hormone-dependent control of developmental timing through...

15
Hormone-dependent control of developmental timing through regulation of chromatin accessibility Christopher M. Uyehara, 1,2,3,4,6 Spencer L. Nystrom, 1,2,3,4,6 Matthew J. Niederhuber, 1,2,3,4 Mary Leatham-Jensen, 1,2,4 Yiqin Ma, 5 Laura A. Buttitta, 5 and Daniel J. McKay 1,2,4 1 Department of Biology, 2 Department of Genetics, 3 Curriculum in Genetics and Molecular Biology, 4 Integrative Program for Biological and Genome Sciences, The University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, 27599, USA; 5 Department of Molecular, Cellular, and Developmental Biology, University of Michigan, Ann Arbor, Michigan 48109, USA Specification of tissue identity during development requires precise coordination of gene expression in both space and time. Spatially, master regulatory transcription factors are required to control tissue-specific gene expression programs. However, the mechanisms controlling how tissue-specific gene expression changes over time are less well understood. Here, we show that hormone-induced transcription factors control temporal gene expression by regu- lating the accessibility of DNA regulatory elements. Using the Drosophila wing, we demonstrate that temporal changes in gene expression are accompanied by genome-wide changes in chromatin accessibility at temporal-spe- cific enhancers. We also uncover a temporal cascade of transcription factors following a pulse of the steroid hormone ecdysone such that different times in wing development can be defined by distinct combinations of hormone-in- duced transcription factors. Finally, we show that the ecdysone-induced transcription factor E93 controls temporal identity by directly regulating chromatin accessibility across the genome. Notably, we found that E93 controls enhancer activity through three different modalities, including promoting accessibility of late-acting enhancers and decreasing accessibility of early-acting enhancers. Together, this work supports a model in which an extrinsic signal triggers an intrinsic transcription factor cascade that drives development forward in time through regulation of chromatin accessibility. [Keywords: temporal gene regulation; open chromatin; ecdysone; pioneer transcription factor; genomics] Supplemental material is available for this article. Received February 28, 2017; revised version accepted April 26, 2017. A defining feature of metazoan development is the organi- zation of cells into tissues. The physiological function of a given tissue is determined by the identity of its constitu- ent cells as well as their arrangement within the tissue. As a result, building tissues during development requires precise spatial control of gene expression over extended periods of time. Whereas many of the genes required for the development of different cell and tissue types have been identified, the mechanisms through which spatial information is coordinated with temporal information re- main incompletely understood. Spatially, a select and conserved group of transcription factors, sometimes termed mastertranscription factors, often specifies the distinct identities of different cell and tissue types (Mann and Carroll 2002; Mullen et al. 2011). Genetic studies from a range of organisms show that loss of function of a given master transcription factor can result in the loss of a given cell type or tissue. Con- versely, ectopic expression of a given master transcription factor can result in transformation of identities. Hence, master transcription factors are major determinants of cell fate. Consistent with their importance in develop- ment, the dysregulation of master transcription factors is associated with a range of diseases. Thus, understanding the mechanisms through which these factors function is an important goal in biomedical research. One proposed mechanism to explain the distinctive power of master transcription factors is that they control where other transcription factors bind in the genome by regulating chromatin accessibility (Fakhouri et al. 2010; Mullen et al. 2011; Pham et al. 2013). In vivo, DNA is wrapped around histone proteins to make nucleosomes, the basic unit of chromatin. Due to their tight as- sociation with DNA, nucleosomes act as barriers to 6 These authors contributed equally to this work. Corresponding author: [email protected] Article published online ahead of print. Article and publication date are online at http://www.genesdev.org/cgi/doi/10.1101/gad.298182.117. Free- ly available online through the Genes & Development Open Access option. © 2017 Uyehara et al. This article, published in Genes & Development, is available under a Creative Commons License (Attribution 4.0 Internation- al), as described at http://creativecommons.org/licenses/by/4.0/. 862 GENES & DEVELOPMENT 31:862875 Published by Cold Spring Harbor Laboratory Press; ISSN 0890-9369/17; www.genesdev.org Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.org Downloaded from

Upload: others

Post on 20-Oct-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

  • Hormone-dependent controlof developmental timing throughregulation of chromatin accessibilityChristopher M. Uyehara,1,2,3,4,6 Spencer L. Nystrom,1,2,3,4,6 Matthew J. Niederhuber,1,2,3,4

