RNA-mediated gene regulation is less evolvable than transcriptional regulation
Edited by Rafael Sanjuán, Universitat de Valencia, Valencia, Spain, and accepted by Editorial Board Member Daniel L. Hartl February 28, 2018 (received for review November 2, 2017)
Significance
Cells regulate the activity of genes in a variety of ways. For example, they regulate transcription through DNA binding proteins called transcription factors, and they regulate mRNA stability and processing through RNA binding proteins. Based on current knowledge, transcriptional regulation is more widespread and is involved in many more evolutionary adaptations than posttranscriptional regulation. The reason could be that transcriptional regulation is studied more intensely. We suggest instead that transcriptional regulation harbors an intrinsic evolutionary advantage: when mutations change transcriptional regulation, they are more likely to bring forth novel patterns of such regulation. That is, transcriptional regulation is more evolvable. Our analysis suggests a reason why a specific kind of gene regulation is especially abundant in the living world.
Abstract
Much of gene regulation is carried out by proteins that bind DNA or RNA molecules at specific sequences. One class of such proteins is transcription factors, which bind short DNA sequences to regulate transcription. Another class is RNA binding proteins, which bind short RNA sequences to regulate RNA maturation, transport, and stability. Here, we study the robustness and evolvability of these regulatory mechanisms. To this end, we use experimental binding data from 172 human and fruit fly transcription factors and RNA binding proteins as well as human polymorphism data to study the evolution of binding sites in vivo. We find little difference between the robustness of regulatory protein–RNA interactions and transcription factor–DNA interactions to DNA mutations. In contrast, we find that RNA-mediated regulation is less evolvable than transcriptional regulation, because mutations are less likely to create interactions of an RNA molecule with a new RNA binding protein than they are to create interactions of a gene regulatory region with a new transcription factor. Our observations are consistent with the high level of conservation observed for interactions between RNA binding proteins and their target molecules as well as the evolutionary plasticity of regulatory regions bound by transcription factors. They may help explain why transcriptional regulation is implicated in many more evolutionary adaptations and innovations than RNA-mediated gene regulation.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
Gene expression is regulated at multiple levels ranging from the accessibility of chromatin to the posttranslational modification of proteins. Much of this regulation is carried out by sequence-specific, nucleotide-binding proteins that target DNA or RNA molecules. Transcription factors (TFs) are one such class of proteins. They bind short DNA sequences to regulate gene expression at the level of transcription by activating or blocking the recruitment of RNA polymerase to the transcription start site (1). RNA binding proteins (RBPs) are another such class of proteins. They bind short RNA sequences to regulate gene expression posttranscriptionally by regulating the splicing of precursor mRNA as well as the stability, transport, translation, and decay of mature mRNA (2).
Mutations that affect the regulation of gene expression are often deleterious. This is evidenced by the numerous diseases associated with mutations in the nucleic acid binding sites of regulatory proteins (3–6). For example, spinal muscular atrophy, a pediatric neurodegenerative disorder, is caused by a point mutation in an exonic splicing element, which abrogates binding of an RBP and results in aberrant exon splicing. Such mutations are not rare. For instance, of 2,931 disease-associated SNPs located within regulatory DNA, 93.2% fall within sequences that bind TFs (5). It is, therefore, important that protein–nucleotide interactions are to some extent robust to mutation.
Changes in the regulation of gene expression need not be deleterious. They can also be adaptive and drive evolutionary change (7–10). For example, single-nucleotide mutations in the binding sites of TFs that regulate the expression of Drosophila Rhodopsin genes led to the restricted expression of these genes in specific subsets of photoreceptors. This change in gene expression facilitated the discrimination of a wide spectrum of optical stimuli and likely provided a selective advantage to the fly (11). It is, therefore, also important that protein–nucleotide interactions are evolvable, meaning that mutations to nucleic acid binding sites have the potential to bring forth new binding phenotypes. That is, they can change which protein a sequence binds, because such a change may lead to an adaptive change in the level, timing, or location of gene expression.
Robustness and evolvability are often studied within the context of a genotype–phenotype map (12, 13), an object of central importance in the biological sciences (14–16). Most of what we know about genotype–phenotype maps comes from computational models of biological systems (17–23). The two most prominent examples are models that predict the secondary structure phenotypes of RNA sequence genotypes (18) and the lattice-based structural phenotypes of simplified amino acid sequence genotypes (17). Analyses of these and other models have revealed three hallmark characteristics of genotype–phenotype maps (24): (i) many genotypes encode the same phenotype, (ii) the number of genotypes per phenotype has a highly nonuniform distribution, and (iii) genotype networks [also known as neutral networks (18)] mutationally connect sets of genotypes that have the same phenotype (25). The existence of genotype networks is important for at least two reasons. First, a mutation to any one genotype on the network has the potential to produce another genotype that is also on the network and therefore, also has the same phenotype. Second, genotype networks tend to spread far throughout the space of all possible genotypes, which provides mutational access to the genotype networks of different phenotypes. Genotype networks, therefore, confer both robustness and evolvability to phenotypes (25).
The study of genotype–phenotype maps is currently being transformed by technological advances in high-throughput sequencing and chip-based technologies (26–43) as well as by synthetic biology (44). These approaches facilitate the assignment of phenotypes to a large number of genotypes and thus, provide the opportunity to empirically characterize a genotype–phenotype map. Examples of such phenotypes include the ability of RNA molecules to bind an aptamer (29), the fluorescent activity of proteins (39), and the formation of a spatial stripe by synthetic gene regulatory circuits (44). For some genotype–phenotype maps, a phenotype can be assigned to all possible genotypes. Specifically, protein binding microarrays provide measurements of the affinity with which a TF binds to all possible 8-nt DNA sequences (45), and RNAcompete provides measurements of the affinity with which an RBP binds to all but two of the possible 7-nt RNA sequences (46). Data like these can be thought of as exhaustively enumerated, empirically derived genotype–phenotype maps, where the genotype is a DNA or RNA sequence and the phenotype is the molecular capacity of the sequence to bind a TF or RBP, respectively (31).
