Skip to main content

A simple metric of promoter architecture robustly predicts expression breadth of human genes suggesting that most transcription factors are positive regulators

Abstract

Background

Conventional wisdom holds that, owing to the dominance of features such as chromatin level control, the expression of a gene cannot be readily predicted from knowledge of promoter architecture. This is reflected, for example, in a weak or absent correlation between promoter divergence and expression divergence between paralogs. However, an inability to predict may reflect an inability to accurately measure or employment of the wrong parameters. Here we address this issue through integration of two exceptional resources: ENCODE data on transcription factor binding and the FANTOM5 high-resolution expression atlas.

Results

Consistent with the notion that in eukaryotes most transcription factors are activating, the number of transcription factors binding a promoter is a strong predictor of expression breadth. In addition, evolutionarily young duplicates have fewer transcription factor binders and narrower expression. Nonetheless, we find several binders and cooperative sets that are disproportionately associated with broad expression, indicating that models more complex than simple correlations should hold more predictive power. Indeed, a machine learning approach improves fit to the data compared with a simple correlation. Machine learning could at best moderately predict tissue of expression of tissue specific genes.

Conclusions

We find robust evidence that some expression parameters and paralog expression divergence are strongly predictable with knowledge of transcription factor binding repertoire. While some cooperative complexes can be identified, consistent with the notion that most eukaryotic transcription factors are activating, a simple predictor, the number of binding transcription factors found on a promoter, is a robust predictor of expression breadth.

Background

Is it possible to predict expression parameters of a gene from knowledge of the promoter architecture of that gene? If, for example, we knew the transcription factors (TF) that bind the promoter of a gene, can we predict the breadth of expression (BoE, i.e. the proportion of tissues/cells within which the gene is expressed) or the mean level of expression of that gene? It is known that expression patterns of gene duplicates diverge over evolutionary time [[1],[2]], but can we predict how different the expression of paralogs will be knowing nothing more than their promoter architecture? What in turn is the relationship between expression breadth and the number of TFs regulating a gene (TfbsNo.)? Given that, in contrast to prokaryotes, the ground state for most eukaryotic genes is inactivity [[3]], we might expect that broadly expressed genes should have very many regulating TFs, assuming eukaryotic TFs are for the most part activating [[4]]. However, some very broadly expressed genes might have reverted to a more prokaryotic state and have activity as the constitutive state and hence not require TF activation. Alternatively, the breadth of expression may be conferred by the ability to bind a few specialist transcription factors or through cooperation of particular TFs, in which case the total number of binders need not predict breadth.

At first sight the answer to many of these questions may appear rather trivial: surely if we know the TFs that bind a gene’s promoter and know when those TFs are present in cells then we must know the expression parameters of a gene [[5]]? However, an in-depth study of STE12 found that expression changes in response to this transcription factor accounted for only half the observed expression fluctuations [[6]]. That the coupling between TF presence/absence need not be such an excellent predictor is indicative of other levels of control. In addition to transcription level regulation (presence/absence of the relevant TFs), genes can be regulated both pre- and post-transcriptionally. Post-transcriptionally, processes such as nonsense-mediated decay (NMD) [[7]], microRNA level regulation [[8]] and modulation of RNA stability [[9]], can also act to reduce the transcript levels below that expected given the transcription rate, potentially buffering larger changes in mRNA levels. Chromatin level pre-transcriptional regulation may be the dominant factor [[10]]. This can mean either higher-level chromatin architecture (open/closed chromatin configuration) [[10]] or other epigenetic marks (histone modification, methylation etc.) [[11],[12]], all of which can modulate the expression of the gene even if the relevant TFs are present.

Much evidence supports a strong role for chromatin in dictating expression profiles. For example, insertion of the same transgene into different regions in the genome leads to different expression levels dependent on the expression profile of the neighbouring genes [[13]]. Similarly, a pair of transgenes can be co-expressed if introduced in tandem (so sharing the same chromatin environment) but have uncoordinated expression when introduced into unlinked locations [[14]]. Upregulation of one gene is similarly thought to cause a time-lagged ripple of chromatin opening which leads to spikes in the expression of neighbours [[15]]. More generally, at least in yeast, physical proximity of genes, is a strong predictor of the degree of co-expression between any two genes [[16]]. Indeed, for unlinked genes, on average two genes with the identical repertoire of TF binders, have only a weak degree of co-expression (r2 ~ 1-2%), much less than the degree of co-expression of two linked genes with no transcription factors in common (r2 ~ 10%) [[16]]. Moreover, DNA methylation was found to increase or decrease BoE depending on the target sequence [[17]]; while CpG islands colocalize with most promoters and are characterized by low methylation [[18]]. These results all suggest that chromatin level effects are not negligible and that extrapolation from TF binding to expression profile might be a relatively futile enterprise. In contrast to this position, however, is a striking counter-example demonstrating that the expression profile of genes involved in Drosophila segmentation is well predicted by the knowledge of TF binding sites and TF levels [[5]].

One approach to determine the extent to which promoter architecture determines expression parameters has been to consider the relationship between expression divergence and promoter divergence between paralogs within a genome or between orthologs in different genomes [[19]–[22]]. The logic is the same in both instances, namely that if the differences from the ancestral expression profile to current expression profile have been owing to changes in the sequence of the promoters, then comparing multiple genes across genomes (for orthologs) or within genomes (for paralogs) should reveal correlations between the degree of promoter divergence and the degree of expression divergence. In the instance of paralogs there is an additional assumption that the duplicate versions of the same gene were generated in a manner that preserved the promoters. These analyses commonly suggest little or no coupling between promoter divergence and expression divergence, consistent with a weak coupling between promoter architecture and gene expression parameters. For example, within yeasts divergence of transcription factor binding sites (Tfbs) has little impact on expression divergence between orthologs [[19]]. Similarly, Park and Makova found in humans that the correspondence of paralog cis-regulatory regions was so weakly correlated with expression divergence in a multiple regression that it wasn’t significant after multi-test correction [[20]]. A further yeast study found that promoter divergence explained only 2-3% of expression variability [[21]]. These results suggest that cis-regulatory effects are not a major influence on expression profile. By contrast, a promoter screen in yeast found evidence for a robust correlation between the number of shared motifs and the degree of expression divergence between paralogs [[22]], although, unexpectedly, the absolute number of motifs the paralogs have is approximately constant over time. Clearly, more analysis is needed to investigate this key question in the field of expression pattern evolution.

While the consensus view is that promoter architecture doesn’t well predict expression parameters, there is also then a lack of perfect agreement on this. One possible reason the studies are not obviously in agreement is that there is much noise in both measures of expression and inference of which proteins bind any given gene’s promoter. In addition, it isn’t immediately clear what metric of, for example, promoter divergence would be most informative. We return to this issue employing a merge of two exceptional data sources, ENCODE and FANTOM5. We used ENCODE ChIP-seq meta data set derived from multi cell-line clustered experiments published in 2012 [[23]]. Whole-genome studies of regulatory evolution in human had been unfeasible before ENCODE [[23]]. Although ENCODE experiments were performed on separate cell lines, standardized experimental protocols and a unified analytical pipeline [[24]] allows one to merge ENCODE data into one meta-dataset [[25],[26]]. FANTOM5 is the most comprehensive expression dataset available, including 952 human and 396 mouse tissues, primary cells and cancer cell-lines (see Table 1). FANTOM5 [[27]] is based on cap analysis of gene expression (CAGE). CAGE characterizes transcriptional start sites across the entire genome in an unbiased fashion, and at a single-base resolution level [[27]].

Table 1 The numbers of samples in distinct FANTOM5 categories

Here then we employ this novel data to ask whether expression profiles can be predicted from promoter architecture. In the first instance we wish to know whether the total number of transcription factors binding a promoter is a good predictor. We follow this up with the analysis of interactants and a more complex machine learning approach. We start by resolving basic parameters of TF binding and promoter architecture.

Results

The number of transcription factors per gene follows a power law

Before attempting to describe any correlations between the number of Tfbs (TfbsNo.) and expression, it is instructive to know what the distribution of the number of transcription factors per gene looks like. Perhaps it is normally distributed? To determine this, proximal promoters were defined by a symmetrical window around the transcription start site―TSS (+ − 500 bps). The distribution of TfbsNo. is not normal, instead it follows a power law (Figure 1). At the Tfbs quality cutoff of 500, ninety percent of genes had between 0 and 26 transcription factor bindings sites, but there was a long-tail of genes with high values (more than 26). The distribution can be defined by Tukey’s five numbers: the minimum 0, the lower-hinge 0, the median 4, the upper hinge 14, and the maximum 58. The ENCODE motif quality cutoff refers to the quality score assigned to all Tfb sites and varying from zero through 1000 [[24]], proportionately to the reliability of the predicted Tfbs. Additional details of the distribution of the number of Tfbs mapping to promoters with varied ENCODE quality cutoff and varied promoter window size are given in Tables 2 and 3.

Figure 1
figure 1

Histograms of the numbers of Tfbs in promoter regions depending on analysis widow size and ENCODE quality cutoff. This figure consists of ten panels identified through row and column margin labels. The top row provides information on Tfbs distributions including all ENCODE sites‚ while the bottom row illustrates distributions at the ENCODE quality cutoff of 500. The motif quality cutoff refers to the quality score assigned to all Tfb sites by the ENCODE consortium, which varies between zero through 1000 (from low to high quality). The promoter window sizes vary from 250 to 10,000 ± TSS (see column labels). The inclusion of all sites and the expansion of the analysis window result in distributions with longer tails in high numbers of mapping Tfbs.

Table 2 The distribution parameters for the number of transcription factor binding sites mapping to proximal promoters depending on the promoter window size and ENCODE quality cutoff
Table 3 The percentages of genes with 1 TF, 2 TFs, and up to 5 TFs depending on the promoter window size and ENCODE quality cutoff

Effective promoter size is about 6kb (±3000 bps from the TSS)

We have assumed above a given size for promoters. Can we use our data to determine an average upper limit to the size of promoters? We expect that TF binding sites should be concentrated near the TSS and as we move ever further away the increase in the number of TF binding sites should tend to a linear function, indicating background/random rates. As expected, the number of Tfbs increases progressively with the window size, transforming gradually to a linear, background, rate of increase (Figure 2a). Using a derivative to determine the point at which the trend linearises, the outer boundary of promoters is estimated at 3kb from the TSS (Figure 2b).

Figure 2
figure 2

Robustness to the variation in the size of the analysis window. This figure consists of three parts identified as ( a ) - ( c ). In part a, the number of transcription factor binding sites depending on the size of the promoter window was shown. As expected, the number of Tfbs was increasing progressively with the window size. However, the rate of the increase gradually decreased and transformed to linear. The point of the transformation was the presumed boundary of the proximal promoter. To localize this boundary more precisely, we fitted a local polynomial regression (loess) model and plotted its first derivative in part b. For all three subsets of ENCODE, there was a clear point of transformation where the rate of change (∆Tfbs) became constant (marked with ‘*’), at the distance from the TSS of approximately 3,000 base pairs (i.e., the promoter window of six thousand base pairs). Thus the outer boundary of promoters was estimated at three thousand base pairs from the transcription start site (TSS). In part c, we show that the correlation between the breadth of expression and the number of transcription factor biding sites was robust to variation in window size, although its strength was decreasing as the size of the analysis window was increasing. This observation suggested that Tfbs controlling the breadth of expression were enriched close to the transcription start site. Note that the analyses described here, used either a 2011 or 2012 ENCODE data-freeze. The 2011 meta data set included 2.7 million peaks for 148 transcription factors, derived from seventy-one cell-types with twenty-four additional experimental cell culture conditions [[31]]. Peak scores varied from zero through 1000. We used either all data or only high-quality peaks with the score above 500. The 2012 data-freeze, a broader dataset, consisted of 161 transcription factors and ninety-one human cell types with various treatment conditions [[32]].

