Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Oct 26.
Published in final edited form as: Proc AAAI Conf Artif Intell. 2021 May 18;35(11):10263–10272.

Gene regulatory network inference as relaxed graph matching

Deborah Weighill 1, Marouen Ben Guebila 1, Camila Lopes-Ramos 1, Kimberly Glass 1,2,3, John Quackenbush 1,2,3, John Platig 2,3, Rebekka Burkholz 1
PMCID: PMC8546743  NIHMSID: NIHMS1725904  PMID: 34707916

Abstract

Bipartite network inference is a ubiquitous problem across disciplines. One important example in the field molecular biology is gene regulatory network inference. Gene regulatory networks are an instrumental tool aiding in the discovery of the molecular mechanisms driving diverse diseases, including cancer. However, only noisy observations of the projections of these regulatory networks are typically assayed. In an effort to better estimate regulatory networks from their noisy projections, we formulate a non-convex but analytically tractable optimization problem called OTTER. This problem can be interpreted as relaxed graph matching between the two projections of the bipartite network. OTTER’s solutions can be derived explicitly and inspire a spectral algorithm, for which we provide network recovery guarantees. We also provide an alternative approach based on gradient descent that is more robust to noise compared to the spectral algorithm. Interestingly, this gradient descent approach resembles the message passing equations of an established gene regulatory network inference method, PANDA. Using three cancer-related data sets, we show that OTTER outperforms state-of-the-art inference methods in predicting transcription factor binding to gene regulatory regions. To encourage new graph matching applications to this problem, we have made all networks and validation data publicly available.

Introduction

Bipartite networks are studied across disciplines ranging from machine learning (Yamanishi 2009), ecology, and economics to biology. They focus on the interaction of different types of nodes like vertices in different graphs, pollinators and plants, countries and products, or compounds and proteins. Another prominent example are gene regulatory networks consisting of transcription factors (TFs) and genes (see Fig. 1). These are fundamental objects of study in molecular biology and their analysis provides insights into mechanisms underlying the progression of various diseases, including cancer (Lopes-Ramos et al. 2018; Burkholz and Quackenbush 2021; Lopes-Ramos et al. 2020).

Figure 1:

Figure 1:

Gene regulation. A. Transcription factors (TFs) are represented by green, blue, and yellow objects that bind to the genome (gray band) in vicinity of the start site of a gene (black arrow) to regulate its expression. B. Representation of A as bipartite gene regulatory network.

Often, we do not observe the bipartite network W (e.g. representing TF–gene interactions) directly but instead have information about its two associated projections WWT (TF–TF cooperation) and WTW (gene–gene interactions, see Fig. 2). The general objective of this work is to infer W based on noisy observations of its projections PWWT and CWTW. We formulate this task as a non-convex optimization problem, OTTER (Optimize to Estimate Regulation). It is related to inexact graph matching, as it seeks agreement between two graphs P and C. W could be interpreted as relaxed permutation matrix that matches nodes in P (TFs) with nodes in C (genes). As relaxed graph matching, OTTER is theoretically tractable but the solutions are non-unique, since information about W is lost as a consequence of projecting. To select a solution, we need a good initial guess W0 of the bipartite network as input in addition to P and C.

Figure 2:

Figure 2:

The unknown bipartite network W is inferred from its observed projections P and C.

Our first contribution is to fully characterize OTTER’s solution space, which depends on the spectral decomposition of C and P. Hence, two natural choices to solve OTTER are (1) a spectral algorithm and (2) gradient descent. For both, we provide theoretical network recovery guarantees. While the spectral method is robust to small noise, gradient descent is more reliable in higher noise settings, which are common in biological applications. As we show on three benchmark data sets related to gene regulation and cancer, optimizing OTTER using gradient descent outperforms state-of-the-art gene regulatory network inference techniques. Among these techniques is PANDA (Passing Attributes between Networks for Data Assimilation (Glass et al. 2013)), an established GRN inference method. OTTER gradient descent resembles the corresponding message passing updates, and it can therefore be interpreted as simplified, theoretically tractable formulation of PANDA. This formulation enables us to provide network recovery guarantees and analyze the effects of noise on the network reconstruction.

Gene regulation

Next generation genome sequencing technology has revolutionized biomedical research and provides data at an unprecedented scale and speed. The low cost of this technology facilitates large, genome-scale studies which provide new insights into gene regulation. The human genome encodes about 25,000 genes, but not all genes are activated, or ”expressed,” in every cell type. Gene expression distinguishes tissues from each other can make the difference between health and disease, as it controls the production of proteins. These proteins influence higher level cellular functions, which are often altered during the development and progression of different diseases, including cancer. To gain an understanding of the gene regulatory mechanisms perturbed by a disease, it is common practice to infer and compare associated gene regulatory networks (GRNs) (Lao et al. 2015; Lopes-Ramos et al. 2018; Qiu et al. 2018; Yung et al. 2019). In many cases, these networks are weighted, bipartite, and have a representation as matrix W. W consist of two types of nodes – transcription factors (TFs) and genes. A TF is a protein that can bind to the DNA in the vicinity of a gene and regulate its expression. When this occurs, the TF and target gene are linked in the gene regulatory network, see Fig. 1. TFs are also known to cooperatively regulate target genes. One actively studied mechanism by which TFs cooperate is through the formation of protein complexes which then go on to bind to DNA. These TF-TF interactions are an area of active study, and public databases with such information are actively maintained. In this work, we denote the TF-TF cooperativity matrix as P.

Genes that are co-regulated are frequently correlated in their expression levels. We can estimate this co-regulation using a gene-gene co-expression matrix C estimated from gene expression data. This is especially attractive because gene expression is widely available and is context-specific, i.e., it depends on the tissue type, disease, etc. A more detailed explanation of gene regulation is given in the supplement. We also recommend the review article by Todeschini et al. (Todeschini, Georges, and Veitia 2014)).

GRN inference is central to deepening our understanding of diseases on the molecular level, but it is a notoriously difficult problem. Contributions by researchers working in different domains like graph matching could therefore have a great impact. For this reason, we provide benchmark data sets for three human tissues in the cancer domain (Guebila et al. 2020). Their use does not require expertise in molecular biology.

