Skip to main content

PERSPECTIVE article

Front. Genet., 22 August 2018
Sec. Epigenomics and Epigenetics
This article is part of the Research Topic Developmental Plasticity and Robustness View all 12 articles

Comparative Studies of Gene Expression Kinetics: Methodologies and Insights on Development and Evolution

Updated
  • Department of Marine Biology, Leon H. Charney School of Marine Sciences, University of Haifa, Haifa, Israel

Across the animal kingdom, embryos of closely related species show high morphological similarity despite genetic and environmental distances. Deciphering the molecular mechanisms that underlie morphological conservation and those that support embryonic adaptation are keys to understand developmental robustness and evolution. Comparative studies of developmental gene regulatory networks can track the genetic changes that lead to evolutionary novelties. However, these studies are limited to a relatively small set of genes and demand extensive experimental efforts. An alternative approach enabled by next-generation sequencing, is to compare the expression kinetic of large sets of genes between different species. The advantages of these comparisons are that they can be done relatively easily, for any species and they provide information of all expressed genes. The challenge in these experiments is to compare the kinetic profiles of thousands of genes between species that develop in different rates. Here we review recent comparative studies that tackled the challenges of accurate staging and large-scale analyses using different computational approaches. These studies reveal how correct temporal scaling exposes the striking conservation of developmental gene expression between morphologically similar species. Different clustering approaches are used to address various comparative questions and identify the conservation and divergence of large gene sets. We discuss the unexpected contribution of housekeeping genes to the interspecies correlations and how this contribution distorts the hourglass pattern generated by developmental genes. Overall, we demonstrate how comparative studies of gene expression kinetics can provide novel insights into the developmental constraints and plasticity that shape animal body plans.

Introduction

Throughout metazoans, closely related species have a highly conserved body plan despite geographic distances and genetic divergence of up to 40 million years of parallel evolution [based on studies of related species of corals (Okubo et al., 2016), nematodes (Levin et al., 2012), flies (Grun et al., 2005; Kalinka et al., 2010; Kuntz and Eisen, 2014), sea urchins (Gildor and Ben-Tabou de-Leon, 2015; Malik et al., 2017), and frogs (Yanai et al., 2011)]. Within the species, embryogenesis can withstand dramatic changes in temperature (Levine et al., 2011; Runcie et al., 2012; Kuntz and Eisen, 2014), acidity (Byrne et al., 2009; Martin et al., 2011; Stumpp et al., 2012; Pespeni et al., 2013a,b; Pan et al., 2015), salinity (Shaw et al., 2014), and other environmental changes (Kitano et al., 2010). Deciphering the molecular mechanisms that underlie morphological conservation is essential for understanding developmental stability; revealing the molecular pathways that support adaptation is a key for comprehending developmental plasticity. These understandings will also help to anticipate the effect of current environmental threats such as ocean acidification and global warming on biodiversity.

Understanding the molecular mechanisms that underlie developmental robustness is quite complex as developmental progression relays on the differential expression of thousands of genes with diverse functions: regulatory, structural, metabolic, and more (Davidson, 2010; Peter and Davidson, 2011; Pan et al., 2015). If we could decipher the structure and function of entire control networks from the upstream regulatory genes to the downstream effector genes and compare these networks between different species we would better understand developmental and evolutionary processes. However, this approach is very demanding experimentally as it requires the perturbations of every gene in the network followed by a study of the morphological phenotype and molecular consequences of the perturbation. The advents in next-generation sequencing allow researchers to study and compare one aspect of developmental progression – the kinetics of gene expression, throughout all the expressed genes, between different species (Wang et al., 2009; Yanai et al., 2011; Hashimshony et al., 2014; Levin et al., 2016; Malik et al., 2017).

Experimentally, genome-wide quantification of developmental gene expression using RNA-seq can be done, relatively easily, for any species that can be cultured in the lab. Novel approaches of de-novo transcriptome assembly enable the initial quantitative analysis of these data even in species that do not have a reference genome (Haas et al., 2013). However, comparing the expression of thousands of genes between different species is a challenging task. The first challenge is to accurately stage the developmental rates of different species (Yanai et al., 2011; Levin et al., 2012; Gildor and Ben-Tabou de-Leon, 2015; Gildor et al., 2017), or even stage the developmental rates of the same species under various environmental conditions (Runcie et al., 2012; Kuntz and Eisen, 2014). This staging is critical for correct comparison as developmental expression patterns are extremely dynamic and rapid changes happen within short time windows (Gildor and Ben-Tabou de-Leon, 2015; Malik et al., 2017). Therefore, mismatching time points could significantly distort the expression kinetics and result with wrong estimation of the conservation of gene expression. Once the staging is done, the next challenge is to compare the expression profiles of thousands of genes with different functions and identify conservation vs. divergence in gene expression. Here we describe some quantitative approaches that we and others applied to tackle these challenges and some of the unexpected insights on developmental constraints and plasticity that were obtained.