Broadly expressed genes have more transcription factor binding sites

Is there something special about those genes with very many TF binding sites? Are they for example broadly expressed, as expected if TFs are dominantly activating? To analyse this we presumed, in the first instance, that a CAGE signal greater than ten tags per million (TPM > 10) classified a gene as expressed, or “on” in a given tissue (this was the consensus definition accepted by the FANTOM5 consortium). The breadth of expression (BoE) is the fraction of tissues or cell-lines in which the gene was “on”, i.e. in which it was transcribed. Figure 3 illustrates the distribution of TPM values in human tissues (Figure 3a), and the consequences of using too high a cutoff for BoE such as 100 or 1000 TPM (Figure 3b). The TPM value of ten is equivalent to ~3 mRNA copies per cell, based on 300,000 mRNAs per cell [[28]]. Using this definition, half of genes are relatively narrowly expressed. If, for example, transcripts are sub-divided into three categories, narrowly expressed (0 < the breadth of expression ≤ 0.33), intermediate (0.33 < the breadth of expression ≤ 0.66), and house-keeping (the breadth of expression > 0.66), nearly half are tissue specific or narrowly expressed (0.46 narrowly expressed, 0.14 intermediate and 0.21 housekeeping). Of the narrowly expressed transcripts, a very small fraction, 0.042 at the cutoff of 10 TPM or 0.053 at the cutoff of 100 TPM, are tissue-specific sensu stricto, i.e. expressed in one tissue only. The remaining 0.19 is the fraction of transcripts which lack evidence for expression in FANTOM5 tissue samples at the cutoff of 10 TPM, owing perhaps to their highly restricted spatial and/or temporal expression in a very limited subset of cells. In comparison to all genes, ENCODE Tfbs have higher average breadth of expression (BoE of 0.46 versus 0.295, Wilcoxon rank sum test P-value = 2.995e-08) with the fractions of tissue-specifc, intermediate and housekeeping Tfbs at 0.32, 0.17, and 0.38. Top ten housekeeping Tfbs included Pol2, JunD, c-Fos, JunB, Rad21, GTF2F1, NELFe, SREBP2, RXRA, and HSF1 (which all had BoE > 0.98). For 17 Tfbs (i.e. 12% of the total) we found no evidence of expression in tissue samples.

Figure 3
figure 3

The definition of the breadth of expression. This figure consists of two panels identified as a and b. Panel a shows a histogram of all TPM values for human tissues. The following are the characteristics of the distribution: n = 5.566005e + 06, mean = 1.807312e + 01, median = 3.090820e + 00, sd = 317,803, min = 0, max = 348,120. The cutoff of 10 TPM is signified with the red vertical line. The breadth of expression was the fraction of samples in which the gene was “on”, i.e. in which it was transcribed. The tags per-million (TPM) value of ten was accepted by the FANTOM5 consortium as the standard threshold for a gene to be “on” in a given library. We considered alternative cutoffs of TPM = 100 and TPM = 1000 in b which compares the density plots for the breadth of expression (BoE) at the cutoffs of 10, 100, and 1000 TPM (see figure legend). It is clear that cutoffs of 100, and 1000 are too high, resulting in almost no intermediate and housekeeping genes (i.e. genes with BoE > 0.33).

Might the correlation between expression breadth and the number of Tfbs be an artifact owing to a correlation with a further parameter? Might indeed the chromatin status or underlying nucleotide content be alternative and better predictors? To explore this we consider a multiway set of correlations and partial correlations, i.e. each variable predicting breadth, controlling for all others (Table 4, see also Figure 4). This suggested a link between BoE and the number of transcription factor binding sites to be the strongest correlation (rho = 0.48, Figure 4 and Table 4), even after controlling for all other parameters (the corresponding partial correlation in Table 4 has rho = 0.40). While the raw data shows some scatter (Figure 5a-b-c) the monotonic trend is easily visualized in a box plot based on deciles of the data by breadth of expression (Figure 5e).

Table 4 Correlations and partial correlations
Figure 4
figure 4

A correlogram of eleven variables describing promoter architecture. In the correlogram, there are four measures of GC-content: GC-content in a 1 kb proximal promoter (GC), GC-content in a 20 kbps window around the promoter (GC_big), GC-content in a third codon position (GC3), and the frequency of CpG sites (CpG); there are also four measures describing the number of transcription factor binding sites in promoters: Tfbs1 (Tfbs_length – straight number of Tfbs), Tfbs2 (Tfbs_length_unique – the number of unique Tfbs), Tfbs3 (Tfbs_length_noPol2 – the number of Tfbs excluding PolII), and Tfbs4 (Tfbs_length_unique_noPol2 - the number of unique Tfbs excluding PolII), a measure of methylation (methyl), a signature of digestion by DNASE1 (DNASE1), and the breadth of expression (BoE).

Figure 5
figure 5

The breadth of expression correlated with the number of transcription factor binding sites. The breadth of expression correlated with the number of transcription factor binding sites in proximal promoters. This figure consists of five parts identified as a - e. Scatterplots were shown for ( a ) FANTOM5 human tissues, ( b ) FANTOM5 human primary cells, ( c ) FANTOM5 human cancer cell lines, ( d ) human data in Gene Expression Atlas [[51]]. The red line signified the linear model for the smoother line, while the blue line signified the non-linear model. Panel ( e ) provided an alternative illustration of the trend using a boxplot for the discretized breadth of expression in FANTOM5 tissues. Outlying tissue-specific genes with many transcription factor binding sites, which were likely enriched in inhibitory TFs, were marked in red. FANTOM5 tissues, primary cells and cancer cell lines were the three subsets of samples in FANTOM5 whose numbers were given in Table 1. Numbers of tags in FANTOM5 were normalized to tags per million (TPM). The TPM value of ten was chosen as a standard cutoff for a gene to be “on”. For Gene Expression Atlas, Affymetrix average difference (AD) higher that 200 classified a gene as “on” or expressed in a given tissue. Proximal promoters were defined by a symmetrical window of one kilobase in size around the transcription start site (+ − 500 bps from the TSS). As an additional control, we performed a randomization procedure where proximal promoters of all genes were shuffled. The value of the t-statistic for the strength of correlation in the observed dataset was compared against 10,000 datasets with randomized assignments between promoters and RefSeqs. The value of t-statistic for observed data (54.29404) was compared with t-statistics for 10,000 randomized datasets (mean −0.00959) and the P-value obtained was lesser than 2.2e-16.

As regards possible chromatin effects we observe (Table 4 and Figure 4), as expected, a positive correlation between BoE and ENCODE DNASE1 signal (Spearman’s rho = 0.19, P-value < 2.2e-16), and a negative correlation between BoE and ENCODE methylation signal (Spearman’s rho = −0.11, P-value < 2.2e-16). There was also a strong correlation of BoE with GC- and CpG-content (rho = 0.33, P-value < 2.2e-16; and rho = 0.42, P-value < 2.2e-16, respectively). There was also a strong correlation between CpG and TfbsNo. (rho = 0.45, P-value < 2.2e-16, Figure 4) and GC-content and the number of Tfb sites (rho = 0.29, P-value < 2.2e-16, Figure 4). Strikingly, however, on multiway partial correlation, the strength of these effects tended to diminish dramatically. Correlation with GC went from a raw correlation of 0.33 to a partial of just 0.03. Correlation with DNASE1 went from 0.19 to just 0.06 and the methyl effect diminished from −0.11 to just −0.04. By contrast the effect of transcription factor number was relatively unchanged (0.48 prior to multiway analysis, 0.4 after). These results suggest that the chromatin effects may mediate the control of gene expression, but the prediction of breadth of expression is best done via transcription factor information (and this is most likely the casual association).

Closer scrutiny of the impact of GC-content as a predictor supports the view that it is GC of the core promoter rather than a more regionalized GC-content that impacts breadth of expression. When we divided promoters into low-GC (less than 50%, n = 5650) and high-GC (more or equal than 50%, n = 25710), the second group had more than three times higher average BoE (the exact ratio was 0.3379/0.099 = 3.41) and on average bound more than four times more TFs (9.55/2.29 = 4.17). The GC-content of proximal promoters (defined as a 1 kbps window) was much higher than that of surrounding DNA sequences (20 kbps window): 0.594 versus 0.463 (Welch Two Sample t-test, P-value < 2.2e-16). Similar results were reported by the ENCODE consortium who found that GC-content of ChIP-seq sites was 61 ± 5% for TSS-proximal peaks [[29]]. The exact cause of this effect is not yet fully understood. Although some TF motifs are GC-rich [[30]], these are usually much smaller that the actual ChIP-seq peaks (8–21 bps vs. 250 bps).

As shown in Table 5, including GC-content-related measures (i.e., GC-content, CpG, and CpGoe) in a SVM learning dataset does not substantially increase prediction accuracy over the simple SVM trained with data on Tfbs numbers. (CpGoe is a measure of observed CpG frequency normalized by the frequencies of G and C nucleotides proposed to work as a proxy of methylation over large evolutionary timescales [[17]].) Nevertheless, partial correlation between BoE and CpG persisted after controlling exclusively for TfbsNo. (rho = 0.214). However, partial correlation between BoE and TfbsNo. was higher, after controlling exclusively for CpG (rho = 0.37), suggesting this effect was dominant. Taken together, these results suggested that CpG was more of a place marker than a key part of the mechanism. Promoter GC-content was clearly distinct from the isochore GC-content or GC3 (while the latter two correlated closely together, see Additional file 1: Figure S7).

Table 5 SVM trained with data on the numbers of interacting Tfbs (SVM-Tfbs) improved on simple correlation, but adding data on GC-content (SVM-Tfbs + GC) did not lead to further improvement of predictions

We note that we see little or no evidence for a class of genes so highly broadly expressed that they dispense with TFs all-together. In fact, there were only thirty-nine broadly expressed genes with fewer than ten high-quality TFs in a broad 10 kb window around the TSS (Additional file 2: Table S1).

Expression level is not well predicted by Tfbs number

The correlation between the number of transcription factor binding sites and the breadth of expression was strongest at the cutoff for a gene to be “on” set at ten TPM. The correlation was much weaker at the cutoff of 100 TPM, and disappeared at the cutoff of 1,000 TPM (Table 6). One interpretation of this result is that TFs control mostly where the gene is expressed, but not at what level. The strength of expression might be regulated predominantly by higher-level chromatin architecture or epigenetic marks. To address this in more detail, we also ask whether Tfbs number predicts the level of expression of a gene.

Table 6 The number of transcription factor binding sites correlated with the breadth of expression, the mean expression, and the median expression, but not with the value of the maximum expression of a transcript

Previous authors suggested a strong correlation between the breadth of expression and average expression of a transcript [[17]]. However, this might be, at least partially, a methodological circularity. If one permits all genes that are unexpressed in a given tissue to score zero for that tissue (definition 1), then tissue specific genes “mean expression” will be dominated by the sum of zeros, hence forcing a tissue specific genes to have low mean level. When instead, we define mean level, as the mean level of expression, in the tissues within which the gene is expressed (definition 2), we find no evidence for a correlation between expression breadth and expression levels in FANTOM5 using parametric statistics (Table 6), and only weak evidence using nonparametric statistics (Table 7).

Table 7 There is evidence for a real correlation between expression breadth and expression levels if non-parametric statistics are used

We can ask how these two definitions also relate to Tfbs number. Using definition 1 of mean/median expression, we find that the number of transcription factor binding sites correlates with the mean expression, and the median expression, but not the maximum expression of a transcript (see Table 6, Figure 6). However, the correlations become very weak (with mean: rho = −0.056, P-value = 8.585e-16; and with median: rho = −0.0151, P-value = 0.0309), when they were calculated only across tissues in which the gene was “on” (at the cutoff of 10 TPM). As definition 1 forces the mean and median across all tissues to covary with the breadth, the strong correlations found using definition 1 were most likely just detecting the primary underlying correlation with the breadth of expression. We conclude that the Tfbs number is a poor predictor of expression rates when expression breadth isn’t a compounding factor.

