Abstract
The combinatorial effect of genetic variants is often assumed to be additive. Although genetic variation can clearly interact non-additively, methods to uncover epistatic relationships remain in their infancy. We develop low-signal signed iterative random forests to elucidate the complex genetic architecture of cardiac hypertrophy. We derive deep learning-based estimates of left ventricular mass from the cardiac MRI scans of 29,661 individuals enrolled in the UK Biobank. We report epistatic genetic variation including variants close to CCDC141, IGF1R, TTN, and TNKS. Several loci where variants were deemed insignificant in univariate genome-wide association analyses are identified. Functional genomic and integrative enrichment analyses reveal a complex gene regulatory network in which genes mapped from these loci share biological processes and myogenic regulatory factors. Through a network analysis of transcriptomic data from 313 explanted human hearts, we found strong gene co-expression correlations between these statistical epistasis contributors in healthy hearts and a significant connectivity decrease in failing hearts. We assess causality of epistatic effects via RNA silencing of gene-gene interactions in human induced pluripotent stem cell-derived cardiomyocytes. Finally, single-cell morphology analysis using a novel high-throughput microfluidic system shows that cardiomyocyte hypertrophy is non-additively modifiable by specific pairwise interactions between CCDC141 and both TTN and IGF1R. Our results expand the scope of genetic regulation of cardiac structure to epistasis.
Main
Heart disease is closely tied to the structure of the heart1. Heart failure, a syndrome characterized by increased pressure within, or decreased output from, the heart is influenced by structural features including atrial and ventricular chamber size and wall thickness2–5. Left ventricular hypertrophy – increased thickness of the left ventricle (LV) – can be the result of mendelian genetic diseases like hypertrophic cardiomyopathy6 but is also a complex phenotypic trait influenced by multiple factors, genetic and environmental. Progressive LV hypertrophy carries significant independent risk for incident heart failure, atrial arrhythmia, and sudden death7–10, highlighting the need to understand genetic determinants of cardiac phenotype.
Recent discoveries leveraging cardiac magnetic resonance imaging in the UK Biobank (UKBB) have revealed that cardiac structure is in part determined by complex genetics11–14. Common genetic variants, many located near genetic loci associated with dilated cardiomyopathy and heart failure, have been found to influence LV size and systolic function11. Further, specific genetic variants that influence LV trabeculation have been shown to impact systolic function and overall risk of cardiomyopathy13. However, these variants remain inadequate to explain the total heritable disease risk15. Indeed, common genetic variants rarely act independently and additively as modeled by most genome-wide association studies (GWAS)16. There is growing biological and clinical evidence17 to support a disease risk model in which multiple genes interact non-additively with each other through epistasis18,19. While some computational studies estimated a minor average epistatic component compared to the additive component within the total genetic variance, these epistatic variance estimates exhibit a large trait-to-trait variation20. In addition, it’s important to distinguish between the concepts of statistical epistasis, estimated through variance components and influenced by allele frequencies, and biological epistasis (e.g., gene actions), which is independent from allele frequencies21. Recent work has shown that common genetic variation influences susceptibility and expressivity of hypertrophic cardiomyopathy14. This raises the possibility that common epistatic interactions drive cardiac phenotype, holding significant potential for uncovering disease mechanisms and developing potential therapeutic strategies.
Several computational and experimental challenges need to be resolved to allow robust identification of epistasis. First, the combinatorial nature of possibly high-order interactions makes an exhaustive search computationally intractable. To reduce the computational burden and ensure stable discoveries, we developed an approach based on signed iterative random forests22,23 to uncover higher-order (not limited to pairwise) nonlinear interactions in a computationally-tractable manner. Second, many previously reported epistatic relationships were not replicated24,25. To achieve more trustworthy results, we adhered to a new framework for veridical data science26, centered around the principles of predictability, computability, and stability (PCS) and the need for transparent documentation of decisions made in data analysis pipelines. A third challenge is the generally small effect size of common genetic variants15,27 which impedes both the data-driven discovery and functional validation of epistatic interactions. In human biobanks, recent advances in deep-learning-enabled phenotyping28 using cardiac magnetic resonance images have led to more refined phenotypes at larger scales. At the cellular level, high-throughput microfluidic technologies29–31 have been integrated with artificial intelligence-based image analysis of single-cell morphology32 and human induced pluripotent stem cell-derived cardiomyocytes33, opening up new possibilities for rapid, label-free detection of the phenotypic consequences of genetic perturbation.
Results
In contrast to many studies18,20,25 that have investigated the statistical significance or causality of epistasis solely from data, we tackle the aforementioned challenges and conceptual gap between statistical epistasis and biological epistasis21 via a multi-stage approach. This approach begins with a data-driven prioritization of promising statistical epistasis followed by extensive functional interpretations and experimental validations to reliably assess the biological epistasis consistency. More specifically, our methodology includes four major stages: derivation of estimates of LV mass (green boxes, Fig. 1); computational prioritization of epistatic drivers (orange boxes, Fig.1); functional interpretation of the hypothesized epistatic genetic loci (purple boxes, Fig.1); and experimental confirmation of epistasis through perturbation (blue boxes, Fig. 1).
Deep learning of UK Biobank cardiac imaging quantifies left ventricular hypertrophy
We accessed all cardiac magnetic resonance images from the UKBB substudy (44,503 people at the time of this analysis)34. We focused on the largest ancestry subset of 29,661 unrelated individuals (summary characteristics in Supplementary Table 1) and analyzed the most recent image per individual. We leveraged a recent deep learning model28 to quantify LV hypertrophy from these 29,661 multislice cine magnetic resonance images (Fig. 2a). A fully convolutional network had been previously trained for image segmentation and was evaluated on manual pixelwise-annotations of images from 4,875 UKBB participants28. This fully convolutional network learns features across five different resolutions through sequential convolutional layers interspersed with non-linearities, and has displayed accurate performance compared to cardiac segmentation by human experts28. Using this segmentation model, we extracted areas of the LV chamber wall in each slice of the short axis image at the end of diastole. Areas extracted from each image slice in the same image stack were then integrated to calculate the heart muscle volume, which we converted to the LV mass using a standard density of 1.05 g/mL35. This was normalized by body surface area, estimated using the Du Bois formula36, to obtain the LV mass index (LVMi, Extended Data Fig. 1). Details regarding this analysis can be found in Methods.
Low-signal signed iterative random forests prioritize epistatic genetic loci
We developed low-signal signed iterative random forests (lo-siRF, Fig. 2a-2e) to prioritize statistical epistatic interactions from the extracted LV mass and single-nucleotide variants (SNVs) from UKBB. Given the inherent low signal-to-noise ratio and aforementioned challenges, lo-siRF aims to recommend reliable candidate interactions for experimental validation rather than directly assessing claims of statistical significance from data. This prioritization pipeline is guided by the PCS framework26 and builds upon signed iterative random forests22,23, a computationally-tractable algorithm to extract predictive and stable nonlinear higher-order interactions that frequently co-occur along decision paths in a random forest. More specifically, lo-siRF proceeds through four steps:
Dimension reduction (Fig. 2b): we combined the results of two initial genome-wide association studies, implemented via PLINK37 and BOLT-LMM38 (Extended Data Fig. 2, Extended Data 1) to reduce the interaction search space from 15 million imputed variants down to 1405 variants (Extended Data 2). Details can be found in the Methods section Lo-siRF step 1: Dimension reduction of variants via genome-wide association studies.
Binarization (Fig. 2c): we partitioned the LV mass measurements into high, middle, and low categories using three different partitioning schemes (Supplementary Table 2). The partitioning enabled us to transform the original low-signal regression problem for a continuous trait into a relatively easier binary classification task for predicting individuals with high versus low LV mass measurements (omitting the middle category). This transformation is necessary to obtain a sufficient prediction signal, ensuring that the model indeed captures pertinent information about reality (Supplementary Table 3). Further justification and details on the partitioning can be found in the Methods section Lo-siRF step 2: Binarization of the left ventricular mass phenotype.
Prediction (Fig. 2d): we trained a signed iterative random forest using the 1405 GWAS-filtered SNVs to predict the binarized LV mass measurements. The learnt model yields on average the highest (balanced) classification accuracy (55%), area under the receiver operator characteristic (0.58), and area under the precision-recall curve (0.57) compared to other common machine learning prediction algorithms (Supplementary Table 4). Details about the model and prediction check can be found in the Methods section Lo-siRF step 3: Prediction.
Prioritization (Fig. 2e): we developed a stability-driven feature importance score (Extended Data Fig. 3), which leveraged the fitted signed iterative random forest and a permutation test, to aggregate SNVs into genetic loci and prioritize interactions between genetic loci. This importance score provides the necessary new interpretable machine learning ingredient to complete the lo-siRF discovery pipeline. Details can be found in the Methods section Lo-siRF step 4: Prioritization.
Additional discussion of the philosophy and modeling decisions driving lo-siRF can be found in Supplementary Note 1, an interactive HTML webpage hosted at https://yu-group.github.io/epistasis-cardiac-hypertrophy/. The webpage also provides a comparison of lo-siRF to alternative epistasis detection methods, including an exhaustive regression-based pairwise interaction search39,40 and MAPIT41, demonstrating the challenges and limitations of existing methods for analyzing low-signal, complex phenotypes.
Lo-siRF identified six genetic risk loci that exhibited stable and reliable associations with LV mass (Fig. 2f). Because these loci are either located within a gene body or in between two genes (Fig. 3a), for convenience we denote these loci by their nearest genes. Notably, out of the six loci, three (TTN, CCDC141, and IGF1R) were prioritized by lo-siRF as epistatic loci. These loci not only interact with other loci, but also marginally affect LV mass. The other three lo-siRF-prioritized loci are LOC157273;TNKS, MIR588;RSPO3, and LSP1. The LOC157273;TNKS locus is located within the intergenic region between genes LOC157273 and TNKS (semicolon indicates intergenic region). This locus was prioritized by lo-siRF to be hypostatic (i.e., effects are deemed stable by lo-siRF only when interacting with the CCDC141 locus). Interestingly, all three identified epistatic interactions involved the CCDC141 locus (Fig. 3a, green links in circle 1). Furthermore, while the MIR588;RSPO3 and LSP1 loci lacked evidence for epistasis by lo-siRF, they were each identified to be marginally associated with LV mass. The specific prioritization order of these loci can be found in Supplementary Table 5, and details regarding the direction or sign of the interactions can be found in Supplementary Note 1. In total, lo-siRF identified 283 SNVs located within the six loci (Extended Data 3, Extended Data Fig. 4). Ninety percent of the 283 SNVs have previously been shown to harbor multiple distinct cardiac function associations42 in phenome-wide analyses (e.g., pulse rate, Extended Data 3), suggesting a strong likelihood that these lo-siRF-prioritized loci contribute to determining cardiac structure and function.
Considering the correlations between LV hypertrophy and hypertension43, we evaluated whether these identified variants affect LV mass through regulating blood pressure. Specifically, we repeated the lo-siRF analysis using only the subset of UKBB individuals without hypertension (details in Methods). All previously highlighted loci and interactions maintained priority in this non-hypertensive subset, except for the MIR588;RSPO3 locus (Fig. 2f) which was not stably prioritized across all three binarization thresholding schemes. Additionally, none of the lo-siRF-prioritized variants showed a strong marginal association with hypertension, failing to meet the genome-wide (p < 5E-8) and even the suggestive (p < 1E-5) significance level. However, the MIR588;RSPO3 locus with lead SNV rs2022479 gave the smallest p-value of 5E-5, which may suggest a possible pleiotropic effect of MIR588;RSPO3 on both LV hypertrophy and blood pressure. In brief, while we cannot completely rule out pleiotropy, the highly stable prioritization of all three epistatic interactions in both analyses with and without hypertensive individuals suggest that the identified epistases on LV mass is not solely driven by blood pressure (additional discussion in Supplementary Note 1).
