Skip to main content

METHODS article

Front. Genet., 21 April 2020
Sec. Evolutionary and Population Genetics

Exact Distribution of Linkage Disequilibrium in the Presence of Mutation, Selection, or Minor Allele Frequency Filtering

Updated
\nJiayi QuJiayi Qu1Stephen D. KachmanStephen D. Kachman2Dorian GarrickDorian Garrick3Rohan L. FernandoRohan L. Fernando4Hao Cheng
Hao Cheng1*
  • 1Department of Animal Science, University of California, Davis, Davis, CA, United States
  • 2Department of Statistics, University of Nebraska Lincoln, Lincoln, NE, United States
  • 3School of Agriculture, Massey University, Wellington, New Zealand
  • 4Department of Animal Science, Iowa State University, Ames, IA, United States

Linkage disequilibrium (LD), often expressed in terms of the squared correlation (r2) between allelic values at two loci, is an important concept in many branches of genetics and genomics. Genetic drift and recombination have opposite effects on LD, and thus r2 will keep changing until the effects of these two forces are counterbalanced. Several approximations have been used to determine the expected value of r2 at equilibrium in the presence or absence of mutation. In this paper, we propose a probability-based approach to compute the exact distribution of allele frequencies at two loci in a finite population at any generation t conditional on the distribution at generation t − 1. As r2 is a function of this distribution of allele frequencies, this approach can be used to examine the distribution of r2 over generations as it approaches equilibrium. The exact distribution of LD from our method is used to describe, quantify, and compare LD at different equilibria, including equilibrium in the absence or presence of mutation, selection, and filtering by minor allele frequency. We also propose a deterministic formula for expected LD in the presence of mutation at equilibrium based on the exact distribution of LD.

1. Introduction

Linkage disequilibrium (LD), the non-random association of alleles at two or more loci, is an important concept in various areas of genetics, including evolutionary, quantitative, and statistical genetics. In evolutionary genetics, LD is used to detect the genomic locations of historical selection and to estimate the time of divergence and geographic subdivision between populations (Sved and Hill, 2018). For example, LD was used to date the divergence of the European population from the African population using the HapMap data (Sved et al., 2008). In the field of quantitative and statistical genetics, methods for genomic prediction and genome-wide association studies (GWAS) rely on the existence of LD between the molecular markers and the unobserved quantitative trait loci (QTL).

Two statistics that have been widely used to quantify the LD between allelic values at two loci are: their covariance (D); or their squared correlation (r2) (Hill and Robertson, 1968). For a diallelic system, they can be expressed as

D=pA1B1-pA1pB1    (1)

and

r2=D2pA1(1-pA1)pB1(1-pB1)    (2)

respectively, where pAi represents the frequency of the ith allele at locus A, pBj represents the frequency of the jth allele at locus B, and pAiBj represents the frequency of haplotype AiBj in the population. Although the covariance of different pairs of alleles depends on their corresponding haplotype frequencies and the respective allele frequencies, a single value of D is sufficient to characterize the disequilibrium between two loci in a diallelic system, i.e., D = DA1B1 = −DA1B2 = −DA2B1 = DA2B2. The value of D depends on how the alleles are coded (i.e., alleles A1 and A2 could be coded as 0 and 1 or as 1 and 0) while the value of r2 is invariant to this choice. Therefore, r2 is increasingly used as a metric to quantify LD in the literature.

In a finite population, allele frequencies at two loci are subject to random fluctuations due to the stochastic sampling of a finite number of gametes across generations. This random change of the gene frequency between one generation and the next is genetic drift. Due to the randomness of the genetic drift, one initial random-mating population can evolve into one of a finite collection of sub-populations or lines with different allele frequencies at the two loci, and thus with different LD. Usually, this finite collection of possible lines is jointly considered to understand the distribution of the allele frequencies and LD of two linked loci over generations. This is equivalent to considering a finite collection of pairs of loci in one line. The consequences of the genetic drift, mutation, and selection for the collection of lines of two loci equally apply to a collection of pairs of loci in one line. The former conceptualization is used in this paper to understand the effects of genetic drift, mutation, selection, and minor allele frequency (MAF) cutoff on the distribution of LD.

Generally, in a finite population, the opposite effects on LD of genetic drift and recombination lead to an equilibrium value for the expectation of r2 when these two forces are counterbalanced. It has been shown that for large values of Nec, the equilibrium value for the expectation of r2 (rE2) is approximately:

rE2=11+4Nec    (3)

where c is the recombination rate and Ne is the effective population size (Sved, 1971; Sved and Feldman, 1973; Hill, 1975).