Here, we use experimental data from protein binding microarrays and RNAcompete to study and compare the robustness and evolvability of the nucleic acid binding sites of TFs and RBPs via a comparative analysis of their genotype–phenotype maps. In these maps, a genotype is a nucleic acid sequence, and a phenotype is the molecular capacity of the sequence to bind a TF or RBP. These genotype–phenotype maps are particularly amenable to comparative analysis because of the many similarities between these two forms of protein–nucleotide interactions. For example, the biophysics of binding are similar for the two nucleic acids, such that binding affinity between nucleotides in the DNA or RNA sequence and amino acids in the protein’s binding domain is largely determined via intermolecular interactions, such as ionic and hydrogen bonding (47, 48). Additionally, the number of DNA sequences profiled by protein binding microarrays is of the same order of magnitude as the number of RNA sequences profiled by RNAcompete, and this number is small enough to facilitate the exhaustive analysis of binding preferences for many TFs and RBPs (49, 50). Finally, both of these datasets exhibit the three hallmark characteristics of genotype–phenotype maps: (i) multiple distinct sequences bind the same protein (49, 50); (ii) some proteins are bound by many distinct sequences, whereas others are bound by very few (49, 50); and (iii) the sequences that bind a protein tend to form a single-genotype network (31, 51). Despite these similarities, we show here that these two genotype–phenotype maps exhibit pronounced architectural differences, which suggest that the nucleic acid binding sites of RNA-mediated gene regulation are less evolvable than those of transcriptional regulation.
Results
Protein Binding Microarray and RNAcompete Data.
We use experimental data from protein binding microarrays (50) and RNAcompete (49) to construct and compare two empirical genotype–phenotype maps. These data are comparable, because both technologies use the same analysis pipeline to transform the fluorescent intensities of microarray spots into an enrichment score (-score) for all possible sequences of a short length. The -score is a variant of the Wilcoxon–Mann–Whitney statistic that ranges from −0.5 to 0.5 and describes the relative binding preferences of a protein to its nucleic acid ligands, with larger numbers indicating increased preference. Protein binding microarrays profile the binding specificity of a TF by assigning an -score to all 32,896 possible 8-nt DNA sequences, whereas RNAcompete profiles the binding specificity of an RBP by assigning an -score to 16,382 7-nt RNA sequences (Materials and Methods). By setting a threshold on the -score distribution, one can delineate sequences that specifically bind a TF or RBP from those that do not. Thus, we can assign the binary phenotype of bound or unbound to each sequence respective to each TF and RBP. To do so, we use an -score threshold of 0.35 following earlier work (31, 41, 52, 53) (Materials and Methods). In SI Appendix, we consider more stringent binding affinity thresholds to show that our findings are qualitatively insensitive to this parameter.
Because there are currently far more protein binding microarray data available than RNAcompete data, we choose our study species according to the availability of the RNAcompete data. Specifically, we study proteins from Drosophila melanogaster and Homo sapiens, because these species have more RNAcompete data available than any other species (Materials and Methods, Table 1, and Dataset S1). These proteins include 38 TFs and 71 RBPs from H. sapiens as well as 30 TFs and 33 RBPs from D. melanogaster.
Table 1.
Type of protein | Species | No. of proteins in the database | No. of proteins in our dataset | No. of binding domain classes in our dataset |
---|---|---|---|---|
TF | H. sapiens | 41 | 38 (29) | 11 (9) |
TF | D. melanogaster | 30 | 30 (23) | 10 (7) |
RBP | H. sapiens | 77 | 71 (38) | 9 (5) |
RBP | D. melanogaster | 48 | 33 (20) | 4 (4) |
A protein was included in our dataset if it bound at least one sequence (E-score > 0.35). Numbers in parentheses correspond to a reduced dataset, in which we only consider proteins that have a genotype network that occupies > 0.5% of the genotype space.
We construct a genotype network for each TF and for each RBP. In such a network, vertices represent sequences that bind the TF or RBP (-score 0.35), and edges connect vertices if their corresponding sequences differ by a single small mutation (Materials and Methods) (31). Since some sequences bind multiple TFs or RBPs, genotype networks may overlap. Additionally, the sequences that bind a TF or RBP sometimes form multiple, disconnected genotype networks (SI Appendix, Table S1). In this case, there is always one genotype network that is much larger than the others (Dataset S1). We restrict our analyses to this largest genotype network for each TF and RBP as in our previous work (31, 41, 51).
Fig. 1 shows the sizes of the genotype networks of TF and RBP binding sites, expressed as fractions of genotype space, for both D. melanogaster and H. sapiens. We consider this fractional size rather than absolute size to facilitate the comparison of these two genotype–phenotype maps, which differ in the sizes of their respective genotype spaces. The range of fractional genotype network sizes is similar for the two classes of proteins, although there are fewer small genotype networks of TF binding sites than of RBP binding sites. For example, 16 genotype networks of TF binding sites (24%) occupy less than 0.5% of genotype space, whereas 46 genotype networks of RBP binding (44%) sites occupy less than 0.5% of genotype space (dashed vertical lines in Fig. 1).
Fig. 1.

Robustness.
The mutational robustness of an individual binding site refers to the likelihood that a mutation in this site leaves binding intact. We quantify this robustness as the fraction of all possible mutations that yield a sequence that is part of the same genotype network. We then generalize this quantity by computing the average robustness of all sites that bind a given TF or RBP (31, 54), which provides a measure of binding site robustness specific to any one nucleic acid binding protein.
The distributions of robustness are similar for binding sites of TFs and RBPs (Fig. 2 A–D). On average, the robustness of D. melanogaster TFs is 0.33, whereas the average robustness of RBPs is 0.27. A nearly identical pattern is found in H. sapiens, where the average robustness for TFs is 0.33 compared with 0.26 for RBPs. However, the left-hand tail of the distribution is longer for RBPs. This means that some RBPs have considerably lower robustness than any of the TFs ( for D. melanogaster and for H. sapiens, one-sided Wilcoxon rank sum test). For example, nine of the D. melanogaster RBPs are less robust than the least robust TF.
Fig. 2.

To determine why robustness can be higher for TFs than for RBPs, we studied the relationship between robustness and genotype network size. Fig. 2E shows that robustness increases approximately logarithmically with the size of the genotype network for both classes of proteins and in both species. This relationship may be a generic property of genotype–phenotype maps as computational models suggest (23, 24, 55, 56). The white symbols in the lower left-hand corner of the plot in Fig. 2E show that the low-robustness RBPs have small genotype networks. This indicates that robustness is often higher for TFs, because they do not have genotype networks as small as RBPs.
Evolvability.
In the context of nucleic acid binding sites, evolvability is the propensity of a mutation to bring forth a new binding phenotype (31). We quantify the evolvability of a TF’s or RBP’s binding sites in two steps. First, we determine the set of sequences that differ by a single mutation from any of the protein’s binding sites but that are not part of the protein’s genotype network. Second, we determine the fraction of TFs or RBPs in our dataset that these one-mutant neighbors bind (31, 54). This fraction is our measure of evolvability. It takes on high values when the one-mutant neighbors of a genotype network bind many TFs or RBPs and low values when the sequences neighboring a genotype network bind few TFs or RBPs.