Loci associated with left ventricular mass exhibit regulatory enrichment
We performed functional mapping and annotation (FUMA)44 for the 283 lo-siRF-prioritized SNVs (Fig. 1, purple and Fig. 3). For linkage disequilibrium (LD), we used a default threshold of r2 = 0.6 and chose the UKBB release 2b reference panel created for British and European subjects to match the population group used for lo-siRF prioritization. FUMA identified 572 additional candidate SNVs (Extended Data 4) in strong LD (r2 > 0.6) with any of the 283 lo-siRF-prioritized SNVs, including 492 SNVs from the input GWAS associations (points in Fig. 3a, circle 8) and 80 non-GWAS-tagged SNVs extracted from the selected reference panel (heatmap tracks in Fig. 3a, circle 8). We then assigned these 572 FUMA-extracted candidate SNVs to a lo-siRF-prioritized locus (Fig. 2f) based on the corresponding lo-siRF-prioritized SNV (out of the 283 SNVs), which has the maximum r2 value with the candidate SNV.
The two loci contributing to the top-ranked epistatic interaction by lo-siRF, the CCDC141 and IGF1R loci (Fig. 2f), both showed a significant enrichment of intronic variants relative to the background reference panel (Fig. 3b, Extended Data 5). Over 88% of the SNVs in or in LD with these two loci were mapped to actively transcribed chromatin states (TxWk) or enhancer states (Enh) in left ventricles based on the ChromHMM Core 15-state model45 (Fig. 3a, circle 7). More than 47% and 76% of the identified SNVs in or in LD with the CCDC141 and IGF1R loci, respectively, showed the highest RegulomeDB44,46 categorical score (ranked within category 1 from the 7 main categories). The Combined Annotation-Dependent Depletion (CADD) score47 was used to judge the deleteriousness of prioritized variants (Extended Data 4). As expected, GTEx48 data revealed that 82% of SNVs in or in LD with the IGF1R locus are expression quantitative trait loci (eQTLs) for the gene IGF1R. In contrast, of the SNVs in or in LD with the CCDC141 locus, only 14% are eQTLs for gene CCDC141 and 22% are splicing quantitative trait loci (sQTLs) for gene FKBP7. Furthermore, Hi-C data indicated that all SNVs identified in or in LD with the IGF1R locus are in 3D chromatin interaction with gene SYNM while more than 54% SNVs identified in or in LD with the CCDC141 locus are in 3D chromatin interaction with gene TTN. These known 3D chromatin interactions could suggest a possibility of higher-order interactions between more than two genes.
The CCDC141 and TTN loci exhibit genomic proximity (Fig. 3a). Their interaction, however, does not appear to stem from this proximity. Indeed, the CCDC141 and TTN genes have been individually associated with LV mass49,50. Due to this proximity, previous studies51,52 have assumed CCDC141 as a secondary gene that affects LV mass through the TTN gene expression. However, we found low LD (r2 < 0.6) between any two of the 283 lo-siRF-prioritized SNVs, suggesting that the identified CCDC141-TTN interaction is unlikely driven by non-random LD associations between SNVs in these two loci. In addition, we compared all the epistasis-contributing SNVs that were aggregated to the TTN locus, including both lo-siRF-prioritized SNVs and their LD-linked variants, with the complementary set of TTN-annotated SNVs in lo-siRF. We found that the TTN locus showed a significant depletion of SNVs located close to (<10 kb) the gene CCDC141 (p = 2.38E-9, two-sided Fisher exact test). Similarly, the CCDC141 locus showed a substantially decreased enrichment of SNVs that are close to gene TTN (p = 0.02, two-sided Fisher exact test). These results suggest that although the CCDC141 and TTN loci are located close to each other in the genome, the prioritized epistatic SNVs are located farther apart relative to randomly selected SNVs from the two loci.
In contrast to the CCDC141 and IGF1R loci, the TTN locus showed a significant enrichment of exonic variants and intronic variants that are transcribed into non-coding RNA (ncRNA_intronic, Fig. 3b). Of those exonic variants, 62% are nonsynonymous. This differential enrichment of exonic variants for the TTN locus may suggest a potential epistatic contribution to the structural alterations in the titin protein. Over 90% of SNVs in or in LD with the TTN locus were mapped to actively transcribed states (Tx, TxWk) in left ventricles (Fig. 3a, circle 7). Interestingly, these SNVs were associated with a quiescent chromatin state (Quies) in the right atrium, indicating that the epistatic effects of the TTN locus may be specific to ventricular tissues. Nearly half of SNVs in or in LD with the TTN locus are eQTLs for the gene FKBP7. In addition, 83% of these SNVs are sQTLs for gene FKBP7 or TTN, suggesting a regulatory effect of the TTN locus on the expression and splicing of gene FKBP7. Moreover, the TTN locus was suggested to impact genes PDE11A, RBM45, PRKRA, and DFNB59 through 3D chromatin interactions.
The hypostatic locus LOC157273;TNKS showed a significant enrichment of variants within non-coding RNA regions of exons and introns (Fig. 3b). Over 95% of identified SNVs in or in LD with this locus were mapped to inactive chromatin states (ReprPCWk, Quies) in left ventricles (Fig. 3a, circle 7). This suggests that in the absence of an epistatic partner, the LOC157273;TNKS locus is epigenetically quiescent or repressed by polycomb group proteins. In addition, of all the SNVs in or in LD with this locus, 66% are eQTLs for MFHAS1 or CLDN23 and 22% are in 3D chromatin interaction with gene TNKS.
Functional annotations for the other two lo-siRF-prioritized loci that were marginally associated with LV mass can be found in Extended Data 4 and 5.
Epistatic loci functionally map to twenty-one protein-coding genes
Three strategies, positional, eQTL, and chromatin interaction, mapped the six LV hypertrophy risk loci to 21 protein-coding genes (Fig. 4a). Genes prioritized by eQTL and chromatin interaction mapping are not necessarily located in the corresponding risk locus, but they are linked to SNVs within or in LD with the locus (Fig. 3a). Among the 21 genes, CCDC141 and IGF1R were prioritized by all the three mapping strategies (Fig. 4a), suggesting that these two genes are very likely involved in determining LV mass. Interestingly, none of the SNVs mapped to IGF1R were statistically significant in our GWAS studies using BOLT-LMM and PLINK (Extended Data Fig. 2 and Extended Data 1). Set-based association tests using SKAT-O53 and MAGMA54 also did not identify the IGF1R locus (details in Methods and Supplementary Note 1). This reveals the potential of lo-siRF to identify risk loci that may be overlooked by GWAS. Based on the expression data from GTEx V8, TTN, TNNT3, and SYNM are up-regulated while CLDN23 and MFHAS1 are down-regulated in both heart and muscle tissues (Fig. 4b). In contrast, CCDC141 is up-regulated specifically in heart tissues whereas RSPO3 is down-regulated in heart but up-regulated in muscle tissues (Fig. 4b).
Ten of twenty-one genes mapped from epistatic loci show strong correlations in network analysis
We performed gene ontology (GO) and pathway enrichment analysis on the 21 genes mapped from lo-siRF loci. We adopted previously established approaches55–57 and integrated enrichment results across libraries from multiple sources to establish a GO and pathway co-association network (Fig. 4c). To evaluate the correlation strength between any two genes in the network, we calculated a co-association score for every possible gene-gene combination (n = 72,771) from both genes prioritized and deprioritized by lo-siRF. Lo-siRF-prioritized genes are the 21 genes functionally mapped from the 283 lo-siRF-prioritized SNVs and their LD-linked SNVs (Fig. 4a). Lo-siRF-deprioritized genes are those functionally mapped from the SNVs that failed to pass the lo-siRF prioritization threshold. Compared to random gene pairs in the network, 10 genes that were functionally mapped from the lo-siRF-prioritized epistatic and hypostatic loci showed significant co-associations with multiple GO/pathways (Fig. 4c, Extended Data 6). Consistent with our hypothesized epistasis (Fig. 2f), gene CCDC141 showed a significant co-association to SYNM (functionally linked to the IGF1R locus) and PDE11A and PLEKHA3 (both functionally linked to the TTN locus) through the GO term of hyperactivity (excessive movement), which has been linked to increased risk of cardiac disease58. Beyond that, TTN, IGF1R, and SYNM are co-associated with kinase activity and cardiac structure related GO terms, indicating that these genes may jointly affect cardiac structure by regulating the process of kinase activity.
Genes mapped from epistatic loci are co-associated with myogenic regulatory factors
We next performed an integrative enrichment analysis to assess transcriptional regulation of genes prioritized and deprioritized by lo-siRF. Due to assay-specific limitations and biases, we integrated the enrichment results across nine distinct gene set libraries55,56 (Fig. 4d, Extended Data 7). We found that the lo-siRF-prioritized epistatic genes shared important myogenic regulatory factors, such as MYOD1, MYF6, and MYOG (Fig. 4d, top). These myogenic regulatory factors coordinate to regulate muscle development and differentiation. In contrast, Transcription factors enriched from lo-siRF-deprioritized genes display a less coordinated regulatory pattern (Fig. 4d, bottom). These analyses enriched transcription factors based on their associations to given sets of individual genes rather than co-association to gene pairs55,56. To further evaluate the correlation strength between any two genes that share transcription factors, we calculated a transcription factor co-association score for all the 72,771 possible gene-gene combinations (see Methods). Compared with random gene pairs,16 gene-gene combinations from the lo-siRF-prioritized genes displayed a significant co-association (empirical p < 0.05, Fig. 4e). These co-associations were found in gene-gene combinations from both intra- and inter-lo-siRF-prioritized loci (Fig. 4e). In particular, pairwise combinations among TTN, TNNT3, CCDC141, and SYNM share a common splicing regulator, RBM20 (Extended Data Fig. 5). RBM20 has been reported to regulate the alternative splicing of genes important for cardiac sarcomere organization59. This suggests that the splicing patterns of these four genes are likely to be co-regulated by RBM20, which is consistent with the exhibited enrichment of sQTLs by the CCDC141, TTN and LSP1 lo-siRF loci (Extended Data 4).
Genes mapped from epistatic loci exhibit strong co-expression and connectivity change in human heart failure transcriptomics
We proceeded to the fourth stage for experimental confirmation (Fig. 1, blue) and evaluated how the identified epistases contribute to the progression of heart failure (Fig. 5). We employed a series of weighted gene co-expression networks derived from human cardiac transcriptomic data from 177 failing hearts isolated at the time of heart transplant and 136 non-failing hearts harvested from cardiac transplant donors whose organs were not able to be placed60 (Fig. 5a). We compared the molecular connectivity of genes identified as statistical epistatic interactors. We defined connectivity as the edge weights between two genes normalized to the distribution of all network edge weights, and compared this to the connectivity of all other available gene-gene combinations in the network. This revealed strong co-expression correlations between CCDC141 and genes functionally linked to the IGF1R locus (SYNM and LYSMD4) and TTN locus (TTN and FKBP7) in the healthy control network (Fig. 5b). In contrast, most of these gene pairs (except for CCDC141-TTN) no longer exhibit a strong connectivity in the heart failure network (Fig. 5c). All of these connectivities showed a significant decrease (indicated by the negative connectivity difference score and p < 0.05 in Fig. 5d) in the differential network, suggesting a declined co-expression correlation between these gene pairs relative to random gene pairs during the progression of failing hearts. This difference is potentially related to the rewired gene modular assignments between the control and heart failure networks60 (Fig. 5e and Extended Data Fig. 6). For instance, CCDC141, SYNM, TTN, and TNNT3 are co-associated with the electron transport chain/metabolism module in the control network. In the failing hearts, SYNM and TTN rewire to the muscle contraction/cardiac remodeling module, whereas CCDC141 and TNNT3 remain associated with the metabolism module (Fig. 5e). In addition, other genes functionally linked to IGF1R and TTN lo-siRF loci are co-associated with the membrane transport or unfolded protein response module in healthy hearts and rewire to the muscle contraction/cardiac remodeling or cell surface/immune/metabolism module in failing hearts.
Perturbation confirms epistatic relationships in cardiomyocyte hypertrophy
We interrogated epistatic associations in a genetic model of cardiac hypertrophy (Fig. 1, blue): induced pluripotent stem cell cardiomyocytes derived from patients with and without hypertrophic cardiomyopathy caused by the cardiac myosin heavy chain (MYH7) p.R403Q variant33 (Fig. 6a). Cardiac myosin heavy chain 7 is a key component of the cardiac sarcomere, and the most common cause of hypertrophic cardiomyopathy33. The patient presented with typical symptoms, and echocardiography revealed severe LV hypertrophy and a small LV cavity33. At the cellular level, cardiomyocytes exhibit an elevated mean cell size and non-Gaussian size distribution with a long tail relative to the unaffected control (Fig. 6d).