The above formula, however, does not consider the variation introduced by mutation in the population. When mutation is considered, a balance is expected between the loss of variation by the fixation of one or other allele, and the replenishment of an extinct allele by mutation. An approximation for the equilibrium value of r2 in the presence of mutation in a finite population has been developed in Ohta and Kimura (1971) and Hill (1975). Instead of using E(r2), they used a quantity (σd2) named the standard linkage deviation (Ohta and Kimura, 1969) to describe the status of this equilibrium, where σd2 is an approximation of the expected value of r2 expressed as the ratio of the expectations of the numerator and the denominator of r2 (Ohta and Kimura, 1971):

σd2=E(D2)E(pA1(1-pA1)pB1(1-pB1)).    (4)

The equilibrium value of σd2 is approximated as:

σd2=10+ρ+4θ22+13ρ+32θ+ρ2+6ρθ+8θ2,    (5)

where ρ = 4Nec, θ = 4Neu and u is the mutation rate per site (Ohta and Kimura, 1971). The equilibrium value of σd2 as an approximation of rE2 has been described in Walsh and Lynch (2018). Similarly, this value is considered in this paper as an approximation of rE2 in the presence of mutation.

Selection is another important force affecting LD that was not considered in either Sved's or Hill's formulas. Selection occurs at causal variants or QTL during evolution (natural selection) or selective breeding (artificial selection). The impact of selection can either decrease or increase the LD in a population given different scenarios (Mitchell-Olds et al., 2007). The equilibrium among mutation and selection is of special interest in terms of the distribution of allele frequencies and of r2.

In contrast to the approximate methods mentioned above, in this paper we will derive the exact distribution of allele frequencies and LD of two linked loci at equilibrium in the absence or presence of mutation, selection, and MAF cutoff. The distribution of LD in the presence of filtering by MAF is considered in our study due to the prevalent practice of filtering out marker loci with low MAF in genomic analyses used for genetic evaluation or QTL discovery (GWAS).

Our exact distributions are first used to validate the approximate deterministic formulas for rE2, such as Sved's and Hill's approximations. Next we describe, quantify and compare rE2 at different equilibria: including equilibrium in the presence of mutation; equilibrium in the presence of mutation and selection; with or without filtering by MAF. Finally, we calibrate Sved's deterministic formula for rE2 in the presence of mutation based on the exact distribution of LD. The objective of this paper is to present a computational approach to derive the exact distribution of LD over generations and use it to study the distribution of LD whether or not there is mutation, selection, or filtering by MAF.

2. Materials and Methods

In this section, we will show how to compute the exact distribution of allele frequencies at two linked loci in generation t, given their distribution in generation t−1, as a function of the effective population size, the recombination rate between the loci, the mutation rate, the selection coefficient, and the MAF threshold. As LD is a function of the allele frequencies at the two loci, its distribution over generations can be calculated based on the distribution of the allele frequencies.

2.1. Transition-Matrix Approach in the Presence of Mutation, Selection, or Filtering by Minor Allele Frequency

Two diallelic loci A and B, with alleles A1 and A2 at the A locus and alleles B1 and B2 at the B locus, are considered. To incorporate selection and MAF threshold, we consider the A locus as the causal variant under selection and B locus as the molecular marker under the MAF filtering. Four possible haplotypes (i.e., A1B1, A1B2, A2B1, and A2B2) are possible at these two loci. In a population of size Ne, the frequency counts of these four haplotypes can take on k=(2Ne+3)!3!(2Ne)! possible values, where for each of these possibilities, the sum of the four counts is 2Ne. For example, when Ne is equal to two, all possible combinations of haplotype frequency counts are given in Figure S1. In general, let X be a k × 4 matrix with each row representing a possible combination of haplotype frequency counts for some value of Ne. Thus, the rows of X represent the collection of k lines with different allele frequencies at two loci. Let Pt denote a k × 1 vector with element i indicating the probability of the frequency counts in row i of X at generation t. Thus, Pt gives the distribution of allele frequencies at generation t. We show below that the distribution in generation t+1 can be written in general as

Pt+1=APt,    (6)

where A is a k × k transition matrix that is derived below in various circumstances of recombination, mutation, and selection.