Figure 6
figure 6

The correlation between the breadth of expression in human tissues, the mean and the maximum expression, and the number of transcription factor binding sites. This figure consists of sixteen parts identified as ( a ) - ( p ). Four measures related to the breadth of expression were considered: ( a-b-c-d ) the breadth of expression at the cutoff of ten TPM, ( e-f-g-h ) the breadth of expression at the cutoff of 100 TPM, ( i-j-k-l ) the mean expression, and ( m-n-o-p ) the maximum expression. The number of transcription factor binding sites was estimated in four different approaches: ( a-e-i-m ) the total number, ( b-f-j-n ) the number of unique binding sites, ( c-g-k-o ) the total number excluding RNA polymerase II binding sites, and ( d-h-l-p ) the number of unique binding sites excluding the polymerase. The red line signified the linear model for the smoother line, while the blue line signified the non-linear model. The correlation between the number of transcription factor binding sites and the breadth of expression at the cutoff of ten TPM was robust under four different approaches to estimating the number of transcription factor binding sites. Interestingly, this correlation was driven by transcripts with between zero to twenty binding sites (r p  = 0.42), and was much weaker for promoters with more than twenty sites (r p  = 0.098). At the value of approximately twenty on the X-axis ( a-b-c-d ), the blue smoother (the non-linear model) reached a plateau and diverged from the red smoother (the linear model). This figure suggests that the correlation presented here was strongest at the cutoff for the breadth of expression of ten TPM, and was not biased by the polymerase or another individual transcription factor. The correlations with the mean expression were likely secondary to the correlation with the breadth of expression (see Results: Broadly expressed genes have more transcription factor binding sites).

The correlation between TF binding sites and expression breadth is robust

The above results strongly support the view that more TF binding is correlated with expression in more tissues. How robust is this result? Is it true in both normal and diseased states? Is it robust to control for whether or not RNA PolII is included in the set of binders? Is it dependent on the assumed size of the promoter? In the three sections below we consider these and other possible confounders.

Correlations are robust to alternative assumptions of the promoter size

Given the decay in the rate of the increase of the number of TFs as the size of promoters expands, we presume that increasing the assumed promoter size should start to cause a decay in the correlation between expression and the number of TFs, just because we are diluting signal (true TF binding) with noise (spurious or unassociated binding). As expected (Figure 2c), the correlation between the breadth of expression and the number of transcription factor binding sites, although robust to the variation in window size, decreases as the size of the analysis window increases. A converse interpretation of this is that Tfbs controlling the breadth of expression are enriched close to the transcription start sites.

The inter-relationship, while strongest for small window sizes, persisted for windows up to 40kb in size (Figure 2c), well beyond the 6kb limit of effective promoter size. We expect this limit to be greater than that derived from the rate of increase of TFs measure (circa 3kb ± TSS) as it takes a considerable dilution of the signal of the TF loaded TSS to remove any correlation. The trend was similar when the ENCODE meta data set from the 2011 freeze [[25]] was compared with the broader 2012 freeze [[26]]. Both these data sets were comprehensive in their coverage of transcription factors (148 and 161, respectively). Both data freezes also covered a wide sample space: the earlier freeze with seventy-one cell-types and twenty-four additional experimental cell culture conditions [[31]], and the later freeze with ninety-one human cell types with various treatments [[32]]. Finally, the trends detected were robust to alterations in the quality cutoff for ENCODE transcription factor binding sites (Figure 2a-b-c).

The correlations are stronger across cell lines than across gross tissues

To control for the possibility of a sample bias or differences between normal and diseased tissues, we tested whether the correlation between the breadth of expression and the number of transcription factor binding sites held across the entire FANTOM5 sample space. Figure 5a and Figure 6 show the results for human tissues. We confirmed that the trends are also seen for primary cells (Figure 5b and Additional file 3: Figure S1) and cancer cell lines (Figure 5c and Additional file 4: Figure S2). The Pearson correlation coefficient (r p ) between the breadth of expression and the number of transcription factor binding sites equaled 0.53 for primary cells, and 0.61 for cancer cell lines. The correlation is strikingly stronger for cell lines than for tissues. It makes sense that the correlation was stronger for primary cells and cell lines (r2 ~ 28-37%) than for tissues (r2 ~ 20%), as tissues are complex mixtures of cell types where some of the cell-type-specific signal might have been lost.

The correlation between the breadth of expression and the number of transcription factor binding sites holds when RNA polymerase II sites are excluded

Above we considered all bindings at promoter regions of genes, including RNA polymerase II binding sites. One might readily object that if one includes PolII binding then highly expressed genes may well have more bindings, if only because they have more PolII. Might then the correlation between the breadth of expression and the number of transcription factor binding sites be driven by RNA polymerase II binding sites, or biased by another abundant transcription factor?

To investigate this we employed four approaches for counting, these being: (1) the total number of binding sites, (2) the number of unique binding sites, (3) the total number of binding sites excluding RNA polymerase II, and (4) the number of unique binding sites excluding the polymerase. We find that the correlation holds regardless of the method (see Figure 6). Indeed, results using these four measures were largely indistinguishable. For example, when polymerase sites were excluded, the correlation between the number of transcription factor binding sites and the breadth of expression in human tissues was 0.434 (t = 84.8393, df = 31093, P-value < 2.2e-16). The correlation was 0.435 when additionally only unique sites were counted (t = 85.1458, df = 31093, P-value < 2.2e-16). In comparison, the original correlation including all sites was 0.448 (t = 88.2645, df = 31093, P-value < 2.2e-16), and when only unique sites were counted 0.45 (t = 88.6194, df = 31093, P-value < 2.2e-16). Indeed, the three derived measures correlated very highly with the original measure with correlation coefficients in pairwise comparisons of 0.994, 0.997, and 0.992 for unique sites, no PolII sites, and unique sites excluding PolII (all P-value < 2.2e-16), respectively.

We can also turn the data the other way around and ask whether sites with more TF bindings also have more PolII. Such a correlation would provide sound evidence that more TFs do indeed result in more transcription. We find this to be the case. After excluding its own sites, the polymerase signal correlated strongly with the total number of transcription factor binding sites (r p  = 0.75). The correlation between the breadth of expression and the number of transcription factor binding sites persisted after controlling for the polymerase signal (partial Spearman’s correlation coefficient equaled 0.3).

The divergence of the promoters of paralogs strongly predicts divergence of their expression patterns

As discussed in the introduction, a common method to approach the problem of the degree of promoter-centred control of gene expression has been to ask about the similarity in gene expression of paralogs as a function of the similarity in their promoter domains. In addition to the correlation between the breadth of expression and the number of transcription factor binding sites, we found that the divergence of proximal promoters (measured via a Jaccard Index on Tfbs repertoire – see Methods) correlated strongly with expression divergence, measured by Pearson’s R. The r p for this trend equaled 0.282 when only the youngest paralog pairs were taken into account. The r p was even higher when all daughter pairs were taken into account 0.54 (t = 239.8391, df = 136608, P-value < 2.2e-16). However, the latter comparisons were not fully independent and the results might have been biased by large gene families with a high number of pairwise comparisons. For example, the core histones of H2A@, H2B@, and H3@ families underwent dramatic expansions in placental mammals, resulting in paralogs which are highly co-expressed in proliferating tissues such as the thymus and the testis (manuscript in preparation).

Pearson’s correlation corresponds well to biologist’s intuitive understanding of what co-expressed genes are and has frequently been used in the past to measure expression divergence [[33]–[35]]. This is because biologists are frequently interested in the identification of tissue-specific or disease-specific biomarkers, and Pearson’s correlation works well for tissues-specific genes. Nonetheless, it is suggested [[36]] that Pearson’s correlation may be affected by the noise present in microarray data. However, alternative measures such as the Euclidean distance may be biased by normalization [[37]]. We used three types of correlation to measure paralog co-expression: Pearson’s, the Kendall rank correlation coefficient, and Spearman's rank correlation coefficient. The correlation between promoter divergence and paralog co-expression held irrespective of the type of the correlation statistic, whether parametric or non-parametric (Figure 7). As expected, the correlation disappeared when duplicate pairs were randomized as a means of negative control (Figure 7d-e-f). The correlations are not simply owing to some paralogs switching from being lowly to broadly expressed (or vice versa). Rather the correlations remains even when we consider paralogs with approximately the same breadth (Table 8, Figure 8).

Figure 7
figure 7

Expression pattern divergence between paralogs correlated with promoter divergence. This figure consists of six parts identified as ( a ) - ( f ). Expression pattern divergence between duplicates was measured with either ( a ) Pearson’s correlation; ( b ) the Kendall rank correlation coefficient; or ( c ) Spearman's rank correlation coefficient. Promoter divergence was measured using Jaccard index (JI). The correlation disappeared when duplicate pairs were randomized ( d-e-f ) proving that it was well defined and specific. The red line signified the linear model for the smoother line, while the blue line signified the non-linear model. This figure suggests that the correlation between the breadth of expression and the number of transcription factor binding sites persisted if alternative non-parametric measures of expression distances between paralogs were used. Details of the correlations were given in Table 8.

Table 8 The correlation between the Jaccard index ( JI ) and paralog co-expression was robust in respect to the breadth of expression
Figure 8
figure 8

The correlation between the promoter divergence of paralogs and paralog co-expression was robust in respect to the breadth of expression of target genes. This figure consists of twelve parts identified as ( a ) - ( l ). Four sets of paralog pairs were considered: ( a-b-c ) both paralogs were tissue-specific with the breadth of expression ≤ 0.33, ( d-e-f ) both genes were intermediate, ( g-h-i ) both genes were housekeeping with the breadth of expression > 0.66, and ( j-k-l ) one of the paralogs was tissue-specific and the other housekeeping. Pearson’s ( a-d-g-j ), the Kendall rank correlation coefficient ( b-e-h-k ) and Spearman's rank correlation coefficient ( c-f-i-l ) correlations were plotted. The correlation between the Jaccard index and paralog co-expression was robust under all these conditions. Paralog promoter divergence was measured using Jaccard index (JI). Paralog expression divergence was measured using Pearson correlation. The red line signified the linear model for the smoother line, while the blue line signified the non-linear model. Numbers of tags in FANTOM5 were normalized to tags per million (TPM). The TPM value of ten was chosen as a standard cutoff for a gene to be “on”, and the breadth of expression was defined as the fraction of FANTOM5 human tissue samples in which a transcript was “on”.

While the divergence of expression between paralogs is predicted by the divergence of transcription factor repertoire, we additionally observe a trend for young duplicates to be preferentially tissue-specific and have fewer transcription factor binding sites in their promoters. Duplicates mapping to the youngest taxa group (i.e. primates) have average BoE almost four-times lower, and average TfbsNo. 2.7 times lower (Figure 9a, Tables 9 and 10) than duplicates mapping to the oldest group (i.e. eukaryotic). Genes that originated though mammalian gene duplication events had intermediate BoE, at approximately 155% of primate BoE, and less than half of the average eukaryotic BoE and TfbsNo. The differences in mean BoE and TfbsNo, were highly statistically significant with all pairwise comparisons having very low P-values (see Additional file 5: Table S3 and Additional file 6: Table S4).

Figure 9
figure 9

The correlation between BoE and TfbsNo. is recapitulated over evolution. This figure consists of four panels identified as a through d. Panel a shows a boxplot for BoE depending on age of gene duplication. Panel b shows a similar boxplot of TfbsNo. Young genes are more tissue-specific and have fewer TF binders. The correlation between TfbsNo. and BoE was strongest for young genes (Spearman rho = 0.531, 0.513, 0.46, 0.447, and 0.403 for increasingly older taxon groups, from primate through to eukaryotic genes). To explain the origins of additional tissue-specific genes in younger taxa, we divided duplication events into three subclasses: “housekeeping conserved”, “tissue − sp. conserved”, and “transformative”. Panels c and d show barplots for the three different mechanistic subtypes of gene duplication events (absolute numbers and relative proportions thereof, respectively).

Table 9 Young duplicates were more tissue-specific
Table 10 Young duplicates had fewer Tfbs regulators

To investigate the origin of new tissue-specific genes, we divided duplication events into three subclasses: “housekeeping conserved” (both paralogs were housekeeping), “tissue − sp. conserved” (both paralogs were tissue-specific), and “transformative” (where one daughter gene was housekeeping while the other was tissue-specific). The relative proportion of “tissue − sp. conserved” events increases for younger taxa indicating that this class of duplication events is responsible for the majority of the increase of tissue-specificity observed for young taxa (Figure 9c and d). This accords with a model suggesting that successful duplication events tend to be those with minimal impact [[38],[39]]. It also accords with the finding that tissue-specific genes are more likely to belong to large gene families [[2]]. The coupling between duplication age and breadth may bias some statistics. If the breadth of expression of a gene in any manner predicts divergence in expression, this bias has the potential to mislead any analysis that considers the degree of divergence between promoters and divergence in expression, as the least diverged duplicates (i.e. the youngest duplicates) will be systematically biased towards the tissue-specific end of the spectrum. However, the trend for gradual expression divergence of paralogs was described in both multicellular [[1],[2]], and unicellular organisms [[40]] where tissue-specificity cannot be an issue.

