[フレーム] [フレーム]
  • Loading metrics

Open Access

Peer-reviewed

Research Article

HMFGraph: Novel Bayesian approach for recovering biological networks

Abstract

Gaussian graphical models (GGM) are powerful tools to examine partial correlation structures in high-dimensional omics datasets. Partial correlation networks can explain complex relationships between genes or other biological variables. Bayesian implementations of GGMs have recently received more attention. Usually, the most demanding parts of GGM implementations are: (i) hyperparameter tuning, (ii) edge selection, (iii) scalability for large datasets, and (iv) the prior choice for Bayesian GGM.

To address these limitations, we introduce a novel Bayesian GGM using a hierarchical matrix-F prior with a fast implementation. We show, with extensive simulations and biological example analyses, that this prior has competitive network recovery capabilities compared to state-of-the-art approaches and good properties for recovering meaningful networks. We present a new way of tuning the shrinkage hyperparameter by constraining the condition number of the estimated precision matrix. For edge selection, we propose using approximated credible intervals (CI) whose width is controlled by the false discovery rate. An optimal CI is selected by maximizing an estimated F1-score via permutations. In addition, a specific choice of hyperparameter can make the proposed prior better suited for clustering and community detection. Our method, with a generalized expectation-maximization algorithm, computationally outperforms existing Bayesian GGM approaches that use Markov chain Monte Carlo algorithms.

The method is implemented in the R package HMFGraph, found on GitHub at https://github.com/AapoKorhonen/HMFGraph. All codes to reproduce the results are found on GitHub at https://github.com/AapoKorhonen/HMFGraph-Supplementary.

Author summary

In this paper, we introduced a new way to recover network structures with biological datasets in mind. Network estimation methods can help research by providing a convenient and easy-to-understand way to explain multivariate data structures and variable interactions. This can show, for example, how different genes co-express. Here, we introduced a new model structure that has competitive network recovery capabilities compared to state-of-the-art methods in a wide range of simulation settings. In addition, we include examples with real datasets and explain how the interpretation changes with different model parameter values. Estimation is performed by using our fast algorithm, which has significant computational advances over conventional estimation methods. Our method has a user-friendly implementation as an R package and is publicly available for download on GitHub at https://github.com/AapoKorhonen/HMFGraph.

Citation: Korhonen AE, Sarala O, Hautamäki T, Kuismin M, Sillanpää MJ (2025) HMFGraph: Novel Bayesian approach for recovering biological networks. PLoS Comput Biol 21(10): e1013614. https://doi.org/10.1371/journal.pcbi.1013614

Editor: Lun Hu, Xinjiang Technical Institute of Physics and Chemistry, CHINA

Received: May 16, 2025; Accepted: October 14, 2025; Published: October 30, 2025

Copyright: © 2025 Korhonen et al. This is an open access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Data Availability: All codes for reproducing the results are publicly available at https://github.com/AapoKorhonen/HMFGraph-Supplementary. The showcased method can be tested with our R package HMFGraph, also publicly available at Github: https://github.com/AapoKorhonen/HMFGraph.

Funding: This research was supported by the Research Council of Finland (grants 349393, 329439), the Finnish Doctoral Program Network in Artificial Intelligence, AI-DOC (decision number VN/3137/2024-OKM-6), and Vilho, Yrjö and Kalle Väisälä Foundation. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Competing interests: The authors have declared that no competing interests exist.

1. Introduction

Data structures and conditional dependencies between multiple variables can be easily explored with graphical models. Gaussian graphical models (GGM) are used for network estimation, and they provide a convenient way of producing partial correlation networks [1]. They have been studied for a long time, and a wide range of methods have been developed, while recently Bayesian methods are increasingly gaining popularity [2,3].

Biological networks estimated with GGM can reveal complex associations between data variables, e.g., finding gene co-expression patterns in gene expression datasets [4,5]. GGMs are general tools that can be applied to various biological datasets, inferring, e.g., cancer pathways, metabolic networks, protein networks, gene networks, or microbiome networks [68]. GGMs provide a convenient and easy-to-interpret representation, which may help researchers to make novel biological discoveries. This includes finding new connections between genes, operational taxonomic units (OTUs), or other omics variables, as well as identifying groups of genes that are potentially operating in related biological processes by inspecting clustering structures between them [9]. In addition, biological networks can help identify important cancer genes [10]. Biological datasets often have properties that must be taken into account during the network estimation procedure. For example, they may be high-dimensional, and the networks could exhibit scale-free or cluster-like structures.

With GGMs, the data are assumed to follow a multivariate normal distribution, and the partial correlation structure is constructed from an estimated inverse covariance (precision) matrix. Non-zero elements of the precision matrix correspond to non-negligible partial correlations, which again determine the edges of the graph [11]. In high-dimensional situations, i.e., when there are fewer samples than variables, we need to introduce regularization to the problem in order to produce well-conditioned estimates for covariance and precision matrices. The most notable frequentist method for GGM is the graphical lasso (Glasso), which uses a lasso penalty to induce sparsity in the estimated precision matrix.

When estimating covariance and precision matrices in the Bayesian framework, we have to use a prior that ensures positive definite estimates and regularizes the estimator. Wishart and inverse-Wishart prior distributions are suitable priors for precision and covariance matrices, respectively, since they fulfil both of these criteria. Wishart distribution is also a conjugate prior for (multivariate) normally distributed data, giving a closed-form expression for the corresponding posterior distribution. Thus, Wishart and inverse-Wishart priors have been used in a wide range of applications for covariance and precision matrix estimation [12]. There are some extensions for Wishart prior because it can be too inflexible due to only having a single tuning parameter [13]. These include hierarchical Wishart [14], matrix-F (scale mixture of Wisharts) [15], and flexible inverse-Wishart prior [13]. This paper aims to extend this line of research further by providing a hierarchical version of the matrix-F prior.

Previously, Wishart prior has been widely used for Bayesian GGMs. Wishart prior introduces a ridge-type regularization to the precision matrix, which will not yield a sparse estimate. If sparsity is preferred, an additional post-selection step (via decision rule) is required to determine which elements can be set to zero. Previously, a couple of decision rules have been used in this context, including the extended Bayesian criterion [16], Bayes factors [17], and posterior probabilities (credible intervals) [18]. In [17], they used an empirical Bayesian method for selecting an optimal hyperparameter for the Wishart prior and analytically calculated values for the Bayes factor to effectively select edges for the graph. This method turns out to be computationally extremely efficient. On the other hand, it has the same possible problems as other Wishart prior models, namely inflexibility.

An extension for Wishart distribution, namely G-Wishart, was developed with graphical models in mind. A graph estimate is produced at each Markov chain Monte Carlo (MCMC) step. A major drawback of G-Wishart is that it necessitates using complicated sampling algorithms. There is a direct sampler for G-Wishart [19], but it is computationally demanding for large graphs.