First, consider a line with haplotype frequency counts xiT from ith row of X at generation t, e.g., xiT=[f11,f12,f21,f22], where the frequency counts for the four haplotypes, A1B1, A1B2, A2B1, and A2B2, are denoted by f11, f12, f21 and f22. Ignoring the effects of recombination, mutation, or selection, sampling of 2Ne gametes from this line can be modeled by a multinomial process with sample size n = 2Ne and probability vector θi=xiT2Ne for the four haplotype probabilities. Thus, the distribution of the frequency count in the next generation is given by the k × 1 vector mi, where the element j of mi is the probability of getting xjT from the Multinomial(n, θi) distribution. Given that element i of Pt is the probability of xiT in generation t, the distribution of allele frequencies in the next generation is given by:

Pt+1=MPt,    (7)

where M is the k × k matrix with columns mi, as described above, for i = 1, …, k. The above formula shows how the distribution of allele frequencies change due to genetic drift, ignoring recombination, mutation, and selection. Now we accommodate recombination in computing Pt+1. In gametes produced from generation t, the probability of a non-recombinant A1B1 is (1-c)×f112Ne, where c is the recombination rate between locus A and locus B. On the other hand, a recombinant A1B1 can be produced in one of four ways. They and their associated probabilities are:

1. Alleles A1 and B1 originate from two different A1B1 haplotypes with probability c×f112Ne×f11-12Ne-1.

2. Allele A1 originates from an A1B1 haplotype and B1 originates from an A2B1 haplotype with probability c×f112Ne×f212Ne-1.

3. Allele A1 originates from an A1B2 haplotype and B1 originates from an A1B1 haplotype with probability c×f122Ne×f112Ne-1.

4. Allele A1 originates from an A1B2 haplotype and B1 originates from an A2B1 haplotype with probability c×f122Ne×f212Ne-1.

Thus, accounting for recombination, the probability of a A1B1 haplotype changes from f112Ne to:

Pr(A1B1)=(1-c)f112Ne+                      cf112Ne[(f11-1)2Ne-1+f212Ne-1]+                     cf122Ne[f112Ne-1+f212Ne-1].    (8)

Similarly, due to recombination, the probabilities of the other three haplotypes become:

Pr(A1B2)=(1-c)f122Ne+                        cf112Ne[f122Ne-1+f222Ne-1]+                      cf122Ne[(f12-1)2Ne-1+f222Ne-1],    (9)
Pr(A2B1)=(1-c)f212Ne+                      cf212Ne[f112Ne-1+(f21-1)2Ne-1]+                      cf222Ne[f112Ne-1+f212Ne-1],    (10)

and

Pr(A2B2)=(1-c)f222Ne+                      cf212Ne[f122Ne-1+f222Ne-1]+                      cf222Ne[f122Ne-1+(f22-1)2Ne-1].    (11)

Now, to see how mutation alters these probabilities, we let θi* be a vector of the four probabilities from Equations (8) through (11) computed using the four haplotype frequencies in xiT. In modeling mutation, we assume that an A1 or B1 allele mutates to an A2 or B2 allele with probability u and an A2 or B2 allele mutates to an A1 or B1 allele with probability v. Then, haplotype probabilities following mutation can be computed as:

θi**=Tθi*,    (12)

where

T=[(1-u)2(1-u)vv(1-u)v2(1-u)u(1-u)(1-v)vuv(1-v)u(1-u)uv(1-v)(1-u)(1-v)vu2u(1-v)(1-v)u(1-v)2].

Furthermore, to incorporate selection in the model, a selection coefficient s that reduces the allele frequency of A1 at locus A, which is a causal variant, is considered. Conditional on the haplotype probabilities in θi**, i.e., pA1B1**, pA1B2**, pA2B1**, and pA2B2**, the haplotype probabilities following selection can be computed as

θi***=[(1-s)pA1B1**(1-s)(pA1B1**+pA1B2**)+pA2B1**+pA2B2**(1-s)pA1B2**(1-s)(pA1B1**+pA1B2**)+pA2B1**+pA2B2**pA2B1**(1-s)(pA1B1**+pA1B2**)+pA2B1**+pA2B2*pA2B2**(1-s)(pA1B1**+pA1B2**)+pA2B1**+pA2B2**]

Finally, to compute the distribution of allele frequencies in generation t+1 using Equation (6), where the forces of the genetic drift, recombination, mutation and selection are simultaneously considered, the matrix A is defined such that element j of column i contains the probability of getting xjT from a Multinomial (2Ne,θi***T) distribution, for i, j = 1, …, k.

