In probability theory , the multinomial distribution is a generalization of the binomial distribution . For example, it models the probability of counts for each side of a k -sided dice rolled n times. For n independent trials each of which leads to a success for exactly one of k categories, with each category having a given fixed success probability, the multinomial distribution gives the probability of any particular combination of numbers of successes for the various categories.
81-654: 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 . In
162-622: A 1 i p ^ i = b 1 , ∑ i a 2 i p ^ i = b 2 , ⋯ , ∑ i a ℓ i p ^ i = b ℓ {\displaystyle {\begin{cases}\sum _{i}{\hat {p}}_{i}=1,\\\sum _{i}a_{1i}{\hat {p}}_{i}=b_{1},\\\sum _{i}a_{2i}{\hat {p}}_{i}=b_{2},\\\cdots ,\\\sum _{i}a_{\ell i}{\hat {p}}_{i}=b_{\ell }\end{cases}}} (notice that
243-537: A simplex with a grid. Similarly, just like one can interpret the binomial distribution as the polynomial coefficients of ( p + q ) n {\displaystyle (p+q)^{n}} when expanded, one can interpret the multinomial distribution as the coefficients of ( p 1 + p 2 + p 3 + ⋯ + p k ) n {\displaystyle (p_{1}+p_{2}+p_{3}+\cdots +p_{k})^{n}} when expanded, noting that just
324-466: A bag, replacing the extracted balls after each draw. Balls of the same color are equivalent. Denote the variable which is the number of extracted balls of color i ( i = 1, ..., k ) as X i , and denote as p i the probability that a given extraction will be in color i . The probability mass function of this multinomial distribution is: for non-negative integers x 1 , ..., x k . The probability mass function can be expressed using
405-429: A categorical distribution is equivalent to a multinomial distribution over a single trial. The goal of equivalence testing is to establish the agreement between a theoretical multinomial distribution and observed counting frequencies. The theoretical distribution may be a fully specified multinomial distribution or a parametric family of multinomial distributions. Let q {\displaystyle q} denote
486-425: A discrete probability distribution to a continuous probability density, we need to multiply by the volume occupied by each point of Δ k , n {\displaystyle \Delta _{k,n}} in Δ k {\displaystyle \Delta _{k}} . However, by symmetry, every point occupies exactly the same volume (except a negligible set on the boundary), so we obtain
567-413: A fixed sample size . The multinomial distribution is normalized according to: where the sum is over all permutations of x j {\displaystyle x_{j}} such that ∑ j = 1 k x j = n {\displaystyle \sum _{j=1}^{k}x_{j}=n} . The expected number of times the outcome i was observed over n trials
648-449: 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) 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
729-554: A head) or failure (obtaining a tail). The binomial distribution generalizes this to the number of heads from performing n independent flips (Bernoulli trials) of the same coin. The multinomial distribution models the outcome of n experiments, where the outcome of each trial has a categorical distribution , such as rolling a k -sided die n times. Let k be a fixed finite number. Mathematically, we have k possible mutually exclusive outcomes, with corresponding probabilities p 1 , ..., p k , and n independent trials. Since
810-438: A matrix with i, j element cov ( X i , X j ) , {\displaystyle \operatorname {cov} (X_{i},X_{j}),} the result is a k × k positive-semidefinite covariance matrix of rank k − 1. In the special case where k = n and where the p i are all equal, the covariance matrix is the centering matrix . The entries of
891-416: A multinomial distribution when a categorical distribution is actually meant. This stems from the fact that it is sometimes convenient to express the outcome of a categorical distribution as a "1-of-k" vector (a vector with one element containing a 1 and all other elements containing a 0) rather than as an integer in the range 1 … k {\displaystyle 1\dots k} ; in this form,
