Skip to main content

ORIGINAL RESEARCH article

Front. Genet., 31 March 2015
Sec. Statistical Genetics and Methodology
This article is part of the Research Topic Genetic prediction View all 10 articles

A re-formulation of generalized linear mixed models to fit family data in genetic association studies

\r\nTao Wang*Tao Wang1*Peng He,Peng He1,2Kwang Woo AhnKwang Woo Ahn1Xujing WangXujing Wang3Soumitra GhoshSoumitra Ghosh4Purushottam LaudPurushottam Laud1
  • 1Division of Biostatistics, Medical College of Wisconsin, Milwaukee, WI, USA
  • 2Global Biostatistical Science, Amgen Inc., Thousand Oaks, CA, USA
  • 3Bioinformatics and Systems Biology Core, National Heart, Lung, and Blood Institute, Bethesda, MD, USA
  • 4Department of Genetics, Quantitative Sciences, GlaxoSmithKline, King of Prussia, PA, USA

The generalized linear mixed model (GLMM) is a useful tool for modeling genetic correlation among family data in genetic association studies. However, when dealing with families of varied sizes and diverse genetic relatedness, the GLMM has a special correlation structure which often makes it difficult to be specified using standard statistical software. In this study, we propose a Cholesky decomposition based re-formulation of the GLMM so that the re-formulated GLMM can be specified conveniently via “proc nlmixed” and “proc glimmix” in SAS, or OpenBUGS via R package BRugs. Performances of these procedures in fitting the re-formulated GLMM are examined through simulation studies. We also apply this re-formulated GLMM to analyze a real data set from Type 1 Diabetes Genetics Consortium (T1DGC).

1. Introduction

Generalized linear mixed model (GLMM) provides a rich class of statistical models to model correlated data with responses from the exponential family of distributions including Gaussian, Binomial, Poisson, etc. (see McCulloch and Searle, 2001). The mixed model approach, which is also called the variance component approach, has long been used in genetic studies to estimate genetic parameters, predict breeding values and model correlated family or pedigree data (Henderson, 1963, 1975; Amos, 1994; Falconer and Mackay, 1996; Almasy and Blangero, 1998; Abecasis et al., 2000; Sham and Purcell, 2001). Due to the diverse genetic correlation structures among families or pedigrees, currently fitting this type of GLMM often relies on special genetic software packages, such as SOLAR (Almasy and Blangero, 1998), Multic (de Andrade et al., 1998, 2006). More recently, as an extension of the R package lme4, Vazquez et al. (2010) also developed an R package pedigreemm following the method of Harville and Callanan (1989). Pedigreemm can handle the additive genetic correlation among all sampled individuals via a Cholesky decomposition of the coancestry (or so-called numerator relationship) matrix. However, despite the popularity of these software packages, they often lack many options such as choosing different algorithms for maximizing the likelihood or specifying various particular type of the covariance structures that are available in standard statistical software packages such as procedures “proc mixed,” “proc glimmix,” and “proc nlmixed” in SAS (SAS Institute, Inc.) or OpenBUGS in R (e.g., via R package BRugs). One major obstacle in using these standard software packages to fit the GLMM is the requirement of the same correlation structures across all the families (or clusters). A few recent studies also suggested fitting the GLMM by using SAS (Feng et al., 2009; Wang et al., 2011). But it appears that these studies have only considered the cases where all the families had the same genetic correlation structure.

The genetic correlation structure among family members for additive effects, dominance and epistasis has been well described (see Lynch and Walsh, 1998). In this study, for the special type of GLMM that models family data with varied family sizes or diverse genetic relatedness, we propose a Cholesky decomposition based re-formulation of the GLMM so that the re-formulated GLMM can be specified conveniently when using some standard statistical software packages. First, a standard GLMM is presented, which can account for the genetic correlation among family members. We briefly discuss the identifiability issue of the variance components from their possible confounding perspective in the GLMM. Next, we explain how the GLMM can be re-formulated into a GLMM with random regression coefficients of equal variances. Assuming that there is no genetic correlation between different families, we start by applying separate Cholesky decompositions on the genetic kinship matrices within each family. Then we stack these Chelosky decomposition matrices from different families into one column and treat the columns of the stacked matrix as fixed covariates. Unlike the regular Cholesky decomposition performed on the whole covariance matrix of the random genetic effects (e.g., Vazquez et al., 2010), here the stacked matrix has its column size being the maximum family size instead of the summation of all family sizes. With the reduced number of columns, this re-formulated GLMM can then be specified conveniently in “proc nlmixed” or “proc glimmix” using SAS, and OpenBUGS via R package BRugs. We provide detailed codes on fitting this re-formulated GLMM by using either “proc nlmixed” or “proc glimmix” in SAS, and OpenBUGS with R (via R package BRugs). The performances of these procedures on fitting the re-formulated GLMM are also examined through some simulation studies. Finally, we apply this re-formulated GLMM to a real data set from Type 1 Diabetes Genetics Consortium (T1DGC).

2. Methods

2.1. A Generalized Linear Mixed Model (GLMM)

Suppose that we have a randomly collected sample of N families from a study population. In the i-th family of size ni, let yij be a (binary or continuous) response variable for a disease phenotype; zij be some fixed environmental covariates that need to be adjusted for; gij denote the observed genotypes at certain targeted genetic marker loci, for family members j = 1, 2, ···, ni and i = 1, 2, ···, N. To model the family data and test for the association of gij with the phenotypic response yij, we need to account for the genetic correlation among family members induced by identity-by-descent (IBD) alleles shared by the family members at some putative disease susceptible loci (DSL). In addition, family members may share certain common environmental factors which could also contribute to the disease phenotypes. Let vij denote the random genetic effect from those putative (unobserved) DSL on the phenotypic response yij, and ei be the shared environmental effect such as diets for members in a family i. Define μij = E(yij|vij, ei). Then a generalized linear mixed model (GLMM) to model the family data can be written as