The value of r2 can be computed for a sub-population or line with haplotype frequency counts in any row of X. For example, consider the frequency counts in row 11 of the X matrix given in Figure S1, where f11 = 0, f12 = 2, f21 = 1 and f22 = 1. From these frequency counts, Pr(A1) = 1/2, Pr(B1) = 1/4, and Pr(A1B1) = 0 are obtained, and r2 calculated using Equations (1) and (2) is 1/3 for a line with haplotype frequency counts in x11T from Figure S1. The distribution of r2 in generation t is given by values of r2 corresponding to the frequency counts in each row of X together with the probabilities for haplotype frequency counts given in Pt. Note that haplotype frequency counts in rows of X that have indeterminate values of r2 (i.e., when one or other allele is extinct) are not used to compute the distribution of r2. Similarly, the distribution of r2 with MAF cutoff is given by considering only the values of r2 and their probabilities corresponding to the rows of X with MAF at the B locus ≥5%. MAF of 0.05 is used as a threshold in the present study.

Starting with an allele frequency of 0.5 at each locus and linkage equilibrium between the two loci, the expected value of r2 was computed over generations given some values of the mutation rate, recombination rate, selection coefficient and effective population size. Mutation rate of u = v = 0 is used to represent the absence of mutation, while u = v = 1 × 10−9 is used to represent the existence of mutation. Similarly, a selection coefficient of s = 0 is used to represent the absence of selection, while s = 0.1 or s = 0.01 is used to represent the existence of selection.

2.2. Data Analysis

In our analysis, different population parameters are considered. Populations with Ne ranging from 5 to 50 in intervals of 5 are compared under different recombination rates, mutation rates (i.e., u = 0 or u = 1e−9) and selection coefficients for the causal variant (i.e., s = 0.1 or s = 0.01 for locus A), in the absence or presence of filtering by MAF (MAF≥0.05 for locus B). The recombination rates ranged from 0.01 to 0.1 in intervals of 0.01 and from 0.1 to 0.5 in intervals of 0.1.

The recombination rate for adjacent markers in a bovine 777 k SNP panel was considered to mimic a realistic scenario. The recombination rate for adjacent markers was estimated to be 6.25 × 10−5 using the Kosambi map function given that the mean distance between adjacent markers is around 5 kb (Espigolan et al., 2013), and the average genetic distance per unit of physical distance in bovine genome is 1.25 cM/Mb (Arias et al., 2009). A period of more than 3,000 generations was used to ensure that the haplotype frequencies had reached their equilibrium values. At equilibrium, the distribution of LD stays constant over generations (i.e., Pt+1 = Pt). That distribution was used to describe, quantify, and compare rE2 at different equilibria in the absence or presence of mutation, selection or MAF filtering. Results from a population similar to the international black and white Holstein dairy cattle are presented in these cases, for which the effective population size is estimated to be about 50 (Kim and Kirkpatrick, 2009; Wray et al., 2019).

Furthermore, the exact distributions of rE2 derived under different scenarios were used to validate the approximate deterministic formulas from Sved and Feldman (1973) and Hill (1975). To generalize our results to larger effective population size, a non-linear regression model of the form 1a+b×Ne×c, following Sved's formula, was considered. The parameters in that model were estimated by non-linear least squares using the data generated from our transition-matrix approach.

The authors state that all data necessary for confirming the conclusions presented in the article are represented fully within the article.

3. Results

3.1. Comparison of Sved's and Hill's Approximations to the Exact Distributions of rE2 Derived From the Transition-Matrix Approach

In this section, Sved's and Hill's approximations were compared to the exact distribution of rE2 derived from our transition-matrix approach. The relationship between equilibrium value of r2 (rE2), recombination rate and mutation rate is shown in Figure 1 for these three approaches. In the absence of mutation and selection, Sved's formula showed consistency with the exact values from our transition-matrix approach (Figure 1A). On the other hand, Hill's formula had a better fit to the exact values of rE2 (Figure 1B) in the presence of mutation. However, neither Sved's nor Hill's deterministic formulae are accurate enough to describe rE2 in the presence of mutation. Therefore, we proposed an adjusted non-linear regression model to correct Sved's approximation to predict rE2 for larger effective population sizes in the presence of mutation. The mean square of errors is used to evaluate the fit of these three models and shows that our calibrated non-linear regression model was significantly better than Sved's or Hill's approximations.

FIGURE 1
www.frontiersin.org

Figure 1. Comparison of Sved's and Hill's approximation to exact distribution of rE2 (scatter points) derived from transition-matrix approach. Mean square errors are shown in the parentheses. “Calibrated” denotes the non-linear regression formula derived from the transition-matrix approach. (A) In the absence of mutation and selection. (B) In the presence of mutation but no selection.