Broad expression is associated with specific transcription factors or groups of cooperating factors

The broad-brush correlations that we have addressed above suggest that the more TFs bind a promoter the more broadly expressed the gene. But are there some TFs that are especially influential in driving broad expression or is the effect simply owing to an accumulation of TFs causing increased likelihood of broad expression? To address this, we clustered the breadth of expression with a matrix of transcription factors to identify key associations (Figure 10).

Figure 10
figure 10

The clustering of the breadth of expression with the number of transcription factor binding sites. The breadth of expression clustered with RNA polymerase II and the transcription initiation factor TFIID (TAF1). More interestingly, in human tissues, the breadth of expression also clustered tightly with Mxi1, YY1, NFKB, HEY1, Sin3A, and c − Myc, suggesting these transcription factors were key in determining the breadth of expression (this cluster was marked as A). Other transcription factors formed two clusters with low and high distance to the breadth of expression (these clusters were marked as B and C, respectively). Many clusters of co-localizing transcription factors could be observed and were annotated (a)—(y). To test the robustness of this analysis, the number of transcription factors was measured in several different ways, which reassuringly clustered together and proved indistinguishable (part of the A cluster). The different measures were: sum of all sites (marked as Tfbs_x_length), sum of unique sites (Tfbs_x_length_unique), sum of all sites without RNA polymerase II (Tfbs_x_length_noPol2), and finally the sum of unique sites without the polymerase (Tfbs_x_length_unique_noPol2). The breadth of expression was also transformed in several ways which proved equivalent by forming a tight cluster (part of the A cluster). Namely, the breadth of expression was encoded as either a continuous variable (marked as breadth_continuous), discretized into three bins (breadth_discrete_3), discretized into ten bins (breadth_discrete_10), or transformed and expressed as a square root (breadth_sqrt_1).

First, the breadth of expression was merged into one matrix with the number of ENCODE transcription factor binding sites. Next, a heatmap was drawn for this matrix in order to determine which transcription factors correlated closest with the breadth of expression. i.e. which transcription factors acted as molecular switches for house-keeping expression (Figure 10). The heatmap in Figure 10 uses Pearson’s correlation as the distance measure. Similar results were obtained for human tissues with both the Kendall rank correlation coefficient and Spearman's rank correlation coefficient (Additional file 7: Figure S5 and Additional file 8: Figure S6). We also investigated distance-based correlations such as Euclidean, Manhattan, or Minkovski distances but these measures did not recover any non-trivial clustering.

Four broad classes of transcripts emerged through this integrative analysis. Class A genes are typical broadly expressed Tfbs-rich genes. Class B genes are unusually broadly expressed Tfbs-poor genes. Class C are unusual tissue-specific Tfbs-rich genes, while class D are typical tissue-specific Tfbs-poor genes. These four classes are marked A—D in Figure 11. The numbers of transcripts in A, B, C and D are 7824, 3206, 3593, and 16170, respectively. The mean numbers of Tfbs per transcript in each class are: 18.77, 5.41, 16.41, and 1.99 (in FANTOM5 human tissues, 500-cutoff, ENCODE 2011 data-freeze). Here the cutoffs of ten for Tfbs and 0.33 for the breadth of expression are used. Differentially distributed transcription factors are listed in Additional file 9: Table S2 with their respective frequencies in clusters A, B, C and D. P-values were calculated using Fisher's exact test with a Bonferroni correction for multiple testing.

Figure 11
figure 11

Conceptual diagram of the four classes of tissue-specific and broadly expressed transcripts rich or poor in transcription factor binding sites. The four classes were marked with A, B, C , and D. The cutoffs were as follows: ten transcription factor binding sites for Tfbs-rich, the breadth of expression of 0.33 for broadly expressed, and the TPM value of ten for a gene to be “on”. The breadth of expression was defined as the fraction of FANTOM5 human tissue samples in which a transcript was “on”. The biological interpretation of this figure was that typical housekeeping genes were Tfbs-rich, while typical tissue-specific genes were Tfbs-poor. The diagram uses Figure 5a as its background to illustrate the number of transcripts in each of the four classes. The four classes of transcripts will facilitate classification of transcription factors and their impact on individual genes as either activatory or inhibitory (Additional file 9: Table S2).

As expected, the breadth of expression clustered with RNA polymerase II and the transcription initiation factor TFIID — TAF1. This was perhaps unsurprising as the polymerase and TAF1 are constitutive components of the transcription apparatus. More interestingly, in human tissues, the breadth of expression also clustered closely with Mxi1, YY1, NFKB, HEY1, Sin3A, and c − Myc (this cluster was marked with an A in Figure 10) suggesting these transcription factors were among control switches. In human primary cells, the breadth of expression also clustered with the polymerase, TFIID, NFKB, HEY1, Sin3A, and c − Myc (Additional file 10: Figure S3). However, in cancer cell lines, the breadth of expression only clustered with RNA polymerase II and TFIID (Additional file 11: Figure S4) suggesting that cancerous transformation interferes with the majority, except the most rudimentary, control switches for the breadth of expression.

Other transcription factors formed two clusters with either low or high distance to the breadth of expression (these clusters were marked with a B and a C in Figure 10). These two clusters could be enriched in either housekeeping transcription factors (B), or tissue-specific transcription factors (C). In cancer cell lines (Additional file 11: Figure S4), clusters with low (B) and high (C) distance to the breadth of expression could be enriched in oncogenic transcription factors, and anti-oncogenic or tumor-specific transcription factors, respectively.

Many clusters of co-interacting transcription factors can be inferred from Figure 10. These clusters were marked with lowercase letters: (a) Nanog and Pou5f1; (b) Srebp1 and Srebp2; (c) STAT1-3; (d) mef2a and mef2b; (e) GATA-1 and GATA-2; (f) MafF and MafK; (g) BAF170 and BAF155; (h) AP-2 alpha and gamma; (i) FOS and FOSL2; (j) Jun and JunD; (k) FOXA1 and FOXA2; (l) HNF4A and HNF4G; (m) CTCF targeted by three different antibodies; (n) Rad21, SMC3 and CTCFL; (o) Pol3, BRF1, RPC155, and BDP1; (p) E2F1, E2F4, and E2F6; (q) SIX5, Znf143 and ETS1; (r) ELK4 and ELF1; (s) PAX5 targeted by two different antibodies; (t) ZBTB33, BRCA1 and CHD2; (u) NELFe, GTF2B, and TAF7; (v) POU2F2 and Oct-2; (w) NF-YB, NF-YA, c-Fos and SP1; (x) USF1 and USF2; (y) PolII and TAF1. Some apparent clusters are the same transcription factors targeted by different antibodies (for example, the CTCF rabbit polyclonal, the CTCF_C − 20 goat polyclonal, and the CTCF_SC − 5916 goat polyclonal antibody) and so should be disregarded. By contrast, some of the clusters are known cooperative complexes. For example, Rad21 and SMC3 form the cohesin complex. The cooperation between Nanog and Pouf1 (alias Oct4) is well described [[41]]. Other clusters suggest entirely new molecular interactions which should provide material for experimental verification.

A support vector machine (SVM) method predicts the breadth of expression better than the correlation alone

Given the evidence for cooperativity between TFs in their association with breadth of expression and for key control TFs, we would expect that more sophisticated statistical tools should improve the predictive ability of a model relating TF binding to expression breadth. To address this we established a machine learning approach. Our intention here is not to produce a better statistical model by incorporating non-causal (e.g. rate of protein evolution) and causal (TF binding) predictors of expression breadth. Rather we simply wish to ask whether incorporation of the likely causal factors in a more sophisticated statistical framework permits better understanding of the control of expression breadth.

The support vector machine is a commonly used machine learning approach to prediction. We trained the support vector machine using a randomly chosen half of the dataset. Each row of the basic SVM training data-frame was a vector representing the number of Tfbs of each gene; although, we also considered SVMs trained using promoter GC- and CpG-contents (Table 5). The resulting SVM model was then applied to predict the value of the breadth of expression for the other half of the dataset. Note that in this mode of operation with continuous data being predicted, an SVM is best regarded as a form of non-linear regression (rather than a discrete classifier) and hence the appropriate metric for considering its ability is the correlation between the observed and predicted variable (rather than AUC, accuracy, etc.). Such a correlation-based appraisal also permits direct comparison with the simple correlation approach examined above.

As noted above, in cancer cell lines, the correlation between the breadth of expression and the number of transcription factor bindings sites is 0.61. The SVM’s prediction improved on this and correlated with the observed value of the breadth of expression with r p  = 0.7460 (Table 5). Results described herein were obtained using the promoter window of 1000 base pairs but essentially identical values were obtained using the promoter window of 4000 base pairs (data not shown). As negative and positive controls we used a teaching dataset where the response variable was scrambled or included as one of the training features, respectively. The SVM is thus capable of explaining 58% of the variation, as opposed to the correlation method’s 37%. The prediction accuracy was not due to polymerase signal alone since the SVM performed equally well when TAF1 and all types of polymerase II sites were removed (r p  = 0.759). Moreover, the predictor did not simply rely on summing up of Tfbs, since an SVM trained using only the sums of Tfbs as features (Tfbs_x_length, Tfbs_x_length_unique, Tfbs_x_length_noPol2, Tfbs_x_length_unique_noPol2) did not improve on the simple correlation (r p  = 0.646, r2 = 0.42). The number of support vectors was high (approximately 60% of the training cases) suggesting that no simple discriminatory features could be found. Taken together, these results underlined a cooperative effect of many Tfbs acting together and clustering in a narrow window around the TSS to control the breadth of expression.

Are promoters in open chromatin “sticky”?

Even though specific interactions can be identified, that a simple correlation approach can capture so much is striking. The fact that a correlation approach works is consistent with the notion that most transcription factors in eukaryotes are activators. There is, however, an alternative interpretation of the correlation between number of TFs and increasing expression, this being that it reflects more a passive process of spurious TF binding. A simple model in which open/transcribed chromatin is to some degree “sticky” could in principle predict the same correlation. At the limit there might be a single TF that forces broad expression, but because of its presence other TFs are recruited, not because they are needed, but because TFs might be attracted to open chromatin and transcriptional hotspots as iron filings are attracted to a magnet. In this context, were for example GC rich sequence “sticky” for transcription factors, this might explain the GC-TF correlations. However, such a “sticky” model would require strong binding of the TFs to the DNA, rather than weak and short-lived non-specific interactions which are most unlikely to resist the processing of the TF-DNA interaction in ChIP-seq methodology.

Several facts argue against the “sticky” model. First, our approach has recovered known interacting complexes, such as cohesin and CTCF [[42],[43]], suggesting that much of the signal is owing to functional rather than spurious effects. Furthermore the correlation between breadth and number of transcription factors is more profound for transcripts with fewer than twenty transcription factor binding sites (r p  = 0.42), than it is for transcripts with more than twenty (r p  = 0.098). We would expect the opposite if Tfbs simply accumulated in a runaway positive feedback loop.

In addition, we can ask whether the TF bindings sites are clustered within promoters. Were the “sticky” spurious binding model correct, we might also expect that TF bindings sites are randomly located within promoters. By contrast, a model of cooperative binding to DNA and the synergistic mode of action in attracting polymerase and activating transcription, might predict TF bindings sites to cluster and overlap more than expected by chance. These two alternative concepts are illustrated in panels a and c of Figure 12. Panel a illustrates a situation where Tfbs are located in a “stacked” arrangement close to the TSS. This arrangement results in high percentage overlap in pairwise Tfbs comparisons. To test for any trend on the whole genome scale, we calculated percentage overlap in all pairwise Tfbs comparisons in each promoter (1 kb) for 25,930 RefSeq transcripts (Figure 12b). The observed overlap (corresponding to the “stacked” arrangement of Tfbs illustrated in Figure 12a) was contrasted against a randomized dataset where Tfbs were assigned random positions within the same proximal promoter (corresponding to the random arrangement of Tfbs illustrated in Figure 12c and resulting in a lower percentage overlap in pairwise Tfbs comparisons). The average observed overlap across all promoters and Tfbs pairs was 0.4586%, much higher than the average overlap in the randomized dataset (0.2295%), suggesting “stacked” rather than dispersed arrangement of Tfbs (t-test, P-value < 2.2e-16). These results are in agreement with the general trend demonstrated by ENCODE for almost all transcription factor binding sites to have highest densities very close to the TSS [[29],[44]]. As an example, we show the SRSF2/MFSD11 locus in Figure 12 panel d. This locus on chromosome 17 (start at 74,732 kbps, end at 74,734 kbps) has the highest number of Tfbs in our dataset, which are clearly overlapping or “stacked”. The SRSF2/MFSD11 locus has 71 Tfbs in the window ± 1000bps from the TSS and drives bidirectional transcription of serine/arginine-rich splicing factor 2 (SRSF2) and major facilitator superfamily domain containing 11 (MFSD11).