Scaling Developmental Rates Between Related Species

Developmental progression varies between different species due to both genetic and environmental factors (Raff, 1996). For example, increasing temperature quickens the developmental rate without morphological changes in the embryos (Yanai et al., 2011; Runcie et al., 2012; Kuntz and Eisen, 2014; Gildor and Ben-Tabou de-Leon, 2015). Hence, to study conservation and change in gene expression kinetics between different species, we first have to accurately scale the developmental rates between the species. This can be done crudely for closely related species by identifying the times where the embryos show the strongest morphological similarity. However, this approach is not quantitative and is limited to short evolutionary distances. We recently developed a simple mathematical analysis of quantitative expression data of a relatively small set of genes, to scale the developmental rates of two closely related sea urchins that separated from their common ancestor 40 million years ago (Gildor and Ben-Tabou de-Leon, 2015). We have also implemented this approach to larger evolutionary distances of up to 500 million years of parallel evolution, within the echinoderm phylum (Gildor et al., 2017).

First, we study the expression kinetics of a small set of zygotically expressed regulatory genes that are known to have a role in early development [for example, 20 genes that encode transcription factors, signaling pathways or differentiation genes (Gildor and Ben-Tabou de-Leon, 2015; Gildor et al., 2017)]. To make sure that we scale the overall developmental rate, we select genes that are expressed in different embryonic lineages. Measuring the expression kinetics could be done by QPCR or nanostring of whole embryos, as long as the spatial expression pattern of these genes is similar in the two species. At least three biological replicates are necessary since expression levels vary between batches (Gildor and Ben-Tabou de-Leon, 2015). The comparison of the initiation times of these genes between the two species is then used to scale the developmental rates (Gildor and Ben-Tabou de-Leon, 2015). For accurate fitting of gene initiation rates it is recommended that the intervals between the time points will be short enough, especially in the dynamic range (the period of gene activation). A rule of thumb is that the interval would be about a tenth of the entire developmental time studied.

Most zygotic genes have a clear activation curve that can be well fitted with the following sigmoidal function: log(mRNA(t))=a-b1+ec(t-ti) (Figure 1Aa; Yanai et al., 2011). Here a is the final expression level, b is the increase in level relative to the basal expression level, c is the slope of the curve, and ti is the initiation time, that is, half-rise time, the time when the expression level is half of the total increase (Figure 1Aa). The initiation times of all measured genes in each species are estimated using this function. The initiation times in one species is then plotted relative to gene initiation times in the other species. In Figure 1Ab we use published measurements of the initiation times of 22 developmental genes in the sea urchins species, Paracentrotus lividus (P. lividus) and Strongylocentrotus purpuratus (S. purpuratus) (Materna et al., 2010; Gildor and Ben-Tabou de-Leon, 2015). These two species diverged from their common ancestor about 40 million years ago and are geographically separated: S. purpuratus occupies the west coasts of the Pacific Ocean and P. lividus occupies the east coasts of the Atlantic Ocean and the Mediterranean Sea. Yet, despite the genetic and geographic distance their embryonic body plan is highly similar. We measured gene initiation times in the two species based on their expression kinetics up to late gastrula stage [30 hpf in P. lividus and 48 hpf in S. purpuratus (Materna et al., 2010; Gildor and Ben-Tabou de-Leon, 2015)]. The trend-line gives the linear relationship between the developmental time in S. purpuratus and P. lividus: TSp = 2.42 + 1.037 ×TPl. Here the constant, 2.42, corresponds to the shift in the maternal to zygotic transition that is about two and a half hours later in S. purpuratus compared to P. lividus. The slope, 1.037, is a little higher than 1, since the developmental rate in S. purpuratus is slower than in P. lividus, possibly due to the lower culture temperature of S. purpuratus [15 vs. 18°C, (Gildor and Ben-Tabou de-Leon, 2015)].