We find that TF binding sites are considerably more evolvable than RBP binding sites in both D. melanogaster and H. sapiens (Fig. 3 A–D) ( for both comparisons, one-sided Wilcoxon rank sum test). For example, in D. melanogaster, 29 of 30 TFs are more evolvable than even the most evolvable RBP. More generally, mutations to TF binding sites have the potential to create binding sites for 96% of the TFs in our dataset (65 TFs) on average, whereas mutations to RBP binding sites have the potential to create binding sites for 60% of the RBPs in our dataset (62 RBPs) on average. This means that the average evolvability of TFs is more than 1.5 times that of RBPs in both species. Mutations to TF binding sites are, therefore, more likely to bring forth new binding phenotypes than are mutations to RBP binding sites.
Fig. 3.

We next studied the relationship between evolvability and genotype network size to better understand how the genotype networks of different binding phenotypes are distributed in the genotype spaces of the two genotype–phenotype maps. Fig. 3E shows that evolvability increases with genotype network size for both TFs and RBPs but that the rate of increase is slower for RBPs. This raises the possibility that nucleic acid binding phenotypes are arranged differently in the two genotype spaces, because genotype networks of similar size confer different levels of evolvability as do phenotypes of similar robustness (SI Appendix, Fig. S1). Perhaps genotype networks of TF binding sites are nearer to one another in genotype space than are those of RBP binding sites? To find out, we calculated the average mutational distance between all pairs of genotypes from different genotype networks for all pairs of TFs and for all pairs of RBPs. This measure provides a sense of the proximity of genotype networks in genotype space. Fig. 4A shows that genotype networks of TF binding sites tend to be separated by fewer mutations than genotype networks of RBP binding sites ( for both comparisons, one-sided Wilcoxon rank sum test). Complementing this analysis, we also characterized the extent of overlap among pairs of genotype networks, where the overlap of two genotype networks is defined as the number of binding sites that they have in common. Fig. 4B shows that the genotype networks of TF binding sites have more sites in common than the genotype networks of RBP binding sites ( for both comparisons, one-sided Wilcoxon rank sum test). These patterns are consistent among pairs of proteins that have the same or different binding domains (SI Appendix, Figs. S2 and S3). Large genotype networks, therefore, confer evolvability in both genotype–phenotype maps, but the extent of this evolvability depends on how the phenotypes are distributed throughout genotype space, a property that varies among the maps.
Fig. 4.

To find out more generally how mutations to binding sites can bring forth novel binding phenotypes, we calculated the number of unique binding phenotypes that are found within mutations of a focal genotype (18). We performed two variants of this analysis. In the first, we only considered genotypes within mutations of any one genotype if they belonged to the same genotype network as the focal genotype (i.e., the nucleic acid sequences bind the same protein). Because any one of these nucleic acid sequences may also bind other proteins, this analysis amounts to a further characterization of genotype network overlap. Fig. 5 shows that, for any number of mutations, a greater proportion of binding phenotypes can be reached in the genotype space of transcriptional regulation than in that of RNA-mediated gene regulation. For example, with just a single mutation (), 39% of D. melanogaster TFs and 23% of H. sapiens TFs can be reached compared with 25% of D. melanogaster RBPs and 9% of H. sapiens RBPs. This difference is even more pronounced at the largest number of mutations ( for TFs and for RBPs), where 75% of D. melanogaster TFs and 67% of H. sapiens TFs can be reached compared with just 44% of D. melanogaster RBPs and 23% of H. sapiens RBPs. In the second variant of this analysis, we considered all genotypes within a given number of mutations of a focal genotype, regardless of whether these genotypes belong to the same genotype network as the focal genotype (i.e., regardless of whether the nucleic acid sequences bind the same protein). This measure, therefore, simultaneously characterizes the overlap and juxtaposition of genotype networks. SI Appendix, Fig. S4 shows that, at any number of mutations, one can reach at least as many and usually more novel binding phenotypes for TFs than for RBPs. As the mutational radius approaches the diameter of the genotype space, the fraction of reachable phenotypes will necessarily approach one. However, the number of mutations at which this occurs is always smaller in the genotype space of TF binding sites than in the genotype space of RBP binding sites. For example, in D. melanogaster, all phenotypes only become reachable for RBP binding sites at the largest possible number of mutations, whereas mutations—just one-half of the maximum distance—suffice to reach all TF binding phenotypes. In sum, both variants of this analysis show that the nucleic acid binding sites of transcriptional regulation are more evolvable, because a given number of mutations can reach a larger number of new binding phenotypes.
Fig. 5.

At least some of the reduced robustness and evolvability of RBP binding sites relative to TF binding sites stems from the presence of small genotype networks in our RBP dataset. This raises the question of whether the abundance of small genotype networks is truly a characteristic feature of this genotype–phenotype map or is actually the result of a sampling bias in our dataset. Such bias could occur if our dataset includes a nonrepresentative proportion of RBPs that have small genotype networks. One way to determine if this bias exists is to compare the distributions of the number of bound sequences per RBP in our dataset with those of the remaining 68 RBPs for which binding data are available (Materials and Methods). The two distributions are statistically indistinguishable (, Wilcoxon rank sum test) (SI Appendix, Fig. S5). To the extent that the RNAcompete data are representative of RBP binding preferences in general, this suggests that small genotype networks are indeed a characteristic feature of the genotype–phenotype map of RBP binding sites. In SI Appendix, we also show that our main conclusions are insensitive to the removal of proteins with small genotype networks (SI Appendix, Figs. S6–S8) and to the single-nucleotide difference in the lengths of the TF and RBP binding sites that we study (SI Appendix, Figs. S9–S12). Additionally, we show that our main conclusions are insensitive to the relaxation of some of our modeling assumptions, including the affinity threshold used to delineate bound from unbound sequences (SI Appendix, Figs. S13–S18) and the exclusion of small genotype network components from our evolvability measure (SI Appendix, Fig. S19).
Binding Site Variants in the Human Population.
The measures of robustness and evolvability that we studied here take into consideration all of the DNA and RNA sequences that bind a TF or RBP, respectively. Moreover, they assume that all types of point mutations to these sequences are equally likely. However, only a subset of all DNA and RNA sequences is used for gene regulation in vivo, and mutations to these sequences may be subject to biases. These include context-dependent mutation rates (57) as well as simple transition:transversion biases (58). In other words, our observations above need not hold for the binding sites encountered in vivo or for their mutational variants.
