Skip to main content
eLife logoLink to eLife
. 2021 Mar 26;10:e65381. doi: 10.7554/eLife.65381

Evidence for additive and synergistic action of mammalian enhancers during cell fate determination

Jinmi Choi 1, Kseniia Lysakovskaia 1, Gregoire Stik 2, Carina Demel 1, Johannes Söding 3, Tian V Tian 2, Thomas Graf 2, Patrick Cramer 1,
Editors: Chris P Ponting4, Kevin Struhl5
PMCID: PMC8004103  PMID: 33770473

Abstract

Enhancer activity drives cell differentiation and cell fate determination, but it remains unclear how enhancers cooperate during these processes. Here we investigate enhancer cooperation during transdifferentiation of human leukemia B-cells to macrophages. Putative enhancers are established by binding of the pioneer factor C/EBPα followed by chromatin opening and enhancer RNA (eRNA) synthesis from H3K4-monomethylated regions. Using eRNA synthesis as a proxy for enhancer activity, we find that most putative enhancers cooperate in an additive way to regulate transcription of assigned target genes. However, transcription from 136 target genes depends exponentially on the summed activity of its putative paired enhancers, indicating that these enhancers cooperate synergistically. The target genes are cell type-specific, suggesting that enhancer synergy can contribute to cell fate determination. Enhancer synergy appears to depend on cell type-specific transcription factors, and such interacting enhancers are not predicted from occupancy or accessibility data that are used to detect superenhancers.

Research organism: Human

Introduction

Enhancers are cis-regulatory DNA elements that drive the transcription activity of target gene promoters (Beagrie and Pombo, 2016; Long et al., 2016; Spitz and Furlong, 2012). Enhancers contain transcription factor (TF) binding sites, recruit TFs, and drive cell type-specific gene expression programs. Previous studies have defined lineage-determining clusters of enhancers (‘superenhancers’) that span several kilobases of DNA and contain a high density of TF- and mediator-binding sites (Hnisz et al., 2013; Whyte et al., 2013). It has been suggested that the individual constituent enhancers of such clusters cooperate in a synergistic manner to activate target genes (Hnisz et al., 2017; Shin et al., 2016). Cooperation may be achieved by liquid–liquid phase separation of general and gene-specific TFs (Hnisz et al., 2017).

Enhancer cooperation has been investigated at the level of individual genes. Available studies suggest that enhancers can cooperate in an additive or a synergistic manner. Genetic in vivo dissection of the enhancer elements of the α-globin superenhancer has shown that the activity of the α-globin genes increased linearly with the number of enhancers used, showing additive enhancer cooperation (Hay et al., 2016). Additive cooperation was also demonstrated for enhancers within the Myc-regulating superenhancer and for limb enhancers (Bahr et al., 2018; Osterwalder et al., 2018). Non-synergistic enhancer cooperation was further observed in mouse embryonic stem cells (Moorthy et al., 2017) where deletion of an individual enhancer resulted in only a small effect on target gene expression (Moorthy et al., 2017). There is also evidence that two enhancers can synergistically activate selected target genes (Fulton and van Ness, 1994; Guerrero et al., 2010; Maekawa et al., 1989; Stine et al., 2011). Moreover, multiple enhancers can interact simultaneously with their target gene promoter in mouse and human cells (Allahyar et al., 2018; Oudelaar et al., 2018). Evidence for enhancer–enhancer interactions was also obtained in Drosophila (Lim et al., 2018).

Despite these studies, the functional cooperation between enhancers over time has not yet been studied in a native genomic context and a genome-wide manner. As a consequence, it is unknown to what extent enhancers cooperate dynamically in cells and whether they do so additively or synergistically or both. To study this, enhancer and promoter activity must be monitored over time with a non-perturbing genome-wide method. We have previously reported such a method called transient transcriptome sequencing (TT-seq). TT-seq combines short-term metabolic RNA labeling (5 min) with sequencing of newly synthesized RNA fragments and provides a genome-wide unbiased view of RNA synthesis activity (Schwalb et al., 2016). The fragments are derived from all RNA species, including short-lived non-coding RNAs such as enhancer RNA (eRNA) and messenger RNA (mRNA) (Schwalb et al., 2016).

TT-seq can monitor changes in enhancer and promoter activities over time with great sensitivity. During T-cell stimulation, transcription from enhancers and promoters of responsive genes is activated simultaneously (Michel et al., 2017). Enhancers can be paired with their putative target gene promoters based on their proximity (Michel et al., 2017). Using eRNA production as a proxy for enhancer transactivation activity (Henriques et al., 2018; Mikhaylichenko et al., 2018), TT-seq is very well suitable to identify active enhancers, to pair enhancers with their putative target promoters, and to measure the transcription activity of enhancers and promoters genome-wide. Putative enhancers can be detected by mapping of chromatin signatures (Creyghton et al., 2010; Heintzman et al., 2007; Robertson et al., 2008). However, these techniques have limitations if time-resolved analysis in a dynamic system is required. Also, enhancers can be removed by genome editing but this is not readily possible for thousands of putative enhancer regions.

To address the question of enhancer cooperation during cell type determination, we used a transdifferentiation system driven by a single TF (Rapino et al., 2013). In this system, induction of the TF C/EBPα converts human leukemic B cells into macrophage-like cells within 7 days in a nearly synchronous and efficient manner (Rapino et al., 2013). Transdifferentiation involves dramatic changes in gene expression, including silencing and activation of cell type-specific genes (Rapino et al., 2013). At the end of the process, cells gain the characteristics of macrophages, including acquisition of adherence, phagocytic activity, and inflammatory response (Gaidt et al., 2016; Rapino et al., 2013). Due to these properties, the C/EBPα-induced transdifferentiation system enables quantitative data acquisition and analysis and is ideally suited for addressing mechanistic questions on enhancer function and cooperation in vivo.

Here we report a genome-wide multi-omics data set for C/EBPα-induced transdifferentiation of human B cells to macrophages . In-depth analysis of the data provided general insights into the order of events that establish active enhancers. Furthermore, TT-seq allowed us to identify transcriptionally active enhancers, to pair these with their putative target promoters, and to analyze the changes in transcription activity of enhancers and promoters over time. Our analyses revealed that most enhancers that drive the expression of a common target gene do so in an additive manner. However, for about one-fifth of the enhancers tested (136 in total), the change in transcription activity of the target gene was larger than the change attributable to the sum of the enhancer activities, indicating that enhancers cooperate synergistically at these loci to drive target gene transcription.

Results

RNA-seq reveals two distinct transitions during transdifferentiation

To study the temporal order of gene regulatory events during transdifferentiation of human precursor leukemia B cells to macrophage-like cells, the master TF C/EBPα fused to estrogen receptor were activated by addition of beta estradiol (Figure 1A). As revealed by FACS analysis of the macrophage marker CD14 and the B cell marker CD19, as well as additional markers monitored by RT-qPCR, transdifferentiation was efficient and occurred in a nearly synchronous manner (Figure 1—figure supplement 1A and B). We then applied several genome-wide techniques (Figure 1B) to monitor RNA metabolism during transdifferentiation. RNA-seq at 0, 12, 24, 30, 36, 48, 72, 96, and 168 hr after induction revealed that a total of 5516 protein-coding mRNAs changed their levels significantly (false discovery rate [FDR] < 0.05) and by at least a twofold change (|log2(FC)|>1) (Figure 1C). Upregulated genes were enriched for the terms signaling and immune system processes, whereas downregulated genes were enriched for chromosome organization and cell cycle functions (Figure 1—figure supplement 1C), consistent with the fact that macrophages become quiescent (Rapino et al., 2013).

Figure 1. RNA-seq monitors two transitions during transdifferentiation.

(A) C/EBPα induces transdifferentiation from precursor leukemia B-cell to macrophage-like cells. (B) Genome-wide data sets collected at different time points during transdifferentiation. (C) Heatmap of differentially expressed genes (n = 5516) from RNA-seq (|log2(FC)|>1, FDR < 0.05) for seven clusters obtained with k-mean clustering. (D) PCA plot of differentially expressed genes from RNA-seq during transdifferentiation.

Figure 1.

Figure 1—figure supplement 1. Transcription activity during transdifferentiation.

Figure 1—figure supplement 1.

(A) B-cell or macrophage marker gene expressions during the transdifferentiation are tested by quantitative PCR (qPCR). C/EBPα was induced for 0 hr, 24 hr, 96 hr, and 168 hr. Gene expressions of indicated genes were quantified using GAPDH expression. Relative expression to 0 hr for B cell markers and to 168 hr for macrophage markers is shown. Standard deviation from three independent experiments are shown as error bars. (B) FACS analysis with antibodies against CD19 (B-cell specific surface marker) and CD14 (macrophage specific surface marker). (C) Gene ontology analysis on the upregulated or downregulated genes during the transdifferentiation, determined by RNA-seq. Bonferroni corrected p-values were transformed by log10. (D) Heatmap of sample distances (similarities) between the indicated time points based on gene expressions during the transdifferentiation. (E) Heatmap of gene expression in principle component 1 or 2 (PC1 or 2) from Figure 1D. (F) Gene ontology analysis on the PC1 or two from E and Figure 1D.

Changes in gene expression occurred in two major transitions, as indicated by principal component analysis and Euclidean distance analysis of RNA-seq data (Figure 1D and Figure 1—figure supplement 1D). The first transition occurred rapidly, between 0 and 12 hr, and was followed by a late transition from 48 to 72 hr (Figure 1D). Therefore, the time points of 12 hr and 48 hr represented intermediate states. The variance of the first principal component (60.9%) corresponded mainly to genes involved in immune system processes, whereas the variance of the second component (35.7%) was mainly due to transiently responsive genes involved in developmental processes (Figure 1—figure supplement 1E and F). The data also identified two sets of genes exclusively expressed either in B-cells (cluster ‘Early’, 896 genes) or in macrophage-like cells (cluster ‘Late1’ and ‘Late2’, 1083 and 1046 genes), which we refer to as 'cell type-specific genes' (Figure 1C). Based on the observed changes in RNA levels, we divided the transdifferentiation process into an early transition, which leads from B-cells to an intermediate state, and a late transition, which leads from an intermediate state to macrophage-like cells. The classification of two distinct transitions facilitated further analyses of cell type-specific gene expression during transdifferentiation.

TT-seq monitors transcriptionally active putative enhancers

We performed TT-seq at 0 hr, 12 hr, 24 hr, and 96 hr post induction to monitor changes in newly synthesized RNA, including eRNAs (Figure 1B). We also measured chromatin accessibility, that is, regions of nucleosome depletion, using the Assay for Transposase-Accessible Chromatin (ATAC-seq). We further used chromatin immunoprecipitation (ChIP-seq) to monitor changes in genome-wide occupancy with C/EBPα, and monomethylation at histone H3 lysine 4 (H3K4me1) as a marker of primed enhancers (Figure 1B).

We then used TT-seq data to annotate transcription units (TUs) as described (Michel et al., 2017), and identified 43,193 TUs (Materials and methods). Of these, 9993 and 586 had been annotated as mRNAs and lincRNA, respectively, by GENCODE (Harrow et al., 2012). Other TUs were classified as downstream (ds) RNAs, upstream antisense (ua) RNAs and convergent (conv) RNAs based on their locations with respect to mRNAs (4033, 1896, and 555 TUs, respectively). Of the remaining 26,130 TUs, 8165 fell into regions that were marked by H3K4me1 and depleted for nucleosomes that do not overlap with more than 20% of transcripts annotated in Gencode (Materials and methods) (Figure 2—figure supplement 1A). We therefore refer to their associated RNAs as eRNAs (Figure 2A). Regions covered by eRNAs that fell within 1 kb of each other were merged, as exemplified for the −148 kb enhancer of KLF4 (Figure 2B).

Figure 2. TT-seq identifies transcriptionally active enhancers.

(A) TT-seq identified various classes of RNAs including previously annotated stable messenger RNAs (mRNAs), novel transient enhancer RNAs (eRNAs), and other noncoding RNAs (downstream RNAs (dsRNAs), upstream antisense RNAs (uaRNAs), long intergenic noncoding RNAs (lincRNAs), convergent RNAs (convRNAs)). (B) TT-seq coverage tracks on KLF4 enhancer at 96 hr exemplify how RNAs that were synthesized within 1 kb of another were merged into a single enhancer. (C) Dynamically regulated eRNAs (|log2(FC)|>1, FDR < 0.05) were clustered using k-means clustering according to the kinetics of synthesis level changes during transdifferentiation. Black line represents median. Red, yellow, and gray represent 0.25, 0.5, and 0.75 quantiles, respectively.

Figure 2.

Figure 2—figure supplement 1. TT-seq identifies transcriptionally active enhancers.

Figure 2—figure supplement 1.

(A) Venn diagram of ncRNAs overlapping with H3K4me1 ChIP-seq or ATAC-seq signal. (B) Width of enhancers identified by TT-seq (Figure 2A and Materials and methods). (C) Location of identified enhancers. (D) Number of enhancers with differentially regulated synthesis levels at various cutoff levels of log2 fold change. (FDR < 0.05).

The resulting 7624 transcriptionally active regions correspond to putative enhancers (referred to as ‘enhancers’ for simplicity). These enhancers had a median length of ~1 kb and were predominantly located in intergenic regions (Figure 2—figure supplement 1B and C). Among these, 3539 eRNAs significantly changed expression during transdifferentiation (|log2(FC)|>1, FDR < 0.05) (Figure 2—figure supplement 1D). We grouped enhancers in six clusters with downregulated, transiently produced, and upregulated eRNAs (Figure 2C). Among the significantly changed eRNAs, 27.1% showed upregulation in the early transition (Cluster 'Inter1', 'Inter2', 'Inter-Late') and 30.7% in the late transition (Cluster 'Late') (Figure 2C). Moreover, 32.6% were downregulated in the early transition (Cluster 'Early') while 9.6% were further downregulated in the late transition (Cluster 'Early-Inter' Figure 2C). In summary, we used a combination of TT-seq, H3K4me1 ChIP-seq, and ATAC-seq to annotate a conservative set of 7624 transcriptionally active enhancers, of which >46% showed significant changes in eRNA synthesis, with 57.8% being initially upregulated and 42.2% being downregulated (Figure 2—figure supplement 1D). As a result, we obtained the dynamics of the enhancer landscape during transdifferentiation.

C/EBPα binding and chromatin opening

In order to investigate how C/EBPα may induce the observed changes in enhancer landscape and gene transcription, we mapped C/EBPα-binding sites genome-wide by ChIP-seq at 0, 12, 24, and 96 hr. The majority of the obtained 14,561 ChIP-seq peaks (FDR < 0.05) fell into intergenic or intronic regions (Figure 3—figure supplement 1A). To investigate the consequences of C/EBPα binding and its impact on enhancers, we studied the changes in H3K4 monomethylation and chromatin accessibility at C/EBPα-binding sites. On 4550 sites with low or undetectable chromatin accessibility at 0 hr, H3K4me1 was present in the absence of C/EBPα binding (Figure 3A and Figure 3—figure supplement 1B). After 12 hr, C/EBPα was bound at these sites, chromatin became more accessible compared to 0 hr, and H3K4 monomethylation was observed in the regions flanking the site. After 24 hr, chromatin showed increased accessibility at these sites relative to previous time points whereas H3K4me1 decreased, likely due to histone depletion (Figure 3A and Figure 3—figure supplement 1B and C).

Figure 3. Enhancer transcription follows chromatin opening.

(A) Average signal of ChIP-seq of C/EBPα (red) and H3K4me1 (purple) or ATAC-seq (blue) on C/EBPα binding sites where chromatin accessibility is limited at 0 hr (n = 4550). Each panel depicts coverage at the indicated time. (B) Coverage plots of genome-wide data sets as indicated. Rows of all panels represent enhancers occupied by C/EBPα, sorted by eRNA synthesis kinetics and intensity. The coverage was aligned to the transcription start site (TSS) of eRNAs. (C) Log2-transformed counts of chromatin accessibility and eRNA synthesis on enhancers within chromatin regions previously inaccessible at 0 hr. (D) Difference between changes of chromatin accessibility and changes in eRNA synthesis at each locus (y-axis) between indicated time points (x-axis). Log2-transformed counts of chromatin accessibility and eRNA synthesis on enhancers within chromatin regions previously inaccessible at 0 hr. (E) Coverage tracks of eRNA synthesis (green) and chromatin accessibility (blue) over indicated time points illustrate that chromatin becomes accessible before enhancer transcription at KDM7A and CD14 enhancers.

Figure 3.

Figure 3—figure supplement 1. C/EBPα binding induces chromatin opening and activate enhancers.

Figure 3—figure supplement 1.

(A) C/EBPα binding location across the genome during transdifferentiation. (B) Average signal of ChIP-seq of C/EBPα (red) and H3K4me1 (purple) or ATAC-seq (blue) on ±5 kb around C/EBPα binding sites where chromatin accessibility is limited at 0 hr (n = 4550). Each panel depicts coverage at the indicated time. (C) Coverage plots of genome-wide data sets as indicated. Rows represent C/EBPα binding sites that appeared at indicated times and were sorted by intensity. The coverage was aligned to the summits of C/EBPα binding sites. (D) Spearman's correlation coefficients between C/EBPα and histone H3K4me1 or chromatin accessibility signal at indicated time points. (E) Exemplary coverage plots for KLF4 and MAFB enhancers show that C/EBPα binding induces chromatin opening. (F) Heatmap of eRNA synthesis changes induced by C/EBPα for six clusters obtained with k-means clustering. (G) Fractions of enhancers where eRNA synthesis was altered upon C/EBPα binding. (H) Spearman's correlation coefficient heatmap of log2 fold changes of eRNA synthesis and C/EBPα binding signal from previous time points.

During the transition from 24 hr to 96 hr binding of C/EBPα occurred at a different set of late activated enhancers. These enhancers showed low levels of H3K4me1 and low chromatin accessibility in the beginning of the transdifferentiation process, both of which however increased strongly over time (Figure 3—figure supplement 1C). C/EBPα occupancy at 24 hr and 96 hr showed the highest correlation with H3K4me1 at 24 hr and with chromatin accessibility at 96 hr, respectively (Figure 3—figure supplement 1D). At these late sites, an increase of H3K4me1 was observed concomitant with C/EBPα binding, but chromatin opening was again delayed as exemplified for the −358 kb KLF4 and the +349 kb MAFB enhancer (Figure 3—figure supplement 1E). Together, these results indicate that C/EBPα initially binds to primed enhancers and subsequently acts as a pioneer factor at de novo enhancers, where it increases H3K4 monomethylation and may induce chromatin opening.

Chromatin opening and eRNA synthesis

Next, we asked whether H3K4 monomethylation or chromatin accessibility is a prerequisite for eRNA synthesis. We clustered 943 significantly upregulated enhancers (log2(FC) >1, FDR < 0.05) occupied by C/EBPα according to changes in eRNA synthesis (Figure 2C), and then sorted them within clusters by eRNA signal (Figure 3B). We found that eRNA synthesis followed C/EBPα binding, H3K4 monomethylation, and chromatin opening, as seen clearly for the late (96 hr) eRNA cluster (Figure 3B). Enhancer transcription always followed chromatin opening. At sites where chromatin was closed at 0 hr and opened after 12 hr, eRNA synthesis was observed only at 24 hr (Figure 3C and D). This was corroborated when comparing changes in chromatin accessibility and eRNA synthesis at specific loci. For example, at the −54.8 kb enhancer of KDM7A, chromatin opening started 12 hr after induction but eRNA synthesis was only detectable after 96 hr. Similarly, the +37.7 kb CD14 enhancer showed nucleosome depletion at 24 hr and eRNA production at 96 hr (Figure 3E).

To further investigate the order of events, we analyzed the temporal relationship between C/EBPα binding and eRNA synthesis for all enhancers bound by C/EBPα (1487 out of 7624 enhancers). Of these, 965 were differentially transcribed during transdifferentiation compared to 0 hr (log2(FC) >1, FDR < 0.05, Figure 3—figure supplement 1F and G), and 76.2% (735 of 965 enhancers) showed increased eRNA synthesis upon C/EBPα binding. Changes in C/EBPα binding from 12 hr to 24 hr correlated with later changes in eRNA synthesis, from 24 hr to 96 hr (Figure 3—figure supplement 1H). At the time of eRNA synthesis, C/EBPα occupancy had decreased relative to the previous time points (Figure 3B), suggesting that C/EBPα is required for chromatin opening but not for transcription of late enhancers. In summary, these analyses provided insights into the temporal order of events at putative enhancers during transdifferentiation. The results are consistent with a pioneering role of C/EBPα, which is able to invade chromatin at de novo enhancers and to induce chromatin opening, which later leads to eRNA synthesis and C/EBPα release at many of these sites.

Multiple enhancers can be paired with target genes

We next aimed at identifying the target promoters for the enhancers. We paired all 7624 enhancers with potential target promoters using two different approaches (Materials and methods, Figure 4A). The ‘neighboring approach’ pairs enhancers with the nearest upstream or downstream promoters that were active at one time point at least. Enhancer–promoter pairs with significant changes in mRNA synthesis were further analyzed (Figure 4B and Figure 4—figure supplement 1A). Changes in eRNA synthesis correlated positively with changes in mRNA synthesis (Figure 4—figure supplement 1C and D), in agreement with what we observed in our previous analysis of T-cell activation (Michel et al., 2017). The ‘1 Mb method’ pairs all enhancers that lie within 1 Mb, the median width of a human topologically associating domain (TAD) (Dixon et al., 2012Yan et al., 2017), of a promoter if the correlation coefficient between synthesis of eRNA and mRNA was greater than 0.4 (Figure 4A). Only pairs with significant changes in both eRNA and mRNA synthesis were further analyzed (Figure 4B and Figure 4—figure supplement 1B).

