Transcriptional enhancers are central to the function and evolution of genes and gene regulation. At the organismal level, enhancers play a crucial role in coordinating tissue- and context-dependent gene expression. At the population level, changes in enhancers are thought to be a major driving force that facilitates evolution of diverse traits. An amazing array of diverse traits seen in insect morphology, physiology and behavior has been the subject of research for centuries. Although enhancer studies in insects outside of Drosophila have been limited, recent advances in functional genomic approaches have begun to make such studies possible in an increasing selection of insect species. Here, instead of comprehensively reviewing currently available technologies for enhancer studies in established model organisms such as Drosophila, we focus on a subset of computational and experimental approaches that are likely applicable to non-Drosophila insects, and discuss the pros and cons of each approach. We discuss the importance of validating enhancer function and evaluate several possible validation methods, such as reporter assays and genome editing. Key points and potential pitfalls when establishing a reporter assay system in non-traditional insect models are also discussed. We close with a discussion of how to advance enhancer studies in insects, both by improving computational approaches and by expanding the genetic toolbox in various insects. Through these discussions, this Review provides a conceptual framework for studying the function and evolution of enhancers in non-traditional insect models.

Understanding how genes are regulated is fundamental to various disciplines of biology. In the field of insect science, molecular mechanisms underlying gene regulation are best studied in the fruit fly Drosophila melanogaster. With a suite of sophisticated genetic tools available in this insect (Hales et al., 2015), scientists have been able to decipher complex interactions among genes and their protein products, revealing comprehensive networks of gene interactions and regulations (i.e. gene regulatory networks, GRNs) for various tissues and contexts.

Two innovations in the last two decades have drastically changed the way we study insects beyond Drosophila. The first is the advancement of next-generation sequencing technology, which allows researchers to gather genomic and transcriptomic information from the insect they study relatively easily and even sometimes prior to detailed ‘wet’ investigation. The second is the application of RNA interference (RNAi), and more recently, CRISPR/Cas9 (clustered regularly interspaced short palindromic repeats/CRISPR associated protein 9) genome editing, in various insects (reviewed in Bellés, 2010; Gilles and Averof, 2014). These gene knockdown/knockout techniques now allow loss-of-function (LOF) analyses in many (albeit not all) insects without the need for creating mutants through traditional means, and often with the capability of controlling the timing of gene disruption. With these critical advances, we can now study various tissues and contexts of non-traditional model insects or even non-model insects at the detailed molecular level (Reardon, 2019).

Although researchers are gradually stepping out from Drosophila to explore molecular mechanisms underlying various intriguing processes found in other insects, the knowledge obtained from Drosophila studies continues to play a critical role in insect science. Researchers often use Drosophila GRNs as a starting point (i.e. Drosophila paradigm) and investigate the function of the genes that are homologous/orthologous to the genes in the Drosophila GRNs through RNAi-based LOF studies, often combined with expression analyses, in their insects. This approach has been very fruitful in gaining new insights into gene function and regulation, as well as into the evolution of GRNs among insects, that are difficult to obtain through studying Drosophila alone (Bellés, 2010).

When discussing GRNs, there are two types of components: trans and cis (Fig. 1A). trans components are transcription factors (TFs) and their upstream regulators that provide instructive cues to cells for patterning, differentiation and various other biological processes. In contrast, cis components are non-coding DNA elements that integrate the upstream trans information and determine the expression of the genes downstream in the GRNs. Enhancers (often also called cis-regulatory elements or cis-regulatory modules) are a class of cis components that play a central role in determining spatial and temporal gene expression (Blackwood and Kadonaga, 1998; Buffry et al., 2016; Cho, 2012; Long et al., 2016; Pennacchio et al., 2013; Rickels and Shilatifard, 2018). As mentioned, most current studies in insects outside of Drosophila utilize RNAi-based LOF analyses (or knocking out coding genes via CRISPR/Cas9) as a central approach. This allows for an investigation of GRNs from the trans point of view (Fig. 1A), by inhibiting the function of trans components and assessing their influence on GRNs. However, although it is at least as important to study GRNs from the cis perspective to gain a comprehensive view of gene regulatory mechanisms, the lack of a reliable method to identify enhancers in non-Drosophila insects has made it difficult to study the function and evolution of cis components beyond Drosophila species.

Fig. 1.

Gene regulation and reporter assay. (A) trans and cis components in gene regulation. RNAi allows functional analyses from the trans point of view. (B) A typical reporter assay configuration. Note that the core promoter (red) is a short stretch (∼80 bp) of DNA sequence where the general transcription factors and RNA polymerase are assembled for gene expression. A core promoter itself is typically not sufficient to initiate transcription unless an active enhancer (either proximally or distally located) facilitates the assembly of transcription initiation factors at the core promoter. The term ‘promoter’ (pink and red) is often used to describe the region immediately upstream of the transcription start site when this region contains both the core promoter (red) and a proximally located enhancer (orange), and is sufficient to drive gene expression.

Fig. 1.

Gene regulation and reporter assay. (A) trans and cis components in gene regulation. RNAi allows functional analyses from the trans point of view. (B) A typical reporter assay configuration. Note that the core promoter (red) is a short stretch (∼80 bp) of DNA sequence where the general transcription factors and RNA polymerase are assembled for gene expression. A core promoter itself is typically not sufficient to initiate transcription unless an active enhancer (either proximally or distally located) facilitates the assembly of transcription initiation factors at the core promoter. The term ‘promoter’ (pink and red) is often used to describe the region immediately upstream of the transcription start site when this region contains both the core promoter (red) and a proximally located enhancer (orange), and is sufficient to drive gene expression.

There are several reasons as to why studying enhancers is so challenging in non-Drosophila insects. First, compared with trans components, cis components, especially enhancers, are extremely labile (Li et al., 2007), which makes identification of enhancers based on sequence conservation challenging even among closely related species (Papatsenko et al., 2006) and nearly impossible among species with divergence time beyond ∼60 million years (Kazemian et al., 2014; Li et al., 2007). This is especially problematic for insects, which underwent an early radiation (winged insects had diversified into at least 10 orders by the early Permian, 300 million years ago; Kukalova-Peck, 1991) and have short generation times, leading to limited non-coding homology beyond the genus or family level. Second, functional validation of enhancers often requires the use of modern genetic and genomic tools, which are currently largely absent from most non-Drosophila insects. Because of these hurdles, investigations into the function and evolution of enhancers have been quite limited in insects outside of Drosophila, despite the clear awareness among researchers that changes in enhancers and other cis-regulatory elements play a crucial role in facilitating evolution and diversification of various traits among insects and other organisms (reviewed in Carroll, 2008).

Typically, an enhancer study consists of two steps: (1) the identification of possible enhancer regions (either focusing on a single gene of interest or genome-wide), and (2) the validation and further downstream functional evaluation of enhancer activity in vivo. Some approaches allow functional validation of enhancer activity in the first step, while others require separate in vivo validation experiments. In this Review, we will first summarize currently available approaches to identify possible enhancer regions in insect genomes. Although many of these approaches are technology and resource intensive (i.e. model system-centered), recent advances in genomics and computational biology have started making some of these approaches more accessible to researchers that use insects other than Drosophila as their model. We will discuss the pros and cons of each of these approaches when applied to non-traditional model insects. We will then turn our attention to in vivo validation of enhancer activity in non-traditional insect models and discuss several possible approaches for enhancer validation, such as reporter assays and CRISPR/Cas9-based genome editing. We will use our recent attempt to establish a cross-species compatible reporter construct as a case study, and discuss some of the key points in establishing a reporter assay system in insects outside of Drosophila. Lastly, we will touch on our current effort to advance enhancer studies in insects, both by improving the computational approach to identify possible enhancer regions in insect genomes and by expanding the genetic toolbox for enhancer studies in various insects.

Classic reporter assay

By definition, enhancers are short DNA sequences that act in cis and increase the transcription of a nearby gene regardless of their orientation and the distance from the gene they regulate (Blackwood and Kadonaga, 1998; Pennacchio et al., 2013). The reporter assay takes advantage of this feature and places a candidate enhancer in front of a marker gene that can be easily visualized (i.e. a reporter gene), such as the IacZ gene of Escherichiacoli or fluorescent protein genes, along with a core promoter (Fig. 1B). The enhancer activity of this ‘reporter construct’ can then be assayed in vivo by visualizing the expression of the reporter gene in various tissues and contexts. This approach was first used to investigate the regulation of several segmentation genes in Drosophila, such as fushi tarazu (ftz) (Hiromi et al., 1985) and even skipped (eve) (Goto et al., 1989; Harding et al., 1989), and now has become the ‘gold standard’ approach when evaluating the activity of enhancers in Drosophila and other traditional model organisms (Suryamohan and Halfon, 2015). However, this approach is often inefficient and incomplete as a method to identify enhancers, as it requires the generation of many transgenic lines to be able to survey a sufficient length of the genome, many of which will inevitably only provide negative results.

Genome-wide reporter assay

Despite the arduous and time-consuming nature of the reporter assay, this approach has been used in a genome-wide fashion in Drosophila. Flylight and Fly Enhancers are the two major projects attempting genome-wide enhancer identification through reporter assays, with a focus on brain development and embryogenesis, respectively (Jenett et al., 2012; Kvon et al., 2014; Pfeiffer et al., 2008). The Flylight collection was also used to describe genome-wide enhancer activities in several different developmental contexts (Jenett et al., 2012; Jory et al., 2012; Tokusumi et al., 2017). These projects have identified thousands of functionally validated enhancers and provided us with an overall outlook of the cis-regulatory landscape in the Drosophila genome. Furthermore, over 10,000 lines generated through these projects use the yeast Gal4 transcription factor as the reporter, which allows application of the Gal4-UAS bipartite expression system (Brand and Perrimon, 1993) to enable researchers to misexpress genes and trace the lineage of cells and tissues with unprecedented precision in Drosophila.

The reporter assay system has also been used in a high-throughput setting. An example of a high-throughput approach used in Drosophila is STARR-seq (self-transcribing active regulatory region sequencing) (Arnold et al., 2013; Muerdter et al., 2015). STARR-seq utilizes a library of reporter constructs, which covers the entirety of the Drosophila genome >10 fold. The reporter constructs are designed with the candidate enhancer sequences placed downstream of the core promoter. As the result, active enhancers are directly transcribed, thus serving double-duty as their own reporter genes when transfected cultured cells are subjected to RNA sequencing (RNA-seq) analysis (Fig. 2C). Furthermore, STARR-seq enables identification of enhancers in a quantitative manner, because the number of RNA-seq reads corresponding to each candidate enhancer sequence is directly proportional to the strength of the enhancer activity of the genome fragment.

Fig. 2.

Enhancer identification methods for non-traditional insect models. (A) Classic reporter assay and (B) reporter assay aided by phylogenetic footprinting. The diagram illustrates an example of searching for a wing enhancer, in which a genomic fragment containing a wing enhancer (orange in the reporter construct) drives reporter gene expression (green) in the wing of the transgenic insect. The classic reporter assay requires a survey of a large genomic region (upstream, downstream, introns and sometimes even exons of the gene of interest) to identify an enhancer (A). In phylogenetic footprinting, pair-wise comparisons of genomic sequences among several closely related species (e.g. species a–d in B) allow identification of evolutionarily conserved regions of the genome (VISTA plot in B; the height of the peak corresponds to the degree of conservation; blue highlights the coding region). Evolutionary conservation outside of the coding sequence (highlighted in pink in the VISTA plot) may imply the presence of functional cis-elements, and therefore could help narrow the search. (C) STARR-seq. The STARR reporter is designed in a way that the inserted genomic fragment is transcribed when the fragment acts as an enhancer (orange in the reporter construct) upon transfection into cultured cells. This allows identification of enhancers in a quantitative and genome-wide manner through RNA-seq, as the number of RNA-seq reads corresponding to each candidate enhancer (depicted as red peaks) is directly proportional to the strength of the enhancer activity of the genome fragment. ORF, open reading frame; pA site, polyadenylation site. (D) Enhancer identification through chromatin profiling and computational approaches. FAIRE-seq, ATAC-seq and DNase-seq allow identification of open chromatin regions, while ChIP-seq can be used either to profile genome-wide epigenetic modifications or to identify the binding sites of a transcription factor (TF) of interest. Outcomes of these analyses are often presented as peaks along the genome, where the peak height represents the number of sequence reads that were mapped to the corresponding genomic region. Note that enhancer regions identified through chromatin profiling or computational approaches are still predictions that require functional validation.

Fig. 2.

Enhancer identification methods for non-traditional insect models. (A) Classic reporter assay and (B) reporter assay aided by phylogenetic footprinting. The diagram illustrates an example of searching for a wing enhancer, in which a genomic fragment containing a wing enhancer (orange in the reporter construct) drives reporter gene expression (green) in the wing of the transgenic insect. The classic reporter assay requires a survey of a large genomic region (upstream, downstream, introns and sometimes even exons of the gene of interest) to identify an enhancer (A). In phylogenetic footprinting, pair-wise comparisons of genomic sequences among several closely related species (e.g. species a–d in B) allow identification of evolutionarily conserved regions of the genome (VISTA plot in B; the height of the peak corresponds to the degree of conservation; blue highlights the coding region). Evolutionary conservation outside of the coding sequence (highlighted in pink in the VISTA plot) may imply the presence of functional cis-elements, and therefore could help narrow the search. (C) STARR-seq. The STARR reporter is designed in a way that the inserted genomic fragment is transcribed when the fragment acts as an enhancer (orange in the reporter construct) upon transfection into cultured cells. This allows identification of enhancers in a quantitative and genome-wide manner through RNA-seq, as the number of RNA-seq reads corresponding to each candidate enhancer (depicted as red peaks) is directly proportional to the strength of the enhancer activity of the genome fragment. ORF, open reading frame; pA site, polyadenylation site. (D) Enhancer identification through chromatin profiling and computational approaches. FAIRE-seq, ATAC-seq and DNase-seq allow identification of open chromatin regions, while ChIP-seq can be used either to profile genome-wide epigenetic modifications or to identify the binding sites of a transcription factor (TF) of interest. Outcomes of these analyses are often presented as peaks along the genome, where the peak height represents the number of sequence reads that were mapped to the corresponding genomic region. Note that enhancer regions identified through chromatin profiling or computational approaches are still predictions that require functional validation.