False discovery rate (FDR) control has previously been used with network estimation methods [17,20,21]. By controlling the target FDR, we can set an a priori expectation for the number of wrong edges. For instance, if we set our target FDR to 0.2, then we can expect 20% of our estimated network’s edges to be incorrect. In some cases, we have to tolerate some level of false positives in order to estimate any network at all. For example, a network estimate that is too sparse can make clusters difficult to recognize. By managing the FDR, we can gain a clear understanding of the extent to which we need to tolerate false edges.

In Bayesian analysis, the posterior distribution can be estimated by generating dependent samples from the posterior density. This can be done with MCMC methods [22] and one of the simplest options is the Gibbs sampler [23]. It is usable in cases where a posterior distribution is not analytically available, but the full conditional distribution for each parameter can be derived. Gibbs sampler is easily available with conjugate priors. By using MCMC methods, the maximum a posteriori (MAP) estimate and the shape of the posterior density (summarizing uncertainty) are attained. In [24] and [15], a Gibbs sampler for matrix-F prior was introduced.

The primary novelties in this paper are:

  • A novel hierarchical matrix-F prior for Bayesian GGM with a user-friendly and flexible parameterization.
  • A generalized expectation-maximization (GEM) algorithm that has remarkable computational advantages over previous MCMC methods.
  • A new way of tuning the hyperparameter by constraining the condition number of the estimated precision matrix.
  • Credible intervals (CI) are approximated with a normal distribution.
  • The optimal CI is selected with permutations and maximizing an estimated -score. This also allows us to control the FDR to a desired level.

We show that our method works well with high-dimensional data and performs considerably better than current state-of-the-art methods. It is flexible and has good network recovery performance in all tested scenarios. Finally, we include examples with real biological datasets and show how a specific choice of hyperparameter can make the proposed prior better suited for clustering and community detection.

The rest of this paper is structured as follows. In Sect 2, we introduce the method in detail. In Sects 3.1 and 3.2, we provide examples and results with simulated and real datasets, respectively. In Sect 4, we discuss the results, the method, and possible future improvements.

2. Methods and implementations

In this section, we will describe our approach in detail. Each major step with the approach is showcased in our flowchart in Fig 1. All important mathematical symbols and notation are explained in Table 1.

Fig 1. Flowchart of the proposed Bayesian approach for network recovery.

https://doi.org/10.1371/journal.pcbi.1013614.g001

Table 1. Mathematical symbols and notation used throughout this article.

https://doi.org/10.1371/journal.pcbi.1013614.t001

2.1. Background on Gaussian graphical models

The data Y is expected to follow a multivariate normal distribution in GGMs with a zero-mean vector and a covariance , and Ω being the precision matrix or inverse of the symmetric and positive-definite covariance matrix. For all samples j: , where is p-length zero vector, where p is the number of variables.

Graphs are represented as a set of vertices V and edges E [11]. If , then vertices and are connected in the graph. With GGMs, connections in the network are determined by off-diagonal elements of the precision matrix . If then . This information is compiled in a symmetric adjacency matrix A = [aij], which is defined as:

(1)

2.2. Bayesian Gaussian graphical models and matrix-F prior

Posterior distribution in the Bayesian framework is obtained from the likelihood and prior distribution:

(2)

where is a likelihood and is a prior. The likelihood function in Bayesian GGM is

(3)

where S is a sample covariance matrix. Here Ω is a positive-definite matrix.

The matrix-F distribution was introduced in [15] as a suitable and flexible prior for covariance and precision matrices. In contrast to one parameter for the degree of freedom in the regular Wishart distribution, the matrix-F has two parameters. The matrix-F prior was used in the Bayesian GGM setting by [24]. The prior is

(4)

where B is a target matrix, and are degree of freedom parameters. The parameter controls the center of the distribution, and the parameter δ controls the tail behavior.

For Bayesian GGM, the matrix-F can be represented as a scale mixture of Wisharts:

(5)

and it can be modeled with two separate priors:

(6) (7)

where Φ is a target matrix for the first Wishart distribution, and it can be thought of as an auxiliary parameter.

We propose a modification to the matrix-F prior so that the parameters are simpler to interpret. We first set the same Wishart scale mixture as in (6) and (7), but we add scaling terms and for target matrices Φ and B, respectively. This modification will simplify the interpretation of the model hyperparameters.

We set the parameter B to be a diagonal matrix, i.e., all off-diagonal elements are zero. For diagonal elements bii, a flat prior on a logarithmic scale was proposed to be used with the Wishart prior in [25]. [14] implemented it with a Metropolis-Hasting-within-Gibbs algorithm. Interestingly, this prior is similar to with small values of and (e.g., 0.001). In fact, when , then .

The gamma prior ensures a proper posterior distribution, but the flat prior on the logarithmic scale is scale-invariant and should be more flexible when working with differently scaled data. In Section M in S1 Text, we include proof that the posterior is proper with the improper prior when n > p. The posterior is proper with the gamma prior even when n < p (see Section N in S1 Text). Therefore, we use the gamma prior for the diagonal elements of the matrix B.

2.3. Hierarchical matrix-F prior for Gaussian graphical models

Our model, using the scaled matrix-F prior and the gamma prior, is

(8) (9) (10) (11)

Hereafter, we call this model formulation a hierarchical matrix-F prior for Gaussian graphical models, or HMFGraph.

A significant benefit of these priors is that the full conditional distributions are known for all parameters Ω, Φ, and B (Section A in S1 Text). The modes of the full conditional distributions are used in our GEM algorithm. For parameters Ω and Φ, they are

(12) (13)

where and are the full conditional distributions of the parameters Ω and Φ, respectively.

Because of our scaling terms, we can present the modes of conditional distributions with new parameters, which are α and β:

(14)

where and (the lower limit of β depends on , see Section A in S1 Text). Similar substitution was done together with ordinary Wishart distribution in [17] and [26]. The modes of conditional distributions of parameters Ω and Φ utilized in our GEM algorithm can now be presented in an interpretable way:

(15) (16)

The mode (15) bears a similarity to Ledoit-Wolf shrinkage (or a linear shrinkage) [27].

2.4. Computational methods for the posterior distribution estimation

We formulated two methods for estimating the posterior distribution of the HMFGraph model. The first is a Gibbs sampler that can estimate the MAP and the shape of the posterior distribution of the parameter Ω. The second is a GEM algorithm that is computationally faster than the Gibbs sampler but is only able to acquire the MAP estimate [28,29]. We use the GEM algorithm in all example analyses unless specified otherwise.

The Gibbs sampler is explained in Section B in S1 Text. The GEM algorithm will be introduced in the next section.

2.4.1. GEM algorithm.

The generalized expectation-maximization algorithm, or GEM algorithm, can be used in various statistical applications (e.g., [30]). Commonly, it is used for maximum likelihood and MAP estimation [28]. GEM and EM algorithms were originally designed for estimation problems with incomplete data or with hidden variables. These algorithms, in general, include two main steps: E-step, where the Q-function is updated with new hidden variables (or incomplete data), and M-step, where parameters are updated by maximization. The main difference between EM and GEM algorithms is the M-step. With the EM algorithm, the Q-function is maximized in terms of parameters in the M-step. With GEM, the Q-function does not need to be fully maximized in the M-step; it just needs to increase it (see [29, p. 24]). This makes the algorithm more usable with multiple parameters.