FIGURE 1
www.frontiersin.org

FIGURE 1. Comparative analyses of developmental expression profiles. (A) Scaling gene expression kinetics. (Aa) The initiation time, ti, of each gene in each species can be measured using a sigmoidal fit, here: a = 0.8; b = 0.6; c = 1 and ti = 8 hpf. (Ab) Gene initiation time in one species vs. initiation time in the other species. (Ac) The estimated linear relationship is used to scale the developmental time points in the two species. Kinetic profiles are shown before (left) and after (right) scaling and expression level normalization. (B) k-means clustering of the temporal profiles of homologous genes in P. lividus and S. purpuratus. Genes are clustered according to their expression profiles in P. lividus (cluster A in the middle). Yellow line indicates the cluster centroid in P. lividus; black lines are expression levels of genes in this cluster in P. lividus (left) and their orthologues in S. purpuratus (right); red lines and blue lines are the median of the temporal profiles of the genes in a cluster in P. lividus and in S. purpuratus, respectively. Secondly, the genes in the clusters are separated into conserved vs. diverged (see text). For example, we detect ribosomal genes in the conserved cluster (left) and ABC transporters in the diverged cluster (right).

Temporal scaling can now be done by matching the time points in the two species using this linear relationship. For proper comparison it is better to also normalize the expression level since different experimental parameters, like primer efficiency and measurement technique can affect the measured absolute levels, regardless of the actual differences. We normalize the expression level of each gene by dividing the level at each time point in the maximal mRNA level measured; so 100% is the maximal expression in this time interval. For example, we plot the expression kinetics before and after scaling and normalization of the regulatory gene, hex, in both species [Figure 1Ac, data from (Gildor and Ben-Tabou de-Leon, 2015)]. hex encodes a transcription factor that participates in the specification of the skeletogenic lineage in the sea urchin (Oliveri et al., 2008). This striking conservation of gene expression dynamics between these two closely related sea urchin species is apparent for most of the 25 developmental genes we studied (Gildor and Ben-Tabou de-Leon, 2015).

We applied this method to larger evolutionary distances, and compared developmental gene expression between sea star and sea urchin that shared a common ancestor about 500 million years ago (Gildor et al., 2017). The sea urchin and the sea star have morphological similarities as well as differences in their larval body plan and accordingly, the expression profiles of regulatory genes show mild heterochronies and loose scaling between these species. In larger evolutionary distances and when the morphologies of the embryos are critically different, it might be possible to use sets of housekeeping genes to scale the developmental rates as the developmental expression of these genes is dynamic and highly conserved between distant species (Kalinka et al., 2010; Levin et al., 2016; Malik et al., 2017). Thus, temporal scaling of developmental rates is essential for correct comparison of expression kinetics between different species and can be done based on a high-resolution quantification of the expression kinetics of a relatively small set of genes.

Comparing the Expression Kinetics Between Thousands of Genes Using Clustering Algorithms

Comparative developmental transcriptomes provide the temporal expression kinetics of all expressed genes and by that significantly broaden the scope of the study (Kalinka et al., 2010; Yanai et al., 2011; Hashimshony et al., 2015; Israel et al., 2016; Levin et al., 2016; Malik et al., 2017). This vast scale of the transcriptomes imposes a challenge to compare the expression of thousands of genes at multiple developmental time points in different species. The use of clustering approaches enables to identify sets of genes/developmental time points that are more similar to each other than to those in other sets. These approaches can be used to answer fundamental questions such as: which time points show the highest similarity in gene expression within and between the species? Which gene sets show conserved expression pattern and which gene sets show divergence? Below we describe how different clustering analyses can be used to address these and other questions.

Principle component analysis (PCA) provides a computational way to identify the experimental time points in different species that show the highest similarity in gene expression, that is, it enables sample grouping (Ringner, 2008). The algorithm identifies directions, called principal components, along which the variation in gene expression throughout the samples, is maximal. Then it uses these components to represent each sample by relatively few numbers instead of by the expression level of thousands of genes. Samples can then be plotted, making it possible to visually assess similarities and differences between experimental time points. This method could be used to guarantee that indeed the temporal staging is accurate and equivalent points in the different species map together (Levin et al., 2012; Israel et al., 2016; Malik et al., 2017). When more than two species are compared, this analysis can be applied to test which species have higher similarity in their gene expression (Israel et al., 2016).

Other approaches group genes together according to the similarity in their expression profiles: Hierarchical clustering uses standard statistical algorithms to group together genes that have similar expression patterns (Eisen et al., 1998). This approach is very informative regarding the different expression profiles that exist in developing embryos (Gildor et al., 2016; Levin et al., 2016). K-means clustering partitions the expression profiles of thousands of genes into k clusters in which each gene belongs to the cluster with the nearest mean (Do and Choi, 2008). The clustering algorithm starts by randomly choosing k expression patterns as initial means for each cluster. Then, each gene is assigned to the cluster that has the closest mean to the gene expression pattern. The new mean is calculated for each cluster and the genes are partitioned again according to the new means. This process is repeated until the cluster means are such that no gene moves from one cluster to another. This analysis results with a partition of all expressed genes into groups of genes that have a similar expression profile, represented by the cluster mean. Both Hierarchical and K-means clustering facilitate the use of gene ontology (GO) enrichment tools to identify functional classes within the clusters, which is very helpful in relating gene expression patterns to gene function (Malik et al., 2017).