To determine if CCDC141 can act both independently and in epistatic interactions with other genes to attenuate the pathologic cellular hypertrophy caused by MYH7-R403Q, we silenced genes CCDC141, IGF1R, TTN, and gene pairs CCDC141-IGF1R and CCDC141-TTN using siRNAs in both diseased and healthy cardiomyocytes and compared them with cells transfected with scramble siRNAs (control) (Fig. 6a and 6e). Phenotypic consequences of these perturbations on cellular morphology were then evaluated in high-throughput using a spiral inertial microfluidic device (Fig. 6b) in combination with automated single-cell image analysis (Fig. 6c). The microfluidic device adopted the Dean flow focusing principle31 (details in Extended Data Fig. 7 and Methods) to mitigate the non-uniform cell focusing61, thereby enhancing the imaging resolution62 affected by the large variations in cardiomyocyte diameter (Fig. 6d).
We first assessed the knockdown effects of the CCDC141-IGF1R interaction on cardiomyocyte size (Fig. 6f). Bootstrapped hypothesis tests were performed, for which the p-values are capped below by p < 1E-4 (Extended Data 8). Silencing IGF1R alone reduces the median cell size by 5.3% ± 0.4% (p < 1E-4) in diseased cells compared to scrambled control and 6.6% ± 0.5% (p < 1E-4) in healthy cells. Silencing CCDC141 alone also decreases median cell size by 3.2% ± 0.5% (p < 1E-4) in diseased cells, but had no impact on healthy cells. Digenic silencing of CCDC141 and IGF1R reveals a synergistic effect on attenuating pathologic cell hypertrophy in diseased cells, resulting in an 8.5% ± 0.3% (p < 1E-4) decrease in the median cell size. This is consistent in healthy cells, where silencing CCDC141 alone fails to affect cell size, but digenic silencing of CCDC141 and IGF1R decreases the median cell size by 9.3% ± 0.5% (p < 1E-4). Moreover, according to our estimated quantile regression analysis (details in Methods), this interaction effect appears to be non-additive for both healthy and diseased cells ( < 0, Fig. 6g; p < 1E-4 for non-additivity, Extended Data 8), consistent with an epistatic mechanism. These findings serve to confirm the strongest epistatic association identified by lo-siRF (Fig. 2f).
We found a comparable non-additive effect for the CCDC141-TTN interaction. Digenic silencing of CCDC141-TTN leads to a pronounced reduction in median cell size (by 5.8% ± 0.6% for healthy cells and 3.3% ± 0.4% for diseased cells, p < 1E-4) relative to monogenic silencing (Fig. 6f). This interaction appears to be non-additive for both healthy and diseased cells (p values in Extended Data 8) yet demonstrating opposite epistatic directions in these two cell states (Fig. 6g). Additionally, CCDC141 and TTN show distinctive independent roles in repressing cardiomyocyte hypertrophy. In healthy cells, monogenic silencing of TTN leads to a larger cell size reduction compared to the case of silencing CCDC141. In contrast, diseased cells display a larger size reduction in response to monogenic silencing of CCDC141.
Furthermore, both CCDC141-IGF1R and CCDC141-TTN interactions show a stronger effect on rescuing larger cardiomyocytes over smaller ones in both cell lines (Extended Data Fig. 8 and 9). In contrast, monogenic silencing does not exhibit such a non-uniform effect on reshaping the cell size distribution, which reinforces the hypothesized non-additivity of these two epistatic interactions (details in Extended Data 8 and Supplementary Note 2).
Recent studies have shown that cellular morphological features, such as cell boundary and textural irregularities, are informative readouts of cytoskeletal structure, which is highly associated with disease state in hypertrophic cardiomyopathy32,63. We analyzed relative changes in cell shape and texture (Fig. 6h) by measuring the counts of peak intensities normalized to the total number of pixels enclosed by the cell boundary (Fig. 6i). Cells with a high normalized peak number display a ruffled texture, which manifests in unevenly distributed 2D intensities (Fig. 6k). Our analysis shows that silencing both CCDC141 and IGF1R (circles in Fig. 6h, left) yields a larger increase in intensity peak number than silencing IGF1R alone (triangles in Fig. 6h, left) for both cell lines, exhibiting a synergistic epistasis between CCDC141 and IGF1R (p < 1E-4 for non-additivity). We also analyzed cell roundness error, a measure of how far radii measured on the cell outline deviate from a perfect circle (Fig. 6i). This parameter increases with an increasing cell boundary waviness or elongation (Fig. 6j). We show that the silencing of CCDC141 and IGF1R synergistically interact to increase roundness error of diseased cardiomyocytes (p < 1E-4 for non-additivity, Fig. 6h, left). In addition, CCDC141 and TTN display antagonistic epistasis and synergistic epistasis in their impact on roundness error for healthy and diseased cells (p < 1E-4 for non-additivity, Fig. 6h, right), respectively.
Discussion
While computational models18,19 have supported epistatic contributions to human complex traits and disease risk, examples in the literature are rare, with even fewer experimentally confirmed. Here, we developed a veridical machine learning26 approach to identify epistatic associations with cardiac hypertrophy derived from a deep learning model that estimates LV mass from cardiac imaging of almost thirty thousand individuals in the UK Biobank. We report novel epistatic effects on LV mass of common genetic variants associated with CCDC141, TTN, and IGF1R. We used established tools to functionally link risk loci to genes, and then confirmed gene level co-associations through network analyses, including via shared transcription factors and pathways enriched against multiple annotated gene set libraries and co-expression networks we built using transcriptomic data from over three hundred healthy and diseased human hearts. Finally, using a cellular disease model incorporating monogenic and digenic silencing of individual genes, we assessed phenotypic changes in cardiomyocyte size and morphology using a novel microfluidic system, confirming the non-additive nature of the interactions.
Our approach advances epistasis discovery in several key ways. First, unlike studies relying on linear-based models64–67, we leverage a more realistic, nonlinear tree-based model that mirrors the thresholding (or switch-like) behavior commonly observed in biomolecular interactions68. Second, in contrast to other tree-based approaches that evaluate interactions on a variant-by-variant basis69–73, our novel stability-driven importance score consolidates individual variants into loci for the assessment of feature importance, allowing for more reliable extraction of epistatic interactions from weak association signals. This is particularly valuable for evaluating non-coding variants and resembles ideas from marginal association mapping with sets of SNVs53,54,74. Moreover, instead of exhaustively searching all possible interactions, signed iterative random forests internally employ a computationally-efficient algorithm, which automatically narrows the search space of interactions to only those that stably appear in the forest and thus achieves a scalability much higher than existing tree-based approaches71,75. This allows lo-siRF to handle larger datasets without the need for LD pruning before the interaction search, which may inadvertently eliminate important epistatic variants, given that epistasis between loci in strong LD has been evidenced by a recent study76. Furthermore, our computational prioritization is rigorously validated through multiple functional network analyses and robust experimental confirmation.
Our results add to a small literature on epistasis in cardiovascular disease. Two recent studies have found epistasis influencing the risk of coronary artery disease18,19. Li et al.19 identified epistasis between ANRIL and TMEM106B in coronary artery tissues. Although their method predicted functionally interpretable interactions between risk loci of interest, they relied heavily on prior knowledge and careful selection of the causal gene pairs,19 making the approach challenging to scale. Zeng et al.18 used population-scale data and performed epistasis scans from regions around 56 known risk loci. This study identified epistasis between variants in cis at the LPA locus without experimental confirmation. In contrast, our approach allows discovery of not only cis-epistasis, but also long-range interactions between interchromosomal loci (e.g., CCDC141 and IGF1R) and is supported by gene perturbation experiments. More importantly, both studies searched for interactions around known risk loci identified by genome-wide association, which can be far away from the possible epistatic or hypostatic loci that are statistically insignificant in linear univariate association studies. In addition, both studies relied on a logistic regression model, which imposes restrictive assumptions that can be avoided using a nonlinear machine learning approach as in lo-siRF.
Our study has limitations. Given our primary interest in biological epistasis rather than statistical epistasis21, we tailored lo-siRF to conservatively prioritize reliable targets for experimental validation as opposed to finding all possible epistatic drivers. Lo-siRF should ideally be used as a first-stage hypothesis generation tool within a broader scientific discovery pipeline. To assess significance of the lo-siRF-prioritized targets, we rely on and encourage follow-up investigations such as the high-throughput gene-silencing experiments conducted here. We focused this analysis on a single ancestry in order to enhance the likelihood of finding reliable interactions from weak association signals. These findings cannot be automatically applied to others. It was not feasible to conduct a formal genetic replication study because the UK Biobank is the only large-scale population cohort with integrated cardiac magnetic resonance images and genetic data. However, to help reduce the possibility of overfitting and increase generalizability, lo-siRF employed numerous stability analyses (see Supplementary Note 1) in addition to a proper training-validation-test data split. Beyond these computational checks, we also present functional supporting evidence and experimental validation. Our computational prioritization via lo-siRF currently groups SNVs based on genomic proximity, without accounting for their functional interdependencies, but this could be addressed by integrating functional annotation into the lo-siRF pipeline. Lo-siRF also relies on a GWAS to reduce the number of SNVs to a computationally manageable size, but this could be improved with more sophisticated epistasis detection algorithms such as MAPIT. Lastly, lo-siRF is not as scalable as linear-based methods, though it is more scalable than alternative tree-based methods for epistasis detection. It also should be noted that although this study did not identify stable higher-order (> order-2) interactions due to the weak association signal between SNVs and LV mass, the method exhibits the capability to detect such interactions for broader phenotypes and complex traits without incurring additional computational cost.
In summary, our work adds to the discovery toolkit for the genomic architecture of complex traits and expands the scope of genetic regulation of cardiac structure to epistasis.
Online Methods
Study participants
The use of human subjects (IRB −4237) and human-derived induced pluripotent stem cells (SCRO −568) in this study has been approved by the Stanford Research Compliance Office. The UK Biobank received ethical approval from the North West -Haydock Research Ethics Committee (21/NW/0157).
The UK Biobank (UKBB) is a biomedical database with detailed phenotypic and genetic data from over half a million UK individuals between ages 40 and 69 years at recruitment77. In this study, we restricted our analysis to the largest ancestry subset (i.e., the White British population) of 29,661 unrelated individuals who have both genetic and cardiac magnetic resonance imaging (MRI) data from the UKBB (Supplementary Table 1). More specifically, we considered only those individuals from the UKBB cohort who self-reported as White British and have similar genotypic backgrounds based on principal components analysis as described in prior work77. We also identified related individuals (i.e., third-degree relatives or closer) via genotyping and omitted all but one individual from each related group in the analysis. Details regarding this cohort refinement have been described and implemented previously77,78. This refinement resulted in a cohort of 337,535 unrelated White British individuals from the UKBB, of which 29,661 have both genetic and cardiac MRI data. We randomly split this data into training, validation, and test sets of size 15,000, 5,000, and 9,661 individuals, respectively.
Genotyping and quality control
For the study cohort of 29,661 individuals described above, we leveraged genotype data from approximately 15 million imputed autosomal SNVs. These have been imputed from 805,426 directly assayed SNVs (obtained by the UKBB from one of two similar Affymetrix arrays) using the Haplotype Reference Consortium and UK10K reference panels77. Imputed variants were subject to several quality-control filters, including outlier-based filtration on effects due to batch, plate, sex, array, and discordance across control replicates. Further, we excluded variants due to extreme heterozygosity, missingness, minor allele frequency (< 10-4), Hardy-Weinburg equilibrium (< 10-10), and poor imputation quality (< 0.9). Further details can be found in previous studies77,78.