In our case, we operate in a complete data situation, thus the E-step becomes an identity operation [29]. This means that all M-steps in the algorithm are parameter-wise conditional maximizations of the posterior distribution. Since we already know the full conditional posterior distributions for each parameter, we can maximize each conditional distribution, and the algorithm turns into a so-called conditional maximization algorithm (see [29, p. 162]). In other words, the GEM algorithm is the same as the Gibbs sampler for our model, but instead of sampling from the distribution, we calculate its mode, which is usually analytically available. In fact, first forming the Gibbs sampler greatly helps to formulate a GEM algorithm [31]. The similarities between the Gibbs sampler and the GEM algorithm are most thoroughly examined in [32]. Also, the expectation conditional maximization algorithm, a special case of GEM algorithms, is analogous to the multi-step Gibbs sampler [29]. Our GEM algorithm is showcased in Algorithm 1 and in Section C in S1 Text. The algorithm can also be called the ECM or iterative conditional modes (ICM) algorithm [33]. It should be noted that for our model, a regular EM algorithm is not possible to implement, but the more flexible nature of the GEM algorithm allows for an easy implementation.

Algorithm 1 GEM algorithm for Bayesian GGM using a hierarchical matrix-F prior.

Require: max-iters, Stop-criterion, , δ, n, p

, , ,

for to max-iters do

for to p do

end for

if Stop-criterion then

break

end if

end for

2.5. Hyperparameter selection

With our model formulation, there are two hyperparameters that need to be tuned. They are α (or ) and β (or δ). Their minimum values correspond to the minimal information priors in the Bayesian framework [15].

The parameter α controls how much we trust the sample covariance matrix. In high-dimensional situations (i.e., p > n), this trust should be low, and α should then be set to a large value (close to one). As the sample size increases, the α should decrease. The exact value is still difficult to select by hand, especially for different data types, even though the interpretation is clear. We propose a method to select a suitable value for α based on a constraint on a condition number, which has been proven to construct well-conditioned estimators for covariance and precision matrices [16,3436]. The condition number is the ratio of the largest and smallest eigenvalues of the estimated precision matrix:

(17)

where is the MAP estimate of the precision matrix acquired with a some α value, and and are the largest and smallest eigenvalues of the precision matrix , respectively.

Our proposed method chooses the smallest α value that produces a smaller condition number than the predetermined limit :

(18)

This raises a new problem: how to choose a correct value. Here, we set the constraint for the condition number to be lower than the condition number of a Ledoit-Wolf estimate (LW) of the same precision matrix [27,37]. The optimal α value is

(19)

where is the condition number for estimated Ω with some α value, and is the condition number of a LW-estimate of the precision matrix. We call this a condition number constraint -method, or CC-method. We calculate the LW-estimate using the R package nlshrink and the function linshrink_cov. We use a simple binary search algorithm (Algorithm A and Section E in S1 Text) to find a solution to Equation (19). In order to speed up the algorithm, we utilized a "warm-start" trick [38,39].

The parameter β controls how much we trust the target matrix B. Because the target matrix B is a diagonal matrix, a large β value will induce a large regularization on off-diagonal elements and shrink them towards zero (see Fig 2). This shrinkage will also reduce the number of connections in the estimated network (see Fig G in S1 Text). The shrinkage is less noticeable if α is selected using the CC-method (see Figs F and G in S1 Text). This can be interpreted such that, if the true network is expected to be sparse, the β value should be set to a large value. The larger the β value, the sparser the obtained network. Based on our experience, the β value does not affect the results as much as α due to the hierarchical structure of our prior and the CC-method (see Figs A-D in S1 Text). Similar behavior was observed with δ when using the matrix-F prior [24]. For all results, we used . We include more discussion on hyperparameter selection in Section D in S1 Text.

Fig 2. Distribution of off-diagonal elements in the estimated precision matrix with hyperparameters: varying values (colors) and fixed = 0.80.

The graphical model considered in this example is a scale-free network generated with the R package huge ().

https://doi.org/10.1371/journal.pcbi.1013614.g002

2.6. Edge selection with credible intervals

Because our model will not produce a sparse estimate for the precision matrix, we have to add an edge selection step after the estimation procedure. This is true for both the Gibbs sampler and the GEM algorithm. Note that this kind of edge selection is also required for models using sparsity-inducing priors, like spike-and-slab and horseshoe priors, that should produce a sparse precision matrix estimate (see e.g., [4042]). Previous methods using Wishart priors have, for instance, used Bayes factors as a decision rule for edge selection [17,24]. Here, we use equal-tailed credible intervals for the edge selection.

The posterior distribution can be used to select nonzero elements from the estimated precision matrix by investigating if zero can be found from the credible interval. This can be expressed as:

(20)

where is (ij) element’s credible interval. For example, we could use a 95% credible interval () as a selection rule to construct our adjacency matrix. If value zero is not included in the 95% credible interval of an off-diagonal element of the precision matrix, the corresponding element will be set to 1 in the adjacency matrix .

2.7. Approximation for credible intervals

When we estimate the MAP for the precision matrix with the GEM algorithm, we neither obtain an estimate for the posterior distributions nor credible intervals. As a solution for this, we chose to use a normal approximation for the posterior distribution of each off-diagonal element of the estimated precision matrix , as was suggested in [18]. Also, the variance-gamma distribution could be used [43] (see Section F in S1 Text). For the mean of the normal approximation, we use the MAP estimate. We know that the full conditional distribution for Ω is the Wishart distribution with known variance (Equations (G) and (P) in S1 Text). We can use this for the normal approximation. The approximation for (ij) off-diagonal element is

(21)

where is the MAP estimate of (ij) off-diagonal element of the precision matrix and is its estimated variance. The estimated variance of the MAP estimate is

(22)

where and is a MAP estimate of the parameter Φ (Equations (G) and (P) in S1 Text).

Credible intervals for off-diagonal elements can now be estimated:

(23)

where is a probit function or the inverse cumulative distribution function for the standard normal distribution.

We compare these approximated credible intervals to the credible intervals that we obtain from the Gibbs sampler and the network recovered based on them in Figs 3 and 4. The GEM algorithm forms almost the same credible intervals as the Gibbs sampler, and this approximation can be considered useful and accurate for our purpose. The recovered networks based on CIs are similar.

Fig 3. Comparisons of 90% credible intervals (CIs) for off-diagonal elements of precision matrix Ω.

For the Gibbs sampler, CIs are calculated from posterior samples. For the GEM algorithm, they are estimated with a normal approximation. The graphical model considered in this example is a scale-free network generated with the R package huge ().

https://doi.org/10.1371/journal.pcbi.1013614.g003

Fig 4. Comparison of recovered networks with the Gibbs sampler and the GEM algorithm using credible intervals.