Figure 12
figure 12

TF bindings sites are clustered (or “stacked”) within promoters. Two theoretically possible alternative Tfbs distributions in proximal promoters are illustrated in panels a and c. Panel a illustrates a situation where Tfbs are located in a “stacked” arrangement close to the TSS. This arrangement results in high percentage overlap in pairwise Tfbs comparisons. In contrast, panel c illustrates a situation where Tfbs are randomly distributed in the proximal promoter. To test for these two trends on the whole genome scale, we calculated percentage overlap in all pairwise Tfbs comparisons in each promoter (1 kb) for 25,930 RefSeq transcripts (panel b ). The observed overlap (corresponding to the “stacked” arrangement of Tfbs illustrated in panel a) was contrasted against a randomized dataset where Tfbs were assigned random positions within the same proximal promoter (corresponding to the random arrangement of Tfbs illustrated in panel c and resulting in a lower percentage overlap in pairwise Tfbs comparisons). As an example, we demonstrate the SRSF2/MFSD11 locus in panel d, which has the highest number of Tfbs (i.e. 71) in our dataset, arranged in the overlapping or “stacked” mode.

A further argument against the “sticky” model of Tfbs binding comes from the examination of the breadth of expression of ENCODE Tfbs and their targets. We detected a strong correlation between the average breadth of expression of Tfbs targets and the breadth of expression of respective regulating Tfbs (Additional file 5: Table S3, Additional file 12: Figure S8) with the Spearman rho = 0.3176 (P-value = 0.0001242). Tissue-specific Tfbs bind on average more tissue-specific genes (mean BoE of targets 0.394), than housekeeping Tfbs (mean BoE = 0.468), Wilcoxon rank sum test P-value = 6.805e-05. Under the pure “sticky” model, we would expect no correlation. The picture is not clear-cut, however. Clearly, tissue-specific transcription factors can have targets in promoters of housekeeping genes. It is possible that housekeeping expression in certain tissue-types demands activation by this tissues unique transcription factors in addition to the standard set of TFs mediating housekeeping expression. Finally, TFs may differ in the degree of “stickiness”. Some TFs may be “sticky” towards open chromatin, while others bind very specifically, perhaps playing a key role in the initial act of the opening of the chromatin.

Another argument against the “sticky” model is that open/close chromatin is only a binary signal which cannot account for all the complexities of spatiotemporal gene regulation in Vertebrates. Moreover, the correlation between the breadth of expression and DNASE1 sensitivity, an excellent marker of open chromatin, was much weaker than that between the breadth of expression and the number of potential interacting transcription factors (Table 4).

Yet another argument against the sticky model is that TfbsNo. describes the general affinity landscape of promoters for Tfbs binding (i.e. the ability to bind TFs). Which TFs are actually bound will depend on a particular tissue and may vary greatly. This is because ENCODE inputs were derived datasets, where individual peaks from different tissues were merged if tagged to the same transcription factor binding site. In fact, counting Tfb sites across different tissues would be a source of a logical error, as a circular association between the breadth of expression and the presence of Tfb sites in multiple tissues would obscure any causal connections.

While multiple lines of evidence argue against the “sticky” chromatin model as the best explanation for the correlation between breadth of expression and number of TF binding sites we cannot entirely refute the hypothesis. Indeed scrutiny of the number of TFs binding the very most broadly expressed genes (right most column in Figure 5e) suggests that these have slightly more TFs than expected given the numbers in the prior bins. There also remains the possibility that the TFs that drive broad expression are themselves “sticky” and attract more TFs. Such a model is more ad hoc than the “sticky” chromatin model, having to make the extra assumption that only TFs associated with broad expression are “sticky”, for which we see no a priori defense. Moreover, this model can’t simply explain some of the above results, such as the stronger correlation for the less broadly expressed genes.

Prediction of the tissue of expression is moderate at best

Above we have asked if we can predict breadth from knowledge of TFBS. A possibly harder question, is whether contained within the promoter architecture is evidence of which tissue or tissues a narrowly expressed gene is expressed in (rather than just the narrowness of that expression). In principle a machine learning approach might be devised to employ the data that we have assembled to tackling such an issue. We took a similar approach to the SVM given above to predict tissues-specific expression (i.e., breadth of expression). The main difference was that the continuous response variable was the preferential expression measure (PEM) in a given tissue instead of the breadth of expression.

For a given transcript Y we consider its expression level in tissue X and divide that by Y’s mean expression in all tissues. This ratio is PEM for that transcript in that tissue. So a high PEM for gene Y in tissue X means it is preferentially expressed in tissue X. For all transcripts we then consider the average PEM for a given tissue (PEM avg ). This provides a metric of the degree to which genes are preferentially expressed in tissue X. We then ask about our ability to predict PEM values. To this end we train our SVM against PEM values and ask it then to predict PEM values of genes outside of the training set. If the SVM works, our predictor should correlate with our observed results. Within each tissue we then correlate predicted PEM for all genes against observed PEM for the same genes in the same tissue. These correlation scores are on the Y-axis in Figure 13.

Figure 13
figure 13

SVMs predict preferential expression for some tissues, such as the brain and the adipose tissue. The average preferential expression measure (PEM avg ) expresses the degree of preferential expression of genes expressed in a given tissue. An individual PEM value for each transcript equals its expression in a given tissue divided by its average expression across all tissues. This being a continuous variable, prediction accuracy is the correlation between SVM’s predicted value of the response variable and its observed value in the half of the dataset designated for prediction (the other half was used for training). The highest prediction accuracy was achieved for the adipose tissue and the brain tissue cluster (which were also the two tissues with the highest degree of tissue-specific expression). Overall, there was a strong correlation between the overall degree of preferential expression in a given tissue (PEM avg ) and the power to predict preferential expression in this tissue (Spearman’s correlation of 0.543). For technical details of the SVMs used see Methods.

We find that SVMs trained using a matrix of TF frequencies can at best modestly predict preferential expression for some tissues (such as the brain and the adipose tissue, the liver, the lung, and the breast) with the correlation between the predicted and the observed expression of up to r p  = 0.36 for brain (Figure 13) (note that as before the SVM is trained against continuous data so the appropriate metric of accuracy is a correlation coefficient). That brain was the best predicted may well reflect the fact that brain is also the tissue with the highest number of preferentially-expressed genes. Indeed, we see an overall correlation between predictive ability (correlation strength) and the mean degree of preferential expression (PEM avg ) for genes expressed in any given tissue (rho = 0.54). That is to say, the tissues for which we fail to predict the preferential expression of transcripts, are the tissues with few preferentially expressed transcripts (e.g. thyroid, salivary gland, skin, bone marrow). Overall, we conclude that at best we have only a moderate ability to predict degree of preferential expression of genes in any given tissue and that this diminishes greatly when the tissues themselves have few tissue specific genes.

An alternative binary classifier for the prediction of tissue-specific expression

An alternative approach to prediction of tissue-specific (i.e. narrow) expression is to divide the transcripts into two categories: tissue-specific and not tissue-specific, at the cutoff of the breadth of expression of 0.33 (i.e., one third of tissues). The approach taken here is identical to the one described above (see: A support vector machine method predicts the breadth of expression better than the correlation alone) except that the input and output measures of the breadth of expression were discretized. Standard validation charts for such a SVM-based binary classifier are shown in Additional file 13: Figure S9. The area under the ROC curve equaled 0.816. Inclusion of GC content does not improve this SVM’s predictive ability, confirming the irrelevance of this feature.

Discussion

The results above support the view that, to a considerable degree, components of the expression profile, notably expression breadth, and in turn expression breadth divergence, can be predicted from knowledge of the TF binders of a given gene. That such a result wasn’t, for the most part, captured previously, may be explained as limitations in the prior data. That we can recover the breadth of expression result using the previously available Gene Expression Atlas [[45]], suggests that it is the high resolution ENCODE transcription factor binding data that is key. Given that the “sticky” chromatin model fails to make a parsimonious explanation of the data, we conclude that the data support the suggestion that most eukaryotic TFs are activating.

Here, while touching on expression level, we have concentrated on expression breadth and divergence in breadth. Our results, however, suggest a series of further questions. How, for example, are we to interpret the evidence for cooperation between partners not previously known to be cooperative? Assuming antibody cross-reactivity is not the explanation, these statistically associated TFs suggest experimental tests for apparent cooperativity. Our analysis of whether it is possible to predict the tissue of expression of tissue specific genes suggests that a support vector approach has limited success at best. Given that we have the best success for the tissue (brain) with the most data, it may yet prove to be the case that with better techniques and more data this problem may become more tractable.

A further open issue is whether we can extrapolate our results to non-human species. An ability to infer mutational changes in promoters that are likely to have a major impact on expression, in poorly studied close relatives would be of considerable value in determining those promoters and TFs that might be (or have been) under selection in humans to switch gene expression “on” or “off” in tissues key to human uniqueness. Given the above result (or the moderate ability to predict brain specific genes) there may be some prospect of identifying some lineage specific changes that might have affected human brains.

Conclusions

We present evidence that expression breadth and paralog expression divergence are strongly predictable with knowledge of transcription factor binding in the proximal promoter. A simple metric, the number of binding transcription factors found on a promoter, is a robust predictor of expression breadth in human tissues. However, prediction of the tissue of expression is moderate at best.

Methods

Data sources

FANTOM5

Primary FANTOM5 CAGE data [[46]] (see: [[47]]) were processed by the consortium Work-package 4 (WP4) to produce expression tables, mapping CAGE tags to genes, contained in the following three files:

(T) human.tissue.hCAGE.hg19.tpm.refgene.osc.txt - human tissues;

(PC) human.primary_cell.hCAGE.hg19.tpm.refgene.osc.txt - human primary cells;

(CCL) human.cell_line.hCAGE.hg19.tpm.refgene.osc.txt - human cancer cell lines.

These were tab-delimited text files with the header section describing columns, library names, and the total number of reads for each library. These files were pre-processed using standard methods to facilitate their import to the R statistical environment with the read.table command with row-names being RefSeq accession numbers and column-names identifying tissue, cell-type, or cell line from which CAGE tags were sequenced.

ENCODE

ChIP-seq data were described in the section below (Assembling TF binding data to define promoter architecture). The links to the files and description of inputs can be found in the following webpage references [[25],[26],[31],[32]]. ENCODE Dnase sites, which mark regions of accessible chromatin across the ENCODE set of cell-lines [[48],[49]], and methylation signal for the HeLa cell line were also retrieved from ENCODE [[50]].

Gene Expression Atlas

Gene Expression Atlas (GEA) data [[51]] were used to independently confirm the correlation between the number of mapping transcription factors and the breadth of expression. We see this as valuable that this strong trend could be detected using two very different expression technologies: one based on next generation sequencing (FANTOM5) and a second one based on microarray hybridization (GEA).