Quantification of left ventricular hypertrophy
We retrieved cardiac MRI images from 44,503 UKBB participants, taken during their most recent imaging visit, and closely followed the method previously described by Bai et al.28. A fully convolutional network28 was previously trained using a dataset of 4,875 subjects with 93,500 pixelwise segmentations of UKBB short-axis cardiac MRI multi-slice images generated manually with quality control checks for inter-operator consistency79. The cardiac MRI image resolution was 1.8 x 1.8 mm2, with a slice thickness of 8.0 mm and gap of 2.0 mm, typically consisting of 10 slices. Each slice was converted to an image and cropped to a 192 x 192 square, and measurements were 0-1 normalized. The network architecture employed multiple convolutional layers to learn image features across five resolution scales. Each scale involved two or three convolutions with kernel size 3 x 3 and stride 1 or 2 (2 appearing every 2 or 3 layers), followed by batch normalization and ReLU transformation. Feature maps from the five scales were upsampled back to the original resolution, combined into a multi-scale feature map, and processed through three additional convolutional layers with kernel size 1 x 1, followed by a softmax function to predict the segmentation label for each pixel. For an exact description of the model architecture, we refer to the original publication on the model28. Notably, each of the pixelwise annotations used for training and evaluation was hand-segmented by a human expert and validated for quality. Furthermore, the model was validated in the UKBB and demonstrated strong concordance with the human-generated gold standard28, ensuring that model predictions in the same dataset are of high quality. To our knowledge, this is the only published model trained in the UKBB on gold standard labels. We thus applied this trained deep learning model to our entire dataset of 44,503 cardiac MRIs. This resulted in segmentations of the LV cavity and myocardium from each short axis frame, which allowed for both an area calculation of each segment as well as the application of quality control checks28 based on consistency within and between slices and time steps. There were 44,219 segmentations that passed the quality control. Using the calculated areas, we computed the volume of the LV myocardium through simple integration over slices. This volume was then converted to a left ventricular mass (LVM) using a standard density estimate of 1.05 g/mL35. LVMi was computed by dividing LVM by an estimate of body surface area based on height and body weight calculated using the Du Bois formula36. From the 44,219 segmentations, we restricted the analysis to LVMi measurements for 29,661 unrelated White British individuals using the measurements from their most recent imaging visit if multiple imaging visits were recorded.
Lo-siRF step 1: Dimension reduction of variants via genome-wide association studies
As the first step in the lo-siRF pipeline, we performed a genome-wide association study (GWAS) on the training data for the rank-based inverse normal-transformed LVMi using two algorithms, PLINK37 and BOLT-LMM38, in order to filter the number of features from over 15 million SNVs to a more computationally-feasible size (Fig. 2b). This step is akin to typical screening phases in fine-mapping80 and other tree-based epistasis detection methods72,81. Since BOLT-LMM and PLINK rely on different statistical models, we chose to employ both implementations to mitigate the dependence of downstream conclusions on this arbitrary choice. Specifically, for the first GWAS run, we fitted a linear regression model, implemented via ‘glm’ in PLINK82. For the second GWAS run, we used BOLT-LMM38, a fast Bayesian-based linear mixed model method. Each GWAS was adjusted for the first five principal components of ancestry, sex, age, height, and body weight. We then ranked the SNVs by significance (i.e., the GWAS p-value) for each GWAS run separately and took the union of the top 1000 SNVs (without clumping) from each of the two GWAS runs. This resulted in a set of 1405 GWAS-filtered SNVs that were used in the remainder of the lo-siRF pipeline. Here, we chose to use the top 1000 SNVs per GWAS method (without clumping) as it yielded the highest validation prediction accuracy compared to choosing other possible thresholds (500 and 2000 SNVs per GWAS with and without clumping). Though the GWAS is not the focus of this work, we provide a summary of the PLINK and BOLT-LMM GWAS results for completeness and for comparison in Extended Data 1. We also provide a list of the 1405 GWAS-filtered SNVs in Extended Data 2. We note that these 1405 GWAS-filtered SNVs strictly contain the SNVs that passed the genome-wide significance threshold (p = 5E-8).
Lo-siRF step 2: Binarization of the left ventricular mass phenotype
Next, we partitioned the raw (continuous) LVMi phenotype into a low, middle, and high LVMi group before fitting signed iterative random forest to classify individuals with low versus high LVMi (Fig. 2c). That is, for a given threshold x, we binned individuals within the top and bottom x% of LVMi values into two classes with the high and low LVMi values, respectively, while omitting the individuals in the middle quantile range. Due to the sex-specific biological variation of LVMi (Supplementary Note 1), we performed this partitioning for males and females separately. For males, low and high LVMi was considered under 43.8-46.0 g/m2 and above 55.4-58.5 g/m2, respectively, depending on our choice of binarization threshold (Supplementary Table 2). For females, low and high LVMi was defined as under 35.1-36.8 g/m2 and above 43.8-46.1 g/m2, respectively, depending on our choice of binarization threshold. We performed this binarization step in order to simplify the original low-signal regression problem into a relatively easier binary classification task: to distinguish between individuals with very high LVMi from those with very low LVMi. This binarization approach was motivated by the observation that the validation R2 values from the original regression problem of predicting each individual’s raw (continuous) LVMi were smaller than 0 (Supplementary Table 3 and Supplementary Note 1), raising the question of whether the regression models were capturing anything relevant to reality. At a minimum however, the PCS framework for veridical data science26 advocates the importance of ensuring that the model fits the data well, as measured by prediction accuracy, before trusting any extracted interpretations from that model. We will see in the next section that the binarization procedure not only strengthened the prediction signal but also helped us more readily interpret and assess the performance of prediction methods with respect to the prediction screening step of the PCS framework26. We importantly note that the practical use of this approach depends heavily on whether understanding the differences between the high and low categories is relevant to the scientific goals. Here, we believe that the connection between cardiac hypertrophy and those with high LVMi helps to justify the binarization approach and that studying how individuals with high LVMi differ from those on the other end of the spectrum may yield relevant scientific insights. Since the specific threshold choice is arbitrary, we ran the remainder of the lo-siRF pipeline using three different binarization thresholds (15%, 20%, 25%) to balance the improvement in prediction signal and amount of data lost. In the end, we aggregated the results that were stable across all three binarization thresholds, described in the Method section Lo-siRF step 4.4: Ranking genetic loci and interactions between loci.
Lo-siRF step 3: Prediction
Lo-siRF step 3.1: Fitting signed iterative random forest on the binarized LV mass index phenotype
For each binarization threshold, we trained a signed iterative random forest (siRF) model23 using the 1405 GWAS-filtered SNVs to predict the binarized LVMi phenotype and generate candidate interactions for further investigation (Fig. 2d). siRF first iteratively grows a sequence of feature-weighted random forests, re-weighting features in each iteration proportional to their feature importance from the previous iteration in order to stabilize the decision paths. Then, provided that the resulting stabilized forest provides reasonable prediction performance (see the Methods section Lo-siRF step 3.2: Prediction check), siRF leverages a computationally-efficient algorithm, random intersection trees83, to identify nonlinear higher-order interaction candidates based on frequently co-occurring features on a decision path. Intuitively, sets of features that frequently co-occur along a decision path together are more likely to interact and are identified by siRF. siRF is particularly attractive for prioritizing epistatic interactions as (1) it offers an interaction search engine that can automatically search for higher-order interactions with the same order of computational cost as a traditional random forest, and (2) the thresholding behavior of its decision trees resembles the thresholding (or switch-like) behavior commonly observed in biomolecular interactions68. Further, siRF improves upon its predecessor, iterative random forests22, by not only tracking which sets of features commonly co-occur on decision paths, but also the sign of the features, i.e., whether low values (denoted X−) or high values (denoted X+) of feature X, appear on the decision path. We refer to Kumbier et al.23 for details, but in brief, the signed feature X− (or respectively, X+) signifies that a decision rule of the form X < t (or respectively, X > t) for some threshold t appeared on the decision path. siRF hence outputs a list of candidate signed interactions, where each signed interaction consists of two or more signed features that frequently co-occur on the same decision path. Note when applying siRF to SNV data in practice, the signed feature SNV+ typically represents a heterozygous or homozygous mutation while the signed feature SNV− typically represents no mutation at the locus. The following hyperparameters were used to train siRF using the iRF2.0 R package: number of iterations = 3, number of trees = 500, number of bootstrap replicates = 50, depth of random intersection tree (RIT) = 3, number of RIT = 500, number of children in RIT = 5, and minimum node size in RIT = 1. We did not perform hyperparameter tuning since siRF has been previously shown to be robust to different choices of hyperparameters23. We fit siRF using 10,000 training samples (randomly sampled out of the 15,000 total training samples) and reserved the remaining 5,000 training samples for selecting genetic loci for the permutation test (see the Method section Lo-siRF step 4.3: Permutation test for difference in local stability importance scores).
Lo-siRF step 3.2: Prediction check
Per the PCS framework for veridical data science26, we next assessed the validation prediction accuracy of siRF (Fig. 2d) to evaluate whether the learnt model is capturing some biologically-relevant phenotypic signal, rather than simply noise, before proceeding to interpret this model in step 4 of lo-siRF. To serve as baseline comparisons, we fit other popular machine learning prediction methods, namely, L1-regularized (LASSO) logistic regression84, L2-regularized (ridge) logistic regression85 using glmnet in R, random forests86 using ranger in R, support vector machines87 with the radial basis kernel using sklearn’s SVC in Python, a multilayer perceptron88 (fully-connected feedforward neural network with one hidden layer and ReLU activations) using sklearn’s MLPClassifier in Python, and AutoGluon TabularPredictor89 (an auto machine learning framework which ensembles multiple models, including neural networks, LightGBM, boosted trees, random forests, and k nearest neighbors, by stacking them in multiple layers) in Python. We used the following hyperparameters and tuned using 5-fold cross-validation where applicable:
L1- and L2-regularized logistic regression: default λ grid from glmnet::cv.glmnet in R;
Random forests: default parameters from ranger::ranger in R;
Support vector machine with radial basis kernel: regularization parameter C = 1E-4, 1E-3, …, 1E3, 1E4;
Multilayer perceptron: number of neurons in the hidden layer = 8, 16, 32, 64, 128, 256; L2-regularization parameter α = 1E-4, 1E-3, 1E-2;
AutoGluon TabularPredictor: trained with the “medium quality” and “good quality” presets.
We also compared siRF to a basic polygenic risk score. Specifically, we used PLINK to construct a polygenic risk score using the lead SNVs from FUMA for the LVMi PLINK GWAS that passed the suggestive significance threshold of 1E-5 (Extended Data 1), and we fit a logistic regression using this polygenic risk score as a predictor of the binarized LVMi. We evaluated prediction performance for each of these methods according to multiple metrics: classification accuracy, area under the receiver operator curve (AUROC), and area under the precision-recall curve (AUPRC). We observed that the prediction power of siRF, though weak (∼55% balanced classification accuracy, ∼0.58 AUROC, and ∼0.57 AUPRC), was greater than these other commonly used prediction methods across all binarization thresholds and evaluation metrics, except for the 15% binarization threshold where siRF performed second-best with respect to classification accuracy (Supplementary Table 4). Since siRF performed better than random guessing (i.e., >50% balanced classification accuracy and >0.5 AUROC/AUPRC, which is not guaranteed given the high phenotypic diversity of the LVMi trait) and demonstrated higher prediction power than alternative popular prediction methods, we deemed that the siRF fit for LVMi passed the prediction screening step of the PCS framework. Hence, we proceeded to interpret this siRF model and prioritize candidate interactions in step 4 of lo-siRF. We note also that this prediction check played a key role in our choice of phenotypic data. Prior to studying LVMi, we attempted to run a similar analysis to predict hypertrophic cardiomyopathy (HCM) diagnosis, defined as any ICD10 billing code diagnosis of I42.1 or I42.2 in the UKBB data. However, neither siRF nor the other aforementioned prediction methods passed the 50% balanced classification accuracy requirement for predicting HCM diagnosis. We thus chose not to proceed with the HCM analysis given the poor prediction accuracy and uncertain relevance between the prediction models and the underlying biological processes. This failed prediction check motivated the need for a more refined phenotypic measure of cardiac hypertrophy, which ultimately led to the deep learning extraction of cardiac MRI-derived LVMi. Further discussion of the HCM analysis can be found in Supplementary Note 1.