The graphical model considered in this example is a scale-free network generated with the R package huge ().

https://doi.org/10.1371/journal.pcbi.1013614.g004

2.8. False discovery rate control with permutations

The false discovery rate (FDR) describes the proportion of false positives among the estimated edges. FDR is generally defined as:

(24)

In multiple testing situations, it is necessary that we can control the FDR [44]. Network estimation can also be considered to suffer from multiple testing issues. With p number of variables, we have to carry out number of tests [45]. As p increases, the number of tests increases greatly. Various ways to control FDR have been used with network estimation methods [17,20,21]. In order to do a frequentist-style multiple-testing correction in the Bayesian framework, we have to adjust the width of the credible interval.

For networks, we define FDR as:

(25)

where Θ includes all off-diagonal elements that are zeros in the hypothetical true adjacency matrix A. The true adjacency matrix A is only acquirable with simulated data or under some specific hypothesis. The estimated adjacency matrix is dependent on the CI(γ). The total number of true and false positives (denominator of Equation (24)) is simply the number of edges or nonzero elements in the upper or lower triangle of the estimated adjacency matrix (). False positives (numerator of Equation (24)) are the number of edges found, of which true partial correlations are zero or negligibly small ().

The number of false positives is estimated using a permutation method. We expect that the permutation eliminates all correlation structures present in the original data. The permuted data will be obtained by permuting variable values within each sample separately, assuming the variables have a similar scale. For permuted data, all off-diagonal elements in the true adjacency matrix AP are hypothesized to be zeros (). Consequently, all connections in the estimated network are false positives. We use this to approximate the number of false positives in the estimated network. The approximation is

(26)

where is the adjacency matrix acquired with permuted data. The expected value is attained by repeating the permutation multiple times and calculating the median of . For all results, we use 50 permutations. A similar permutation method for controlling FDR has been used with a fused lasso latent feature model [46], and permutations have also been used with network models [47,48]. The estimate of true positives is now .

We control FDR by adjusting the CI such that the target FDR is reached:

(27)

where is the CI that produces the network with the specified target FDR. We illustrate the precision of this FDR control with multiple α values in Fig 5 and with different β values in Figs H and I in S1 Text. The FDR control seems to be the most accurate with α values selected using the CC-method. The accuracy of FDR control also depends on the β value, but a suitably selected α seems to be more important accuracy-wise.

Fig 5. Controlling FDR with multiple different α values.

Averages of 50 different simulated datasets. For all datasets, p = 100 and n = 300.

https://doi.org/10.1371/journal.pcbi.1013614.g005

2.9. An optimal credibility interval

With permutations, we get estimates for the number of false positives () and true positives (). This enables us to use other metrics than FDR to select an optimal CI(γ). In our assessment, maximizing the -score will provide a graph that has an informative number of true positives compared to false positives [49]. An estimated is

(28)

where and K is the expected number of real connections in the network (see Section G in S1 Text). Our rule of thumb is to set it to the number of variables p, which we use for all results.

The optimal credible interval is achieved by maximizing the :

(29)

We want to point out that the maximization metric can be changed from to some other value depending on the situation. For example, the Matthews Correlation Coefficient (MCC) can be used instead of .

3. Results

3.1. Examples with simulated data

Simulated data enables us to compare our method to other network estimation methods. We opted to simulate our test data with two R packages, huge and BDgraph. In our experience, data simulated with huge is more difficult to estimate than BDgraph’s data in high-dimensional situations. We used both R packages to generate scale-free and cluster-type networks. Both packages generate the scale-free network structures in the same way, but the cluster structures differ. Both huge and BDgraph generate a network consisting of five disconnected clusters, meaning there are no connections between the clusters. The clusters are internally sparser with BDgraph, and the average clustering coefficient (ACC) is smaller with the network structure generated with BDgraph.

The main difference between huge and BDgraph is in how the data samples are generated and how the network structure is incorporated into the partial correlations. huge sets all partial correlation values to the same value that depends on the number of variables. BDgraph randomly samples the partial correlations, and they can be anything between –1 and 1. To our knowledge, no established practice exists for simulating data with a specific network structure in mind. Thus, we are eager to use both data generation methods to illustrate how our method works well with both generators, which is not true for all GGM methods.

With both R packages and for both network structures, we generated datasets with sample sizes (n) of 35, 75, 150, and 300. For all sample sizes, the number of variables is 100 (p). With each sample size, we generated 50 different datasets. In Figs A-E in S1 Text, all these datasets were used.

For comparison we also picked six GGM methods: (i) graphical lasso (Glasso) [50] with stability approach for regularization selection (StARS) [51]; (ii) fast Bayesian GGM (FBGGM) [17]; (iii) G-Wishart using Birth-Death MCMC sampling algorithm [52]; (iv) CLEVEL [21,53]; (v) thresholded adaptive validation (thAV) for Glasso [54]; (vi) tuning-insensitive graph estimation and regression (TIGER) [55]; and (vii) Bayesian Gaussian Graphical Models using matrix-F prior (BGGM) [24,56]. All of these have convenient R packages that make the comparison easy to implement. For Glasso with StARS, we used R packages pulsar [57] and huge [58], FBGGM is implemented with R package beam [17], G-Wishart with BDgraph [52], TIGER with R package flare [59], BGGM with R package BGGM [60], CLEVEL with R package PCGII [61], which also includes a new method called PCGII, that has the same properties than CLEVEL, but user can additionally incorporated a prior information to the estimation. BGGM implementation does not work with high-dimensional datasets and thus results with and are not available. BGGM is designed for use with psychological datasets that often have a large number of samples. For more details, see Section I in S1 Text.

We measured the network-recovery performance with four metrics for scale-free networks and three metrics for cluster networks. For scale-free, we report Matthews Correlation Coefficient (MCC), -score (), false discovery rate (FDR), and true positive rate (TPR). For the cluster network, we report the average clustering coefficient (ACC), normalized mutual information (NMI), and -score. All metrics range from 0 to 1, except MCC, which has a value range from –1 to 1. For NMI, , MCC, and TPR, values closer to 1 indicate better network recovery performance. For FDR, the lower value is considered preferable, and ACC is data-dependent. We considered the most informative metric for network estimation due to the highly uneven number of edges and absence of edges [49]. We described all metrics in greater detail in Section H in S1 Text.

3.1.1. Computation time comparison between MCMC samplers and GEM algorithm.

In order to showcase the computational advantages of our GEM algorithm, we compared it with our implementation of Gibbs sampler, Gibbs sampler for matrix-F prior found in the R package BGGM, and Birth-Death MCMC sampler for G-Wishart prior in the R package BDgraph. In addition, we included computation times for our method with permutations for selecting an optimal credible interval and with the CC-method. We tested all methods with 100, 200, 300, and 400 variables. The test datasets were generated with the R package BDgraph. We repeated each run 5 times. The test computer uses an Intel Core i7-11700F processor with 32 GB of RAM. Computation time tests are heavily affected by the computer’s capabilities, and we recognize the difficulty of reproducing these results. Regardless, this should still show the clear computational benefit of using the GEM algorithm over MCMC methods. The results are found in Table 2.

Table 2. Comparisons of median computation times (in seconds) based on 5 runs for the GEM algorithm and the MCMC methods.