{yij|vij,ei~fyij|vij,ei(y|vij,ei), j=1,,niE(yij|vij,ei)=μijg(μij)=m+zijα+x(gij)β+vij+ei,  j=1,,nivi=(vi1,,vini)T~N(0,Σi)ei~N(0,σc2),eivi    (1)

where g(·) is a known link function, m is an intercept for the baseline, α is a p-dimensional vector of parameters for the fixed effects of environmental covariates zij, x(gij) is a q-dimensional vector with its components being defined by certain coding functions for marker genotypes (see Wang, 2011), and β is a q-dimensional vector of parameters for the fixed genetic effects contributed by the observed genotypes gij. Typically, the random effects {vi, i = 1, ···, n} from different families are assumed to be independent. Given the random effects vi and ei, yij, j = 1, ···, ni, are also assumed to be conditionally independent. As each vij represents an aggregated polygenic effect from multiple putative DSLs, vij tends to be normally distributed based on the Central Limit Theorem. Therefore, Σi denotes the genetic covariance matrix among the g-transformed conditional means gi1), ···, gini), which is often induced by identity-by-descent (IBD) alleles shared by the i-th family members at the putative DSL.

For a quantitative phenotypic trait, the link function g is often chosen as an identity function (i.e., g(x) = x). Assuming that there is no inbreeding between parents, it has been known that the genetic covariance between a pair of relatives j, k within a family i can be expressed as Kempthorne (1955), Amos (1994), Lynch and Walsh (1998), and Yu et al. (2006)

Cov(vij,vik)=2ϕjkσA2+δjkσD2    (2)

where σ2A and σ2D are the so-called additive and dominant genetic variance components, which are contributed by the additive allelic effects and allelic interactions from those unknown DSL, respectively; ϕjk and δjk are the so-called kinship and double coancestry coefficients between the two relatives. Similarly, for a general link function g, we can also define σ2A and σ2D as the additive and dominant variance components contributed by the allelic effects and allelic interactions from those unknown DSL to the variation of the g-transformed conditional means gij). Let μi = (μi1, ···, μini). From model Equations (1), (2), we have

Cov(g(μi))=2ΦiσA2+ΔiσD2+σc2Jni

where Jni is a ni × ni matrix of 1's, Φi = (ϕjk) and Δi = (δjk) are ni × ni kinship and double coancestry matrices, respectively, for the i-th family members. In the absence of inbreeding, the expected kinship and double coancestry coefficients for various common relatedness have been well established (e.g., see Table 7.1 in Lynch and Walsh, 1998). In practice, the actual kinship and double coancestry coefficients could deviate from their expected values because the realized IBD status could vary for a particular pair of family members due to the randomness in their parents' inheritance segregation. Two parents could also be related to each other due to possible inbreeding from their common ancestry. Therefore, the genetic covariance matrix Σi = 2Φiσ2A + Δi σ2D may have varied sizes across different families. For families with the same size and relatedness, their actual kinship and double coancestry matrices could also be different. Nowadays, with the high density of genome-wide genetic markers available, such as single nucleotide polymorphism (SNPs), it is possible to estimate the actual genome-wide kinship and double coancestry coefficients among family members using external programs such as PLINK (Purcell et al., 2007). The genome-wide kinship matrix can also be estimated as half of the genomic relationship matrix (see VanRaden, 2008).

The above GLMM provides a rich class of statistical models, which is applicable to both quantitative and qualitative traits. Depending on the family structures, however, not all the variance components in a GLMM are always estimable. For example, for a quantitative trait with g(·) being the identity link function, the GLMM becomes a linear mixed model (LMM).

yij=m+zijα+x(gij)β+vij+ei+ϵij,

where i = 1, …, N, j = 1, … ni, ϵij ~ N(0, σ2) are the model residuals, with σ2 being the residual variance of random effects from other risk factors not captured by zij, gij, vij, and ei. When each family comprises two parents and one offspring, the expected Δi = I3, which is a 3 × 3 identity matrix for any family i. We have

Cov(yi)=2ΦiσA2+(σD2+σ2)I3+σc2J3, i=1,,N

Thus, the dominant variance component σ2D and the residual variance σ2 are completely confounded. Similarly, if each family consists of only siblings (e.g., in a sib-pair design), then

V(yij)=σA2+σD2+σc2+σ2, i=1,,NCov(yij,yik)=12σA2+14σD2+σc2, jk

In this case, we cannot distinguish the four variance components from each other unless some of them are negligible (e.g., assuming σ2D = σ2c = 0). For a sample of unrelated individuals, it is also easy to see that all the four variance components σ2A, σ2D, σ2c, and σ2 are inseparable.

2.2. In Fitting the GLMM

The main goal in fitting GLMM Equation (1) is to make statistical inference on the fixed effects α, β as well as assessing the variance components σ2A, σ2D, and σ2c. Based on model Equation (1), the full likelihood is

L(α,β|{yij,zij,gij})=i=1Nvieifyi|vi,ei(yi|vi,ei)fvi(vi)                                                                                    fei(ei) deidvi

We need to calculate a multi-dimensional integration for each family, which in most cases cannot be analytically evaluated in closed forms. Various methods in fitting a standard GLMM have been proposed based on either numerical approximations to the integrations or a linearization of the regression model. Traditional methods for numerical integral approximation include Laplace approximation (Wolfinger, 1993), the adaptive Gauss-Hermite quadrature (Pinheiro and Bates, 1995), Monte Carlo integration, and Bayesian method via Markov chain Monte Carlo. The model linearization is often made via Taylor expansion on the inverse of the nonlinear link function g(·), based on which the pseudo-likelihood or restricted pseudo-likelihood for optimization can then be derived (Wolfinger and O'Connell, 1993). It has been known that the numerical integral approximations could become computationally intractable when a large number of random effects are involved. On the other hand, the model linearization approach could encounter severe unconvergence problems especially for binary outcomes with small cluster (family) sizes.