Lo-siRF step 4: Prioritization
To lastly interpret the siRF fit for LVMi, we developed a novel stability-driven importance score to prioritize genetic loci and more interestingly, interactions between loci for follow-up experimental validation (Fig. 2e). The assessment of importance at the level of genetic loci, instead of individual variants, is necessary since variant-level importances here are incredibly unstable (detailed in Supplementary Note 1). This is due to the high correlation between SNVs in LD and the weak phenotypic signal. Consequently, our new importance score aims to aggregate weak, unstable variant-level importances into stronger, more stable locus-level importances via three steps: (1) assigning each variant to a genetic locus, (2) evaluating the local (or per-individual) importance of each genetic loci or interaction between loci in the siRF fit via a stability-driven measure, and (3) conducting a permutation test to summarize the importance of the genetic locus or interaction between loci across all individuals. We provide details for each step next.
Lo-siRF step 4.1: Aggregation of SNVs into loci
We aggregated SNVs into a genetic locus based on genomic proximity. Specifically, we used ANNOVAR90 to assign each SNV that appears in the siRF fit to a genetic locus according to the hg19 refSeq Gene annotations (i.e., given by the ‘Gene.refGene’ column in the ANNOVAR output).
ANNOVAR uses a default of 1 kb as the maximum distance between SNVs and gene boundaries. Note that from these annotations, each SNV is assigned to exactly one genetic locus. Thus, herein in the context of lo-siRF, a genetic locus is a (non-overlapping) group of SNVs, and a signed genetic locus is a (non-overlapping) group of signed SNVs with the specified sign (i.e., Locus+ consists of SNV1+, …, SNVp+ while Locus-consists of SNV1-, …, SNVp-).
Lo-siRF step 4.2: Local stability importance score
We next measured the importance of a genetic locus or interaction between loci based on their stability, or frequency of occurrence, within the siRF fit (i.e., the total number of times that SNVs from a particular locus or interaction were split upon in the fitted forest). However, because the number of variants assigned to each genetic locus can vary, the raw frequency of occurrence will be biased towards larger loci (i.e., those with more variants). A more detailed discussion is provided in Supplementary Note 1. To address this issue, we developed a local (or per-individual) stability importance score, which quantifies the importance of a signed locus or interaction between loci for making the prediction for each individual. Let G = {g1,…, gK} denote a signed order-K interaction involving the signed genetic loci g1,…, gK, and let , . . . ) denote the signed SNVs belonging to the signed genetic locus gj. Then given a forest T, a signed interaction between loci G, and individual i, the local stability importance score, LSIT(G, i), is defined as DT(G, i) / |T|, where |T| is the number of trees in the forest T, and DT(G, i) is the number of decision paths in the forest T for which two criteria are satisfied: (1) individual i appears in its terminal node and (2) for each j = 1,…, K, there exists an l ∈ {1,.. ., pj} such that was used in a decision split along the path (Extended Data Fig. 3a). In other words, LSIT(G, i) is the proportion of trees in the forest T for which at least one signed variant from each signed locus in the signed interaction G was used in making the prediction for individual i. A high score indicates that the signed interaction G was frequently used to predict individual i’s response and is an important interaction for individual i. Note that a genetic locus can be viewed as an order-1 interaction, and thus, this local stability importance score can also be applied to assess the (marginal) importance of a single genetic locus.
Lo-siRF step 4.3: Permutation test for difference in local stability importance scores
Once we obtained these local stability importance scores for each individual, we performed a two-sample permutation test (Extended Data Fig. 3a) to assess whether the local stability importance scores for a given signed locus or interaction between loci, G, are different between individuals with high and low LVMi (conditional on the rest of the fitted forest). More formally, the proposed permutation test tests the null hypothesis L = H versus the alternative hypothesis L ≠ H, where L and H are the distributions of local stability importance scores for individuals with low and high LVMi, respectively. If the local stability importance scores are indeed different between high and low LVMi individuals (thus giving a small permutation p-value), this indicates that G can differentiate between individuals with high versus low LVMi given the fitted siRF and hence is an important locus or interaction between loci for LVMi. We performed this permutation test using 10,000 permutations, the difference in means as the test statistic, and the 5,000 validation samples. To bolster the reliability of our findings, we only tested a conservative subset of genetic loci and interactions between loci that passed predictive and stability checks in accordance with the PCS framework. Namely, we tested:
The top 25 genetic loci, ranked by their average local stability importance scores across 5,000 samples. These 5,000 samples were previously set-aside from within the 15,000 training samples and were not used in fitting the siRF (see the Methods section Lo-siRF step 3.1: Fitting signed iterative random forest on the binarized LV mass index phenotype);
The signed interactions between loci that were stably identified by siRF across 50 bootstrap replicates. Here, we performed the random intersection trees search within siRF at the locus-level (i.e., using the variant-to-locus assignment as the hyper-features or ‘varnames.grp’ argument when running siRF in R), and we defined a “stable” interaction as one that passed the following siRF stability metric thresholds: stability score > 0.5, stability score for mean increase in precision > 0, and stability score for independence of feature selection > 022,23 (Supplementary Table 6). Briefly, the stability score measures how frequently the interaction appears in siRF. The stability score for mean increase in precision threshold requires that the interaction is predictive of the response. The stability score for feature selection dependence threshold helps to filter out additive interactions (as opposed to the desired non-additive interactions). Details on the siRF interaction and stability metrics can be found in previous work23.
We reiterate that given the complexities and challenges associated with the low-signal data under study, we utilize these permutation p-values primarily as a summary statistic to rank candidate loci and interactions (detailed next), rather than as an assessment of statistical significance, which relies heavily on untestable model assumptions that often do not hold in practice.
Lo-siRF step 4.4: Ranking genetic loci and interactions between loci
Before ranking the top lo-siRF recommendations for follow-up experimental validation, we incorporated one final stability check, recommending only those signed loci and interactions between loci that underwent the permutation test and yielded a p-value < 0.1 in all three binarization runs. For these signed loci and interactions between loci that were stably important in all three binarization runs, we ranked them by the mean permutation p-value, averaged across the three binarization thresholds (Supplementary Table 5). Because of our emphasis on prioritizing candidates for experimental validation, if both the + and -version of the signed locus (or interaction) appear, the final prioritized loci (or interaction) are ranked according to the smaller one of the two p-values (Fig. 2f). We note that though the signed information is not pertinent to our goal of recommending candidates for experiments, the signed information from siRF provides more granular information that can improve our interpretation of the fit, and we discuss this further in Supplementary Note 1. We also provide the permutation p-values for all conducted permutation tests (including the loci and interactions between loci that were unstable across binarization thresholds) in Supplementary Note 1.
Lo-siRF: PCS documentation and additional stability analyses
We acknowledge that many human judgment calls were inevitably made throughout our veridical machine learning pipeline and that alternative choices could have been made (e.g., different dimension reduction techniques, binarization procedures, and prediction models). In an effort to facilitate transparency of these human judgment calls, we provide extensive documentation, discussion, and justification in Supplementary Note 1. In particular, Supplementary Note 1 includes a discussion of our reasoning and motivation behind the choice of phenotypic data as well as choices in the dimension reduction, binarization, prediction, and prioritization steps. We also performed additional stability analyses in accordance with the PCS framework26, to ensure that our findings are stable and robust to these human judgment calls (e.g., the choice of GWAS method and binarization threshold) and to bolster the reproducibility of our findings. Supplementary Note 1 is an HTML document, which can be downloaded and displayed in a browser or found at https://yu-group.github.io/epistasis-cardiac-hypertrophy/.
Non-hypertensive cohort analysis
We defined hypertensive individuals as anyone with self-reported hypertension, high blood pressure as diagnosed by a doctor, or any ICD10 billing code diagnosis in I10-I16. Out of the 29,661 UKBB participants in the original lo-siRF analysis, 7,371 individuals had hypertension, leaving 22,290 individuals for the non-hypertensive analysis. Specifically, using the same set of 1405 GWAS-filtered SNVs as in the original lo-siRF analysis, we performed steps 2-4 of the lo-siRF analysis using only the non-hypertension cohort. We also assessed the marginal effect of each of the 1405 GWAS-filtered SNV on hypertension. Here, we fit logistic regression models, regressing hypertension (i.e., a binary indicator of whether or not one has hypertension) on each SNV marginally, while adjusting for the first five principal components of ancestry, sex, age, height, and body weight. A more detailed discussion of the non-hypertension analysis results can be found in Supplementary Note 1.
Implementation of existing epistasis detection methods
Exhaustive regression-based pairwise interaction scan39,40
For each pair of SNVs that passed the GWAS filter in lo-siRF step 1, we fit the follow regression:
where yi is the rank-based inverse normal-transformed LVMi for individual i, gij is the genotype of SNV j for individual i, zi is a vector of covariates for individual i (i.e., sex, age, height, body weight, and the first five principal components of ancestry), and εi is the random error or noise term for individual i. Under this regression model, we tested the null hypothesis of β12 = 0 versus the alternative hypothesis of β12 ≠ 0 via the traditional t-test. We also repeated this exhaustive interaction search using the binarized LVMi response for each of the three different binarization thresholds (15%, 20%, and 25%). For the binarized LVMi, we used a logistic regression in lieu of the linear regression and tested for a non-zero β12 coefficient via the traditional Wald z-test. For brevity, we defer results to Supplementary Note 1.
MAPIT41
MAPIT leverages a variance component model to first identify candidate variants with non-zero marginal epistatic effects, defined as the total pairwise interaction effect between the variant and all other variants41. By focusing on these marginal epistatic effects, MAPIT can advantageously search for epistatic variants without enduring the computational and statistical burdens associated with pinpointing their epistatic partners. We performed MAPIT using the mvMAPIT (v2.0.3) R package. For inputs, we used the 1405 GWAS-filtered SNVs with minor allele frequency > 0.05, adjusted for sex, age, height, body weight, and the first five principal components of ancestry, and used the rank-based inverse normal-transformed LVMi as the response. We used the default settings in the mvmapit function and chose the “normal” test to minimize the computational burden. We provide the results in Supplementary Note 1.
Implementation of existing set-based genome-wide association tests
To investigate the importance of the IGF1R locus using existing set-based association methods, we performed SKAT-O53 using the subset of 1405 GWAS-filtered SNVs with minor allele frequency > 0.05 as input and the rank-based inverse normal-transformed LVMi as the response. We also adjusted for sex, age, height, body weight, and the first five principal components of ancestry in the SKAT-O null model. This analysis was carried out using the SKAT (v2.2.5) R package. In addition to SKAT-O, we also ran the gene-based test as computed by MAGMA54 (v1.6) using the LVMi PLINK GWAS results as input. This MAGMA analysis was carried out using FUMA with the default settings. Results of these two analyses are detailed in Supplementary Note 1.
Functional interpretation of lo-siRF-prioritized variants
Functional interpretation step 1: Extraction of candidate SNVs and LD structures
Our lo-siRF approach described above identified a total of 283 SNVs located within 6 LVMi genetic risk loci (Fig. 2f). In order to explore the functional consequences of these prioritized genetic variants and identify genes that are potentially involved in the trait of LV hypertrophy, we performed functional mapping and annotation using a web-based platform, FUMA (v1.5.4)44. The SNP2GENE function in FUMA was used to incorporate LD structure and prioritize candidate genes. Taking the GWAS summary statistics from PLINK37 and BOLT-LMM38 as an input, we submitted the 283 lo-siRF-prioritized SNVs into SNP2GENE as predefined SNVs. This allows SNP2GENE to define LD blocks for each of the 283 lo-siRF-prioritized SNV and use the given 283 SNVs and SNVs in LD with them for further annotations. We adopted the default r2 threshold (i.e., 0.6) for defining independent significant SNVs. Because any two of the 283 lo-siRF-prioritized SNVs are in LD with each other at r2 < 0.6, all of the 283 SNVs were defined as independent significant SNVs by FUMA. In order to match the population group used for our lo-siRF prioritization, the reference panel from UKBB release 2b that FUMA created for British and European subjects was chosen for the computation of r2 and minor allele frequencies. A total of 572 candidate SNVs in strong LD (r2 < 0.6) with any of the 283 independent significant SNVs were extracted from both the inputted GWAS (with the maximum p-value threshold being 0.05) and the reference panel. These 572 candidate SNVs were then assigned to one of the six lo-siRF-identified loci (Fig. 2f) based on its corresponding independent significant SNV, which showed the maximum r2 value in LD with the given candidate SNV. A combination of the 283 independent significant SNVs and the 572 FUMA-extracted candidate SNVs in LD with the independent significant SNVs (details in Extended Data 4) was defined as the lo-siRF-prioritized SNV set, which was used to generate the list of lo-siRF-prioritized genes (Fig. 4a) for the following enrichment analysis (Fig. 4c-4e). As a comparison to the lo-siRF-prioritized SNV set, we uploaded all 1405 GWAS-filtered SNVs (Extended Data 2) as the predefined SNVs in a separate SNP2GENE job. Using the same approach and parameter settings, 929 independent significant SNVs were identified within the given 1405 GWAS-filtered SNVs, and 5771 candidate SNVs in LD with the 929 independent significant SNVs were extracted by FUMA. A combination of the 929 independent significant SNVs and the 5771 candidate SNVs were defined as the reference SNV set. This reference SNV set is purely generated from GWAS prioritization and excludes the evaluation of epistatic effects between genetic variants by lo-siRF. Genes functionally mapped from the reference SNV set was used as a comparison group for the lo-siRF-prioritized gene list in the following enrichment analysis to explore the specific contribution of the identified epistatic genes in the enriched gene ontologies, pathways, and transcription factors (Fig. 4c-4e).