HMFGraph Gibbs, G-Wishart, and BGGM were run with 5000 iterations. For all p, .

https://doi.org/10.1371/journal.pcbi.1013614.t002

Computationally, our GEM algorithm is considerably faster than any MCMC method. Even with permutations and α-selection using parallelization, our method outperforms the MCMC methods. Without parallelization, our method has computation times similar to those of the Gibbs sampler and BGGM. Parallelizing our method is straightforward and helps computations immensely.

Memory usage is also a concern for larger networks. MCMC methods naturally use more memory because all samples of the precision matrix are saved. In R, this means an array with size of (number of iterations). The GEM algorithm does not use a lot of memory because only two iterations have to be kept in memory simultaneously. Multiple GEM algorithms can be run in parallel without requiring a large amount of system memory.

3.1.2. Simulated scale-free network datasets.

We present the results for datasets with a scale-free network structure in Table 3. Figs J-M in S1 Text illustrate how the recovered networks visually differ between all tested methods. With HMFGraph, we selected the best α value based on the CC-method (see Fig E in S1 Text) and the optimal credible interval by maximizing the estimated -score. In addition, results with a target FDR of 0.2 are included. For all methods, default values were used, except for CLEVEL. A target FDR can be selected for CLEVEL, and we used the same 0.2 as with our method, but we also included the results with the default target FDR used in the R package (0.05). It should be noted that only CLEVEL and our HMFGraph have a built-in FDR control procedure. FDR control provides both HMFGraph and CLEVEL a benefit over all other methods. We can now also compare our FDR control procedure to CLEVEL’s and find out how accurate the permutation-based approach is.

Table 3. Means of performance metrics.

The metrics are Matthews correlation coefficient (MCC), false discovery rate (FDR), -score (), and true positive rate (TPR). Averages of 50 different datasets. Standard deviations are in parentheses. The simulated networks have a scale-free structure and the datasets are generated with the R packages huge and BDgraph.

https://doi.org/10.1371/journal.pcbi.1013614.t003

In general, our method performs well with all sample sizes and with both data generators when considering the and MCC values. In the worst case, our method has the third-best values with 75 and 150 samples with huge datasets. The other methods tested do not show this level of consistency with their network recovery capabilities. For example, Glasso with StARS has a rather competitive performance with huge datasets, but it has the second-worst performance with BDgraph datasets. The same can be observed with thAV, which works well with BDgraph, but not so well with huge. TIGER performs well with high-dimensional datasets, but the performance decreases relative to other methods when the sample size grows. In contrast, our method has either the best performance or close to the best with all sample sizes and with both generators.

The real FDR of our method is closer to the set target FDR than CLEVEL, especially with datasets generated with BDgraph. For instance, when n = 75, our method achieves, on average, an FDR value of 0.35 (target FDR = 0.2), but for CLEVEL, the corresponding FDR value is 0.69. This indicates that our permutation method is more flexible and can work with different data types.

BGGM has the worst network recovery capability out of all tested methods. This is perhaps due to that the implementation is designed for low-dimensional problems (). As noted previously, the implementation in BGGM does not work with high-dimensional datasets and thus only results with n = 150 and n = 300 are presented. Our method is noticeably better than BGGM when considering F1 and MCC performance metrics, even though BGGM uses a matrix-F prior as our method, but without the hierarchical part. The results show that our hierarchical matrix-F prior and hyperparameter tuning is an immense improvement for high-dimensional datasets, and even when n > p.

3.1.3. Simulated cluster network datasets.

For cluster network datasets, the results are found in Table 4. We ran our method with two different α values: the first results in the table are acquired with the CC-method, and the latter with , which produces larger α values than the CC-method (see Fig E in S1 Text).

Table 4. Means of performance metrics.

The metrics are average clustering coefficient (ACC), normalized mutual information (NMI), and -score (). Averages of 50 different datasets. Standard deviations are in parentheses. The simulated networks have a cluster structure and the datasets are generated with the R packages huge and BDgraph.

https://doi.org/10.1371/journal.pcbi.1013614.t004

For our method, a large α value helps with clustering. NMI values are better for both data generators with larger α values. A major downside is that graph recovery capabilities suffer. For data generated with huge, the best values are still acquired with , but for the data generated with BDgraph, the CC-method produces the best values, but NMI is considerably worse. The negative correlation between NMI and is true for other methods as well. For instance, Glasso with StARS has the best NMI for BDgraph dataset with n = 300, but its is the fourth lowest. Our method produces more clustered networks (larger ACC) with large α values than with α selected with the CC-method. values are better with our method when α is selected with the CC-method. This can be interpreted such that with large α values, the estimated networks are denser and less accurate than with α values selected with the CC-method, but the denser graph helps with identifying the clusters and cluster members.

As with scale-free networks, HMFGraph outperforms BGGM based on NMI and F1 performance metrics. This further suggests that our method has superior network recovery capabilities in these situations.

3.2. Examples with real datasets

Simulated data rarely behaves in a similar way to a real dataset, and tests with real datasets can be considered necessary. We apply our method to two real datasets: a riboflavin dataset and an American gut dataset [62]. The riboflavin dataset is publicly available in the R package hdi [63], and the American gut dataset (amgut2.filt.phy) is available in the R package SpiecEasi [64]. With the riboflavin dataset, only 100 genes with the highest variance in expression were selected for the analysis, similar to [65] and [54]. The datasets have been tested with two other GGM methods before, namely with the Meinshausen and Bühlmann (MB) graph estimation method and with thAV [54,65,66]. This gives us an easy way to compare our results with those methods. As these datasets are not normally distributed, we first normalize both datasets with nonparanormal transformation using the R package huge [67].

3.2.1. Riboflavin dataset.

The riboflavin dataset has 100 gene expression measurements and riboflavin production amounts collected from 71 observations (p = 101, n = 71). The study species is a bacterium called bacillus subtilis, from which gene expression and riboflavin production were measured. Riboflavin production was included as an extra node to the analysis. For our analysis, CI width was selected using permutations, and the target FDR was set to 0.2. The resulting graph is shown in Fig 6. For the same dataset, thAV produced a sparser graph than our method, but some similar structures can still be found [54]. The MB method recovered a denser graph in turn, which makes it much more difficult to structurally compare the graphs. They used StARS for parameter selection, which tends to produce dense graphs [51,65].

Fig 6. The network recovered from the riboflavin dataset using the HMFGraph method.

The network was constructed with a target FDR set to 0.2. The node describing the riboflavin production is highlighted in red. Genes linked to riboflavin production are highlighted with colors describing the target FDR level, triggering a connection.

https://doi.org/10.1371/journal.pcbi.1013614.g006

In Table 5, we listed all genes that are connected to the riboflavin production node, which is labeled as 1 in Fig 6 with varying target FDR values (see Fig N in S1 Text). When the target FDR increased, the number of connections to ribolavin production increased. Interestingly, almost all listed genes have been previously linked to riboflavin production in earlier analyses with this dataset [63,65,6870]. The first gene that connects to the riboflavin production is YXLE_at. The same gene was found to be linked to riboflavin production with several multiple regression methods [68,70]. The other genes are less commonly found by other methods, and those that require a larger target FDR value to be recovered are more likely to be false positives or falsely identified partial correlations.