Figure 4. Multiple enhancers activate target gene promoters.

(A) Schematic diagram describing ‘neighboring’ and ‘1 Mb’ enhancer–promoter pairing methods. With the neighboring method, enhancers are paired to the nearest upstream and downstream transcribed promoters. Pairs with differentially regulated promoters were taken for further analysis. With the 1 Mb method, enhancers were paired with all promoters within 1 Mb, which corresponds to the median width of a human topologically associating domain (TAD) (Dixon et al., 2012Yan et al., 2017). Pairs with promoters and enhancers that are differentially regulated in a correlated manner were analyzed further (Materials and methods). (B) Number of enhancer–promoter pairs, enhancers and genes per pair and method. (C) 59.1% and 65.7% of genes are regulated by more than one enhancer when using the neighboring and 1 Mb method, respectively. (D) Multiple enhancer cooperation at CD14 gene is exemplified by the coverage tracks of indicated data sets at 96 hr. CD14 gene is regulated by six enhancers. BRD4, H3K4me1, and H3K27Ac coverage tracks at 0 hr and 96 hr are depicted in Figure 4—figure supplement 1E. (E) Scatterplot of log2 fold change of mRNA synthesis level at 96 hr compared to 0 hr and log2 fold change of eRNA synthesis from individual (ind.) or sum of all enhancers. Genes regulated by three enhancers (left) or by four enhancers (right) are shown. Spearman's correlation coefficients at the top left corner and adjusted R-squared (R2) values at the bottom right corner indicate that the sum of enhancer activity changes explains changes in mRNA synthesis better.

Figure 4.

Figure 4—figure supplement 1. Enhancer–promoter pairing.

Figure 4—figure supplement 1.

(A and B) Distance between enhancer–promoter paired by ‘neighboring’ (A) or ‘1 Mb’ (B) methods. (C and D) Spearman's correlation heatmap (C) and scatter plot (D) of log2 fold changes of eRNA and mRNA synthesis from previous time points, using the neighboring method. (E) The exemplary coverage tracks of indicated data sets at CD14 gene are depicted. H3K27Ac coverage track of 72 hr is shown since H3K27Ac ChIP-seq at 96 hr is not available. (F) Fraction of EP pairs from indicated method overlapping with TAD obtained from Hi-C data. (G and H) Positions of enhancers paired with CD14 (G) and LMO2 (H) with chromatin contact information (HiC data from THP-1 cells [Phanstiel et al., 2017]) indicate that these enhancers have high contact frequency (orange: genes, green: enhancers, black lines: pairs).
Figure 4—figure supplement 2. Multiple enhancers activate target gene promoters.

Figure 4—figure supplement 2.

(A) mRNA synthesis level (log2 transformed) was greater as number of paired enhancers was increased, both by neighboring and 1 Mb methods. The regression coefficients and p-value of linear regression model (lm) are indicated. (B) Regression coefficients indicating changes in mRNA synthesis over individual eRNA synthesis. Neighboring (left) or 1 Mb (right) methods. (C and D) Spearman's correlation coefficients of mRNA synthesis and individual or the sum of 2, 3, or all of paired enhancers by neighboring (C) and 1 Mb methods (D). (E and F) Adjusted coefficient of determination (R-squared, R2) was calculated from linear regression models (lm) of mRNA synthesis against individual enhancers or against the sum of the indicated number of paired enhancers by neighboring (E) and 1 Mb methods (F).

We found that 59.1% or 65.7% of putative target genes (1057 and 1446 genes, respectively) were paired with more than one enhancer when we used the neighboring or 1 Mb pairs, respectively, as exemplified for CD14 (Figure 4C and D, Figure 4—figure supplement 1E). Obtaining similar results in parallel analyses with two different pairing methods strengthens our conclusions. To gain additional support for the promoter–enhancer pairing, we tested whether the pairs were located within the same TAD obtained from Hi-C data (Stik et al., 2020Phanstiel et al., 2017). This showed that 98.3% of the promoter–enhancer pairs obtained by the neighboring method were found within the same TAD, and 96.2% of the pairs obtained with 1 Mb method were found within the same TAD at one or more time points (Figure 4—figure supplement 1F), as exemplified by CD14 and LMO2, (Figure 4—figure supplement 1G and H). In summary, we identified putative target genes for the majority of enhancers and found that many genes can be paired with multiple enhancers, in line with previous reports (Beagrie et al., 2017; Rubin et al., 2017).

Multiple enhancers cooperate to activate target genes

To investigate the temporal changes in transcription activity of target genes that were paired with multiple enhancers, we used the amount of eRNA production as a proxy for enhancer activity. We first examined whether target gene transcription changes relative to 0 hr can be explained by the activity changes of a single enhancer or whether they can only be explained when multiple enhancers were taken into account. Generally, target genes paired with more than one enhancer showed higher mRNA synthesis (Figure 4—figure supplement 2A), consistent with published observations (Rubin et al., 2017). We sorted target genes by the number of their paired enhancers, and calculated regression coefficients of log2(FC) of eRNA synthesis (explanatory variable) against log2(FC) of mRNA synthesis (target variable) (Figure 4—figure supplement 2B). As the number of paired enhancers increased, the regression coefficient increased, showing that stronger changes in mRNA synthesis were detected when more paired, active enhancers were present (Figure 4—figure supplement 2B).

Next, we tested whether the changes in promoter activity can be explained by the cumulative changes of all paired enhancers. We calculated the correlation between the log2(FC)s in mRNA synthesis and the sum of eRNA synthesis of the paired enhancers. There was a clear trend that promoter activity changes are better explained by the sum of activity changes in paired enhancers (Figure 4E). Generally, the correlation between promoter and enhancer activity changes improved with the number of paired enhancers (Figure 4—figure supplement 2C and D), as did the coefficient of determination (R2) of a linear regression, and this was independent of the method of promoter–enhancer pairing (Figure 4—figure supplement 2E and F). These results indicate that multiple paired and active enhancers can contribute to target promoter activation and should be taken into account to understand changes in target gene transcription.

Enhancer cooperation can be additive or synergistic

We next asked whether the contributions of individual enhancers to target gene activation are always additive or whether they can also be non-additive. This requires an analysis at the level of individual loci, which is only possible with time course data as available here. For robust curve fitting, we included TT-seq data for three additional time points, resulting in a total of seven time points (0, 12, 24, 30, 36, 72, and 96 hr). We fitted these data to three models (Dukler et al., 2017), an additive, a synergistic (or exponential), and a logistic model (Figure 5A,B and Figure 5—figure supplement 1A, respectively; Materials and methods). The additive model explains promoter activity with the sum of the activities of all enhancers paired with the neighboring method. The synergistic model assumes that changes in promoter activity over time are greater than what is predicted from the changes in the sum of the enhancer activities. The logistic model postulates that promoter activity reaches a limit and cannot increase any further even if the sum of the enhancer activities increases more, reflecting the known upper limit of promoter activity (Gressel et al., 2017; Ikeda et al., 1992). To determine the best-fitting model, we computed the Bayesian information criterion (BIC) score for each model and plotted the relative BIC, that is, the difference between the BIC score for the additive model and the BIC score of the tested models (Figure 5—figure supplement 1B and C).

Figure 5. Enhancer cooperation can be additive or synergistic.

(A and B) Exemplary curve fits of the additive (A) and the synergistic models (B) to the observed size factor and length normalized TT-seq RNA synthesis level for the indicated mRNA and the sum of the paired eRNAs. (C) The relative Bayesian Information Criterion (BIC) for the curve fits for a total of 523 promoters reveals that RNA synthesis at most loci follows either the additive or the synergistic model. The synergistic model fitted better for 17% of tested loci, showing a relative BIC greater than 2 (red). Any loci with a relative BIC between 0 and 2 were left as ambiguous because they fitted both models (gray). (D and E) Heatmap of gene expression (RNA-seq) regulated in additive (n = 348) (D) or synergistic (n = 136) (E) manner. Clusters are as shown in Figure 1C (k-means). (F) ChIP signal for BRD4, C/EBPα, and H3K27Ac, and ATAC-seq signal, ranked by signal strength on the peaks within 1 kb of our enhancers (Materials and methods, black dots) with cut-off values for superenhancers (red dashed lines). Synergistic enhancers are depicted with gray dots. (G) Additive (blue) and synergistic (red) enhancers are compared for their occupancy with BRD4, C/EBPα, H3K27Ac, or their chromatin accessibility (ATAC-seq). The comparison was carried out for the time point where the highest signal was observed. Fold differences in median signal of BRD4, C/EBPα, H3K27Ac, and chromatin accessibility between additive and synergistic enhancers are 1.02, 0.96, 1.07, and 1.09, respectively.

Figure 5.

Figure 5—figure supplement 1. Enhancer cooperation can be additive or synergistic.

Figure 5—figure supplement 1.

(A) The observed normalized TT-seq RNA synthesis level for the indicated mRNA and the sum of the paired eRNAs were plotted for PIM1 and PDGFA, those best-fitted with a logistic model as shown with curve fits (gray line). (B and C) The relative Bayesian Information Criterion (BIC) reveals that RNA synthesis follows either the additive (B) or the synergistic model (C). A higher relative BIC indicates better fitting of the model. (D) Using the neighboring pairing method, each of the tested genes was assigned to three theoretical models according to the BIC with seven data points. Numbers of the genes assigned with the indicated models are depicted. (E) Using the neighboring pairing method, BIC of the genes that were best fitted with logistic model were recalculated with six data points, without the data point with the highest sum of enhancer activities. Number of the genes assigned with the indicated models according to BIC is depicted.
Figure 5—figure supplement 2. Enhancer synergy is a robust phenomenon.

Figure 5—figure supplement 2.

(A) The relative Bayesian Information Criterion (BIC) of logarithmic model against additive model. The logarithmic model fitted better only for a few tested loci, showing a relative BIC greater than 2 (green) showing distinct curve shape from Figure 5C. (B) Using the 1 Mb pairing, each of the tested genes was assigned to three theoretical models according to the BIC with seven data points (upper). BIC of the genes that were best fitted with logistic model were recalculated with six data points, without the data point with the highest sum of enhancer activities (lower). Numbers of the genes assigned with the indicated models are depicted. (C) Overlap of additive and synergistic enhancers obtained from neighboring and 1 Mb methods. (D) 68.9% of genes are regulated by more than one enhancer when using the TAD method. (E) Overlap of additive and synergistic enhancers obtained from neighboring and TAD methods. (F) Overlap of additive and synergistic enhancers obtained using 7624 EUs defined by H3K4me1 ChIP-seq and ATAC-seq signals (indicated as EU) and 15487 extended EUs defined by TT-seq signals regardless of epigenetic state (indicated as extEU) (Materials and methods). (G) Number of synergistic enhancers obtained using simulatively increased eRNA transcription level. (H) Fraction of additive (blue) or synergistic (red) genes paired with different numbers of enhancers. (I) Fraction of the maximum (highest) eRNA synthesis to the total eRNA signal for all paired enhancers of each gene. (J) Number of enhancers regulating two genes in the indicated regulatory manner. Nearest pairing method was used. p-values from binomial test are indicated.
Figure 5—figure supplement 3. Synergistic enhancers drive cell type-specific genes.

Figure 5—figure supplement 3.

(A) mRNA synthesis of the upregulated genes at 96 hr in additive (blue) or synergistic (red) manner. (B) GO analysis of mRNAs regulated in additive (blue) or synergistic (red) manner. (C) Heatmap of chromatin accessibility on additive (left) or synergistic (right) enhancers upregulating the genes at 96 hr. (D) Chromatin accessibility at promoters at 0 hr was not significantly different between additive and synergistic genes in late cluster.
Figure 5—figure supplement 4. Synergistic enhancers drive cell type-specific genes.

Figure 5—figure supplement 4.

(A) ChIP signal for BRD4, C/EBPα, and H3K27Ac, and ATAC-seq signal, ranked by signal strength on the peaks within 1 kb of our enhancers (Materials and methods, black dots) with cutoff value for superenhancers (red dashed line). Additive enhancers are shown with gray dots. (B) Coverage tracks of exemplary superenhancers, EBF1 and IRF8 at 0 hr and 96 hr. H3K27Ac at 72 hr instead of 96 hr is depicted. (C) Heatmap of RNA synthesis on target genes (upper panel) or eRNAs (lower panel) of superenhancers determined by ChIP occupancy with BRD4 or C/EBPα, H3K27Ac, and chromatin accessibility. (D) Fraction of our additive (blue) or synergistic (red) enhancers overlapping with superenhancers from dbSUPER (Hnisz et al., 2013; Khan and Zhang, 2016). All of our additive or synergistic enhancers (136 enhancers) were compared with superenhancers from all available cells in dbSUPER (first column), from CD14+ cells, which are macrophage cells (second column), and from CD19+ cells, which are B cells (third column). Fisher’s exact test p-value is shown.

To investigate which of the three models best explains transcription changes at target genes that are paired with multiple enhancers we estimated the relative BIC for 773 target genes that were paired with 2–20 enhancers according to the neighboring method. Of these target genes, 277, 92, and 250 were best described by the additive, synergistic, and logistic models, respectively (Figure 5—figure supplement 1D, Materials and methods). We ignored 154 genes with a relative BIC between 0 and 2, because they were ambiguous, showing reasonable fits to both additive and synergistic models (Figure 5C). For the 250 genes fitted with the logistic model, we observed that promoter activity depended on enhancer activity either in an additive or in a synergistic way before reaching a plateau (Figure 5—figure supplement 1A). To distinguish between these models, we excluded the data point with the highest sum of enhancer activities and calculated the BIC with the remaining six data points. Of the 250 genes, 71 and 44 genes could now be fitted with the additive or synergistic model, whereas 109 genes remained in the logistic class (Figure 5—figure supplement 1E).

Taken together, our analysis indicates how enhancers cooperate to drive gene transcription over time. A total of 348 genes (~45%) were regulated in an additive manner by multiple enhancers. These genes included Myc, which is known to be regulated in an additive manner (Bahr et al., 2018). We also found a total of 136 genes (92 + 44, ~17%) that were regulated by multiple enhancers in a synergistic manner. These genes showed an exponential change in mRNA synthesis over time relative to the sum of eRNA synthesis from paired enhancers (Figure 5B), and included macrophage-related genes such as CITED2, LYZ (Keshav et al., 1991; Kranc et al., 2009), and B cell-related genes such as BCL7A, LEF1, and TLE1 (Haddad et al., 2004). In summary, enhancers cooperate in an additive or a synergistic manner to regulate their target genes during the transdifferentiation process.

Enhancer synergy is a robust phenomenon

To investigate the phenomenon of synergistic enhancer activity further, we carried out additional analyses. To exclude that the observed synergistic behavior is a consequence of a noisy relative BIC distribution, we computed the relative BIC of an additive model with respect to a logarithmic model, which is not meaningful (Figure 5—figure supplement 2A). Strikingly, data from only three genes could be fitted with the logarithmic model, providing a negative control for our curve fitting. Next, we checked whether we missed distant enhancers that contribute to target gene activation. We carried out the analysis with enhancer–promoter pairs obtained with the 1 Mb pairing method. We observed 603 and 194 genes regulated by additive and synergistic enhancers, respectively (Figure 5—figure supplement 2B). The synergistically regulated genes found with the two different pairing methods strongly overlapped. Of the 57 synergistically regulated genes that were found with the neighboring method and could be tested, 33 were classified as synergistically regulated with the 1 Mb pairing method (Fisher’s exact test p-value = 3.5e-11) (Figure 5—figure supplement 2C).

Since the number of enhancers paired to each gene varies between the neighboring and 1 Mb methods, only a small number of genes could be tested for enhancer synergy. We therefore used the third method of promoter–enhancer pairing referred to as ‘TAD pairing’. TAD pairing pairs all promoters and enhancers that lie within a TAD that was experimentally identified with Hi-C data (Stik et al., 2020). We assumed pairing if the correlation coefficient between changes in synthesis of eRNA and mRNA was greater than 0.4. Pairs with significant changes in both eRNA and mRNA synthesis were further analyzed. We found 7355 pairs with 1882 promoters and 2567 enhancers. Similar to other pairing methods (Figure 4C), 69% of putative target genes (1296 genes) were paired with more than one enhancer (Figure 5—figure supplement 2D). We observed 561 and 238 genes regulated by additive or synergistic enhancers, respectively. Out of 68 synergistically regulated genes that were found with the neighboring method and could be tested, 52 were also classified as synergistically regulated with TAD pairing (Fisher’s exact test p-value = 2.9e-19) (Figure 5—figure supplement 2E). Thus, the synergistically regulated genes found with the different pairing approaches overlapped significantly, providing further support of our findings.

We further tested whether we had underestimated the number of enhancers because this could lead to incorrect assignment of synergistic behavior. We extended our list of putative enhancers by including 7863 additional merged ncRNA TUs that lacked H3K4me1 or ATAC-seq signals. Using the neighboring method, this increased the number of paired enhancers to 1184 out of 1790 TUs. We found that 97 of the 136 previously identified synergistically regulated genes (71%) were again detected to be regulated in a synergistic manner (Figure 5—figure supplement 2F). This indicates that the widespread synergistic behavior we observed was not due to an underestimation of the number of enhancers. In order to check whether the observed enhancer synergy was due to the underestimation of enhancer activities, we simulated an increase of eRNA transcription and determined synergistically regulated genes. Synergistically regulated genes were not affected by the artificial increase of eRNA transcription (Figure 5—figure supplement 2G). We further found no difference in the number of enhancers between additive and synergistic loci (median 3.5 and 3, respectively) (Figure 5—figure supplement 2H).

We also asked whether one highly active enhancer could dominate gene activation for synergistic enhancers. This was not the case because the median contribution from the strongest enhancer to the total eRNA signal was the same for additive and synergistic enhancers (Figure 5—figure supplement 2I). As an additional control, we asked whether synergistic enhancers could regulate two target genes, and whether in this case both target genes are regulated in the same manner. We found that 694 additive or synergistic enhancers were paired with two target gene promoters using the neighboring method. Of these, 463 enhancers were paired with target genes that were either both regulated in an additive manner (463 enhancers, 254 promoters) or both in a synergistic manner (139 enhancers, 84 promoters) (Figure 5—figure supplement 2J, binomial p.val <2.2e-16). This is consistent with coordinated transcriptional bursting of two promoters regulated by one enhancer (Fukaya et al., 2016). These controls and analyses confirm the observed enhancer synergy and indicate that co-operative action is a property of enhancers, not promoters.

Synergistic enhancers are involved in cell type-specific gene expression

We next investigated the nature of the target genes of synergistic enhancers. We found that most target genes of synergistic enhancers (85%) were cell type-specific genes, that is, genes that were specifically expressed either in B cells (60 TUs) or in macrophage-like cells (56 TUs) according to RNA-seq data (Figure 5D and E). B cell-specific genes controlled by synergistic enhancers include MYB, BLNK, VPREB1, and IGLL5 (Cobaleda et al., 2007; Thomas et al., 2005). Macrophage-specific genes regulated by synergistic enhancers include CEBPB, VSIG4, and ITGAX (Friedman, 2007; Lavin et al., 2014; Vogt et al., 2006). Macrophage-specific, synergistically regulated genes often remained inactive until 36 hr, while additive genes showed activation from 12 hr onwards (Figure 5—figure supplement 3A). Gene ontology analysis of the genes regulated by additive and synergistic enhancers confirmed the largely cell type-specific nature of the latter ones (Figure 5—figure supplement 3B). Thus, the majority of synergistic enhancers drive cell type-specific genes.

Further analysis showed that the delayed but rapid induction of synergistically regulated genes was not a consequence of late chromatin opening. Although the average chromatin accessibility changed more at synergistic enhancers compared to additive enhancers (45% vs 38% of overlapping ATAC-seq peaks, Fisher’s exact test p-value = 0.01), the kinetics of chromatin opening were similar for both types of enhancers at late upregulated genes (Figure 5—figure supplement 3C). The initial chromatin accessibility at 0 hr was also comparable for genes regulated by additive or synergistic enhancers (Figure 5—figure supplement 3D). Thus, there is no evidence that enhancer synergy at target genes that are strongly induced at later time points is simply a consequence of delayed chromatin opening at these enhancers. Instead, these observations point to another mechanistic basis for enhancer synergy.

We further tested whether our synergistic enhancers may be described as ‘superenhancers’, because these can also target cell type-specific genes. Superenhancers are generally defined by their high occupancy with specific TFs and coactivators (Hnisz et al., 2013). We collected ChIP-seq data for the Mediator subunit MED1, the chromatin regulator BRD4 (Materials and methods), and used these and recent ChIP-seq data for H3K27Ac (Stik et al., 2020) for superenhancer calling. The MED1 signal was weak and did not allow for a robust analysis. However, the BRD4 and H3K27Ac signal could be used to generate a typical ‘superenhancer plot’ (Hnisz et al., 2013; Sabari et al., 2018; Whyte et al., 2013; Figure 5F). We also generated such plots for C/EBPα ChIP-seq data and for the ATAC-seq data. All four plots identified enhancers with high signals (Figure 5F, Figure 5—figure supplement 4A and B). Although the superenhancers derived this way show high occupancy or accessibility at 96 hr, RNA synthesis from these superenhancers and their paired promoters was not always the highest at 96 hr compared to other time points (Figure 5—figure supplement 4C). Consistent with this, the genes paired with superenhancers included not only cell type-specific genes such as CD14, DDIT4, JUNB, and FOSL2, but also genes that are commonly expressed across different cell types, such as DNMT1, ARID5B, and MAP3K1. Gene ontology analysis on the paired genes of superenhancers includes cell type-specific GO terms such as immune response and inflammatory response as well as general GO terms such as regulation of cellular process and cell–cell adhesion (Supplementary files 25).