SECTION 10
#1732780224593972-575: A multinomial distribution with parameters n and p , where p = ( p 1 , ..., p k ). While the trials are independent, their outcomes X i are dependent because they must be summed to n. n ∈ { 0 , 1 , 2 , … } {\displaystyle n\in \{0,1,2,\ldots \}} number of trials k > 0 {\displaystyle k>0} number of mutually exclusive events (integer) Suppose one does an experiment of extracting n balls of k different colors from
1053-517: 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. The simple derivation above can be generalized for more than two alleles and polyploidy . Consider an extra allele frequency, r . The two-allele case
1134-537: 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 f 0 (A) = p and f 0 (a) = q , respectively. The allele frequencies at each generation are obtained by pooling together
1215-420: 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. In real world genotype data, deviations from Hardy–Weinberg Equilibrium may be a sign of genotyping error. Where the A gene is sex linked , the heterogametic sex ( e.g. , mammalian males; avian females) have only one copy of
1296-456: A probability density ρ ( p ^ ) = C e − n 2 ∑ i ( p ^ i − p i ) 2 p i {\displaystyle \rho ({\hat {p}})=Ce^{-{\frac {n}{2}}\sum _{i}{\frac {({\hat {p}}_{i}-p_{i})^{2}}{p_{i}}}}} , where C {\displaystyle C}
1377-514: A rational number, whereas p 1 , p 2 , . . . , p k {\displaystyle p_{1},p_{2},...,p_{k}} may be chosen from any real number in [ 0 , 1 ] {\displaystyle [0,1]} and need not satisfy the Diophantine system of equations. Only asymptotically as n → ∞ {\displaystyle n\rightarrow \infty } ,
1458-449: A theoretical multinomial distribution and let p {\displaystyle p} be a true underlying distribution. The distributions p {\displaystyle p} and q {\displaystyle q} are considered equivalent if d ( p , q ) < ε {\displaystyle d(p,q)<\varepsilon } for a distance d {\displaystyle d} and
1539-535: A tolerance parameter ε > 0 {\displaystyle \varepsilon >0} . The equivalence test problem is H 0 = { d ( p , q ) ≥ ε } {\displaystyle H_{0}=\{d(p,q)\geq \varepsilon \}} versus H 1 = { d ( p , q ) < ε } {\displaystyle H_{1}=\{d(p,q)<\varepsilon \}} . The true underlying distribution p {\displaystyle p}
1620-411: Is The covariance matrix is as follows. Each diagonal entry is the variance of a binomially distributed random variable, and is therefore The off-diagonal entries are the covariances : for i , j distinct. All covariances are negative because for fixed n , an increase in one component of a multinomial vector requires a decrease in another component. When these expressions are combined into
1701-477: 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 . The 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}}} denote
SECTION 20
#17327802245931782-412: Is 1, it is the categorical distribution . The term "multinoulli" is sometimes used for the categorical distribution to emphasize this four-way relationship (so n determines the suffix, and k the prefix). The Bernoulli distribution models the outcome of a single Bernoulli trial . In other words, it models whether flipping a (possibly biased ) coin one time will result in either a success (obtaining
1863-488: Is 3.84, and since the χ value is less than this, the null hypothesis that the population is in Hardy–Weinberg frequencies is not rejected. 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
1944-424: Is a constant. Finally, since the simplex Δ k {\displaystyle \Delta _{k}} is not all of R k {\displaystyle \mathbb {R} ^{k}} , but only within a ( k − 1 ) {\displaystyle (k-1)} -dimensional plane, we obtain the desired result. The above concentration phenomenon can be easily generalized to