Phylogenetic footprinting

Phylogenetic footprinting is based on the concept that functionally important sequences within a genome, even outside of the coding regions (such as TF binding sites), should be evolutionarily conserved. The fast-evolving nature of insect genomes appears to make the alignment of genomic sequences among multiple insect species challenging. Nonetheless, this approach can be powerful at identifying enhancers when genome sequences from a set of closely related species are available. For example, multiple sequenced genomes within the genus Drosophila, along with the available genome alignments for these species, have made it possible to quickly identify blocks of conserved sequence outside of the coding regions (Frazer et al., 2004; Mayor et al., 2000; Papatsenko et al., 2006; Sosinsky et al., 2007; Stark et al., 2007). However, several studies functionally validating the conserved non-coding sequences point toward a consensus that conservation alone might not be sufficient to efficiently identify enhancers (i.e. not all demonstrated enhancers are well conserved and not all conserved non-coding sequences appear to function as enhancers) (Bergman et al., 2002; Kharchenko et al., 2011; Li et al., 2007; Richards et al., 2005; Roy et al., 2010).

Chromatin profiling

Changes in chromatin status through epigenetic modifications are critical to facilitate precise gene regulation (reviewed in Klemm et al., 2019). Various cis-regulatory elements, including enhancers, are ‘open’ (i.e. nucleosome free) when they are active, so TFs have access to these regions. Several methods exploit this feature of the genome and identify possible enhancer regions through chromatin profiling (reviewed in Klemm et al., 2019; Meyer and Liu, 2014; Suryamohan and Halfon, 2015). DNase-seq (DNase I hypersensitive sites sequencing) uses high sensitivity to DNase as the indicator of open chromatin regions (Boyle et al., 2008). FAIRE-seq (formaldehyde-assisted isolation of regulatory elements, combined with sequencing) and ATAC-seq (assay for transposase-accessible chromatin using sequencing) also identify open chromatin regions. FAIRE-seq uses organic phase separation chemistry to isolate nucleosome-free chromatin away from nucleosome-containing DNA (Giresi et al., 2007; McKay, 2019; McKay and Lieb, 2013), while ATAC-seq uses transposase accessibility as the indicator of open chromatin (Buenrostro et al., 2013).

Another type of chromatin-related method that might be useful to identify enhancers is chromosome conformation capture (3C) (reviewed in de Wit and de Laat, 2012). Hi-C (3C combined with high-throughput sequencing) allows genome-wide investigation of the spatial chromatin organization, including long-distance interactions among multiple loci (Lieberman-Aiden et al., 2009). This technique can be useful in identifying enhancers by analyzing promoter–enhancer interactions (Ron et al., 2017).

Antibody-based enhancer identification

A number of antibody-based methods have proven useful when identifying possible enhancer regions (reviewed in Suryamohan and Halfon, 2015). ChIP-seq (chromatin immunoprecipitation followed by sequencing) is a widely used technique to either (1) identify the binding sites of a specific TF or (2) gain a genome-wide chromatin profile (Ghavi-Helm and Furlong, 2012; Ghavi-Helm et al., 2016; Park, 2009). For the former application, antibodies that specifically recognize the TF of interest are used to identify the regions that are occupied by the TF throughout the genome. Those binding sites are often indicative of the enhancers that are regulated by the investigated TF. For the latter application of ChIP-seq, antibodies against global chromatin modification markers are used. For example, antibodies against histone H3 with its lysine at position 27 acetylated (H3 K27Ac) can be used to identify active chromatin regions, while antibodies against histone H3 with its K27 trimethylated (H3 K27me3) are often useful to identify inactive regions in the genome (Bannister and Kouzarides, 2011). Antibodies against histone acetyltransferase p300 are also often used to identify active chromatin regions (Kharchenko et al., 2011; Nègre et al., 2011; Visel et al., 2009).

More recently, a new antibody-based method, CUT&RUN (cleavage under targets and release using nuclease combined with sequencing), has been developed (Meers et al., 2019; Skene and Henikoff, 2017; Skene et al., 2018). Briefly, in this method, unfixed permeabilized tissues/cells are incubated with antibodies that target a protein of interest (such as TFs or histones with a specific modification). Then, protein A conjugated micrococcal nuclease (MNase) is added, which binds to the antibody and cuts the DNA in its vicinity. The released DNA fragments are isolated through size selection and used for sequencing. CUT&RUN allows researchers to obtain data equivalent to ChIP-seq, but with fewer procedures and much less input tissue.

CRISPR/Cas9-based screening

Unlike RNAi, CRISPR/Cas9-based genome disruption can interrogate not only the function of transcriptionally active regions of the genome but also the non-coding portions, such as enhancers. Several high-throughput strategies have been established to identify possible enhancer regions through CRISPR/Cas9-based genome disruption, many of which use a tiling approach in a cultured cell setting and comprehensively survey a locus of interest with a collection of short guide RNAs (sgRNAs) designed to cover the entirety of the locus (reviewed in Catarino and Stark, 2018; Klein et al., 2018; Lopes et al., 2016). More recently, the next generation of CRISPR/Cas9 technologies, such as CRISPR-based transcriptional activation (CRISPRa) or interference (CRISPRi), have allowed researchers to manipulate the transcription of endogenous loci by taking advantage of the sequence specificity of the CRISPR/Cas9 system (reviewed in Adli, 2018; Pickar-Oliver and Gersbach, 2019). In brief, these techniques utilize a nuclease-inactive version of the Cas9 protein (dCas9) fused with either a transcription activation domain, such as VP64 or p300, or a repressive chromatin modifier domain, such as Krüppel-associated box (KRAB). These dCas9–effector fusion proteins can facilitate transcriptional regulation at any desired genomic site guided by an sgRNA (Gilbert et al., 2013). The initial studies utilizing the dCas9–effector fusion proteins focused on the transcribed regions of the genome [coding regions as well as long non-coding RNA (lncRNA) loci] (e.g. Ewen-Campen et al., 2017; Jia et al., 2018; Lin et al., 2015); however, this technique was later successfully used to modulate enhancer functions (e.g. Thakore et al., 2015, reviewed in Klein et al., 2018; Lopes et al., 2016). Considering that dCas9–effector techniques have successfully been used in a genome-wide fashion (albeit currently limited to a cultured cell setting), these techniques should be adoptable to identify endogenous enhancers in insects, especially if the context of interest can be studied in cell culture.

Evolution of computational approaches

Early approaches to computational enhancer prediction often relied on a limited degree of knowledge of enhancer features, such as evolutionary conservation (i.e. phylogenetic footprints) and/or the tendency of TF binding motifs to cluster within an enhancer (Berman et al., 2002; Halfon et al., 2002; Markstein et al., 2002; also reviewed in Halfon and Michelson, 2002; Markstein and Levine, 2002). These studies resulted in successful identification of enhancers in Drosophila, especially during embryogenesis, but success rates were low and false-positive prediction rates high. In recent years, the field has started to coalesce around supervised machine learning approaches that are trained using one or more features from a known set of enhancers. These features can include the DNA sequence itself, epigenetic information such as histone methylation and acetylation status, DNA methylation status and nucleosome positioning, transcription factor and co-factor binding, and evidence of transcription (e.g. of ‘enhancer RNAs’), among others. Support vector machines (SVMs) and random forest classifiers remain common approaches (e.g. Arbel et al., 2019; Chen et al., 2018; He et al., 2017; Le et al., 2019; Liu et al., 2018), although ‘deep learning’ approaches using artificial neural networks (ANNs) have been increasing in popularity as these methods become more mature and more feasible with current advances in computing power (e.g. Chen et al., 2018; Li et al., 2018; Liu et al., 2016; Min et al., 2017; Yang et al., 2017). In-depth reviews of computational enhancer discovery approaches have been provided elsewhere (Kleftogiannis et al., 2016; Lim et al., 2018; Suryamohan and Halfon, 2015), and the interested reader is directed to these for detailed treatment.

Generic versus specific enhancer prediction

In general, the current computational approaches can be classified into two types: ‘generic’ and ‘specific.’ Generic approaches rely on characteristics likely to be common among all enhancers regardless of particular spatio-temporal specificity, such as histone modifications and chromatin accessibility. This broad applicability means that a method trained on a single set of known enhancers in a particular cell line, or functioning under a given set of biological conditions, is still likely to be effective for enhancer discovery in a different tissue, cell line or physiological milieu. Indeed, these characteristics may be able to carry over across vast evolutionary distances, allowing models trained on insect enhancers to be used for mammalian enhancer discovery (Sethi et al., 2018 preprint), and presumably vice-versa. However, generic approaches primarily provide a large list of sequences with predicted enhancer function, but no information as to what spatial, temporal or physiological characteristics these putative enhancers may have. Specific approaches, in contrast, attempt to discover discrete subsets of enhancers with common activity. While these may include general features such as chromatin accessibility or histone modification status obtained through the use of chromatin profiling techniques (such as FAIRE-seq or ATAC-seq) performed on a specific tissue of interest, they also include specific features such as presence of particular bound TFs or their binding sites, or the DNA sequence itself. Although specific approaches in general find fewer enhancers overall than generic approaches, true-positive prediction rates tend to be similar for both types of methods.

Enhancer prediction independent of experimentally derived features

When considering application to non-traditional insect models, methods that require training based on multiple experimentally derived features are of considerably less utility, as these data sets are often not available, and rarely, if ever, exist for multiple cell types or conditions. Therefore, approaches that rely solely on genome sequence are likely to be the most appealing to researchers that use non-traditional insect models. A number of these are available, most of which fall into the ‘specific’ enhancer discovery class (Chen et al., 2018; Kazemian and Halfon, 2019; Le et al., 2019; Liu et al., 2018). In general, these approaches deconstruct the training sequences into a set of small (e.g. 4–8 nucleotides) subsequences, or ‘k-mers’, which are then evaluated against a similarly deconstructed set of non-enhancer background sequences. With the notable exception of SCRMshaw (Kantorovitz et al., 2009; Kazemian and Halfon, 2019; Kazemian et al., 2011, 2014), most such approaches have not been tested with respect to insect genomes, including that of Drosophila (a somewhat ironic situation given the unmatched availability of empirically confirmed Drosophila enhancers for use as training data; Rivera et al., 2019). Although methods demonstrated to work using vertebrate genomes are expected to function equally well in insects, comparing efficacies is difficult given the different training and validation regimens applied. An evaluation platform for assessing methods using a uniform set of Drosophila training and validation data has recently been described (Asma and Halfon, 2019), and a critical comparison of various approaches would be a valuable addition to the field.

As showcased above, there are a variety of approaches that allow identification of possible enhancer regions from the genomes of insects (and many more that we could not cover here; see Suryamohan and Halfon, 2015 for a more comprehensive review on currently available techniques). However, the options are limited when using non-traditional insect models owing to the early stage of genetic and genomic resource development in these insects. Below, we focus on several options that are more likely applicable to non-traditional insect models, and discuss key points to consider when choosing a method depending on the insect used or the context studied (Figs 2 and 3).

Fig. 3.

A flow chart on enhancer prediction and validation. Blue boxes indicate the approaches that allow identification of candidate enhancer regions. The function of these enhancer candidates can be validated through the approaches indicated by green boxes. Some approaches allow both identification and validation simultaneously (pink box), especially when the context of interest can be studied in cultured cells.

Fig. 3.

A flow chart on enhancer prediction and validation. Blue boxes indicate the approaches that allow identification of candidate enhancer regions. The function of these enhancer candidates can be validated through the approaches indicated by green boxes. Some approaches allow both identification and validation simultaneously (pink box), especially when the context of interest can be studied in cultured cells.

Choice of experimental approaches

Reporter assays

Perhaps the first factor that influences the decision as to which approach to take would be whether one is interested in analyzing multiple loci or focusing on just one locus. A brute-force classic reporter assay-based survey for enhancers is a feasible option when focusing on analyzing the regulation of just one gene (Fig. 2A), assuming that either a valid reporter assay system is available in the study insect or the assay can be performed in Drosophila (see ‘Validating and investigating enhancer function’ below for more detailed discussion). However, the position of enhancers in relation to the gene of interest is quite unpredictable, from tens of thousands of base pairs upstream or downstream of the gene they regulate to inside of an intron or even sometimes within an exon (for example, see Arnold et al., 2013; Kvon et al., 2014), making the reporter assay-based enhancer search time-consuming, tedious and quite risky. Therefore, considering that genome sequences of many insects are now available and long-read sequencing technologies continue to advance (reviewed in Levy and Myers, 2016), it is beneficial to utilize some additional experimental approaches to identify possible enhancer regions even when focusing on only a single locus.

When genome sequences of a set of closely related species (such as within the same genus) are available or can be obtained, phylogenetic footprinting can mitigate the risk of a reporter assay-based enhancer search by providing candidate regions based on evolutionary conservation (Fig. 2B). However, as mentioned, conservation is not always reliable for predicting enhancers. Also, phylogenetic footprinting does not provide any context-dependent information (such as which enhancers are active in which tissues, at which time points or under which physiological conditions). Nonetheless, evolutionary conservation may be informative in a certain genus/family of insects, making phylogenetic footprinting a possible option to consider, particularly as a means of refining the boundaries of a putative enhancer sequence predicted by other methods.