Furthermore, the superenhancers detected by occupancy or accessibility signals did not correspond to the synergistic enhancers we identified here based on RNA synthesis changes. Indeed, previously identified superenhancers from the dbSUPER database (Hnisz et al., 2013; Khan and Zhang, 2016) overlapped to similar extents with our additive or synergistic enhancers (Figure 5—figure supplement 4D). Additionally, there was no difference in the occupancy signals between synergistic enhancers and additive enhancers (Figure 5G). In conclusion, the synergistic enhancers we identified based on functional data (RNA synthesis) are involved in cell type-specific gene expression and do not generally correspond to classically defined superenhancers.

Synergistic enhancers are regulated by cell type-specific TFs

We finally investigated whether enhancer synergy depends on binding of specific TFs to enhancer regions. According to our RNA-seq data, different sets of TFs are expressed in cascades from 0 hr to 96 hr during transdifferentiation (Figure 6—figure supplement 1A). We located sites in our enhancers that contained DNA motifs for these TFs. Then we determined which of these sites are occupied by TFs, using our ATAC-seq data to perform TF footprinting on the enhancers regulating late upregulated genes (Sherwood et al., 2014). This analysis revealed that motifs of 23 TF subfamilies are significantly enriched in our synergistic enhancers compared to the additive enhancers (Fisher's exact test p-value < 0.05, Figure 6—figure supplement 1B). These enriched TFs include C/EBP, PU.1, CREB1, ETS, KLF, and RUNT family factors (Figure 6A). TFs in these categories are known to function in macrophages (Friedman, 2007). In particular, C/EBPα/β/ε are required for granulocyte–monocyte progenitors (GMP) and macrophages (Friedman, 2007). Knockdown of PU.1 and C/EBPβ was shown to impair transdifferentiation (van Oevelen et al., 2015), and C/EBPβ dimerizes with CREB1 to activate macrophage genes (Ruffell et al., 2009). Members of the KLF family, such as KLF1-4 and KLF6, are important for monocyte and macrophage activation (Cao et al., 2010; Date et al., 2014). C/EBPα, PU.1, and RUNX1 are frequently mutated in acute myeloid leukemia (AML), which can arise from reduced transcription activity and impede myeloid differentiation (Rosenbauer and Tenen, 2007).

Figure 6. Cell type-specific transcription factors are enriched in synergistic enhancers.

(A) Motifs of TFs enriched in synergistic enhancers from TF footprinting analysis (Materials and methods). (B) Expression of differentially expressed TFs (| log2(FC) |>1, FDR < 0.05, RNA-seq) enriched in synergistic (left panel), additive (middle panel), and both synergistic and additive (right panel) enhancers.

Figure 6.

Figure 6—figure supplement 1. Specific sets of TFs are enriched on synergistic enhancers.

Figure 6—figure supplement 1.

(A) Heatmap of differentially expressed TFs from RNA-seq (| log2(FC) |>1, FDR < 0.05) for seven clusters obtained with k-mean clustering (Figure 1C). (B) Bound TF motif occurrences were calculated from TF footprinting with ATAC-seq at each indicated time point. Each dot represents a fraction of synergistic enhancers with a TF motif to a fraction of additive enhancers with a TF motif. TF motifs that are bound significantly more within the synergistic or additive enhancers are indicated in red or blue, respectively (Fisher’s exact test p-value < 0.05). (C) Motifs of TFs enriched in additive enhancers upregulating genes at 96 hr from TF footprinting analysis (Materials and methods).

In contrast, additive enhancers contained motifs of only two TF families that were enriched compared to synergistic enhancers, in particular the E2F and FOS (Figure 6—figure supplement 1C). TFs of the E2F family have general regulatory functions in macrophages, such as cell cycle and apoptosis regulation (Trikha et al., 2011). TFs of the FOS family also have regulatory functions in macrophages (Friedman, 2007). Binding to FOS motifs was enriched in additive enhancers only at 12 hr (Figure 6—figure supplement 1B), indicating that at later time points FOS family factors bind to additive and synergistic enhancers similarly.

In general, synergistic enhancers were enriched for macrophage-specific TFs, whereas additive enhancers were not. To confirm this, we investigated the expression of TFs that belong to the same subfamilies with the enriched TFs in additive or synergistic enhancers. Indeed, of the TFs that were enriched in synergistic enhancers and differentially expressed, about half were macrophage specific, whereas TFs of E2F family that were enriched in additive enhancers were not (Figure 6B). This suggests that additional TFs are needed to activate macrophage-specific additive genes during the late transition (96 hr) (Figure 6B). This is possibly due to binding of TFs at both additive and synergistic enhancers such as those for the IRF and MAF family factors that are also known to regulate macrophage genes (Friedman, 2007). Many of these shared TFs are indeed upregulated at 96 hr (Figure 6B). These results indicate that a set of shared TFs bind to both types of enhancers, whereas a specific set of macrophage-specific TFs additionally and preferentially binds to synergistic enhancers to drive macrophage-specific target gene promoters that establish the new cell type.

Discussion

How enhancers cooperate to drive target gene expression and to determine cell types is a central question in the field of genomic regulation. However, the question of functional enhancer cooperativity could thus far not be addressed in a systematic manner in a dynamic system. A route to such an analysis has now been provided by the availability of an efficient and simple transdifferentiation system and the development of TT-seq to measure genome-wide RNA synthesis from both enhancers and target gene promoters. Here we investigated enhancer cooperation over time during transdifferentiation of human B-cells to macrophage-like cells. We examined this process in a quantitative manner, assuming eRNA synthesis to be a proxy for enhancer activity. We identified transcriptionally active enhancers, their temporal activity changes, and used correlations to predict how multiple enhancers cooperate to regulate their target genes. These analyses indicated that multiple enhancers often act in an additive manner, and that synergistic enhancer action occurs at ~20% of the tested putative target genes. Target genes of synergistic enhancers were almost exclusively cell type-specific.

The synergistic enhancers identified here were different from previously reported superenhancers. We identified synergistic enhancers from changes in RNA synthesis activity during the time course, whereas identification of superenhancers generally uses TF and coactivator occupancy measurements at one time point. Whereas our synergistic enhancers are not necessarily arranged next to each other, superenhancers are generally linear clusters of constituent enhancers. In contrast to our time-resolved study, previous studies investigated enhancer cooperation in steady state or at an end point, and at selected genes, rather than genome-wide (Hay et al., 2016; Osterwalder et al., 2018; Stine et al., 2011). It was suggested that superenhancers retain stable interactions via nuclear condensation that involves phase separation (Hnisz et al., 2017).

Although the molecular mechanisms underlying enhancer cooperation remain to be investigated, simple models may explain our observations (Figure 7). An additive effect of enhancers on gene expression is expected if the enhancers act independently of each other and do not need each other for productive transcription. This additive mode of regulation is predicted to allow for fine-tuning of promoter activity and to enable robust expression. It would better tolerate the loss of one of the enhancer activities, for instance due to mutations in TF-binding sites or due to changes in the expression levels of TFs that bind the enhancers. Indeed, expression of gap genes in Drosophila are regulated by multiple enhancers, and deleting one of these enhancers has little effect on the gene expression pattern (Perry et al., 2011). Thus, additive enhancers would be advantageous for genes that are expressed and function in multiple cell types, and indeed we observe that additive enhancers target genes commonly expressed throughout multiple transitions during transdifferentiation.

Figure 7. Two models of how multiple enhancers regulate target genes.

Figure 7.

Simple models may explain additive and synergistic enhancer cooperation during target gene activation. Additive cooperation may result from alternative activation of the target promoter by each constituent enhancer at different time points. Synergistic cooperation may be achieved if multiple enhancers contact the target gene promoter simultaneously. For details compare text.

A synergistic effect of enhancers on target gene transcription is expected when the enhancers function together (Figure 7). Such enhancer cooperation could be simultaneous and may rely on cooperative nuclear condensation based on liquid–liquid phase separation (Boehning et al., 2018; Boija et al., 2018; Chong et al., 2018; Hnisz et al., 2017; Murray et al., 2017; Nair et al., 2019; Sabari et al., 2018). Indeed, some of the TFs with enriched motifs in our synergistic enhancers have been shown to undergo phase separation with Mediator (MYC and TP53) or to interact with Mediator (C/EBPβ) (Boija et al., 2018; Li et al., 2008). Nuclear condensation may also allow for a higher frequency of transcriptional bursting (Larsson et al., 2019). Additionally or alternatively, enhancer synergy may also stem from sequential cooperation if different enhancers target different steps of transcription such as initiation and elongation (Beagrie and Pombo, 2016; Haberle and Stark, 2018; Henriques et al., 2018; Kim et al., 1998; Sawado et al., 2003; Spicuglia et al., 2002). The effect of synergistic enhancer cooperation on target gene activity is expected to be switch-like and would be beneficial for cell type-specific genes. Such switches would allow for efficient determination of cell types during differentiation, because genes that are active in the old cell type could be rapidly switched off and genes expressed in the new cell type could suddenly be switched on at a certain stage of differentiation.

Finally, our results improve our understanding of step-wise enhancer and gene activation during a transdifferentiation process. Previous work showed that a subset of TFs is critical for enhancer priming and chromatin remodeling (Heinz et al., 2010; Takaku et al., 2016). C/EBPα can prime enhancers and initiate chromatin opening by interactions with MLL3/4 complexes (Lee et al., 2013) and with SWI/SNF chromatin remodeling complexes (Pedersen et al., 2001). Consistent with this, we find that chromatin opening follows C/EBPα binding and that this results in subsequent enhancer transcription. These observations are consistent with the capacity of the pioneering factor C/EBPα to induce DNA opening (Iberg-Badeaux et al., 2017; van Oevelen et al., 2015), and its ability to reorganize chromatin states and genome architecture before gene expression changes (Stadhouders et al., 2018). It remains controversial whether enhancer transcription is required for enhancer priming (Calo and Wysocka, 2013; Dorighi et al., 2017; Kaikkonen et al., 2013), but our data support that in our system the chromatin modification H3K4me1 generally precedes nucleosome depletion and enhancer transcription (Bonn et al., 2012; Creyghton et al., 2010; Rada-Iglesias et al., 2011; Wamstad et al., 2012). Our results also explain how activation of the glucocorticoid receptor can induce chromatin decompaction in the presence of the transcription inhibitor α-amanitin (Jubb et al., 2017).

In conclusion, our data reveal the stepwise activation of enhancers and genes during transdifferentiation following binding of the pioneering factor C/EBPα. We provide evidence from time course analysis that enhancers can cooperate in an additive or synergistic manner to alter the activity of target genes. Synergistic enhancers tend to bind cell type-specific TFs and regulate cell type-specific genes, and we speculate that this enables a switch-like expression behavior for such genes. The synergistic enhancers identified here are generally distinct from previously reported superenhancers. Finally, our approach can be used to investigate enhancer cooperation in many other cellular processes.

Materials and methods

Key resources table.

Reagent type (species) or resource Designation Source or reference Identifiers Additional information
Cell line (Homo sapiens; female) BLaER1
B-cell precursor leukemia
Laboratory of Thomas Graf RRID:CVCL_VQ57 RCH-ACV stably expressing estrogen inducible C/EBPα
Commercial assay, kit Plasmo Test Mycoplasma Detection Kit InvivoGen rep-pt1
Commercial assay, kit NUGEN Ovation V2 Kit NUGEN 0343
Commercial assay, kit NEB Ultra DNA Library kit NEB E7370S
Commercial assay, kit KAPA Real-Time Library Amplification Kit Peqlab KK2701
Commercial assay, kit Nextera Tn5 Transposase Illumina FC-121–1030
Chemical compound, drug Human CSF-1 PEPROTECH 300–25
Chemical compound, drug Human IL-3 PEPROTECH 200–03
Chemical compound, drug β-estradiol CALBIOCHEM 3301
Chemical compound, drug 4-thiouridine Carbosynth 13957-31-8
Antibody Anti-C/EBPα (rabbit polyclonal) Santa Cruz Cat# sc-61, RRID:AB_631233 ChIP-seq (5 μg for 50 μg of chromatin)
Antibody Anti-H3K4me1 (rabbit polyclonal) Abcam Cat# ab8895, RRID:AB_306847 ChIP-seq (5 μg for 30 μg of chromatin)
Antibody Anti-BRD4 (rabbit polyclonal) Bethyl Laboratories Cat# A301-985A100, RRID:AB_2620184 ChIP-seq (5 μg for 100 μg of chromatin)
Antibody Anti-human CD19-APC-cy7APC-cy7 mouse anti-human CD19(mouse monoclonal) BD Biosciences Cat# 557791, RRID:AB_396873 FACS (2.5 μL per test)
Antibody Anti-human CD14-PEPE mouse anti-human CD14 (mouse monoclonal) BD Biosciences Cat# 555398, RRID:AB_395799 FACS (5 μL per test)
Sequenced-based reagent CD19 forward Rapino et al., 2013 qPCR primers GATGCAGACTCTTATGAGAAC
Sequenced-based reagent CD19 reverse Rapino et al., 2013 qPCR primers TCAGATTTCAGAGTCAGGTG
Sequenced-based reagent IGJ forward Rapino et al., 2013 qPCR primers TGTTCATGTGAAAGCCCAAG
Sequenced-based reagent IGJ reverse Rapino et al., 2013 qPCR primers TCGGATGTTTCTCTCCACAA
Sequenced-based reagent VPREB3 forward Rapino et al., 2013 qPCR primers GGGGACCTTCCTGTCAGTTT
Sequenced-based reagent VPREB3 reverse Rapino et al., 2013 qPCR primers ACCGTAGTCCCTGATGGTGA
Sequenced-based reagent CD14 forward Rapino et al., 2013 qPCR primers GATTACATAAACTGTCAGAGGC
Sequenced-based reagent CD14 reverse Rapino et al., 2013 qPCR primers TCCATGGTCGATAAGTCTTC
Sequenced-based reagent FCGR1B forward Rapino et al., 2013 qPCR primers CCTTGAGGTGTCATGCGTG
Sequenced-based reagent FCGR1B reverse Rapino et al., 2013 qPCR primers AAGGCTTTGCCATTTCGATAGT
Sequenced-based reagent ITGAM forward Rapino et al., 2013 qPCR primers GGGGTCTCCACTAAATATCTC
Sequenced-based reagent ITGAM reverse Rapino et al., 2013 qPCR primers CTGACCTGATATTGATGCTG
Sequenced-based reagent GAPDH forward This paper qPCR primers TCTCTGCTCCTCCTGTTCGAC
Sequenced-based reagent GAPDH reverse This paper qPCR primers GGCGCCCAATACGACCAAAT
Software, algorithm Cutadapt Martin, 2012 RRID:SCR_011841 Version 1.9.1
Software, algorithm Trim Galore https://www.bioinformatics.babraham.ac.uk/projects/trim_galore/ RRID:SCR_011847 Version 0.4.1
Software, algorithm FastQC http://www.bioinformatics.bbsrc.ac.uk/projects/fastqc RRID:SCR_014583
Software, algorithm SAMTOOLS Li et al., 2009 RRID:SCR_002105 Version 1.2
Software, algorithm STAR Dobin et al., 2013 RRID:SCR_004463 Version 2.4.2
Software, algorithm DESeq2 Love et al., 2014 RRID:SCR_015687
Software, algorithm Bowtie 2 Langmead and Salzberg, 2012 RRID:SCR_016368 Version 2.2.5
Software, algorithm MACS Zhang et al., 2008 RRID:SCR_013291 Version 2.1.1
Software, algorithm Bowtie Langmead et al., 2009 RRID:SCR_005476 Version 1.0.0
Software, algorithm HTSeq Anders et al., 2015 RRID:SCR_005514 Version 0.6.1p1
Software, algorithm ggplot2 http://ggplot2.org RRID:SCR_014601
Software, algorithm Pheatmap https://CRAN.R-project.org/package=pheatmap RRID:SCR_016418
Software, algorithm DAVID Huang et al., 2009 RRID:SCR_001881
Software, algorithm GenoSTAN Zacher et al., 2017
Software, algorithm MEME Suite - motif-based sequence analysis tools Bailey et al., 2009 RRID:SCR_001783 Version 4.11.2
Software, algorithm 3D Genome http://promoter.bx.psu.edu/hi-c/ RRID:SCR_017525
Software, algorithm DEoptim Dukler et al., 2017
Software, algorithm ROSE Lovén et al., 2013; Whyte et al., 2013 RRID:SCR_017390
Software, algorithm PIQ Sherwood et al., 2014
Software, algorithm TFBSTool Tan and Lenhard, 2016

Cell line

We used BLaER1 cells derived from RCH-ACV, precursor leukemia B cells, stably expressing estrogen-inducible C/EBPα (Rapino et al., 2013); these cells contain the Squirrel Monkey Retrovirus (SMRV). Cells were cultured in RPMI 1640 (GIBCO, 31870–025) with 10% FBS (GIBCO, 10500–064), 2% Glutamax (GIBCO, 35050–038), 2% Penicillin/Streptomycin (GIBCO, 15140122), and HEPES (GIBCO, 15630–056) at 37°C under 5% CO2. Cells were regularly checked and tested negative for mycoplasma infection using Plasmo Test Mycoplasma Detection Kit (InvivoGen, rep-pt1). For our time course experiments, cells were treated with beta estradiol (100 nM) (CALBIOCHEM, 3301), human IL-3 (10 ng/ml) and human CSF-1 (10 ng/ml) (PEPROTECH, 200–03 and 300–25). Cells were harvested at various time points after treatment (0, 12, 24, 30, 36, 48, 72, 96, and 168 hr).

RNA-seq and TT-seq

RNA-seq and TT-seq were performed from two biological replicates at nine and seven time points, respectively. At each time point, 1–2.2 × 108 cells were labeled with 0.5 mM 4-thiouridine (4sU, Carbosynth, 13957-31-8) for 5 min and harvested by centrifuging at 500 g for 2 min. Harvested pellets were lysed with QIAzol (Qiagen, 79306). We generated RNA spike-ins based on sequences of six ERCC RNA spike-in mix – ERCC-00043, ERCC-00170, ERCC-00136, ERCC-00145, ERCC-00092, and ERCC-00002, and prepared spike-in mix as described previously (Huang et al., 2009). Following addition of 5 ng of spike-in mix per 1 × 108 cells, total RNA was purified using QIAzol (Qiagen, 79306) according to the manufacturer’s instructions. Total RNA (300 μg) was fragmented to 1500–5000 bp in size using Covaris S220 Ultrasonicator. Fraction of total RNA was set aside for total-RNA-seq. Newly synthesized RNAs were purified as described (Schwalb et al., 2016) with the following modifications. After streptavidin pull down, the resulting RNA was purified with RNeasy Micro Kit (Qiagen, 74004), together with DNase treatment (Qiagen, 79254). Libraries were generated with Nugen Ovation Universal RNA-seq System (Nugen, 0343), and amplified by nine cycles in addition to two initial cycles according to the manufacturer's instructions. Libraries were sequenced at 2 × 50 bp on HiSeq2000 (TAL, Göttingen) or at 2 × 75 bp on an Illumina NextSeq machine in house. The average read numbers obtained for total RNA-seq were 1 × 107 and 2 × 108 for TT-seq.

Chromatin immunoprecipitation sequencing (ChIP-seq)

ChIP-seq was performed from two biological replicates at four time points. At each time point, 3 × 107 cells were harvested and fixed with 1% of formaldehyde for 8 min (Thermo Scientific, 28908). Formaldehyde was quenched with 0.2 M glycine for 5 min. Chromatin was sheared to 200–500 bp in size using the Covaris S220 ultrasonicator. A fraction of the resulting chromatin was used as input control. Sheared chromatin (50 μg for C/EBPα, 30 μg for H3K4me1, 100 μg for BRD4) was incubated with 5 μg of antibodies – anti-C/EBPα (Santa Cruz, sc-61), anti-H3K4me1 (abcam, ab8895), and anti-BRD4 (Bethyl Laboratories, A301–985A100) – bound to protein-A beads (Life Technologies, 9999–01). Samples were incubated at 65°C overnight to reverse the crosslinks, followed by RNase and proteinase K treatments. Purified ChIP and input DNA were quantified by Qubit 2.0 Fluorometer (Life Technologies, Q32866). Libraries were generated using NEB Ultra DNA Library kit (NEB, E7370S) and amplified by nine cycles with KAPA Real-Time Library Amplification Kit (Peqlab, KK2701). The libraries of BRD4-ChIP-seq were sequenced at 2 × 75 bp on an Illumina NextSeq machine in house. The other libraries were sequenced at 2 × 50 bp on HiSeq2000 (TAL, Göttingen or LAFUGA, LMU München). The average read numbers obtained were 4 × 107 for C/EBPα ChIP-seq, 7 × 107 for H3K4me1 ChIP-seq, and 2 × 107 for BRD4 ChIP-seq.

ATAC-seq

ATAC-seq was performed from two biological replicates at four time points as described (Buenrostro et al., 2013). Briefly, 5 × 104 cells at each time point were harvested and treated with Nextera Tn5 Transposase (Illumina, FC-121–1030) for 45 min at 37°C. Library fragments were amplified using 1× NEBNext High-Fidelity 2× PCR Master Mix (NEB, M0541S) and 1.25 μM of custom Nextera PCR primers 1 and 2. PCR amplification was done with 11 cycles, determined by KAPA Real-Time Library Amplication Kit (Peqlab, KK2701) to stop prior to saturation. Then, the samples were purified using Qiagen MinElute PCR Purification Kit (Qiagen, 28004) and with Agencourt AMPure XP beads (Beckman Coulter, A63881) in 3:1 ratio. The libraries were sequenced at 2 × 50 bp on HiSeq2000 (TAL, Göttingen). The average read numbers obtained were 3 × 107.