In practice, several common statistical software packages are available in fitting a standard GLMM. These include but not limited to “proc nlmixed” and “proc glimmix” procedures in SAS, and OpenBUGS for Bayesian approach. The “proc nlmixed” mainly conducts integral approximations using an adaptive Gauss-Hermite quadrature as default, and then directly maximizes the approximately integrated likelihood. In contrast, the “proc glimmix” primarily performs several model linearization based pseudo-likelihood methods, although it can also fit the GLMM using Laplace or adaptive Gauss-Hermite quadratures for integral approximations (see documentations supported by SAS Institute Inc, Raleigh, NC). In addition, the model fitting algorithm in R package “lme4” consists of an iteration between two sub-optimization procedures. One is to determine the conditional modes of the fixed and random effects, given the current deviance and variance components using penalized iteratively re-weighted least squares (PIRLS). The other is to obtain MLE of the deviance and variance components based on a profile likelihood from Laplace approximation, given the current conditional modes of the fixed and random effects (Bates et al., 2012). Nonetheless, these software packages typically require the random effects to have the same covariance (or correlation) structures across all the clusters. When families have different sizes or varied kinship or double coancestry matrices, it is difficult to directly specify GLMM Equation (1) using these software packages. It should be pointed out that the R package pedigreemm can fit a GLMM to family data with complex family structures without separating families into clusters. It also allows individuals from different families to be correlated. Here, we focus on fitting GLMM (1) using “proc nlmixed” and “proc glimmix” procedures in SAS, or OpenBUGS as an alternative choice.

To deal with different families sizes or varied kinship or double coancestry matrices, we apply a Cholesky decomposition based re-parameterization of the random genetic effects vi to re-formulate GLMM Equation (1) into a GLMM with random regression coefficients. First, we apply separate Cholesky decompositions on the kinship and double coancestry matrices: 2Φi = LΦiLTΦi, Δi = LΔiLTΔi, for i = 1, ···, N. We then re-parameterize the random genetic effects within family i as vi = LΦiai + LΔidi, where ai = (ai1, ···, airi)T ~ N(0, σ2AIri) with ri = ranki), and di = (di1, ···, disi)T ~ N(0, σ2DIsi) with si = ranki). Besides, we assume that ai's are independent of di's. By replacing vi by LΦiai + LΔidi in model Equation (1), we have

Cov(vi)=Cov(LΦiai+LΔidi)=2ΦiσA2+ΔiσD2=Σi

Note that the random effects ai and di are orthogonal within families as well as across families, even though the dimension of these random effects may still vary from family to family because the families may have different sizes.

Next, to deal with possible different family sizes, let r = maxi{ri} and s = maxi{si} be the maximum number of columns in {LΦi, i = 1, ···, N} and {LΔi, i = 1, ···, N}, respectively. For those families with ri < r (or si < s), we further expand their design matrices LΦi (or LΔi) to r (or s) columns by adding rri (or ssi) columns of 0's at the right end (or any places). Then, we obtain a re-formulated GLMM with all the families having the same dimension r (or s) for their random effects ai (or di) as the following.

g(μi)=1nim+Ziα+X(gi)β+LΦiai+LΔidi+1niei

where Zi = (zi1, ···, zini)T, X(gi) = (x(gi1), ···, x(gini))T, ai ~ N(0, σ2AIr), di ~ N(0, σ2DIs) and ei ~ N(0, σ2c). Finally, we construct the design matrix LΦ of ai from LΦi, i = 1, ···, N, by stacking them one above the other; and similarly build the design matrix LΔ of di by stacking LΔi, i = 1, ···, N, one above the other. Now, the columns of the two matrices LΦ and LΔ can be treated as (r + s) ordinary fixed continuous covariates with ai's and di's being their random regression coefficients. Within each family, all the family members share the same set of slopes. Across different families, the ai, i = 1, ···, N (or di, i = 1, ···, N), are independent but share the same variance component σ2A (or σ2D).

The re-formulated GLMM above can be easily specified by “proc nlmixed” or “proc glimmix” procedures in SAS. With “proc glimmix,” we can use three separate “random” commands “random e/subject = famid,” “random La1Lar/subject = famid type = TOEP(1);” and “random Ld1Lds/subject = famid type = TOEP(1);” to specify the correlation structures for the random effects ei, ai, and di, respectively. Here, La1,…, Lar represent the columns of the stacked matrix LΦ; Ld1Lds denote the columns of the stacked matrix LΔ. The option “TOEP(1)” can force all the elements in ai (or di) to share the same variance component σ2A (or σ2D). For “proc nlmixed,” currently it only allows to have one “random” command. But we can specify a joint multivariate normal distribution for ei, ai, and di via “random e a1ar d1ds ~ normal(mu, v) subject = famid,” where v is a diagonal matrix with one σ2c, r elements of σ2A's and s elements of σ2D's on its diagonal. As an example, the SAS codes for specification of a GLMM using both “proc glimmix” and “proc nlmixed” for families with two parents and two full sibs (i.e., r = s = 4) are provided in Appendices A, B (Supplementary Material), respectively. We also explored using the R package lme4 to fit the re-formulated GLMM. But it appears that the functions “lmer” and “glmer” provided in lme4 do not have an option that can force all the elements in ai (or di) to share the same variance component σ2A (or σ2D).

This re-formulation also makes it more convenient to fit GLMM Equation (1) using the Markov chain Monte Carlo (MCMC) based Bayesian approach. We use R package BRugs to get access to OpenBUGS software (Christensen et al., 2011), which has the Bayesian approach implemented. It is noticed that OpenBUGS has a weak support for matrix operations. In the original GLMM Equation (1), the genetic covariance matrix Σi = iσ2A + Δi σ2D involves two unknown variance components σ2A and σ2D. As a result, we cannot directly specify the covariance matrix Σi in OpenBUGS. With the re-formulated GLMM, we can pass LΦ and LΔ as fixed covariates to OpenBUGS with ai and di being their random regression coefficients. Since the Bayesian approach often treats all the model parameters as random, it appears especially suitable for fitting the re-formulated GLMM. We can also extract from the MCMC the posterior distributions of random effects for each individual, which allow us to assess the variation contribution from the putative random genetic effects {vij} to the total variance of {gij)}. The R codes for specification of a re-formulated GLMM using BRugs + OpenBUGS are also provided in Appendix C (Supplementary Material).