Unfortunately, a genome-wide reporter assay is currently not an option for most insects, as it requires a large workforce, a well-annotated genome, and a highly established and efficient transgenic technique (the Flylight project in Drosophila, for example). However, some unique circumstances make a high-throughput reporter assay a feasible option for genome-wide enhancer identification. For instance, STARR-seq is an option when the context of interest can be studied using a cultured cell line or perhaps even with a tissue that is culturable and transfectable in vitro (Fig. 2C). Therefore, although contexts are limited, when applicable, a high-throughput reporter assay can be a very powerful approach that will allow a comprehensive identification of functionally validated enhancers.

When performing a reporter assay-based enhancer search, whether a brute-force survey, aided by phylogenetic footprinting, or a high-throughput approach, there is a significant caveat in regard to the backbone structure of the reporter construct used in the assay, such as the choice of core promoter. There is no guarantee that reporter constructs previously established in Drosophila are transferable to a different insect of interest. We discuss this point, along with other potential challenges of establishing a reporter assay system in non-traditional insect models, in a later section (see ‘Validating and investigating enhancer function’).

Chromatin profiling

When a well-annotated genome sequence is available, chromatin profiling can provide rich information about the cis-regulatory landscape of the genome of the study insect. Currently, FAIRE-seq and ATAC-seq appear to be the primary approaches when investigating non-traditional insect models (Fig. 2D) [e.g. Aedes (Behura et al., 2016), Anopheles (Pérez-Zamorano et al., 2017), Tribolium (Lai et al., 2018), Heliconius (Lewis and Reed, 2019), Junonia (van der Burg et al., 2019), Bombyx (Zhang et al., 2017c, 2019)] as they do not require any special reagents and can be performed with a relatively small amount of input tissue. For example, we previously performed FAIRE-seq with tissues of the red flour beetle (Tribolium castaneum) and successfully obtained genome-wide chromatin profiles from various tissues and stages of this insect (Fig. 4) (Lai et al., 2018). More than 40,000 open chromatin regions in the Tribolium genome are detected, and comparison of the profiles across the samples revealed a distinct set of open chromatin regions in each tissue and at each stage. Many of these context-dependent openings likely correspond to tissue- and timing-specific enhancers, thus demonstrating the usefulness of chromatin profiling when studying non-traditional model insects.

Fig. 4.

A significant overlap between FAIRE peaks and SCRMshaw predictions. FAIRE profiles and SCRMshaw predictions at the Tribolium sog locus in six different tissues/stages. More examples of the overlap between FAIRE peaks and SCRMshaw predictions can be found in fig. S3 of Lai et al. (2018).

Fig. 4.

A significant overlap between FAIRE peaks and SCRMshaw predictions. FAIRE profiles and SCRMshaw predictions at the Tribolium sog locus in six different tissues/stages. More examples of the overlap between FAIRE peaks and SCRMshaw predictions can be found in fig. S3 of Lai et al. (2018).

With the use of antibodies against global chromatin modification markers, antibody-based methods, such as ChIP-seq, are also powerful at obtaining a genome-wide chromatin landscape (Fig. 2D). These techniques can reveal context-dependent and/or tissue-specific chromatin profiles, which is very useful when identifying possible enhancer regions that are active uniquely in a certain context. The requirement of a large amount of input tissues has been a significant limiting factor when using ChIP-seq (for instance, thousands of discs are likely required for one biological replicate if ChIP-seq is performed with Drosophila imaginal discs); however, CUT&RUN might now allow researchers to perform an equivalent analysis with a much smaller amount of input tissue. Through a combination of these chromatin profiling techniques, the Reed lab has revealed the genome-wide chromatin landscape of Heliconius butterfly wings, a beautiful example of the use of chromatin profiling in a non-traditional insect model (Lewis and Reed, 2019; Lewis et al., 2016).

ChIP-seq can also be used for identifying the binding sites of a particular TF throughout the genome (Fig. 2D). TF binding sites detected by ChIP-seq are often instructive when identifying context-dependent enhancers that are under the regulation of the investigated TF; thus this approach is quite advantageous when you know which TF to study, or which TF possibly regulates the gene of interest. The requirement for a high-quality ‘ChIP-compatible’ antibody against the TF of interest and the need for a large amount of input tissue have been significant drawbacks of this technique; however, the latter can now be bypassed by using CUT&RUN. One important point that requires attention when using TF binding sites to identify enhancers relates to the affinity of TFs to DNA. Recent studies have revealed that low-affinity binding of TFs to DNA can also be critical for gene regulation (Crocker et al., 2015, 2016). These low-affinity TF binding sites might not be readily detected by ChIP-seq and other antibody-based methods (as these techniques rely on strong binding of TFs to DNA), presenting a risk of missing biologically relevant TF binding sites. Moreover, low-affinity TF binding sites are often not evolutionary conserved (Crocker et al., 2015), further compounding the difficulty of finding enhancers.

It is worth emphasizing that the ‘enhancers’ identified through chromatin profiling described in this section, as well as the computationally identified enhancers (Fig. 2D, next section), are all still predictions. Therefore, it is imperative to functionally validate these candidate enhancer regions.

Computational approaches for non-traditional insect models

Computational approaches are an attractive option for use with non-traditional insect models in that they are quick, inexpensive, and in many cases do not rely on extensive empirically derived genomic data. As mentioned, approaches that rely solely on genome sequence (see Box 1 for discussion about how the status of genome assemblies and gene annotations influence enhancer prediction), such as SCRMshaw, are the most appealing. However, there is a significant caveat when applying supervised sequence-based approaches to non-traditional insect models: the acute dearth of training data, as few insect enhancers are known outside of Drosophila. Interestingly, SCRMshaw trained with known Drosophila enhancers was demonstrated to effectively discover enhancers throughout the 345 Mya range of holometabolous insects (Kazemian et al., 2014), indicating that Drosophila enhancers can be useful as training data at least for the genomes of the Holometabola. When SCRMshaw-predicted enhancers from other insects, including bees, wasps, beetles and mosquitoes, are tested in reporter gene assays in transgenic Drosophila, they validate at rates similar to those seen from within-species prediction of Drosophila enhancers (Kazemian et al., 2014; Suryamohan et al., 2016). Direct testing of Tribolium enhancers in transgenic Tribolium confirms that SCRMshaw can find bona fide enhancers cross-species (Lai et al., 2018). Although not tested in insects, Chen et al. (2018) similarly demonstrate that a k-mer-based prediction method trained using data from a single species can be used for enhancer discovery across a range of mammalian genomes. When trained on a tissue-specific enhancer set, their method performed better at discovering enhancers in the same tissue in other species than in different tissues of the same species. Together, these studies indicate that specific enhancer characteristics could be learned and applied in a cross-species setting, and therefore k-mer-based enhancer predictions will be useful when studying non-traditional insect models.

Box 1. Influence of the status of genome assemblies on computational enhancer prediction

An important point to consider when applying computational enhancer prediction to non-traditional insect models is the status of their genome assemblies and gene annotations. Although the count of sequenced insect species is currently ∼470 (i5k: Sequencing Five Thousand Arthropod Genomes; http://i5k.github.io/arthropod_genomes_at_ncbi), assemblies are of varying quality, ranging from the extremely well-assembled Drosophila melanogaster (contig N50=21 Mb) to the poorly assembled meadow spittlebug Philaenus spumarius (contig N50=319 bp), and fewer than 40% have accompanying gene annotation (Li et al., 2019). How effective is enhancer discovery when genome assemblies are highly incomplete? Testing SCRMshaw with simulated dis-assembly of the Drosophila genome has revealed that contig N50s of at least 23,000 bp (which encompasses the upper 50% of current insect assemblies) are sufficient for effective SCRMshaw prediction, with minor loss of sensitivity and negligible increase in false-positive rates (Asma and Halfon, 2019). Therefore, highly complete genome assembly does not appear to be a prerequisite for successful enhancer prediction by SCRMshaw. Requirements for gene annotation are more difficult to assess. Annotation is not strictly necessary for enhancer prediction, but can certainly facilitate it. For example, SCRMshaw disregards coding sequences to focus on the regions that more likely contain enhancers, i.e. non-coding regions. The effect of gene annotation quality on computational enhancer prediction has not been explored.

Integrating experimental and computational approaches

As discussed above, each approach has its strengths and weaknesses. Therefore, the use of multiple strategies (ideally both experimental and computational), and comparison across the outcomes of several different approaches, is likely to be the most fruitful in narrowing down candidate regions to be functionally validated. In Tribolium, we compared the FAIRE profiles with SCRMshaw predictions and found surprisingly high overlaps between these two datasets (Fig. 4) (Lai et al., 2018). However, in the case of Tribolium (but we think this can be generalizable), chromatin profiling provided too many candidate regions (>40,000 peaks across samples), while k-mer-based computational prediction was too stringent and identified a relatively small number of candidate enhancers (∼1200 regions). Nonetheless, having two independent enhancer prediction approaches greatly helped us narrow down the enhancers for functional validation. Adding more tissues and/or using more homogeneous tissues/cell types for chromatin profiling, along with enhancing k-mer-based computational prediction through the use of improved training data and other refinements, will help increase resolution when identifying candidate regions for context-specific enhancers.

Unless a reporter assay system is used to screen for enhancers (Fig. 2A–C), the enhancer regions identified through the methods described above (Fig. 2D), either chromatin profiling or computational approaches, are still predictions that require functional validation. In this section, we discuss several possible validation approaches when studying enhancers of non-traditional insect models, such as the reporter assay and CRISPR/Cas9-based genome editing. We also highlight some key issues and potential pitfalls when establishing a reporter assay system in insects outside of Drosophila.

Testing activity of non-Drosophila enhancers in Drosophila

As mentioned, confirmation of enhancer activity in vivo via a reporter assay is widely considered to be the gold standard when validating enhancer function (Fig. 1B). Since the first application of a reporter assay in Drosophila in the 1980s (Goto et al., 1989; Harding et al., 1989; Hiromi et al., 1985), reporter assays have been used to investigate the regulation and evolution of numerous genes in Drosophila, identifying over 20,000 enhancers (Rivera et al., 2019) and generating a large variety of useful reporter constructs. Although now feasible in a growing number of species (Fraser, 2012), making transgenic lines is a laborious task when using non-traditional insect models. Therefore, considering the ease of making transgenic lines in Drosophila (in part thanks to low-cost commercial injection services) and the availability of established reporter assay systems, the logical first step is to test the activity of possible enhancer regions identified from non-D.melanogaster insects (including various species in the genus Drosophila) in D.melanogaster. This approach has been quite successful when studying enhancer evolution among multiple Drosophila species (e.g. Frankel et al., 2011; Gompel et al., 2005; also see Rebeiz and Williams, 2017; Stern and Frankel, 2013 for review). Some studies have even demonstrated that enhancers from insect orders outside of Diptera work in Drosophila. For example, enhancers of some developmental genes in beetles, honeybees and even spiders were demonstrated to be active in their expected contexts in Drosophila (e.g. Ayyar et al., 2010; Cande et al., 2009a,b; Kazemian et al., 2014; Lai et al., 2018; Prasad et al., 2016; Wolff et al., 1998; Zinzen et al., 2006). These studies show the power of the cross-species reporter assay using Drosophila as an in vivo test tube, but with one unavoidable concern: are these non-Drosophila enhancers really showing biologically relevant activities in Drosophila? Owing to this obvious caveat of using a cross-species reporter assay, it is ideal if the enhancer activity is also tested in the native species.

Establishing a reporter assay in non-traditional insect models

We often assume that the reporter constructs and other genetic tools established in Drosophila are readily transferable to other insects. However, considering the deep divergence and the vast diversity among insect orders, there is no guarantee that these reporter constructs will function properly in other insects, especially those outside of the order Diptera. In fact, several groups including ourselves have encountered various interesting issues when transferring Drosophila constructs to Tribolium (Lai et al., 2018; Schinko et al., 2010). One of the major issues was related to the choice of core promoter (also known as ‘minimal’ or ‘basal’ promoter; reviewed in Kadonaga, 2012; Vo Ngoc et al., 2019). The most widely used core promoter in insects, the core promoter of the Drosophila Heat Shock Protein 70 (Hsp70) gene, did not work reliably in Tribolium, forcing researchers to look for an alternative core promoter. Schinko et al. (2010) identified that a Tribolium-native promoter, the core promoter of Tc-hsp68 (Tc-bhsp68), works well for their Gal4/UAS system, while Lai et al. (2018) determined that a variation of the DSCP (Drosophila Synthetic Core Promoter) properly drives gene expression when placed in a reporter construct (see Box 2 for details). Other components of transgenic constructs, such the choice of untranslated regions (UTRs) or the inclusion of exogenous genes that are often used in modern genetics, could also present problems. For instance, the yeast Gal4 gene has been used routinely in the gene misexpression system (Gal4/UAS system) in Drosophila (Brand and Perrimon, 1993) and has been successfully transferred to mosquitoes (Kokoza and Raikhel, 2011; Lynd and Lycett, 2012; O'Brochta et al., 2012) and silk moths (Imamura et al., 2003) without any major modification. However, when Schinko et al. (2010) worked on transferring the Gal4/UAS system to Tribolium, they noticed that the full-length Gal4 gene does not work in Tribolium. We also confirmed this, even though the Gal4 gene is transcribed in Tribolium (K. D. Deem and Y. Tomoyasu, unpublished data). Schinko et al. (2010) also tried two shorter and more active versions of Gal4, Gal4-VP16 and Gal4Δ (Ma and Ptashne, 1987; Viktorinová and Wimmer, 2007), both of which worked in Tribolium. These outcomes regarding the core promoter and other components highlight the potential difficulty of establishing a reporter assay in non-traditional insect models; however, we hope that the various issues we and others have encountered (such as the deep rabbit hole of core promoters; Box 2) will serve as guidance when working on other insects.

