow from fundamental concepts in probability specifically independence and conditional probability For example consider the probability of an offspring from the generation t displaystyle textstyle t nbsp being homozygous dominant Alleles are inherited independently from each parent A dominant allele can be inherited from a homozygous dominant parent with probability 1 or from a heterozygous parent with probability 0 5 To represent this reasoning in an equation let A t displaystyle textstyle A t nbsp represent inheritance of a dominant allele from a parent Furthermore let A A t 1 displaystyle textstyle AA t 1 nbsp and A a t 1 displaystyle textstyle Aa t 1 nbsp represent potential parental genotypes in the preceding generation p t P A t A t P A t 2 P A t A A t 1 P A A t 1 P A t A a t 1 P A a t 1 2 1 p t 1 0 5 2 q t 1 2 p t 1 q t 1 2 displaystyle begin aligned p t amp P A t A t P A t 2 amp left P A t mid AA t 1 P AA t 1 P A t mid Aa t 1 P Aa t 1 right 2 amp left 1 p t 1 0 5 2q t 1 right 2 amp left p t 1 q t 1 right 2 end aligned nbsp The same reasoning applied to the other genotypes yields the two remaining recurrence relations Equilibrium occurs when each proportion is constant between subsequent generations More formally a population is at equilibrium at generation t displaystyle textstyle t nbsp when 0 p t p t 1 displaystyle textstyle 0 p t p t 1 nbsp 0 q t q t 1 displaystyle textstyle 0 q t q t 1 nbsp and 0 r t r t 1 displaystyle textstyle 0 r t r t 1 nbsp By solving these equations necessary and sufficient conditions for equilibrium to occur can be determined Again consider the frequency of homozygous dominant animals Equilibrium implies 0 p t p t 1 p t 1 2 2 p t 1 q t 1 q t 1 2 p t 1 displaystyle begin aligned 0 amp p t p t 1 amp p t 1 2 2p t 1 q t 1 q t 1 2 p t 1 end aligned nbsp First consider the case where p t 1 0 displaystyle textstyle p t 1 0 nbsp and note that it implies that q t 1 0 displaystyle textstyle q t 1 0 nbsp and r t 1 1 displaystyle textstyle r t 1 1 nbsp Now consider the remaining case where p t 1 0 displaystyle textstyle p t 1 neq textstyle 0 nbsp 0 p t 1 p t 1 2 q t 1 q t 1 2 p t 1 1 q t 1 2 p t 1 r t 1 displaystyle begin aligned 0 amp p t 1 p t 1 2q t 1 q t 1 2 p t 1 1 amp q t 1 2 p t 1 r t 1 end aligned nbsp where the final equality holds because the allele proportions must sum to one In both cases q t 1 2 p t 1 r t 1 displaystyle textstyle q t 1 2 p t 1 r t 1 nbsp It can be shown that the other two equilibrium conditions imply the same equation Together the solutions of the three equilibrium equations imply sufficiency of Hardy s condition for equilibrium Since the condition always holds for the second generation all succeeding generations have the same proportions Numerical example edit Estimation of genotype distribution edit An example computation of the genotype distribution given by Hardy s original equations is instructive The phenotype distribution from Table 3 above will be used to compute Hardy s initial genotype distribution Note that the p and q values used by Hardy are not the same as those used above sum o b s AA 2 o b s Aa o b s aa 1469 2 138 5 1750 displaystyle begin aligned text sum amp mathrm obs text AA 2 times mathrm obs text Aa mathrm obs text aa 1469 2 times 138 5 5pt amp 1750 end aligned nbsp p 1469 1750 0 83943 2 q 2 138 1750 0 15771 r 5 1750 0 00286 displaystyle begin aligned p amp 1469 over 1750 0 83943 5pt 2q amp 2 times 138 over 1750 0 15771 5pt r amp 5 over 1750 0 00286 end aligned nbsp As checks on the distribution compute p 2 q r 0 83943 0 15771 0 00286 1 00000 displaystyle p 2q r 0 83943 0 15771 0 00286 1 00000 nbsp and E 0 q 2 p r 0 00382 displaystyle E 0 q 2 pr 0 00382 nbsp For the next generation Hardy s equations give q 0 15771 2 0 07886 p 1 p q 2 0 84325 2 q 1 2 p q q r 0 15007 r 1 q r 2 0 00668 displaystyle begin aligned q amp 0 15771 over 2 0 07886 p 1 amp p q 2 0 84325 5pt 2q 1 amp 2 p q q r 0 15007 5pt r 1 amp q r 2 0 00668 end aligned nbsp Again as checks on the distribution compute