Assembling TF binding data to define promoter architecture

The 2011 meta data set included 2,750,490 ChIP-seq peaks for 148 transcription factors, derived from seventy-one cell-types with twenty-four additional experimental cell culture conditions [[25]]. Peaks were called and merged using UCSC clustering tools (encodeMergeReplicates, regClusterBedExpCfg and hgBedsToBedExps). Crucially, this procedure merges peaks for the same TF across replicates, cell lines, and from different labs into one. Peak scores varied between zero to 1000 (proportionately to the Tfbs prediction reliability). We used either all data, or only high-quality peaks with the score over half the maximum (i.e., 500) to avoid noisy data associated with multi-mapping next-generation sequencing tags. For the correlation between the breadth of expression and the number of transcription factor binding sites, we verified that results analogous to those for the January, 2011, data-freeze [[25]] were obtained using a broader September, 2012, data-freeze [[26]]. The 2012 data-freeze consisted of 161 transcription factors and ninety-one human cell types with various treatment conditions [[32]]. The ENCODE preprocessing pipeline merges all overlapping binding sites for a TF in different cell lines into a single site. Therefore, the cumulative metric which we calculate corresponds to the total capacity of a promoter to bind TFs (i.e., promoter architecture), rather than to the actual number of sites in any particular cell- or tissue-type or across the sample space. This is crucial, as our approach relies on defining the promoter architecture, that is to say the binding landscape, rather than summing up Tfbs across the tissues which would be dangerously tautologous to measuring breadth of expression. However, we took great care not to fall into this trap. On top of the ENCODE’s merging procedure, we employ an additional safeguard by introducing a metric which only counts each TF once even if the promoter can bind this TF at alternative genomic locations (i.e., Tfbs_x_length_unique). The correlation between Tfbs_x_length and Tfbs_x_length_unique was almost perfect (r p  = 0.9929) and these two measures supported identical biological conclusions in all analyses (ensuring that no single Tfbs could introduce a bias).

The Jaccard index for promoter divergence

The Jaccard index (JI) measured the overlap between two sets of genomic features [[52]]. The index was calculated as the ratio of the intersection over the union (JI = I/U). JI equaled one when the intersection equaled the union (that is when there was a perfect overlap). JI equaled zero when there was no overlap.

FANTOM5 gene expression tables

CAGE tags were mapped to RefSeq transcripts +/− 500 bps from their transcription start sites. The numbers of tags were normalized to tags per million (TPM). Finally, the TPM value of ten was chosen as the default cutoff for a gene to be “on” (unless stated otherwise). We downloaded RefSeq in BED format from the hg19 UCSC genome browser using the table browser tool. This dataset included 40,856 human transcripts, including messenger RNAs (NM-accession) and non-coding transcripts (NR-accession). Non-coding transcripts included structural RNAs and transcribed pseudogenes. The dataset was later processed with BEDtools [[53]].

Gene Expression Atlas

Gene Expression Atlas [[45]] was employed to confirm the correlation between the breadth of expression and the number of transcription factor binding sites in proximal promoters (Figure 5d). Affymetrix average difference (AD) higher that 200 classified a gene as “on” or expressed in a given tissue. Affymetrix ids were mapped to RefSeq ids using R annotation object hgu95aREFSEQ from the hgu95a.db package. ENCODE transcription factors were mapped to RefSeq just as in the FANTOM5 analysis.

TreeFam and the inference of gene duplications

TreeFam [[54]] used evolutionary histories of individual genes to construct phylogenetic trees and to date gene duplications [[54]]. In our hands, TreeFam consistently delivered high quality phylogenetic trees which were congruent with the insights of molecular biologists [[55],[56]].

TreeBeST was the tree-building engine behind TreeFam. TreeBeST merged a maximum likelihood tree from PHYML [[57]] with neighbor-joining trees based on P-distance, Ka, and Ks. TreeBeST used smart heuristics intended to maximize the similarity between the gene tree and the species tree and to minimize the number of predicted gene duplications and losses.

TreeFam release eight was based on Ensembl version 54 and included 79 species, 1,539,621 genes, and 16,064 families. Phylogenetic timing was used to associate gene duplications with the emergence of different taxa. The Vertebrata and the Bilateria were consistently linked with the most numerous waves of duplications in animals [[56]]. The algorithm for speciation and duplication inference (SDI) reconciled gene trees with the species dendrogram. SDI also inferred duplications, speciation events, paralogy, and orthology.

The comparison of all paralog pairs versus the comparison of youngest pairs only

Combinatorics explains why there are more possible paralog comparisons than paralogs. In a set of n elements, the number of k-combinations was equal to the binomial coefficient ("n choose k"). K equals two for pairwise comparisons. For example, there were ten possible pairwise comparisons in a family of five paralogs. There were 45 legitimate pairwise comparisons in a family of ten paralogs. However, these comparisons might not be regarded as independent data points, and the results might have been biased by a few large gene families. The best alternative was to perform a comparison between each paralog with only its closest relative (i.e., only the most recent duplicate).

The duplicator

The dataset used in the preparation of this study was released to the public domain as an R package called the Duplicator [[58]]. R data-frames with data on duplications mapping to different taxa can be found in R helper environments for the package (env_duplicator_base and env_duplicator_vectors), which are located in the path duplicator/data/duplicator. Naming conventions indicate the species of origin. For instance, the R data-frames containing human data are called: dupEvent12_genes_LL_hs (individual genes) and dupEvent12_genes_LL_hs2 (paralog pairs).

Data-frames in the package follow the denormalized data model. For example, the R data-frame dupEvent12_genes_LL_hs2 has the following fields: family (TreeFam family ID), node (unique identifier for each gene duplication node), taxon (taxon of duplication), gene.x (ENSEMBL ENSG ID for paralog x), familySide.x, primary_acc.x (Entrez IDs for paralog x), gene.y (ENSEMBL ENSG ID for paralog y), familySide.y, primary_acc.y (Entrez IDs for paralog y). FamilySide was a flag with values of one or two defining on which side of the duplication node the gene was located; this flag was used to prevent multiple comparisons on the “same side” of the primary duplication node if later duplications occurred.

The analytical pipeline

The analytical pipeline consisted of three stages: (A) data retrieval and remodeling; (B) the detection of the overlap between promoters and transcription factor binding sites; and (C) parsing, statistical tests, and figure generation.

  1. (A)

    Data retrieval and remodeling (the pre-BEDtools stage). This stage consisted of data retrieval from TreeFam and remodeling into R data-frames. Helper environments held these data-frames for later use, and stored intermediate results.

  2. (B)

    The detection of the overlap between promoters and transcription factor binding sites (the BEDtools stage). We checked if two sets of genomic features overlapped using BEDtools. We used mostly coverageBed and intersectBed. CoverageBed calculates the depth of coverage of features in the file A against the file B. Herein, coverageBed was used to calculate the depth of coverage by ENCODE transcription factor binding sites in promoters (coverageBed -a ENCODE_Tfbs -b RefSeq_promoters > result). IntersectBed writes the original entry in the file A for each overlap when used with the “-wa” option (intersectBed -a ENCODE_Tfbs -b RefSeq_promoters -wa > result). This identified the exact type of transcription factor binding sites in the overlap. An alternative analysis pipeline was constructed using R/BioC packages GenomicRanges and IRanges. However, the R/BioC pipeline proved prohibitively slow even on SNIC Supercomputers. The alternative pipeline was used only to verify the results of the main BEDtools-based pipeline.

  3. (C)

    Parsing, statistical tests, and figure generation (the post-BEDtools stage). Post-processing was performed in R and Bioconductor (version 2.11). Standard Bioconductor packages such as Biodist, gplot, ggplot, rtracklayer, TxDb.Hsapiens.UCSC.hg19.knownGene, and GOstats were used.

The post-BEDtoolsanalysis was divided into the four steps listed below.

Step (1): Parsing BEDtools output into an R list.

BEDtools output in the browser extensible data (BED) file format was read into R as a data-frame. The data-frame was then remodeled into an R list called resJaccard_ENCODE_Tfbs_substrate. Each element of the list was indexed with a RefSeq transcript id as the accession key. A vector of ENCODE transcription factor binding sites was contained within each element of the list. Transcription factor parsing and creation of the list were performed by the script called env_bed_jaccard_make_1.R. Intermediate results were stored in the R helper environment named env_promoter_bed_500_res_env.

Step (2): The Jaccard index calculation.

The Jaccard index was calculated using the script resJaccard_ ENCODE_Tfbs_substrate. Additionally, a randomized dataset was generated using sampling without replacement. The randomized dataset was used to calculate a control distribution of the Jaccard index.

Step (3): Additional calculations.

The following additional variables were calculated: relative frequencies of different transcription factor binding sites, intersection, union, and the Jaccard index depending on the taxon of duplication. These calculations were performed by scripts called env_jaccard_analyse_result_bed_sampled.R and env_jaccard_analyse_random_bed_sampled.R.

Step (4): Figure generation.

Manuscript figures were generated automatically using ggplot2 from intermediate results stored in R helper environments.

The support vector machine

The support vector machine is a commonly used machine learning technique for regression. We used the R package e1071, an R implementation of libsvm. Each row of the basic SVM training data-frame was a vector representing the number of Tfbs for each gene; although, we also considered SVMs trained using promoter GC- and CpG-contents (Table 5). The training dataset was scaled and centered. We used a radial kernel as a regression machine. After a grid search for optimal parameter values, the following SVM parameters were used: cost = 1, gamma = 0.01, and epsilon = 0.1. The continuous response variable was either the breadth of expression (BoE) or the preferential expression measure (PEM).

Data access

Access to FANTOM5 is provided at the FANTOM5 public website, including the UCSC genome browser mirror, and FANTOM5’s own CAGE-focused ZENBU genome browser.

Additional files

References

  1. Makova KD, Li WH: Divergence in the spatial pattern of gene expression between human duplicate genes. Genome Res. 2003, 13: 1638-1645. 10.1101/gr.1133803.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  2. Huminiecki L, Wolfe KH: Divergence of spatial gene expression profiles following species-specific gene duplications in human and mouse. Genome Res. 2004, 14: 1870-1879. 10.1101/gr.2705204.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  3. Struhl K: Fundamentally different logic of gene regulation in eukaryotes and prokaryotes. Cell. 1999, 98: 1-4. 10.1016/S0092-8674(00)80599-1.

    Article  PubMed  CAS  Google Scholar 

  4. Ptashne MGA: Genes and signals. 2002, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York

    Google Scholar 

  5. Segal E, Raveh-Sadka T, Schroeder M, Unnerstall U, Gaul U: Predicting expression patterns from regulatory sequence in Drosophila segmentation. Nature. 2008, 451: 535-U531. 10.1038/nature06496.

    Article  PubMed  CAS  Google Scholar 

  6. Tirosh I, Weinberger A, Bezalel D, Kaganovich M, Barkai N: On the relation between promoter divergence and gene expression evolution. Mol Syst Biol. 2008, 4: 159-

    Article  PubMed  PubMed Central  Google Scholar 

  7. Ge Y, Porse BT: The functional consequences of intron retention: Alternative splicing coupled to NMD as a regulator of gene expression. Bioessays. 2013, 36: 236-243.

    Article  PubMed  Google Scholar 

  8. Stroynowska-Czerwinska A, Fiszer A, Krzyzosiak WJ: The panorama of miRNA-mediated mechanisms in mammalian cells. Cell Mol Life Sci. 2014, 71: 2253-2270. 10.1007/s00018-013-1551-6.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  9. Cheadle C, Fan J, Cho-Chung YS, Werner T, Ray J, Do L, Gorospe M, Becker KG: Control of gene expression during T cell activation: alternate regulation of mRNA transcription and mRNA stability. BMC Genomics. 2005, 6: 75-10.1186/1471-2164-6-75.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Gilbert N, Boyle S, Fiegler H, Woodfine K, Carter NP, Bickmore WA: Chromatin architecture of the human genome: Gene-rich domains are enriched in open chromatin fibers. Cell. 2004, 118: 555-566. 10.1016/j.cell.2004.08.011.

    Article  PubMed  CAS  Google Scholar 

  11. Chen T, Dent SY: Chromatin modifiers and remodellers: regulators of cellular differentiation. Nat Rev Genet. 2014, 15: 93-106.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  12. Agalioti T, Chen G, Thanos D: Deciphering the transcriptional histone acetylation code for a human gene. Cell. 2002, 111: 381-392. 10.1016/S0092-8674(02)01077-2.

    Article  PubMed  CAS  Google Scholar 

  13. Gierman HJ, Indemans MH, Koster J, Goetze S, Seppen J, Geerts D, van Driel R, Versteeg R: Domain-wide regulation of gene expression in the human genome. Genome Res. 2007, 17: 1286-1295. 10.1101/gr.6276007.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  14. Raj A, Peskin CS, Tranchina D, Vargas DY, Tyagi S: Stochastic mRNA synthesis in mammalian cells. Plos Biology. 2006, 4: e309-10.1371/journal.pbio.0040309.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Ebisuya M, Yamamoto T, Nakajima M, Nishida E: Ripples from neighbouring transcription. Nat Cell Biol. 2008, 10: 1106-1113. 10.1038/ncb1771.