Box 2. A quest to identify a proper core promoter in Tribolium

Since the early time of reporter assays, the core promoter of the Heat Shock Protein 70 (Hsp70) gene has been widely used in Drosophila when an exogenous core promoter is required (Goto et al., 1989; Harding et al., 1989; Hiromi and Gehring, 1987). This core promoter functioned properly in other insects when 3xP3, a synthetic enhancer composed of three repeats of the P3 Pax6 homeodomain binding site (Mishra et al., 2010; Sheng et al., 1997), was used to drive expression in eye- and nervous-related tissues as a marker of transgenesis [e.g. Drosophila virilis (Horn and Wimmer, 2000), Tribolium (Berghammer et al., 1999; Lorenzen et al., 2003), Bombyx mori (Thomas et al., 2002)]. The Dm-hsp70 core promoter was also used for a reporter assay in Tribolium to identify embryonic enhancers that regulate the Tribolium hairy gene (Eckert et al., 2004). However, when establishing the Gal4/UAS system in Tribolium, Schinko et al. (2010) found that the Dm-hsp70 core promoter does not reliably work in their constructs. SCP1 (Super Core Promoter 1, a composition of Drosophila and viral core promoter motifs that was designed to drive a high level of transcription in Hela cells; Juven-Gershon et al., 2006) also did not work well when used in a UAS construct either in Tribolium or in Drosophila (Schinko et al., 2010). These outcomes led Schinko et al. (2010) to try Tribolium-native promoters, the core promoters of Tc-hsp68 (Tc-bhsp68) and Tc-hairy. Although the Tc-hairy core promoter failed to work properly (it drove an unexpected nervous-system-specific expression), the Tc-bhsp68 core promoter worked well with their UAS constructs in Tribolium. Based on the extensive characterization of core promoter compatibility in Tribolium by Schinko et al. (2010), we thought that the Tc-bhsp68 core promoter would be a safe choice when we attempted to establish a reporter assay system in Tribolium. However, surprisingly, the Tc-bhsp68 core promoter failed to work reliably in Tribolium in a reporter construct with a wing enhancer of the nubbin gene (Tc-nub), even though the same construct worked well in Drosophila (Lai et al., 2018). We decided to try DSCP (Drosophila Synthetic Core Promoter), the core promoter that was established for a genome-wide reporter assay in Drosophila (the FlyLight project) (Pfeiffer et al., 2008). This core promoter is a chimera of the SCP and the Drosophila eve gene core promoter, which was shown to work more efficiently, and with a more diverse array of developmental enhancers, as compared with gene-specific core promoters in Drosophila (Pfeiffer et al., 2008). The DSCP also was found to work preferentially with developmental gene enhancers (versus housekeeping gene enhancers) when tested with STARR-seq (Zabidi et al., 2015). The DSCP we used was a variation of this promoter, containing a fragment of Dm-hsp70 promoter downstream of the transcription initiation site in addition to the motifs taken from SCP and the eve core promoter (cloned from the construct used in McKay and Lieb, 2013; see fig. 6 of Lai et al., 2018 for the sequence and annotation of this promoter). This DSCP worked well in our reporter construct both in Tribolium and Drosophila and in two very different developmental contexts (wing development and embryogenesis) in Tribolium, thus allowing us to establish a cross-species compatible reporter assay system (Lai et al., 2018).

It is currently unknown why the Dm-hsp70 and Tc-bhsp68 core promoters did not work properly when used in some transgenic constructs in Tribolium. Interestingly (and confusingly), these core promoters drove various patterns of enhancer–trap expression when inserted in the Tribolium genome (Lai et al., 2018; Lorenzen et al., 2003; Trauner et al., 2009). This indicates that these core promoters are capable of working with a diverse array of enhancers in Tribolium, even though they did not work properly in the tested artificially configured transgenic constructs. Also worth mentioning is the use of gene-specific promoters (not to be confused with ‘core’ promoters, see Fig. 1A). Promoters of several housekeeping genes have been successfully used to drive gene expression in Tribolium (Gilles et al., 2019; Lai et al., 2018; Lorenzen et al., 2002; Rylee et al., 2018; Sarrazin et al., 2012; Schinko et al., 2012; Siebert et al., 2008; Strobl et al., 2018). However, when we tested the Tc-nub promoter with the Tc-nub wing enhancer in a reporter construct, this construct failed to drive any expression either in Tribolium or in Drosophila (Lai et al., 2018). These confusing outcomes regarding promoters might be related to enhancer–core promoter compatibility and/or an optimal distance between the enhancer and the core promoter, which will require detailed investigation in the future.

One aspect that often makes this type of technology transfer so challenging is the absence of reliable positive controls in non-traditional insect models. For instance, we did not have any known Tribolium enhancers that work in the context we study, forcing us to assume that the potential enhancer we identified through a cross-species assay was in fact a true functional Tribolium enhancer and blindly use it as a positive control when we were troubleshooting our reporter constructs (Lai et al., 2018). The enhancers we validated through our cross-species reporter assay (such as Tc-Nub1L) are functional in both Drosophila (Diptera) and Tribolium (Coleoptera); thus these enhancers might serve as positive controls in a wide range of insects (at least in Holometabola). Hopefully the number of positive control enhancers will quickly increase as more enhancer studies are performed in new insect models.

Functional analysis of enhancers through genome editing

Although validation by a reporter assay continues to be the gold standard when studying enhancers, reporter assays do come with several caveats, even beyond the species-specific issues discussed above. For instance, the distance between the enhancer and the core promoter within a reporter construct has been observed in some cases to affect proper enhancer activity (e.g. Small et al., 1993; Swanson et al., 2010). Compatibility between enhancers and core promoters is another potential issue, which can drastically influence the outcome of the assay (e.g. Pfeiffer et al., 2008; also see Zabidi et al., 2015 for an extreme case of compatibility issues between two core promoters, reviewed in Atkinson and Halfon, 2014). Considering these caveats, in addition to the potential hassles described above when establishing a reporter assay system in non-Drosophila insects, an LOF approach through CRISPR/Cas9-based disruption of enhancers is an attractive alternative when validating enhancer function (also discussed in Duester, 2019).

Several CRISPR/Cas9-based methods have been used to investigate the function of enhancers in Drosophila, some of which are described in the previous section. For instance, Xu et al. (2017) have used a split-drive configuration of a CRISPR/Cas9-based gene drive system (dubbed as CopyCat) to replace the wing vein enhancer of the knirps (kni) gene in Drosophila with that of a mutant allele or the homologous enhancers of other dipteran species. Also, some next-generation CRISPR/Cas technologies, such as CRISPRa, have been successfully used in vivo using Drosophila (Ewen-Campen et al., 2017; Jia et al., 2018; Lin et al., 2015). Although not CRISPR/Cas9-based, Crocker and Stern (2013) used a method that is conceptually similar to dCas9-effector technologies (transcription activator-like effectors, TALEs) to interrogate enhancer function in Drosophila, demonstrating that the dCas9-effector technologies can be quite useful when studying insect enhancers.

Although these and other elaborated CRISPR/Cas technologies (reviewed in Bier et al., 2018) are attractive, a simple CRISPR/Cas9-based knockout via NHEJ (non-homologous end joining) may be most appealing to researchers that use non-traditional insect models owing to the challenging nature of implementing these new technologies in insects outside of Drosophila. Since the first breakthrough application of the CRISPR/Cas9 system in genome editing in 2012 and 2013 (Cong et al., 2013; Jinek et al., 2012), CRISPR/Cas9-based knockout techniques have already been applied to various orders of non-traditional model insects and other arthropod species [e.g. Coleoptera (Gilles et al., 2015), Lepidoptera (Connahs et al., 2019; Mazo-Vargas et al., 2017; Prakash and Monteiro, 2018; Wei et al., 2014; Zhang and Reed, 2016), Hymenoptera (Trible et al., 2017; Yan et al., 2017), Orthoptera (Watanabe et al., 2017), Zygentoma (Ohde et al., 2018) and crustacean species (Martin et al., 2016; Nakanishi et al., 2014), reviewed in Gantz and Akbari, 2018; Gilles and Averof, 2014], showing the relative ease of adopting this technique in insects.

The idea of using CRISPR/Cas9 knockout to validate enhancer function is straightforward: remove/disrupt the candidate enhancer from the genome and evaluate the resulting phenotype. There are several points that require attention when designing an enhancer knockout experiment using non-traditional insect models. From a technical point of view, researchers need to decide whether to analyze the mutant phenotype in the somatic cells of G0 insects or in established mutant lines. Analyzing in G0 is beneficial for insects with difficulties in husbandry and/or a longer generation time, as it allows bypassing multiple generations of crosses to establish a mutant strain and analyzing the outcome immediately in the individuals that received CRISPR/Cas9 injection. This approach has been very successful in butterflies (e.g. Connahs et al., 2019; Mazo-Vargas et al., 2017; Prakash and Monteiro, 2018; Zhang and Reed, 2016; Zhang et al., 2017a,b) as well as in other species such as in the crustacean Parhyale (Bruce and Patel, 2018 preprint; Clark-Hachtel and Tomoyasu, 2017 preprint; Martin et al., 2016). However, to be able to detect mutant phenotypes in G0, (1) genome editing events need to happen in a large enough number of somatic cells and (2) mutant phenotypes need to be clearly visible (e.g. pigmentation and patterning defect, loss of tissues) in the context of study. In other words, the outcome is interpretable likely only when genome editing works properly and results in a clear mutant phenotype, thus presenting a potential caveat of analyzing mutant phenotypes in G0. Establishing a mutant line is a safer option if genetics of your insects allows, and if a valid screening scheme is available to identify and track mutations. A stable line allows for more quantitative measures to be brought to bear, such as qRT-PCR to measure changes in gene expression. Replacing and/or disrupting the targeted enhancer with a marker construct (such as a fluorescent gene driven by 3xP3) might allow an easier tracking of the mutation compared with the mutation created through a simple NHEJ knockout, which likely requires genomic PCR-based methods to identify the mutation (unless the mutation results in a haplo-insufficient, visible and viable phenotype). Although the HDR (homology directed repair) knock-in approach appears to suffer from low efficiency when used in non-traditional insect models (C. M. Clark-Hachtel, K. D. Deem and Y. Tomoyasu, unpublished data), various strategies have been developed to improve success rates (e.g. Aird et al., 2018; Savic et al., 2018, also see Bier et al., 2018; Liu et al., 2019 for review of additional methods to increase the HDR efficiency), some of which might be worth pursuing when performing HDR knock-in in insects. In addition, NHEJ knock-in (Auer et al., 2014; Watanabe et al., 2017), including the CRISPaint (CRISPR-assisted insertion tagging) system (Bosch et al., 2020; Schmid-Burgk et al., 2016), or MMEJ (microhomology mediated end joining) knock-in (e.g. Nakade et al., 2014) might facilitate a more efficient disruption of enhancers while also tagging the genome-editing event with a visible marker.

A further aspect that requires attention when knocking out enhancers is related to the nature of enhancers. It has been shown that multiple enhancers sometimes act redundantly (dubbed as ‘shadow enhancers’), which fosters a robust and stable gene expression (Perry et al., 2010) and might also facilitate evolution of novel traits (Hong et al., 2008). Enhancer redundancy appears to be pervasive among developmental genes (Cannavò et al., 2016). This may cause an issue when performing an enhancer knockout, as targeting one enhancer might not be sufficient to cause any visible abnormalities, which is especially problematic when analyzing in G0. Whether these features of enhancers found in Drosophila are conserved in other insects is not known, which is one of the very reasons why it is essential to study enhancers in various organisms. Also, considering that knocking out enhancers also presents some caveats, it would be ideal to validate enhancer function via multiple methods, such as a reporter assay and knockout, through which both necessity and sufficiency of an enhancer can be evaluated (discussed in detail in Catarino and Stark, 2018; Halfon, 2019).

With the recent confluence of effective enhancer-discovery approaches established in Drosophila, and the sequencing of numerous insect genomes (Thomas et al., 2018 preprint), the time is ripe to start broadening the investigation of enhancers and other cis-regulatory mechanisms to a wider range of insects. There are several critical areas that should receive high priority for active development to make enhancer studies readily possible in non-traditional insect models.

On the computational side, these areas include (1) the use of improved training data by incorporating the latest available data (such as the data from the REDfly database; Rivera et al., 2019), and (2) better integration of computational and experimental enhancer-discovery methods (such as ATAC-seq and FAIRE-seq). Another interesting area to pursue would be to leverage comparative genomics in improving the accuracy of computational predictions. As mentioned, enhancer sequences are frequently alignable between closely related species, but conservation diminishes with increasing evolutionary distance. Nevertheless, enhancer locations are often maintained in equivalent locations despite the inability to directly align the enhancer sequences themselves (Cande et al., 2009a; Kazemian et al., 2014). By prioritizing those enhancer predictions among moderately related species that fall into the same approximate genomic location, we should be able to not only reduce false-positive predictions, but also build up sets of evolutionarily related but non-alignable enhancers. The latter will provide an unprecedented resource for probing the evolution of regulatory sequences. With its average 75% success rate (Kazemian et al., 2014), SCRMshaw has emerged as a promising first-line tool for identifying enhancer sequences in non-traditional insect models, whose prediction capacity will be further improved through these areas of focus.