3. Simulation Study

In this section, we examine the performances of procedures “proc nlmixed” and “proc glimmix” in SAS (version 9.3) as well as R packages BRugs + OpenBUGS in fitting the re-formulated GLMM through simulation. We consider three biallelic genetic markers with alleles “0” or “1,” and one fixed explanatory covariate Z ~ Bernoulli(0.5). The three genetic markers are assumed to be unlinked (i.e., independent) and have allele frequencies p1 = 0.5, p2 = 0.2, p3 = 0.1 for alleles “1” at locus 1, 2, and 3, respectively. We first generate a pairs of haplotypes independently for each parent, where each haplotype is comprised of three alleles randomly generated from Bernoulli(pj) for j = 1, 2, and 3, respectively. Each child inherits one haplotype from father, and the other from mother. The haplotype from father (or mother) consists of three alleles with each allele being selected from the two paternal (or maternal) alleles at the same locus with 50% chance. In case where a family has more than one child, the above random process is repeated for each child independently. For simplicity, we use the expected values to construct the kinship and double coancestry matrices for each family.

We first consider simulating quantitative traits from a LMM. Let gij = (gij1, gij2, gij3) be the genotypes of the j-th subject in family i, where gijk ϵ {0, 1, 2} counts the number of alleles “1” at locus k = 1, 2, 3. The quantitative trait values are simulated from the following LMM.

yij=m+zijα+x(gij1)β1+x(gij2)β2+x(gij3)      β3+vij+ei+ϵij,i=1,,n, j=1,ni

where x(gijk) = (xa(gijk), xd(gijk))βk with xa(gijk) = gijk, xd(gijk) = 1 (or 0) if gijk = 2 (or otherwise) based on the allele-coding (see Wang, 2011), and βk = (βka, βkd)T being the effects of marker k for k = 1, 2, 3. We further set σ2c = σ2 = σ2A = 1 and σ2D = 0.5. The other true values of parameters are listed in Table 1. Each simulation data set contains n1 families with two parents and one child and n2 families with two parents and two children. We consider three cases: (a) n1 = 500, n2 = 500; (b) n1 = 1000, n2 = 1000; (c) n1 = 0, n2 = 2000.

TABLE 1
www.frontiersin.org

Table 1. The true values of model parameters in simulation.

For each simulation data set, we fit the GLMM by using two methods: (1) adaptive Gaussian quadrature (AGQ) via “proc nlmixed” in SAS; (2) Bayesian approach via BRugs + OpenBUGS. Based on 200 simulation data sets, the parameter estimates are summarized in Table 3. We also explored using “proc glimmix” to fit the simulation data but abandoned it due to some severe unconvergence problems. In running BRugs + OpenBUGS, we choose the following priors for initialization of the parameters: m ~ N(0, 10), α ~ N(0, 10), βka ~ N(0, 10) and βkd ~ N(0, 10) for k = 1, 2, 3, τa ~ Gamma(1, 1), σD ~ Uniform(0, 5), τc ~ Gamma(1, 1), and τ ~ Gamma(1, 1), where τa = 1/σ2A, τd = σ2D, τc = 1/σ2c, and τ = 1/σ2. We use the first 10,000 updates as burn-in, and another 10,000 updates to estimate parameters as posterior means. The parameter estimates and their standard deviations (SD) from the 200 simulations are summarized in Table 2. For each simulation data, we also calculate the length of the 95% confidence (or probability) interval for each parameter in running “proc nlmixed” (or BRugs + OpenBUGS). The average length of these intervals and their coverage rate for each true parameter value from the 200 simulations are summarized in Table 3.

TABLE 2
www.frontiersin.org

Table 2. Mean and standard deviation (SD) of the parameter estimates from 200 simulations for linear mixed models.

TABLE 3
www.frontiersin.org

Table 3. The average length of the 95% confidence (or probability) intervals and the coverage rate for true parameters from 200 simulations for linear mixed models.

From Table 2 we can see that both methods can provide reasonable estimates of the fixed effects α, βka, and βkd (k = 1, 2, 3) as well as the variance components with improved accuracy as sample size increases. As expected, the estimate of allelic interaction βkd has a larger SD than that of the additive allelic effect βka at each locus k, for k = 1, 2, 3. The estimates of σ2D and σ2 are slightly biased, which could be caused by the potential confounding between these two variance components. Besides, it appears that “proc nlmixed” tends to over-estimate the dominant variance σ2D and under-estimate the residual variance σ2. Meanwhile, the Bayesian method heads to the opposite way. From Table 3 the coverage probabilities are close to the nominal level of 95% for most parameters except σ2D and σ2. The average lengths of the 95% confidence (or probability) intervals for σ2D and σ2 are also greater than the ones for other two variance components. In addition, the average length of the 95% confidence (or probability) intervals for β3d is substantially larger than that for β1d and β2d, which is likely caused by the fact that the allele “1” at marker locus 3 is rare and the homozygous genotypes “11” at marker locus 3 are much less present in a simulation data set than those at the other two loci.

Overall, the results from “proc nlmixed” and BRugs + OpenBUGS are quite comparable in all three cases. As we have mentioned before, the variance components σ2 and σ2D are confounded in single-child families, and only two-child families are informative for distinguishing them. In case (b), where there is an increased number of 2-child families, the estimates of these two variance components are improved with both methods. In case (c), the accuracy in estimates of these two variance components is further improved. In terms of the computational speed, it appears that the “proc nlmixed” runs much faster than BRugs + OpenBUGS. We ran BRugs + OpenBUGS on one laptop installed with Intel(R) Core(TM)i7-3520M CPU @ 2.90GHz. It took about 19 h, 50 h and 38 min, and 59 h and 15 min of CPU time to complete the 200 simulations (including data generation and MCMC iterations) in cases (a), (b), and (c), respectively. The “proc nlmixed” procedure in SAS was performed on a UNIX workstation which has a compatible speed with the laptop, and it took about 2 and a half hours, 6 and a half hours, and 7 h 18 min of CPU time to complete the 200 simulations in cases (a), (b), and (c), respectively. By contrast, the Bayesian approach can provide the posterior distributions for all the parameters rather than just the modes (MLE) and their variance or covariance estimates. When σ2 and σ2D are almost completely confounded, we found that the “proc nlmixed” may give unreliable estimates of the model parameters or encounter unconvergence problem caused by the nearly singular Hessian matrix, while the Bayesian approach can still provide reasonable estimates at least for other model parameters except σ2 and σ2D.