2025-601: Is a tolerance parameter. If the hypothesis H 0 {\displaystyle H_{0}} 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). The equivalence tests for the case of multiple alleles are proposed in Ostrovski (2020). The inbreeding coefficient, F {\displaystyle F} (see also F -statistics ),
2106-435: 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, F = 0 {\displaystyle F=0} . The inbreeding coefficient
2187-526: Is proposed in Frey (2009). The distance between the true underlying distribution p {\displaystyle p} and a family of the multinomial distributions M {\displaystyle {\mathcal {M}}} is defined by d ( p , M ) = min h ∈ M d ( p , h ) {\displaystyle d(p,{\mathcal {M}})=\min _{h\in {\mathcal {M}}}d(p,h)} . Then
2268-473: 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). Today, tests for Hardy–Weinberg genotype frequencies are used primarily to test for population stratification and other forms of non-random mating. Consider
2349-558: 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 n 11 , n 12 , n 22 are the observed numbers of the three genotypes, AA, Aa, and aa, respectively, and n 1 is the number of A alleles, where n 1 = 2 n 11 + n 12 {\displaystyle n_{1}=2n_{11}+n_{12}} . An example Using one of
2430-659: Is restricted to a ( k − ℓ − 1 ) {\displaystyle (k-\ell -1)} -dimensional plane. In particular, expanding the KL divergence D K L ( p ^ | | p ) {\displaystyle D_{KL}({\hat {p}}\vert \vert p)} around its minimum q {\displaystyle q} (the I {\displaystyle I} -projection of p {\displaystyle p} on Δ k , n {\displaystyle \Delta _{k,n}} ) in
2511-505: Is some distance. The equivalence test problem is given by H 0 = { d ( p , M ) ≥ ε } {\displaystyle H_{0}=\{d(p,{\mathcal {M}})\geq \varepsilon \}} and H 1 = { d ( p , M ) < ε } {\displaystyle H_{1}=\{d(p,{\mathcal {M}})<\varepsilon \}} , where ε > 0 {\displaystyle \varepsilon >0}
Hardy–Weinberg principle - Misplaced Pages Continue
2592-474: Is that the population is in Hardy–Weinberg proportions, and the alternative hypothesis is that the population is not in Hardy–Weinberg proportions. 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
2673-538: Is the binomial expansion of ( p + q ), and thus the three-allele case is the trinomial expansion of ( p + q + r ). More generally, consider the alleles A 1 , ..., A n given by the allele frequencies p 1 to p n ; giving for all homozygotes : and for all heterozygotes : 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
2754-448: 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): 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. For n {\displaystyle n} distinct alleles in c {\displaystyle c} -ploids,
2835-494: Is the intersection between Δ k {\displaystyle \Delta _{k}} and the lattice ( Z k ) / n {\displaystyle (\mathbb {Z} ^{k})/n} . As n {\displaystyle n} increases, most of the probability mass is concentrated in a subset of Δ k , n {\displaystyle \Delta _{k,n}} near p {\displaystyle p} , and
2916-440: Is the intersection of ( Z k ) / n {\displaystyle (\mathbb {Z} ^{k})/n} with Δ k {\displaystyle \Delta _{k}} and ℓ {\displaystyle \ell } hyperplanes, all linearly independent, so the probability density ρ ( p ^ ) {\displaystyle \rho ({\hat {p}})}
2997-441: Is unknown. Instead, the counting frequencies p n {\displaystyle p_{n}} are observed, where n {\displaystyle n} is a sample size. An equivalence test uses p n {\displaystyle p_{n}} to reject H 0 {\displaystyle H_{0}} . If H 0 {\displaystyle H_{0}} can be rejected then
3078-530: 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 } ; F | E = 0 , O > 0 {\displaystyle F{\big |}_{E=0,O>0}} is undefined. Mendelian genetics were rediscovered in 1900. However, it remained somewhat controversial for several years as it
3159-419: The p ^ i {\displaystyle {\hat {p}}_{i}} 's can be regarded as probabilities over [ 0 , 1 ] {\displaystyle [0,1]} . Away from empirically observed constraints b 1 , … , b ℓ {\displaystyle b_{1},\ldots ,b_{\ell }} (such as moments or prevalences)