Although advances in computational and genomics approaches have begun to enable enhancer studies in insects outside of Drosophila, the underdeveloped nature of functional genetics and genomics tools in non-traditional insect models still hinders researchers from functionally dissecting diverse mechanisms underlying cis-regulation and investigating how changes in cis-regulation have contributed (and are contributing) to the evolution of various traits at the detailed molecular level. The cross-species compatible reporter assay construct we previously established (Lai et al., 2018) is a step forward towards performing enhancer studies in various insects, and there are several areas that we can explore to continue our progress for a better implementation of a reporter assay system and other functional genetics tools in non-traditional insect models. The first area is related to core promoters. As we discussed extensively in Box 2, it is crucial to choose the right core promoter that fits the gene, context and species of the study. The modified DSCP we used in our study worked well in two species (a coleopteran and a dipteran, representing a large span of the Holometabola) and in two developmental contexts (appendage development and embryogenesis), suggesting that this core promoter can be used in a wide taxonomy of insects. Nonetheless, the DSCP is mainly constructed with Drosophila core promoter motifs (Juven-Gershon et al., 2006; Lai et al., 2018; Pfeiffer et al., 2008); it might therefore be possible to tailor a more efficient core promoter for each species by a similar strategy. It may also be possible to design a universal core promoter that works across multiple orders of insects. The second area is to expand the genetic toolkit for non-Drosophila insects. Tissue- and context-specific enhancers identified through enhancer studies will be essential when developing additional genetic tools and resources. With the use of these enhancers combined with some modifications to reporter constructs, we will be able to build various modern genetic tools useful for lineage tracing, gene misexpression, tissue-specific RNAi and CRISPR, and beyond. Developing these functional genetic tools will further accelerate investigation of enhancers as well as of many other aspects of biology in non-traditional insect models.

Armed with ample genetic, genomic and computational tools and resources, studies in Drosophila have revealed a wealth of intriguing aspects of enhancer function and evolution. In this Review, we mainly focused on how to ‘find’ enhancers in non-traditional insect models. This is just the first step in investigating the amazing array of diverse traits found in insects from the cis point of view, a widely unexplored area of biology.

Y.T. thanks Leslie Vosshall, Michael Dickinson and Julian Dow for the kind invitation to the JEB Symposium on Genome Editing for Comparative Physiology, Michaela Handel for her help on facilitating the symposium, and the editors of JEB for organizing the special issue derived from the symposium. We also thank Kevin Deem and other members of the Tomoyasu lab for helpful comments.

Funding

This work was supported by the National Science Foundation (NSF) (grant IOS1557936 to Y.T.) and the U.S. Department of Agriculture (USDA) (grant 2018-08230 to M.S.H. and Y.T.).