Fluorescence-activated cell sorting (FACS)

Cells were harvested, washed with PBS, and blocked with Human FC Receptor Binding Inhibitor (ThermoFisher, 16–9161). Cells were incubated with antibodies, APC-cy7 Mouse Anti-Human CD19 (BD Pharmingen, 557791), PE Mouse Anti-Human CD14 (BD Pharmingen, 555398). After washing with PBS, cells were resuspended in 7AAD (ThermoFisher, 00–6993) diluted with PBS. The analysis was performed on BD Accuri C6 flow cytometer (BD Biosciences) according to standard protocols with following changes in filters. We used filter 675/25 at F3 and 670/LP at F4.

Quantitative RT-PCR (qRT-PCR)

Cells were harvested with QIAzol and purified using RNeasy Mini Kit (Qiagen, 74104). The purified RNA was quantified using a Nanodrop spectrophotometer (Thermofisher), and cDNA was generated by reverse transcription PCR. The resulting cDNA was used as a template for quantitative PCR (qPCR, SybrGreen). Ct values were normalized with GAPDH expression. Relative expression levels were calculated at each time points compared to 0 hr for B cell markers or to 168 hr for macrophage markers. Primers used for qRT-PCR are listed in Supplementary file 6. Error bars represent standard deviation from three biological replicates.

Quantification and statistical analysis

All sequencing data were trimmed using Cutadapt 1.9.1 (Martin, 2012) and trim_galore_v0.4.1 (https://www.bioinformatics.babraham.ac.uk/projects/trim_galore/). Read quality was screened using FastQC (http://www.bioinformatics.bbsrc.ac.uk/projects/fastqc). After alignments, Samtools1.2 (Li et al., 2009) was used to filter the alignments with MAPQ smaller than 7 (-q 7), and to retain only the proper pairs (-f99, -f147, -f83, -f163). Boxplots show inter-quantile ranges, with the median indicated with black line. Whiskers of boxplots represent maximum and minimum values excluding outliers; outliers were calculated as values greater or lower than 1.5 times the interquartile range. Significance was tested by Wilcoxon test, unless otherwise stated.

RNA-seq and TT-seq data processing

Trimmed reads were aligned to GRCh38 genome assembly (Human Genome Reference Consortium) using STAR2.4.2 (Dobin et al., 2013). For coverage profiles and visualization, reads were uniquely mapped, antisense corrected, and normalized with size factors calculated from DESeq2 (Love et al., 2014; Michel et al., 2017). Spearman's correlation coefficients between replicates were 0.99 for total RNA-seq and TT-seq on protein coding genes at all time points.

ChIP-seq data processing

Trimmed reads were aligned to GRCh38 genome assembly using Bowtie2.2.5 (Langmead and Salzberg, 2012). ChIP-seq peaks were called using the 'callpeak' function of MACS2.1.1 (Zhang et al., 2008). We used ‘--BAMPE’ options of MACS2.1.1. To catalogue the binding sites over all the time points, we merged and reduced all peaks, and removed those falling within ±4 Mb of centromeric heterochromatin regions. Duplicates were removed. Read numbers, Spearman's correlation coefficients between replicates, FRiP, and peak numbers are tabulated in Supplementary file 7.

ATAC-seq data processing

Trimmed reads were aligned to GRCh38 genome assembly without mitochondrial chromosome (ChrM) using Bowtie1.0.0 (Langmead et al., 2009). Peaks were called using the 'callpeak' function, with ‘--shift 37 --extsize 73’ options of MACS2.1.1. Peaks called with ‘--broad’ and ‘--narrow’ were combined and used for further analysis. Duplicates were removed.

RNA-seq data analysis

For each time points, read count data were generated on RefSeq annotation using HTSeq0.6.1p1 (Anders et al., 2015). For Figure 1C, we used DESeq2 to obtain normalized count data and differentially expressed genes at each time point compared to any other time points with a threshold of |log2 fold change| >1 and false discovery rate <0.05. The resulting genes were clustered using k-means clustering. Principal component analysis was carried out using the built-in R function, prcomp(). PCA plot in Figure 1D was generated using ggplot2 (http://ggplot2.org). Loadings were computed, and top 1000 genes with high loadings were selected for component 1 and 2. Among 1000 genes, the ones that were not found in the other group were retained for further analysis. Heatmaps were generated from gene expression data with pheatmap (https://CRAN.R-project.org/package=pheatmap). Gene ontology (GO) analysis was performed with DAVID (Huang et al., 2009).

TU annotation

TUs were annotated from TT-seq data as described (Michel et al., 2017), with modifications. Briefly, we used the GenoSTAN package in R/Bioconductor (Zacher et al., 2017) for segmenting the genome into two states, ‘expressed’ and ‘not expressed’, based on combined TT-seq data at 0 hr, 12 hr, 24 hr, and 96 hr. Posterior state probabilities were calculated, and the most likely state path (Viterbi) was obtained. TUs within 200 bp as well as TUs mapping to exons within GENCODE (Human, Release 25) annotated genes were merged together. The Jaccard Index was computed based on GENCODE (all support levels) and used as an expression cutoff.

Classification of gene types

Gene types from GENCODE were taken if TUs overlapped with more than 60% of GENCODE annotation. Downstream RNAs (dsRNAs) were ncRNAs that lie within 1 kb downstream of protein-coding genes or products from stepwise termination of RNA polymerase II (Schwalb et al., 2016); ncRNAs that lie within 10 kb downstream of protein-coding genes with diminishing synthesis level over the genomic position. TUs within 1000 bp upstream and 500 bp downstream of the transcription start site (TSS) of protein-coding genes were assigned as upstream antisense RNA (uaRNA) if synthesized on upstream of the opposite strand as protein-coding gene TSS or convergent RNA (convRNA) if synthesized on the overlapping opposite strand as protein-coding gene TSS. Among rest of TUs, TUs overlapping with GENCODE (gene_type = gene) more than 20% of GENCODE were discarded. Of the remaining, ncRNAs within 1 kb of ATAC-seq and H3K4me1-ChIP-seq signals at any time point were assigned as eRNAs. The eRNAs within 1 kb of each other were merged regardless of synthesized strands, assuming to be synthesized from one enhancer. To compute length normalized counts of merged enhancers, counts were normalized to lengths of each eRNA TUs before merging. These normalized counts were added together for merged enhancers.

Enhancer–promoter pairing

Each eRNA was paired with the putative target mRNAs by two methods; the neighboring method and 1 Mb method. For both methods, the eRNAs and mRNAs were taken from our annotation with TT-seq data and had to be synthesized at least at one time point. The neighboring method paired the eRNAs with mRNAs based on distances. First, intragenic eRNAs were paired with the mRNA in whose transcribed region they lay. The rest of the eRNAs was paired with both the closest upstream and downstream mRNAs. Of 12,900 pairs, 5786 pairs (45%) with differentially synthesized mRNAs were used (|log2FC| > 1, FDR < 0.05, computed by DESeq2). As a result, 4721 of 7624 enhancers were paired to 1790 genes (Figure 2B). The 1 Mb method paired the enhancers with the mRNAs that were differentially synthesized (|log2FC| > 1, FDR < 0.05, computed by DESeq2), and within 1 Mb of the TSS of eRNAs. Pairs that had a Spearman's correlation coefficient between mRNA and the eRNA read counts higher than 0.4 were used. TAD pairing was carried out as for 1 Mb pairing, except that experimentally determined TADs (Stik et al., 2020) were used instead of the 1 Mb linear distance.

TAD analysis

TAD information of BLaER cell line during transdifferentiation was obtained from Stik et al., 2020. We also confirmed that the enhancers and the paired target genes are within the same TAD, using previously published Hi-C data from the human monocytic leukemia cell line THP-1 (Phanstiel et al., 2017). We could use the data from the related cell line because TADs are known to be well conserved across cell types (Rao et al., 2014). Figures were generated using ‘The 3D Genome Browser’ (http://promoter.bx.psu.edu/hi-c/).

Correlation between enhancer–promoter pairs

Enhancer–promoter (EP) pairs were classified based on the number of enhancers paired with the same promoter. We obtained the log2 fold change of length normalized mRNA and eRNA synthesis from 0 hr. For Figure 4 and S4, we computed the Spearman's correlation coefficient using a sum of changes in eRNA synthesis as explanatory variable and a change in mRNA synthesis as target variable. The log2 fold change of eRNA and mRNA synthesis were regressed using the lm() function in R, and regression coefficient and coefficient of determination (R-squared, R2) were obtained.

Model to predict promoter activity based on the activity of multiple enhancers

Theoretical models were suggested by Dukler et al., 2017.

  • The additive relationship is described as follows:

  • Promoter activity = α × (sum of enhancer activities) + β

  • The synergistic (exponential) relationship is described as:

  • Promoter activity = exp[α × (sum of enhancer activities) + β]

  • The logistic relationship is described as follows:

  • Promoter activity = δ/(1 + exp[−(α × (sum of enhancer activities) + β)])

The constants α, β, and δ describe the effect size of enhancer activities on their target promoter and vary between promoters. It can be assumed that these constants are fixed for each enhancer−promoter pair, even upon stimulation. The promoter activity was predicted from the sum of enhancer activities using generalized linear models. The additive and synergistic models were fitted with the use of normal error mode using the DEoptim package in R/Bioconductor and codes adapted from Dukler et al., 2017. The goodness of fit of the models was assessed using the BIC as in Dukler et al., 2017. Enhancer–promoter pairs with a good correlation of mRNA and eRNA synthesis (top 75% of Spearman’s correlation coefficients) were further analyzed. First, we identified the genes with lowest BIC among each model, additive, synergistic, or logistic (Figure 5—figure supplement 1D,E and G). For the genes with the lowest BIC for additive or synergistic model (n = 523, 1017 with the neighboring or 1 Mb method, respectively), relative BIC was computed as a difference of BIC of the additive model compared to the exponential model. The synergistic regulatory mode was favored over the additive mode if the relative BIC was greater than 2. The regulatory mode was ambiguous if the relative BIC was between 0 and 2, and was considered to be additive if the relative BIC was less than 0. If BIC for the logistic model was the lowest (n = 250, 429 with the neighboring or 1 Mb method, respectively), we excluded the data point with the highest sum of enhancer activities and repeated from assessing BIC for each model.

Superenhancer prediction using ROSE

Superenhancers are predicted as suggested in Lovén et al., 2013; Whyte et al., 2013. Briefly, ATAC-seq peaks, C/EBPα, BRD4, and H3K27Ac ChIP-seq peaks within 1 kb of our enhancers were stitched together if they fell within 12.5 kb. Signals were normalized to the peak widths. Highly ranked peaks were identified as superenhancers. The putative target genes of superenhancers are listed and compared with the genes predicted to be regulated by superenhancers in dbSUPER, superenhancer database (Khan and Zhang, 2016; Supplementary file 1). Gene ontology analysis was performed on the putative target genes of superenhancers with DAVID (Huang et al., 2009; Supplementary files 25). Previously identified superenhancers from all available cell lines from dbSUPER, CD14+ cells, or CD19+ cells (Hnisz et al., 2013) were used to compare with the enhancers identified in this study. Specifically, acetylated H3K27 were used to classify superenhancers in CD14+ and CD19+ cells (Hnisz et al., 2013). Before the comparison, genome coordinates of our enhancers were converted to GRCh37/hg19 genome assembly (Human Genome Reference Consortium) using liftOver (Hinrichs et al., 2006).

TF footprinting analysis

TF bound motifs were identified by PIQ, a TF footprinting method (Sherwood et al., 2014). First, GRCh38 genome was scanned with the motifs in Jaspar2018 CORE vertebrates database (Khan et al., 2018). With the combined coverage of two replicates of ATAC-seq, putative TF binding sites were computed at each time point, 0 hr, 12 hr, 24 hr, and 96 hr with a default threshold of PIQ. Significantly enriched motifs in late synergistic enhancers relative to additive enhancers were determined using Fisher's exact test (p-value < 0.05) for each time point. The SeqLogo function in the TFBSTools package (Tan and Lenhard, 2016) was used to generate motif logos. Using TFClass classification (Wingender et al., 2018), TFs in the same subfamily as enriched motifs were further analyzed.

Acknowledgements

We thank the past and current members of the Cramer laboratory, in particular B Schwalb, M Lidschreiber, and S Sohrabi-Jahromi for help with bioinformatical analysis, K Maier and P Rus for sequencing and A Sawicka for advice on gene reporter assays. We also thank LAFUGA at the Gene Center, Ludwig-Maximilians-University of Munich, and TAL at University Medical Center Göttingen for sequencing. PC was supported by the Deutsche Forschungsgemeinschaft (SFB860, SPP1935, SPP2191), the European Research Council Advanced Investigator Grant TRANSREGULON (grant agreement No 693023), and the Volkswagen Foundation.

Funding Statement

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Contributor Information

Patrick Cramer, Email: patrick.cramer@mpibpc.mpg.de.

Chris P Ponting, University of Edinburgh, United Kingdom.

Kevin Struhl, Harvard Medical School, United States.

Funding Information

This paper was supported by the following grants:

  • Max-Planck-Gesellschaft Open-access funding to Patrick Cramer.

  • Deutsche Forschungsgemeinschaft SFB860 to Patrick Cramer.

  • Deutsche Forschungsgemeinschaft SPP1935 to Patrick Cramer.

  • Deutsche Forschungsgemeinschaft SPP2191 to Patrick Cramer.

  • European Research Council 693023 to Patrick Cramer.

  • Volkswagen Foundation to Patrick Cramer.

Additional information

Competing interests

No competing interests declared.

Author contributions

Conceptualization, Data curation, Formal analysis, Validation, Investigation, Visualization, Methodology, Writing - original draft, Writing - review and editing, Designed research, performed all experiments except when stated otherwise, analysed and interpreted the data.

Validation, Investigation, Methodology, Writing - review and editing, Set up and performed CRISPR validation and gene reporter assays.

Validation, Investigation, Methodology, Carried out CRISPR-based enhancer deletion and provided Hi-C data prior to publication.

Data curation, Formal analysis, Carried out TU annotation.

Conceptualization, Formal analysis, Supervision, Writing - review and editing, Advised on data analysis and manuscript writing.

Methodology, Provided BLaER1 cell line and instructions for transdifferentiation.

Conceptualization, Methodology, Writing - review and editing, Provided BLaER1 cell line and instructions for transdifferentiation, advised on manuscript writing.

Conceptualization, Resources, Supervision, Funding acquisition, Investigation, Visualization, Methodology, Writing - original draft, Project administration, Writing - review and editing.

Additional files

Supplementary file 1. List of superenhancers identified from chromatin accessibility, BRD4, C/EBPα, and H3K27Ac occupancies.
elife-65381-supp1.xlsx (48.4KB, xlsx)
Supplementary file 2. GO analysis on superenhancers identified from BRD4 signal.
elife-65381-supp2.xlsx (86.3KB, xlsx)
Supplementary file 3. GO analysis on superenhancers identified from H3K27Ac ChIP-seq signal.
elife-65381-supp3.xlsx (99.2KB, xlsx)
Supplementary file 4. GO analysis on superenhancers identified from C/EBPα signal.
elife-65381-supp4.xlsx (122.6KB, xlsx)
Supplementary file 5. GO analysis on superenhancers identified from chromatin accessibility signal.
elife-65381-supp5.xlsx (63.9KB, xlsx)
Supplementary file 6. Primer list.
elife-65381-supp6.docx (12.9KB, docx)
Supplementary file 7. Quality assessment of ChIP seq data sets.
elife-65381-supp7.xlsx (15.6KB, xlsx)
Transparent reporting form

Data availability

RNA-seq, TT-seq, ChIP-seq, ATAC-seq data reported in this study were deposited with the National Center for Biotechnology Information Gene Expression Omnibus (accession number GSE131620). Hi-C data and H3K27Ac ChIP-seq in BLaER and Hi-C data in THP-1 cell lines that support the findings of this study are available with the National Center for Biotechnology Information Gene Expression Omnibus (accession GSE141226) and BioProject (accession PRJNA385337).

The following dataset was generated:

Choi J, Lysakovskaia K, Stik G, Demel C, Soeding J, Tian TV, Graf T, Cramer P. 2020. Evidence for additive and synergistic action of mammalian enhancers during cell fate determination. NCBI Gene Expression Omnibus. GSE131620

The following previously published datasets were used:

Stik G, Casadesus MV, Graf T. 2020. CTCF is dispensable for cell fate conversion but facilitates acute cellular responses [Hi-C] NCBI Gene Expression Omnibus. GSE141226

UNC Chapel Hill. 2017. in situ Hi-C data of THP-1 cells untreated and treated with PMA. NCBI BioProject. PRJNA385337