Related work

The OTTER objective is inspired by a state-of-the-art GRN inference method, PANDA (Glass et al. 2013). PANDA integrates multiple data sources through a message passing approach, which is similar to the gradient descent of OTTER. A derivation is given in the supplement. PANDA has been used to investigate gene regulatory relationships in both tissue specific (Sonawane et al. 2017) as well as several disease contexts, including chronic obstructive pulmonary disease (Lao et al. 2015), asthma (Qiu et al. 2018), beta cell differentiation (Yung et al. 2019), and colon cancer (Lopes-Ramos et al. 2018). OTTER can be seen as a theoretically tractable simplification of PANDA, which is amenable to modern optimization techniques and draws connections to graph matching.

Many methods try to infer regulatory relationships solely based on gene expression with two possible (non-exclusive) objectives: structure learning and gene expression prediction. Usually, gene expression prediction makes indirect statements about the interaction structure of variables as well and thus forms an hypothesis about which TFs regulate which genes. TFs are proteins that are created from the mRNA expressed by their corresponding genes. Hence, predicting target gene expression from the expression of the genes coding for the TFs assumes a biologically reasonable structure. The most common and basic approach is to analyse the Pearson correlation (COR) matrix or, if feasible, partial correlations (PARTIAL COR). Spearman correlations usually lead to similar conclusions. Another popular approach is Weighted Gene Co-expression Network Analysis (WGCNA) (Zhang and Horvath 2005; Langfelder and Horvath 2008), in particular the TOM subroutine. It also starts from a gene expression correlation matrix but down-weighs connections if they are not consistent with neighborhood information. Other pruning heuristics take also different types of node similarities resulting from graph embeddings into account (Pio et al. 2020). Alternatives are based on mutual information, where ARACNe (Lachmann et al. 2016) is one of the most commonly used representatives. Among graphical models, mainly Gaussian graphical models are used because the learning algorithms have to scale to a large number of genes. The GLASSO (Friedman, Hastie, and Tibshirani 2008) method is among the best performing candidates and uses LASSO regularization to enforce sparsity. However, it still does not scale to our setting (approximately 20,000–30,000 genes for human tissue), so that we had to omit it from the benchmarks in our experiments. Linear models (Haury et al. 2012) and random forests (Huynh-Thu et al. 2010) have been used for a similar purpose, where TIGRESS (Haury et al. 2012) and GENIE3 (Huynh-Thu et al. 2010) were top scorers at the DREAM5 challenge (Marbach et al. 2012) (although the challenge was somewhat different from the GRN modeling we study here). Both methods have high computational requirements and are less suitable for the human genome. An alternative approach is to treat the binding of TFs to the promoter region of a gene as supervised learning problem (Karimzadeh and Hoffman 2019; Yuan and Bar-Joseph 2019). While such models can be quite accurate, they are limited to the small number of TFs for which the relevant data is available, which is provided by ChIP-seq experiments. Hence, supervised approaches cannot discover new gene regulatory relationships. Transfer learning algorithms can utilize more data from different domains, for instance, GRNs related to mice (Mignone et al. 2019), but might also inherit unrealistic biases. Note that, in contrast, the data required to define OTTER are widely available, related to the relevant domain, and include a much larger set of known TFs.

In addition, OTTER is related to established problems in graph matching (Yan et al. 2016), which have strong theoretical foundations (Jiang et al. 2017; Barak et al. 2019). The quadratic assignment problem (QAP) (Aflalo, Bronstein, and Kimmel 2015) and its variants (Maron and Lipman 2018) have a direct link to OTTER and can support a similar biological theory. Graph matching has broad applications in computer science ranging from machine learning (Cour, Srinivasan, and Shi 2007), pattern matching (Zhou and De la Torre 2016), vision (Berg, Berg, and Malik 2005; Zhou and De la Torre 2013), and protein network alignment (Singh, Xu, and Berger 2008) to social network analysis (Fan 2012). However, it has not been applied to gene regulatory network inference to the best of our knowledge. As we show, simple relaxed graph matching techniques are competitive with established GRN inference methods.

Contributions

1) We pose a novel optimization problem, OTTER, for the inference of bipartite networks in general and gene regulatory networks (GRNs). Importantly, OTTER is analytically tractable. 2) We gain insights into a state-of-the-art GRN inference method, PANDA (Glass et al. 2013), as OTTER gradient descent resembles the related message passing equations. 3) We characterize OTTER’s solution space and derive a spectral algorithm on its basis, for which we give network recovery guarantees. 4) We solve the gradient flow dynamics associated with gradient descent for OTTER. 5) We draw a connection from OTTER to relaxed graph matching and open a new application area for related algorithms. 6) We show that OTTER gradient descent outperforms the current state of the art in GRN inference on three challenging biological data sets related to cancer. 7) We make the processed data publicly available to ease the use for researchers without a computational biology background and to foster further innovation in relaxed graph matching and GRN inference.

OTTER

Biological motivation

Our objective is to infer a gene regulatory network (GRN) represented by a matrix W. Entries wij with larger values indicate a higher probability that TF i regulates gene j. OTTER and PANDA refine an initial guess W0 of a GRN by increasing its consistency with protein-protein interactions P and observed gene expression with correlation matrix C. In our experiments, a TF–gene edge exists in W0 if the sequence motif for that TF is present in the promoter region of the target gene. This information depends only on the human reference genome and provides a reasonable estimate of where TFs bind. Yet, it is context agnostic. TF binding changes between different tissues, allowing cells to assume their specific functions, and can become disrupted by diseases like cancer.