To find out if they do, we studied putative TF and RBP binding sites in humans and the single-nucleotide mutants of these sequences that exist as standing variation (Materials and Methods). Specifically, we collected DNaseI footprint data from 41 diverse cell and tissue types (59). These data demarcate protein-bound regions of the genome at single-nucleotide resolution genome-wide and can, therefore, be used to predict TF binding sites. We focused on footprints that are likely to be involved in gene regulation by filtering the footprints to only include those that overlap the promoter regions of protein-coding genes. We also collected RNase footprints from HeLa cells, which demarcate protein-bound regions of the transcriptome at single-nucleotide resolution transcriptome-wide and can, therefore, be used to predict RBP binding sites (60). We focused on footprints that are likely to be involved in gene regulation by filtering the footprints to only include those that were in the 5′ or 3′ UTRs of protein-coding transcripts, because these are common regulatory targets of RBPs (49). For each of the DNaseI and RNase footprints in these regulatory regions, we determined the likely bound TFs or RBPs based on the available binding data (49, 50) (Materials and Methods). This resulted in a list of putative binding sites for each TF and RBP. To determine which mutational variants of these binding sites exist as standing variation in the human population, we queried the 1000 Genomes Project Consortium data for SNPs (61). This allowed us to determine which of a binding site’s mutational variants likely abrogate binding to the focal protein and which do not. For those that are likely to abrogate binding, we determined which other TFs or RBPs the mutational variant might bind. In this way, we determined the robustness and evolvability of the set of nucleic acid sequences that are encountered as putative binding sites in vivo and how these properties relate to the size of a protein’s genotype network.
Fig. 6A shows that, as the size of a genotype network increases, so does the number of a binding site’s mutational variants that exist as standing variation in the human population and that do not abrogate binding to the focal protein (i.e., the mutational variants of a binding site are on the same genotype network as the binding site). This pattern is similar to that of Fig. 2E, indicating that the relationship between robustness and genotype network size is consistent among the subset of binding sites encountered in vivo and the superset of sequences characterized in vitro. Moreover, the distributions of robustness are statistically indistinguishable among TFs and RBPs (, one-sided Wilcoxon rank sum test), consistent with our observation that their binding sites are similarly robust to mutation (Fig. 2). Much of the standing variation in human nucleic acid binding sites is, therefore, unlikely to abrogate binding, especially for TFs and RBPs with large genotype networks.
Fig. 6.

Fig. 6B shows that, as the size of a genotype network increases, so does the number of TFs or RBPs that are bound by the one-mutant neighbors of binding sites that abrogate binding to the focal protein (i.e., the mutational variants of a binding site are not on the same genotype network as the binding site but rather, are on the genotype network of another protein). This pattern is similar to that of Fig. 3E, indicating that, like robustness, the relationship between evolvability and genotype network size is consistent among the subset of binding sites encountered in vivo and the superset of sequences characterized in vitro. Therefore, mutational variants of TF binding sites in the human population are more likely to have new binding phenotypes than those of RBP binding sites (, one-sided Wilcoxon rank sum test). This provides support for our main conclusion that the nucleic acid binding sites of RNA-mediated gene regulation are less evolvable than those of transcriptional regulation.
Discussion
The nucleic acid sequences that bind TFs dictate when, where, and to what extent genes are transcribed. The resulting transcripts comprise nucleic acid sequences that bind RBPs to regulate much of the transcripts’ lifecycle, including splicing, transport, and decay. Interactions between nucleic acid binding sites and their cognate regulatory proteins are, therefore, fundamental to the regulation of gene expression. Here, we used experimental data to study the robustness and evolvability of these interactions. We did so via a comparative analysis of their genotype–phenotype maps. In these maps, a genotype is a nucleic acid sequence, and a phenotype is the molecular capacity of the sequence to bind a protein (31). For each TF and RBP, we constructed a genotype network from the sequences that bind the protein, and we studied how these genotype networks confer robustness and evolvability to the binding sites they harbor.
These empirical genotype–phenotype maps exhibit two architectural differences. First, the genotype networks of RBP binding sites are often much smaller than those of TF binding sites, which means that the bindings sites of some RBPs are less robust than those of TFs. Second, the genotype networks of TF binding sites interface and overlap with one another more than the genotype networks of RBP binding sites, rendering the space of TF binding sites more conducive to the evolution of new binding phenotypes. This observation is consistent with our analysis of standing genetic variation in the human population as reported by the 1000 Genomes Project Consortium (61). Specifically, standing variation in TF binding sites is more likely to confer new binding phenotypes than standing variation in RBP binding sites.
There are additional facets to the robustness and evolvability of transcriptional and RNA-mediated gene regulation that we do not study here. For example, the robustness of transcriptional regulation can be enhanced by homotypic clusters of TF binding sites, shadow enhancers, and redundant TFs (62), whereas the evolvability of RNA-mediated gene regulation can be enhanced by alternative splicing, alternative polyadenylation, and RNA editing (63). Even in the context of nucleic acid binding sites, which are the focus of this study, there is an additional facet to evolvability that we do not consider but that may further contribute to the greater evolvability of TF binding sites. It is the propensity for binding sites to evolve de novo in regulatory regions (64, 65). The reason is that, in higher eukaryotes, such as the species studied here, the transcriptional regulation of gene expression is typically implemented by a promoter and several enhancers (66, 67), and each of these can span well over 1 kb of DNA (68, 69). This is a larger mutational target than the 5′ and 3′ UTRs of transcripts, which in humans, tend to span 200 bp and 1 kbp pair, respectively (70). While RBP binding sites are known to evolve de novo in these regions (71), the larger size of promoters and enhancers makes it likely that TF binding sites are also more evolvable in the context of de novo evolution.
Our evolvability measure is specifically concerned with mutations to binding sites that change the binding partner from one protein to another. Such mutations can indeed lead to adaptive changes in gene expression as examples from transcriptional regulation show (11, 72, 73). For instance, the initiation and progression of cancer are parts of an evolutionary process, in which mutations facilitate the uncontrolled division and spread of abnormal cells throughout the body. Such mutations commonly arise in regulatory regions (72, 74, 75) as evidenced by recurrent mutations in the binding sites of CEBP TFs, which create high-affinity binding sites for other TFs (72). These mutations are found across a diversity of cancer types, which suggests that they drive a change in gene expression that is selectively advantageous for cancer cells. We do not know of an analogous example for RBP binding sites. The reason may be that they are not as well-studied, that they show lower evolvability, or both.