Functional interpretation step 2: ANNOVAR enrichment test
To evaluate the functional consequences of the lo-siRF-prioritized genetic loci, we performed ANNOVAR enrichment test of the aforementioned 283 independent significant SNVs and 572 candidate SNVs in LD with them against the selected reference panel in FUMA. The FUMA SNP2GENE process generated unique ANNOVAR90 annotations for all the identified SNVs. The enrichment score for a given annotation in a given lo-siRF-prioritized genetic locus (Fig. 3b) was computed as the proportion of SNVs associated to that locus with the given annotation divided by the proportion of SNVs with the same annotation relative to all available SNVs in the reference panel. For the ith ANNOVAR annotation in the jth lo-siRF-prioritized locus, the enrichment p-value was computed by performing a two-sided Fisher’s exact test on the 2-by-2 contingency table containing nj (i), , and . Here, nj (i) is the number of SNVs with the ith annotation in the jth lo-siRF-prioritized locus, N (i) is the number of SNVs with the ith annotation in the reference panel, is the summation of nj (i) for all available annotations in the jth lo-siRF-prioritized locus, and is the summation of N (i) for all available annotations in the reference panel. Detailed information can be found in Extended Data 5.
Functional interpretation step 3: Functional annotations
In addition to ANNOVAR annotations, FUMA annotated all 283 independent significant SNVs and 572 SNVs in LD with them for functional consequences on potential regulatory functions (core-15 chromatin state prediction and RegulomeDB score) and deleterious effects (CADD score). In particular, the core-15 chromatin state was annotated to all SNVs of interest by ChromHMM45 derived from 5 chromatin markers (H3K4me3, H3K4me1, H3K36me3, H3K27me3, and H3K9me3) for 127 tissue/cell types, of which left ventricle (E095), right ventricle (E105), right atrium (E104), and fetal heart (E083) were taken into consideration in this study (Fig. 3a, circle 7). Data and a corresponding description of the core-15 chromatin state model can be found at https://egg2.wustl.edu/roadmap/web_portal/chr_state_learning.html. RegulomeDB44,46 annotations guide interpretation of regulatory variants through a seven-level categorical score, of which the category 1 (including 6 subcategories ranging from 1a to 1f) indicates the strongest evidence for a variant to result in a functional consequence. Because the RegulomeDB database (v1.1) used in FUMA has not been updated, we queried all SNVs identified by lo-siRF and FUMA in the RegulomeDB database v2.2 (https://regulomedb.org/regulome-search). Annotations for deleteriousness were obtained from the CADD database (v1.4)47 by matching chromosome, position, reference, and alternative alleles of all SNVs. High CADD scores indicate highly deleterious effects of a given variant. A minimum threshold CADD score of 12.37 was suggested by Kircher et al.47. In addition to the aforementioned functional annotations, we extracted information of eQTLs and sQTLs for all independent significant SNVs and SNVs that are in LD with one of the independent significant SNVs from GTEx v848. The eQTL information was used for eQTL gene mapping as described in the following section.
Functional interpretation step 4: Functional gene mapping
In SNP2GENE, we performed three functional gene mapping strategies – positional, eQTL, and 3D chromatin interaction mapping – using the lo-siRF-prioritized SNV set and the reference SNV set described in the Methods section Functional interpretation step 1: Extraction of candidate SNVs and LD structures. For positional mapping44,46, a default value of 10 kb was used as the maximum distance between SNVs and genes. For eQTL mapping, cis-eQTL information of heart left ventricle, heart atrial appendage, and muscle skeletal tissue types from GTEx v848 was used. Only significant SNV-gene pairs (FDR < 0.05 and p < 1E-3) were used for eQTL mapping. For 3D chromatin interaction mapping, Hi-C data of left ventricle tissue from GSE87112 was chosen with a default threshold of FDR < 1E-6. A default promoter region window was defined as 250 bp upstream and 500 bp downstream of TSS44,46. Using these three gene mapping strategies, we mapped the lo-siRF-prioritized SNV set to 21 protein-coding genes (Fig. 4a), of which 20 are HGNC-recognizable. Each of the 21 genes was also functionally linked to a specific lo-siRF-prioritized LV hypertrophy risk locus (Fig. 2f), to which the highest proportion of SNVs mapped to the given gene were assigned. A Circos plot (Fig. 3a) showing comprehensive information of the lo-siRF-prioritized epistatic interactions, FUMA-prioritized eQTL SNV-to-gene connections and 3D chromatin interactions, as well as LD structures and prioritized genes was created by TBtools91. We then submitted these 21 genes to the GENE2FUNC process in FUMA and obtained GTEx gene expression data for 19 (out of the 21) genes across multiple tissue types (Fig. 4b). In addition, we used the same approach and mapped the reference SNV set (mentioned in the Methods section Functional interpretation step 1: Extraction of candidate SNVs and LD structures) to a separate gene set that contains 382 HGNC-approved genes. The lo-siRF-prioritized gene set and the reference gene set were used for gene set enrichment analysis that are described in the following sections.
Gene ontology and pathway enrichment analysis
Genes that co-associate to shared gene ontology (GO) and pathway terms are likely to be functionally related. To assess the differential GO and pathway co-association among the lo-siRF-prioritized genes relative to their counterparts that were deprioritized by lo-siRF, we performed an integrative GO and pathway enrichment analysis followed by an exhaustive permutation of co-association scores between any possible gene-gene combinations found in the aforementioned 382 HGNC-approved genes (see the Methods section Functional interpretation step 4: Functional gene mapping).
In order to improve GO and pathway prioritization, we adopted the concept from Enrichr-KG57 and ChEA356 to assess enrichment analysis results across libraries and domains of knowledge as an integrated network of genes and their annotations. We first queried the 382 HGNC-approved genes from the reference gene set against various prior-knowledge gene set libraries in Enrichr55 (https://maayanlab.cloud/Enrichr/). We selected five representative libraries from the GO and pathway Enrichr categories as follows: GO biological process92,93, GO molecular function92,93, MGI Mammalian Phenotypes94, Reactome pathways95, and KEGG pathways96. Other FUMA-extracted genes that were not approved by HGNC using synonyms or aliases were discarded. This enrichment analysis allowed us to search for a union of enriched GO or pathway terms and their correspondingly annotated gene sets, from which we built a co-association network. According to the method by Enrichr-KG57, nodes in the co-association network are either the enriched GO and pathway terms or genes.
To measure the degree of co-association to specific GO and pathway terms for two given interactor genes, we computed a co-association score for each of the 72,771 possible gene-gene combinations (from the 382 queried genes). The co-association score was calculated by R = N(A⋂B)⧸N(A⋃B). Here, N(A⋂B) denotes the number of GO or pathway terms that were significantly enriched for both gene A and gene B in the proposed gene-gene combination, and N(A⋃B) is the number of GO or pathway terms that were enriched for either gene A or gene B. For cases where N(A⋃B) = 0, we defined R = 0 to indicate that no GO or pathway terms were found to be co-associated with the respective gene-gene combinations. Of the 382 HGNC-approved genes, 20 genes were mapped to lo-siRF-prioritized loci by FUMA functional gene mapping (one of the 21 lo-siRF-prioritized genes is HGNC-unrecognizable and is discarded). We compared the co-association scores R for gene-gene combinations in the lo-siRF-prioritized gene set relative to the full distribution of R provided by an exhaustive permutation of all possible gene-gene combinations in the set of 382 HGNC-approved genes. The ranking of gene-gene combinations was determined by the two-sided empirical p-values. Fig. 4c displays significant co-associations (empirical p < 0.05) between enriched GO or pathway terms and genes functionally mapped to lo-siRF-prioritized epistatic and hypostatic loci (Fig. 2f). Further details can be found in Extended Data 6.
Transcription factor enrichment analysis
Owing to the limitations and biases of various specific assays, we performed an integrative transcription factor (TF) enrichment analysis against multiple annotated gene set libraries in ChEA356 and Enrichr55. To preserve the variety of library types, we assembled 9 gene set libraries (Fig. 4d) from distinct sources as follows:
Putative TF target gene sets determined by ChIP-seq experiments from ENCODE97;
Putative TF target gene sets determined by ChIP-seq experiments from ReMap98;
Putative TF target gene sets determined by ChIP-seq experiments from individual publications56;
TF co-expression with other genes based on RNA-seq data from GTEx48;
TF co-expression with other genes based on RNA-seq data from ARCHS499;
Single TF perturbations followed by gene differential expression55;
Putative target gene sets determined by scanning PWMs from JASPAR100 and TRANSFAC101 at promoter regions of all human genes;
Gene sets predicted by transcriptional regulatory relationships unraveled by sentence-based text-mining (TRRUST)102;
Top co-occurring genes with TFs in a large number of Enrichr queries56.
Of the mentioned 9 gene set libraries, libraries 1, 3, and 5 were assembled by combining gene set libraries downloaded from both ChEA356 and Enrichr55. Libraries 2, 4, and 9 were downloaded from ChEA356. Libraries 6, 7, and 8 were downloaded from Enrichr55. According to the integration method by ChEA356, for libraries in which multiple gene sets were annotated to the same TF, the unique gene set with the lowest FET p-value was used. As mentioned in previous sections, we used separate FUMA SNP2GENE processes and functionally mapped lo-siRF-prioritized SNVs and all GWAS-filtered SNVs to a lo-siRF-prioritized gene set (20 HGNC-recognizable genes) and a reference gene set (382 HGNC-recognizable genes), respectively. Because the lo-siRF-prioritized gene set is a subset of the reference gene set, we considered the 362 genes complementary to the lo-siRF-prioritized gene set as the lo-siRF-deprioritized gene set. Taking the 20 lo-siRF-prioritized genes and 263 lo-siRF-deprioritized genes as two separate input gene sets, we performed enrichment analysis against the 9 gene set libraries. For each of the 9 libraries, we ranked the significance of overlap between the input gene set and the TF-annotated gene sets in that library by FET p-values. Those TFs with identical FET p-values were ranked by the same integer number. A scaled rank was then assigned to each TF by dividing the corresponding integer rank by the maximum integer rank in its respective library. We then integrated the 9 sets of TF rankings and re-ordered the TFs by two sequential criteria: (1) the number of libraries that display a significant overlap with the input gene set (FET p < 0.05) and (2) the mean scaled rank across all libraries containing that TF. Using this method, we prioritized two distinct sets of TFs for the lo-siRF-prioritized genes (Fig. 4d, top) and lo-siRF-deprioritized genes (Fig. 4d, bottom).
The above analyses aim to enrich TFs based on their associations to given sets of individual genes rather than co-associations to gene pairs55,56. To evaluate the differential TF co-association between the lo-siRF-prioritized genes relative to the lo-siRF-deprioritized genes, we used the same approach described in the Methods section Gene ontology and pathway enrichment analysis. A TF co-association score was computed for each of the 72,771 possible gene-gene combinations from the 382 genes. Still, the TF co-association score was computed by R = N(A⋂B)⧸N(A⋃B), except that N(A⋂B) and N(A⋃B) denote the number of enriched TF terms instead of GO or pathway terms. Pairwise interactions between lo-siRF-prioritized genes were extracted and ranked by the empirical p-values (Fig. 4e) from an exhaustive permutation of TF co-association scores for the 72,771 possible gene-gene combinations. Details regarding TF enrichment from both lo-siRF-prioritized and lo-siRF-deprioritized genes and TF co-association strengths can be found in Extended Data 7.