To estimate condition specific GRNs, we solve the OTTER objective. In doing so, we assume that P and C agree partially with the projections of the actual gene regulatory network W. We do not require perfect equality of WWT = P and WTW = C but improve W0 according to the following three central elements of gene regulation: (1) TFs that can bind to the promoter region of a gene are more likely to regulate that gene (WW0) (Spitz and Furlong 2012; Lambert et al. 2018; Ouyang, Zhou, and Wong 2009), (2) genes that are correlated in their expression are more likely to be co-regulated by similar TFs (WTWC) (Lambert et al. 2018; Shi, Fornes, and Wasserman 2018; Hobert 2008), and (3) TFs that interact (for example, by forming complexes) are more likely to target the same genes (WWTP). TF cooperation is often mediated through protein-protein interactions (Spitz and Furlong 2012; Morgunova and Taipale 2017; Deplancke, Alpern, and Gardeux 2016). For example, the TFs Msn2 and Msn4 bind together to form a complex before binding to the DNA of their target genes (Chapal et al. 2019).

This reasoning motivates our general study of bipartite network inference based on observed noisy projections (P and C). As we will show, a considerable amount of information is lost by projecting. This explains partially why GRN inference is challenging. Central to our success is a good initialization W0 and the choice of algorithm that picks a solution among the many different options. Specifically, we study two approaches: (a) a spectral algorithm and (b) a gradient descent variant optimizing the OTTER objective, which we introduce formally in the next section. As we show, the spectral approach has excellent recovery guarantees in low noise settings, while gradient descent is more reliable in high noise applications, which is common in high throughput sequencing data in biology. Gradient descent also has the advantage that it allows us to stay closer to the initial W0 with early stopping. For this reason, it enables us to outperform the state of the art in GRN inference.

Theoretical framework

In this section, we analyse the general problem of learning a bipartite and weighted network with matrix representation Wnp×nc from its symmetric projections Pnp×np and Cnc×nc. By analogy with our motivation of GRN inference, we call the nodes of one type transcription factors (TFs) and of the other type genes. We have np TFs and nc genes, where the number of genes is usually much larger (npnc). Minimizing the following OTTER objective

f(W)=(1λ)4WWTP2+λ4WTWC2+γ2W2 (1)

with respect to W seeks agreement between the projections of W and P and C. λ ∈ [0, 1] denotes a tuning parameter that moderates the influence of P versus C, and γ corresponds to a potential regularization. As we will see later, this choice of regularization compensates for a bias that noise in P and C introduces. In principle, we could choose any matrix norm but limit our following discussion to the Frobenius norm A2:=i=1nj=1maij2=tr(ATA) for a matrix A = (aij). For this choice, gradient descent resembles most closely the related message passing equations of PANDA and we can derive the solutions of the minimization problem.

These solutions depend on the spectral decomposition of P=UpDpUpT and C=VcDcVcT, which exist with respect to orthogonal Up and Vc, as P and C are symmetric. Otherwise, the same results hold for the spectral decomposition of (P + PT)/2 and (C + CT)/2. Dp and Dc are diagonal matrices containing the eigenvalues of the respective matrix. In a slight abuse of notation, we denote with Dp a matrix Dpnp×np and, if convenient, a matrix Dpnc×nc, which is padded with zeros accordingly. Furthermore, let M[np]=(mij)inp,jnp denote a submatrix of a larger matrix M with dimension np × np. Without loss of generality, we assume that the eigenvalues dp,ii of P are indexed in descending order; dp,iidp,jj for i < j. For C however, we require a good matching with P. We therefore assume implicitly that the distance of Dc to Dp is minimized with respect to permutations of the eigenvalues of C, that is DcDp2=minπPDc,πDp2, where P denotes the set of permutations of {1, · · · , nc} and Dc,π the corresponding ordering of eigenvalues on the diagonal. If Dp and Dc show little discrepancy, this will result in the eigenvalues of C being in descending order as well. Now, everything is in place to characterize the solution space 𝒮.

Theorem 1.

For given Pnp×np with P = PT and Cnc×nc with C = CT, for any spectral decomposition P=UpDpUpT and C=VcDcVcT, λ ∈ [0, 1] the minimization problem (1) has solutions W*S with singular value decomposition W*=UpDwVcT, where

dw,ii=max((1λ)dp,ii+λdc,iiγ,0) (2)

for inp. For dw,ii = 0, the corresponding columns of Uw and Vw are not restricted to the eigenvectors of P and C. The eigenvalues of C are ordered such that Dc = Dc,π, where the permutation solves the minimization problem

π=argminπP(λ(1λ)2DpDc,π[np]2λ2Dc,π[np]2+(1λ)γtr(Dc,π[np])). (3)

For Δ:=Dc[np]Dp, we further assume that

1λ2(P2+1nptr(P)2)+1np(tr(Δ))2>Δ2. (4)

Condition (4) is usually met and it is a minor technicality to exclude alternative global minima of Objective (1) that defy our intuition. The nature of these alternatives is discussed in detail in the proof of Thm. 1 in the supplement.

According to Thm. 1, OTTER (Eq. (1)) has at least 2np different solutions. Each column u:i of Up has two optional signs that do not alter the spectral decomposition of P but can lead to a different solution W*. The same applies to columns v:i of Vc. Only the product of corresponding columns (u:i and v:,i) determines the respective solution W*, as we have wij*=kdw,kkuikvjk. This leaves us with 2np alternatives. If the (non-zero) spectra are not simple, such that some eigenspaces have multiple choices of basis functions, we have additional degrees of freedom in constructing the solutions.

As a consequence, we face a model selection problem and require additional information to make an informed decision. In the following, we propose two natural algorithmic choices to identify a solution: (a) a spectral approach based on Thm. 1 and (b) gradient descent minimizing OTTER.

Both rely on additional input W0np×nc, an initial guess of a gene regulatory matrix. The choice of W0 is crucial for the performance of both algorithms. To understand some of their advantages and limitations, we provide theoretical recovery results when W0 is a random perturbation of the correct W and compare both algorithms on synthetic data with increasing levels of noise.