We also consider binary traits and simulate phenotypic values from the following mixed logistic regression model.

logitP(yij=1|zij,gij1,gij2,gij3,vij,ei)=m+zijα       +x(gij1)β1+x(gij2)β2+x(gij3)β3+vij+ei,

for i = 1, …, n and j = 1, …, ni. In order to have enough number of events, we choose m = −3. Meanwhile, we keep the same true values as before for other model parameters. We explored various options on using “proc nlmixed” and “proc glimmix” but failed to fit the above model appropriately due to severe unconvergence problems. By contrast, BRugs + OpenBUGS can still provide reasonable estimates for most of the model parameters. Using the similar priors in the previous setting, 10,000 burn-in and 10,000 updates for parameter estimation, we compute the means and standard deviations (SD) of the parameter estimates and the average lengths (AL) and coverage rates (CR) of the 95% probability intervals from 200 simulations as summarized in Table 4.

TABLE 4
www.frontiersin.org

Table 4. Means (SD) of the parameter estimates and AL (CR) of the 95% probability intervals from 200 simulations for mixed logistic regression models.

Under our simulation setting, it appears that the SAS procedures “proc nlmixed” and “proc glimmix” perform unexpectedly poorly especially for binary outcomes. But we have to admit that, although we have explored many options that are available in running “proc nlmixed” and “proc glimmix,” our exploration is surely not exclusive based on our limited knowledge. Besides, we use SAS version 9.3 on our Linux computer system. Recently, a newer version 9.4 of SAS has become available for PC users, which may provide improved performance in fitting the GLMM. Unfortunately, we do not have access to this new version of SAS yet based on our current Linux computer system. The performance of “proc nlmixed” and “proc glimmix” in SAS 9.4 on fitting the re-formulated GLMM needs further exploration.

While the frequencist approach implemented in “proc nlmixed” and “proc glimmix” often require numerical approximations to the full likelihood, the Bayesian approach directly maximizes the full likelihood via MCMC. From our simulation study, it seems that the Bayesian approach can better handle the re-formulated GLMM especially for binary traits. However, it should be pointed out that the choice of priors in using the Bayesian approach could have a significant impact on convergence of an MCMC procedure. Besides the gamma priors for all the τ's, we also tested using uniform priors on σ's and obtained similar results. Throughout our simulation, the Monte-Carlo errors for all the parameter estimates appear to be acceptable. But some auto-correlations in certain Markov chains (e.g., the ones for σ2D and σ2) are noticed. Typically, the auto-correlation could be reduced by thinning the Markov chains. Otherwise, appropriate adjustment is needed in computing the SD of parameter estimates.

4. Analysis of T1DGC data for Type I Diabetes

As an example, we consider fitting a real family data set obtained from the Type 1 Diabetes Genetics Consortium (T1DGC). The data set includes five cohorts: Asia-Pacific (AP), Danish Steno Diabetes Center (DAN), European (EUR), Sardinian (SAR) and United Kingdom Genetic Resource Investigating Diabete (UK). For simplification, we only adopt nuclear families and exclude some grand-parents or grand children (less than 1% of the total subjects). Most of the families (about 80.4%) consist of 2 parents and 2–3 children. There are 13 families that have more than 9 family members, and they all belong to the DAN cohort. The actual numbers of subjects and families we used in the five cohorts are listed in Table 5.

TABLE 5
www.frontiersin.org

Table 5. Number of subjects and families in T1DGC by cohorts.

Our research interest is to test for association of HLA-DQB1 locus with Type 1 Diabetes (T1D) incidence, while appropriately controlling for other potential genetic risk factors on the incidence of T1D. The adjustment for random additive and dominance effects is important because it has been known that T1D is a polygenic disease. Some studies have suggested that other genes such as INS and CTLA4 could be implicated with T1D (Anjos and Polychronakos, 2004; McGinnis et al., 2009). One article “Genetics and Diabetes” from the World Health Organization (WHO) web site “http://www.who.int/genomics/about/Diabetis-fin.pdf” also provides a nice review of the T1D. From our previous study (Glisic et al., 2009), we classify the subjects into 4 groups: low risk (DQrisk = 0), moderate risk (DQrisk = 1), high risk (DQrisk = 2), and very high risk (DQrisk = 3) based on the HLA-DQB1 genotypes and CD4 + CD25 + highT-cell apoptosis. Gender and age are also known risk factors for T1D. We categorize age into 6 categories: age≤18, 18<age≤30, 30<age≤40, 40<age≤50, 50<age≤60, and age>60. By choosing the “high risk” (the largest group across all cohorts) at HLA-DQB1, male and “age≤18” as a baseline, we fit each cohort separately using the following mixed logistic regression model.

logitP(yij=1|vij,ei)=μ+α11(18<age30)+α21(30<age40)+α31(40<age50)+α41(50<age60)+α51(age>60)+α61(sex=F)+β11(DQrisk=0)+β21(DQrisk=1)+β31(DQrisk=3)+vij+ei

where ei ~ N(0, σ2c), eαi (i = 1, ···, 5) are odds ratios of having T1D in different age groups comparing with the youngest age group of age ≤18, eβj (j = 1, 2, 3) are odds ratios of having T1D in different HLA-DQB1 risk groups comparing with the high risk group, eβ6 is the odds ratio of T1D in females versus males, and {vij} have the covariance structures as specified in GLMM Equation (1).