k-means can also be used to identify interspecies conservation/divergence in the expression of orthologous genes across several related species. Genes are first clustered according to their expression kinetics in one species and then this clustering is used to identify conservation or change in their orthologues gene expression in other species (Figure 1B; Israel et al., 2016; Malik et al., 2017). Conservation can be defined as the mapping of a gene to the same cluster in multiple species and divergence as the mapping of a gene to different clusters in different species (Israel et al., 2016). It can also be defined by the distance of a gene expression profile in one species from the mean of the cluster of its orthologue in another species, as we exemplify in Figure 1B (Malik et al., 2017). Here we used k-means to compare the developmental transcriptomes measured at seven equivalent developmental time points in P. lividus (between 0 and 36 hpf) and S. purpuratus (between 0 and 48 hpf) (Tu et al., 2012; Malik et al., 2017). First k-means clustering was used to group genes that have similar expression kinetics in P. lividus, into 10 clusters (for example, cluster A, Figure 1B). A gene was identified as conserved if the Euclidian distance of its expression profile in S. purpuratus from its P. lividus cluster-centroid was shorter than a certain cutoff and diverged otherwise (conserved/diverged cluster A in Figure 1B; Malik et al., 2017). The cutoff was defined based on the maximal distances of P. lividus gene expression from the P. lividus centroids (Malik et al., 2017). We then used various GO-enrichment tools to identify functional classes that were over-represented in each conserved or diverged cluster. We discovered that the developmental expression of common housekeeping genes is dynamic and highly conserved between the two species, probably due to developmental constraints on these genes. For example, the expression of ribosomal genes is highly conserved and increases throughout development, possibly to provide the growing embryo with sufficient translation machinery (e.g., rpl23 in conserved cluster A, Figure 1B). On the other hand we discovered that some homeostasis and environmental response genes show diverged expression, either due to evolutionary drift or to allow for adaptation to the local environmental conditions (e.g., the ATP-potassium channel, abcc9h in diverged cluster A, Figure 1B). Thus, k-means clustering is a useful tool to identify typical gene expression patterns, test for interspecies conservation and identify the functional gene classes that show this typical expression pattern.

The Interspecies Correlations of Housekeeping Genes Has a Funnel Shape That Distorts the Developmental Hourglass Pattern

A fundamental quest in evolution and development is to identify the phylotypic stage, a stage in which developing embryos of species in the same phylum display maximal morphological similarity (Raff, 1996). If such a stage exists in different phyla, identifying it could illuminate the developmental bottlenecks that constraint the evolution of body plans (Slack et al., 1993). Furthermore, this would indicate that evolutionary processes within a phylum are somehow distinct form evolutionary processes between phyla (Hejnol and Dunn, 2016).

The phylotypic stage concept was based on comparative morphological studies of selected vertebrates’ embryos by Haeckel (Haeckel, 1874) and Baer (von Baer, 1828), that showed high similarity at the tailbud stage (Slack et al., 1993). These and other studies gave rise to two conservation models; the funnel model assumes the highest conservation at early embryonic stages whereas the hourglass model suggests that intermediate developmental stages are most resistant to evolutionary changes (Slack et al., 1993; Raff, 1996). Later morphological studies, that extended Baer and Haeckel comparison to larger sets of vertebrates’ embryos, showed considerable variability in multiple morphological characteristics during the tailbud stage (Richardson et al., 1997). The advent of next generation sequencing provided the opportunity to resolve this debate by searching for a developmental time that shows the highest interspecies conservation of gene expression that could be defined as the molecular phylotypic stage.

Different computational approaches were used to assess the level of interspecies conservation of gene expression. A method that is commonly applied for this measurement is the Pearson or Spearman correlations between gene expression in equivalent time points of two different species (Irie and Kuratani, 2011; Yanai et al., 2011; Levin et al., 2016; Malik et al., 2017). Pearson correlation measures the linear relationship between two variables while Spearman correlation measures how well the relationship between two variables can be described as a monotonic function, even if it is not linear. In both methods, the higher the correlation, the stronger the conservation of the ratios between the expression levels of different genes. The advantage of these pair-wise comparisons is that they are relatively simple to use and intuitive to interpret. However, when interpolating the results of multiple pair-wise comparisons to assess the conservation patterns between various species, the phylogenetic distances should be explicitly considered, otherwise distant species could lead to a pattern distortion (Dunn et al., 2018). Other methods compute the interspecies divergence in gene expression as the variance in gene expression in each time point across different species (Kalinka et al., 2010; Yanai et al., 2011). These different approaches were used to quantify the interspecies conservation in gene expression between various species across the animal kingdom.