A spectral method for solving OTTER

Assuming that W0 provides good evidence, our first proposal for a network inference algorithm selects the closest solution (W^S) to W0 in a spectral approach: W^=minWSWW02. If P and C have simple spectra so that the non-zero eigenvalues correspond to 1-dimensional eigenspaces, the solution to this minimization problem can be computed easily. Note that this assumption is satisfied in our applications. From our previous derivation of the solution space, we know that the only ambiguity lies in the sign of the eigenvectors or, equivalently, the singular eigenvalues. Essentially, for fixed spectral decomposition P=UpDpUpT and C=VcDcVcT, our candidate solutions are of the form W=UpDwDsVcT, where Ds contains the unknown sign information. Ds is a diagonal matrix with entries ds,ii ∈ {−1, 1} on the diagonal. These are our only degrees of freedom. Hence, our problem turns into

minWSWW02=minds,ii{1,1}DwDsUpTW0Vc2 (5)

For simplicity, we write M0=UpTW0Vc. The solution W^ is unique and given by W^=UpD^wVcT with d^w,ii=dw,iisign(m0,ii), where dw,ii is defined as in Thm. 1 and sign(x) = 1 for x ≥ 0 and sign(x) = −1 for x < 0.

An important question in many applications is how well this approach performs under noise. First, we study a simplified scenario, in which only W0 is noise corrupted so that we know the singular values Dw, as we can deduce them from the correct P and C. Even in this simplified case, perfect recovery of W is unlikely for large-scale problems, as the next proposition states. Let Φ denote the cumulative distribution function (cdf) of a standard normal and X ∼ Ber(p) a Bernoulli random variable with success probability p.

Proposition 2.

Assume that we observe P = W*TW*, C = W*W*T, and W0 = W* + E for a true underlying W*np×np and noise Enp×np with independent identically normally distributed components eij~N(0,σ2). Further assume that P and C have a simple spectrum {d1,,dnp}. Then, for the spectral approach W^=argminWSWW02 with γ = 0, the recovery loss is distributed as W^W*2=4i=1npdi2Ri, where Ri ∼ Ber(Φ(−di)) for di > 0 and Ri = 0 for di = 0 are independent. For any ϵ > 0, the following holds with the usual Chernoff bound:

P(W^W*2ϵ)1exp(ϵμϵ4δlog(ϵμ)),

where μ=ipi and δ=1maxi(dw,ii2) for ϵμ and δ=1mini(dw,ii2) otherwise.

The proof is given in the supplement. The insight that Ri ∼ Ber(Φ(−di)) allows us also to analyze the probability of perfect recovery (ϵ = 0). We have P(W^=W*)=i=1np(1Φ(di/σ)). In our examples, np = 1636 and dmin ≈ 0.0001. To achieve a probability of at least 0.5, we could allow for a noise variance of σ ≈ 3 · 10−5. In many applications, this would be a reasonable range, considering that we have npnc ≈ 4.4 · 107 matrix entries.

Yet, biological data is known to be very noisy. In addition, we also expect high noise in P and C. To compensate for this additional noise, we need regularization.

Regularization

To motivate the need for regularization (γ > 0), we show that, depending on the source of the noise, the spectrum of P and C becomes biased. Typical noise matrices Ep and Ec have iid entries with zero mean, variance σp/c2, and a symmetric distribution. They could distort the true projections W*W*T and W*TW* as P = (W* +Ep)(W* +Ep)T and C = (W* +Ec)T(W* +Ec) or P=W*W*T+EpEpT and C=W*TW*+EcTEc, respectively. In both cases, P and C become biased as E(P)=W*W*T+σp2I and γ=λσc2+(1λ)σp2, where I denotes the respective identity matrix. The choice γ=λσc2+(1λ)σp2 in Eq. (1) can compensate for this spectral shift.

It should be noted that, Thm. 1 states that such a l2-regularization alters the solutions to Problem (1) in two ways. Not only are the singular values of W* shifted by −γ to compensate for the biases introduced by the noise, also the matching of the eigenvalues of P and C is influenced by the additional penalty γ(1λ)tr(Dc,π[np]) in Eq. (3). Consequently, it may be optimal to pair the eigenvalues of P with smaller eigenvalues of C rather than larger ones if γ is large.

The spectral method can be powerful in a setting in which noise is well controlled such that our assumptions are met approximately. Our second solution proposal, gradient descent, however, gives us more tuning options, including the step size and early stopping, that will allow us to stay closer to the initial guess W0.

Gradient descent for solving OTTER

The message passing equations of PANDA resemble a gradient descent procedure minimizing Objective (1). We explain this relationship in detail in the supplement. In our experiments, we used the ADAM method (Kingma and Ba 2014) for gradient descent, but alternatives are equally applicable. To better understand the approach from a theoretical perspective and reason about its response to noisy data, we take the continuous time approximation (corresponding to infinitesimally small step size) and study the corresponding gradient flow:

τdWdt=f(W)=WWTW+(1λ)PW+λWCγW, (6)

where we set the time unit τ = 1 in the following for simplicity. If the initial W0 has a similar singular value decomposition as a solution, the differential equation decouples and we can solve the resulting one-dimensional ordinary differential equations for the diagonal elements explicitly.

Proposition 3.

For initial W0=UpD0VcT with UpDpUpT and VcDcVcT, the solution of the gradient flow (6) is given by W(t)=UpDtVcT with

dt,ii=sign(d0,ii)dw,ii×12h(dw,ii2t+h1(2d0,ii2/dw,ii21))+1,

where h(x) = tanh(x) if d0,ii2<dw,ii2 and h(x) = coth(x) otherwise.

The proof is provided in the supplement. Note that the square root factor converges to 1 for t → ∞ in both cases. Hence, the final solution inherits the signs sign(d0,kk) of the initialization, which is similar to our spectral approach. Thus, if we start from a reasonable guess W0 that diagonalizes with respect to the same U and V as the global minima, gradient descent will converge to the closest global minimum (for small enough learning rate). For general W0, however, it is important to keep in mind that gradient descent can converge to different solutions, since it optimizes a non-convex objective (Kingma and Ba 2014; Burkholz and Dubatovka 2019). It does not necessarily stay close to our initialization and can even get stuck in local minima. But it also provides us with additional tuning options and early stopping, which will enable us to outperform the state of the art in GRN inference.