Table 5. Genes connected to the riboflavin production node with different target FDR levels.

https://doi.org/10.1371/journal.pcbi.1013614.t005

An advantage of using the GGM method is that one can visualize all gene co-expressions in addition to the genes connected with the riboflavin production. For example, from the graph we can see that YXLE_at is a part of a larger cluster (including nodes 42, 95-99 and 101, see Section K in S1 Text). This cluster includes a gene YXLD_at (node 96), which is often linked to riboflavin production [65,68,69,71]. Based on our analysis, we can postulate that the gene YXLD_at is not directly linked or partially correlated with riboflavin production, but the apparent connection is due to the gene YXLE_at. Also, the whole cluster has been identified previously (see STRING database [72,73]) and many genes in it have been shown to interact biologically with each other (e.g., for YXLE_at and YXLD_at [74]).

3.2.2. American gut dataset.

The gut dataset consists of 296 samples of 138 operational taxonomic units (OTUs). In addition, each OTU’s respective taxon is known. We used two methods for selecting suitable α values. Fig 7A illustrates the network recovered with HMFGraph, whose α value was selected with the CC-method. Furthermore, we used a larger α value () in Fig 7B. We selected the optimal credible interval for both α values based on the estimated -score. With a larger α value, the recovered graph has a more clustered structure. The obtained network clusters closely follow the biological taxon level order. This behavior is similar to the one observed with the simulated datasets. Based on the remarks with the cluster dataset results, we can postulate that the network recovered with an optimal α value is closer to the real partial correlation structure in Fig 7A, but the underlying clusters are easier to recognize in Fig 7B, where a larger α value is used.

Fig 7. The network recovered from the gut dataset using HMFGraph.

An optimal credible interval was selected using 50 permutations. Different OTU groups (in order taxonomic rank) are illustrated with different colors. (A) An optimal value of parameter α was selected based on the CC-method (). (B) The value of parameter α was set to be large (e.g., ).

https://doi.org/10.1371/journal.pcbi.1013614.g007

The found clusters in Fig 7B are similar to those recovered with thAV [54]. For example, OTUs from Enterobacteriales order are closely grouped together, i.e., they have a lot of connections between each other in the same order, but almost no connections to OTUs in other orders. We additionally noticed that the clustering behavior between different taxa levels has been observed before. For instance, in [6], different OTUs on the taxon level family were found to have clustering properties. For comparison, we added our networks in Fig 7 with a different coloring scheme to highlight the family taxon (see Fig O in S1 Text). We can observe that with our analysis, we also see that OTUs from different taxa familiae do not partially correlate. A good example is familiae Bifidobacteriaceae and Bacteroidaceae, which do not have any connection between them, as was also observed in [6]. This can be interpreted such that OTUs from familiae Bifidobacteriaceae and Bacteroidaceae are found independently of each other in the human gut.

Generally, on phylum taxonomic level, the found network clusters do not follow phyla groups as closely as on other taxa levels [7578]. For instance, in [77], where they analyzed another gut microbiome dataset, the clustering between taxa phyla was noticeable, but there were connections between OTUs in different phyla. In Fig P in S1 Text, phyla taxa are highlighted with different colors in our recovered network. The results show that our method clusters the same phyla as in [77], but there are fewer connections between the clusters. A similar phenomenon was observed with other human microbiome datasets [76,78], where found clusters were similar to phyla groups. Our results can be interpreted as follows: if one OTU is found from some phylum, then it is more likely that other OTUs from the same phylum are also present.

4. Discussion

In this article, we introduced a novel hierarchical matrix-F prior for GGMs and showed its capabilities with high-dimensional datasets. The results demonstrated that a model using our hierarchical matrix-F prior is able to recover network structures in most cases better, or at least similarly, than competing methods. In addition, we provided a fast GEM algorithm with computational advantages over previous MCMC methods.

Many biological data do not fully follow the Gaussian assumption and also exhibit heterogeneity, such as the single-cell gene expressions, and may also be zero-inflated and compositional in nature (e.g., microbiome datasets). However, our method and often other GGMs are based on the assumption that the analyzed data are normally distributed, exceptions being, e.g., BLGGM [8]. In example analysis, we followed [54] and handled this problem by using a nonparanormal transformation [67]. Our analysis results illustrated that this approach can produce meaningful networks with non-Gaussian datasets. If a case study of the specific type of observations (e.g., zero-inflated microbiome data) is conducted, then a more thorough consideration of the validity of the data normalization procedure and model assumptions is recommended.

In our view, there is still work to be done in selecting the correct values for the parameters α and β (or and δ). Thanks to our reparameterization, interpretations for α and β are now easier and more intuitive than with and δ. We also explained a suitable value range of α when dealing with high-dimensional datasets and showed that the CC-method based on Ledoit-Wolf covariance shrinkage seems to work adequately. More detailed research is still required for the parameter selection.

We believe the optimal credible interval selection based on permutation can be used flexibly. Here, we used the credible interval that maximized the estimated -score, but other criteria can also be considered. Depending on the situation, this may yield a more informative graph recovery.

We want to expand our method for dynamic Gaussian graphical models in the future. A possible way of introducing a time dependency for this model is by setting the B matrix in the prior (10) to be a precision matrix of a previous time point.

Supporting information

S1 Text. Supporting information document.

Includes all supporting information materials.

https://doi.org/10.1371/journal.pcbi.1013614.s001

(PDF)