3240-480: The chi-squared distribution χ 2 ( k − 1 − ℓ ) {\displaystyle \chi ^{2}(k-1-\ell )} . An analogous proof applies in this Diophantine problem of coupled linear equations in count variables n p ^ i {\displaystyle n{\hat {p}}_{i}} , but this time Δ k , n {\displaystyle \Delta _{k,n}}
3321-728: The chi-squared distribution χ 2 ( k − 1 ) {\displaystyle \chi ^{2}(k-1)} . The space of all distributions over categories { 1 , 2 , … , k } {\displaystyle \{1,2,\ldots ,k\}} is a simplex : Δ k = { ( y 1 , … , y k ) : y 1 , … , y k ≥ 0 , ∑ i y i = 1 } {\displaystyle \Delta _{k}=\left\{(y_{1},\ldots ,y_{k})\colon y_{1},\ldots ,y_{k}\geq 0,\sum _{i}y_{i}=1\right\}} , and
Hardy–Weinberg principle - Misplaced Pages Continue
3402-551: The gamma function as: This form shows its resemblance to the Dirichlet distribution , which is its conjugate prior . Suppose that in a three-way election for a large country, candidate A received 20% of the votes, candidate B received 30% of the votes, and candidate C received 50% of the votes. If six voters are selected randomly, what is the probability that there will be exactly one supporter for candidate A, two supporters for candidate B and three supporters for candidate C in
3483-431: 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. The seven assumptions underlying Hardy–Weinberg equilibrium are as follows: Violations of the Hardy–Weinberg assumptions can cause deviations from expectation. How this affects the population depends on the assumptions that are violated. If
3564-468: The k outcomes are mutually exclusive and one must occur we have p i ≥ 0 for i = 1, ..., k and ∑ i = 1 k p i = 1 {\displaystyle \sum _{i=1}^{k}p_{i}=1} . Then if the random variables X i indicate the number of times outcome number i is observed over the n trials, the vector X = ( X 1 , ..., X k ) follows
3645-418: 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. f 1 (AA) ≠ f 0 (AA) . However, the genotype frequencies for all future times will equal the Hardy–Weinberg frequencies, e.g. f t (AA) = f 1 (AA) for t > 1 . This follows since the genotype frequencies of
3726-412: 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
3807-421: 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: 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. The sum of
3888-437: The asymptotic formula, the probability that empirical distribution p ^ {\displaystyle {\hat {p}}} deviates from the actual distribution p {\displaystyle p} decays exponentially, at a rate n D K L ( p ^ ‖ p ) {\displaystyle nD_{KL}({\hat {p}}\|p)} . The more experiments and
3969-587: The case where we condition upon linear constraints. This is the theoretical justification for Pearson's chi-squared test . Theorem. Given frequencies x i ∈ N {\displaystyle x_{i}\in \mathbb {N} } observed in a dataset with n {\displaystyle n} points, we impose ℓ + 1 {\displaystyle \ell +1} independent linear constraints { ∑ i p ^ i = 1 , ∑ i
4050-1238: The coefficients must sum up to 1. By Stirling's formula , at the limit of n , x 1 , . . . , x k → ∞ {\displaystyle n,x_{1},...,x_{k}\to \infty } , we have ln ( n x 1 , ⋯ , x k ) + ∑ i = 1 k x i ln p i = − n D K L ( p ^ ‖ p ) − k − 1 2 ln ( 2 π n ) − 1 2 ∑ i = 1 k ln ( p ^ i ) + o ( 1 ) {\displaystyle \ln {\binom {n}{x_{1},\cdots ,x_{k}}}+\sum _{i=1}^{k}x_{i}\ln p_{i}=-nD_{KL}({\hat {p}}\|p)-{\frac {k-1}{2}}\ln(2\pi n)-{\frac {1}{2}}\sum _{i=1}^{k}\ln({\hat {p}}_{i})+o(1)} where relative frequencies p ^ i = x i / n {\displaystyle {\hat {p}}_{i}=x_{i}/n} in
4131-470: The column vector p . Just like one can interpret the binomial distribution as (normalized) one-dimensional (1D) slices of Pascal's triangle , so too can one interpret the multinomial distribution as 2D (triangular) slices of Pascal's pyramid , or 3D/4D/+ (pyramid-shaped) slices of higher-dimensional analogs of Pascal's triangle. This reveals an interpretation of the range of the distribution: discretized equilateral "pyramids" in arbitrary dimension—i.e.