Disease-state-specific gene co-expression network analysis
In order to evaluate the connectivity between genes and their potential roles in the transition from healthy to failing myocardium, we compared gene-gene connectivity and changes in the topological structure between gene co-expression networks for healthy and failing human heart tissues (Fig. 5). To construct gene co-expression networks, cardiac tissue samples from 177 failing hearts and 136 donor, non-failing (control) hearts were collected from operating rooms and remote locations for RNA expression measurements. We performed weighted gene co-expression network analysis (WGCNA) on the covariate-corrected RNA microarray data for the control and heart failure networks separately (Fig. 5a). Detailed steps for generating these co-expression networks, which included calculating the correlation matrix, TOM transformation, and Dynamic Tree Cut module finding, are described in our previous study60, and data for these networks is available at https://doi.org/10.5281/zenodo.2600420. To evaluate the degree of connectivity between correlating genes in each of the networks, we compared the edge weights between the lo-siRF-prioritized genes demonstrated in this study relative to the distribution of all possible pairwise combinations of genes (Fig. 5b and 5c). We also evaluated the difference of edge weights (Z-score normalized) between the control and heart failure networks to understand how these gene-gene connectivities change between non-failing and failing hearts (Fig. 5d). The two-tailed empirical p-value represents the proportion of the absolute difference in edge weights of all gene pairs that exceed the absolute difference score for gene pairs of interest. We then compared the structure of modules derived from dendrograms on the WGCNA control and heart failure networks (Extended Data Fig. 6). Modules were labeled according to Reactome enrichment analysis of genes within each module. The full gene module descriptions and Benjamini-Hochberg-adjusted enrichment p-values can be found in the Supplementary Data 5 and 6 in the study by Cordero et al.60.
Induced pluripotent stem cell cardiomyocytes differentiation
The studied patient-specific human induced pluripotent stem cells (hiPSCs) were derived from a 45-year-old female proband with a heterozygous MYH7-R403Q mutation. Derivation and maintenance of hiPSC lines were performed following Dainis et al.33. Briefly, hiPSCs were maintained in MTeSR (StemCell Technologies) and split at a low density (1:12) onto fresh 1:200 matrigel-coated 12 well plates. Following the split, cells were left in MTeSR media supplemented with 1 μM Thiazovivin. The hiPSCs were maintained in MTeSR until cells reached 90% confluency, which began Day 0 of the cardiomyocyte differentiation protocol. Cardiomyocytes were differentiated from hiPSCs using small molecule inhibitors. For Days 0-5, cells were given RPMI 1640 medium + L-glutamine and B27 -insulin. On Days 0 and 1, the media was supplemented with 6 μM of the GSK3β inhibitor, CHIR99021. On Days 2 and 3, the media was supplemented with 5 μM of the Wnt inhibitor, IWR-1. Media was switched to RPMI 1640 medium + L-glutamine and B27 + insulin on Days 6-8. On Days 9-12, cells were maintained in RPMI 1640 medium + L-glutamine -glucose, B27 + insulin, and sodium lactate. On Day 13, cells were detached using Accutase for 7-10 minutes at 37 ℃ and resuspended in neutralizing RPMI 1640 medium + L-glutamine and B27 + insulin. This mixture was centrifuged for 5 minutes at 1000 rpm (103 rcf). The cell pellet was resuspended in 1 μM thiazovivin supplemented RPMI 1640 medium + L-glutamine and B27 + insulin. For the rest of the protocol (Days 14-40), cells were exposed to RPMI 1640 medium + L-glutamine -glucose, B27 + insulin, and sodium lactate. Media changes occurred every other day on Days 14-19 and every three days for Days 20-40. On Day 40, cardiomyocytes reached maturity.
RNA silencing in induced pluripotent stem cell-derived cardiomyocytes
Mature hiPSC-derived cardiomyocytes were transfected with Silencer Select siRNAs (Thermofisher) using TransIT-TKO Transfection reagent (Mirus Bio). Cells were incubated for 48 hours with 75 nM siRNA treatments. Four wells of cells were transfected with each of the six siRNAs: scramble, CCDC141 (ID s49797), IGF1R (ID s223918), TTN (ID s14484), CCDC141 and IGF1R, and CCDC141 and TTN. After 2 days, hiPSC-CMs were collected for RNA extraction.
RT-qPCR analysis for siRNA gene silencing efficiency
Following cell morphology measurement, all cells for each condition were centrifuged for 5 minutes at 1000 rpm (103 rcf). Cell pellets were frozen at −80 ℃ prior to RNA extraction. RNA was extracted using Trizol reagent for RT-qPCR to confirm gene knockdown occurred. Reverse Transcription of RNA was done using High-Capacity cDNA Reverse Transcription Kit (Thermofisher). qPCR of the single stranded cDNA was performed using TaqMan Fast Advanced MM (Thermofisher) with the following annealing temperatures: 95°C 20” and 40 cycles of 95°C 1” and 60°C 20”. qPCR of the silenced genes was performed using TaqMan® Gene Expression Assays, including CCDC141 (Hs00892642_m1), IGF1R (Hs00609566_m1), and TTN (Hs00399225_m1). For gene-silencing efficiency analysis, gene RPLP0 (Hs00420895_gH) was used as a reference gene. Data were analyzed using the delta-delta Ct method.
Cell sample preparation for cell morphology measurement
Following siRNA treatments, cells were detached for microfluidic single-cell imaging using a mixture of 5 parts Accutase and 1 part TrypLE, treated for 6 minutes at 37 ℃. Cells were then added to the neutralizing RPMI 1640 medium + L-glutamine and B27 + insulin. These mixtures were centrifuged for 5 minutes at 1000 rpm (103 rcf). For each gene-silencing condition, the four wells of cells were resuspended in 4 mL of the MEM medium, which is composed of MEM (HBSS balanced) medium, 10% FBS, and 1% Pen Strep (Gibco). Cells were filtered with 100 μm strainers (Corning) before adding into the microfluidic devices.
Microfluidic inertial focusing device
We developed a new spiral inertial microfluidics system on the basis of the study by Guan et al.31 to focus randomly suspended cells into separate single streams based on cell size for high-resolution and high-throughput single-cell imaging. The microfluidic device (Extended Data Fig. 7) contains 5 loops of spiral microchannel with a radius increasing from 3.3 mm to 7.05 mm. The microchannel has a cross-section with a slanted ceiling, resulting in 80 μm and 150 μm depths at the inner and outer side of the channel, respectively. The channel width is fixed to 600 μm. The 495 μm wide slanted region of the channel ceiling is composed of ten 7 μm deep stairs. This particular geometry induces strong Dean vortices in the outer half of the channel cross-section, leading to high sensitivity of size separation and cell focusing. The device has two inlets at the spiral center to introduce cell suspensions and sheath flow of fresh medium. At the outlet region, the channel is expanded in width and split into two outlet channels with a width of 845 μm for the top outlet and 690 μm for the bottom outlet. Depths of the two outlet channels are designed to create equal hydraulic resistance. The top and bottom outlet channels are connected to 80 μm and 50 μm deep straight observation channels for high-throughput cell imaging.
Microdevice fabrication
The spiral microchannel was fabricated by CNC micromachining a piece of laser-cut poly (methyl methacrylate) (PMMA) sheet, which was bonded with a PMMA chip machined only with the inlet channels and another blank PMMA chip using a solvent-assisted thermal binding process to form the enclosed channel29. Before bonding, PMMA chips were cleaned with acetone, methanol, isopropanol and deionized water in sequence. Droplets of a solvent mixture (47.5% DMSO, 47.5% water, 5% methanol) were evenly spread over the cleaned chips. The PMMA chips were assembled appropriately and clamped using a customized aluminum fixture, and then heated in a ThermoScientific Lindberg Blue M oven at 96 ℃ for 2 hrs. After bonding, fluid reservoirs (McMaster) were then attached to the chips using a two-part epoxy (McMaster). Microchannels were flushed with 70% ethanol followed by DI water for sterilization.
High-throughput single-cell imaging
Before each experiment, microchannels were flushed with 3 mL of the MEM medium. Prepared cell samples and fresh MEM medium were loaded into 3 mL syringes, which were connected to the corresponding microchannel inlets using Tygon PVC tubing (McMaster). Both cells and the fresh MEM medium were infused into the microchannel using a Pico Plus Elite syringe pump (Harvard Apparatus) at 1.2 mL/min. Microscope image sequences of cells focused to the top and bottom observation channels were captured using a VEO 710S high-speed camera (Phantom) with a sampling rate of 700 fps and a 5 μsec light exposure.
Image analysis for cell feature extraction
For each gene-silencing condition of each biological repeat, 21,000 images were processed to extract cell morphology features. To analyze cell size and shape changes induced by gene silencing, we developed a MATLAB-based image analysis pipeline, which includes three major steps: image preparation, feature extraction, and image post-processing (Fig. 6c). In step one, image sequences were fed into the MATLAB program and subtracted from the corresponding background image to correct any inhomogeneous illumination. The program automatically generates background images, in which each pixel value is computed as the mode pixel intensity value among the same pixel of the entire corresponding image sequence. After illumination correction, step two detects cell edges by looking for the local maxima of the bright field intensity gradient, following which the program closes edge gaps, removes cells connected to the image borders, cleans small features (noise), and then fills holes to generate binary images and centroid positions for each single cell. Cell locations were then traced and stuck cells were removed by a double-counting filter if present. The double-counting filter excludes measurements collected around the same location with similar cell sizes using a Gaussian kernel density method (bandwidth = 0.09) when the estimated density for a certain location and size exceeds a particular threshold. The maximal density value for experimental runs where no repeated measurements were observed was used as the threshold. This procedure was manually validated using visual inspection of the removed cells. Binary images passing the double-counting filter were used to create coordinates (X, Y) of cell outlines, which leads to a range of cell size and shape parameters, including cell diameter and area, solidity, roundness error, circularity, and intensity spatial relationship enclosed within the cell boundary. Cell area was computed as the 2D integration of the cell outline, and the cell diameter was computed as . Solidity is the ratio of cell area to the area of the smallest convex polygon that contains the cell region. Roundness error was computed as the ratio between the standard deviation and mean of radii on the cell outline measured from the centroid. Circularity was calculated as 4Areaπ/Perimeter2. The 2D intensity distributions within cell outlines were used to derive peak locations and count peak numbers, which is a measure of intensity spatial relationship and a gating parameter to remove clumped cells. In the post-processing step, data were cleaned using three filters with the following gating threshold. To remove large clumps, the peak-solidity filter removes data outside of the polygonal region defined by {(0.9, 0), (0.9, 3.2), (0.934, 8.26), (1, 28), (1, 0)} in the (solidity, peak No.) space. Then, the roundness filter removes cells with weird shapes by excluding data with a roundness error higher than 0.3 or a circularity lower than 0.6. Finally, the small size filter removes cell debris whose major diameter is lower than 15 μm (12 μm) or minor diameter is lower than 12 μm (10 μm) for images photographed at the top (bottom) outlet microchannels.