References

  1. Allahyar A, Vermeulen C, Bouwman BAM, Krijger PHL, Verstegen M, Geeven G, van Kranenburg M, Pieterse M, Straver R, Haarhuis JHI, Jalink K, Teunissen H, Renkens IJ, Kloosterman WP, Rowland BD, de Wit E, de Ridder J, de Laat W. Enhancer hubs and loop collisions identified from single-allele topologies. Nature Genetics. 2018;50:1151–1160. doi: 10.1038/s41588-018-0161-5. [DOI] [PubMed] [Google Scholar]
  2. Anders S, Pyl PT, Huber W. HTSeq--a Python framework to work with high-throughput sequencing data. Bioinformatics. 2015;31:166–169. doi: 10.1093/bioinformatics/btu638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Bahr C, von Paleske L, Uslu VV, Remeseiro S, Takayama N, Ng SW, Murison A, Langenfeld K, Petretich M, Scognamiglio R, Zeisberger P, Benk AS, Amit I, Zandstra PW, Lupien M, Dick JE, Trumpp A, Spitz F. A myc enhancer cluster regulates normal and leukaemic haematopoietic stem cell hierarchies. Nature. 2018;553:515–520. doi: 10.1038/nature25193. [DOI] [PubMed] [Google Scholar]
  4. Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, Ren J, Li WW, Noble WS. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Research. 2009;37:W202–W208. doi: 10.1093/nar/gkp335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Beagrie RA, Scialdone A, Schueler M, Kraemer DCA, Chotalia M, Xie SQ, Barbieri M, de Santiago I, Lavitas L-M, Branco MR, Fraser J, Dostie J, Game L, Dillon N, Edwards PAW, Nicodemi M, Pombo A. Complex multi-enhancer contacts captured by genome architecture mapping. Nature. 2017;543:519–524. doi: 10.1038/nature21411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Beagrie RA, Pombo A. Gene activation by metazoan enhancers: diverse mechanisms stimulate distinct steps of transcription. BioEssays. 2016;38:881–893. doi: 10.1002/bies.201600032. [DOI] [PubMed] [Google Scholar]
  7. Boehning M, Dugast-Darzacq C, Rankovic M, Hansen AS, Yu T, Marie-Nelly H, McSwiggen DT, Kokic G, Dailey GM, Cramer P, Darzacq X, Zweckstetter M. RNA polymerase II clustering through carboxy-terminal domain phase separation. Nature Structural & Molecular Biology. 2018;25:833–840. doi: 10.1038/s41594-018-0112-y. [DOI] [PubMed] [Google Scholar]
  8. Boija A, Klein IA, Sabari BR, Dall'Agnese A, Coffey EL, Zamudio AV, Li CH, Shrinivas K, Manteiga JC, Hannett NM, Abraham BJ, Afeyan LK, Guo YE, Rimel JK, Fant CB, Schuijers J, Lee TI, Taatjes DJ, Young RA. Transcription factors activate genes through the Phase-Separation capacity of their activation domains. Cell. 2018;175:1842–1855. doi: 10.1016/j.cell.2018.10.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bonn S, Zinzen RP, Girardot C, Gustafson EH, Perez-Gonzalez A, Delhomme N, Ghavi-Helm Y, Wilczyński B, Riddell A, Furlong EE. Tissue-specific analysis of chromatin state identifies temporal signatures of enhancer activity during embryonic development. Nature Genetics. 2012;44:148–156. doi: 10.1038/ng.1064. [DOI] [PubMed] [Google Scholar]
  10. Buenrostro JD, Giresi PG, Zaba LC, Chang HY, Greenleaf WJ. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nature Methods. 2013;10:1213–1218. doi: 10.1038/nmeth.2688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Calo E, Wysocka J. Modification of enhancer chromatin: what, how, and why? Molecular Cell. 2013;49:825–837. doi: 10.1016/j.molcel.2013.01.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Cao Z, Sun X, Icli B, Wara AK, Feinberg MW. Role of Kruppel-like factors in leukocyte development, function, and disease. Blood. 2010;116:4404–4414. doi: 10.1182/blood-2010-05-285353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Chong S, Dugast-Darzacq C, Liu Z, Dong P, Dailey GM, Cattoglio C, Heckert A, Banala S, Lavis L, Darzacq X, Tjian R. Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science. 2018;361:eaar2555. doi: 10.1126/science.aar2555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Cobaleda C, Schebesta A, Delogu A, Busslinger M. Pax5: the guardian of B cell identity and function. Nature Immunology. 2007;8:463–470. doi: 10.1038/ni1454. [DOI] [PubMed] [Google Scholar]
  15. Creyghton MP, Cheng AW, Welstead GG, Kooistra T, Carey BW, Steine EJ, Hanna J, Lodato MA, Frampton GM, Sharp PA, Boyer LA, Young RA, Jaenisch R. Histone H3K27ac separates active from poised enhancers and predicts developmental state. PNAS. 2010;107:21931–21936. doi: 10.1073/pnas.1016071107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Date D, Das R, Narla G, Simon DI, Jain MK, Mahabeleshwar GH. Kruppel-like transcription factor 6 regulates inflammatory macrophage polarization. Journal of Biological Chemistry. 2014;289:10318–10329. doi: 10.1074/jbc.M113.526749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Dixon JR, Selvaraj S, Yue F, Kim A, Li Y, Shen Y, Hu M, Liu JS, Ren B. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature. 2012;485:376–380. doi: 10.1038/nature11082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, Gingeras TR. STAR: ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29:15–21. doi: 10.1093/bioinformatics/bts635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Dorighi KM, Swigut T, Henriques T, Bhanu NV, Scruggs BS, Nady N, Still CD, Garcia BA, Adelman K, Wysocka J. Mll3 and Mll4 facilitate enhancer RNA synthesis and transcription from promoters independently of H3K4 monomethylation. Molecular Cell. 2017;66:568–576. doi: 10.1016/j.molcel.2017.04.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Dukler N, Gulko B, Huang Y-F, Siepel A. Is a super-enhancer greater than the sum of its parts? Nature Genetics. 2017;49:2–3. doi: 10.1038/ng.3759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Friedman AD. Transcriptional control of granulocyte and monocyte development. Oncogene. 2007;26:6816–6828. doi: 10.1038/sj.onc.1210764. [DOI] [PubMed] [Google Scholar]
  22. Fukaya T, Lim B, Levine M. Enhancer control of transcriptional bursting. Cell. 2016;166:358–368. doi: 10.1016/j.cell.2016.05.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Fulton R, van Ness B. Selective synergy of immunoglobulin enhancer elements in B-cell development: a characteristic of kappa light chain enhancers, but not heavy chain enhancers. Nucleic Acids Research. 1994;22:4216–4223. doi: 10.1093/nar/22.20.4216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Gaidt MM, Ebert TS, Chauhan D, Schmidt T, Schmid-Burgk JL, Rapino F, Robertson AA, Cooper MA, Graf T, Hornung V. Human monocytes engage an alternative inflammasome pathway. Immunity. 2016;44:833–846. doi: 10.1016/j.immuni.2016.01.012. [DOI] [PubMed] [Google Scholar]
  25. Gressel S, Schwalb B, Decker TM, Qin W, Leonhardt H, Eick D, Cramer P. CDK9-dependent RNA polymerase II pausing controls transcription initiation. eLife. 2017;6:e29736. doi: 10.7554/eLife.29736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Guerrero L, Marco-Ferreres R, Serrano AL, Arredondo JJ, Cervera M. Secondary enhancers synergise with primary enhancers to guarantee fine-tuned muscle gene expression. Developmental Biology. 2010;337:16–28. doi: 10.1016/j.ydbio.2009.10.006. [DOI] [PubMed] [Google Scholar]
  27. Haberle V, Stark A. Eukaryotic core promoters and the functional basis of transcription initiation. Nature Reviews Molecular Cell Biology. 2018;19:621–637. doi: 10.1038/s41580-018-0028-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Haddad R, Guardiola P, Izac B, Thibault C, Radich J, Delezoide AL, Baillou C, Lemoine FM, Gluckman JC, Pflumio F, Canque B. Molecular characterization of early human T/NK and B-lymphoid progenitor cells in umbilical cord blood. Blood. 2004;104:3918–3926. doi: 10.1182/blood-2004-05-1845. [DOI] [PubMed] [Google Scholar]
  29. Harrow J, Frankish A, Gonzalez JM, Tapanari E, Diekhans M, Kokocinski F, Aken BL, Barrell D, Zadissa A, Searle S, Barnes I, Bignell A, Boychenko V, Hunt T, Kay M, Mukherjee G, Rajan J, Despacio-Reyes G, Saunders G, Steward C, Harte R, Lin M, Howald C, Tanzer A, Derrien T, Chrast J, Walters N, Balasubramanian S, Pei B, Tress M, Rodriguez JM, Ezkurdia I, van Baren J, Brent M, Haussler D, Kellis M, Valencia A, Reymond A, Gerstein M, Guigó R, Hubbard TJ. GENCODE: the reference human genome annotation for the ENCODE project. Genome Research. 2012;22:1760–1774. doi: 10.1101/gr.135350.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Hay D, Hughes JR, Babbs C, Davies JOJ, Graham BJ, Hanssen L, Kassouf MT, Marieke Oudelaar AM, Sharpe JA, Suciu MC, Telenius J, Williams R, Rode C, Li PS, Pennacchio LA, Sloane-Stanley JA, Ayyub H, Butler S, Sauka-Spengler T, Gibbons RJ, Smith AJH, Wood WG, Higgs DR. Genetic dissection of the α-globin super-enhancer in vivo. Nature Genetics. 2016;48:895–903. doi: 10.1038/ng.3605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Heintzman ND, Stuart RK, Hon G, Fu Y, Ching CW, Hawkins RD, Barrera LO, Van Calcar S, Qu C, Ching KA, Wang W, Weng Z, Green RD, Crawford GE, Ren B. Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the human genome. Nature Genetics. 2007;39:311–318. doi: 10.1038/ng1966. [DOI] [PubMed] [Google Scholar]
  32. Heinz S, Benner C, Spann N, Bertolino E, Lin YC, Laslo P, Cheng JX, Murre C, Singh H, Glass CK. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Molecular Cell. 2010;38:576–589. doi: 10.1016/j.molcel.2010.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Henriques T, Scruggs BS, Inouye MO, Muse GW, Williams LH, Burkholder AB, Lavender CA, Fargo DC, Adelman K. Widespread transcriptional pausing and elongation control at enhancers. Genes & Development. 2018;32:26–41. doi: 10.1101/gad.309351.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Hinrichs AS, Karolchik D, Baertsch R, Barber GP, Bejerano G, Clawson H, Diekhans M, Furey TS, Harte RA, Hsu F, Hillman-Jackson J, Kuhn RM, Pedersen JS, Pohl A, Raney BJ, Rosenbloom KR, Siepel A, Smith KE, Sugnet CW, Sultan-Qurraie A, Thomas DJ, Trumbower H, Weber RJ, Weirauch M, Zweig AS, Haussler D, Kent WJ. The UCSC genome browser database: update 2006. Nucleic Acids Research. 2006;34:D590–D598. doi: 10.1093/nar/gkj144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Hnisz D, Abraham BJ, Lee TI, Lau A, Saint-André V, Sigova AA, Hoke HA, Young RA. Super-enhancers in the control of cell identity and disease. Cell. 2013;155:934–947. doi: 10.1016/j.cell.2013.09.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Hnisz D, Shrinivas K, Young RA, Chakraborty AK, Sharp PA. A phase separation model for transcriptional control. Cell. 2017;169:13–23. doi: 10.1016/j.cell.2017.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Huang daW, Sherman BT, Lempicki RA. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nature Protocols. 2009;4:44–57. doi: 10.1038/nprot.2008.211. [DOI] [PubMed] [Google Scholar]
  38. Iberg-Badeaux A, Collombet S, Laurent B, van Oevelen C, Chin KK, Thieffry D, Graf T, Shi Y. A transcription factor pulse can prime chromatin for heritable transcriptional memory. Molecular and Cellular Biology. 2017;37:e00372-16. doi: 10.1128/MCB.00372-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Ikeda RA, Lin AC, Clarke J. Initiation of transcription by T7 RNA polymerase as its natural promoters. Journal of Biological Chemistry. 1992;267:2640–2649. doi: 10.1016/S0021-9258(18)45929-7. [DOI] [PubMed] [Google Scholar]
  40. Jubb AW, Boyle S, Hume DA, Bickmore WA. Glucocorticoid receptor binding induces rapid and prolonged Large-Scale chromatin decompaction at multiple target loci. Cell Reports. 2017;21:3022–3031. doi: 10.1016/j.celrep.2017.11.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Kaikkonen MU, Spann NJ, Heinz S, Romanoski CE, Allison KA, Stender JD, Chun HB, Tough DF, Prinjha RK, Benner C, Glass CK. Remodeling of the enhancer landscape during macrophage activation is coupled to enhancer transcription. Molecular Cell. 2013;51:310–325. doi: 10.1016/j.molcel.2013.07.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Keshav S, Chung P, Milon G, Gordon S. Lysozyme is an inducible marker of macrophage activation in murine tissues as demonstrated by in situ hybridization. Journal of Experimental Medicine. 1991;174:1049–1058. doi: 10.1084/jem.174.5.1049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Khan A, Fornes O, Stigliani A, Gheorghe M, Castro-Mondragon JA, van der Lee R, Bessy A, Chèneby J, Kulkarni SR, Tan G, Baranasic D, Arenillas DJ, Sandelin A, Vandepoele K, Lenhard B, Ballester B, Wasserman WW, Parcy F, Mathelier A. JASPAR 2018: update of the open-access database of transcription factor binding profiles and its web framework. Nucleic Acids Research. 2018;46:D1284. doi: 10.1093/nar/gkx1188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Khan A, Zhang X. dbSUPER: a database of super-enhancers in mouse and human genome. Nucleic Acids Research. 2016;44:D164–D171. doi: 10.1093/nar/gkv1002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Kim TK, Kim TH, Maniatis T. Efficient recruitment of TFIIB and CBP-RNA polymerase II holoenzyme by an interferon-beta enhanceosome in vitro. PNAS. 1998;95:12191–12196. doi: 10.1073/pnas.95.21.12191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Kranc KR, Schepers H, Rodrigues NP, Bamforth S, Villadsen E, Bouriez-Jones T, Sigvardsson M, Bhattacharya S, Jacobsen SE, Jacobsen SE, Enver T. Cited2 is an essential regulator of adult hematopoietic stem cells. Cell Stem Cell. 2009;5:659–665. doi: 10.1016/j.stem.2009.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Langmead B, Trapnell C, Pop M, Salzberg SL. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biology. 2009;10:R25. doi: 10.1186/gb-2009-10-3-r25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Langmead B, Salzberg SL. Fast gapped-read alignment with bowtie 2. Nature Methods. 2012;9:357–359. doi: 10.1038/nmeth.1923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Larsson AJM, Johnsson P, Hagemann-Jensen M, Hartmanis L, Faridani OR, Reinius B, Segerstolpe Å, Rivera CM, Ren B, Sandberg R. Genomic encoding of transcriptional burst kinetics. Nature. 2019;565:251–254. doi: 10.1038/s41586-018-0836-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Lavin Y, Winter D, Blecher-Gonen R, David E, Keren-Shaul H, Merad M, Jung S, Amit I. Tissue-Resident macrophage enhancer landscapes are shaped by the local microenvironment. Cell. 2014;159:1312–1326. doi: 10.1016/j.cell.2014.11.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Lee JE, Wang C, Xu S, Cho YW, Wang L, Feng X, Baldridge A, Sartorelli V, Zhuang L, Peng W, Ge K. H3K4 mono- and di-methyltransferase MLL4 is required for enhancer activation during cell differentiation. eLife. 2013;2:e01503. doi: 10.7554/eLife.01503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Li H, Gade P, Nallar SC, Raha A, Roy SK, Karra S, Reddy JK, Reddy SP, Kalvakolanu DV. The Med1 subunit of transcriptional mediator plays a central role in regulating CCAAT/enhancer-binding protein-beta-driven transcription in response to interferon-gamma. Journal of Biological Chemistry. 2008;283:13077–13086. doi: 10.1074/jbc.M800604200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, Marth G, Abecasis G, Durbin R, 1000 Genome Project Data Processing Subgroup The sequence alignment/Map format and SAMtools. Bioinformatics. 2009;25:2078–2079. doi: 10.1093/bioinformatics/btp352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Lim B, Fukaya T, Heist T, Levine M. Temporal dynamics of pair-rule stripes in living Drosophila embryos. PNAS. 2018;115:8376–8381. doi: 10.1073/pnas.1810430115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Long HK, Prescott SL, Wysocka J. Ever-Changing landscapes: transcriptional enhancers in development and evolution. Cell. 2016;167:1170–1187. doi: 10.1016/j.cell.2016.09.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biology. 2014;15:550. doi: 10.1186/s13059-014-0550-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Lovén J, Hoke HA, Lin CY, Lau A, Orlando DA, Vakoc CR, Bradner JE, Lee TI, Young RA. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell. 2013;153:320–334. doi: 10.1016/j.cell.2013.03.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Maekawa T, Imamoto F, Merlino GT, Pastan I, Ishii S. Cooperative function of two separate enhancers of the human epidermal growth factor receptor proto-oncogene. Journal of Biological Chemistry. 1989;264:5488–5494. doi: 10.1016/S0021-9258(18)83571-2. [DOI] [PubMed] [Google Scholar]
  59. Martin M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal. 2012;17:10–12. doi: 10.14806/ej.17.1.200. [DOI] [Google Scholar]
  60. Michel M, Demel C, Zacher B, Schwalb B, Krebs S, Blum H, Gagneur J, Cramer P. TT-seq captures enhancer landscapes immediately after T-cell stimulation. Molecular Systems Biology. 2017;13:920. doi: 10.15252/msb.20167507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Mikhaylichenko O, Bondarenko V, Harnett D, Schor IE, Males M, Viales RR, Furlong EEM. The degree of enhancer or promoter activity is reflected by the levels and directionality of eRNA transcription. Genes & Development. 2018;32:42–57. doi: 10.1101/gad.308619.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Moorthy SD, Davidson S, Shchuka VM, Singh G, Malek-Gilani N, Langroudi L, Martchenko A, So V, Macpherson NN, Mitchell JA. Enhancers and super-enhancers have an equivalent regulatory role in embryonic stem cells through regulation of single or multiple genes. Genome Research. 2017;27:246–258. doi: 10.1101/gr.210930.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Murray DT, Kato M, Lin Y, Thurber KR, Hung I, McKnight SL, Tycko R. Structure of FUS protein fibrils and its relevance to Self-Assembly and phase separation of Low-Complexity domains. Cell. 2017;171:615–627. doi: 10.1016/j.cell.2017.08.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Nair SJ, Yang L, Meluzzi D, Oh S, Yang F, Friedman MJ, Wang S, Suter T, Alshareedah I, Gamliel A, Ma Q, Zhang J, Hu Y, Tan Y, Ohgi KA, Jayani RS, Banerjee PR, Aggarwal AK, Rosenfeld MG. Phase separation of ligand-activated enhancers licenses cooperative chromosomal enhancer assembly. Nature Structural & Molecular Biology. 2019;26:193–203. doi: 10.1038/s41594-019-0190-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Osterwalder M, Barozzi I, Tissières V, Fukuda-Yuzawa Y, Mannion BJ, Afzal SY, Lee EA, Zhu Y, Plajzer-Frick I, Pickle CS, Kato M, Garvin TH, Pham QT, Harrington AN, Akiyama JA, Afzal V, Lopez-Rios J, Dickel DE, Visel A, Pennacchio LA. Enhancer redundancy provides phenotypic robustness in mammalian development. Nature. 2018;554:239–243. doi: 10.1038/nature25461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Oudelaar AM, Davies JOJ, Hanssen LLP, Telenius JM, Schwessinger R, Liu Y, Brown JM, Downes DJ, Chiariello AM, Bianco S, Nicodemi M, Buckle VJ, Dekker J, Higgs DR, Hughes JR. Single-allele chromatin interactions identify regulatory hubs in dynamic compartmentalized domains. Nature Genetics. 2018;50:1744–1751. doi: 10.1038/s41588-018-0253-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Pedersen TA, Kowenz-Leutz E, Leutz A, Nerlov C. Cooperation between C/EBPalpha TBP/TFIIB and SWI/SNF recruiting domains is required for adipocyte differentiation. Genes & Development. 2001;15:3208–3216. doi: 10.1101/gad.209901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Perry MW, Boettiger AN, Levine M. Multiple enhancers ensure precision of gap gene-expression patterns in the Drosophila embryo. PNAS. 2011;108:13570–13575. doi: 10.1073/pnas.1109873108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Phanstiel DH, Van Bortle K, Spacek D, Hess GT, Shamim MS, Machol I, Love MI, Aiden EL, Bassik MC, Snyder MP. Static and dynamic DNA loops form AP-1-Bound activation hubs during macrophage development. Molecular Cell. 2017;67:1037–1048. doi: 10.1016/j.molcel.2017.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Rada-Iglesias A, Bajpai R, Swigut T, Brugmann SA, Flynn RA, Wysocka J. A unique chromatin signature uncovers early developmental enhancers in humans. Nature. 2011;470:279–283. doi: 10.1038/nature09692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Rao SS, Huntley MH, Durand NC, Stamenova EK, Bochkov ID, Robinson JT, Sanborn AL, Machol I, Omer AD, Lander ES, Aiden EL. A 3D map of the human genome at Kilobase resolution reveals principles of chromatin looping. Cell. 2014;159:1665–1680. doi: 10.1016/j.cell.2014.11.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Rapino F, Robles EF, Richter-Larrea JA, Kallin EM, Martinez-Climent JA, Graf T. C/EBPα induces highly efficient macrophage transdifferentiation of B lymphoma and leukemia cell lines and impairs their tumorigenicity. Cell Reports. 2013;3:1153–1163. doi: 10.1016/j.celrep.2013.03.003. [DOI] [PubMed] [Google Scholar]
  73. Robertson AG, Bilenky M, Tam A, Zhao Y, Zeng T, Thiessen N, Cezard T, Fejes AP, Wederell ED, Cullum R, Euskirchen G, Krzywinski M, Birol I, Snyder M, Hoodless PA, Hirst M, Marra MA, Jones SJ. Genome-wide relationship between histone H3 lysine 4 mono- and tri-methylation and transcription factor binding. Genome Research. 2008;18:1906–1917. doi: 10.1101/gr.078519.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Rosenbauer F, Tenen DG. Transcription factors in myeloid development: balancing differentiation with transformation. Nature Reviews Immunology. 2007;7:105–117. doi: 10.1038/nri2024. [DOI] [PubMed] [Google Scholar]
  75. Rubin AJ, Barajas BC, Furlan-Magaril M, Lopez-Pajares V, Mumbach MR, Howard I, Kim DS, Boxer LD, Cairns J, Spivakov M, Wingett SW, Shi M, Zhao Z, Greenleaf WJ, Kundaje A, Snyder M, Chang HY, Fraser P, Khavari PA. Lineage-specific dynamic and pre-established enhancer-promoter contacts cooperate in terminal differentiation. Nature Genetics. 2017;49:1522–1528. doi: 10.1038/ng.3935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Ruffell D, Mourkioti F, Gambardella A, Kirstetter P, Lopez RG, Rosenthal N, Nerlov C. A CREB-C/EBPbeta cascade induces M2 macrophage-specific gene expression and promotes muscle injury repair. PNAS. 2009;106:17475–17480. doi: 10.1073/pnas.0908641106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Sabari BR, Dall'Agnese A, Boija A, Klein IA, Coffey EL, Shrinivas K, Abraham BJ, Hannett NM, Zamudio AV, Manteiga JC, Li CH, Guo YE, Day DS, Schuijers J, Vasile E, Malik S, Hnisz D, Lee TI, Cisse II, Roeder RG, Sharp PA, Chakraborty AK, Young RA. Coactivator condensation at super-enhancers links phase separation and gene control. Science. 2018;361:eaar3958. doi: 10.1126/science.aar3958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Sawado T, Halow J, Bender MA, Groudine M. The beta -globin locus control region (LCR) functions primarily by enhancing the transition from transcription initiation to elongation. Genes & Development. 2003;17:1009–1018. doi: 10.1101/gad.1072303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Schwalb B, Michel M, Zacher B, Frühauf K, Demel C, Tresch A, Gagneur J, Cramer P. TT-seq maps the human transient transcriptome. Science. 2016;352:1225–1228. doi: 10.1126/science.aad9841. [DOI] [PubMed] [Google Scholar]
  80. Sherwood RI, Hashimoto T, O'Donnell CW, Lewis S, Barkal AA, van Hoff JP, Karun V, Jaakkola T, Gifford DK. Discovery of directional and nondirectional pioneer transcription factors by modeling DNase profile magnitude and shape. Nature Biotechnology. 2014;32:171–178. doi: 10.1038/nbt.2798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Shin HY, Willi M, HyunYoo K, Zeng X, Wang C, Metser G, Hennighausen L. Hierarchy within the mammary STAT5-driven wap super-enhancer. Nature Genetics. 2016;48:904–911. doi: 10.1038/ng.3606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Spicuglia S, Kumar S, Yeh JH, Vachez E, Chasson L, Gorbatch S, Cautres J, Ferrier P. Promoter activation by enhancer-dependent and -independent loading of activator and coactivator complexes. Molecular Cell. 2002;10:1479–1487. doi: 10.1016/S1097-2765(02)00791-8. [DOI] [PubMed] [Google Scholar]
  83. Spitz F, Furlong EEM. Transcription factors: from enhancer binding to developmental control. Nature Reviews Genetics. 2012;13:613–626. doi: 10.1038/nrg3207. [DOI] [PubMed] [Google Scholar]
  84. Stadhouders R, Vidal E, Serra F, Di Stefano B, Le Dily F, Quilez J, Gomez A, Collombet S, Berenguer C, Cuartero Y, Hecht J, Filion GJ, Beato M, Marti-Renom MA, Graf T. Transcription factors orchestrate dynamic interplay between genome topology and gene regulation during cell reprogramming. Nature Genetics. 2018;50:238–249. doi: 10.1038/s41588-017-0030-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Stik G, Vidal E, Barrero M, Cuartero S, Vila-Casadesús M, Mendieta-Esteban J, Tian TV, Choi J, Berenguer C, Abad A, Borsari B, le Dily F, Cramer P, Marti-Renom MA, Stadhouders R, Graf T. CTCF is dispensable for immune cell transdifferentiation but facilitates an acute inflammatory response. Nature Genetics. 2020;52:655–661. doi: 10.1038/s41588-020-0643-0. [DOI] [PubMed] [Google Scholar]
  86. Stine ZE, McGaughey DM, Bessling SL, Li S, McCallion AS. Steroid hormone modulation of RET through two estrogen responsive enhancers in breast Cancer. Human Molecular Genetics. 2011;20:3746–3756. doi: 10.1093/hmg/ddr291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Takaku M, Grimm SA, Shimbo T, Perera L, Menafra R, Stunnenberg HG, Archer TK, Machida S, Kurumizaka H, Wade PA. GATA3-dependent cellular reprogramming requires activation-domain dependent recruitment of a chromatin remodeler. Genome Biology. 2016;17:36. doi: 10.1186/s13059-016-0897-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Tan G, Lenhard B. TFBSTools: an R/bioconductor package for transcription factor binding site analysis. Bioinformatics. 2016;32:1555–1556. doi: 10.1093/bioinformatics/btw024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Thomas MD, Kremer CS, Ravichandran KS, Rajewsky K, Bender TP. c-Myb is critical for B cell development and maintenance of follicular B cells. Immunity. 2005;23:275–286. doi: 10.1016/j.immuni.2005.08.005. [DOI] [PubMed] [Google Scholar]
  90. Trikha P, Sharma N, Opavsky R, Reyes A, Pena C, Ostrowski MC, Roussel MF, Leone G. E2f1-3 are critical for myeloid development. Journal of Biological Chemistry. 2011;286:4783–4795. doi: 10.1074/jbc.M110.182733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. van Oevelen C, Collombet S, Vicent G, Hoogenkamp M, Lepoivre C, Badeaux A, Bussmann L, Sardina JL, Thieffry D, Beato M, Shi Y, Bonifer C, Graf T. C/EBPα activates Pre-existing and de novo macrophage enhancers during induced Pre-B cell transdifferentiation and myelopoiesis. Stem Cell Reports. 2015;5:232–247. doi: 10.1016/j.stemcr.2015.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Vogt L, Schmitz N, Kurrer MO, Bauer M, Hinton HI, Behnke S, Gatto D, Sebbel P, Beerli RR, Sonderegger I, Kopf M, Saudan P, Bachmann MF. VSIG4, a B7 family-related protein, is a negative regulator of T cell activation. Journal of Clinical Investigation. 2006;116:2817–2826. doi: 10.1172/JCI25673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Wamstad JA, Alexander JM, Truty RM, Shrikumar A, Li F, Eilertson KE, Ding H, Wylie JN, Pico AR, Capra JA, Erwin G, Kattman SJ, Keller GM, Srivastava D, Levine SS, Pollard KS, Holloway AK, Boyer LA, Bruneau BG. Dynamic and coordinated epigenetic regulation of developmental transitions in the cardiac lineage. Cell. 2012;151:206–220. doi: 10.1016/j.cell.2012.07.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Whyte WA, Orlando DA, Hnisz D, Abraham BJ, Lin CY, Kagey MH, Rahl PB, Lee TI, Young RA. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell. 2013;153:307–319. doi: 10.1016/j.cell.2013.03.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Wingender E, Schoeps T, Haubrock M, Krull M, Dönitz J. TFClass: expanding the classification of human transcription factors to their mammalian orthologs. Nucleic Acids Research. 2018;46:D343–D347. doi: 10.1093/nar/gkx987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Yan KK, Lou S, Gerstein M. MrTADFinder: a network modularity based approach to identify topologically associating domains in multiple resolutions. PLOS Computational Biology. 2017;13:e1005647. doi: 10.1371/journal.pcbi.1005647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Zacher B, Michel M, Schwalb B, Cramer P, Tresch A, Gagneur J. Accurate promoter and enhancer identification in 127 ENCODE and roadmap epigenomics cell types and tissues by GenoSTAN. PLOS ONE. 2017;12:e0169249. doi: 10.1371/journal.pone.0169249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Zhang Y, Liu T, Meyer CA, Eeckhoute J, Johnson DS, Bernstein BE, Nusbaum C, Myers RM, Brown M, Li W, Liu XS. Model-based analysis of ChIP-Seq (MACS) Genome Biology. 2008;9:R137. doi: 10.1186/gb-2008-9-9-r137. [DOI] [PMC free article] [PubMed] [Google Scholar]