Adli
,
M.
(
2018
).
The CRISPR tool kit for genome editing and beyond
.
Nat. Commun.
9
,
1911
.
Aird
,
E. J.
,
Lovendahl
,
K. N.
,
St Martin
,
A.
,
Harris
,
R. S.
and
Gordon
,
W. R.
(
2018
).
Increasing Cas9-mediated homology-directed repair efficiency through covalent tethering of DNA repair template
.
Commun. Biol.
1
,
54
.
Arbel
,
H.
,
Basu
,
S.
,
Fisher
,
W. W.
,
Hammonds
,
A. S.
,
Wan
,
K. H.
,
Park
,
S.
,
Weiszmann
,
R.
,
Booth
,
B. W.
,
Keranen
,
S. V.
,
Henriquez
,
C.
, et al. 
(
2019
).
Exploiting regulatory heterogeneity to systematically identify enhancers with high accuracy
.
Proc. Natl Acad. Sci. USA
116
,
900
-
908
.
Arnold
,
C. D.
,
Gerlach
,
D.
,
Stelzer
,
C.
,
Boryn
,
L. M.
,
Rath
,
M.
and
Stark
,
A.
(
2013
).
Genome-wide quantitative enhancer activity maps identified by STARR-seq
.
Science
339
,
1074
-
1077
.
Asma
,
H.
and
Halfon
,
M. S.
(
2019
).
Computational enhancer prediction: evaluation and improvements
.
BMC Bioinformatics
20
,
174
.
Atkinson
,
T. J.
and
Halfon
,
M. S.
(
2014
).
Regulation of gene expression in the genomic context
.
Comput. Struct. Biotechnol. J.
9
,
e201401001
.
Auer
,
T. O.
,
Duroure
,
K.
,
De Cian
,
A.
,
Concordet
,
J.-P.
and
Del Bene
,
F.
(
2014
).
Highly efficient CRISPR/Cas9-mediated knock-in in zebrafish by homology-independent DNA repair
.
Genome Res.
24
,
142
-
153
.
Ayyar
,
S.
,
Negre
,
B.
,
Simpson
,
P.
and
Stollewerk
,
A.
(
2010
).
An arthropod cis-regulatory element functioning in sensory organ precursor development dates back to the Cambrian
.
BMC Biol.
8
,
127
.
Bannister
,
A. J.
and
Kouzarides
,
T.
(
2011
).
Regulation of chromatin by histone modifications
.
Cell Res.
21
,
381
-
395
.
Behura
,
S. K.
,
Sarro
,
J.
,
Li
,
P.
,
Mysore
,
K.
,
Severson
,
D. W.
,
Emrich
,
S. J.
and
Duman-Scheel
,
M.
(
2016
).
High-throughput cis-regulatory element discovery in the vector mosquito Aedes aegypti
.
BMC Genomics
17
,
341
.
Bellés
,
X.
(
2010
).
Beyond Drosophila: RNAi in vivo and functional genomics in insects
.
Annu. Rev. Entomol.
55
,
111
-
128
.
Berghammer
,
A. J.
,
Klingler
,
M.
and
Wimmer
,
E. A.
(
1999
).
A universal marker for transgenic insects
.
Nature
402
,
370
-
371
.
Bergman
,
C. M.
,
Pfeiffer
,
B. D.
,
Rincón-Limas
,
D. E.
,
Hoskins
,
R. A.
,
Gnirke
,
A.
,
Mungall
,
C. J.
,
Wang
,
A. M.
,
Kronmiller
,
B.
,
Pacleb
,
J.
,
Park
,
S.
, et al. 
(
2002
).
Assessing the impact of comparative genomic sequence data on the functional annotation of the Drosophila genome
.
Genome Biol.
3
,
RESEARCH0086
.
Berman
,
B. P.
,
Nibu
,
Y.
,
Pfeiffer
,
B. D.
,
Tomancak
,
P.
,
Celniker
,
S. E.
,
Levine
,
M.
,
Rubin
,
G. M.
and
Eisen
,
M. B.
(
2002
).
Exploiting transcription factor binding site clustering to identify cis-regulatory modules involved in pattern formation in the Drosophila genome
.
Proc. Natl Acad. Sci. USA
99
,
757
-
762
.
Bier
,
E.
,
Harrison
,
M. M.
,
O'Connor-Giles
,
K. M.
and
Wildonger
,
J.
(
2018
).
Advances in engineering the fly genome with the CRISPR-Cas system
.
Genetics
208
,
1
-
18
.
Blackwood
,
E. M.
and
Kadonaga
,
J. T.
(
1998
).
Going the distance: a current view of enhancer action
.
Science
281
,
60
-
63
.
Bosch
,
J. A.
,
Colbeth
,
R.
,
Zirin
,
J.
and
Perrimon
,
N.
(
2020
).
Gene knock-ins in Drosophila using homology-independent insertion of universal donor plasmids
.
Genetics
214
,
75
-
89
.
Boyle
,
A. P.
,
Davis
,
S.
,
Shulha
,
H. P.
,
Meltzer
,
P.
,
Margulies
,
E. H.
,
Weng
,
Z.
,
Furey
,
T. S.
and
Crawford
,
G. E.
(
2008
).
High-resolution mapping and characterization of open chromatin across the genome
.
Cell
132
,
311
-
322
.
Brand
,
A. H.
and
Perrimon
,
N.
(
1993
).
Targeted gene expression as a means of altering cell fates and generating dominant phenotypes
.
Development
118
,
401
-
415
.
Bruce
,
H. S.
and
Patel
,
N. H.
(
2018
).
Insect wings and body wall evolved from ancient leg segments
.
bioRxiv
.
Buenrostro
,
J. D.
,
Giresi
,
P. G.
,
Zaba
,
L. C.
,
Chang
,
H. Y.
and
Greenleaf
,
W. J.
(
2013
).
Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position
.
Nat. Methods
10
,
1213
-
1218
.
Buffry
,
A. D.
,
Mendes
,
C. C.
and
McGregor
,
A. P.
(
2016
).
The functionality and evolution of eukaryotic transcriptional enhancers
.
Adv. Genet.
96
,
143
-
206
.
Cande
,
J.
,
Goltsev
,
Y.
and
Levine
,
M. S.
(
2009a
).
Conservation of enhancer location in divergent insects
.
Proc. Natl. Acad. Sci. USA
106
,
14414
-
14419
.
Cande
,
J. D.
,
Chopra
,
V. S.
and
Levine
,
M.
(
2009b
).
Evolving enhancer-promoter interactions within the tinman complex of the flour beetle, Tribolium castaneum
.
Development
136
,
3153
-
3160
.
Cannavò
,
E.
,
Khoueiry
,
P.
,
Garfield
,
D. A.
,
Geeleher
,
P.
,
Zichner
,
T.
,
Gustafson
,
E. H.
,
Ciglar
,
L.
,
Korbel
,
J. O.
and
Furlong
,
E. E. M.
(
2016
).
Shadow enhancers are pervasive features of developmental regulatory networks
.
Curr. Biol.
26
,
38
-
51
.
Carroll
,
S. B.
(
2008
).
Evo-devo and an expanding evolutionary synthesis: a genetic theory of morphological evolution
.
Cell
134
,
25
-
36
.
Catarino
,
R. R.
and
Stark
,
A.
(
2018
).
Assessing sufficiency and necessity of enhancer activities for gene expression and the mechanisms of transcription activation
.
Genes Dev.
32
,
202
-
223
.
Chen
,
L.
,
Fish
,
A. E.
and
Capra
,
J. A.
(
2018
).
Prediction of gene regulatory enhancers across species reveals evolutionarily conserved sequence properties
.
PLoS Comput. Biol.
14
,
e1006484
.
Cho
,
K. W. Y.
(
2012
).
Enhancers
.
Wiley Interdiscip. Rev. Dev. Biol.
1
,
469
-
478
.
Clark-Hachtel
,
C. M.
and
Tomoyasu
,
Y.
(
2017
).
Two sets of wing homologs in the crustacean, Parhyale hawaiensis
.
bioRxiv
,
1
-
9
.
Cong
,
L.
,
Ran
,
F. A.
,
Cox
,
D.
,
Lin
,
S.
,
Barretto
,
R.
,
Habib
,
N.
,
Hsu
,
P. D.
,
Wu
,
X.
,
Jiang
,
W.
,
Marraffini
,
L. A.
, et al. 
(
2013
).
Multiplex genome engineering using CRISPR/Cas systems
.
Science
339
,
819
-
823
.
Connahs
,
H.
,
Tlili
,
S.
,
van Creij
,
J.
,
Loo
,
T. Y. J.
,
Banerjee
,
T. D.
,
Saunders
,
T. E.
and
Monteiro
,
A.
(
2019
).
Activation of butterfly eyespots by Distal-less is consistent with a reaction-diffusion process
.
Development
146
,
dev169367
.
Crocker
,
J.
and
Stern
,
D. L.
(
2013
).
TALE-mediated modulation of transcriptional enhancers in vivo
.
Nat. Methods
10
,
762
-
767
.
Crocker
,
J.
,
Abe
,
N.
,
Rinaldi
,
L.
,
McGregor
,
A. P.
,
Frankel
,
N.
,
Wang
,
S.
,
Alsawadi
,
A.
,
Valenti
,
P.
,
Plaza
,
S.
,
Payre
,
F.
, et al. 
(
2015
).
Low affinity binding site clusters confer hox specificity and regulatory robustness
.
Cell
160
,
191
-
203
.
Crocker
,
J.
,
Noon
,
E. P.-B.
and
Stern
,
D. L.
(
2016
).
The soft touch: low-affinity transcription factor binding sites in development and evolution
.
Curr. Top. Dev. Biol.
117
,
455
-
469
.
de Wit
,
E.
and
de Laat
,
W.
(
2012
).
A decade of 3C technologies: insights into nuclear organization
.
Genes Dev.
26
,
11
-
24
.
Duester
,
G.
(
2019
).
Knocking out enhancers to enhance epigenetic research
.
Trends Genet.
35
,
89
.
Eckert
,
C.
,
Aranda
,
M.
,
Wolff
,
C.
and
Tautz
,
D.
(
2004
).
Separable stripe enhancer elements for the pair-rule gene hairy in the beetle Tribolium
.
EMBO Rep.
5
,
638
-
642
.
Ewen-Campen
,
B.
,
Yang-Zhou
,
D.
,
Fernandes
,
V. R.
,
González
,
D. P.
,
Liu
,
L.-P.
,
Tao
,
R.
,
Ren
,
X.
,
Sun
,
J.
,
Hu
,
Y.
,
Zirin
,
J.
, et al. 
(
2017
).
Optimized strategy for in vivo Cas9-activation in Drosophila
.
Proc. Natl. Acad. Sci. USA
114
,
9409
-
9414
.
Frankel
,
N.
,
Erezyilmaz
,
D. F.
,
McGregor
,
A. P.
,
Wang
,
S.
,
Payre
,
F.
and
Stern
,
D. L.
(
2011
).
Morphological evolution caused by many subtle-effect substitutions in regulatory DNA
.
Nature
474
,
598
-
603
.
Fraser
,
M. J.
(
2012
).
Insect transgenesis: current applications and future prospects
.
Annu. Rev. Entomol.
57
,
267
-
289
.
Frazer
,
K. A.
,
Pachter
,
L.
,
Poliakov
,
A.
,
Rubin
,
E. M.
and
Dubchak
,
I.
(
2004
).
VISTA: computational tools for comparative genomics
.
Nucleic Acids Res.
32
,
W273
-
W279
.
Gantz
,
V. M.
and
Akbari
,
O. S.
(
2018
).
Gene editing technologies and applications for insects
.
Curr. Opin. Insect Sci.
28
,
66
-
72
.
Ghavi-Helm
,
Y.
and
Furlong
,
E. E. M.
(
2012
).
Analyzing transcription factor occupancy during embryo development using ChIP-seq
.
Methods Mol. Biol.
786
,
229
-
245
.
Ghavi-Helm
,
Y.
,
Zhao
,
B.
and
Furlong
,
E. E. M.
(
2016
).
Chromatin immunoprecipitation for analyzing transcription factor binding and histone modifications in Drosophila
.
Methods Mol. Biol.
1478
,
263
-
277
.
Gilbert
,
L. A.
,
Larson
,
M. H.
,
Morsut
,
L.
,
Liu
,
Z.
,
Brar
,
G. A.
,
Torres
,
S. E.
,
Stern-Ginossar
,
N.
,
Brandman
,
O.
,
Whitehead
,
E. H.
,
Doudna
,
J. A.
, et al. 
(
2013
).
CRISPR-mediated modular RNA-guided regulation of transcription in eukaryotes
.
Cell
154
,
442
-
451
.
Gilles
,
A. F.
and
Averof
,
M.
(
2014
).
Functional genetics for all: engineered nucleases, CRISPR and the gene editing revolution
.
Evodevo
5
,
43
.
Gilles
,
A. F.
,
Schinko
,
J. B.
and
Averof
,
M.
(
2015
).
Efficient CRISPR-mediated gene targeting and transgene replacement in the beetle Tribolium castaneum
.
Development
142
,
2832
-
2839
.
Gilles
,
A. F.
,
Schinko
,
J. B.
,
Schacht
,
M. I.
,
Enjolras
,
C.
and
Averof
,
M.
(
2019
).
Clonal analysis by tunable CRISPR-mediated excision
.
Development
146
,
dev170969
.
Giresi
,
P. G.
,
Kim
,
J.
,
McDaniell
,
R. M.
,
Iyer
,
V. R.
and
Lieb
,
J. D.
(
2007
).
FAIRE (Formaldehyde-Assisted Isolation of Regulatory Elements) isolates active regulatory elements from human chromatin
.
Genome Res.
17
,
877
-
885
.
Gompel
,
N.
,
Prud'homme
,
B.
,
Wittkopp
,
P. J.
,
Kassner
,
V. A.
and
Carroll
,
S. B.
(
2005
).
Chance caught on the wing: cis-regulatory evolution and the origin of pigment patterns in Drosophila
.
Nature
433
,
481
-
487
.
Goto
,
T.
,
Macdonald
,
P.
and
Maniatis
,
T.
(
1989
).
Early and late periodic patterns of even skipped expression are controlled by distinct regulatory elements that respond to different spatial cues
.
Cell
57
,
413
-
422
.
Hales
,
K. G.
,
Korey
,
C. A.
,
Larracuente
,
A. M.
and
Roberts
,
D. M.
(
2015
).
Genetics on the fly: a primer on the Drosophila model system
.
Genetics
201
,
815
-
842
.
Halfon
,
M. S.
(
2019
).
Studying transcriptional enhancers: the founder fallacy, validation creep, and other biases
.
Trends Genet.
35
,
93
-
103
.
Halfon
,
M. S.
and
Michelson
,
A. M.
(
2002
).
Exploring genetic regulatory networks in metazoan development: methods and models
.
Physiol. Genomics
10
,
131
-
143
.
Halfon
,
M. S.
,
Grad
,
Y.
,
Church
,
G. M.
and
Michelson
,
A. M.
(
2002
).
Computation-based discovery of related transcriptional regulatory modules and motifs using an experimentally validated combinatorial model
.
Genome Res.
12
,
1019
-
1028
.
Harding
,
K.
,
Hoey
,
T.
,
Warrior
,
R.
and
Levine
,
M.
(
1989
).
Autoregulatory and gap gene response elements of the even-skipped promoter of Drosophila
.
EMBO J.
8
,
1205
-
1212
.
He
,
Y.
,
Gorkin
,
D. U.
,
Dickel
,
D. E.
,
Nery
,
J. R.
,
Castanon
,
R. G.
,
Lee
,
A. Y.
,
Shen
,
Y.
,
Visel
,
A.
,
Pennacchio
,
L. A.
,
Ren
,
B.
, et al. 
(
2017
).
Improved regulatory element prediction based on tissue-specific local epigenomic signatures
.
Proc. Natl. Acad. Sci. USA
114
,
E1633
-
E1640
.
Hiromi
,
Y.
and
Gehring
,
W. J.
(
1987
).
Regulation and function of the Drosophila segmentation gene fushi tarazu
.
Cell
50
,
963
-
974
.
Hiromi
,
Y.
,
Kuroiwa
,
A.
and
Gehring
,
W. J.
(
1985
).
Control elements of the Drosophila segmentation gene fushi tarazu
.
Cell
43
,
603
-
613
.
Hong
,
J.-W.
,
Hendrix
,
D. A.
and
Levine
,
M. S.
(
2008
).
Shadow enhancers as a source of evolutionary novelty
.
Science
321
,
1314-1314
.
Horn
,
C.
and
Wimmer
,
E. A.
(
2000
).
A versatile vector set for animal transgenesis
.
Dev. Genes Evol.
210
,
630
-
637
.
Imamura
,
M.
,
Nakai
,
J.
,
Inoue
,
S.
,
Quan
,
G. X.
,
Kanda
,
T.
and
Tamura
,
T.
(
2003
).
Targeted gene expression using the GAL4/UAS system in the silkworm Bombyx mori
.
Genetics
165
,
1329
-
1340
.
Jenett
,
A.
,