The relationship between gene expression conservation and the morphological phylotypic stage seem to be quite complex. The conservation in gene expression between equivalent time points of two closely related Xenopus species increases with developmental progression, even after the morphological phylotypic stage (Yanai et al., 2011). However, the Spearman correlation of gene expression between four vertebrates’ species were the highest at the mid-developmental stage, in agreement with Haeckel and Baer observation of the morphological phylotypic stage in the vertebrate phylum (Irie and Kuratani, 2011). Relatedly, the interspecies divergence of gene expression between six Drosophila species is the lowest at mid-development, which could indicate that this developmental stage is the phylotypic stage in flies (Kalinka et al., 2010). These vertebrates and drosophila studies support the hourglass model of conservation within a phylum. A recent comparative study of 10 species representing 10 different phyla shows that mid-developmental transition corresponds to the time of lowest interspecies correlations between the phyla (Levin et al., 2016). Thus, intriguingly, the time of highest conservation in gene expression within the phylum is the time of lowest conservation between different phyla (Levin et al., 2016). This conclusion was criticized due to the inclusion of a distant species (ctenophores) that distorts the conservation pattern (Dunn et al., 2018). Regardless of methodological disagreements, the relationship between interspecies conservation of gene expression and the interspecies morphological similarity or divergence is still not clear.

We believe that some of the conflicting observations could result from the different contribution of different functional classes to the pattern of gene expression conservation and how these contributions change with increasing evolutionary distances (Figure 2; Malik et al., 2017). Comparing gene expression between two closely related sea urchin species we noticed that the conservation of developmental gene expression is tightly related to morphological similarity and therefore is diagonal and picks at equivalent developmental stages (Figure 2A). Thus, within the phylum the conservation of these genes is highest at mid-development, resulting with an hourglass pattern (Figure 2A). On the other hand, the conservation of certain sets of housekeeping genes like ribosomal genes, RNA-processing genes and mitochondrial genes was relatively low at early time points and significantly increased after the maternal to zygotic transition (Figure 2B; Malik et al., 2017). Importantly, within these sets, high interspecies correlation was observed between all zygotic time points, regardless of morphological similarity and the correlation increases with developmental time. Other sets of genes with GO terms related to homeostasis and environmental response show intermediate levels of interspecies correlation between morphologically similar and dissimilar time points, possibly reflecting evolutionary drift or adaptation (Figure 2C; Malik et al., 2017). When all expressed genes are considered together, the different conservation patterns merge into a hourglass/funnel shape (Figure 2D; Malik et al., 2017).

FIGURE 2
www.frontiersin.org

FIGURE 2. Different functional classes have different interspecies conservation patterns. (A) The interspecies correlations between genes with developmental GO-terms are strongest between equivalent time points and pick at intermediate developmental times (upper panel) supporting the hourglass model (bottom panel). (B) The interspecies correlation of different housekeeping genes rapidly increases after the maternal to zygotic transition and increases with developmental progression giving rise to a funnel pattern of conservation. (C) The interspecies correlation of homeostasis and environmental response genes are intermediate throughout development regardless of morphological similarity resulting in a broad funnel conservation pattern. (D) When all genes are combined the different contributions merge, and the hourglass pattern is superimposed with the funnel pattern.

With increasing evolutionary distances and diverged morphologies we expect that developmental gene expression would be less conserved and will contribute less to the interspecies correlation patterns. On the other hand, the expression dynamics of housekeeping genes is related to universal embryonic transitions, such as the maternal to zygotic transition or gastrulation that demand a boost in the expression of these genes. Therefore, housekeeping expression dynamics in development is expected to be highly conserved, even between highly divergent species. Thus, at high evolutionary distances, the main contribution to the conservation pattern could come from the housekeeping genes and reflect major embryonic transitions and not morphological similarity vs. divergence. Thus to better understand the evolutionary constraints of gene expression and identify a molecular phylotypic stage, the interspecies correlation should be done separately to different classes of genes.

Conclusion

Large-scale comparative analyses of developmental gene expression inflict a great computational challenge that requires the use of appropriate methods. Novel insights on developmental constraints and flexibility can be gained from a careful analysis of the data. Attention should be given to the dynamic expression of genes that are not considered developmental and seem to evolve under a different set of constraints that is not tightly related to morphology and could be very influential with increasing evolutionary distances.

Author Contributions

Both authors have made a substantial, direct and intellectual contribution to the work, and approved it for publication.