Relation to inexact graph matching

As we show in this section, OTTER can also be interpreted as relaxed graph matching. If W solves the OTTER objective for γ = 0 perfectly (f(W) = 0), P and C are its projections, i.e., P = WWT and C = WTW. It follows that P, C, and W also fulfill the relation PW = WWTW = WC. Hence, it would also be reasonable infer a bipartite network from its projections by minimizing the objective

g1(W)=12PWWC2+γ2W2 (7)

(with additional l2-regularization). This is the well known quadratic assignments problem (QAP), a standard objective in graph matching (Aflalo, Bronstein, and Kimmel 2015). In this setting, P and C are usually assumed to have the same dimension (nP = nC). The dimensions can differ for inexact graph matching, but the smaller network is then supposed to be similar to a subgraph of the bigger one. Thus, the minimization is performed under the constraint that W is a permutation matrix. In contrast, we are not interested in a permutation matrix, but in a weighted network Wnp×nc that solves the relaxed QAP. As OTTER, QAP has different solutions and thus solution techniques. Gradient descent and spectral approaches are common choices.

In particular, GRAMPA (Fan et al. 2019a,b) is a variant of QAP with strong recovery guarantees. It adds the term −δ1TW1 to the QAP objective (7), where 1 denotes a vector with all entries equal to one:

g2(W)=12PWWC2+γ2W2δ1TW1 (8)

As a consequence, the GRAMPA minimization problem has a unique solution and becomes explicitly solvable by a spectral approach.

As for OTTER, the spectral approach performs worse in estimating GRNs than the optimization by gradient descent. We therefore only report the latter in our experiments, where we explore the utility of graph matching techniques for GRN inference in comparison with OTTER. The precise gradient descent algorithms minimizing QAP or GRAMPA are detailed in the supplement.

Graph matching can also be studied within the optimal transport framework (Peyré, Cuturi, and Solomon 2016; Titouan et al. 2019). We could formulate the OTTER objective with respect to a nonstandard metric and regularization term. Since we are not searching for stochastic matrices W, this does not serve our purpose and we leave the transfer of related methods to gene regulatory network inference to future explorations.

Experiments

Experiments on synthetic data

To showcase the performance of OTTER for cases in which our assumptions are met and to study the influence of noise, we create synthetic data based on a ground truth W* that we try to recover from noise corrupted inputs W0 = W* + E0, P = (W*+Ep)(W*+Ep)T, and C = (W*+Ec)T(W*+ Ec). All noise entries are Gaussian and independently distributed with e0,ii~N(0,σ02), ep,ii~N(0,σp2), and ec,ii~N(0,σc2). To obtain a realistic ground truth for which we can repeat each experiment 10 times conveniently, we sub-sample (in each repetition) the ChIP-Seq network for the liver tissue to np = 100 and nc = 200. (See the next section for more details.) As this is unweighted, we draw the weights iid from N(105,1). For each network, we use the spectral and the gradient descent version of OTTER and report the obtained recovery error ∥|WW*∥2.

We align the eigenvalues of P and C by arranging them in descending order in the spectral approach. ADAM gradient descent for the OTTER objective is run for 104 steps with the default ADAM parameters, as detailed in the supplement. For both the gradient decent and the spectral approach, we use parameters γ=σp/c2=σp2=σc2 and λ = 0.5.

The results are shown in Fig. 3. For small levels of noise in P and C, the spectral approach performs reliably and better than gradient descent. However, for high σp/c2 gradient descent outperforms the spectral method. Since biological data is inherently noisy, gradient descent seems to be the method of choice. Furthermore, it provides us with additional tuning options that we can leverage to outperform state-of-the-art methods.

Figure 3:

Figure 3:

OTTER recovery error using (top) gradient descent and (bottom) spectral decomposition for artificial networks of size np = 100, nc = 200 and Gaussian noise with variance σp/c2 for P and C and σ02 for W0. Shaded regions correspond to the 0.95 confidence interval and lines to the average over 10 repetitions. The legend applies to both figures.

Experiments on cancer data

The most abundant data source for studying gene regulation is gene expression data. These data are often measured using bulk RNA-sequencing (RNA-seq) with samples corresponding to different individuals.

Datasets and experimental set-up

We obtained bulk RNA-seq data from the Cancer Genome Atlas (TCGA) (Tomczak, Czerwińska, and Wiznerowicz 2015). The data is downloaded from recount2 (Collado-Torres et al. 2017) for liver, cervical, and breast cancer tumors and normalized and filtered as described in the supplement. The corresponding Pearson correlation matrix defines the gene-gene co-expression matrix C consisting of nc = 31,247 genes for breast cancer, nc = 30,181 for cervix cancer and 27,081 for liver cancer. The protein-protein interaction matrix P is derived using laboratory experiments and represents possible interactions; we use the version of (Sonawane et al. 2017) and fill unavailable information with zeros. P consists of np = 1,636 potential TFs. Our initial guess of a gene regulatory network, W0, is derived from the human reference genome. It is almost identical across tissues. It only varies slightly according to the number of genes (nc) included after filtering and normalization. W0 is a binary matrix with w0,ij ∈ {0, 1} where “1” indicates a TF sequence motif in the promoter of the target gene. Sequence motif mapping was performed using the FIMO software (Grant, Bailey, and Noble 2011) from the MEME suite (Bailey et al. 2009) and the GenomicRanges R package (Lawrence et al. 2013). Note that neither W0 or P carry sign information about edge weights so that we cannot infer whether TFs inhibit or activate the expression of a gene. We therefore focus on the prediction of link existence with the understanding that the type of interaction can be estimated post hoc.

Validation of gene regulatory networks is a major challenge. Data from chromatin immunoprecipitation followed by sequencing (ChIP-seq) experiments, which measure the binding of TFs to DNA in the genome, provide a validation standard against which to benchmark our results. Each ChIP-seq experiment assays only one TF. Because of the assay’s relatively high cost, there are only few data sets that have ChIP-seq data for many TFs from the same cells. We used ChIP-seq data from the HeLa cell line (cervical cancer, 48 TFs), HepG2 cell line (liver cancer, 77 TFs) and MCF7 cell line (breast cancer, 62 TFs) available in the ReMap2018 database (Chèneby et al. 2018), a database collection of publicly available ChIP-seq datasets from available studies. This database contains identified ChIP-seq peaks, representing our target TF binding sites. Further details are given in the supplementary material. Based on this data, we measure the performance of link classification on the subnetwork in each tissue that is constrained to the available TFs and report the AUC-ROC (area under the receiver operating characteristic curve) and AUPR (or AUC-PR) (area under the precision recall curve).