Rubin
,
G. M.
,
Ngo
,
T.-T. B.
,
Shepherd
,
D.
,
Murphy
,
C.
,
Dionne
,
H.
,
Pfeiffer
,
B. D.
,
Cavallaro
,
A.
,
Hall
,
D.
,
Jeter
,
J.
, et al. 
(
2012
).
A GAL4-driver line resource for Drosophila neurobiology
.
Cell Rep.
2
,
991
-
1001
.
Jia
,
Y.
,
Xu
,
R.-G.
,
Ren
,
X.
,
Ewen-Campen
,
B.
,
Rajakumar
,
R.
,
Zirin
,
J.
,
Yang-Zhou
,
D.
,
Zhu
,
R.
,
Wang
,
F.
,
Mao
,
D.
, et al. 
(
2018
).
Next-generation CRISPR/Cas9 transcriptional activation in Drosophila using flySAM
.
Proc. Natl. Acad. Sci. USA
115
,
4719
-
4724
.
Jinek
,
M.
,
Chylinski
,
K.
,
Fonfara
,
I.
,
Hauer
,
M.
,
Doudna
,
J. A.
and
Charpentier
,
E.
(
2012
).
A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity
.
Science
337
,
816
-
821
.
Jory
,
A.
,
Estella
,
C.
,
Giorgianni
,
M. W.
,
Slattery
,
M.
,
Laverty
,
T. R.
,
Rubin
,
G. M.
and
Mann
,
R. S.
(
2012
).
A survey of 6,300 genomic fragments for cis-regulatory activity in the imaginal discs of Drosophila melanogaster
.
Cell Rep.
2
,
1014
-
1024
.
Juven-Gershon
,
T.
,
Cheng
,
S.
and
Kadonaga
,
J. T.
(
2006
).
Rational design of a super core promoter that enhances gene expression
.
Nat. Methods
3
,
917
-
922
.
Kadonaga
,
J. T.
(
2012
).
Perspectives on the RNA polymerase II core promoter
.
Wiley Interdiscip. Rev. Dev. Biol.
1
,
40
-
51
.
Kantorovitz
,
M. R.
,
Kazemian
,
M.
,
Kinston
,
S.
,
Miranda-Saavedra
,
D.
,
Zhu
,
Q.
,
Robinson
,
G. E.
,
Göttgens
,
B.
,
Halfon
,
M. S.
and
Sinha
,
S.
(
2009
).
Motif-blind, genome-wide discovery of cis-regulatory modules in Drosophila and mouse
.
Dev. Cell
17
,
568
-
579
.
Kazemian
,
M.
and
Halfon
,
M. S.
(
2019
).
CRM discovery beyond model insects
.
Methods Mol. Biol.
1858
,
117
-
139
.
Kazemian
,
M.
,
Zhu
,
Q.
,
Halfon
,
M. S.
and
Sinha
,
S.
(
2011
).
Improved accuracy of supervised CRM discovery with interpolated Markov models and cross-species comparison
.
Nucleic Acids Res.
39
,
9463
-
9472
.
Kazemian
,
M.
,
Suryamohan
,
K.
,
Chen
,
J.-Y.
,
Zhang
,
Y.
,
Samee
,
M. A. H.
,
Halfon
,
M. S.
and
Sinha
,
S.
(
2014
).
Evidence for deep regulatory similarities in early developmental programs across highly diverged insects
.
Genome Biol. Evol.
6
,
2301
-
2320
.
Kharchenko
,
P. V.
,
Alekseyenko
,
A. A.
,
Schwartz
,
Y. B.
,
Minoda
,
A.
,
Riddle
,
N. C.
,
Ernst
,
J.
,
Sabo
,
P. J.
,
Larschan
,
E.
,
Gorchakov
,
A. A.
,
Gu
,
T.
, et al. 
(
2011
).
Comprehensive analysis of the chromatin landscape in Drosophila melanogaster
.
Nature
471
,
480
-
485
.
Kleftogiannis
,
D.
,
Kalnis
,
P.
and
Bajic
,
V. B.
(
2016
).
Progress and challenges in bioinformatics approaches for enhancer identification
.
Brief. Bioinform.
17
,
967
-
979
.
Klein
,
J. C.
,
Chen
,
W.
,
Gasperini
,
M.
and
Shendure
,
J.
(
2018
).
Identifying novel enhancer elements with CRISPR-based screens
.
ACS Chem. Biol.
13
,
326
-
332
.
Klemm
,
S. L.
,
Shipony
,
Z.
and
Greenleaf
,
W. J.
(
2019
).
Chromatin accessibility and the regulatory epigenome
.
Nat. Rev. Genet.
20
,
207
-
220
.
Kokoza
,
V. A.
and
Raikhel
,
A. S.
(
2011
).
Targeted gene expression in the transgenic Aedes aegypti using the binary Gal4-UAS system
.
Insect Biochem. Mol. Biol.
41
,
637
-
644
.
Kukalova-Peck
,
J.
(
1991
).
Fossil history and the evolution of hexapod structures
. In
The Insects of Australia: A Textbook for Students and Research Workers
(ed.
I. D.
Naumann
), pp.
141
-
179
.
Melbourne University Press
.
Kvon
,
E. Z.
,
Kazmar
,
T.
,
Stampfel
,
G.
,
Yáñez-Cuna
,
J. O.
,
Pagani
,
M.
,
Schernhuber
,
K.
,
Dickson
,
B. J.
and
Stark
,
A.
(
2014
).
Genome-scale functional characterization of Drosophila developmental enhancers in vivo
.
Nature
512
,
91
-
95
.
Lai
,
Y.-T.
,
Deem
,
K. D.
,
Borràs-Castells
,
F.
,
Sambrani
,
N.
,
Rudolf
,
H.
,
Suryamohan
,
K.
,
El-Sherif
,
E.
,
Halfon
,
M. S.
,
McKay
,
D. J.
and
Tomoyasu
,
Y.
(
2018
).
Enhancer identification and activity evaluation in the red flour beetle, Tribolium castaneum
.
Development
145
,
dev160663
.
Le
,
N. Q. K.
,
Yapp
,
E. K. Y.
,
Ho
,
Q.-T.
,
Nagasundaram
,
N.
,
Ou
,
Y.-Y.
and
Yeh
,
H.-Y.
(
2019
).
iEnhancer-5Step: Identifying enhancers using hidden information of DNA sequences via Chou's 5-step rule and word embedding
.
Anal. Biochem.
571
,
53
-
61
.
Levy
,
S. E.
and
Myers
,
R. M.
(
2016
).
Advancements in next-generation sequencing
.
Annu. Rev. Genomics Hum. Genet.
17
,
95
-
115
.
Lewis
,
J. J.
and
Reed
,
R. D.
(
2019
).
Genome-wide regulatory adaptation shapes population-level genomic landscapes in Heliconius
.
Mol. Biol. Evol.
36
,
159
-
173
.
Lewis
,
J. J.
,
van der Burg
,
K. R. L.
,
Mazo-Vargas
,
A.
and
Reed
,
R. D.
(
2016
).
ChIP-Seq-annotated Heliconius erato genome highlights patterns of cis-regulatory evolution in Lepidoptera
.
Cell Rep.
16
,
2855
-
2863
.
Li
,
L.
,
Zhu
,
Q.
,
He
,
X.
,
Sinha
,
S.
and
Halfon
,
M. S.
(
2007
).
Large-scale analysis of transcriptional cis-regulatory modules reveals both common features and distinct subclasses
.
Genome Biol.
8
,
R101
.
Li
,
Y.
,
Shi
,
W.
and
Wasserman
,
W. W.
(
2018
).
Genome-wide prediction of cis-regulatory regions using supervised deep learning methods
.
BMC Bioinformatics
19
,
202
.
Li
,
F.
,
Zhao
,
X.
,
Li
,
M.
,
He
,
K.
,
Huang
,
C.
,
Zhou
,
Y.
,
Li
,
Z.
and
Walters
,
J. R.
(
2019
).
Insect genomes: progress and challenges
.
Insect Mol. Biol.
28
,
739
-
758
.
Lieberman-Aiden
,
E.
,
van Berkum
,
N. L.
,
Williams
,
L.
,
Imakaev
,
M.
,
Ragoczy
,
T.
,
Telling
,
A.
,
Amit
,
I.
,
Lajoie
,
B. R.
,
Sabo
,
P. J.
,
Dorschner
,
M. O.
, et al. 
(
2009
).
Comprehensive mapping of long-range interactions reveals folding principles of the human genome
.
Science
326
,
289
-
293
.
Lim
,
L. W. K.
,
Chung
,
H. H.
,
Chong
,
Y. L.
and
Lee
,
N. K.
(
2018
).
A survey of recently emerged genome-wide computational enhancer predictor tools
.
Comput. Biol. Chem.
74
,
132
-
141
.
Lin
,
S.
,
Ewen-Campen
,
B.
,
Ni
,
X.
,
Housden
,
B. E.
and
Perrimon
,
N.
(
2015
).
In vivo transcriptional activation using CRISPR/Cas9 in Drosophila
.
Genetics
201
,
433
-
442
.
Liu
,
F.
,
Li
,
H.
,
Ren
,
C.
,
Bo
,
X.
and
Shu
,
W.
(
2016
).
PEDLA: predicting enhancers with a deep learning-based algorithmic framework
.
Sci. Rep.
6
,
28517
.
Liu
,
B.
,
Li
,
K.
,
Huang
,
D.-S.
and
Chou
,
K.-C.
(
2018
).
iEnhancer-EL: identifying enhancers and their strength with ensemble learning approach
.
Bioinformatics
34
,
3835
-
3842
.
Liu
,
M.
,
Rehman
,
S.
,
Tang
,
X.
,
Gu
,
K.
,
Fan
,
Q.
,
Chen
,
D.
and
Ma
,
W.
(
2019
).
Methodologies for improving HDR efficiency
.
Front. Genet.
9
,
691
.
Long
,
H. K.
,
Prescott
,
S. L.
and
Wysocka
,
J.
(
2016
).
Ever-changing landscapes: transcriptional enhancers in development and evolution
.
Cell
167
,
1170
-
1187
.
Lopes
,
R.
,
Korkmaz
,
G.
and
Agami
,
R.
(
2016
).
Applying CRISPR-Cas9 tools to identify and characterize transcriptional enhancers
.
Nat. Rev. Mol. Cell Biol.
17
,
597
-
604
.
Lorenzen
,
M. D.
,
Brown
,
S. J.
,
Denell
,
R. E.
and
Beeman
,
R. W.
(
2002
).
Transgene expression from the Tribolium castaneum Polyubiquitin promoter
.
Insect Mol. Biol.
11
,
399
-
407
.
Lorenzen
,
M. D.
,
Berghammer
,
A. J.
,
Brown
,
S. J.
,
Denell
,
R. E.
,
Klingler
,
M.
and
Beeman
,
R. W.
(
2003
).
piggyBac-mediated germline transformation in the beetle Tribolium castaneum
.
Insect Mol. Biol.
12
,
433
-
440
.
Lynd
,
A.
and
Lycett
,
G. J.
(
2012
).
Development of the bi-partite Gal4-UAS system in the African malaria mosquito, Anopheles gambiae
.
PLoS ONE
7
,
e31552
.
Ma
,
J.
and
Ptashne
,
M.
(
1987
).
Deletion analysis of GAL4 defines two transcriptional activating segments
.
Cell
48
,
847
-
853
.
Markstein
,
M.
and
Levine
,
M.
(
2002
).
Decoding cis-regulatory DNAs in the Drosophila genome
.
Curr. Opin. Genet. Dev.
12
,
601
-
606
.
Markstein
,
M.
,
Markstein
,
P.
,
Markstein
,
V.
and
Levine
,
M. S.
(
2002
).
Genome-wide analysis of clustered Dorsal binding sites identifies putative target genes in the Drosophila embryo
.
Proc. Natl Acad. Sci. USA
99
,
763
-
768
.
Martin
,
A.
,
Serano
,
J. M.
,
Jarvis
,
E.
,
Bruce
,
H. S.
,
Wang
,
J.
,
Ray
,
S.
,
Barker
,
C. A.
,
O'Connell
,
L. C.
and
Patel
,
N. H.
(
2016
).
CRISPR/Cas9 mutagenesis reveals versatile roles of Hox genes in crustacean limb specification and evolution
.
Curr. Biol.
26
,
14
-
26
.
Mayor
,
C.
,
Brudno
,
M.
,
Schwartz
,
J. R.
,
Poliakov
,
A.
,
Rubin
,
E. M.
,
Frazer
,
K. A.
,
Pachter
,
L. S.
and
Dubchak
,
I.
(
2000
).
VISTA: visualizing global DNA sequence alignments of arbitrary length
.
Bioinformatics
16
,
1046
-
1047
.
Mazo-Vargas
,
A.
,
Concha
,
C.
,
Livraghi
,
L.
,
Massardo
,
D.
,
Wallbank
,
R. W. R.
,
Zhang
,
L.
,
Papador
,
J. D.
,
Martinez-Najera
,
D.
,
Jiggins
,
C. D.
,
Kronforst
,
M. R.
, et al. 
(
2017
).
Macroevolutionary shifts of WntA function potentiate butterfly wing-pattern diversity
.
Proc. Natl. Acad. Sci.
114
,
10701
-
10706
.
McKay
,
D. J.
(
2019
).
Using Formaldehyde-Assisted Isolation of Regulatory Elements (FAIRE) to identify functional regulatory DNA in insect genomes
.
Methods Mol. Biol.
1858
,
89
-
97
.
McKay
,
D. J.
and
Lieb
,
J. D.
(
2013
).
A common set of DNA regulatory elements shapes Drosophila appendages
.
Dev. Cell
27
,
306
-
318
.
Meers
,
M. P.
,
Bryson
,
T. D.
,
Henikoff
,
J. G.
and
Henikoff
,
S.
(
2019
).
Improved CUT&RUN chromatin profiling tools
.
eLife
8
,
e46314
.
Meyer
,
C. A.
and
Liu
,
X. S.
(
2014
).
Identifying and mitigating bias in next-generation sequencing methods for chromatin biology
.
Nat Rev. Genet.
15
,
709
-
721
.
Min
,
X.
,
Zeng
,
W.
,
Chen
,
S.
,
Chen
,
N.
,
Chen
,
T.
and
Jiang
,
R.
(
2017
).
Predicting enhancers with deep convolutional neural networks
.
BMC Bioinformatics
18
,
478
.
Mishra
,
M.
,
Oke
,
A.
,
Lebel
,
C.
,
McDonald
,
E. C.
,
Plummer
,
Z.
,
Cook
,
T. A.
and
Zelhof
,
A. C.
(
2010
).
Pph13 and Orthodenticle define a dual regulatory pathway for photoreceptor cell morphogenesis and function
.
Development
137
,
2895
-
2904
.
Muerdter
,
F.
,
Boryń
,
Ł. M.
and
Arnold
,
C. D.
(
2015
).
STARR-seq - principles and applications
.
Genomics
106
,
145
-
150
.
Nakade
,
S.
,
Tsubota
,
T.
,
Sakane
,
Y.
,
Kume
,
S.
,
Sakamoto
,
N.
,
Obara
,
M.
,
Daimon
,
T.
,
Sezutsu
,
H.
,
Yamamoto
,
T.
,
Sakuma
,
T.
, et al. 
(
2014
).
Microhomology-mediated end-joining-dependent integration of donor DNA in cells and animals using TALENs and CRISPR/Cas9
.
Nat. Commun.
5
,
5560
.
Nakanishi
,
T.
,
Kato
,
Y.
,
Matsuura
,
T.
and
Watanabe
,
H.
(
2014
).
CRISPR/Cas-mediated targeted mutagenesis in Daphnia magna
.
PLoS ONE
9
,
e98363
.
Nègre
,
N.
,
Brown
,
C. D.
,
Ma
,
L.
,
Bristow
,
C. A.
,
Miller
,
S. W.
,
Wagner
,
U.
,
Kheradpour
,
P.
,
Eaton
,
M. L.
,
Loriaux
,
P.
,
Sealfon
,
R.
, et al. 
(
2011
).
A cis-regulatory map of the Drosophila genome
.
Nature
471
,
527
-
531
.
O'Brochta
,
D. A.
,
Pilitt
,
K. L.
,
Harrell
,
R. A.
,
Aluvihare
,
C.
and
Alford
,
R. T.
(
2012
).
Gal4-based enhancer-trapping in the malaria mosquito Anopheles stephensi
.
G3 (Bethesda).
2
,
1305
-
1315
.
Ohde
,
T.
,
Takehana
,
Y.
,
Shiotsuki
,
T.
and
Niimi
,
T.
(
2018
).
CRISPR/Cas9-based heritable targeted mutagenesis in Thermobia domestica: a genetic tool in an apterygote development model of wing evolution
.
Arthropod Struct. Dev.
47
,
362
-
369
.
Papatsenko
,
D.
,
Kislyuk
,
A.
,
Levine
,
M.
and
Dubchak
,
I.
(
2006
).
Conservation patterns in different functional sequence categories of divergent Drosophila species
.
Genomics
88
,
431
-
442
.
Park
,
P. J.
(
2009
).
ChIP–seq: advantages and challenges of a maturing technology
.
Nat. Rev. Genet.
10
,
669
-
680
.
Pennacchio
,
L. A.
,
Bickmore
,
W.
,
Dean
,
A.
,
Nobrega
,
M. A.
and
Bejerano
,
G.
(
2013
).
Enhancers: five essential questions
.
Nat. Rev. Genet.
14
,
288
-
295
.
Pérez-Zamorano
,
B.
,
Rosas-Madrigal
,
S.
,
Lozano
,
O. A. M.
,
Castillo Méndez
,
M.
and
Valverde-Garduño
,
V.
(
2017
).
Identification of cis-regulatory sequences reveals potential participation of lola and Deaf1 transcription factors in Anopheles gambiae innate immune response
.
PLoS ONE
12
,
e0186435
.
Perry
,
M. W.
,
Boettiger
,
A. N.
,
Bothma
,
J. P.