3.2. Trajectory of LD Over the Generations Until Equilibrium Is Reached

The trajectory of evolving LD over generations computed from our transition-matrix approach is shown in this section, in the absence or presence of mutation and selection, where selection was applied with either a low selection coefficient of 0.01 or a high selection coefficient of 0.1. Starting with an allele frequency of 0.5 at each locus and linkage equilibrium between the two loci, expected values of LD [i.e., E(r2)] over generations are displayed in Figure 2. Note that the trajectory may be quite different, depending on the initial conditions used.

FIGURE 2
www.frontiersin.org

Figure 2. Expected value of LD E(r2) over generations under four different conditions at different recombination rates (c) for a population of effective size Ne = 50.

In the absence of mutation and selection, the expected value of r2 over generations increases to its maximum and reaches its “equilibrium” status. Value of rE2 increases as c decreases due to the decreasing breakdown of LD with lower recombination rates. Typically, when the recombination rate is relatively small (e.g., c = 6.25 × 10−5), the expected value of LD at “equilibrium” is almost equal to 1. However, at this “equilibrium” stage, frequencies of lines (Pt) keep changing though the distribution of allele frequencies remain unchanged. This is because forces of recombination and drift are balanced in lines with determinate r2, and frequencies of lines with determinate r2 almost proportionally decrease toward 0.

In the absence of selection but with mutation, the expected value of r2 increases and stays at an apparent equilibrium for several generations and then decreases to its true equilibrium value, where the mutation-drift equilibrium is reached. During the apparent equilibrium that is initially observed, the expected value of r2 is identical to the equilibrium value in the absence of mutation and selection, where the forces of drift and recombination are balanced. When this stage is first reached, the effect of mutation is negligible because the mutation rate is low relative to the frequency of lines with segregating loci. The apparent equilibrium ends when the frequency is high for lines where one or both loci are fixed (r2 is indeterminate) and the frequency of lines with segregating loci becomes close to the rate of mutation. Then, alleles introduced due to mutation into these lines have a noticeable effect on the distribution of allele frequencies and expected LD changes. At the end of the apparent equilibrium, the most frequent lines have only one haplotype. In the lines with two haplotypes, which have low frequencies, r2 is either indeterminate or has value 1.0. When mutation introduces a third haplotype in a line where r2 is 1.0, it drops in value. Further, when mutation introduces a third haplotype in a line where r2 is indeterminate, the value of r2 will be close to zero. Thus, as mutation becomes noticeable, the expected value of LD decreases. When the true equilibrium is reached, the loss of variation by fixation is balanced by its replenishment by mutation. In other words, the frequency of lines where loci are segregating will remain non-zero. Therefore, when mutation is present, the probabilities of allele frequencies (Pt) stay constant over generations when the true equilibrium is reached.

In the presence of both mutation and selection, lines with segregating loci become low in frequency at an earlier stage due to selection, particularly when the selection coefficient is large (e.g., s = 0.1), and thus, mutation starts to have a noticeable effect on allele frequency at an earlier stage in the presence of selection relative to when selection is absent. When selection coefficient is relatively large (e.g., s = 0.1), the expected value of r2 reaches its maximum value early and then decreases sharply to two plateaus. When the recombination rate is low, e.g., c ≤ 0.01, the difference between these two plateaus is small.

In the absence of selection, allele frequencies at two loci tend to be similar. However, when locus A is under selection, allele frequencies at these two loci tend to be different such that LD is lower than when selection is absent. This effect, however, is negligible when the selection coefficient is small (e.g., s = 0.01). Thus, a selection coefficient of 0.1 is used to present results when mutation and selection are present in the next section.

The expected values of r2 over generations in the presence of filtering by MAF (i.e., MAF ≥0.05 for locus B) are shown in Figure S2, where a similar pattern as for E(r2) over generations are observed.

3.3. Distributions of LD at Equilibrium

In a population where mutation is present and Ne = 50, the equilibrium value of r2 (rE2) is shown in Figure 3, when selection is absent or present, for different recombination rates. As explained in last section, the equilibrium values of LD in the absence of selection are always higher than those in the presence of selection. The equilibrium value of LD (i.e., rE2) with filtering by MAF (i.e., MAF ≥5% at locus B) is higher than its corresponding value in the absence of filtering.