Funding

This work was supported by the Israel Science Foundation grant number 41/14 to SB-Td-L.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank Gregory Cary for his insightful comments on the manuscript and Yarden Ben-Tabou de-Leon for the embryo drawings in Figure 2. We also thank the two reviewers for their thorough review of the manuscripts and their helpful comments that helped us improve the depth and rigor of the paper.

References

Byrne, M., Ho, M., Selvakumaraswamy, P., Nguyen, H. D., Dworjanyn, S. A., Davis, A., et al. (2009). But not pH, compromises sea urchin fertilization and early development under near-future climate change scenarios. Proc. Biol. Sci. 276, 1883–1888. doi: 10.1098/rspb.2008.1935

PubMed Abstract | CrossRef Full Text | Google Scholar

Davidson, E. H. (2010). Emerging properties of animal gene regulatory networks. Nature 468, 911–920. doi: 10.1038/nature09645

PubMed Abstract | CrossRef Full Text | Google Scholar

Do, J. H., and Choi, D. K. (2008). Clustering approaches to identifying gene expression patterns from DNA microarray data. Mol. Cells 25, 279–288.

Google Scholar

Dunn, C. W., Zapata, F., Munro, C., Siebert, S., and Hejnol, A. (2018). Pairwise comparisons across species are problematic when analyzing functional genomic data. Proc. Natl. Acad. Sci. U.S.A. 115, E409–E417. doi: 10.1073/pnas.1707515115

PubMed Abstract | CrossRef Full Text | Google Scholar

Eisen, M. B., Spellman, P. T., Brown, P. O., and Botstein, D. (1998). Cluster analysis and display of genome-wide expression patterns. Proc. Natl. Acad. Sci. U.S.A. 95, 14863–14868. doi: 10.1073/pnas.95.25.14863

CrossRef Full Text | Google Scholar

Gildor, T., and Ben-Tabou de-Leon, S. (2015). Comparative study of regulatory circuits in two sea urchin species reveals tight control of timing and high conservation of expression dynamics. PLoS Genet. 11:e1005435. doi: 10.1371/journal.pgen.1005435

PubMed Abstract | CrossRef Full Text | Google Scholar

Gildor, T., Hinman, V., and Ben-Tabou-De-Leon, S. (2017). Regulatory heterochronies and loose temporal scaling between sea star and sea urchin regulatory circuits. Int. J. Dev. Biol. 61, 347–356. doi: 10.1387/ijdb.160331sb

PubMed Abstract | CrossRef Full Text | Google Scholar

Gildor, T., Malik, A., Sher, N., Avraham, L., and Ben-Tabou de-Leon, S. (2016). Quantitative developmental transcriptomes of the Mediterranean sea urchin Paracentrotus lividus. Mar. Genom. 25, 89–94. doi: 10.1016/j.margen.2015.11.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Grun, D., Wang, Y. L., Langenberger, D., Gunsalus, K. C., and Rajewsky, N. (2005). microRNA target predictions across seven Drosophila species and comparison to mammalian targets. PLoS Comput. Biol. 1:e13. doi: 10.1371/journal.pcbi.0010013

PubMed Abstract | CrossRef Full Text | Google Scholar

Haas, B. J., Papanicolaou, A., Yassour, M., Grabherr, M., Blood, P. D., Bowden, J., et al. (2013). De novo transcript sequence reconstruction from RNA-seq using the Trinity platform for reference generation and analysis. Nat. Protoc. 8, 1494–1512. doi: 10.1038/nprot.2013.084

PubMed Abstract | CrossRef Full Text | Google Scholar

Haeckel, E. (1874). Anthropogenie Oder Entwickelungsgeschichte des Menschen (“Anthropogeny: Or, the Evolutionary History of Man”). Leipzig: W. Engelmann.

Google Scholar

Hashimshony, T., Feder, M., Levin, M., Hall, B. K., and Yanai, I. (2014). Spatiotemporal transcriptomics reveals the evolutionary history of the endoderm germ layer. Nature 519, 219–222. doi: 10.1038/nature13996

PubMed Abstract | CrossRef Full Text | Google Scholar

Hashimshony, T., Feder, M., Levin, M., Hall, B. K., and Yanai, I. (2015). Spatiotemporal transcriptomics reveals the evolutionary history of the endoderm germ layer. Nature 519, 219–222. doi: 10.1038/nature13996

PubMed Abstract | CrossRef Full Text | Google Scholar

Hejnol, A., and Dunn, C. W. (2016). Animal evolution: are phyla real? Curr. Biol. 26, R424–R426. doi: 10.1016/j.cub.2016.03.058

PubMed Abstract | CrossRef Full Text | Google Scholar

Irie, N., and Kuratani, S. (2011). Comparative transcriptome analysis reveals vertebrate phylotypic period during organogenesis. Nat. Commun. 2:248. doi: 10.1038/ncomms1248