To simplify the calculation, we use the expected values to construct the kinship and double coancestry matrices for each family. As the actual kinship and double coancestry matrices should be close to their expected matrices, we would expect only minor deviations from the fitted GLMM. In running “proc glimmix” and “proc nlmixed” to fit the GLMMs, we encountered some severe unconvergence problems (data not shown). In running BRugs + OpenBUGS, we use 10,000 burn-in and another 10,000 updates to estimate the parameters. For the DAN cohort, the current BRugs + OpenBUGS cannot accommodate more than 12 additive or dominant random coefficients in the model specification—see part (2) of Appendix C in Supplementary Material, where “maxsize” (i.e., the number of columns of the stacked matrices LΦ and LΔ) cannot exceed 12 even though the total number of family members can still exceed 12. We also find that some of the parameter estimates become unstable when we actually use 10–12 columns. So we adopt using 8 columns of the stacked matrices LΦ and LΔ in fitting the Dan cohort. The estimates of odds ratios and variance components in the fitted models from running BRugs + OpenBUGS are summarized in Table 6.

TABLE 6
www.frontiersin.org

Table 6. Posterior means and 2.5%, 97.5% percentiles of the odds ratios and variance components for type I diabetes in five cohorts of the T1DGC data set.

The results have confirmed that age is significantly associated with T1D incidence. In most cohorts except AP, the odds of T1D occurrence reaches the highest in the youngest group of age < 18 and then decreases quickly as age increases. The females appear to have a less chance of having T1D than males in DAN and EUR cohorts. For HLA-DQB1, it appears that in each cohort there is a significant increase of T1D risk in the HLA-DQB1 very high risk group comparing with the high risk group, and the high risk group is also significantly different from the low risk and moderate risk groups after adjusting for the age and gender effects. The odds of T1D for the high or very high risk groups appear significantly higher than that for the low or moderate risk groups, which are consistent with the relative risk of 3–45 of the HLA-DQB1 susceptibility variant reported in “Genetics and Diabetes” from WHO.

Regarding the variance components, the estimates of the additive and family shared variances appear reasonably well, although the dominance variances in SAR and AP cohorts have relatively large variation due to perhaps the lack of information in their family data. To see whether we should not include σ2D in the models, we also fit the GLMM models without σ2D and compare them with our previous models. Based on the deviance information criterion (DIC), which is an estimate of the expected predictive error (lower deviance is better) and it can account for both model fitness and model complexity, the models with both additive and dominance variances included have their DIC values of 658.6, 647.3, 1525, 342.2, and 316.6 which are lower than the DIC values of 669.5, 658.9. 1543. 354.1, and 322.3 in the reduced models without σ2D in AP, DAN, EUR, SAR, and UK cohorts, respectively. So the models with both additive and dominance variances included are preferable. We also estimate the variations contributed by the random putative genetic effects {vij}, which account for 22%, 26%, 9%, 27%, and 12% of the total variation in logitP(yij = 1|vij, ei) for the AP, DAN, EUR, SAR, and UK cohorts, respectively.

The T1DGC family data was collected from the observational retrospective sampling, which likely had over-sampled families with T1D children. It is well known that fitting a mixed logistic regression model with random effects to a retrospective data set may no longer provide the equivalent maximum likelihood estimates of the model parameters as the ones defined in the same model for a prospective cohort. Therefore, the results above are only applicable to the families we obtained from T1DGC. For general populations, an adjustment for the retrospective sampling strategy is needed in order to avoid the bias in the parameter estimates.

5. Discussion

In this study, we propose a Cholesky decomposition based re-formulation of the GLMM to fit family data with varied sizes and diverse genetic relatedness. Assuming that there is no genetic correlation between different families, we first apply separate Cholesky decompositions on the genetic kinship (or double co-ancestry) matrices within each family. Next, we stack these Chelosky decomposition matrices from different families into columns to form two stacked matrices. The columns of the stacked matrices can then be treated as fixed covariates with random regression coefficients of equal variances. It should be pointed out that applying separate Cholesky decompositions on each family does not provide computational or storage benefits comparing with the regular Cholesky decomposition on the whole covariance matrix of the random genetic effects because the non-diagonal blocks in the sparse matrix decomposition are actually not saved. However, with the reduced number of columns in our stacked matrices, this re-formulated GLMM can be specified more conveniently using “proc nlmixed” and “proc glimmix” in SAS, or OpenBUGS via R package BRugs. Theoretically, the re-formulated GLMM is equivalent to the original GLMM. Therefore, it should retain the same validity and power in testing the fixed and random genetic effects as the original GLMM. From our simulation, it appears that this re-formulated GLMM can be fitted reasonably well by “proc nlmixed” and BRugs + OpenBUGS for quantitative traits with moderate family sizes. For binary traits, our simulation and real data example shows that at least BRugs + OpenBUGS can appropriately fit the re-formulated GLMM for families of sizes not exceeding 12.

This Cholesky decomposition based re-formulation of GLMM in fitting family data is somewhat analogous to Haseman and Elston's regression method for sib-pairs (Haseman and Elston, 1972). While Haseman and Elston's algorithm regresses the squared sib-pair's phenotypic difference on the kinship and double coancestry coefficients at a targeted locus with fixed effects, our Cholesky decomposition based GLMM regresses all family members phenotypic values on the square roots of the kinship and double coancestry coefficient matrices with random regression coefficients across families.

Except SOLAR, Multic, and pedigreemm, many other special software packages are currently available for analyzing pedigree data. For examples, VITESSE implemented the well-known Elston–Stewart's peeling algorithm for computing the likelihood of pedigrees in linkage analysis (Elston and Stewart, 1971; O'Connell and Weeks, 1995). Mendel can run pedigree analysis for quantitative traits (Lange et al., 2013). SAGE SAGE (2012) can be used for pedigree analysis of binary traits. Based on the score statistics and generalized estimating equations, FBAT and its extension PBAT can also be used in association testing for both quantitative and binary traits (Rabinowitz and Laird, 2000; Lange and Laird, 2002; Lange et al., 2004; Laird and Lange, 2006). Nevertheless, our study provides an alternative choice for analyzing the pedigree data by using standard statistical software, which could be useful for statisticians who are not very familiar with these special genetic software packages. The standard statistical software packages often provide many options on choosing different optimization techniques to maximize the likelihood. As long as the optimization procedure converges appropriately, the standard statistical software packages can provide reliable results in most of the cases.