Decision letter

Editor: Chris P Ponting1
Reviewed by: Harinder Singh2

In the interests of transparency, eLife publishes the most substantive revision requests and the accompanying author responses.

Acceptance summary:

The authors provide insight into genome control mechanisms that are used to generate context-specific patterns of gene activity that underlie the distinct functions of cells and tissues. The study's novelty lies in the application of eRNA kinetics during differentiation and the finding that synergistic enhancer/promoter sets are enriched for both lineage-determining genes and binding of cell type-specifying transcription factors.

Decision letter after peer review:

[Editors’ note: the authors submitted for reconsideration following the decision after peer review. What follows is the decision letter after the first round of review.]

Thank you for submitting your work entitled "Enhancer synergy can drive human cell fate determination" for consideration by eLife. Your article has been reviewed by three peer reviewers, and the evaluation has been overseen by a Reviewing Editor and a Senior Editor. The following individual involved in review of your submission has agreed to reveal their identity: Harinder Singh (Reviewer #2).

Our decision has been reached after extensive consultation among the reviewers. Based on these discussions and the individual reviews below, we regret to inform you that your work will not be considered further for publication in eLife.

The reviewers appreciated the study's attempt at addressing several important and timely questions regarding the developmental control of mammalian gene transcription. In particular, they were supportive of the rich data sets, the kinetic analyses and the use of TT-seq to analyze eRNAs (as a proxy of endogenous enhancer activity) and promoter output. Nevertheless, the three reviewers expressed a variety of concerns (see below) which, following extensive discussion, led to a decision to decline your submission. From reading the reviewers' comments you will see that a variety of further experiments would have been necessary to satisfy them that some conclusions are fully warranted. When coming to this decision, we and the reviewers took into consideration the likely time required to perform these experiments. Reviewers expressed an interest in seeing a new submission containing additional data that support (or otherwise) your initial manuscript's predictions, should you choose to submit this to eLife.

Reviewer #1:

The manuscript tackles an interesting question, but it completely lacks experimental verification of the hypothesis generated by the very intricate but poorly explained analysis. Without experimental validation of the prediction that some enhancer sets act synergistically while other are only additive, the major conclusions of the manuscript remain preliminary.

1) The manuscript is devoid of performance metrics for the sequencing experiments, making it impossible to gauge data quality. No comparison/reproducibility analysis of the replicates is presented, nor data on IP quality, peak numbers, read counts, IP efficiency. The super-enhancer analysis is not internally validated by assessing whether the super-enhancers identified using BRD4, C/EBPa or ATAC-seq make biological sense (are the identified super-enhancers lineage genes?), which would be important for the conclusion that synergistic enhancers are distinct from super-enhancers.

2) For validation, enhancer cooperativity could be tested in reporter assays with a handful of enhancers of either type, alone or in combination. Alternatively, Cas9-mediated knockout of increasing numbers of enhancers in a given gene/enhancer set would be predicted to have linear or non-linear effects on their target gene, depending on whether they are synergistic or additive.

3) The use of estrogen to activate the fusion protein could lead to estrogen receptor-dependent enhancer activation. The authors should clarify whether they see these effects, and if so, how they discern them from the C/EBPa-mediated effects.

4) The Discussion lacks discussion of previous findings; for example, inducible TF binding has previously been shown induce histone methylation (H3K4me1) at enhancers and nucleosome remodeling (doi:10.1186/s13059-016-0897-0, doi:10.1016/j.molcel.2010.05.004), which is corroborated by the current manuscript. For short-term transcriptional responses, eRNA transcription elongation has been shown to be required for MLL4-mediated H3K4me1/2 deposition at enhancers, while in this manuscript eRNA transcription occurs after H3K4me1 deposition – how do the authors explain this mechanistic difference?

Reviewer #2:

The authors are attempting to address several fundamental issues concerning the developmental control of mammalian gene transcription by deploying various genome wide methodologies and computationally integrating their findings. They have previously described a useful model system to analyze such regulatory processes. The experimental system involves a B cell line that undergoes transdifferentiation into a macrophage by the ectopic and inducible expression of the transcription factor C/EBPa. Using this system and performing transient transcriptome sequencing (TT-seq), to analyze eRNAs and promoter output, as well as C/EBPa ChIP-seq, H3K4me1 ChIP-seq and ATAC-seq they are able to put together a temporal sequence of molecular events for C/EBPa targeted genes. The experiments are rigorous and enhanced by the kinetic analyses.

Such temporal analysis has been performed in other systems and several of the key conclusions reached are similar to those of earlier studies. In particular that C/EBPa binds to a class of genomic sites which are accessible as well as those that are inaccessible (ATAC-seq). For the latter set, chromatin accessibility follows C/EBPa binding and H3K4me1.

The new element in this work is the genome-wide analysis of eRNAs, used as a proxy for endogenous enhancer activity, and their correlation with the activities of nearest (or within I Mb) promoters. Based on such analysis the authors conclude that for promoters that are acted on by multiple enhancers, most enhancers function additively. However for a smaller set of genes, enhancers appear to function synergistically and these genes are enriched for cell type specific functions.

A current model for super enhancer action involving liquid-liquid phase transition (molecular condensate formation with TFs, chromatin modifiers and the RNA Pol II machinery) is suggested to underlie the basis of enhancer synergy.

There are two major issues to be addressed:

1) In this study, enhancers are inferred to be connected to their cognate promoters on the basis of nearest neighbors or by using an arbitrary genomic distance (I Mb). Although such criteria are widely used they have been supplanted by Hi-C or promote capture Hi-C analysis as a means of directly analyzing enhancer-promoter interactions. Given that this connectivity is critical for the subsequent analysis and conclusions focusing on enhancer action (additive versus synergistic) on different types of target genes, the authors need to use of one of the above approaches.

2) Although not emphasized in the Abstract the authors fail to uncover evidence for super enhancers (clusters of enhancers spanning several kb) operating in their cellular system in the selective control of key differentiation genes. They suggest that their method of utilizing eRNAs rather than mediator binding revealed by ChIP-seq is better at distinguishing functionally distinct classes of enhancers. This major conclusion needs to be elaborated and deserves to be featured in the Abstract. From a mechanistic standpoint there appears to be a problem that needs to be better discussed by the authors. As noted above a current model for super enhancer action involving liquid-liquid phase transition is invoked by the authors to underlie the basis of enhancer synergy. However in their study the identified super enhancers appear to function either additively or synergistically. Can these different types of analyses be reconciled or perhaps not?

Reviewer #3:

In this manuscript, Choi and colleagues study enhancer cooperativity – i.e. how multiple enhancers collaboratively regulate the transcription of their target genes – during transdifferentiation of human B cells to macrophages at a genome-wide scale. To do so, they used an estrogen-inducible system driven by the C/EBPα pioneer factor and monitored transcriptional and chromatin dynamics over time. By using RNA-Seq, they first show that two main waves of transcriptional changes underline B cell to macrophage transdifferentiation. Then, they combined Transient Transcriptome Sequencing (TT-Seq) with newly generated ATAC-Seq and ChIP-seq for the H3K4me1 enhancer mark to identify enhancer elements that are transcribed in at least one of the timepoints, highlighting a subset of enhancers whose transcription changes upon transdifferentiation. To further disentangle the order of events underpinning transdifferentiation they performed ChIP-seq for the C/EBPα TF, whose binding is associated with chromatin opening and later transcriptional activation, further supporting the proposed role of C/EBPα as a pioneering factor.

In order to tackle enhancer cooperative behaviors, the authors used two distinct assignment strategies to pair enhancers to their target promoter(s) and report that (1) mRNA transcription of target genes correlate with the number of assigned enhancers and that (2) the sum of eRNA transcription at all assigned enhancers outperforms eRNA transcription at individual loci in predicting mRNA transcription at multiple enhancer genes. By fitting additive, synergistic and logistic models to their time-course datasets, the authors find enhancers predicted to cooperate in an additive but also synergistic fashion, meaning that mRNA output of their assigned target gene(s) is higher than expected from the sum of eRNA transcription at cognate enhancers. Finally, they show that synergistic enhancers are mostly associated with cell-type specific genes and are enriched for cell-type specific TF binding motifs, although they do not show higher C/EBPα, chromatin accessibility or BRD4 levels, as previously reported for cell-type specific "super-enhancer" hubs.

The authors address an important and timely question in the field of gene regulation, i.e. how do multiple enhancers cooperate to regulate their target genes. To achieve this goal, the authors produce a compelling and rich dataset, analyze various aspects of it and apply a modeling approach to assign enhancers to additive and synergistic modes.

Overall, this study presents a comprehensive characterization of transcription and various chromatin aspects (chromatin accessibility, H3K4me1, factor binding) for a transdifferentiation timecourse from human B lineage cells to macrophages – a true tour-de-force. We however find that the authors partly overstate their findings and draw overly strong conclusions without adequately discussing caveats. We therefore recommend that the authors carefully revise their manuscript prior to publication to address two major concerns. The two major concerns relate to 1. some important untested assumptions that underlie the key finding of the manuscript, i.e. enhancer synergy and 2. partly overly strong conclusions that for example incorrectly imply causation from correlative analyses.

1) Potential caveats to the key finding – enhancer synergy

The additive and synergistic cooperative behaviors reported by the authors heavily rely on (i) eRNAs – or TT-seq signals – accurately reflecting enhancer activity, (ii) a truly comprehensive set of enhancers, and (iii) enhancer-promoter assignment strategies. The authors need to address or discuss these points in more depth, as their key finding, i.e. the existence of enhancer synergy, crucially relies on all three assumptions being correct – any underestimation of the number of enhancers or their activities will unavoidably make the enhancers appear synergistic even if they are not. We comment on each of these points separately and then discuss some implications in more depth below.

(i) The authors treat this caveat rather superficially, referring to two recent publications (Henriques et al., 2018 and Mikhaylichenko et al., 2018) that study transcription at enhancers in Drosophila. Whether or not transcription rates truly reflect enhancer activities is far from clear: Mikhaylichenko et al. for example state "the levels and directionality of transcription are highly varied among active enhancers. […] enhancer RNA (eRNA) production and activity are not always strictly coupled", and in human HeLa and HepG2 cells, CAGE-eRNA-signals do not quantitatively correlate with enhancer activities (Andersson et al., Nature 2014; see supplementary figures 9a and 10a). In addition (more minor), at the core of TT-Seq is the transient sparse labeling of newly synthetized RNAs and their subsequent enrichment by affinity purification. We recommend discussing in more detail TT-seq's sensitivity and ability to reliably detect short, unstable and lowly abundant eRNAs (compared to more stable and longer mRNA molecules).

ii) The authors define enhancers as TT-Seq Transcription Units (TUs) and further require them to be enriched for both H3K4me1 and ATAC-Seq signals. Through these latter criteria, only 8,165 of 26,130 TUs are defined as enhancers, leaving the vast majority of about 18,000 TUs unassigned. While stringent filtering makes much sense in general, a systematic underestimation of the number of enhancers might strongly confound the analysis of enhancer cooperativity. As the H3K4me1 mark is not required for enhancer activity (Dorighi et al., 2017; Rickels et al., Nature Genetics, 2017), the authors should consider whether these unannotated TUs might also stem from enhancer elements and e.g. from gene loci that show synergistic effects.

iii) In addition to potentially missing enhancers by an overly stringent enhancer definition (point ii), the number of enhancers per gene can also be underestimated at the step of enhancer-to-gene assignment. To mitigate this problem, the authors use two complementary assignment strategies, "neighboring" and "1Mb", which make sense yet have their caveats. While the 1Mb method should be more inclusive, it for example imposes a 0.4 correlation coefficient between eRNA and mRNA changes, thus limiting the pertinence of this approach for further hypothesis testing (see major concern 2). More importantly though, the two approaches don't seem to yield very consistent results: 136 genes (92+44; subsection “Enhancer cooperation can be additive or synergistic”) were classified as synergistic by the neighboring method and 194 by the 1Mb method (subsection “Enhancer synergy is a robust phenomenon”), yet only 57 could be tested by both methods of which 33 were consistently classified as synergistic. While the authors claim that this corresponds to a strong overlap (a Fisher exact p-value is provided with 3.5e-11), it is unclear why so few genes can be tested by both methods and why of the 136 or 194 synergistically regulated genes, respectively, only 33 are consistent between methods. At the very least, these numbers suggest that the classification to the additive or synergistic models strongly depends on the enhancer-to-gene assignment.

With these caveats in mind, it is not clear whether the manuscript conclusively demonstrates the existence of synergistic enhancer cooperativity and what to recommend. A systematical under-estimation of enhancer activities or their dynamics (Figure 3B indicates that eRNAs are only detected rather late) would bias the analysis towards seemingly synergistic enhancer cooperativity, especially for genes that are highly expressed and/or strongly upregulated.

To reach strong conclusions, one would probably have to directly test combinations of enhancers or perform CRISPR-Cas9-mediated deletion of additive and synergistic enhancers with similar individual activities (as has been done in papers claiming or refuting enhancer synergy of e.g. superenhancers). One would expect that the deletion of an additive enhancer would have a limited impacted on downstream mRNA transcription and transdifferentiation efficiency, while the deletion of a synergistic enhancer would lead to stronger effects and might potentially impede transdifferentation.

However, especially given the data-richness of the manuscript, such major additional efforts seem to be beyond the scope of this project. Maybe the authors could more openly discuss the caveats listed above and how they might confound the analyses and/or perform additional sanity checks, maybe some of the following:

– simulations to assess the potential influence of underestimating enhancer activities or enhancer number on the additivity vs. synergy estimates. For example, the authors could determine how a simulated increase of eRNA transcription (e.g. by 5%, 10%, 20% or 50%) impacts the assignment of gene loci to additive, synergistic and/or logistic classes.

– a systematic analysis of potential correlations of synergistic genes with the number of enhancers. The author used gene loci containing 2 to 20 enhancers for model fitting and it would be informative whether synergistic loci overall contain more enhancers compared to additive ones.

2) Data presentation and overly strong conclusions

The authors describe and analyze their datasets by a combination of rule-based filtering and direct comparisons. This strategy is valid, yet can easily suffer from circular logic if the rules influence the comparisons, which can render results trivial. We recommend that the authors carefully revise their manuscript to remove analyses that might suffer from circular logic or discuss such caveats and tone down their conclusions.

For example, the authors require "nucleosome depletion" to define enhancers and subsequently study 4550 C/EBPa binding sites "with low or undetectable chromatin accessibility at 0h", which unavoidably leads to the observation that chromatin opens. Similarly, the "1Mb" method to assign enhancers to promoters includes a requirement that the activities need to be correlated, which means that correlations (e.g. those shown in Figure 4—figure supplement 1 E-F) cannot be seen as results.

Similarly, the authors draw strong causal conclusions from correlative analyses (e.g. "Chromatin opening enables enhancer RNA synthesis”, or "Enhancer synergy can drive cell fate determination", title), which should be avoided.

[Editors’ note: further revisions were suggested prior to acceptance, as described below.]

Thank you for submitting your article "Additive and synergistic action of mammalian enhancers during cell fate determination" for consideration by eLife. Your article has been reviewed by three peer reviewers, and the evaluation has been overseen by a Reviewing Editor and Kevin Struhl as the Senior Editor. The following individual involved in review of your submission has agreed to reveal their identity: Harinder Singh (Reviewer #1).

The reviewers have discussed the reviews with one another and the Reviewing Editor has drafted this decision to help you prepare a revised submission.

We would like to draw your attention to changes in our revision policy that we have made in response to COVID-19 (https://elifesciences.org/articles/57162). Specifically, we are asking editors to accept without delay manuscripts, like yours, that they judge can stand as eLife papers without additional data, even if they feel that they would make the manuscript stronger. Thus the revisions requested below only address clarity and presentation.

Summary:

Choi et al. analyze the transcription changes at enhancers and promoters during estrogen-induced transdifferentiation of a human CEBPa-ER-expressing B cell lymphoma cell line into a macrophage state. By integrating diverse data (C/EBPa binding pattern, H3K4me1 ChIP-seq, ATAC-seq and TT-seq) at enhancers, the authors propose that C/EBPa binding opens chromatin, increases H3K4me1 in its vicinity, and is followed by eRNA transcription. They conclude that C/EBPa functions as a pioneer factor. The manuscript studies gene transcription, enhancer activities and chromatin properties during transdifferentiation of human B cells to macrophages. While the sums of enhancer activities explain gene transcription for most genes, they do not for over 130 gene loci, suggesting that the enhancers in these loci might cooperate synergistically. The study's novelty lies in the application of eRNA kinetics during differentiation and the finding that synergistic enhancer/promoter sets are enriched for both lineage-determining genes and binding of cell type-specifying transcription factors.

The authors define temporally co-regulated pairs and cohorts of enhancers and promoters, and find 3-fold more mRNA/eRNA sets that fit an additive abundance model over those that exhibit an exponential (synergistic) relationship. Although the synergistic sets of enhancers are enriched for cell type-specific genes, similar to what has been described for super-enhancers, they appear to be distinct from super-enhancers defined previously in the same cells. Motif analysis of deeply sequenced ATAC-seq data indicates that synergistic enhancer sets are enriched of additive enhancer sets for cell type-specific transcription factors involved in lineage determination.