Statistical assessment of gene-silencing effects in high-throughput single-cell experiments
To analyze the experimental results, we began by examining how and where cell size distributions differ between each of the gene/gene-pair silenced groups and their respective scrambled control groups. We thus conducted various statistical analyses to investigate these size distribution disparities. First, we compared the difference in median cell size (i.e., diameter measured in μm) between the gene/gene-pair silenced cells and their scrambled controls. We performed two different tests – a Wilcoxon signed rank test (Fig. 6f) and a bootstrap quantile test at the 0.5 quantile level. In accordance with the PCS framework, we used two different tests to ensure that our findings are robust to this arbitrary modeling choice and that the underlying assumptions do not drive our results. We note that the difference in median cell size was of greater interest than the difference in mean cell size due to the heavy right-skewness of the cell size distribution. Still, as an additional stability check, we performed a bootstrap-t test for the difference in trimmed means with varying levels of trimming (ranging from 0-0.3). These differences in trimmed mean tests (data not shown) yielded similar results to the tests for difference in medians, providing further evidence for the robustness of our conclusions. Secondly, in addition to comparing differences in central behavior, we compared differences in upper quantiles of size distributions for the gene/gene-pair silenced cells versus the scrambled controls. Identifying cell size differences at these upper quantiles, which focus on the larger, hypertrophic cells, is particularly relevant for the pathologic phenotype of cardiac hypertrophy and its clinical implications. To assess these differences in cell sizes at upper quantiles, we performed a bootstrap quantile test at the 0.6, 0.7, 0.8, and 0.9 quantile levels (Extended Data Fig. 8). All tests are performed on each experimental batch separately. To be as conservative as possible when claiming a significant effect, the maximum p-value across batches is reported in the main text. Similar analyses were conducted for assessing differences in morphological features (i.e., cell roundness error and normalized peak number).
Statistical assessment of non-additivity in high-throughput single-cell experiments
We also assessed whether the silenced gene pairs (i.e., CCDC141-TTN and CCDC141-IGF1R) are interacting in an additive or non-additive (i.e., epistatic) way to affect cell size in the high-throughput single-cell experiments (Fig. 6g). More formally, to assess this (non-)additivity for a given gene pair, say gene 1 and gene 2, we fit the following quantile regression:
where yi is the diameter (μm) of cell i, xi1 is an indicator whether gene 1 was silenced in cell i, xi2 is an indicator whether gene 2 was silenced in cell i, gi encodes the batch identifier from which cell i came (so that µgi is a batch effect term), and εi is the random error or noise term for cell i. This regression was fitted using the scrambled control cells (xi1 = xi2 = 0), gene 1-silenced cells (xi1 = 1; xi2 = 0), gene 2-silenced cells (xi1 = 0; xi2 = 1), and gene 1 – gene 2 jointly silenced cells (xi1 = xi2 = 1). Under this regression model, we tested the null hypothesis of β12 = 0 versus the alternative hypothesis of β12 ≠ 0 via a percentile bootstrap t-test and a traditional t-test (Extended Data 8) for varying quantile levels (0.5, 0.6, 0.7, 0.8, 0.9). A small p-value suggest that the gene pair is a non-additive epistatic interaction under the above model. We note again that two different tests were performed to check the robustness of our conclusions against modeling assumptions associated with each statistical test. To further bolster the robustness of our conclusions, we repeated this assessment of epistasis using the rank-based inverse normal-transformed cell diameter as the response y and under an ordinary linear regression model, finding that both the p-values and the direction of the non-additive interaction effects are similar to the reported quantile regression results. We thus omit these results for brevity. Since these regression models require comparisons between gene-silencing conditions (e.g., silencing CCDC141 and TTN vs only silencing CCDC141 vs only silencing TTN), and gene-silencing efficiency varied across silencing conditions, we only include cells from experimental batches with high gene-silencing efficiencies (>60%) for each regression. This helps to mitigate the possibility that the gene-silencing efficiency differences are driving spurious epistatic signals. We also conducted a simulation study to better understand how differences in gene-silencing efficiencies across batches might impact conclusions in Supplementary Note 3 (available on the website: https://yu-group.github.io/epistasis-cardiac-hypertrophy/simulations_efficiency). In general, we found that differences in our observed gene-silencing efficiencies do not typically lead to high false positive rates under the low signal-to-noise regimes that we expect in reality. We defer further discussion to Supplementary Note 3. Similar analyses were conducted for assessing differences in morphological features (i.e., cell roundness error and normalized peak number).
Data Availability
All genotype and cardiac MRI data used as input to the lo-siRF pipeline are available from the UK Biobank (https://www.ukbiobank.ac.uk/). This work was conducted under the UK Biobank application 22282. GWAS-filtered SNVs using PLINK and BOLT-LMM are summarized in Extended Data 2. Data for the gene co-expression networks from 313 explanted human hearts is available at https://doi.org/10.5281/zenodo.2600420. All other data produced in this study are available upon reasonable request to the authors.
Extended Data Figures
Supplementary Tables
Data availability
All genotype and cardiac MRI data used as input to the lo-siRF pipeline are available from the UK Biobank (https://www.ukbiobank.ac.uk/). This work was conducted under the UK Biobank application 22282. GWAS-filtered SNVs using PLINK34 and BOLT-LMM35 are summarized in Extended Data 2. Data for the gene co-expression networks from 313 explanted human hearts is available at https://doi.org/10.5281/zenodo.2600420.
Code availability
All code for running the lo-siRF analysis and analyzing the experimental results can be found on GitHub (https://github.com/Yu-Group/epistasis-cardiac-hypertrophy). This lo-siRF analysis was conducted using R version 3.6.1, Python 3.6.1, and iRF2.0 (https://github.com/karlkumbier/iRF2.0). The LVMi derivation from cardiac MRI images and corresponding deep learning model have been published elsewere25 (https://github.com/baiwenjia/ukbb_cardiac). PLINK34 (https://www.cog-genomics.org/plink/) and BOLT-LMM35 (https://alkesgroup.broadinstitute.org/BOLT-LMM/BOLT-LMM_manual.html) were used to perform the GWAS dimension reduction. ANNOVAR78 (https://annovar.openbioinformatics.org/en/latest/) was used to map each SNV to a genetic locus within lo-siRF. FUMA GWAS37 (https://fuma.ctglab.nl/) version 1.5.4 was used to functionally annotate SNVs and map to genes.
Contributions
Q.W., T.M.T., C.S.W., J.W.H., A.J.B., R.A., J.B.B., J.R.P., V.N.P., B.Y., and E.A.A. conceived and designed research. J.W.H. performed LVMi extraction from cardiac MRI images. T.M.T., A.A., X.L., M.B., and K.K. performed exploratory data investigations leading to development of lo-siRF. T.M.T., A.A., and B.Y. developed the lo-siRF pipeline; T.M.T. performed the lo-siRF analysis. Q.W. and T.M.T. performed the FUMA SNP2GENE process using lo-siRF-prioritized SNVs and GWAS-filtered SNVs. Q.W. evaluated functional annotation results from FUMA, performed ANNOVAR enrichment test for each lo-siRF loci and functional gene mapping. Q.W. performed integrative biological enrichment analyses and evaluated the co-associations between lo-siRF hypothesized gene-gene interactions and the enriched GOs, pathways, and TFs. E.T.C., C.S.M., W.H.W.T., K.B.M., T.P.C., V.N.P., and E.A.A. contributed to the data collection and construction of WGCNA healthy and heart failure coexpression networks. E.T.C. evaluated connectivity differences of epistatic genes hypothesized by lo-siRF between cardiac co-expression networks of failing and non-failing hearts. Q.W. designed and created microfluidic devices; N.Y., S.C.S. and V.N.P. created gene-silenced hiPSC-CM lines; Q.W. and N.Y. performed the microfluidic single cell imaging experiments. Q.W., O.R., and A.M.K. performed single cell image analysis and morphological feature extraction. Q.W., T.M.T., A.M.K., O.R., C.S.W., V.N.P., J.W.H., B.Y., and E.A.A. interpreted results of experiments; Q.W., T.M.T., and A.M.K. prepared figures; Q.W., T.M.T., C.S.W. and J.W.H. drafted manuscript; All authors contributed in editing and revising manuscript.
Competing interests
E.A.A. is a Founder of Personalis, Deepcell, Svexa, RCD Co, and Parameter Health; Advisor to Oxford Nanopore, SequenceBio, and Pacific Biosciences; and a non-executive director for AstraZeneca. C.S.W. is a consultant for Tensixteen Bio and Renovacor. V.N.P. is an SAB member for and receives research support from BioMarin, Inc, and is a consultant for Constantiam, Inc. and viz.ai. The remaining authors declare no competing interests.
Acknowledgements
The authors would like to acknowledge Dr. David Amar, Dr. Srigokul Upadhyayula, Dr. Haiyan Huang, and Dr. Ziad Obermeyer for their critical comments and discussions on this work, Stanford nanofabrication facility and Elmer Enriquez for the technical support of microfluidic device fabrication, and Dr. Anna Shcherbina and Dr. Manuel Rivas for their technical help with the UK Biobank data. This work was supported by the Chan Zuckerberg Biohub – San Francisco through the Intercampus Research Awards (2019-2022) to R.A., J.R.P., J.B.B., A.J.B., E.A.A., and B.Y. E.A.A. received funding from National Institutes of Health (NIH) through grant number 1R01HL144843. B.Y. received support from National Science Foundation (NSF) through grants DMS-1613002 and IIS 1741340, an NIH grant R01GM152718, and a Weill Neurohub grant. V.N.P. received funding from K08HL143185. T.M.T. was supported by the National Science Foundation (NSF) Graduate Research Fellowship Program DGE-2146752. Q.W. received funding from American Heart Association Postdoctoral Fellowship through grant number 23POST1023278. C.S.W. received support from NIH through grants F32HL160067 and L30HL159413.
Footnotes
We have made the following key changes to improve our study in both computational and experimental perspectives. They include: ● A broader comparison with various commonly used machine learning methods, polygenic risk scores, regression-based pairwise interaction scans, MAPIT, and marginal gene-based methods (e.g., SKAT-O and MAGMA), all of which consistently validate the superior prediction accuracy and prioritizations using lo-siRF. ● Further analyses that underscore the necessity and importance of the binarization step in lo-siRF. ● An additional non-hypertensive cohort analysis which reveals weak pleiotropic effects of the three identified epistatic interactions on hypertension, strengthening the evidence for epistasis. ● New results from additional co-expression network analyses, which reinforced our findings about a strong molecular connectivity between statistical epistasis contributors compared to random gene pairs in healthy hearts and a significantly weakened connectivity in failing hearts at the transcriptomic level. ● Additional assessment of the non-additive (epistatic) effects in the gene-silencing experiments that confirmed the consistency between predicted and biologically observed epistases, demonstrating the existence, strength, and directions of epistatic effects. ● Expanded evaluation of gene knockdown effects on cell size in different scales, which confirmed a stable epistatic effect across different choices of scales and statistical evaluation methods. ● New simulation study which confirmed a small likelihood that variations in gene-silencing efficiencies across experimental batches lead to spurious epistasis signals, particularly in the signal-to-noise ratio regimes relevant to this study. These additional results and extensive justification of our modeling have been incorporated into the revised manuscript and in the following interactive HTML webpage available at: https://yu-group.github.io/epistasis-cardiac-hypertrophy/
Reference
- 1.↵
- 2.↵
- 3.
- 4.
- 5.↵
- 6.↵
- 7.↵
- 8.
- 9.
- 10.↵
- 11.↵
- 12.
- 13.↵
- 14.↵
- 15.↵
- 16.↵
- 17.↵
- 18.↵
- 19.↵
- 20.↵
- 21.↵
- 22.↵
- 23.↵
- 24.↵
- 25.↵
- 26.↵
- 27.↵
- 28.↵
- 29.↵
- 30.
- 31.↵
- 32.↵
- 33.↵
- 34.↵
- 35.↵
- 36.↵
- 37.↵
- 38.↵
- 39.↵
- 40.↵
- 41.↵
- 42.↵
- 43.↵
- 44.↵
- 45.↵
- 46.↵
- 47.↵
- 48.↵
- 49.↵
- 50.↵
- 51.↵
- 52.↵
- 53.↵
- 54.↵
- 55.↵
- 56.↵
- 57.↵
- 58.↵
- 59.↵
- 60.↵
- 61.↵
- 62.↵
- 63.↵
- 64.↵
- 65.
- 66.
- 67.↵
- 68.↵
- 69.↵
- 70.
- 71.↵
- 72.↵
- 73.↵
- 74.↵
- 75.↵
- 76.↵
- 77.↵
- 78.↵
- 79.↵
- 80.↵
- 81.↵
- 82.↵
- 83.↵
- 84.↵
- 85.↵
- 86.↵
- 87.↵
- 88.↵
- 89.↵
- 90.↵
- 91.↵
- 92.↵
- 93.↵
- 94.↵
- 95.↵
- 96.↵
- 97.↵
- 98.↵
- 99.↵
- 100.↵
- 101.↵
- 102.↵