Sometimes we may want to perform hypothesis tests on the existence of certain variance components. For non-Bayesian methods in fitting the GLMM, it has been known that the likelihood ratio statistics (LRS) usually do not asymptotically follow the standard Chi-square distributions under the null because the zeros under the null hypothesis are located on the boundary of the parameter space for the variance components, where the standard regularity conditions no longer hold. As pointed out in Wang et al. (2011), in the hypothesis testing of a single variance component H0: σ2A = 0, the asymptotic distribution of LRS has 0.5χ20 + 0.5χ21 under the null. In the hypothesis testing of more than one variance components such as H0: σ2A = σ2D = 0, the asymptotic distribution of LRS could be a mixture of several chi-square distributions with their weights of the mixture depending on the number of family types. Therefore, directly using LRS to test for the existence of variance components could be incorrect. On the other hand, the Bayesian method can always provide appropriate estimates of the variance components as well as their variances without relying on the asymptotic results.

It should be pointed out that this Cholesky decomposition based re-formulation in fitting the GLMM has some limitations. For example, in order to synchronize the varied family sizes, the number of random effects in smaller families needs to be expanded to match that in the largest family. The Cholesky decomposition also requires that the kinship and double coancestry matrices be positive. When the kinship matrix is calculated from genome-wide genotypes, the kinship or double coancestry matrices could become singular for some of the families. One possible solution to this problem is to apply a different type of decomposition to the kinship and double coancestry matrices for these families. Note that the decomposition matrices can have a reduced number of columns as long as a good approximation to the kinship and double coancestry matrices is maintained. Finally, it appears that the computational speed in fitting the re-formulated GLMM via SAS or OpenBUGs is slow. The proposed GLMM re-formulation is probably more suitable for a refined analysis on certain targeted loci rather than a genome-wide scan for a large number of genetic markers.

Conflict of Interest Statement

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

This work is partially supported by the National Institute of Diabetes and Digestive and Kidney Diseases under grant number R01 DK080100. This research also uses the data resource provided by T1DGC, a collaborative clinical study sponsored by the National Institute of Diabetes and Digestive and Kidney Diseases, National Institute of Allergy and Infectious Diseases, National Human Genome Research Institute, National Institute of Child Health and Human Development, and Juvenile Diabetes Research Foundation International (JDRF), which is supported by U01 DK-062418.

Supplementary Material

The Supplementary Material for this article can be found online at: http://www.frontiersin.org/journal/10.3389/fgene.2015.00120/abstract

References

Abecasis, G. R., Cardon, L. R., and Cookson, W. O. (2000). A general test of association for quantitative traits in nuclear families. Am. J. Hum. Genet. 66, 279–292. doi: 10.1086/302698

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Almasy, L., and Blangero, J. (1998). Multipoint quantitative trait linkage analysis in general pedigrees. Am. J. Hum. Genet. 62, 1198–1211. doi: 10.1086/301844

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Amos, C. (1994). Robust variance-components approach for assessing genetic linkage in pedigrees. Am. J. Hum. Genet. 54, 535–543.

PubMed Abstract | Full Text | Google Scholar

Anjos, S., and Polychronakos, C. (2004). Mechanisms of genetic susceptibility to type i diabetes: beyond HLA. Mol. Genet. Metab. 81, 187–195. doi: 10.1016/j.ymgme.2003.11.010

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Bates, D., Maechler, M., and Bolker, B. (2012). Linear Mixed-Effects Models Using S4 Classes. lme4 Version 0.999999-0.

Christensen, R., Johnson, W., Branscum, A., and Hanson, T. E. (2011). Bayesian Ideas and Data Analysis. Boca Raton, FL: CRC Press; Taylor and Francis Group, LL.

Google Scholar

de Andrade, M., Amos, C. I., and Foulkes, W. D. (1998). Segregation analysis of squamous cell carcinoma of the head and neck: evidence for a major gene determining risk. Ann. Hum. Genet. 62(Pt 6), 505–510. doi: 10.1017/S0003480099007204

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

de Andrade, M., Atkinson, E. J., Lunde, E., Amos, C. I., and Chen, J. F. (2006). Estimating Genetic Components of Variance for Quantitative Traits in Family Studies Using the MULTIC. Technical Report 78, Mayo Foundation.

Elston, R. C., and Stewart, J. (1971). A general model for the genetic analysis of pedigree data. Hum. Hered. 21, 523–542. doi: 10.1159/000152448

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Falconer, D. S., and Mackay, T. F. C. (1996). Introduction to Quantitative Genetics, 4th Edn. Harlow: Longman.

Feng, R., Zhou, G., Zhang, M., and Zhang, H. (2009). Analysis of twin data using SAS. Biometrics 65, 584–589. doi: 10.1111/j.1541-0420.2008.01098.x

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Glisic, S., Klinker, M., Waukau, J., Jailwala, P., Jana, S., Basken, J., et al. (2009). Genetic association of hla dqb1 with cd4+cd25+(high) t-cell apoptosis in type 1 diabetes. Genes. Immun. 10, 334–340. doi: 10.1038/gene.2009.14

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Harville, D., and Callanan, T. (1989). “Computational aspects of likelihood-based inference for variance component,” in Advances in Statistical Methods for Genetic Improvement of Livestock, eds D. Gianola and K. Hammond (Berlin, Germany: Springer-Verlag), 136–176. doi: 10.1007/BF01066731

CrossRef Full Text

Haseman, J. K., and Elston, R. C. (1972). The investigation of linkage between a quantitative trait and a marker locus. Behav. Genet. 2, 3–19. doi: 10.1007/BF01066731

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Henderson, C. R. (1963). “Selection index and expected genetic advance,” in Statistical Genetics and Plant Breeding, eds W. D. Hanson and H. F. Robison (Washington, DC: National Academy of Sciences), 141–163.