The authors have substantially revised their manuscript in response to previous comments. In the light that experimental verification of the predicted additive versus synergistic enhancer action was not feasible, the message of the manuscript could be strengthened further in multiple ways. First, the authors should disclose lists of (i) identified additive and synergistic enhancer cohorts and associated genes and (ii) the super-enhancers they identified, including their genomic coordinates. Similarly, the authors should disclose the full results of the GO analysis for all enhancer/super-enhancer set. Additionally, they could provide genome browser tracks and/or more screenshots of example loci. Second, the authors could verify enrichment of lineage-determining transcription factors in synergistic enhancers by ChIP-seq. Finally, the manuscript would benefit from focusing further on the novel aspects it presents and attenuating remaining claims that could be perceived as overstatements.

Revisions:

1) As the authors could not directly validate enhancer synergy, they indirectly infer its existence by the following logic: They show that the sum of enhancer activities does not match the transcription levels for some genes but leaves some gap. They use this gap to argue that the identified enhancers must be super-additive, i.e. synergistic, in order to fill this gap. While plausible, such an inference is different from the unambiguous demonstration of enhancer synergy. This fact will need to be better reflected in the Title and Abstract. Otherwise, a superficial reader might assume that synergy has been demonstrated directly. This wording could for example directly state the authors' reasoning, namely that a gap in observable enhancer activities suggests the existence of enhancer synergy for some genes. The last sentence of the revised Abstract needs to provide clarity on the comparison being made between synergistically functioning enhancer clusters and super-enhancers. It could read: "Enhancer synergy appears to be dependent on binding of cell type-specific transcription factors and such interacting enhancers are not predicted from occupancy or accessibility data that are used to detect super-enhancers."

2) The claim that enhancer synergy "could not be predicted from occupancy or accessibility data that are used to detect super-enhancers" hinges on the comparison of eRNA-derived enhancer cohorts with super-enhancers. This comparison retains two issues that still need to be addressed. First, super-enhancers represent regions of high enhancer activity, originally defined by Med1 or H3K27ac ChIP-seq. Consequently, open chromatin (ATAC-seq) or the locations of a pioneer factor (C/EBPa) that induces open and poised – but not necessarily active – chromatin is not necessarily expected to yield meaningful results in terms of defining genomic regions with cell type-specific and lineage-defining enhancer and gene activity. This would leave the BRD4 ChIP-seq data as the sole signal to define meaningful super-enhancers. However, its high peak count of >98,000 raises concerns about potential ChIP quality issues that unfortunately cannot be assessed because the authors fail to provide additional information on these ChIPs other than the peak count. That the Spearman rho values for BRD4 ChIP replicates fluctuate between 0.72 and 0.99 also do not inspire confidence. Here, a H3K27ac as a broader and more robust ChIP-seq marker of enhancer and transcription activity might help to improve super-enhancer calling, should BRD4 turn out to be sub-optimal. Second, save for a handful of super-enhancer-associated genes that have been included in the revised manuscript, the biological relevance (i.e. gene ontology term enrichment) of the identified super-enhancer-associated genes cannot be verified by readers, since the authors provide neither the corresponding list of genes nor an analysis of the super-enhancer-enriched GO terms. At a minimum, the authors should provide these gene lists for the different super-enhancer methods and additional quality information of the ChIP-seqs used to identify super-enhancers, perform GO term analyses and provide the full list of significantly enriched GO terms, to be able to gauge whether they identify true super-enhancers with the corresponding expected characteristics (cell type specificity, enrichment near lineage-determining genes and transcription factors), and whether these are indeed distinct from synergistic enhancers.

3) The analysis of transcription factors in synergistic enhancers versus additive ones is not fully developed and contains inaccuracies and confusing interpretations of the data. Specifically, there is confusion over the two motifs that were determined to be enriched in additive enhancers, for FOS and E2F family factors.

a) The motif for "FOS family" factors (TGAc/gTCA) is actually the motif to which most AP-1 family heterodimers (Fonseca, Nat Comm 2019). Indeed, AP-1 family factors (and not just FOS) play cell type-specific roles in macrophage development and function and their activity is low or undetectable at B cell enhancers by motif analysis (Heinz et al., 2010). Consequently, the statement that "additive enhancers were not [enriched for macrophage-specific TFs]" is incorrect, especially in the context of this study of B cell (-like cell) trans-differentiation into macrophage-like cells, and it is not questionable whether FOSL1 as given in Figure 6—figure supplement 1C is indeed the factor that binds these motifs, and which of the Jun-like factors that become upregulated are involved in AP-1 heterodimer binding to them.

b) There appears to be a disconnect between the interpretation of the authors' finding that E2F motifs are enriched in additive enhancers, their statement that E2F factors have regulatory functions in macrophages and the expression data in Figure 6B showing that E2F factors are rapidly downregulated upon estrogen addition, which is in line with the cell cycle arrest observed in these cells when leaving the B cell-like state (Rapino et al., 2013). Together, this indicates that E2F factors would only be able to bind to additive enhancers and play a role in their activation in the proliferative B cell-like state, but not in the macrophage-like state. If so, then why are E2F motifs enriched in late-time point enhancers?

4) The number of C/EBPa ChIP-seq peaks is surprisingly low for an overexpressed pioneer transcription factor, especially while other ChIPs in the data set have unusually large numbers of peaks (>98,000 peaks for BRD4 seems very high for a non-DNA-binding coregulator). These peak numbers are unusual and raise data quality concerns, and for C/EBPa might also explain the small number of C/EBPa-defined "super-enhancers". Further, the peak numbers identified by ATAC-seq are not provided, and here the Spearman's rho was only calculated for promoters, which is problematic in a manuscript dedicated to enhancers and raises additional data QC concerns. Providing additional QC data such as FRiP and missing peak numbers, preferentially in a tabulated form, as well as additional genome browser track screenshots at different resolutions (also of the BRD4 ChIP-seqs) are recommended.

eLife. 2021 Mar 26;10:e65381. doi: 10.7554/eLife.65381.sa2

Author response


[Editors’ note: the authors resubmitted a revised version of the paper for consideration. What follows is the authors’ response to the first round of review.]

Reviewer #1:

The manuscript tackles an interesting question, but it completely lacks experimental verification of the hypothesis generated by the very intricate but poorly explained analysis. Without experimental validation of the prediction that some enhancer sets act synergistically while other are only additive, the major conclusions of the manuscript remain preliminary.

1) The manuscript is devoid of performance metrics for the sequencing experiments, making it impossible to gauge data quality. No comparison/reproducibility analysis of the replicates is presented, nor data on IP quality, peak numbers, read counts, IP efficiency.

We have added these metrics to the Materials and methods section. The average read numbers obtained for total RNA-seq were 1x107, and 2x108 for TT-seq. Spearman's correlation coefficients between replicates were 0.99 for total RNA-seq and TT-seq on protein coding genes at all time points. The average read numbers obtained were 4x107 for C/EBPa ChIP-seq, 7x107 for H3K4me1 ChIP-seq and 2x107 for BRD4 ChIP-seq. Spearman's correlation coefficients between replicates were 0.94-0.98 for C/EBPα ChIP-seq, 0.96-1 for H3K4me1 ChIP-seq, and 0.72-0.99 for BRD4 ChIP-seq. The number of peaks were 14,561, 55,904 and 98,757 for C/EBPa, H3K4me1 and BRD4, respectively. The average read numbers obtained were 3x107 for ATAC-seq. Spearman's correlation coefficients between replicates were 0.93-0.98 for ATAC-seq on promoters.

The super-enhancer analysis is not internally validated by assessing whether the super-enhancers identified using BRD4, C/EBPa or ATAC-seq make biological sense (are the identified super-enhancers lineage genes?), which would be important for the conclusion that synergistic enhancers are distinct from super-enhancers.

Superenhancers are identified according to the common procedure as enhancers that show high signals of BRD4 ChIP-seq, C/EBPa ChIP-seq or ATAC-seq. As previously reported, some of these enhancers target cell type-specific genes, such as CD14, DDIT4, JUNB, FOSL2 etc. However, there are some superenhancers commonly found at different cell states, for example at DNMT1, JMJD8 and POLE genes. As shown in Figure 5—figure supplement 3B, some of the superenhancers identified from macrophage-like cells and their target genes are transcribed more in B-cells (0h) than macrophage-like cells (96h). Altogether, the superenhancers identified using BRD4, C/EBPa ChIP-seq and ATAC-seq include both lineage genes and common genes.

2) For validation, enhancer cooperativity could be tested in reporter assays with a handful of enhancers of either type, alone or in combination. Alternatively, Cas9-mediated knockout of increasing numbers of enhancers in a given gene/enhancer set would be predicted to have linear or non-linear effects on their target gene, depending on whether they are synergistic or additive.

We have spent one year to conduct the suggested experiments, unfortunately without success. We have now decided to submit the revised version without the suggested experiments. We have changed the text and conclusions to reflect the lack of validation by enhancer knockouts/reporter gene assays. Our attempts are summarized below as an explanation to the reviewer. We note we are not able to continue these efforts.

a) CRISPR/Cas9 enhancer knockout experiments:

We first tried to obtain stable cell lines containing Cas9-mediated knockouts of increasing numbers of enhancers but this turned out to be impossible. Sometimes heterozygous cell lines resulted, and in other cases we could not obtain the desired knockouts. We believe that the enhancers we target are critical for cell growth, preventing analysis by CRISPR knockout.

b) Reporter gene assays

Due to these difficulties, we switched to reporter gene assays as suggested by the reviewer. For the luciferase reporter assay, we selected two representative gene candidates with three additive or synergistic enhancers (FTH1 and PELI1 respectively). We cloned target gene promoters alone or in different combinations with one, two or three enhancers (5 constructs per each candidate) into luciferase gene containing plasmid. We performed transdifferentiation of BLaER cells until 96 h and collected induced macrophage-like cells (iMacs). We transfected iMacs with different constructs and measured luciferase activity after 24 h of transfection using Dual-Luciferase Reporter Assay System (Promega, E1910). In total, we tested two transfection systems: electroporation with Neon Transfection System (Thermo Fisher Scientific) and lipofection with FuGENE HD Transfection Reagent (Promega, E2311). Unfortunately, it was not possible to get reproducible results for the experimental constructs from the technical and biological replicates due to the complexity of the system. Another complication stems from the need to use (as we did) the endogenous target promoter. Please understand we have to investigate reporter gene activity under transdifferentiation conditions and to our knowledge this has never been achieved. Provided the fluctuations often observed with reporter gene assays, in retrospect it is not surprising that we could not obtain reproducible signals.

3) The use of estrogen to activate the fusion protein could lead to estrogen receptor-dependent enhancer activation. The authors should clarify whether they see these effects, and if so, how they discern them from the C/EBPa-mediated effects.

We focus our enhancer activation analysis on C/EBPa binding sites (Figure 3). Additionally, the enhancer cooperation we observe would not be a consequence of C/EBPa activation only, but rather a general behavior of enhancers that comes from the cooperative actions of TFs (as shown in Figure 6). Thus, estrogen receptor dependent enhancer activation, if any, would not change our conclusion.

In addition, a corresponding inducible mouse cell (C10) is available with the exogenous estrogen receptor fused to C/EBPa (Rapino et al., 2013). The effect of estrogen receptor activation could be tested by comparing the gene expression between an inducible cell line (C10) and the mother cell line (HAFTL) treated with b-estrogen treatment. As shown in a MA (ratio intensity) plot (GSE17316), no significant changes in gene expression were observed (adjusted p.val < 0.05). (Figure 1—figure supplement 1C). This is now mentioned in the text.

4) The Discussion lacks discussion of previous findings; for example, inducible TF binding has previously been shown induce histone methylation (H3K4me1) at enhancers and nucleosome remodeling (doi:10.1186/s13059-016-0897-0, doi:10.1016/j.molcel.2010.05.004), which is corroborated by the current manuscript.

We thank the reviewer for pointing this out and we added the missing references and included these points in discussion. It is well agreed that TF binding induces histone methylation (H3K4me1) and nucleosome remodeling at enhancers. However, whether eRNA transcription is required for H3K4me1 deposition at enhancers is still controversial. The study mentioned in the reviewer's comment suggests that inhibition of Pol II elongation reduces H3K4me2 (Kaikkonen et al., 2013).

For short-term transcriptional responses, eRNA transcription elongation has been shown to be required for MLL4-mediated H3K4me1/2 deposition at enhancers, while in this manuscript eRNA transcription occurs after H3K4me1 deposition – how do the authors explain this mechanistic difference?

We thank the reviewer for pointing this out. It has also been shown that knockout of MLL3/4 causes reduction of enhancer Pol II occupancy and eRNA synthesis (Dorighi et al., 2017). Moreover, H3K4me1 can precede nucleosomal depletion, which is necessary step prior to transcription (current study, Jubb, Boyle, Hume, and Bickmore, 2017, Creyghton et al., 2011; Rada-Iglesias et al., 2011; Zentner et al., 2011; Bogdanovic et al., 2012; Bonn et al., 2012; Mercer et al., 2011; Rada-Iglesias et al., 2012; Wamstad et al., 2012). Collectively, H3K4me1 and enhancer transcription likely affect each other. We edited the manuscript accordingly.

Reviewer #2:

The authors are attempting to address several fundamental issues concerning the developmental control of mammalian gene transcription by deploying various genome wide methodologies and computationally integrating their findings. They have previously described a useful model system to analyze such regulatory processes. The experimental system involves a B cell line that undergoes transdifferentiation into a macrophage by the ectopic and inducible expression of the transcription factor C/EBPa. Using this system and performing transient transcriptome sequencing (TT-seq), to analyze eRNAs and promoter output, as well as C/EBPa ChIP-seq, H3K4me1 ChIP-seq and ATAC-seq they are able to put together a temporal sequence of molecular events for C/EBPa targeted genes. The experiments are rigorous and enhanced by the kinetic analyses.

Such temporal analysis has been performed in other systems and several of the key conclusions reached are similar to those of earlier studies. In particular that C/EBPa binds to a class of genomic sites which are accessible as well as those that are inaccessible (ATAC-seq). For the latter set, chromatin accessibility follows C/EBPa binding and H3K4me1.

The new element in this work is the genome-wide analysis of eRNAs, used as a proxy for endogenous enhancer activity, and their correlation with the activities of nearest (or within I Mb) promoters. Based on such analysis the authors conclude that for promoters that are acted on by multiple enhancers, most enhancers function additively. However for a smaller set of genes, enhancers appear to function synergistically and these genes are enriched for cell type specific functions.

A current model for super enhancer action involving liquid-liquid phase transition (molecular condensate formation with TFs, chromatin modifiers and the RNA Pol II machinery) is suggested to underlie the basis of enhancer synergy.

There are two major issues to be addressed:

1) In this study, enhancers are inferred to be connected to their cognate promoters on the basis of nearest neighbors or by using an arbitrary genomic distance (I Mb). Although such criteria are widely used they have been supplanted by Hi-C or promote capture Hi-C analysis as a means of directly analyzing enhancer-promoter interactions. Given that this connectivity is critical for the subsequent analysis and conclusions focusing on enhancer action (additive versus synergistic) on different types of target genes, the authors need to use of one of the above approaches.

Although we used two criteria that are widely accepted, we followed the reviewer’s request and now also incorporated Hi-C data to confirm the enhancer-promoter interactions. TAD boundaries were computed from the Hi-C data and used to study whether the enhancer promoter pairs (EP pairs) are located within the same TAD. Confirming our analysis, 98.3% of the EP pairs obtained by the neighboring method were found within the same TAD, and 96.2% of the EP pairs obtained with 1Mb method were found within the same TAD at one or more time points. We also employed the third pairing method by pairing enhancers to the correlating promoters within the same TAD obtained from Hi-C data. The synergistically regulated genes found with the neighboring and TAD pairing methods significantly overlapped, supporting our conclusions on synergistic enhancer action. This information was added to the text and figures.

2) Although not emphasized in the Abstract the authors fail to uncover evidence for super enhancers (clusters of enhancers spanning several kb) operating in their cellular system in the selective control of key differentiation genes. They suggest that their method of utilizing eRNAs rather than mediator binding revealed by ChIP-seq is better at distinguishing functionally distinct classes of enhancers. This major conclusion needs to be elaborated and deserves to be featured in the Abstract. From a mechanistic standpoint there appears to be a problem that needs to be better discussed by the authors. As noted above a current model for super enhancer action involving liquid-liquid phase transition is invoked by the authors to underlie the basis of enhancer synergy. However in their study the identified super enhancers appear to function either additively or synergistically. Can these different types of analyses be reconciled or perhaps not?

We agree this is an important finding. As suggested by the reviewer, we have now mentioned this point in the Abstract and discuss it better in the manuscript, but avoiding speculative conclusions. Briefly, superenhancers are identified as enhancers with high signals of BRD4 ChIP-seq, C/EBPa ChIP-seq or ATAC-seq. As previously reported, some of these enhancers target cell type specific genes, such as CD14, DDIT4, JUNB, FOSL2 etc. However, there are some superenhancers commonly found at different cell states, for example at DNMT1, JMJD8 and POLE genes. As shown in Figure 5—figure supplement 3B, some of the superenhancers identified from macrophage-like cells and their target genes are transcribed more in B-cells (0h) than macrophage-like cells (96h). Altogether, the superenhancers identified using BRD4, C/EBPa ChIP-seq and ATAC-seq include both lineage genes and common genes. It has been shown that not all super enhancers act synergistically (see Introduction). Meanwhile, the synergistic enhancers identified with RNA synthesis clearly showed higher frequency of long-range EP interactions than additive enhancers (see above). A recent report also showed that BRD4 inhibition does not disrupt EP interactions (biorxiv: doi: https://doi.org/10.1101/848325), and this was included in the Discussion.

Reviewer #3:

In this manuscript, Choi and colleagues study enhancer cooperativity – i.e. how multiple enhancers collaboratively regulate the transcription of their target genes – during transdifferentiation of human B cells to macrophages at a genome-wide scale. To do so, they used an estrogen-inducible system driven by the C/EBPα pioneer factor and monitored transcriptional and chromatin dynamics over time. By using RNA-Seq, they first show that two main waves of transcriptional changes underline B cell to macrophage transdifferentiation. Then, they combined Transient Transcriptome Sequencing (TT-Seq) with newly generated ATAC-Seq and ChIP-seq for the H3K4me1 enhancer mark to identify enhancer elements that are transcribed in at least one of the timepoints, highlighting a subset of enhancers whose transcription changes upon transdifferentiation. To further disentangle the order of events underpinning transdifferentiation they performed ChIP-seq for the C/EBPα TF, whose binding is associated with chromatin opening and later transcriptional activation, further supporting the proposed role of C/EBPα as a pioneering factor.

In order to tackle enhancer cooperative behaviors, the authors used two distinct assignment strategies to pair enhancers to their target promoter(s) and report that (1) mRNA transcription of target genes correlate with the number of assigned enhancers and that (2) the sum of eRNA transcription at all assigned enhancers outperforms eRNA transcription at individual loci in predicting mRNA transcription at multiple enhancer genes. By fitting additive, synergistic and logistic models to their time-course datasets, the authors find enhancers predicted to cooperate in an additive but also synergistic fashion, meaning that mRNA output of their assigned target gene(s) is higher than expected from the sum of eRNA transcription at cognate enhancers. Finally, they show that synergistic enhancers are mostly associated with cell-type specific genes and are enriched for cell-type specific TF binding motifs, although they do not show higher C/EBPα, chromatin accessibility or BRD4 levels, as previously reported for cell-type specific "super-enhancer" hubs.

The authors address an important and timely question in the field of gene regulation, i.e. how do multiple enhancers cooperate to regulate their target genes. To achieve this goal, the authors produce a compelling and rich dataset, analyze various aspects of it and apply a modeling approach to assign enhancers to additive and synergistic modes.

Overall, this study presents a comprehensive characterization of transcription and various chromatin aspects (chromatin accessibility, H3K4me1, factor binding) for a transdifferentiation timecourse from human B lineage cells to macrophages – a true tour-de-force. We however find that the authors partly overstate their findings and draw overly strong conclusions without adequately discussing caveats. We therefore recommend that the authors carefully revise their manuscript prior to publication to address two major concerns. The two major concerns relate to 1. some important untested assumptions that underlie the key finding of the manuscript, i.e. enhancer synergy and 2. partly overly strong conclusions that for example incorrectly imply causation from correlative analyses.

We thank the reviewer for the careful analysis and kind words. We carefully revised the manuscript to avoid overstating our findings and we now adequately discuss potential caveats.

1) Potential caveats to the key finding – enhancer synergy

The additive and synergistic cooperative behaviors reported by the authors heavily rely on (i) eRNAs – or TT-seq signals – accurately reflecting enhancer activity, (ii) a truly comprehensive set of enhancers, and iii) enhancer-promoter assignment strategies. The authors need to address or discuss these points in more depth, as their key finding, i.e. the existence of enhancer synergy, crucially relies on all three assumptions being correct – any underestimation of the number of enhancers or their activities will unavoidably make the enhancers appear synergistic even if they are not. We comment on each of these points separately and then discuss some implications in more depth below.

(i) The authors treat this caveat rather superficially, referring to two recent publications (Henriques et al., 2018 and Mikhaylichenko et al., 2018) that study transcription at enhancers in Drosophila. Whether or not transcription rates truly reflect enhancer activities is far from clear: Mikhaylichenko et al. for example state "the levels and directionality of transcription are highly varied among active enhancers. […] enhancer RNA (eRNA) production and activity are not always strictly coupled", and in human HeLa and HepG2 cells, CAGE-eRNA-signals do not quantitatively correlate with enhancer activities (Andersson et al., Nature 2014; see supplementary figures 9a and 10a). In addition (more minor), at the core of TT-Seq is the transient sparse labeling of newly synthetized RNAs and their subsequent enrichment by affinity purification. We recommend discussing in more detail TT-seq's sensitivity and ability to reliably detect short, unstable and lowly abundant eRNAs (compared to more stable and longer mRNA molecules).