Our study has three limitations that are worth highlighting. The first is the size of our dataset. The human genome encodes 1,400 TFs and 700 mRNA-binding RBPs (6, 76), whereas our dataset comprises 38 human TFs and 71 human RBPs. The inclusion of more proteins will necessarily affect the architecture of the genotype–phenotype maps that we study. However, the consistency of our conclusions across the D. melanogaster and H. sapiens datasets, which comprise different binding domains and different numbers of proteins per binding domain, provides reassurance that our findings are general. The second limitation is that we use indirect evidence of in vivo protein–nucleotide interactions, because footprinting assays do not reveal which proteins are bound by a specific nucleic acid sequence. We ameliorate this limitation by only studying footprints that contain sequences that bind the proteins in our dataset. Ideally, however, we would study data from assays that provide direct evidence of protein–nucleotide interactions (77, 78), but these data are available neither for most of the proteins in our dataset nor for the 1000 Genomes Project Consortium data. The third limitation is the length of the binding sites that we study, which are representative of the binding preferences of many but not all TFs and RBPs (49, 50). For example, Cys2-His2 zinc finger proteins typically bind longer sequences (79), to which we cannot extrapolate our findings. As our ability to predict (80) and measure (81) the binding preferences of such proteins continues to advance, it will become possible to extend our analyses to longer nucleic acid ligands.
In sum, our comparative analysis of two empirical genotype–phenotype maps suggests that the nucleic acid binding sites of RNA-mediated gene regulation are less evolvable than those of transcriptional regulation. This observation is consistent with the high levels of mRNA target conservation for RBPs (10, 82, 83) and the evolutionary plasticity of the regulatory regions involved in transcriptional regulation (84–86) as well as their role in the evolution of myriad adaptations and innovations (11, 87, 88).
Materials and Methods
Protein Binding Microarray and RNAcompete Data.
The largest databases for protein binding microarray and RNAcompete data are CIS-BP (50) and CISBP-RNA (49), respectively. There is currently far more protein binding microarray data available than RNAcompete data. Specifically, the most recent build of the CIS-BP database (version 1.02) contains protein binding microarray data for 1,665 TFs from 132 species, whereas the most recent build of the CISBP-RNA database (version 0.6) contains RNAcompete data for 194 RBPs from 24 eukaryotic species. Our choice of study species was, therefore, guided by the availability of the RNAcompete data, because they are more limited. We chose to study proteins from D. melanogaster and H. sapiens, because these species have more RNAcompete data available than any of the other species (Table 1 and Dataset S1). We limited our study to two species, because the species with the next largest number of RBPs in the CISBP-RNA database was Caenorhabditis elegans, which had only 15 RBPs profiled.
More specifically, we downloaded protein binding microarray data for 30 D. melanogaster and 38 H. sapiens TFs from the CIS-BP database (version 1.02) (50), and we downloaded RNAcompete data for 33 D. melanogaster and 71 H. sapiens RBPs from the web supplement of ref. 49. Both the protein binding microarray and RNAcompete data include a nonparametric rank-based enrichment score (-score) that can be used to delineate sequences that specifically bind a TF or RBP from those that do not. The protein binding microarray data include an -score for all possible 8-nt dsDNA sequences. The total number of such sequences is rather than , because each sequence is merged with its reverse complement and because there are sequences that are identical to their reverse complement and therefore, cannot be merged (89). The RNAcompete data include an -score for nearly all possible 7-nt ssRNA sequences. The total number of such sequences is , where the accounts for the two sequences (GCTCTTC and GAAGAGC) that contain an SapI restriction site, which is used during the RNA pool synthesis to remove linker sequences from PCR products (46). The data from these two experimental protocols are comparable, because the numbers of nucleic acid sequences profiled are of the same order of magnitude and because the -score is calculated in the same way and has the same meaning in the two datasets. Following earlier work (31, 41, 52), we consider a sequence to bind a TF or an RBP if its -score exceeds . We chose this threshold not only because it has precedent (31, 41, 52, 53), but also because an analysis of the relationship between -score and false discovery rate for 104 mouse TFs revealed that sequences with an -score greater than 0.35 had a false discovery rate less than 0.001 (31). In SI Appendix, we perform a sensitivity analysis of our results to this binding affinity threshold.
For each TF and RBP, -scores were provided from two distinct array designs. We considered the -score of a DNA or RNA sequence to be the average of the two -scores. To include a TF or RBP in our dataset, we required that it bind at least one sequence (i.e., that at least one sequence had an average -score ). For 1 TF and for 12 RBPs, data were available from more than one experiment. For these proteins, we chose the experiment with the largest number of bound sequences (Dataset S1) to avoid any bias toward small genotype networks in the genotype–phenotype map of RBP binding sites. In a supplementary analysis reported in SI Appendix, Fig. S5, we studied RNAcompete data for an additional 68 RBPs. We processed these in the same way as the data from the 33 D. melanogaster and 71 H. sapiens RBPs. In supplementary analyses reported in SI Appendix, Figs. S9–S12, we studied the space of all possible 7-nt dsDNA sequences (note that, in this case, we do not have to account for sequences that are identical to their reverse complement, because such sequences only exist when the sequence length is an even number). We used the same analysis pipeline to calculate -scores for these sequences as used for the 8-nt DNA sequences studied in the text.
Both the CIS-BP and the CISBP-RNA databases have been updated since their original release. This is why our calculation of the total number of TFs and RBPs in these databases does not match the numbers reported in the original manuscripts (49, 50). We calculated the number of TFs and species in the CIS-BP database by entering “PBM” as “By Evidence Type” on the homepage of the CIS-BP website and downloading the resulting .csv file. We removed all TFs labeled as “PBM_CONTRUCTS” in the “Species” column and then counted the number of TFs and the number of unique species. Analogously, we calculated the number of RBPs and species in the CISBP-RNA database by entering “RNAcompete” as By Evidence Type in the homepage of the CISBP-RNA website and downloading the resulting .csv file. We removed all RBPs labeled as “RNAcompete_CONSTRUCTS” in the Species column and counted the number of RBPs and the number of unique species.
Human Polymorphism Data.
We used data from phase 3 of the 1000 Genomes Project, which includes 84.7 million SNPs from 2,504 individuals representing 26 human populations (61). We filtered the dataset to only include SNPs that passed quality control. These data and the data described below pertain to build hg19 of the human genome.
We sought to determine the amount of variation in TF and RBP binding sites. To do so, we first identified regions of the human genome that may be involved in gene regulation by TFs or RBPs. For TFs, we identified protein-bound regions of gene promoters using digital footprints from DNaseI hypersensitivity assays (59). These data provide genome-wide evidence of protein–DNA interactions across 41 cell and tissue types at single-nucleotide resolution. We used BEDTools to filter the footprints (90), such that we only retained footprints that overlapped promoter regions by at least 1 bp. We defined the promoter region of a gene to be the DNA sequence 1 kb upstream of the gene’s transcription start site. We restricted our attention to genes encoding mRNA as indicated by the prefix NM in the RefSeq database (91). For these same genes, we identified 5′ and 3′ UTRs as potential regulatory targets for RBPs. We identified protein-bound regions in mRNA transcripts using footprints from an RNase hypersensitivity assay (60), which provides transcriptome-wide evidence of protein–RNA interactions in HeLa cells. We used bedtools to filter the footprints, such that we only retained footprints that overlapped 5′ and 3′ UTRs by at least 1 bp.