and
Levine
,
M.
(
2010
).
Shadow enhancers foster robustness of Drosophila gastrulation
.
Curr. Biol.
20
,
1562
-
1567
.
Pfeiffer
,
B. D.
,
Jenett
,
A.
,
Hammonds
,
A. S.
,
Ngo
,
T.-T. B.
,
Misra
,
S.
,
Murphy
,
C.
,
Scully
,
A.
,
Carlson
,
J. W.
,
Wan
,
K. H.
,
Laverty
,
T. R.
, et al. 
(
2008
).
Tools for neuroanatomy and neurogenetics in Drosophila
.
Proc. Natl Acad. Sci. USA
105
,
9715
-
9720
.
Pickar-Oliver
,
A.
and
Gersbach
,
C. A.
(
2019
).
The next generation of CRISPR–Cas technologies and applications
.
Nat. Rev. Mol. Cell Biol.
20
,
490
-
507
.
Prakash
,
A.
and
Monteiro
,
A.
(
2018
).
apterous A specifies dorsal wing patterns and sexual traits in butterflies
.
Proc. R. Soc. B
285
,
20172685
.
Prasad
,
N.
,
Tarikere
,
S.
,
Khanale
,
D.
,
Habib
,
F.
and
Shashidhara
,
L. S.
(
2016
).
A comparative genomic analysis of targets of Hox protein Ultrabithorax amongst distant insect species
.
Sci. Rep.
6
,
27885
.
Reardon
,
S.
(
2019
).
CRISPR gene-editing creates wave of exotic model organisms
.
Nature
568
,
441
-
442
.
Rebeiz
,
M.
and
Williams
,
T. M.
(
2017
).
Using Drosophila pigmentation traits to study the mechanisms of cis-regulatory evolution
.
Curr. Opin. insect Sci.
19
,
1
-
7
.
Richards
,
S.
,
Liu
,
Y.
,
Bettencourt
,
B. R.
,
Hradecky
,
P.
,
Letovsky
,
S.
,
Nielsen
,
R.
,
Thornton
,
K.
,
Hubisz
,
M. J.
,
Chen
,
R.
,
Meisel
,
R. P.
, et al. 
(
2005
).
Comparative genome sequencing of Drosophila pseudoobscura: chromosomal, gene, and cis-element evolution
.
Genome Res.
15
,
1
-
18
.
Rickels
,
R.
and
Shilatifard
,
A.
(
2018
).
Enhancer logic and mechanics in development and disease
.
Trends Cell Biol.
28
,
608
-
630
.
Rivera
,
J.
,
Keränen
,
S. V. E.
,
Gallo
,
S. M.
and
Halfon
,
M. S.
(
2019
).
REDfly: the transcriptional regulatory element database for Drosophila
.
Nucleic Acids Res.
47
,
D828
-
D834
.
Ron
,
G.
,
Globerson
,
Y.
,
Moran
,
D.
and
Kaplan
,
T.
(
2017
).
Promoter-enhancer interactions identified from Hi-C data using probabilistic models and hierarchical topological domains
.
Nat. Commun.
8
,
2237
.
Roy
,
S.
,
Ernst
,
J.
,
Kharchenko
,
P. V.
,
Kheradpour
,
P.
,
Negre
,
N.
,
Eaton
,
M. L.
,
Landolin
,
J. M.
,
Bristow
,
C. A.
,
Ma
,
L.
,
Lin
,
M. F.
, et al. 
(
2010
).
Identification of functional elements and regulatory circuits by Drosophila modENCODE
.
Science
330
,
1787
-
1797
.
Rylee
,
J. C.
,
Siniard
,
D. J.
,
Doucette
,
K.
,
Zentner
,
G. E.
and
Zelhof
,
A. C.
(
2018
).
Expanding the genetic toolkit of Tribolium castaneum
.
PLoS ONE
13
,
e0195977
.
Sarrazin
,
A. F.
,
Peel
,
A. D.
and
Averof
,
M.
(
2012
).
A segmentation clock with two-segment periodicity in insects
.
Science
336
,
338
-
341
.
Savic
,
N.
,
Ringnalda
,
F. C.
,
Lindsay
,
H.
,
Berk
,
C.
,
Bargsten
,
K.
,
Li
,
Y.
,
Neri
,
D.
,
Robinson
,
M. D.
,
Ciaudo
,
C.
,
Hall
,
J.
, et al. 
(
2018
).
Covalent linkage of the DNA repair template to the CRISPR-Cas9 nuclease enhances homology-directed repair
.
eLife
7
,
e33761
.
Schinko
,
J. B.
,
Weber
,
M.
,
Viktorinova
,
I.
,
Kiupakis
,
A.
,
Averof
,
M.
,
Klingler
,
M.
,
Wimmer
,
E. A.
and
Bucher
,
G.
(
2010
).
Functionality of the GAL4/UAS system in Tribolium requires the use of endogenous core promoters
.
BMC Dev. Biol.
10
,
53
.
Schinko
,
J. B.
,
Hillebrand
,
K.
and
Bucher
,
G.
(
2012
).
Heat shock-mediated misexpression of genes in the beetle Tribolium castaneum
.
Dev. Genes Evol.
222
,
287
-
298
.
Schmid-Burgk
,
J. L.
,
Höning
,
K.
,
Ebert
,
T. S.
and
Hornung
,
V.
(
2016
).
CRISPaint allows modular base-specific gene tagging using a ligase-4-dependent mechanism
.
Nat. Commun.
7
,
12338
.
Sethi
,
A.
,
Gu
,
M.
,
Gumusgoz
,
E.
,
Chan
,
L.
,
Yan
,
K.-K.
,
Rozowsky
,
J.
,
Barozzi
,
I.
,
Afzal
,
V.
,
Akiyama
,
J.
,
Plajzer-Frick
,
I.
, et al. 
(
2018
).
A cross-organism framework for supervised enhancer prediction with epigenetic pattern recognition and targeted validation
.
bioRxiv
,
385237
.
Sheng
,
G.
,
Thouvenot
,
E.
,
Schmucker
,
D.
,
Wilson
,
D. S.
and
Desplan
,
C.
(
1997
).
Direct regulation of rhodopsin 1 by Pax-6/eyeless in Drosophila: evidence for a conserved function in photoreceptors
.
Genes Dev.
11
,
1122
-
1131
.
Siebert
,
K. S.
,
Lorenzen
,
M. D.
,
Brown
,
S. J.
,
Park
,
Y.
and
Beeman
,
R. W.
(
2008
).
Tubulin superfamily genes in Tribolium castaneum and the use of a Tubulin promoter to drive transgene expression
.
Insect Biochem. Mol. Biol.
38
,
749
-
755
.
Skene
,
P. J.
and
Henikoff
,
S.
(
2017
).
An efficient targeted nuclease strategy for high-resolution mapping of DNA binding sites
.
eLife
6
,
e21856
.
Skene
,
P. J.
,
Henikoff
,
J. G.
and
Henikoff
,
S.
(
2018
).
Targeted in situ genome-wide profiling with high efficiency for low cell numbers
.
Nat. Protoc.
13
,
1006
-
1019
.
Small
,
S.
,
Arnosti
,
D. N.
and
Levine
,
M.
(
1993
).
Spacing ensures autonomous expression of different stripe enhancers in the even-skipped promoter
.
Development
119
,
762
-
772
.
Sosinsky
,
A.
,
Honig
,
B.
,
Mann
,
R. S.
and
Califano
,
A.
(
2007
).
Discovering transcriptional regulatory regions in Drosophila by a nonalignment method for phylogenetic footprinting
.
Proc. Natl Acad. Sci. USA
104
,
6305
-
6310
.
Stark
,
A.
,
Lin
,
M. F.
,
Kheradpour
,
P.
,
Pedersen
,
J. S.
,
Parts
,
L.
,
Carlson
,
J. W.
,
Crosby
,
M. A.
,
Rasmussen
,
M. D.
,
Roy
,
S.
,
Deoras
,
A. N.
, et al. 
(
2007
).
Discovery of functional elements in 12 Drosophila genomes using evolutionary signatures
.
Nature
450
,
219
-
232
.
Stern
,
D. L.
and
Frankel
,
N.
(
2013
).
The structure and evolution of cis-regulatory regions: the shavenbaby story
.
Philos. Trans. R. Soc. B
368
,
20130028
.
Strobl
,
F.
,
Anderl
,
A.
and
Stelzer
,
E. H. K.
(
2018
).
A universal vector concept for a direct genotyping of transgenic organisms and a systematic creation of homozygous lines
.
eLife
7
,
e31677
.
Suryamohan
,
K.
and
Halfon
,
M. S.
(
2015
).
Identifying transcriptional cis -regulatory modules in animal genomes
.
Wiley Interdiscip. Rev. Dev. Biol.
4
,
59
-
84
.
Suryamohan
,
K.
,
Hanson
,
C.
,
Andrews
,
E.
,
Sinha
,
S.
,
Scheel
,
M. D.
and
Halfon
,
M. S.
(
2016
).
Redeployment of a conserved gene regulatory network during Aedes aegypti development
.
Dev. Biol.
416
,
402
-
413
.
Swanson
,
C. I.
,
Evans
,
N. C.
and
Barolo
,
S.
(
2010
).
Structural rules and complex regulatory circuitry constrain expression of a Notch- and EGFR-regulated eye enhancer
.
Dev. Cell
18
,
359
-
370
.
Thakore
,
P. I.
,
D'Ippolito
,
A. M.
,
Song
,
L.
,
Safi
,
A.
,
Shivakumar
,
N. K.
,
Kabadi
,
A. M.
,
Reddy
,
T. E.
,
Crawford
,
G. E.
and
Gersbach
,
C. A.
(
2015
).
Highly specific epigenome editing by CRISPR-Cas9 repressors for silencing of distal regulatory elements
.
Nat. Methods
12
,
1143
-
1149
.
Thomas
,
J.-L.
,
Da Rocha
,
M.
,
Besse
,
A.
,
Mauchamp
,
B.
and
Chavancy
,
G.
(
2002
).
3xP3-EGFP marker facilitates screening for transgenic silkworm Bombyx mori L. from the embryonic stage onwards
.
Insect Biochem. Mol. Biol.
32
,
247
-
253
.
Thomas
,
G. W. C.
,
Dohmen
,
E.
,
Hughes
,
D. S. T.
,
Murali
,
S. C.
,
Poelchau
,
M.
,
Glastad
,
K.
,
Anstead
,
C. A.
,
Ayoub
,
N. A.
,
Batterham
,
P.
,
Bellair
,
M.
, et al. 
(
2018
).
The genomic basis of arthropod diversity
.
bioRxiv
,
382945
.
Tokusumi
,
T.
,
Tokusumi
,
Y.
,
Brahier
,
M. S.
,
Lam
,
V.
,
Stoller-Conrad
,
J. R.
,
Kroeger
,
P. T.
and
Schulz
,
R. A.
(
2017
).
Screening and analysis of Janelia FlyLight Project enhancer-Gal4 strains identifies multiple gene enhancers active during hematopoiesis in normal and Wasp-challenged Drosophila larvae
.
G3 (Bethesda)
7
,
437
-
448
.
Trauner
,
J.
,
Schinko
,
J.
,
Lorenzen
,
M. D.
,
Shippy
,
T. D.
,
Wimmer
,
E. A.
,
Beeman
,
R. W.
,
Klingler
,
M.
,
Bucher
,
G.
and
Brown
,
S. J.
(
2009
).
Large-scale insertional mutagenesis of a coleopteran stored grain pest, the red flour beetle Tribolium castaneum, identifies embryonic lethal mutations and enhancer traps
.
BMC Biol.
7
,
73
.
Trible
,
W.
,
Olivos-Cisneros
,
L.
,
McKenzie
,
S. K.
,
Saragosti
,
J.
,
Chang
,
N.-C.
,
Matthews
,
B. J.
,
Oxley
,
P. R.
and
Kronauer
,
D. J. C.
(
2017
).
orco mutagenesis causes loss of antennal lobe glomeruli and impaired social behavior in ants
.
Cell
170
,
727
-
735.e10
.
van der Burg
,
K. R. L.
,
Lewis
,
J. J.
,
Martin
,
A.
,
Nijhout
,
H. F.
,
Danko
,
C. G.
and
Reed
,
R. D.
(
2019
).
Contrasting roles of transcription factors spineless and EcR in the highly dynamic chromatin landscape of butterfly wing metamorphosis
.
Cell Rep.
27
,
1027
-
1038.e3
.
Viktorinová
,
I.
and
Wimmer
,
E. A.
(
2007
).
Comparative analysis of binary expression systems for directed gene expression in transgenic insects
.
Insect Biochem. Mol. Biol.
37
,
246
-
254
.
Visel
,
A.
,
Blow
,
M. J.
,
Li
,
Z.
,
Zhang
,
T.
,
Akiyama
,
J. A.
,
Holt
,
A.
,
Plajzer-Frick
,
I.
,
Shoukry
,
M.
,
Wright
,
C.
,
Chen
,
F.
, et al. 
(
2009
).
ChIP-seq accurately predicts tissue-specific activity of enhancers
.
Nature
457
,
854
-
858
.
Vo Ngoc
,
L.
,
Kassavetis
,
G. A.
and
Kadonaga
,
J. T.
(
2019
).
The RNA polymerase II core promoter in Drosophila
.
Genetics
212
,
13
-
24
.
Watanabe
,
T.
,
Noji
,
S.
and
Mito
,
T.
(
2017
).
Genome editing in the cricket, Gryllus bimaculatus
.
Methods Mol. Biol.
1630
,
219
-
233
.
Wei
,
W.
,
Xin
,
H.
,
Roy
,
B.
,
Dai
,
J.
,
Miao
,
Y.
and
Gao
,
G.
(
2014
).
Heritable genome editing with CRISPR/Cas9 in the silkworm, Bombyx mori
.
PLoS ONE
9
,
e101210
.
Wolff
,
C.
,
Schröder
,
R.
,
Schulz
,
C.
,
Tautz
,
D.
and
Klingler
,
M.
(
1998
).
Regulation of the Tribolium homologues of caudal and hunchback in Drosophila: evidence for maternal gradient systems in a short germ embryo
.
Development
125
,
3645
-
3654
.
Xu
,
X.-R. S.
,
Gantz
,
V. M.
,
Siomava
,
N.
and
Bier
,
E.
(
2017
).
CRISPR/Cas9 and active genetics-based trans-species replacement of the endogenous Drosophila kni-L2 CRM reveals unexpected complexity
.
eLife
6
,
e30281
.
Yan
,
H.
,
Opachaloemphan
,
C.
,
Mancini
,
G.
,
Yang
,
H.
,
Gallitto
,
M.
,
Mlejnek
,
J.
,
Leibholz
,
A.
,
Haight
,
K.
,
Ghaninia
,
M.
,
Huo
,
L.
, et al. 
(
2017
).
An engineered orco mutation produces aberrant social behavior and defective neural development in ants
.
Cell
170
,
736
-
747.e9
.
Yang
,
B.
,
Liu
,
F.
,
Ren
,
C.
,
Ouyang
,
Z.
,
Xie
,
Z.
,
Bo
,
X.
and
Shu
,
W.
(
2017
).
BiRen: predicting enhancers with a deep-learning-based model using the DNA sequence alone
.
Bioinformatics
33
,
1930
-
1936
.
Zabidi
,
M. A.
,
Arnold
,
C. D.
,
Schernhuber
,
K.
,
Pagani
,
M.
,
Rath
,
M.
,
Frank
,
O.
and
Stark
,
A.
(
2015
).
Enhancer–core-promoter specificity separates developmental and housekeeping gene regulation
.
Nature
518
,
556
-
559
.
Zhang
,
L.
and
Reed
,
R. D.
(
2016
).
Genome editing in butterflies reveals that spalt promotes and Distal-less represses eyespot colour patterns
.
Nat. Commun.
7
,
11769
.
Zhang
,
L.
,
Martin
,
A.
,
Perry
,
M. W.
,
van der Burg
,
K. R. L.
,
Matsuoka
,
Y.
,
Monteiro
,
A.
and
Reed
,
R. D.
(
2017a
).
Genetic basis of melanin pigmentation in butterfly wings
.
Genetics
205
,
1537
-
1550
.
Zhang
,
L.
,
Mazo-Vargas
,
A.
and
Reed
,
R. D.
(
2017b
).
Single master regulatory gene coordinates the evolution and development of butterfly color and iridescence
.
Proc. Natl Acad. Sci. USA
114
,
10707
-
10712
.
Zhang
,
Q.
,
Cheng
,
T.
,
Jin
,
S.
,
Guo
,
Y.
,
Wu
,
Y.
,
Liu
,
D.
,
Xu
,
X.
,
Sun
,
Y.
,
Li
,
Z.
,
He
,
H.
, et al. 
(
2017c
).
Genome-wide open chromatin regions and their effects on the regulation of silk protein genes in Bombyx mori
.
Sci. Rep.
7
,
12919
.
Zhang
,
Q.
,
Cheng
,
T.
,
Sun
,
Y.
,
Wang
,
Y.
,
Feng
,
T.
,
Li
,
X.
,
Liu
,
L.
,
Li
,
Z.
,
Liu
,
C.
,
Xia
,
Q.
, et al. 
(
2019
).
Synergism of open chromatin regions involved in regulating genes in Bombyx mori
.
Insect Biochem. Mol. Biol.
110
,
10
-
18
.
Zinzen
,
R. P.
,
Cande
,
J.
,
Ronshaugen
,
M.
,
Papatsenko
,
D.
and
Levine
,
M.
(
2006
).
Evolution of the ventral midline in insect embryos
.
Dev. Cell
11
,
895
-
902
.

Competing interests

The authors declare no competing or financial interests.