PubMed Abstract | CrossRef Full Text | Google Scholar

Israel, J. W., Martik, M. L., Byrne, M., Raff, E. C., Raff, R. A., McClay, D. R., et al. (2016). Comparative developmental transcriptomics reveals rewiring of a highly conserved gene regulatory network during a major life history switch in the sea urchin genus heliocidaris. PLoS Biol. 14:e1002391. doi: 10.1371/journal.pbio.1002391

PubMed Abstract | CrossRef Full Text | Google Scholar

Kalinka, A. T., Varga, K. M., Gerrard, D. T., Preibisch, S., Corcoran, D. L., Jarrells, J., et al. (2010). Gene expression divergence recapitulates the developmental hourglass model. Nature 468, 811–814. doi: 10.1038/nature09634

PubMed Abstract | CrossRef Full Text | Google Scholar

Kitano, J., Lema, S. C., Luckenbach, J. A., Mori, S., Kawagishi, Y., Kusakabe, M., et al. (2010). Adaptive divergence in the thyroid hormone signaling pathway in the stickleback radiation. Curr. Biol. 20, 2124–2130. doi: 10.1016/j.cub.2010.10.050

PubMed Abstract | CrossRef Full Text | Google Scholar

Kuntz, S. G., and Eisen, M. B. (2014). Drosophila embryogenesis scales uniformly across temperature in developmentally diverse species. PLoS Genet. 10:e1004293. doi: 10.1371/journal.pgen.1004293

PubMed Abstract | CrossRef Full Text | Google Scholar

Levin, M., Anavy, L., Cole, A. G., Winter, E., Mostov, N., Khair, S., et al. (2016). The mid-developmental transition and the evolution of animal body plans. Nature 531, 637–641. doi: 10.1038/nature16994

PubMed Abstract | CrossRef Full Text | Google Scholar

Levin, M., Hashimshony, T., Wagner, F., and Yanai, I. (2012). Developmental milestones punctuate gene expression in the Caenorhabditis embryo. Dev. Cell 22, 1101–1108. doi: 10.1016/j.devcel.2012.04.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Levine, M. T., Eckert, M. L., and Begun, D. J. (2011). Whole-genome expression plasticity across tropical and temperate Drosophila melanogaster populations from Eastern Australia. Mol. Biol. Evol. 28, 249–256. doi: 10.1093/molbev/msq197

PubMed Abstract | CrossRef Full Text | Google Scholar

Malik, A., Gildor, T., Sher, N., Layous, M., and Ben-Tabou de-Leon, S. (2017). Parallel embryonic transcriptional programs evolve under distinct constraints and may enable morphological conservation amidst adaptation. Dev. Biol. 430, 202–213. doi: 10.1016/j.ydbio.2017.07.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Martin, S., Richier, S., Pedrotti, M. L., Dupont, S., Castejon, C., Gerakis, Y., et al. (2011). Early development and molecular plasticity in the Mediterranean sea urchin Paracentrotus lividus exposed to CO2-driven acidification. J. Exp. Biol. 214, 1357–1368. doi: 10.1242/jeb.051169

PubMed Abstract | CrossRef Full Text | Google Scholar

Materna, S. C., Nam, J., and Davidson, E. H. (2010). High accuracy, high-resolution prevalence measurement for the majority of locally expressed regulatory genes in early sea urchin development. Gene Expr. Patterns 10, 177–184. doi: 10.1016/j.gep.2010.04.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Okubo, N., Hayward, D. C., Foret, S., and Ball, E. E. (2016). A comparative view of early development in the corals Favia lizardensis, Ctenactis echinata, and Acropora millepora – Morphology, transcriptome, and developmental gene expression. BMC Evol. Biol. 16:48. doi: 10.1186/s12862-016-0615-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Oliveri, P., Tu, Q., and Davidson, E. H. (2008). Global regulatory logic for specification of an embryonic cell lineage. Proc. Natl. Acad. Sci. U.S.A. 105, 5955–5962. doi: 10.1073/pnas.0711220105

PubMed Abstract | CrossRef Full Text | Google Scholar

Pan, T. C., Applebaum, S. L., and Manahan, D. T. (2015). Experimental ocean acidification alters the allocation of metabolic energy. Proc. Natl. Acad. Sci. U.S.A. 112, 4696–4701. doi: 10.1073/pnas.1416967112

PubMed Abstract | CrossRef Full Text | Google Scholar

Pespeni, M. H., Chan, F., Menge, B. A., and Palumbi, S. R. (2013a). Signs of adaptation to local pH conditions across an environmental mosaic in the California current ecosystem. Integr. Compar. Biol. 53, 857–870. doi: 10.1093/icb/ict094