For each TF, we scanned each promoter footprint for potential TF binding sites, and for each RBP, we scanned each 5′ UTR footprint and each 3′ UTR footprint for potential RBP binding sites. The scanning worked as follows. If a footprint contained at least one sequence with an -score for the TF or RBP, then we considered this sequence to be a potential binding site. If the footprint contained more than one binding site, then we randomly chose one of them for later analysis. We chose to assign only one binding site per footprint per TF or RBP to avoid any bias that may be introduced by the different lengths of DNaseI footprints and RNase footprints.
The above procedure resulted in a list of genomic coordinates of potential binding sites for each TF and RBP. The number of such TF binding sites ranged from 266 for the TF POU6F1 in HFF cells to a maximum of 50,126 for the TF DNMT1 in NB4 cells. The number of such RBP binding sites ranged from 0 for the RBPs ENOX1, FXR2, and ZNF638 to a maximum of 32,272 for the RBP HNRNPH2. For each of these potential binding sites, we used the Samtools function tabix to query the 1000 Genomes Consortium data for SNPs (92). We used the results of these queries to calculate the proportion of binding site variants that is on the focal TF’s or RBP’s genotype network and the proportion that is not. For those variants that are not on the genotype network, we calculated the fraction of TFs or RBPs in our dataset that these variants have the potential to bind.
Genotype Networks.
We constructed a genotype network for each TF and RBP following the procedure of Payne and Wagner (31). Specifically, for each protein, we used the Smith–Waterman algorithm to perform a pairwise alignment on all pairs of bound sequences (-score ), in which we prohibited gaps. We then calculated the mutational distance between two sequences and as the number of mismatches in the alignment of these sequences. We represented bound sequences as vertices in the network, and we connected two vertices by an edge if their corresponding sequences differed by a single small mutation. We considered point mutations and small indels that shift an entire contiguous binding site by a single base (31). For TFs, we calculated the mutational distance between two DNA sequences and as the minimum of the mutational distance and , where is the reverse complement of . For RBPs, we calculated the mutational distance between two RNA sequences and as . We constructed genotype networks in this way using the Genonets Server (51), which we also used to measure robustness and evolvability as well as the overlap among genotype networks.
Statistical Analysis.
We use the nonparametric, one-sided Wilcoxon rank sum test of the null hypothesis that the distributions of samples and come from distributions with equal medians against the alternative hypothesis that is sampled from a distribution with a median that is greater than that of the distribution from which was sampled. We do not distinguish between values that are less than .
Acknowledgments
We thank Mihai Albu, Alejandro V. Cano, Paco Majic, and Matt Weirauch for discussions as well as three anonymous reviewers for their critical reading and feedback. J.L.P. acknowledges support from Swiss National Science Foundation Grants PZ00P3_154773 and PP00P3_170604. A.W. acknowledges support from European Research Council Advanced Grant 739874, Swiss National Science Foundation Grant 31003A_172887, and the University Priority Research Program in Evolutionary Biology at the University of Zurich.
Supporting Information
Appendix (PDF)
- Download
- 1.04 MB
Dataset_S01 (XLSX)
- Download
- 52.53 KB
Dataset_S02 (XLSX)
- Download
- 40.23 KB
References
1
VM Weake, JL Workman, Inducible gene expression: Diverse regulatory mechanisms. Nat Rev Genet 11, 426–437 (2010).
2
JD Keene, RNA regulons: Coordination of post-transcriptional events. Nat Rev Genet 8, 533–543 (2007).
3
KE Lukong, K Chang, EW Khandjian, S Richard, RNA-binding proteins in human genetic disease. Trends Genet 24, 416–425 (2008).
4
TA Cooper, L Wan, G Dreyfuss, RNA and disease. Cell 136, 777–793 (2009).
5
MT Maurano, et al., Systematic localization of common disease-associated variation in regulatory DNA. Science 337, 1190–1195 (2012).
6
S Gerstberger, M Hafner, T Tuschl, A census of human RNA-binding proteins. Nat Rev Genet 15, 829–845 (2014).
7
M King, AC Wilson, Evolution at two levels in humans and chimpanzees. Science 188, 107–116 (1975).
8
B Prud’homme, N Gompel, SB Carroll, Emerging principles of regulatory evolution. Proc Natl Acad Sci USA 104, 8605–8612 (2007).
9
GA Wray, The evolutionary significance of cis-regulatory mutations. Nat Rev Genet 8, 206–216 (2007).
10
GJ Hogan, PO Brown, D Herschlag, Evolutionary conservation and diversification of Puf RNA binding proteins and their mRNA targets. PLoS Biol 13, e1002307 (2015).
11
J Rister, et al., Single-base pair differences in a shared motif determine differential Rhodopsin expression. Science 350, 1258–1261 (2015).
12
J Burns, The synthetic problem and the genotype-phenotype relation in cellular metabolism. Towards a Theoretical Biology. Volume 3: Drafts. An IUBS Symposium, ed CH Waddington (Aldine Publ Co, Chicago), pp. 47–51 (1970).
13
P Alberch, From genes to phenotype: Dynamical systems and evolvability. Genetica 84, 5–11 (1991).
14
M Pigliucci, Genotype-phenotype mapping and the end of the ‘genes as blueprint’ metaphor. Proc R Soc Lond B Biol Sci 365, 557–566 (2010).
15
GP Wagner, J Zhang, The pleiotropic structure of the genotype-phenotype map: The evolvability of complex organisms. Nat Rev Genet 12, 204–213 (2011).
16
B Lehner, Genotype to phenotype: Lessons from model organisms for human genetics. Nat Rev Genet 14, 168–178 (2013).
17
DJ Lipman, JW Wilbur, Modelling neutral and selective evolution of protein folding. Proc R Soc Lond B Biol Sci 245, 7–11 (1991).
18
P Schuster, W Fontana, PF Stadler, IL Hofacker, From sequences to shapes and back: A case study in RNA secondary structures. Proc R Soc Lond B Biol Sci 255, 279–284 (1994).
19
S Ciliberti, OC Martin, A Wagner, Innovation and robustness in complex regulatory gene networks. Proc Natl Acad Sci USA 104, 13591–13596 (2007).