Hyperparameter tuning of OTTER was assisted by MATLAB’s bayesopt function utilizing a Gaussian process prior to maximize the joint AUC-PR for breast and cervix cancer, max AUPRbreast · AUPRcervix. Breast and cervix data serve therefore as training data while the liver cancer data is an independent test set. The parameters of all compared methods are reported in the supplementary information.

Results

Table 1 compares the feasible GRN inference and relaxed graph matching methods based on comparison with experimental ChIP-seq binding data. Note that we also report the performance of our initialization W0, which is based on motif data. OTTER GRAD, PANDA, QAP GRAD, and GRAMPA GRAD greatly improve this initial guess and make it tissue specific. Overall, OTTER gradient descent (OTTER GRAD) achieves the best performance on all tissues, in particular, on the liver test set. An enrichment analysis of Gene Ontology terms between networks for healthy and cancerous liver tissue in the supplement provides additional evidence that OTTER GRAD is biologically meaningful.

Table 1:

TF binding prediction for different cancer tissues.

AUC-ROC (AUC)
Method Breast Cervix Liver
COR 0.5900 0.5758 0.5637
Partial COR 0.5366 0.5209 0.5175
ARACNE 0.6150 0.5234 0.5636
GENIE3 0.4818 0.4832 0.4846
TIGRESS 0.4945 0.4808 0.5018
OTTER spectral 0.5787 0.5420 0.5345
GRAMPA grad 0.6301 0.6328 0.6072
QAP grad 0.6373 0.6287 0.6081
WGCNA TOM 0.6146 0.5842 0.5946
W0 0.6282 0.6261 0.5982
PANDA 0.6739 0.6642 0.6211
OTTER grad 0.6936 0.6833 0.6600

AUC-PR (AUPR)

COR 0.2772 0.2247 0.3057
Partial COR 0.2361 0.1952 0.2525
ARACNE 0.2858 0.2027 0.2986
GENIE3 0.2064 0.1836 0.2437
TIGRESS 0.2088 0.1845 0.2523
OTTER spectral 0.2555 0.2024 0.2614
GRAMPA grad 0.3162 0.2763 0.3223
QAP grad 0.3215 0.2637 0.3425
WGCNA TOM 0.2834 0.2229 0.3140
W0 0.2865 0.2523 0.3045
PANDA 0.3481 0.2960 0.3503
OTTER grad 0.3752 0.3179 0.3746

The symbol † indicates that binding predictions were made only for TFs with ChIP-seq data due to high computational demands. The highest score for each data set is shown in bold.

Interestingly, ADAM gradient descent solving alternative graph matching problems, i.e., QAP GRAD and GRAMPA GRAD, achieve better results than established GRN inference algorithms, even though they were not originally designed for this purpose. They succeed based on similar hyperparameters as OTTER GRAD.

In general, we observe better performance for the methods that incorporate additional biological evidence such as (transformed) protein-protein interactions and binding motifs, even though P and W0 are not tissue-specific. A reason for this is that correlations in gene expression can be caused by many factors. Many TFs are expressed at very low levels but strongly activate their target genes, obscuring correlations between TFs and their targets. Hence, graph matching approaches are a promising alternative to models that make predictions based on gene expression alone.

Discussion

We formulated the inference of a bipartite network from its two projections as a non-convex but analytically tractable optimization problem, OTTER. The projections alone do not provide enough information for network inference, as OTTER has multiple solutions that we derived explicitly. We proposed two natural inference algorithms for model selection, a spectral approach and gradient descent, and derived sufficient conditions for network recovery. Both rely on an additional initial guess of the bipartite network, W0, which has to be close to the correct network to guarantee good network recovery. We find the spectral approach to be more reliable in low noise settings, while gradient descent seems to be more robust with respect to higher amounts of noise and therefore more suitable for our application of interest: gene regulatory network inference.

As we have shown, gradient descent also resembles in part an established gene regulatory network inference method, PANDA. OTTER can therefore be interpreted as a theoretically tractable simplification of PANDA that provides an intriguing connection to relaxed graph matching. OTTER also outperforms state-of-the art gene regulatory network inference approaches on real world data sets corresponding to three human cancer tissues. We make these data sets publicly available to benchmark the use of general graph matching algorithms for gene regulatory network inference (Guebila et al. 2020). As highlighted, relaxed graph matching approaches apply to this setting and achieve competitive performance. They have the advantage that they can integrate additional information about a gene regulatory network in the form of W0 and protein interactions P. We therefore see great potential in transferring other graph matching techniques to gene regulatory network inference in future investigations.

Supplementary Material

Supplement_proofs

Acknowledgements

The results shown here are in part based upon data generated by the TCGA Research Network: https://www.cancer.gov/tcga. DW, MG, CL, JQ, RB were supported by a grant from the US National Cancer Institute (1R35CA220523). JP acknowledges support from the US National Heart, Lung, and Blood Institute (NHLBI): K25HL140186 and KG from the K25 grant: K25HL133599.

We thank Alkis Gotovos for helpful feedback on the manuscript.

Footnotes

We provide a tutorial to walk the users through the usage of OTTER in R (https://netzoo.github.io/netZooR/).

The raw and processed data are accessible through netZoo (https://netzoo.github.io/zooanimals/otter/) and the networks can be downloaded from the GRAND database (https://grand.networkmedicine.org/cancers/).

Supplement

A supplementary file with proofs of theorems and details of the presented algorithms is available at https://netzoo.s3.useast-2.amazonaws.com/supData/otter/OtterAAAI2021-12.pdf.

Data and code availability

OTTER is available in R, Python, and MATLAB through the netZoo packages: netZooR v0.7 (https://github.com/netZoo/netZooR), netZooPy v0.7 (https://github.com/netZoo/netZooPy), and netZooM v0.5 (https://github.com/netZoo/netZooM).