References

  1. 1. Pourahmadi M. High-dimensional covariance estimation: With high-dimensional data. John Wiley & Sons; 2013.
  2. 2. Kuismin MO, Sillanpää MJ. Estimation of covariance and precision matrix, network structure, and a view toward systems biology. WIREs Computat Stat. 2017;9(6).
  3. 3. Vogels L, Mohammadi R, Schoonhoven M, Birbil Şİ. Bayesian structure learning in undirected Gaussian graphical models: Literature review with empirical comparison. J Am Stat Assoc. 2024;119(548):3164–82.
  4. 4. Barabási A-L, Oltvai ZN. Network biology: Understanding the cell’s functional organization. Nat Rev Genet. 2004;5(2):101–13. pmid:14735121
  5. 5. Ideker T, Krogan NJ. Differential network biology. Mol Syst Biol. 2012;8:565. pmid:22252388
  6. 6. Kurtz ZD, Müller CL, Miraldi ER, Littman DR, Blaser MJ, Bonneau RA. Sparse and compositionally robust inference of microbial ecological networks. PLoS Comput Biol. 2015;11(5):e1004226. pmid:25950956
  7. 7. Layeghifard M, Hwang DM, Guttman DS. Disentangling interactions in the microbiome: A network perspective. Trends Microbiol. 2017;25(3):217–28. pmid:27916383
  8. 8. Wu Q, Luo X. Estimating heterogeneous gene regulatory networks from zero-inflated single-cell expression data. Ann Appl Stat. 2022;16(4).
  9. 9. van Dam S, Võsa U, van der Graaf A, Franke L, de Magalhães JP. Gene co-expression analysis for functional classification and gene-disease predictions. Brief Bioinform. 2018;19(4):575–92. pmid:28077403
  10. 10. Su X, Hu P, Li D, Zhao B, Niu Z, Herget T, et al. Interpretable identification of cancer genes across biological networks via transformer-powered graph representation learning. Nat Biomed Eng. 2025;9(3):371–89. pmid:39789329
  11. 11. Lauritzen SL. Graphical models. Clarendon Press; 1996.
  12. 12. Zhang Z. A note on Wishart and inverse Wishart priors for covariance matrix. JBDS. 2021;1(2).
  13. 13. Hsu C-W, Sinay MS, Hsu JSJ. Bayesian estimation of a covariance matrix with flexible prior specification. Ann Inst Stat Math. 2010;64(2):319–42.
  14. 14. Bouriga M, Féron O. Estimation of covariance matrices based on hierarchical inverse-Wishart priors. J Stat Plan Inference. 2013;143(4):795–808.
  15. 15. Mulder J, Pericchi LR. The matrix-F prior for estimating and testing covariance matrices. Bayesian Anal. 2018;13(4).
  16. 16. Kuismin M, Sillanpää MJ. Use of Wishart prior and simple extensions for sparse precision matrix estimation. PLoS One. 2016;11(2):e0148171. pmid:26828427
  17. 17. Leday GGR, Richardson S. Fast Bayesian inference in large Gaussian graphical models. Biometrics. 2019;75(4):1288–98. pmid:31009060
  18. 18. Williams DR. Bayesian estimation for Gaussian graphical models: Structure learning, predictability, and network comparisons. Multivariate Behav Res. 2021;56(2):336–52. pmid:33739907
  19. 19. Lenkoski A. A direct sampler for G-Wishart variates. Stat. 2013;2(1):119–28.
  20. 20. Yu L, Kaufmann T, Lederer J. False discovery rates in biological networks. In: Banerjee A, Fukumizu K, editors. Proceedings of The 24th International Conference on Artificial Intelligence and Statistics. vol. 130 of Proceedings of Machine Learning Research. PMLR; 2021. p. 163–171. Available from: https://proceedings.mlr.press/v130/yu21a.html
  21. 21. Wang H, Qiu Y, Guo H, Yin Y, Liu P. Information-incorporated gene network construction with FDR control. Bioinformatics. 2024;40(3):btae125. pmid:38430463
  22. 22. Robert CP, Casella G. Introducing Monte Carlo methods with R. Springer; 2010.
  23. 23. Casella G, George EI. Explaining the Gibbs sampler. Am Stat. 1992;46(3):167–74.
  24. 24. Williams DR, Mulder J. Bayesian hypothesis testing for Gaussian graphical models: Conditional independence and order constraints. J Math Psychol. 2020;99:102441.
  25. 25. Daniels MJ, Kass RE. Nonconjugate Bayesian estimation of covariance matrices and its use in hierarchical models. J Am Stat Assoc. 1999;94(448):1254–63.
  26. 26. Hannart A, Naveau P. Estimating high dimensional covariance matrices: A new look at the Gaussian conjugate framework. J Multivariate Anal. 2014;131:149–62.
  27. 27. Ledoit O, Wolf M. Honey, I shrunk the sample covariance matrix. JPM. 2004;30(4):110–9.
  28. 28. Neal RM, Hinton GE. A view of the em algorithm that justifies incremental, sparse, and other variants. Learning in Graphical Models. Springer Netherlands; 1998. p. 355–68. https://doi.org/10.1007/978-94-011-5014-9_12
  29. 29. McLachlan GJ, Krishnan T. The EM algorithm and extensions. John Wiley & Sons; 2007.
  30. 30. Kärkkäinen HP, Sillanpää MJ. Back to basics for Bayesian model building in genomic selection. Genetics. 2012;191(3):969–87. pmid:22554888
  31. 31. Meng XL, Rubin DB. Maximum likelihood estimation via the ECM algorithm: A general framework. Biometrika. 1993;80(2):267–78.
  32. 32. Sahu SK, Roberts GO. On convergence of the EM algorithmand the Gibbs sampler. Stat Comput. 1999;9(1):55–64.
  33. 33. Pungpapong V, Zhang M, Zhang D. Selecting massive variables using an iterated conditional modes/medians algorithm. Electron J Statist. 2015;9(1).
  34. 34. Won JH, Kim S-J. Maximum likelihood covariance estimation with a condition number constraint. In: 2006 Fortieth Asilomar Conference on Signals, Systems and Computers; 2006. p. 1445–9. https://doi.org/10.1109/acssc.2006.354997
  35. 35. Won J-H, Lim J, Kim S-J, Rajaratnam B. Condition number regularized covariance estimation. J R Stat Soc Series B Stat Methodol. 2013;75(3):427–50. pmid:23730197
  36. 36. Tabeart JM, Dance SL, Lawless AS, Nichols NK, Waller JA. Improving the condition number of estimated covariance matrices. Tellus A: Dyn Meteorol Oceanogr. 2020;72(1):1696646.
  37. 37. Ledoit O, Wolf M. A well-conditioned estimator for large-dimensional covariance matrices. J Multivariate Anal. 2004;88(2):365–411.
  38. 38. Sambharya R, Hall G, Amos B, Stellato B. Learning to warm-start fixed-point optimization algorithms. J Mach Learn Res. 2024;25(166):1–46.
  39. 39. Kim J, Kim S, Choi S. Learning to warm-start Bayesian hyperparameter optimization. arXiv preprint. 2018; doi:10.48550/arXiv.1710.06219
  40. 40. Li Y, Craig BA, Bhadra A. The graphical horseshoe estimator for inverse covariance matrices. J Computat Graph Stat. 2019;28(3):747–57.
  41. 41. Gan L, Yang X, Narisetty N, Liang F. Bayesian joint estimation of multiple graphical models. In: Advances in Neural Information Processing Systems; 2019. https://proceedings.neurips.cc/paper_files/paper/2019/file/94130ea17023c4837f0dcdda95034b65-Paper.pdf
  42. 42. Byrd M, Nghiem LH, McGee M. Bayesian regularization of Gaussian graphical models with measurement error. Computat Stat Data Anal. 2021;156:107085.