20
J Cotterell, J Sharpe, An atlas of gene regulatory networks reveals multiple three-gene mechanisms for interpreting morphogen gradients. Mol Syst Biol 6, 425 (2010).
21
JFM Rodrigues, A Wagner, Evolutionary plasticity and innovations in complex metabolic reaction networks. PLoS Computat Biol 5, e1000613 (2009).
22
I Salazar-Ciudad, M Marín-Riera, Adaptive dynamics under development-based genotype-phenotype maps. Nature 497, 361–364 (2013).
23
SF Greenbury, IG Johnson, AA Louis, SE Ahnert, A tractable genotype-phenotype map modelling the self-assembly of protein quaternary structure. J R Soc Interface 11, 20140249 (2014).
24
SE Ahnert, Structural properties of genotype-phenotype maps. J R Soc Interface 14, 20170275 (2017).
25
A Wagner, Neutralism and selectionism: A network-based reconciliation. Nat Rev Genet 9, 965–974 (2008).
26
JN Pitt, AR Ferré-D’Amaré, Rapid construction of empirical RNA fitness landscapes. Science 330, 376–379 (2010).
27
W Rowe, et al., Analysis of a complete DNA-protein affinity landscape. J R Soc Interface 7, 397–408 (2010).
28
T Hinkley, et al., A systems analysis of mutational effects in HIV-1 protease and reverse transcriptase. Nat Genet 43, 487–489 (2011).
29
JI Jiménez, R Xulvi-Brunet, GW Campbell, R Turk-MacLeod, IA Chen, Comprehensive experimental fitness landscape and evolutionary network for small RNA. Proc Natl Acad Sci USA 110, 14984–14989 (2013).
30
JD Buenrostro, et al., Quantitative analysis of RNA-protein interactions on a massively parallel array reveals biophysical and evolutionary landscapes. Nat Biotechnol 32, 562–568 (2014).
31
JL Payne, A Wagner, The robustness and evolvability of transcription factor binding sites. Science 466, 714–719 (2014).
32
DW Anderson, AN McKeown, JW Thornton, Intermolecular epistasis shaped the function and evolution of an ancient transcription factor and its DNA binding sites. eLife 4, e07864 (2015).
33
MGJ de Vos, A Dawid, V Sunderlikova, SJ Tans, Breaking evolutionary constraint with a tradeoff ratchet. Proc Natl Acad Sci USA 112, 14906–14911 (2015).
34
AI Podgornaia, MT Laub, Pervasive degeneracy and epistasis in a protein-protein interface. Science 347, 673–677 (2015).
35
P Julien, B Miñana, P Baeza-Centurion, J Valcárcel, B Lehner, The complete local genotype-phenotype landscape for the alternative splicing of a human exon. Nat Commun 7, 11558 (2016).
36
C Li, W Qian, CJ Maclean, J Zhang, The fitness landscape of a tRNA gene. Science 352, 837–840 (2016).
37
O Puchta, et al., Network of epistatic interactions within a yeast snoRNA. Science 352, 840–844 (2016).
38
C Qiu, et al., High-resolution phenotypic landscape of the RNA polymerase II Trigger Loop. PLoS Genet 12, e1006321 (2016).
39
KS Sarkisyan, et al., Local fitness landscape of the green fluorescent protein. Nature 533, 397–401 (2016).
40
B Steinberg, M Ostermeier, Environmental changes bridge evolutionary valleys. Sci Adv 2, e1500921 (2016).
41
J Aguilar-Rodríguez, JL Payne, A Wagner, A thousand adaptive landscapes and their navigability. Nat Ecol Evol 1, 0045 (2017).
42
T Starr, LK Picton, JW Thornton, Alternative evolutionary histories in the sequence space of an ancient protein. Nature 549, 409–413 (2017).
43
EE Wrenbeck, LR Azouz, TA Whitehead, Single-mutation fitness landscapes for an enzyme on multiple substrates reveal specificity is globally encoded. Nat Commun 8, 15695 (2017).
44
Y Schaerli, et al., A unified design space of synthetic stripe-forming networks. Nat Commun 5, 4905 (2014).
45
G Badis, et al., Diversity and complexity in DNA recognition by transcription factors. Science 324, 1720–1723 (2009).
46
D Ray, et al., Rapid and systematic analysis of the RNA recognition specificities of RNA-binding proteins. Nat Biotechnol 27, 667–670 (2009).
47
D Elliott, M Ladomery Molecular Biology of RNA (Oxford Univ Press, Oxford, 2011).
48
GD Stormo Introduction to Protein-DNA Interactions (Cold Spring Harbor Lab Press, Plainview, NY, 2013).
49
D Ray, et al., A compendium of RNA-binding motifs for decoding gene regulation. Nature 499, 172–177 (2013).
50
MT Weirauch, et al., Determination and inference of eukaryotic transcription factor sequence specificity. Cell 158, 1431–1443 (2014).
51
F Khalid, J Aguilar-Rodriguez, A Wagner, JL Payne, Genonets server — A web server for the construction, analysis and visualization of genotype networks. Nucleic Acids Res 44, W70–W76 (2016).
52
C Zhu, et al., High-resolution DNA-binding specificity analysis of yeast transcription factors. Genome Res 19, 556–566 (2009).
53
S Nakagawa, SS Gisselbrecht, JM Rogers, DL Hartl, ML Bulyk, DNA-binding specificity changes in the evolution of forkhead transcription factors. Proc Natl Acad Sci USA 110, 12349–12354 (2013).
54
A Wagner, Robustness and evolvability: A paradox resolved. Proc R Soc Lond B Biol Sci 275, 91–100 (2008).
55
J Aguirre, J Buldú, M Stich, S Manrubia, Topological structure of the space of phenotypes: The case of RNA neutral networks. PLoS One 6, e26324 (2011).
56
SF Greenbury, S Schaper, SE Ahnert, AA Louis, Genetic correlations greatly increase mutational robustness and can both reduce and enhance evolvability. PLoS Comput Biol 12, e1004773 (2016).
57
JW Schroeder, WG Hirst, GA Szewczyk, LA Simmons, The effect of local sequence context on mutational bias of genes encoded on the leading and lagging strands. Curr Biol 26, 692–697 (2016).
58
J Wakeley, The excess of transitions among nucleotide substitutions: New methods of estimating transition bias underscore its significance. Trends Ecol Evol 11, 158–162 (1996).
59
S Neph, et al., An expansive human regulatory lexicon encoded in transcription factor footprints. Nature 489, 83–90 (2012).
60
IM Silverman, et al., RNase-mediated protein footprinting sequencing reveals protein-binding sites throughout the human transcriptome. Genome Biol 15, R3 (2014).