FIGURE 3
www.frontiersin.org

Figure 3. Relationship between the expectation of LD at equilibrium (rE2) and recombination rate (c) in the absence or presence of selection or filtering by MAF at locus B for a population of effective size Ne = 50 with mutation rate u = 1.0 × 10−9. Only locus A is under selection with selection coefficient s = 0.1.

To understand why filtering by MAF increases rE2, the k lines with different haplotype frequencies were divided into two groups: lines where MAF at the locus B is <5% and lines where MAF ≥5%. Next, the transition-matrix approach was used to compute the frequency of each of these lines at equilibrium. In lines where r2 is defined, the equilibrium frequency of each line is plotted against its r2 value in Figure 4; the color blue is used for the first group of lines where MAF at the locus B is <5% and red is used for the second group of lines where MAF is ≥5%; frequencies in the absence of selection are given in Figure 4A and those in the presence of selection are given in Figure 4B.

FIGURE 4
www.frontiersin.org

Figure 4. Distribution of LD (r2) with recombination rate (c) of 6.25 × 10−5 at equilibrium in the presence of mutation but no selection or in the presence of mutation and selection in a population of effective size Ne = 50. Only locus A is under selection. (A) Presence of mutation but no selection (s = 0). (B) Presence of mutation and selection (s = 0.1).

In the absence of selection (Figure 4A), a proportion of about 0.41 of the lines were in the first group and a proportion of about 0.59 were in the second group. It can be seen from this figure that most of the r2 values in the first group were small, where around 45% of the lines had an r2 value < 0.001. In contrast, the second group had many lines with large r2 values; around 15% of the lines had r2 = 1. This difference in the distribution of r2 in the two groups shows why filtering by MAF at the B locus (removing lines from the first group, which had an abundance of low r2 values) would increase the expected value of r2.

The reason for the lower value of r2 in the first group is that, due to filtering by MAF at the B locus, this group has a large proportion of lines with recent mutations at the B locus. Consider a line where alleles are segregating at the A locus but fixed at the B locus. Such a line would belong to neither of the groups because r2 is not defined in a line where one locus is fixed. However, the introduction of a new allele at the B locus due to mutation will result in r2 becoming defined, but, typically, at a very low level because it results from the association in a single haplotype. The MAF in this line with the new mutation will be 12Ne, and it will belong to the first group provided that 12Ne<0.05.

Figure 4B gives the distribution of r2 for the two groups in the presence of selection, and it can be seen that still there is a greater abundance of low r2 values in the first group. The difference between the two groups, however, is smaller than in the absence of selection, and this explains the smaller effect of filtering on equilibrium value of r2 when selection is present (Figure 3).

3.4. Extrapolation of Exact rE2 to Larger Population Size by Non-linear Modeling

The two existing deterministic formulas (i.e., Sved's and Hill's formula) for rE2 are not accurate as shown in Figure 1. Thus, a non-linear regression formula by recombination rate was calibrated using the exact values of rE2 calculated from our transition-matrix approach. rE2 is computed as the expectation of r2 at equilibrium (i.e., mean r2 weighted by the frequency of each of the k possible lines at equilibrium). In the following, this formula is referred to as the calibrated non-linear regression formula. The form of the non-linear regression formula follows that of Sved's formula as presented below:

rE2=1/(β0+β1Nec).    (13)

To study the extrapolation of rE2 from small population sizes (i.e., Ne ≤ 50) to predict those for a larger population size, we split the values of rE2 calculated from our transition-matrix approach into training and validation sets. Exact values of rE2 with Ne ≤ 40 are included in the training set, and the remaining with Ne >40 are used for validation. The prediction accuracy is assessed using mean square error (MSE). As shown in Table S1, prediction accuracy from the calibrated non-linear regression formula is substantially higher than those from Sved's or Hill's formulas. Finally, parameters in Equation (13) are estimated using all available values of rE2, and parameter estimates are shown in Table 1. Typically, the estimates of β0 and β1 are inversely related to the values of recombination rates.

TABLE 1
www.frontiersin.org

Table 1. Calibration (estimation of βis) of the non-linear regression model under different recombination rates (c).

To further validate the accuracy of the extrapolation approach, the expected value of LD at equilibrium between adjacent markers in a 777k SNPs chip (i.e.,c = 6.25 × 10−5) was predicted for a Nellore cattle population, for which the effective population size is about 100 (Brito et al., 2013), using Equation (13) with β0 = 2.26 and β1 = 1557.11. In a real genotyped Nellore cattle population (Espigolan et al., 2013), the estimated mean and standard deviation for rE2 was 0.17 and 0.20, respectively. The predicted rE2 is 0.08, 0.449, and 0.976 using the calibrated non-linear regression formula, Hill's formula, and Sved's formula, respectively. Only the predicted rE2 from the calibrated non-linear regression formula falls in the range 0.17±0.2, while the other two are out of the measured range from this real population.