SECTION 50
#17327802245934212-712: The constrained problem ensures by the Pythagorean theorem for I {\displaystyle I} -divergence that any constant and linear term in the counts n p ^ i {\displaystyle n{\hat {p}}_{i}} vanishes from the conditional probability to multinationally sample those counts. Notice that by definition, every one of p ^ 1 , p ^ 2 , . . . , p ^ k {\displaystyle {\hat {p}}_{1},{\hat {p}}_{2},...,{\hat {p}}_{k}} must be
4293-422: The corresponding correlation matrix are Note that the number of trials n drops out of this expression. Each of the k components separately has a binomial distribution with parameters n and p i , for the appropriate value of the subscript i . The support of the multinomial distribution is the set Its number of elements is In matrix notation, and with p = the row vector transpose of
4374-515: 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
4455-461: The data can be interpreted as probabilities from the empirical distribution p ^ {\displaystyle {\hat {p}}} , and D K L {\displaystyle D_{KL}} is the Kullback–Leibler divergence . This formula can be interpreted as follows. Consider Δ k {\displaystyle \Delta _{k}} ,
4536-477: The details below. Request from 172.68.168.237 via cp1104 cp1104, Varnish XID 209399794 Upstream caches: cp1104 int Error: 429, Too Many Requests at Thu, 28 Nov 2024 07:50:24 GMT Multinomial distribution When k is 2 and n is 1, the multinomial distribution is the Bernoulli distribution . When k is 2 and n is bigger than 1, it is the binomial distribution . When k is bigger than 2 and n
4617-461: The entries is p + 2 pq + q = 1 , as the genotype frequencies must sum to one. Note again that as p + q = 1 , the binomial expansion of ( p + q ) = p + 2 pq + q = 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: These frequencies define
4698-401: The equivalence between p {\displaystyle p} and q {\displaystyle q} is shown at a given significance level. The equivalence test for Euclidean distance can be found in text book of Wellek (2010). The equivalence test for the total variation distance is developed in Ostrovski (2017). The exact equivalence test for the specific cumulative distance
4779-507: The equivalence test problem is given by H 0 = { d ( p , M ) ≥ ε } {\displaystyle H_{0}=\{d(p,{\mathcal {M}})\geq \varepsilon \}} and H 1 = { d ( p , M ) < ε } {\displaystyle H_{1}=\{d(p,{\mathcal {M}})<\varepsilon \}} . The distance d ( p , M ) {\displaystyle d(p,{\mathcal {M}})}
4860-457: The examples from Emigh (1980), 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. 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
4941-416: The expected genotype contributions of each such mating. 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,
SECTION 60
#17327802245935022-1317: The exponential decay, at large n {\displaystyle n} , almost all the probability mass is concentrated in a small neighborhood of p {\displaystyle p} . In this small neighborhood, we can take the first nonzero term in the Taylor expansion of D K L {\displaystyle D_{KL}} , to obtain ln ( n x 1 , ⋯ , x k ) p 1 x 1 ⋯ p k x k ≈ − n 2 ∑ i = 1 k ( p ^ i − p i ) 2 p i = − 1 2 ∑ i = 1 k ( x i − n p i ) 2 n p i {\displaystyle \ln {\binom {n}{x_{1},\cdots ,x_{k}}}p_{1}^{x_{1}}\cdots p_{k}^{x_{k}}\approx -{\frac {n}{2}}\sum _{i=1}^{k}{\frac {({\hat {p}}_{i}-p_{i})^{2}}{p_{i}}}=-{\frac {1}{2}}\sum _{i=1}^{k}{\frac {(x_{i}-np_{i})^{2}}{np_{i}}}} This resembles
5103-485: The family of the genotype distributions under the assumption of Hardy Weinberg equilibrium. The distance between a genotype distribution p {\displaystyle p} 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)} , where d {\displaystyle d}
5184-489: The first constraint is simply the requirement that the empirical distributions sum to one), such that empirical p ^ i = x i / n {\displaystyle {\hat {p}}_{i}=x_{i}/n} satisfy all these constraints simultaneously. Let q {\displaystyle q} denote the I {\displaystyle I} -projection of prior distribution p {\displaystyle p} on
5265-699: The gaussian distribution, which suggests the following theorem: Theorem. At the n → ∞ {\displaystyle n\to \infty } limit, n ∑ i = 1 k ( p ^ i − p i ) 2 p i = ∑ i = 1 k ( x i − n p i ) 2 n p i {\displaystyle n\sum _{i=1}^{k}{\frac {({\hat {p}}_{i}-p_{i})^{2}}{p_{i}}}=\sum _{i=1}^{k}{\frac {(x_{i}-np_{i})^{2}}{np_{i}}}} converges in distribution to
5346-539: 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 p , 2 pq and q 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 q = 0.007), very close to Hardy–Weinberg proportions. If