61
; The 1000 Genomes Project Consortium, A global reference for human genetic variation. Nature 526, 68–74 (2015).
62
JL Payne, A Wagner, Mechanisms of mutational robustness in transcriptional regulation. Front Genet 6, 332 (2015).
63
DD Licatalosi, RB Darnell, RNA processing and its regulation: Global insights into biological networks. Nat Rev Genet 11, 75–87 (2010).
64
S MacArthur, JFY Brookfield, Expected rates and modes of evolution of enhancer sequences. Mol Biol Evol 21, 1064–1073 (2004).
65
M Tuğrul, T Paixao, NH Barton, G Tkac̆ik, Dynamics of transcription factor binding site evolution. PLoS Genet 11, e1005639 (2015).
66
D Marbach, et al., Tissue-specific regulatory circuits reveal variable modular perturbations across complex diseases. Nat Genet 13, 366–370 (2016).
67
Q Cao, et al., Reconstruction of enhancer-target networks in 935 samples of human primary cells, tissues and cell lines. Nat Genet 49, 1428–1436 (2017).
68
KY Yip, et al., Classification of human genomic regions based on experimentally determined binding sites of more than 100 transcription-related factors. Genome Biol 13, R48 (2012).
69
EZ Kvon, et al., Genome-scale functional characterization of Drosophila developmental enhancers in vivo. Nature 512, 91–95 (2014).
70
F Mignone, C Gissi, S Liuni, G Pesole, Untranslated regions of mRNAs. Genome Biol 3, REVIEWS0004.1 (2002).
71
D Wilinski, et al., Recurrent rewiring and emergence of RNA regulatory networks. Proc Natl Acad Sci USA 114, E2816–E2825 (2017).
72
C Melton, JA Reuter, DV Spacek, M Snyder, Recurrent somatic mutations in regulatory regions of human cancer genomes. Nat Genet 47, 710–716 (2015).
73
A Cvekl, et al., Evolutionary origins of Pax6 control of crystallin genes. Genome Biol Evol 9, 2075–2092 (2017).
74
FW Huang, et al., Highly recurrent TERT promoter mutations in human melanoma. Science 339, 957–959 (2013).
75
N Weinhold, A Jacobsen, N Schultz, C Sander, W Lee, Genome-wide analysis of noncoding regulatory mutations in cancer. Nat Genet 46, 1160–1165 (2014).
76
JM Vaquerizas, SK Kummerfeld, SA Teichmann, NM Luscombe, A census of human transcription factors: Function, expression, and evolution. Nat Rev Genet 10, 252–263 (2009).
77
DS Johnson, A Mortazavi, RM Meyers, B Wold, Genome-wide mapping of in vivo protein-DNA interactions. Science 316, 1497–1502 (2007).
78
EL Van Nostrand, et al., Robust transcriptome-wide discovery of RNA-binding protein binding sites with enhanced CLIP (eCLIP). Nat Biotechnol 13, 508–514 (2016).
79
HS Najafabadi, et al., C2h2 zinc finger proteins greatly expand the human regulatory lexicon. Nat Biotechnol 33, 555–562 (2015).
80
AV Persikov, M Singh, De novo prediction of DNA-binding specificities for Cys2His2 zinc finger proteins. Nucleic Acids Res 42, 97–108 (2014).
81
A Isakova, et al., SMiLE-seq identifies binding motifs of single and dimeric transcription factors. Nat Methods 14, 316–322 (2017).
82
S Lebedeva, et al., Transcriptome-wide analysis of regulatory interactions of the RBA-Binding protein HuR. Mol Cell 43, 340–352 (2011).
83
MT Locvi, et al., Rbfox proteins regulate alternative mRNA splicing through evolutionarily conserved RNA bridges. Nat Struct Mol Biol 20, 1434–1442 (2013).
84
MZ Ludwig, C Bergman, NH Patel, M Kreitman, Evidence for stabilizing selection in a eukaryotic enhancer element. Nature 403, 564–567 (2000).
85
AE Tsong, BB tuch, H Li, AD Johnson, Evolution of alternative transcriptional circuits with identical logic. Nature 443, 415–420 (2006).
86
MT Weirauch, T Hughes, Conserved expression without conserved regulatory sequence: The more things change, the more they stay the same. Trends Genet 26, 66–74 (2010).
87
N Gompel, B Prud’homme, PJ Wittkopp, VA Kassner, SB Carroll, Chance caught on the wing: cis-Regulatory evolution and the origin of pigment patterns in Drosophila. Nature 433, 481–487 (2005).
88
I Guerreiro, et al., Role of a polymorphism in a hox/pax-responsive enhancer in the evolution of the vertebrate spine. Proc Natl Acad Sci USA 110, 10682–10686 (2013).
89
MF Berger, et al., Compact, universal DNA microarrays to comprehensively determine transcription-factor binding site specificities. Nat Biotechnol 24, 1429–1435 (2006).
90
AR Quinlan, IM Hall, BEDTools: A flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).
91
NA O’Leary, et al., Reference sequence (RefSeq) database at NCBI: Current status, taxonomic expansion, and functional annotation. Nucleic Acids Res 44, D733–D745 (2016).
92
H Li, Tabix: Fast retrieval of sequence features from generic TAB-delimited files. Bioinformatics 27, 718–719 (2010).
Information & Authors
Information
Published in
Classifications
Copyright
Copyright © 2018 the Author(s). Published by PNAS. This open access article is distributed under Creative Commons Attribution-NonCommercial-NoDerivatives License 4.0 (CC BY-NC-ND).
Submission history
Published online: March 26, 2018
Published in issue: April 10, 2018
Keywords
Acknowledgments
We thank Mihai Albu, Alejandro V. Cano, Paco Majic, and Matt Weirauch for discussions as well as three anonymous reviewers for their critical reading and feedback. J.L.P. acknowledges support from Swiss National Science Foundation Grants PZ00P3_154773 and PP00P3_170604. A.W. acknowledges support from European Research Council Advanced Grant 739874, Swiss National Science Foundation Grant 31003A_172887, and the University Priority Research Program in Evolutionary Biology at the University of Zurich.
Notes
This article is a PNAS Direct Submission. R.S. is a guest editor invited by the Editorial Board.
Authors
Competing Interests
The authors declare no conflict of interest.
Metrics & Citations
Metrics
Altmetrics
Citations
Cite this article
RNA-mediated gene regulation is less evolvable than transcriptional regulation, Proc. Natl. Acad. Sci. U.S.A.
115 (15) E3481-E3490,
https://doi.org/10.1073/pnas.1719138115
(2018).
Copied!
Copying failed.
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.