  43. 43. Fischer A, Gaunt RE, Sarantsev A. The variance-gamma distribution: A review. Statist Sci. 2025;40(2).
  44. 44. Benjamini Y, Hochberg Y. Controlling the false discovery rate: A practical and powerful approach to multiple testing. J R Stat Soc Ser B: Stat Methodol. 1995;57(1):289–300.
  45. 45. Li S, Cai TT, Li H. Transfer learning in large-scale Gaussian graphical models with false discovery rate control. J Am Stat Assoc. 2023;118(543):2171–83. pmid:38143788
  46. 46. Nowak G, Hastie T, Pollack JR, Tibshirani R. A fused lasso latent feature model for analyzing multi-sample aCGH data. Biostatistics. 2011;12(4):776–91. pmid:21642389
  47. 47. Kuismin M, Dodangeh F, Sillanpää MJ. Gap-com: General model selection criterion for sparse undirected gene networks with nontrivial community structure. G3 (Bethesda). 2022;12(2):jkab437. pmid:35100338
  48. 48. van Borkulo CD, van Bork R, Boschloo L, Kossakowski JJ, Tio P, Schoevers RA, et al. Comparing network structures on three aspects: A permutation test. Psychol Methods. 2023;28(6):1273–85. pmid:35404628
  49. 49. Chicco D, Tötsch N, Jurman G. The Matthews correlation coefficient (MCC) is more reliable than balanced accuracy, bookmaker informedness, and markedness in two-class confusion matrix evaluation. BioData Min. 2021;14(1):13. pmid:33541410
  50. 50. Friedman J, Hastie T, Tibshirani R. Sparse inverse covariance estimation with the graphical lasso. Biostatistics. 2008;9(3):432–41. pmid:18079126
  51. 51. Liu H, Roeder K, Wasserman L. Stability approach to regularization selection (StARS) for high dimensional graphical models. Adv Neural Inf Process Syst. 2010;24(2):1432–40. pmid:25152607
  52. 52. Mohammadi R, Wit EC. BDgraph: An R package for Bayesian structure learning in graphical models. J Stat Soft. 2019;89(3).
  53. 53. Qiu Y, Zhou X-H. Estimating c-level partial correlation graphs with application to brain imaging. Biostatistics. 2020;21(4):641–58. pmid:30596883
  54. 54. Laszkiewicz M, Fischer A, Lederer J. Thresholded adaptive validation: Tuning the graphical lasso for graph recovery. In: Banerjee A, Fukumizu K, editors. Proceedings of The 24th International Conference on Artificial Intelligence and Statistics. vol. 130 of Proceedings of Machine Learning Research. PMLR; 2021. p. 1864–72. Available from: https://proceedings.mlr.press/v130/laszkiewicz21a.html
  55. 55. Liu H, Wang L. TIGER: A tuning-insensitive approach for optimally estimating Gaussian graphical models. Electron J Statist. 2017;11(1).
  56. 56. Williams DR, Rast P, Pericchi LR, Mulder J. Comparing Gaussian graphical models with the posterior predictive distribution and Bayesian model selection. Psychol Methods. 2020;25(5):653–72. pmid:32077709
  57. 57. Müller CL, Bonneau RA, Kurtz ZD. Generalized stability approach for regularized graphical models. arXiv preprint; 2016. doi:10.48550/arXiv.1605.07072
  58. 58. Zhao T, Liu H, Roeder K, Lafferty J, Wasserman L. The huge package for high-dimensional undirected graph estimation in R. J Mach Learn Res. 2012;13:1059–62. pmid:26834510
  59. 59. Li X, Zhao T, Yuan X, Liu H. The flare package for high dimensional linear regression and precision matrix estimation in R. J Mach Learn Res. 2015;16:553–7. pmid:28337074
  60. 60. Williams D, Mulder J. BGGM: Bayesian Gaussian graphical models in R. JOSS. 2020;5(51):2111.
  61. 61. Wang H, Qiu Y, Liu P. PCGII: Partial correlation graph with information incorporation; 2024. https://haowang47.github.io/PCGII/
  62. 62. McDonald D, Hyde E, Debelius JW, Morton JT, Gonzalez A, Ackermann G, et al. American gut: An open platform for citizen science microbiome research. mSystems. 2018;3(3):e00031-18. pmid:29795809
  63. 63. Dezeure R, Bühlmann P, Meier L, Meinshausen N. High-dimensional inference: Confidence intervals, p-values and R-software hdi. Statist Sci. 2015;30(4).
  64. 64. Kurtz ZD, Bonneau R, Müller CL. Disentangling microbial associations from hidden environmental and technical factors via latent graphical models. Cold Spring Harbor Laboratory; 2019. https://doi.org/10.1101/2019.12.21.885889
  65. 65. Bühlmann P, Kalisch M, Meier L. High-dimensional statistics with a view toward applications in biology. Annu Rev Stat Appl. 2014;1(1):255–78.
  66. 66. Meinshausen N, Bühlmann P. High-dimensional graphs and variable selection with the Lasso. Ann Statist. 2006;34(3).
  67. 67. Liu H, Lafferty J, Wasserman L. The nonparanormal: Semiparametric estimation of high dimensional undirected graphs. J Mach Learn Research. 2009;10(80):2295–328.
  68. 68. Lederer J, Müller C. Don’t fall for tuning parameters: Tuning-free variable selection in high dimensions with the TREX. AAAI. 2015;29(1).
  69. 69. Bar HY, Booth JG, Wells MT. A scalable empirical Bayes approach to variable selection in generalized linear models. J Comput Graph Stat. 2020;29(3):535–46. pmid:38919169
  70. 70. Oliveira DCR, Schumacher FL, Oliveira MS, Zuanetti DA, Lachos VH. Revolutionizing high-dimensional regularization. In: Proceeding series of the Brazilian society of computational and applied mathematics; 2025. https://doi.org/10.5540/03.2025.011.01.0472
  71. 71. Chichignoud M, Lederer J, Wainwright MJ. A practical scheme and fast algorithm to tune the lasso with optimality guarantees. J Mach Learn Res. 2016;17(229):1–20.
  72. 72. Szklarczyk D, Nastou K, Koutrouli M, Kirsch R, Mehryary F, Hachilif R, et al. The STRING database in 2025 : Protein networks with directionality of regulation. Nucleic Acids Res. 2025;53(D1):D730–7. pmid:39558183
  73. 73. Szklarczyk D, Nastou K, Koutrouli M, Kirsch R, Mehryary F, Hachilif R, et al. Database: The STRING version 12; 2025. https://version-12-0.string-db.org/cgi/network?networkId=bvizcBOxWXjQ. Accessed 2025 September 3.
  74. 74. Cao M, Salzberg L, Tsai CS, Mascher T, Bonilla C, Wang T, et al. Regulation of the Bacillus subtilis extracytoplasmic function protein sigma(Y) and its target promoters. J Bacteriol. 2003;185(16):4883–90. pmid:12897008
  75. 75. Cougoul A, Bailly X, Wit EC. MAGMA: Inference of sparse microbial association networks. Cold Spring Harbor Laboratory; 2019. https://doi.org/10.1101/538579
  76. 76. Osborne N, Peterson CB, Vannucci M. Latent network estimation and variable selection for compositional data via variational EM. J Comput Graph Stat. 2022;31(1):163–75. pmid:36776345
  77. 77. Vinciotti V, Behrouzi P, Mohammadi R. Bayesian structural learning with parametric marginals for count data: An application to microbiota systems. J Mach Learn Res. 2024;25(339):1–26.
  78. 78. Zeng Z, Li M, Vannucci M. Bayesian covariate-dependent graph learning with a dual group spike-and-slab prior. Biometrics. 2025;81(2):ujaf053. pmid:40322851

AltStyle によって変換されたページ (->オリジナル) /