    Article  PubMed  CAS  Google Scholar 

  16. Batada NN, Urrutia AO, Hurst LD: Chromatin remodelling is a major source of coexpression of linked genes in yeast. Trends in genetics : TIG. 2007, 23: 480-484. 10.1016/j.tig.2007.08.003.

    Article  PubMed  CAS  Google Scholar 

  17. Park J, Xu K, Park T, Yi SV: What are the determinants of gene expression levels and breadths in the human genome?. Hum Mol Genet. 2012, 21: 46-56. 10.1093/hmg/ddr436.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Illingworth RS, Bird AP: CpG islands–'a rough guide'. FEBS Lett. 2009, 583: 1713-1720. 10.1016/j.febslet.2009.04.012.

    Article  PubMed  CAS  Google Scholar 

  19. Borneman AR, Gianoulis TA, Zhang ZD, Yu H, Rozowsky J, Seringhaus MR, Wang LY, Gerstein M, Snyder M: Divergence of transcription factor binding sites across related yeast species. Science. 2007, 317: 815-819. 10.1126/science.1140748.

    Article  PubMed  CAS  Google Scholar 

  20. Park C, Makova KD: Coding region structural heterogeneity and turnover of transcription start sites contribute to divergence in expression between duplicate genes. Genome Biol. 2009, 10: R10-10.1186/gb-2009-10-1-r10.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Zhang Z, Gu J, Gu X: How much expression divergence after yeast gene duplication could be explained by regulatory motif evolution?. Trends in genetics : TIG. 2004, 20: 403-407. 10.1016/j.tig.2004.07.006.

    Article  PubMed  Google Scholar 

  22. Papp B, Pal C, Hurst LD: Evolution of cis-regulatory elements in duplicated genes of yeast. Trends in genetics : TIG. 2003, 19: 417-422. 10.1016/S0168-9525(03)00174-4.

    Article  PubMed  CAS  Google Scholar 

  23. Dunham I, Kundaje A, Aldred SF, Collins PJ, Davis CA, Doyle F, Epstein CB, Frietze S, Harrow J, Kaul R, Khatun J, Lajoie BR, Landt SG, Lee BK, Pauli F, Rosenbloom KR, Sabo P, Safi A, Sanyal A, Shoresh N, Simon JM, Song L, Trinklein ND, Altshuler RC, Birney E, Brown JB, Cheng C, Djebali S, Dong X, Ernst J, et al: An integrated encyclopedia of DNA elements in the human genome. Nature. 2012, 489: 57-74. 10.1038/nature11247.

    Article  CAS  Google Scholar 

  24. Landt SG, Marinov GK, Kundaje A, Kheradpour P, Pauli F, Batzoglou S, Bernstein BE, Bickel P, Brown JB, Cayting P, Chen Y, DeSalvo G, Epstein C, Fisher-Aylor KI, Euskirchen G, Gerstein M, Gertz J, Hartemink AJ, Hoffman MM, Iyer VR, Jung YL, Karmakar S, Kellis M, Kharchenko PV, Li Q, Liu T, Liu XS, Ma L, Milosavljevic A, Myers RM, et al: ChIP-seq guidelines and practices of the ENCODE and modENCODE consortia. Genome Res. 2012, 22: 1813-1831. 10.1101/gr.136184.111.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  25. ENCODE-TfbsV2. ᅟ, ᅟ, ᅟ

  26. ENCODE-TfbsV3. ᅟ, ᅟ, ᅟ

  27. Forrest AR, Kawaji H, Rehli M, Baillie JK, de Hoon MJ, Lassmann T, Itoh M, Summers KM, Suzuki H, Daub CO, Kawai J, Heutink P, Hide W, Freeman TC, Lenhard B, Bajic VB, Taylor MS, Makeev VJ, Sandelin A, Hume DA, Carninci P, Hayashizaki Y: A promoter-level mammalian expression atlas. Nature. 2014, 507: 462-470. 10.1038/nature13182.

    Article  PubMed  CAS  Google Scholar 

  28. Velculescu VE, Madden SL, Zhang L, Lash AE, Yu J, Rago C, Lal A, Wang CJ, Beaudry GA, Ciriello KM, Cook BP, Dufault MR, Ferguson AT, Gao Y, He TC, Hermeking H, Hiraldo SK, Hwang PM, Lopez MA, Luderer HF, Mathews B, Petroziello JM, Polyak K, Zawel L, Kinzler KW: Analysis of human transcriptomes. Nat Genet. 1999, 23: 387-388.

    Article  PubMed  CAS  Google Scholar 

  29. Wang J, Zhuang J, Iyer S, Lin X, Whitfield TW, Greven MC, Pierce BG, Dong X, Kundaje A, Cheng Y, Rando OJ, Birney E, Myers RM, Noble WS, Snyder M, Weng Z: Sequence features and chromatin structure around the genomic regions bound by 119 human transcription factors. Genome Res. 2012, 22: 1798-1812. 10.1101/gr.139105.112.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  30. Spivakov M, Akhtar J, Kheradpour P, Beal K, Girardot C, Koscielny G, Herrero J, Kellis M, Furlong EE, Birney E: Analysis of variation at transcription factor binding sites in Drosophila and humans. Genome Biol. 2012, 13: R49-10.1186/gb-2012-13-9-r49.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  31. ENCODE-InputsV2. ᅟ, ᅟ, ᅟ

  32. ENCODE-InputsV3. ᅟ, ᅟ, ᅟ

  33. Makova KD, Li WH: Divergence in the spatial pattern of gene expression between human duplicate genes. Genome Res. 2003, 13: 1638-1645. 10.1101/gr.1133803.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  34. Huminiecki L, Wolfe KH: Divergence of spatial gene expression profiles following species-specific gene duplications in human and mouse. Genome Res. 2004, 14: 1870-1879. 10.1101/gr.2705204.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  35. Jordan IK, Marino-Ramirez L, Koonin EV: Evolutionary significance of gene expression divergence. Gene. 2005, 345: 119-126. 10.1016/j.gene.2004.11.034.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  36. Pereira V, Waxman D, Eyre-Walker A: A problem with the correlation coefficient as a measure of gene expression divergence. Genetics. 2009, 183: 1597-1600. 10.1534/genetics.109.110247.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Piasecka B, Robinson-Rechavi M, Bergmann S: Correcting for the bias due to expression specificity improves the estimation of constrained evolution of expression between mouse and human. Bioinformatics. 2012, 28: 1865-1872. 10.1093/bioinformatics/bts266.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  38. He X, Zhang J: Higher duplicability of less important genes in yeast genomes. Mol Biol Evol. 2006, 23: 144-151.

    Article  PubMed  CAS  Google Scholar 

  39. Woods S, Coghlan A, Rivers D, Warnecke T, Jeffries SJ, Kwon T, Rogers A, Hurst LD, Ahringer J: Duplication and Retention Biases of Essential and Non-Essential Genes Revealed by Systematic Knockdown Analyses. Plos Genetics. 2013, 9: e1003330-10.1371/journal.pgen.1003330.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  40. Gu Z, Nicolae D, Lu HH, Li WH: Rapid divergence in expression between duplicate genes inferred from microarray data. Trends in genetics : TIG. 2002, 18: 609-613. 10.1016/S0168-9525(02)02837-8.

    Article  PubMed  CAS  Google Scholar 

  41. Loh YH, Wu Q, Chew JL, Vega VB, Zhang W, Chen X, Bourque G, George J, Leong B, Liu J, Wong KY, Sung KW, Lee CW, Zhao XD, Chiu KP, Lipovich L, Kuznetsov VA, Robson P, Stanton LW, Wei CL, Ruan Y, Lim B, Ng HH: The Oct4 and Nanog transcription network regulates pluripotency in mouse embryonic stem cells. Nat Genet. 2006, 38: 431-440. 10.1038/ng1760.

    Article  PubMed  CAS  Google Scholar 

  42. Gause M, Schaaf CA, Dorsett D: Cohesin and CTCF: cooperating to control chromosome conformation?. Bioessays. 2008, 30: 715-718. 10.1002/bies.20787.

    Article  PubMed  CAS  Google Scholar 

  43. Wendt KS, Peters JM: How cohesin and CTCF cooperate in regulating gene expression. Chromosome Res. 2009, 17: 201-214. 10.1007/s10577-008-9017-7.

    Article  PubMed  CAS  Google Scholar 

  44. Factorbook. ᅟ, ᅟ, ᅟ

  45. Su AI, Cooke MP, Ching KA, Hakak Y, Walker JR, Wiltshire T, Orth AP, Vega RG, Sapinoso LM, Moqrich A, Patapoutian A, Hampton GM, Schultz PG, Hogenesch JB: Large-scale analysis of the human and mouse transcriptomes. Proc Natl Acad Sci U S A. 2002, 99: 4465-4470. 10.1073/pnas.012025199.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  46. WP4 expression tables. ᅟ, ᅟ, ᅟ

  47. Primary FANTOM5 CAGE data. ᅟ, ᅟ, ᅟ

  48. Dnase sites slustered. ᅟ, ᅟ, ᅟ

  49. Dnase-Inputs. ᅟ, ᅟ, ᅟ

  50. Methylation in HeLa. ᅟ, ᅟ, ᅟ

  51. Su AI, Wiltshire T, Batalov S, Lapp H, Ching KA, Block D, Zhang J, Soden R, Hayakawa M, Kreiman G, Cooke MP, Walker JR, Hogenesch JB: A gene atlas of the mouse and human protein-encoding transcriptomes. Proc Natl Acad Sci U S A. 2004, 101: 6062-6067. 10.1073/pnas.0400782101.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  52. Favorov A, Mularoni L, Cope LM, Medvedeva Y, Mironov AA, Makeev VJ, Wheelan SJ: Exploring massive, genome scale datasets with the GenometriCorr package. PLoS Comput Biol. 2012, 8: e1002529-10.1371/journal.pcbi.1002529.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  53. Quinlan AR, Hall IM: BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010, 26: 841-842. 10.1093/bioinformatics/btq033.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  54. Li H, Coghlan A, Ruan J, Coin LJ, Heriche JK, Osmotherly L, Li R, Liu T, Zhang Z, Bolund L, Wong GK, Zheng W, Dehal P, Wang J, Durbin R: TreeFam: a curated database of phylogenetic trees of animal gene families. Nucleic Acids Res. 2006, 34: D572-D580. 10.1093/nar/gkj118.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  55. Huminiecki L, Goldovsky L, Freilich S, Moustakas A, Ouzounis C, Heldin CH: Emergence, development and diversification of the TGF-beta signalling pathway within the animal kingdom. BMC Evol Biol. 2009, 9: 28-10.1186/1471-2148-9-28.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Huminiecki L, Heldin CH: 2R and remodeling of vertebrate signal transduction engine. BMC Biol. 2010, 8: 146-10.1186/1741-7007-8-146.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Guindon S, Gascuel O: A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol. 2003, 52: 696-704. 10.1080/10635150390235520.

    Article  PubMed  Google Scholar 

  58. The Duplicator. ᅟ, ᅟ, ᅟ

Download references

Acknowledgments and funding

Lukasz Huminiecki was partially funded by Bioinformatics Infrastructure for Life Sciences (BILS). BILS was a distributed national research infrastructure supported by the Swedish Research Council (Vetenskapsrådet), providing bioinformatics support to life science researchers in Sweden. FANTOM5 was made possible by a Research Grant for RIKEN Omics Science Center from MEXT to Yoshihide Hayashizaki and a Grant of the Innovative Cell Biology by Innovative Technology (Cell Innovation Program) from the MEXT, Japan to YH. We would like to thank all members of the FANTOM5 consortium for contributing to generation of samples and analysis of the data-set and thank GeNAS for data production. The Swedish National Infrastructure for Computing (SNIC) provided computational resources for the Swedish research community. Our computations were performed on SNIC resources, UPPMAX and kappa.

Author information

Authors and Affiliations

Authors

Consortia

Corresponding author

Correspondence to Lukasz Huminiecki.

Additional information

Competing interests

The authors declare no competing interests.