p 1 2 q 1 r 1 0 84325 0 15007 0 00668 1 00000 displaystyle p 1 2q 1 r 1 0 84325 0 15007 0 00668 1 00000 nbsp and E 1 q 1 2 p 1 r 1 0 00000 displaystyle E 1 q 1 2 p 1 r 1 0 00000 nbsp which are the expected values The reader may demonstrate that subsequent use of the second generation values for a third generation will yield identical results Estimation of carrier frequency edit The Hardy Weinberg principle can also be used to estimate the frequency of carriers of an autosomal recessive condition in a population based on the frequency of suffers Let us assume an estimated 1 2500 displaystyle textstyle frac 1 2500 nbsp babies are born with cystic fibrosis this is about the frequency of homozygous individuals observed in Northern European populations We can use the Hardy Weinberg equations to estimate the carrier frequency the frequency of heterozygous individuals 2 p q displaystyle textstyle 2pq nbsp q 2 1 2500 q 1 50 p 1 q displaystyle begin aligned amp q 2 frac 1 2500 5pt amp q frac 1 50 5pt amp p 1 q end aligned nbsp As 1 50 displaystyle textstyle frac 1 50 nbsp is small we can take p 1 1 50 displaystyle textstyle 1 frac 1 50 nbsp to be 1 2 p q 2 1 50 2 p q 1 25 displaystyle begin aligned 2pq 2 cdot frac 1 50 5pt 2pq frac 1 25 end aligned nbsp We therefore estimate the carrier rate to be 1 25 displaystyle textstyle frac 1 25 nbsp which is about the frequency observed in Northern European populations This can be simplified to the carrier frequency being about twice the square root of the birth frequency Graphical representation edit nbsp A de Finetti diagram representing a distribution of genotype frequencies It is possible to represent the distribution of genotype frequencies for a bi allelic locus within a population graphically using a de Finetti diagram This uses a triangular plot also known as trilinear triaxial or ternary plot to represent the distribution of the three genotype frequencies in relation to each other It differs from many other such plots in that the direction of one of the axes has been reversed 16 The curved line in the diagram is the Hardy Weinberg parabola and represents the state where alleles are in Hardy Weinberg equilibrium It is possible to represent the effects of natural selection and its effect on allele frequency on such graphs 17 The de Finetti diagram was developed and used extensively by A W F Edwards in his book Foundations of Mathematical Genetics 18 See also editRegression toward the mean Multinomial distribution Hardy Weinberg is a trinomial distribution with probabilities 8 2 2 8 1 8 1 8 2 displaystyle theta 2 2 theta 1 theta 1 theta 2 nbsp Additive disequilibrium and z statistic Population genetics Genetic diversity Founder effect Population bottleneck Genetic drift Inbreeding depression Natural selection Fitness Genetic loadNotes edit The term frequency usually refers to a number or count but in this context it is synonymous with probability References editCitations edit Edwards A W F 2008 G H Hardy 1908 and Hardy Weinberg Equilibrium Genetics 179 3 1143 1150 doi 10 1534 genetics 104 92940 ISSN 0016 6731 PMC 2475721 PMID 18645201 Carr Dr Steven M Hardy Weinberg in dioecious organisms www mun ca Hartl DL Clarke AG 2007 Principles of population genetics Sunderland MA Sinauer Hosking Louise Lumsden Sheena Lewis Karen Yeo Astrid McCarthy Linda Bansal Aruna Riley John Purvis Ian Xu Chun Fang May 2004 Detection of genotyping errors by Hardy Weinberg equilibrium testing European Journal of Human Genetics 12 5 395 399 doi 10 1038 sj ejhg 5201164 ISSN 1018 4813 PMID 14872201 Pompanon Francois Bonin Aurelie Bellemain Eva Taberlet Pierre November 2005 Genotyping errors causes consequences and solutions Nature Reviews Genetics 6 11 847 859 doi 10 1038 nrg1707 ISSN 1471 0064 PMID 16304600 S2CID 14031116 Cox David G Kraft Peter 2006 Quantification of the Power of Hardy Weinberg Equilibrium Testing to Detect Genotyping Error Human Heredity 61 1 10 14 doi 10 1159 000091787 ISSN 0001 5652 PMID 16514241 S2CID 37599930 a b