References

  1. Aflalo Y; Bronstein A; and Kimmel R 2015. On convex relaxation of graph isomorphism. Proceedings of the National Academy of Sciences 112(10): 2942–2947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Bailey TL; Boden M; Buske FA; Frith M; Grant CE; Clementi L; Ren J; Li WW; and Noble WS 2009. MEME SUITE: tools for motif discovery and searching. Nucleic acids research 37(suppl 2): W202–W208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Barak B; Chou C-N; Lei Z; Schramm T; and Sheng Y 2019. (Nearly) Efficient Algorithms for the Graph Matching Problem on Correlated Random Graphs. In Advances in Neural Information Processing Systems 32, 9190–9198. NIPS 2019. [Google Scholar]
  4. Berg AC; Berg TL; and Malik J 2005. Shape matching and object recognition using low distortion correspondences. In 2005 IEEE Computer Society Conference on Computer Vision and Pattern Recognition (CVPR’05), volume 1, 26–33 vol. 1. [Google Scholar]
  5. Burkholz R; and Dubatovka A 2019. Initialization of Re-LUs for Dynamical Isometry. In Advances in Neural Information Processing Systems 32, 2385–2395. NeurIPS’ 2019. [Google Scholar]
  6. Burkholz R; and Quackenbush J 2021. Cascade size distributions: Why they matter and how to compute them efficiently. In Proceedings of the AAAI Conference on Artificial Intelligence. AAAI’ 2021. [Google Scholar]
  7. Chapal M; Mintzer S; Brodsky S; Carmi M; and Barkai N 2019. Resolving noise–control conflict by gene duplication. PLoS Biology 17(11). [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Chèneby J; Gheorghe M; Artufel M; Mathelier A; and Ballester B 2018. ReMap 2018: an updated atlas of regulatory regions from an integrative analysis of DNA-binding ChIP-seq experiments. Nucleic acids research 46(D1): D267–D275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Collado-Torres L; Nellore A; Kammers K; Ellis SE; Taub MA; Hansen KD; Jaffe AE; Langmead B; and Leek JT 2017. Reproducible RNA-seq analysis using recount2. Nature biotechnology 35(4): 319–321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Cour T; Srinivasan P; and Shi J 2007. Balanced Graph Matching. In Schölkopf B; Platt JC; and Hoffman T, eds., Advances in Neural Information Processing Systems 19, 313–320. NIPS 2007. [Google Scholar]
  11. Deplancke B; Alpern D; and Gardeux V 2016. The Genetics of Transcription Factor DNA Binding Variation. Cell 166(3): 538–554. [DOI] [PubMed] [Google Scholar]
  12. Fan W 2012. Graph Pattern Matching Revised for Social Network Analysis. In Proceedings of the 15th International Conference on Database Theory, ICDT ‘12, 8–21. [Google Scholar]
  13. Fan Z; Mao C; Wu Y; and Xu J 2019a. Spectral Graph Matching and Regularized Quadratic Relaxations I: The Gaussian Model.
  14. Fan Z; Mao C; Wu Y; and Xu J 2019b. Spectral Graph Matching and Regularized Quadratic Relaxations II: Erdős-Rényi Graphs and Universality.
  15. Friedman J; Hastie T; and Tibshirani R 2008. Sparse inverse covariance estimation with the graphical lasso. Biostatistics 9(3): 432–441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Glass K; Huttenhower C; Quackenbush J; and Yuan G-C 2013. Passing messages between biological networks to refine predicted interactions. PloS one 8(5). [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Grant CE; Bailey TL; and Noble WS 2011. FIMO: scanning for occurrences of a given motif. Bioinformatics 27(7): 1017–1018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Guebila MB; Lopes-Ramos C; Sonawane A; Burkholz R; Weighill D; Platig J; Kuijjer M; Glass K; and Quackenbush. 2020. GRAND: A database of gene regulatory models across human conditions. https://grand.networkmedicine.org/. [DOI] [PMC free article] [PubMed]
  19. Haury A-C; Mordelet F; Vera-Licona P; and Vert J-P 2012. TIGRESS: trustful Inference of Gene REgulation using Stability Selection. BMC systems biology 6: 145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Hobert O 2008. Gene regulation by transcription factors and microRNAs. Science 319(5871): 1785–1786. [DOI] [PubMed] [Google Scholar]
  21. Huynh-Thu VA; Irrthum A; Wehenkel L; and Geurts P 2010. Inferring Regulatory Networks from Expression Data Using Tree-Based Methods. PLOS ONE 5(9): 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Jiang B; Tang J; Ding C; Gong Y; and Luo B 2017. Graph Matching via Multiplicative Update Algorithm. In Advances in Neural Information Processing Systems 30, 3187–3195. NIPS 2017. [Google Scholar]
  23. Karimzadeh M; and Hoffman MM 2019. Virtual ChIP-seq: predicting transcription factor binding by learning from the transcriptome. bioRxiv doi: 10.1101/168419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Kingma D; and Ba J 2014. Adam: A Method for Stochastic Optimization. International Conference on Learning Representations. [Google Scholar]
  25. Lachmann A; Giorgi FM; Lopez G; and Califano A 2016. ARACNe-AP: gene network reverse engineering through adaptive partitioning inference of mutual information. Bioinformatics 32(14): 2233–2235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Lambert SA; Jolma A; Campitelli LF; Das PK; Yin Y; Albu M; Chen X; Taipale J; Hughes TR; and Weirauch MT 2018. The Human Transcription Factors. Cell 172(4): 650–665. [DOI] [PubMed] [Google Scholar]
  27. Langfelder P; and Horvath S 2008. WGCNA: an R package for weighted correlation network analysis. BMC Bioinformatics 1: 559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Lao T; Glass K; Qiu W; Polverino F; Gupta K; Morrow J; Mancini JD; Vuong L; Perrella MA; Hersh CP; et al. 2015. Haploinsufficiency of Hedgehog interacting protein causes increased emphysema induced by cigarette smoke through network rewiring. Genome medicine 7(1): 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Lawrence M; Huber W; Pages H; Aboyoun P; Carlson M; Gentleman R; Morgan MT; and Carey VJ 2013. Software for computing and annotating genomic ranges. PLoS computational biology 9(8). [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Lopes-Ramos CM; Kuijjer M; Glass K; DeMeo D; and Quackenbush J 2020. Abstract 6569: Regulatory networks of liver carcinoma reveal sex specific patterns of gene regulation. Cancer Research 80: 6569–6569. [Google Scholar]