5427-416: 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}} : Testing deviation from the HWP is generally performed using Pearson's chi-squared test , using the observed genotype frequencies obtained from
5508-1014: The growth rate of P r ( p ^ ∈ A ϵ ) {\displaystyle Pr({\hat {p}}\in A_{\epsilon })} on each piece A ϵ {\displaystyle A_{\epsilon }} , we obtain Sanov's theorem , which states that lim n → ∞ 1 n ln P r ( p ^ ∈ A ) = − inf p ^ ∈ A D K L ( p ^ ‖ p ) {\displaystyle \lim _{n\to \infty }{\frac {1}{n}}\ln Pr({\hat {p}}\in A)=-\inf _{{\hat {p}}\in A}D_{KL}({\hat {p}}\|p)} Due to
5589-464: The more different p ^ {\displaystyle {\hat {p}}} is from p {\displaystyle p} , the less likely it is to see such an empirical distribution. If A {\displaystyle A} is a closed subset of Δ k {\displaystyle \Delta _{k}} , then by dividing up A {\displaystyle A} into pieces, and reasoning about
5670-487: 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
5751-716: The probability distribution near p {\displaystyle p} becomes well-approximated by ( n x 1 , ⋯ , x k ) p 1 x 1 ⋯ p k x k ≈ e − n 2 ∑ i ( p ^ i − p i ) 2 p i {\displaystyle {\binom {n}{x_{1},\cdots ,x_{k}}}p_{1}^{x_{1}}\cdots p_{k}^{x_{k}}\approx e^{-{\frac {n}{2}}\sum _{i}{\frac {({\hat {p}}_{i}-p_{i})^{2}}{p_{i}}}}} From this, we see that
5832-585: The probability of the mating combination (AA,aa) is 2 f t (AA) f t (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
5913-506: 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": Population genetics Too Many Requests If you report this error to the Wikimedia System Administrators, please include
5994-413: The sample? Note: Since we’re assuming that the voting population is large, it is reasonable and permissible to think of the probabilities as unchanging once a voter is selected for the sample. Technically speaking this is sampling without replacement, so the correct distribution is the multivariate hypergeometric distribution , but the distributions converge as the population grows large in comparison to
6075-614: The set of all possible empirical distributions after n {\displaystyle n} experiments is a subset of the simplex: Δ k , n = { ( x 1 / n , … , x k / n ) : x 1 , … , x k ∈ N , ∑ i x i = n } {\displaystyle \Delta _{k,n}=\left\{(x_{1}/n,\ldots ,x_{k}/n)\colon x_{1},\ldots ,x_{k}\in \mathbb {N} ,\sum _{i}x_{i}=n\right\}} . That is, it
6156-531: 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) = p for the AA homozygotes , f (aa) = q for the aa homozygotes, and f (Aa) = 2 pq 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
6237-519: The space of all possible distributions over the categories { 1 , 2 , . . . , k } {\displaystyle \{1,2,...,k\}} . It is a simplex . After n {\displaystyle n} independent samples from the categorical distribution p {\displaystyle p} (which is how we construct the multinomial distribution), we obtain an empirical distribution p ^ {\displaystyle {\hat {p}}} . By
6318-800: The sub-region of the simplex allowed by the linear constraints. At the n → ∞ {\displaystyle n\to \infty } limit, sampled counts n p ^ i {\displaystyle n{\hat {p}}_{i}} from the multinomial distribution conditional on the linear constraints are governed by 2 n D K L ( p ^ | | q ) ≈ n ∑ i ( p ^ i − q i ) 2 q i {\displaystyle 2nD_{KL}({\hat {p}}\vert \vert q)\approx n\sum _{i}{\frac {({\hat {p}}_{i}-q_{i})^{2}}{q_{i}}}} which converges in distribution to
6399-403: The subset upon which the mass is concentrated has radius on the order of 1 / n {\displaystyle 1/{\sqrt {n}}} , but the points in the subset are separated by distance on the order of 1 / n {\displaystyle 1/n} , so at large n {\displaystyle n} , the points merge into a continuum. To convert this from
6480-521: The theorem can be generalized: Theorem. In the case that all p ^ i {\displaystyle {\hat {p}}_{i}} are equal, the Theorem reduces to the concentration of entropies around the Maximum Entropy. In some fields such as natural language processing , categorical and multinomial distributions are synonymous and it is common to speak of
6561-473: 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. The American William E. Castle (1903) showed that without selection , the genotype frequencies would remain stable. Karl Pearson (1903) found one equilibrium position with values of p = q = 0.5. Reginald Punnett , unable to counter Yule's point, introduced
#592407