Google Scholar

Henderson, C. R. (1975). Best linear unbiased estimation and prediction under a selection model. Biometrics 31, 423–447.

PubMed Abstract | Full Text | Google Scholar

Kempthorne, O. (1955). The theoretical values of correlations between relatives in random mating populations. Genetics 40, 153–167.

PubMed Abstract | Full Text | Google Scholar

Laird, N. M., and Lange, C. (2006). Family-based designs in the age of large-scale gene-association studies. Nat. Rev. Genet. 7, 385–394. doi: 10.1038/nrg1839

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Lange, C., and Laird, N. M. (2002). Power calculations for a general class of family-based association tests: dichotomous traits. Am. J. Hum. Genet. 71, 575–584. doi: 10.1086/342406

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Lange, C., Blacker, D., and Laird, N. M. (2004). Family-based association tests for survival and times-to-onset analysis. Stat. Med. 23, 179–189. doi: 10.1002/sim.1707

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Lange, K., Papp, J. C., Sinsheimer, J. S., Sripracha, R., Zhou, H., and Sobel, E. M. (2013). Mendel: the swiss army knife of genetic analysis programs. Bioinformatics 29, 1568–1570. doi: 10.1093/bioinformatics/btt187

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Lynch, M., and Walsh, B. (1998). Genetics and Analysis of Quantitative Traits. Sunderland, MA: Sinauer.

Google Scholar

McCulloch, C., and Searle, S. (2001). Generalized, Linear, and Mixed Models. New York, NY: Wiley & Sons, INC.

Google Scholar

McGinnis, R., McLaren, W., Ranganath, V., Whittaker, P., Hunt, S., Deloukas, P., et al. (2009). Haplotype-based search for snps associated with differential type 1 diabetes risk among chromosomes carrying a specific hla drb1-dqa1-dqb1 haplotype. Diabetes Obes. Metab. 11 Suppl. 1, 8–16. doi: 10.1111/j.1463-1326.2008.00998.x

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

O'Connell, J. R., and Weeks, D. E. (1995). The vitesse algorithm for rapid exact multilocus linkage analysis via genotype set-recoding and fuzzy inheritance. Nat. Genet. 11, 402–408. doi: 10.1038/ng1295-402

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Pinheiro, J., and Bates, D. (1995). Approximations to the log-likelihood function in the nonlinear mixed-effects model. J. Comput. Graph. Stat. 4, 12–35.

PubMed Abstract | Full Text | Google Scholar

Purcell, S., Neale, B., Todd-Brown, K., Thomas, L., Ferreira, M. A. R., Bender, D., et al. (2007). Plink: a tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 81, 559–575. doi: 10.1086/519795

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Rabinowitz, D., and Laird, N. (2000). A unified approach to adjusting association tests for population admixture with arbitrary pedigree structure and arbitrary missing marker information. Hum. Hered. 50, 211–223. doi: 10.1159/000022918

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

SAGE (2012). Statistical Analysis for Genetic Epidemiology, Release 6.3. Available online at: http://darwin.cwru.edu.

Sham, P. C., and Purcell, S. (2001). Equivalence between haseman-elston and variance-components linkage analyses for sib pairs. Am. J. Hum. Genet. 68, 1527–1532. doi: 10.1086/320593

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

VanRaden, P. M. (2008). Efficient methods to compute genomic predictions. J. Dairy Sci. 91, 4414–4423. doi: 10.3168/jds.2007-0980

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Vazquez, A., Bates, D., Rosa, G., Gianola, D., and Weigel, K. (2010). Technical note: an r package for fitting generalized linear mixed models in animal breeding. J. Anim. Sci. 88, 497–504. doi: 10.2527/jas.2009-1952

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Wang, X., Guo, X., He, M., and Zhang, H. (2011). Statistical inference in mixed models and analysis of twin and family data. Biometrics 67, 987–995. doi: 10.1111/j.1541-0420.2010.01548.x

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Wang, T. (2011). On coding genotypes for genetic markers with multiple alleles in genetic association study of quantitative traits. BMC Genet. 12:82. doi: 10.1186/1471-2156-12-82

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Wolfinger, R., and O'Connell, M. (1993). Generalized linear mixed models: a pseudo-likelihood approach. J. Stat. Comput. Simul. 48, 233–243. doi: 10.1080/00949659308811554

CrossRef Full Text | Google Scholar

Wolfinger, R. (1993). Laplace approximation for nonlinear mixed models. Biometrika 80, 791–795. doi: 10.1093/biomet/80.4.791

CrossRef Full Text | Google Scholar

Yu, J., Pressoir, G., Briggs, W. H., Vroh Bi, I., Yamasaki, M., Doebley, J. F., et al. (2006). A unified mixed-model method for association mapping that accounts for multiple levels of relatedness. Nat. Genet. 38, 203–208. doi: 10.1038/ng1702

PubMed Abstract | Full Text | CrossRef Full Text | Google Scholar

Keywords: family data, generalized linear mixed models (GLMM), genetic correlation, genetic variance components, random genetic effects, re-parameterization, Cholesky decomposition, Bayesian methods

Citation: Wang T, He P, Ahn KW, Wang X, Ghosh S and Laud P (2015) A re-formulation of generalized linear mixed models to fit family data in genetic association studies. Front. Genet. 6:120. doi: 10.3389/fgene.2015.00120

Received: 18 November 2014; Accepted: 12 March 2015;
Published: 31 March 2015.

Edited by:

Eduardo Manfredi, Institut National de la Recherche Agronomique, France

Reviewed by:

Solomon K. Musani, Egerton University, Kenya
Ana I. Vazquez, University of Alabama at Birmingham, USA
Dalin Li, Cedars-Sinai Medical Center, USA

Copyright © 2015 Wang, He, Ahn, Wang, Ghosh and Laud. 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) or licensor 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: Tao Wang, Division of Biostatistics, Institute for Health and Society, Medical College of Wisconsin, 8701 Watertown Plank Road, Milwaukee, WI 53226-0509, USA taowang@mcw.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.