PubMed Abstract | CrossRef Full Text | Google Scholar

Pespeni, M. H., Sanford, E., Gaylord, B., Hill, T. M., Hosfelt, J. D., Jaris, H. K., et al. (2013b). Evolutionary change during experimental ocean acidification. Proc. Natl. Acad. Sci. U.S.A. 110, 6937–6942. doi: 10.1073/pnas.1220673110

PubMed Abstract | CrossRef Full Text | Google Scholar

Peter, I. S., and Davidson, E. H. (2011). Evolution of gene regulatory networks controlling body plan development. Cell 144, 970–985. doi: 10.1016/j.cell.2011.02.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Raff, R. A. (1996). The Shape of Life: Genes. Development and the Evolution of Animal. Chicago, IL: The University of Chicago Press.

Google Scholar

Richardson, M. K., Hanken, J., Gooneratne, M. L., Pieau, C., Raynaud, A., Selwood, L., et al. (1997). There is no highly conserved embryonic stage in the vertebrates: implications for current theories of evolution and development. Anat Embryol. (Berl.) 196, 91–106. doi: 10.1007/s004290050082

PubMed Abstract | CrossRef Full Text | Google Scholar

Ringner, M. (2008). What is principal component analysis? Nat. Biotechnol. 26, 303–304. doi: 10.1038/nbt0308-303

PubMed Abstract | CrossRef Full Text | Google Scholar

Runcie, D. E., Garfield, D. A., Babbitt, C. C., Wygoda, J. A., Mukherjee, S., and Wray, G. A. (2012). Genetics of gene expression responses to temperature stress in a sea urchin gene network. Mol. Ecol. 21, 4547–4562. doi: 10.1111/j.1365-294X.2012.05717.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Shaw, J. R., Hampton, T. H., King, B. L., Whitehead, A., Galvez, F., Gross, R. H., et al. (2014). Natural selection canalizes expression variation of environmentally induced plasticity-enabling genes. Mol. Biol. Evol. 31, 3002–3015. doi: 10.1093/molbev/msu241

PubMed Abstract | CrossRef Full Text | Google Scholar

Slack, J. M., Holland, P. W., and Graham, C. F. (1993). The zootype and the phylotypic stage. Nature 361, 490–492. doi: 10.1038/361490a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Stumpp, M., Hu, M. Y., Melzner, F., Gutowska, M. A., Dorey, N., Himmerkus, N., et al. (2012). Acidified seawater impacts sea urchin larvae pH regulatory systems relevant for calcification. Proc. Natl. Acad. Sci. U.S.A. 109, 18192–18197. doi: 10.1073/pnas.1209174109

PubMed Abstract | CrossRef Full Text | Google Scholar

Tu, Q., Cameron, R. A., Worley, K. C., Gibbs, R. A., and Davidson, E. H. (2012). Gene structure in the sea urchin Strongylocentrotus purpuratus based on transcriptome analysis. Genome Res. 22, 2079–2087. doi: 10.1101/gr.139170.112

PubMed Abstract | CrossRef Full Text | Google Scholar

von Baer, K. E. (1828). Über Entwickelungsgeschichte der Thiere. Beobachtung und Reflexion. Königsberg: Bornträger.

Google Scholar

Wang, Z., Gerstein, M., and Snyder, M. (2009). RNA-Seq: a revolutionary tool for transcriptomics. Nat. Rev. Genet. 10, 57–63. doi: 10.1038/nrg2484

PubMed Abstract | CrossRef Full Text | Google Scholar

Yanai, I., Peshkin, L., Jorgensen, P., and Kirschner, M. W. (2011). Mapping gene expression in two Xenopus species: evolutionary constraints and developmental flexibility. Dev. Cell 20, 483–496. doi: 10.1016/j.devcel.2011.03.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: comparative developmental biology, developmental robustness and plasticity, development and evolution, gene expression kinetics, scaling & modeling, clustering algorithm

Citation: Gildor T and Ben-Tabou de-Leon S (2018) Comparative Studies of Gene Expression Kinetics: Methodologies and Insights on Development and Evolution. Front. Genet. 9:339. doi: 10.3389/fgene.2018.00339

Received: 25 June 2018; Accepted: 09 August 2018;
Published: 22 August 2018.

Edited by:

Frédérique Peronnet, Centre National de la Recherche Scientifique (CNRS), France

Reviewed by:

Pedro Martinez, University of Barcelona, Spain
Jose Maria Martin-Duran, Queen Mary University of London, United Kingdom

Copyright © 2018 Gildor and Ben-Tabou de-Leon. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Smadar Ben-Tabou de-Leon, sben-tab@univ.haifa.ac.il

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.