We further studied the accuracy of the extrapolation approach for very large effective population size, e.g., Ne = 10, 000 in human. However, rE2 did not align with the biological expectation, which indicates that our non-linear formulas for extrapolation of rE2 to extremely large population size does not work because the non-linear regression formulas are derived from data with Ne ≤ 50.

4. Discussion

A contribution of this article is to propose a computational transition-matrix approach to deriving the distribution of LD between two loci over generations in the presence of multiple genetic forces including drift, mutation, and selection. These distributions of LD are also studied in the presence of filtering by MAF (i.e., MAF ≥0.05). In addition to deriving exact distribution of LD, several critical results emerge from the proposed approach.

1. The expected value of LD at equilibrium decreases in the presence of selection. Decrease of LD caused by mutation is magnified in the presence of selection due to a higher fixation rate of the favorable allele. In the presence of mutation but no selection (or in the presence of mutation and weak selection), allele frequencies at both loci are similar due to genetic drift, and the LD between them tends to remain high. Conversely, in the presence of mutation and strong selection, this phenomenon is disrupted by the selection process resulting in diverse gene frequencies at the two loci,and low LD between them is observed.

2. Caution is needed when LD between a causal variant and a marker is inferred after filtering out marker loci with low MAF. LD between a causal variant and a marker variant with high MAF (i.e., MAF ≥5%) is higher than that between a causal variant and all marker variants, especially in the presence of mutation but no selection. That is, rE2 attributed to LD between a causal variant and marker variants with high MAF is relatively higher than that between a causal variant and marker variant with low MAF which leads to the reduction of the overall expected LD. In practice, LD is sometimes inferred from molecular marker panel with MAF cutoff applied. The inferred value is usually used to estimate important population parameters (e.g., effective population size). Here we have demonstrated the potential increase of LD brought by MAF cutoff and caution is needed when inferring LD in the presence of filtering by MAF.

3. “Fake” equilibrium may appear in the presence of mutation. Two equilibrium stages are observed in the presence of mutation. The first “equilibrium” indicates the balance between fixation increasing LD and recombination decreasing LD in lines with determinant r2 given the effect of mutation is negligible. When most loci become fixed at the end of the first “equilibrium” stage, change of allele frequency caused by mutation is of importance and reduces the expected LD. At the second equilibrium stage, the balance between mutation decreasing LD and fixation increasing LD is reached. Note that linkage equilibrium between two loci of allele frequency 0.5 is assumed in the initial population in Figure 2, and distribution of allele frequencies in the initial population affects the existence of the “fake” equilibrium.

4.1. Exact rE2 for Large Effective Population Sizes

The maximum effective population size (Ne) presented in this paper is 50 (Kim and Kirkpatrick, 2009; Wray et al., 2019), which is the estimated Ne for the international black and white Holstein dairy cattle population. Larger Ne was not studied in this paper due to computational limitations. When Ne = 50, the transition matrix A is square of order k=4Ne2+8Ne=176,851, which requires more than 250 gigabytes memory to store. In addition, it takes around 5 h to complete the computational process for the analysis of 3,500 generations. Note that the memory complexity to store A is O(Ne6). When Ne = 100, the estimated processing memory to store the transition matrix A is more than 14 terabytes. This computational problem may be addressed by parallel computing. This idea, however, needs further investigation and is out of the scope of this paper. Thus, to generalize the transition-matrix approach a non-linear regression model was calibrated using exact values of rE2. The proposed calibrations may provide a better description of the relationship between LD, effective population size, recombination rate, and mutation.

Data Availability Statement

The scripts used to generate the data can be found at https://github.com/Jiayi-Qu/Distribution_of_LD.

Author Contributions

HC and RF conceived the study. JQ, HC, and RF undertook the analysis and wrote the draft. DG and SK contributed to the analysis. All authors contributed to the final version of manuscript, read, and approved the final manuscript.