As recommend, we are now discussing in more detail TT-seq's sensitivity and ability to reliably detect short, unstable and lowly abundant eRNAs (compared to more stable and longer mRNA molecules. We also point out better the advantages of our approach.

Briefly, the generally used method to measure enhancer activity is to assess gene expression as a readout upon alterations of enhancers, for instance reporter gene assays or CRISPR editing of enhancers. Yet, the limitation of these methods is that we cannot easily quantify the enhancer activities and we cannot do the analysis genome-wide. We therefore used TT-seq to get genome-wide quantitative readouts for enhancer transcription, which is a good proxy for enhancer activity. TT-seq is highly sensitive to detect newly synthesized RNAs including unstable short RNAs such as eRNAs without perturbation (Schwalb et al., 2016). During T-cell stimulation, transcription from enhancers and promoters of responsive genes is activated simultaneously (Michel et al., 2017). Enhancers can be paired with their putative target gene promoters based on their proximity (Michel et al., 2017). In addition, we have shown that eRNA synthesis measured with TT-seq highly correlated with the putative target gene synthesis (Michel et al., 2017), implicating that enhancer transcription activity can indeed reflect enhancer activities. As eRNA production is a very good proxy for enhancer transactivation activity (Henriques et al., 2018; Mikhaylichenko et al., 2018), TTseq can be used to identify active enhancers, to pair enhancers with their putative target promoters, and to measure the transcription activity of enhancers and promoters genome wide.

ii) The authors define enhancers as TT-Seq Transcription Units (TUs) and further require them to be enriched for both H3K4me1 and ATAC-Seq signals. Through these latter criteria, only 8,165 of 26,130 TUs are defined as enhancers, leaving the vast majority of about 18,000 TUs unassigned. While stringent filtering makes much sense in general, a systematic underestimation of the number of enhancers might strongly confound the analysis of enhancer cooperativity. As the H3K4me1 mark is not required for enhancer activity (Dorighi et al., Mol Cell 2017; Rickels et al., Nature Genetics, 2017), the authors should consider whether these unannotated TUs might also stem from enhancer elements and e.g. from gene loci that show synergistic effects.

As suggested, we investigated whether unannotated TUs might also stem from enhancer elements. Among non-coding TUs, 13,952 TUs overlap with H3K4me1 signals, whereas 8,529 TUs overlap with ATAC-seq signals. The restriction rather comes from ATAC-seq, which is required for eRNA synthesis (Figure 2—figure supplement 1A). However, to exclude the possibility that the unassigned TUs are left out due to the low sensitivity of chromatin accessibility detection etc, we have paired these unannotated TUs with the nearest genes, and repeated the analysis to identify synergistic enhancers. After excluding ncRNAs that are downstream RNAs, convergent RNAs, upstream antisense RNAs or that overlap with more than 20% of transcripts annotated in Gencode, 16360 TUs (or 15487 extended EUs after merging TUs within 1kb) remained for further analysis (see Materials and methods for details). Using the neighboring pairing method, 1184 of 1790 TUs were paired with more enhancers than previous analysis. In total, we observed 544 and 209 genes were regulated by additive and synergistic enhancers, respectively. Of 136 synergistically regulated genes that have been previously identified, 71% (97 genes) were regulated in synergistic manner (Figure 5—figure supplement 1H). This indicates that the widespread synergistic behavior we observed was not due to an underestimation of the number of enhancers. We edited the text accordingly and this clarifies the reviewer’s concern.

iii) In addition to potentially missing enhancers by an overly stringent enhancer definition (point ii), the number of enhancers per gene can also be underestimated at the step of enhancer-to-gene assignment. To mitigate this problem, the authors use two complementary assignment strategies, "neighboring" and "1Mb", which make sense yet have their caveats. While the 1Mb method should be more inclusive, it for example imposes a 0.4 correlation coefficient between eRNA and mRNA changes, thus limiting the pertinence of this approach for further hypothesis testing (see major concern 2). More importantly though, the two approaches don't seem to yield very consistent results: 136 genes (92+44; subsection “Enhancer cooperation can be additive or synergistic”) were classified as synergistic by the neighboring method and 194 by the 1Mb method (subsection “Enhancer synergy is a robust phenomenon”), yet only 57 could be tested by both methods of which 33 were consistently classified as synergistic. While the authors claim that this corresponds to a strong overlap (a Fisher exact p-value is provided with 3.5e-11), it is unclear why so few genes can be tested by both methods and why of the 136 or 194 synergistically regulated genes, respectively, only 33 are consistent between methods. At the very least, these numbers suggest that the classification to the additive or synergistic models strongly depends on the enhancer-to-gene assignment.

The reviewer is correct and the relatively low number of consistent E-P pairs depends on the type of assignment. Briefly, to test enhancer synergy rigorously, we can only use genes that are paired with more than two enhancers. Since the number of enhancers paired to each gene vary between the neighboring and 1Mb methods, only a small number of genes could be tested for enhancer synergy. Nevertheless, our results are consistent between two methods. As in the response to the reviewer #2, we also employed a third pairing method by pairing enhancers to the correlating promoters within the same TAD obtained from Hi-C data. The synergistically regulated genes found with the neighboring and TAD pairing methods significantly overlapped, supporting our conclusions on synergistic enhancer action.This information was added to the text and figures.

With these caveats in mind, it is not clear whether the manuscript conclusively demonstrates the existence of synergistic enhancer cooperativity and what to recommend. A systematical under-estimation of enhancer activities or their dynamics (Figure 3B indicates that eRNAs are only detected rather late) would bias the analysis towards seemingly synergistic enhancer cooperativity, especially for genes that are highly expressed and/or strongly upregulated.

Although it is true that Figure 3B shows eRNAs are strongly synthesized at 96h in the LATE Cluster, eRNA synthesis is also observed at earlier time points in other clusters. In the originally submitted manuscript we had already excluded that there is a bias towards highly expressed genes by showing that synergistically regulated genes do not have systematically higher expression levels (Figure 5—figure supplement 2A), and that not all LATE upregulated genes are synergistically regulated. In summary, with all these controls, we conclusively demonstrate enhancer synergy at selected loci.

To reach strong conclusions, one would probably have to directly test combinations of enhancers or perform CRISPR-Cas9-mediated deletion of additive and synergistic enhancers with similar individual activities (as has been done in papers claiming or refuting enhancer synergy of e.g. superenhancers). One would expect that the deletion of an additive enhancer would have a limited impacted on downstream mRNA transcription and transdifferentiation efficiency, while the deletion of a synergistic enhancer would lead to stronger effects and might potentially impede transdifferentiation.

Please compare our answers to reviewer #1. We spent one year trying both the enhancer knockout approach and reporter gene assays, without success, due to the complex nature of the system, which combines transdifferentiation and enhancer cooperativity. We now resubmit a revised version that is substantially edited and hope the reviewer is fine with publication of this strongly revised version.

However, especially given the data-richness of the manuscript, such major additional efforts seem to be beyond the scope of this project. Maybe the authors could more openly discuss the caveats listed above and how they might confound the analyses and/or perform additional sanity checks, maybe some of the following:

– simulations to assess the potential influence of underestimating enhancer activities or enhancer number on the additivity vs. synergy estimates. For example, the authors could determine how a simulated increase of eRNA transcription (e.g. by 5%, 10%, 20% or 50%) impacts the assignment of gene loci to additive, synergistic and/or logistic classes.

We thank the reviewer for stating that major additional efforts are beyond the scope of this project. We have carried out the suggested test. In the original manuscript we had included more enhancers around a few genes and showed that this did not change conclusions with respect to enhancer synergy. We have now done such testing more systematically and added a statement to the text. Briefly, we simulated increase of eRNA transcription and determined synergistic genes. Synergistically regulated genes were hardly affected by the increase of eRNA transcription (Figure 5—figure supplement 1G).

– a systematic analysis of potential correlations of synergistic genes with the number of enhancers. The author used gene loci containing 2 to 20 enhancers for model fitting and it would be informative whether synergistic loci overall contain more enhancers compared to additive ones.

We carried out this test and found that there is no difference in enhancer numbers between additive and synergistic loci (median 3.5 vs 3 enhancers, respectively) (ECDF plot). We added this to the manuscript (Figure 5—figure supplement 1H).

2) Data presentation and overly strong conclusions

The authors describe and analyze their datasets by a combination of rule-based filtering and direct comparisons. This strategy is valid, yet can easily suffer from circular logic if the rules influence the comparisons, which can render results trivial. We recommend that the authors carefully revise their manuscript to remove analyses that might suffer from circular logic or discuss such caveats and tone down their conclusions.

We understand the concern and went through the manuscript and made sure there are not analyses that might suffer from circular logic.

For example, the authors require "nucleosome depletion" to define enhancers and subsequently study 4550 C/EBPa binding sites "with low or undetectable chromatin accessibility at 0h", which unavoidably leads to the observation that chromatin opens.

This may be a misunderstanding. We define enhancers based on enhancer transcription, which can only occur after nucleosome depletion because the Pol II machinery needs to access DNA. Many of these enhancers showed changes in activities (Figure 2C). Our question was how enhancers are newly activated from closed chromatin to active transcription. The basic premise here is that enhancers undergo chromatin opening. Then the question is on the temporal relationship of transcription factor binding, H3K4monomethylation, chromatin accessibility and transcription activities at the sites undergoing chromatin opening. In fact, this cannot be observed if chromatin does not open.

Similarly, the "1Mb" method to assign enhancers to promoters includes a requirement that the activities need to be correlated, which means that correlations (e.g. those shown in Figure 4—figure supplement 1 E-F) cannot be seen as results.

We agree and thank the reviewer for spotting this. We removed these figure panels and the corresponding text.

Similarly, the authors draw strong causal conclusions from correlative analyses (e.g. "Chromatin opening enables enhancer RNA synthesis”, or "Enhancer synergy can drive cell fate determination", title), which should be avoided.

We changed the title and trust the reviewer agrees to it. We edited the text to make sure there are no overly strong conclusions.

[Editors’ note: what follows is the authors’ response to the second round of review.]

Revisions:

1) As the authors could not directly validate enhancer synergy, they indirectly infer its existence by the following logic: They show that the sum of enhancer activities does not match the transcription levels for some genes but leaves some gap. They use this gap to argue that the identified enhancers must be super-additive, i.e. synergistic, in order to fill this gap. While plausible, such an inference is different from the unambiguous demonstration of enhancer synergy. This fact will need to be better reflected in the Title and Abstract. Otherwise, a superficial reader might assume that synergy has been demonstrated directly. This wording could for example directly state the authors' reasoning, namely that a gap in observable enhancer activities suggests the existence of enhancer synergy for some genes. The last sentence of the revised Abstract needs to provide clarity on the comparison being made between synergistically functioning enhancer clusters and super-enhancers. It could read: "Enhancer synergy appears to be dependent on binding of cell type-specific transcription factors and such interacting enhancers are not predicted from occupancy or accessibility data that are used to detect super-enhancers."

We appreciate the comment and changed the Abstract as suggested. We also changed the title to: “Evidence for additive and synergistic action of mammalian enhancers during cell fate determination”.

2) The claim that enhancer synergy "could not be predicted from occupancy or accessibility data that are used to detect super-enhancers" hinges on the comparison of eRNA-derived enhancer cohorts with super-enhancers. This comparison retains two issues that still need to be addressed. First, super-enhancers represent regions of high enhancer activity, originally defined by Med1 or H3K27ac ChIP-seq. Consequently, open chromatin (ATAC-seq) or the locations of a pioneer factor (C/EBPa) that induces open and poised – but not necessarily active – chromatin is not necessarily expected to yield meaningful results in terms of defining genomic regions with cell type-specific and lineage-defining enhancer and gene activity. This would leave the BRD4 ChIP-seq data as the sole signal to define meaningful super-enhancers. However, its high peak count of >98,000 raises concerns about potential ChIP quality issues that unfortunately cannot be assessed because the authors fail to provide additional information on these ChIPs other than the peak count. That the Spearman rho values for BRD4 ChIP replicates fluctuate between 0.72 and 0.99 also do not inspire confidence. Here, a H3K27ac as a broader and more robust ChIP-seq marker of enhancer and transcription activity might help to improve super-enhancer calling, should BRD4 turn out to be sub-optimal. Second, save for a handful of super-enhancer-associated genes that have been included in the revised manuscript, the biological relevance (i.e. gene ontology term enrichment) of the identified super-enhancer-associated genes cannot be verified by readers, since the authors provide neither the corresponding list of genes nor an analysis of the super-enhancer-enriched GO terms. At a minimum, the authors should provide these gene lists for the different super-enhancer methods and additional quality information of the ChIP-seqs used to identify super-enhancers, perform GO term analyses and provide the full list of significantly enriched GO terms, to be able to gauge whether they identify true super-enhancers with the corresponding expected characteristics (cell type specificity, enrichment near lineage-determining genes and transcription factors), and whether these are indeed distinct from synergistic enhancers.

To provide more support that classically defined superenhancers do not correspond to the synergistic enhancers, we used H3K27Ac ChIP-seq data (Stik et al., 2020) as suggested by the reviewers. Consistent with our previous results, the superenhancers derived from H3K27Ac ChIP-seq signal included both synergistic and additive enhancers (Figure 5F, Figure 5—figure supplement 4A). Similar to superenhancers derived from other markers for high enhancer/transcription activity, RNA synthesis from the superenhancers which show high H3K27Ac levels at 96h and their paired promoters was not always the highest at 96h compared to other time points (Figure 5—figure supplement 4C). The H3K27Ac level at synergistic enhancers was similar to the level at additive enhancers (Figure 5G).

Furthermore, we included a list of genes regulated by superenhancers derived from various ChIP-seq data in Supplementary file 1. To ensure these superenhancers correspond to the previously reported superenhancers, we compared the list to dbSUPER (Khan and Zhang, 2016). Most of the superenhancers identified in this study could be found in the database (Supplementary file 1). Additionally, GO analysis was performed using superenhancers. GO terms with p values less than 0.05 were included for superenhancers derived from eRNA signal, ATAC-seq signal, BRD4 and C/EBPa ChIP-seq signal in Supplementary files 2, 4 and 5. GO terms with Bonferroni p value less than 0.05 were included for superenhancers derived from H3K27Ac ChIP-seq signal in Supplementary file 3. The resulting GO terms included immune response, inflammatory response and defense response. Finally, FRiP of ChIP-seqs can be found in Supplementary file 7. Additional genome browser screenshots are included in Figure 4—figure supplement 1E and Figure 5—figure supplement 4B.

3) The analysis of transcription factors in synergistic enhancers versus additive ones is not fully developed and contains inaccuracies and confusing interpretations of the data. Specifically, there is confusion over the two motifs that were determined to be enriched in additive enhancers, for FOS and E2F family factors.

a) The motif for "FOS family" factors (TGAc/gTCA) is actually the motif to which most AP-1 family heterodimers (Fonseca, Nat Comm 2019). Indeed, AP-1 family factors (and not just FOS) play cell type-specific roles in macrophage development and function and their activity is low or undetectable at B cell enhancers by motif analysis (Heinz et al., Mol Cell 2010). Consequently, the statement that "additive enhancers were not [enriched for macrophage-specific TFs]" is incorrect, especially in the context of this study of B cell (-like cell) trans-differentiation into macrophage-like cells, and it is not questionable whether FOSL1 as given in Figure 6—figure supplement 1C is indeed the factor that binds these motifs, and which of the Jun-like factors that become upregulated are involved in AP-1 heterodimer binding to them.

The reviewers are right that the motif for FOS family factors is bound by most AP-1 family heterodimers, which have cell type specific roles in macrophage development. We had mentioned this in the originally submitted manuscript. However, this motif was enriched in additive enhancers only at 12h when the cells did not establish macrophage specific program (Figure 6—figure supplement 1B). At later time points in macrophage-like cells, these motifs were found both in additive and synergistic enhancers as no enrichment was shown in Figure 6—figure supplement 1B. We interpret this to mean that the TFs in FOS family and AP-1 family bind both in additive and synergistic enhancers. We improved the text to avoid any misunderstanding.

b) There appears to be a disconnect between the interpretation of the authors' finding that E2F motifs are enriched in additive enhancers, their statement that E2F factors have regulatory functions in macrophages and the expression data in Figure 6B showing that E2F factors are rapidly downregulated upon estrogen addition, which is in line with the cell cycle arrest observed in these cells when leaving the B cell-like state (Rapino, 2013). Together, this indicates that E2F factors would only be able to bind to additive enhancers and play a role in their activation in the proliferative B cell-like state, but not in the macrophage-like state. If so, then why are E2F motifs enriched in late-time point enhancers?

E2F factors have general regulatory functions in many cell types including macrophages. Indeed, E2F1, a member of E2F factors, is rapidly downregulated upon estrogen addition. Yet, E2F3, E2F4 and E2F6, other members of E2F factors, maintains their expression level during transdifferentiation. Thus, it is possible that E2F motifs were enriched in late-time point enhancers due to the E2F factors without significant expression changes in macrophage-like cells such as E2F3, E2F4 and E2F6.

4) The number of C/EBPa ChIP-seq peaks is surprisingly low for an overexpressed pioneer transcription factor, especially while other ChIPs in the data set have unusually large numbers of peaks (>98,000 peaks for BRD4 seems very high for a non-DNA-binding coregulator). These peak numbers are unusual and raise data quality concerns, and for C/EBPa might also explain the small number of C/EBPa-defined "super-enhancers". Further, the peak numbers identified by ATAC-seq are not provided, and here the Spearman's rho was only calculated for promoters, which is problematic in a manuscript dedicated to enhancers and raises additional data QC concerns. Providing additional QC data such as FRiP and missing peak numbers, preferentially in a tabulated form, as well as additional genome browser track screenshots at different resolutions (also of the BRD4 ChIP-seqs) are recommended.

As additional QC data, peak numbers for ATAC-seq, Spearman’s rho for all ATAC-seq peaks and FRiP of ChIP-seq data are provided in Supplementary file 7. Genome browser screenshots of the datasets including BRD4 and H3K27Ac around CD14 are shown in Figure 4D and Figure 4—figure supplement 1). Genome browser screenshots around IRF8 and EBF1 are also shown in Figure 5—figure supplement 4B (Also mentioned in the additional revision point).

Associated Data

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

    Data Citations

    1. Choi J, Lysakovskaia K, Stik G, Demel C, Soeding J, Tian TV, Graf T, Cramer P. 2020. Evidence for additive and synergistic action of mammalian enhancers during cell fate determination. NCBI Gene Expression Omnibus. GSE131620 [DOI] [PMC free article] [PubMed]
    2. Stik G, Casadesus MV, Graf T. 2020. CTCF is dispensable for cell fate conversion but facilitates acute cellular responses [Hi-C] NCBI Gene Expression Omnibus. GSE141226
    3. UNC Chapel Hill. 2017. in situ Hi-C data of THP-1 cells untreated and treated with PMA. NCBI BioProject. PRJNA385337

    Supplementary Materials

    Supplementary file 1. List of superenhancers identified from chromatin accessibility, BRD4, C/EBPα, and H3K27Ac occupancies.
    elife-65381-supp1.xlsx (48.4KB, xlsx)
    Supplementary file 2. GO analysis on superenhancers identified from BRD4 signal.
    elife-65381-supp2.xlsx (86.3KB, xlsx)
    Supplementary file 3. GO analysis on superenhancers identified from H3K27Ac ChIP-seq signal.
    elife-65381-supp3.xlsx (99.2KB, xlsx)
    Supplementary file 4. GO analysis on superenhancers identified from C/EBPα signal.
    elife-65381-supp4.xlsx (122.6KB, xlsx)
    Supplementary file 5. GO analysis on superenhancers identified from chromatin accessibility signal.
    elife-65381-supp5.xlsx (63.9KB, xlsx)
    Supplementary file 6. Primer list.
    elife-65381-supp6.docx (12.9KB, docx)
    Supplementary file 7. Quality assessment of ChIP seq data sets.
    elife-65381-supp7.xlsx (15.6KB, xlsx)
    Transparent reporting form

    Data Availability Statement

    RNA-seq, TT-seq, ChIP-seq, ATAC-seq data reported in this study were deposited with the National Center for Biotechnology Information Gene Expression Omnibus (accession number GSE131620). Hi-C data and H3K27Ac ChIP-seq in BLaER and Hi-C data in THP-1 cell lines that support the findings of this study are available with the National Center for Biotechnology Information Gene Expression Omnibus (accession GSE141226) and BioProject (accession PRJNA385337).

    The following dataset was generated:

    Choi J, Lysakovskaia K, Stik G, Demel C, Soeding J, Tian TV, Graf T, Cramer P. 2020. Evidence for additive and synergistic action of mammalian enhancers during cell fate determination. NCBI Gene Expression Omnibus. GSE131620

    The following previously published datasets were used:

    Stik G, Casadesus MV, Graf T. 2020. CTCF is dispensable for cell fate conversion but facilitates acute cellular responses [Hi-C] NCBI Gene Expression Omnibus. GSE141226

    UNC Chapel Hill. 2017. in situ Hi-C data of THP-1 cells untreated and treated with PMA. NCBI BioProject. PRJNA385337


    Articles from eLife are provided here courtesy of eLife Sciences Publications, Ltd

    RESOURCES