Emigh Ted H 1980 A Comparison of Tests for Hardy Weinberg Equilibrium Biometrics 36 4 627 642 doi 10 2307 2556115 JSTOR 2556115 PMID 25856832 Wellek Stefan September 2004 Tests for establishing compatibility of an observed genotype distribution with Hardy Weinberg equilibrium in the case of a biallelic locus Biometrics 60 3 694 703 doi 10 1111 j 0006 341X 2004 00219 x PMID 15339292 S2CID 12028776 Official web link subscription required Ostrovski Vladimir February 2020 New equivalence tests for Hardy Weinberg equilibrium and multiple alleles Stats 3 34 39 doi 10 3390 stats3010004 Official web link Yule 1902 Castle 1903 Pearson 1903 Hardy 1908 Crow James F 1999 Hardy Weinberg and language impediments Genetics 152 3 821 825 doi 10 1093 genetics 152 3 821 PMC 1460671 PMID 10388804 Stern Curt 1962 Wilhelm Weinberg Genetics 47 1 5 Cannings C Edwards A W F 1968 Natural selection and the de Finetti diagram Annals of Human Genetics 31 4 421 428 doi 10 1111 j 1469 1809 1968 tb00575 x PMID 5673165 S2CID 8863631 See e g Ineichen amp Batschelet 1975 Edwards 1977 Sources edit Castle W E 1903 The laws of Galton and Mendel and some laws governing race improvement by selection Proceedings of the American Academy of Arts and Sciences 35 233 242 Crow Jf July 1999 Hardy Weinberg and language impediments Genetics 152 3 821 5 doi 10 1093 genetics 152 3 821 ISSN 0016 6731 PMC 1460671 PMID 10388804 Edwards A W F 1977 Foundations of Mathematical Genetics Cambridge University Press Cambridge 2nd ed 2000 ISBN 0 521 77544 2 Emigh T H 1980 A comparison of tests for Hardy Weinberg equilibrium Biometrics 36 4 627 642 doi 10 2307 2556115 JSTOR 2556115 PMID 25856832 Ford E B 1971 Ecological Genetics London Guo Sw Thompson Elizabeth A June 1992 Performing the exact test of Hardy Weinberg proportion for multiple alleles Biometrics 48 2 361 72 doi 10 2307 2532296 ISSN 0006 341X JSTOR 2532296 PMID 1637966 Hardy G H July 1908 Mendelian Proportions in a Mixed Population PDF Science 28 706 49 50 Bibcode 1908Sci 28 49H doi 10 1126 science 28 706 49 ISSN 0036 8075 PMC 2582692 PMID 17779291 Ineichen Robert Batschelet Eduard 1975 Genetic selection and de Finetti diagrams Journal of Mathematical Biology 2 33 39 doi 10 1007 BF00276014 S2CID 123415153 Masel Joanna 2012 Rethinking Hardy Weinberg and genetic drift in undergraduate biology BioEssays 34 8 701 10 doi 10 1002 bies 201100178 PMID 22576789 S2CID 28513167 Pearson K 1903 Mathematical contributions to the theory of evolution XI On the influence of natural selection on the variability and correlation of organs Philosophical Transactions of the Royal Society A 200 321 330 1 66 Bibcode 1903RSPTA 200 1P doi 10 1098 rsta 1903 0001 Stern C 1943 The Hardy Weinberg law Science 97 2510 137 138 Bibcode 1943Sci 97 137S doi 10 1126 science 97 2510 137 JSTOR 1670409 PMID 17788516 Weinberg W 1908 Uber den Nachweis der Vererbung beim Menschen Jahreshefte des Vereins fur vaterlandische Naturkunde in Wurttemberg 64 368 382 Wigginton Je Cutler Dj Abecasis Gr May 2005 A Note on Exact Tests of Hardy Weinberg Equilibrium American Journal of Human Genetics 76 5 887 93 doi 10 1086 429864 ISSN 0002 9297 PMC 1199378 PMID 15789306 Yule G U 1902 Mendel s laws and their probable relation to intra racial heredity New Phytol 1 193 207 222 238 doi 10 1111 j 1469 8137 1902 tb07336 x External links edit nbsp Wikimedia Commons has media related to Hardy Weinberg law EvolutionSolution at bottom of page Hardy Weinberg Equilibrium Calculator genetics Population Genetics Simulator permanent dead link HARDY C implementation of Guo amp Thompson 1992 Source code C C Fortran R for Wigginton et al 2005 Online de Finetti Diagram Generator and Hardy Weinberg equilibrium tests Online Hardy Weinberg equilibrium tests and drawing of de Finetti diagrams Archived 26 May 2015 at the Wayback Machine Hardy Weinberg Equilibrium Calculator Retrieved from https en wikipedia org w index php title Hardy Weinberg principle amp oldid 1187129917, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.