Authors’ contributions

LH and LDH designed the study (mostly LH) and wrote the manuscript. LH performed all analyses. ARRF and CD were involved in the FANTOM5 concepts and management. OS provided technical assistance for some of the analyses. All authors read and approved the final manuscript.

Electronic supplementary material

13059_2014_413_MOESM1_ESM.pdf

Additional file 1: Figure S7 Promoter GC-content was a place marker distinct from isochore GC-content or GC3. This figure consists of four parts. In the upper-left part, we show that isochore GC-content and GC3 corresponded closely to each other. However, promoter GC-content was quite distinct from the isochore GC-content with a large population of high GC-promoters located in low-GC chromosomal regions (the upper-right part). There was no simple direct relationship between methylation and BoE (see bottom-left, the red line is the fitted loess curve). Low-GC promoters formed a clearly distinct group from high-GC promoters in terms of average BoE when a correlation between core promoter-GC and BoE was plotted on a scatterplot (bottom-right, the blue line is the fitted loess curve). (PDF 4911 kb) (PDF 5 MB)

13059_2014_413_MOESM2_ESM.txt

Additional file 2: Table S1 There is little or no evidence for a class of genes so highly broadly expressed that they dispense with TFs all-together. There were only thirty-nine broadly expressed transcripts (the breadth of expression in human tissues > 0.9) with fewer than ten TFs in the promoter (defined as a broad 10 kb window around the TSS). Concerted action of multiple TFs appears necessary for housekeeping expression. The columns of the table encode as follows: RefSeq, EntrezID, gene symbol, and gene name. (TXT 2 kb) (TXT 2 KB)

13059_2014_413_MOESM3_ESM.pdf

Additional file 3: Figure S1 The relationship between the breadth of expression, the mean and the maximum expression in human primary cells, and the number of transcription factor binding sites. This figure consists of sixteen parts identified as (a) - (p). Four measures related to the breadth of expression were considered: (a-b-c-d) the breadth of expression at the cutoff of ten TPM, (e-f-g-h) the breadth of expression at the cutoff of 100 TPM, (i-j-k-l) the mean expression, and (m-n-o-p) the maximum expression. The number of transcription factor binding sites was estimated in four different approaches: (a-e-i-m) the total number, (b-f-j-n) the number of unique binding sites, (c-g-k-o) the total number excluding polymerase binding sites, and (d-h-l-p) the number of unique binding sites excluding the polymerase. The red line signified the linear model for the smoother line, while the blue line signified the non-linear model. This figure confirms the robustness of the findings presented in Figure 6 across the FANTOM5 sample space (i.e., in human primary cells). (PDF 11913 kb) (PDF 12 MB)

13059_2014_413_MOESM4_ESM.pdf

Additional file 4: Figure S2 The relationship between the breadth of expression, the mean and the maximum expression in human cancer cell lines, and the number of transcription factor binding sites. This figure consists of sixteen parts identified as (a) - (p). Four measures related to the breadth of expression were considered: (a-b-c-d) the breadth of expression at the cutoff of ten TPM, (e-f-g-h) the breadth of expression at the cutoff of 100 TPM, (i-j-k-l) the mean expression, and (m-n-o-p) maximum expression. The number of transcription factor binding sites was estimated in four different approaches: (a-e-i-m) the total number, (b-f-j-n) the number of unique binding sites, (c-g-k-o) the total number excluding polymerase binding sites, and (d-h-l-p) the number of unique binding sites excluding the polymerase. The red line signified the linear model for the smoother line, while the blue line signified the non-linear model. This figure confirms the robustness of the findings presented in Figure 6 across the FANTOM5 sample space (i.e., in human cancer cell lines). (PDF 11905 kb) (PDF 12 MB)

13059_2014_413_MOESM5_ESM.pdf

Additional file 5: Table S3 P-values for pairwise BoE comparisons using Wilcoxon rank sum test for data in Table 9. NOTE: P-value adjustment method: holm. (PDF 41 kb) (PDF 41 KB)

13059_2014_413_MOESM6_ESM.pdf

Additional file 6: Table S4 P-values for pairwise TfbsNo. comparisons using Wilcoxon rank sum test for data in Table 10. NOTE: P-value adjustment method: holm. (PDF 42 kb) (PDF 42 KB)

13059_2014_413_MOESM7_ESM.pdf

Additional file 7: Figure S5 The clustering of the breadth of expression with the number of transcription factor binding sites in human tissues (Kendall rank correlation coefficient). Transcription factors clustering closest with the breadth of expression were marked as A. Other transcription factors formed two clusters with low and high distance to the breadth of expression (these clusters were marked as B and C, respectively). To test the robustness of this analysis, the number of transcription factors was measured in several different ways, which reassuringly clustered together and proved indistinguishable. The different measures were: sum of all sites (marked as Tfbs_x_length), sum of unique sites (Tfbs_x_length_unique), sum of all sites without RNA polymerase II (Tfbs_x_length_noPol2), and finally the sum of unique sites without the polymerase (Tfbs_x_length_unique_noPol2). The breadth of expression was also transformed in several ways which proved equivalent by forming a tight cluster. Namely, the breadth of expression was encoded as either a continuous variable (marked as breadth_continuous), discretized into three bins (breadth_discrete_3), discretized into ten bins (breadth_discrete_10), or transformed and expressed as a square root (breadth_sqrt_1). (PDF 885 kb) (PDF 885 KB)

13059_2014_413_MOESM8_ESM.pdf

Additional file 8: Figure S6 The clustering of the breadth of expression with the number of transcription factor binding sites in human tissues (Spearman's rank correlation coefficient). Transcription factors clustering closest with the breadth of expression were marked as A. Other transcription factors formed two clusters with low and high distance to the breadth of expression (these clusters were marked as B and C, respectively). To test the robustness of this analysis, the number of transcription factors was measured in several different ways, which reassuringly clustered together and proved indistinguishable. The different measures were: sum of all sites (marked as Tfbs_x_length), sum of unique sites (Tfbs_x_length_unique), sum of all sites without RNA polymerase II (Tfbs_x_length_noPol2), and finally the sum of unique sites without the polymerase (Tfbs_x_length_unique_noPol2). The breadth of expression was also transformed in several ways which proved equivalent by forming a tight cluster. Namely, the breadth of expression was encoded as either a continuous variable (marked as breadth_continuous), discretized into three bins (breadth_discrete_3), discretized into ten bins (breadth_discrete_10), or transformed and expressed as a square root (breadth_sqrt_1). (PDF 885 kb) (PDF 886 KB)

13059_2014_413_MOESM9_ESM.txt

Additional file 9: Table S2 Transcription factor frequencies in the four classes of transcripts (A, B, C, and D) in respect to the breadth of expression and the number of TFs. Because of the large number of rows, this is a supplementary table available as a text file in the supplementary material. NOTE: The four classes of transcripts were illustrated in Figure 11. Typical broadly expressed Tfbs-rich genes were summarized in column A. Unusual broadly expressed Tfbs-poor genes were summarized in column B. Unusual tissue-specific Tfbs-rich genes were summarized in column C. Typical tissue-specific Tfbs-poor genes were summarized in column D. The numbers of transcripts in respective classes were 7824, 3206, 3593, and 16170. The frequencies of transcription factors were expressed as fractions of all sites in a given class. The total sums of Tfbs in A, B, C and D were: 146863, 17336, 58978, and 32134. The mean numbers of Tfbs per a transcript in each class were: 18.77, 5.41, 16.41, and 1.99. P-values of Fisher’s test were given. All data were calculated for FANTOM5 human tissues. Future work should facilitate classification of all transcription factors and their impact on individual genes as either activatory or inhibitory. (TXT 4 kb) (TXT 4 KB)

13059_2014_413_MOESM10_ESM.pdf

Additional file 10: Figure S3 The clustering of the breadth of expression with the number of transcription factor binding sites in human primary cells. In human primary cells, the breadth of expression clustered with the polymerase, TFIID, NFKB, HEY1, Sin3A, and c − Myc (this cluster was marked as A). Other transcription factors formed two clusters with low and high distance to the breadth of expression (these clusters were marked as B and C, respectively). To test the robustness of this analysis, the number of transcription factors was measured in several different ways, which reassuringly clustered together and proved indistinguishable. The different measures were: sum of all sites (marked as Tfbs_x_length), sum of unique sites (Tfbs_x_length_unique), sum of all sites without RNA polymerase II (Tfbs_x_length_noPol2), and finally the sum of unique sites without the polymerase (Tfbs_x_length_unique_noPol2). The breadth of expression was also transformed in several ways which proved equivalent by forming a tight cluster. Namely, the breadth of expression was encoded as either a continuous variable (marked as breadth_continuous), discretized into three bins (breadth_discrete_3), discretized into ten bins (breadth_discrete_10), or transformed and expressed as a square root (breadth_sqrt_1). (PDF 885 kb) (PDF 886 KB)

13059_2014_413_MOESM11_ESM.pdf

Additional file 11: Figure S4 The clustering of the breadth of expression with the number of transcription factor binding sites in human cancer cell lines. In cancer cell lines, the breadth of expression only clustered with RNA polymerase II and TFIID (this cluster was marked as A) suggesting that cancerous transformation disables most normal control switches for the breadth of expression. Other transcription factors formed two clusters with low and high distance to the breadth of expression (these clusters were marked as B and C, respectively). To test the robustness of this analysis, the number of transcription factors was measured in several different ways, which reassuringly clustered together and proved indistinguishable. The different measures were: sum of all sites (marked as Tfbs_x_length), sum of unique sites (Tfbs_x_length_unique), sum of all sites without RNA polymerase II (Tfbs_x_length_noPol2), and finally the sum of unique sites without the polymerase (Tfbs_x_length_unique_noPol2). The breadth of expression was also transformed in several ways which proved equivalent by forming a tight cluster. Namely, the breadth of expression was encoded as either a continuous variable (marked as breadth_continuous), discretized into three bins (breadth_discrete_3), discretized into ten bins (breadth_discrete_10), or transformed and expressed as a square root (breadth_sqrt_1). (PDF 889 kb) (PDF 889 KB)

13059_2014_413_MOESM12_ESM.pdf

Additional file 12: Figure S8 There was a positive correlation between the breadth of expression of transcription factors and the average breadth of expression of their targets. The breadth of expression of transcription factors in FANTOM5 tissues was plotted on the X-axis of the scatterplot (signified by Tfbs_BoE). The unweighted mean of the breadth of expression of all target genes (i.e. all genes that have a given Tfbs in their proximal promoter) was plotted on the Y-axis (signified by Mean_target_BoE). As both the independent and dependent variables were highly non-normally distributed, we used non-parametric correlation (Spearman’s rho = 0.3176). An alternative measure, a weighted mean, in which that data points were weighted according to the actual number of transcription factor binding sites was also considered and gave almost exactly the same correlation. The blue line is the fitted loess curve. (PDF 7 kb) (PDF 7 KB)

13059_2014_413_MOESM13_ESM.pdf

Additional file 13: Figure S9 Performance of SVM models as the predictor of tissue-specific (i.e. narrow) expression. Tissue-specific expression was defined as breadth of expression lower than 0.33 (i.e., a gene that was expressed in less than one third of tissues) and transcripts were categorized as either tissue-specific or not (i.e. in a binary classification). This figure consists of four panels identified as a through d. The panels display standard predictor validation charts for the basic SVM model (SVM-Tfbs): (a) the ROC curve, (b) the precision/recall graph, (c) the sensitivity/specificity plot, and (d) the lift chart. The curves were averages from 10 different cross-validation runs. The area under the ROC curve equaled 0.816 (standard deviation equaled 0.002325). The parameterization of the curves was performed using the value of the linear SVM output and visualized by printing cutoff values at the corresponding curve positions (the curve was also colored according to the cutoff). The curves were plotted using R package ROCR. Essentially identical results were obtained a more complex SVM model with added data on GC-content (SVM-Tfbs + GC), with AUC = 0.816. (PDF 210 kb) (PDF 211 KB)

Authors’ original submitted files for images

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hurst, L.D., Sachenkova, O., Daub, C. et al. A simple metric of promoter architecture robustly predicts expression breadth of human genes suggesting that most transcription factors are positive regulators. Genome Biol 15, 413 (2014). https://doi.org/10.1186/s13059-014-0413-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13059-014-0413-3

Keywords