Funding

Financial support was provided by the United States Department of Agriculture, Agriculture and Food Research Initiative National Institute of Food and Agriculture Competitive grant no. 2018-67015-27957.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

RF is grateful to Prof. William G. Hill for extensive discussions of early results from the approach presented here. This manuscript has been released as a Pre-Print at https://www.biorxiv.org/content/10.1101/794347v1.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fgene.2020.00362/full#supplementary-material

References

Arias, J. A., Keehan, M., Fisher, P., Coppieters, W., and Spelman, R. (2009). A high density linkage map of the bovine genome. BMC Genet. 10:18. doi: 10.1186/1471-2156-10-18

PubMed Abstract | CrossRef Full Text | Google Scholar

Brito, F., Sargolzaei, M., Braccini Neto, J., Cobuci, J., Pimentel, C., Barcellos, J., et al. (2013). In-depth pedigree analysis in a large Brazilian nellore herd. Genet. Mol. Res. 12, 5758–5765. doi: 10.4238/2013.November.22.2

PubMed Abstract | CrossRef Full Text | Google Scholar

Espigolan, R., Baldi, F., Boligon, A. A., Souza, F. R., Gordo, D. G., Tonussi, R. L., et al. (2013). Study of whole genome linkage disequilibrium in nellore cattle. BMC Genomics 14:305. doi: 10.1186/1471-2164-14-305

PubMed Abstract | CrossRef Full Text | Google Scholar

Hill, W., and Robertson, A. (1968). Linkage disequilibrium in finite populations. Theor. Appl. Genet. 38, 226–231. doi: 10.1007/BF01245622

PubMed Abstract | CrossRef Full Text | Google Scholar

Hill, W. G. (1975). Linkage disequilibrium among multiple neutral alleles produced by mutation in finite population. Theor. Popul. Biol. 8, 117–126. doi: 10.1016/0040-5809(75)90028-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, E. S., and Kirkpatrick, B. W. (2009). Linkage disequilibrium in the North American Holstein population. Anim. Genet. 40, 279–288. doi: 10.1111/j.1365-2052.2008.01831.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Mitchell-Olds, T., Willis, J. H., and Goldstein, D. B. (2007). Which evolutionary processes influence natural genetic variation for phenotypic traits? Nat. Rev. Genet. 8, 845–856. doi: 10.1038/nrg2207

PubMed Abstract | CrossRef Full Text | Google Scholar

Ohta, T., and Kimura, M. (1969). Linkage disequilibrium at steady state determined by random genetic drift and recurrent mutation. Genetics 63, 229–238.

PubMed Abstract | Google Scholar

Ohta, T., and Kimura, M. (1971). Linkage disequilibrium between two segregating nucleotide sites under the steady flux of mutations in a finite population. Genetics 68, 571–580.

PubMed Abstract | Google Scholar

Sved, J. (1971). Linkage disequilibrium and homozygosity of chromosome segments in finite populations. Theor. Popul. Biol. 2, 125–141. doi: 10.1016/0040-5809(71)90011-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Sved, J., and Feldman, M. (1973). Correlation and probability methods for one and two loci. Theor. Popul. Biol. 4, 129–132. doi: 10.1016/0040-5809(73)90008-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Sved, J. A., and Hill, W. G. (2018). One hundred years of linkage disequilibrium. Genetics 209, 629–636.

PubMed Abstract | Google Scholar

Sved, J. A., McRae, A. F., and Visscher, P. M. (2008). Divergence between human populations estimated from linkage disequilibrium. Am. J. Hum. Genet. 83, 737–743. doi: 10.1016/j.ajhg.2008.10.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Walsh, B., and Lynch, M. (2018). Evolution and Selection of Quantitative Traits. Oxford: Oxford University Press.

Google Scholar

Wray, N. R., Kemper, K. E., Hayes, B. J., Goddard, M. E., and Visscher, P. M. (2019). Complex trait prediction from genome data: contrasting EBV in livestock to PRS in humans. Genetics 211, 1131–1141. doi: 10.1534/genetics.119.301859

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: linkage disequilibrium, effective population size, mutation rate, selection, minor allele frequency filtering

Citation: Qu J, Kachman SD, Garrick D, Fernando RL and Cheng H (2020) Exact Distribution of Linkage Disequilibrium in the Presence of Mutation, Selection, or Minor Allele Frequency Filtering. Front. Genet. 11:362. doi: 10.3389/fgene.2020.00362

Received: 08 November 2019; Accepted: 25 March 2020;
Published: 21 April 2020.

Edited by:

Jacob A. Tennessen, Harvard University, United States

Reviewed by:

Guanglin He, Sichuan University, China
Dan Skelly, Jackson Laboratory, United States

Copyright © 2020 Qu, Kachman, Garrick, Fernando and Cheng. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Hao Cheng, qtlcheng@ucdavis.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.