  31. Lopes-Ramos CM; Kuijjer ML; Ogino S; Fuchs CS; DeMeo DL; Glass K; and Quackenbush J 2018. Gene regulatory network analysis identifies sex-linked differences in colon cancer drug metabolism. Cancer research 78(19): 5538–5547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Marbach D; Costello JC; Küffner R; Vega NM; Prill RJ; Camacho DM; Allison KR; Aderhold A; Bonneau R; Chen Y; et al. 2012. Wisdom of crowds for robust gene network inference. Nature methods 9(8): 796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Maron H; and Lipman Y 2018. (Probably) Concave Graph Matching. In Advances in Neural Information Processing Systems 31, 408–418. NIPS 2018. [Google Scholar]
  34. Mignone P; Pio G; D’Elia D; and Ceci M 2019. Exploiting transfer learning for the reconstruction of the human gene regulatory network. Bioinformatics 36(5): 1553–1561. [DOI] [PubMed] [Google Scholar]
  35. Morgunova E; and Taipale J 2017. Structural perspective of cooperative transcription factor binding. Curr Opin Struct Biol 47: 1–8. [DOI] [PubMed] [Google Scholar]
  36. Ouyang Z; Zhou Q; and Wong WH 2009. ChIP-Seq of transcription factors predicts absolute and differential gene expression in embryonic stem cells. Proceedings of the National Academy of Sciences 106(51): 21521–21526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Peyré G; Cuturi M; and Solomon J 2016. Gromov-Wasserstein Averaging of Kernel and Distance Matrices. In Proceedings of the 33rd International Conference on International Conference on Machine Learning - Volume 48, ICML16, 2664–2672. [Google Scholar]
  38. Pio G; Ceci M; Prisciandaro F; and Malerba D 2020. Exploiting causality in gene network reconstruction based on graph embedding. Machine Learning 109(6): 1231–1279. [Google Scholar]
  39. Qiu W; Guo F; Glass K; Yuan GC; Quackenbush J; Zhou X; and Tantisira KG 2018. Differential connectivity of gene regulatory networks distinguishes corticosteroid response in asthma. Journal of Allergy and Clinical Immunology 141(4): 1250–1258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Shi W; Fornes O; and Wasserman WW 2018. Gene expression models based on transcription factor binding events confer insight into functional cis-regulatory variants. Bioinformatics 35(15): 2610–2617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Singh R; Xu J; and Berger B 2008. Global alignment of multiple protein interaction networks with application to functional orthology detection. Proceedings of the National Academy of Sciences 105(35): 12763–12768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Sonawane AR; Platig J; Fagny M; Chen C-Y; Paulson JN; Lopes-Ramos CM; DeMeo DL; Quackenbush J; Glass K; and Kuijjer ML 2017. Understanding tissue-specific gene regulation. Cell reports 21(4): 1077–1088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Spitz F; and Furlong EEM 2012. Transcription factors: from enhancer binding to developmental control. Nat Rev Genet 13(9): 613–626. [DOI] [PubMed] [Google Scholar]
  44. Titouan V; Courty N; Tavenard R; Laetitia C; and Flamary R 2019. Optimal Transport for structured data with application on graphs. In Chaudhuri K; and Salakhutdinov R, eds., Proceedings of the 36th International Conference on Machine Learning, volume 97, 6275–6284. [Google Scholar]
  45. Todeschini A-L; Georges A; and Veitia RA 2014. Transcription factors: specific DNA binding and specific gene regulation. Trends in genetics 30(6): 211–219. [DOI] [PubMed] [Google Scholar]
  46. Tomczak K; Czerwińska P; and Wiznerowicz M 2015. The Cancer Genome Atlas (TCGA): an immeasurable source of knowledge. Contemporary oncology 19(1A): A68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Yamanishi Y 2009. Supervised Bipartite Graph Inference. In Koller D; Schuurmans D; Bengio Y; and Bottou L, eds., Advances in Neural Information Processing Systems 21, 1841–1848. NIPS 2009. [Google Scholar]
  48. Yan J; Yin X-C; Lin W; Deng C; Zha H; and Yang X 2016. A Short Survey of Recent Advances in Graph Matching. In Proceedings of the 2016 ACM on International Conference on Multimedia Retrieval, ICMR ‘16, 167–174. [Google Scholar]
  49. Yuan Y; and Bar-Joseph Z 2019. Deep learning for inferring gene relationships from single-cell expression data. Proceedings of the National Academy of Sciences 116(52): 27151–27158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Yung T; Poon F; Liang M; Coquenlorge S; McGaugh EC; Hui C.-c.; Wilson MD; Nostro MC; and Kim T-H 2019. Sufu-and Spop-mediated downregulation of Hedgehog signaling promotes beta cell differentiation through organ-specific niche signals. Nature communications 10(1): 1–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Zhang B; and Horvath S 2005. A general framework for weighted gene coexpression network analysis. In Statistical Applications in Genetics and Molecular Biology 4: Article 17. [DOI] [PubMed] [Google Scholar]
  52. Zhou F; and De la Torre F 2013. Deformable Graph Matching. In Proceedings of the 2013 IEEE Conference on Computer Vision and Pattern Recognition, CVPR ‘13, 2922–2929. USA: IEEE Computer Society. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Zhou F; and De la Torre F 2016. Factorized Graph Matching. IEEE Transactions on Pattern Analysis and Machine Intelligence 38(9): 1774–1789. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplement_proofs

Data Availability Statement

OTTER is available in R, Python, and MATLAB through the netZoo packages: netZooR v0.7 (https://github.com/netZoo/netZooR), netZooPy v0.7 (https://github.com/netZoo/netZooPy), and netZooM v0.5 (https://github.com/netZoo/netZooM).

RESOURCES