    Mary Leatham-Jensen,1,2,4 Yiqin Ma,5 Laura A. Buttitta,5 and Daniel J. McKay1,2,4

    1Department of Biology, 2Department of Genetics, 3Curriculum in Genetics and Molecular Biology, 4Integrative Program forBiological and Genome Sciences, The University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, 27599, USA;5Department of Molecular, Cellular, and Developmental Biology, University of Michigan, Ann Arbor, Michigan 48109, USA

    Specification of tissue identity during development requires precise coordination of gene expression in both spaceand time. Spatially, master regulatory transcription factors are required to control tissue-specific gene expressionprograms. However, themechanisms controlling how tissue-specific gene expression changes over time are less wellunderstood. Here, we show that hormone-induced transcription factors control temporal gene expression by regu-lating the accessibility of DNA regulatory elements. Using the Drosophila wing, we demonstrate that temporalchanges in gene expression are accompanied by genome-wide changes in chromatin accessibility at temporal-spe-cific enhancers.We also uncover a temporal cascade of transcription factors following a pulse of the steroid hormoneecdysone such that different times in wing development can be defined by distinct combinations of hormone-in-duced transcription factors. Finally, we show that the ecdysone-induced transcription factor E93 controls temporalidentity by directly regulating chromatin accessibility across the genome. Notably, we found that E93 controlsenhancer activity through three different modalities, including promoting accessibility of late-acting enhancers anddecreasing accessibility of early-acting enhancers. Together, this work supports amodel in which an extrinsic signaltriggers an intrinsic transcription factor cascade that drives development forward in time through regulation ofchromatin accessibility.

    [Keywords: temporal gene regulation; open chromatin; ecdysone; pioneer transcription factor; genomics]

    Supplemental material is available for this article.

    Received February 28, 2017; revised version accepted April 26, 2017.

    A defining feature ofmetazoan development is the organi-zation of cells into tissues. The physiological function of agiven tissue is determined by the identity of its constitu-ent cells as well as their arrangement within the tissue.As a result, building tissues during development requiresprecise spatial control of gene expression over extendedperiods of time. Whereas many of the genes required forthe development of different cell and tissue types havebeen identified, the mechanisms through which spatialinformation is coordinated with temporal information re-main incompletely understood.

    Spatially, a select and conserved group of transcriptionfactors, sometimes termed “master” transcription factors,often specifies the distinct identities of different cell andtissue types (Mann and Carroll 2002; Mullen et al.2011). Genetic studies from a range of organisms show

    that loss of function of a given master transcription factorcan result in the loss of a given cell type or tissue. Con-versely, ectopic expression of a givenmaster transcriptionfactor can result in transformation of identities. Hence,master transcription factors are major determinants ofcell fate. Consistent with their importance in develop-ment, the dysregulation of master transcription factorsis associatedwith a range of diseases. Thus, understandingthe mechanisms through which these factors function isan important goal in biomedical research.

    One proposed mechanism to explain the distinctivepower of master transcription factors is that they controlwhere other transcription factors bind in the genome byregulating chromatin accessibility (Fakhouri et al. 2010;Mullen et al. 2011; Pham et al. 2013). In vivo, DNA iswrapped around histone proteins to make nucleosomes,the basic unit of chromatin. Due to their tight as-sociation with DNA, nucleosomes act as barriers to6These authors contributed equally to this work.

    Corresponding author: [email protected] published online ahead of print. Article and publication date areonline at http://www.genesdev.org/cgi/doi/10.1101/gad.298182.117. Free-ly available online through the Genes & Development Open Accessoption.

    © 2017Uyehara et al. This article, published inGenes&Development, isavailable under a Creative Commons License (Attribution 4.0 Internation-al), as described at http://creativecommons.org/licenses/by/4.0/.

    862 GENES & DEVELOPMENT 31:862–875 Published by Cold Spring Harbor Laboratory Press; ISSN 0890-9369/17; www.genesdev.org

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    mailto:[email protected]:[email protected]:[email protected]://www.genesdev.org/cgi/doi/10.1101/gad.298182.117http://www.genesdev.org/cgi/doi/10.1101/gad.298182.117http://genesdev.cshlp.org/site/misc/terms.xhtmlhttp://creativecommons.org/licenses/by/4.0/http://creativecommons.org/licenses/by/4.0/http://genesdev.cshlp.org/site/misc/terms.xhtmlhttp://genesdev.cshlp.org/http://www.cshlpress.com

  • transcription factor binding. For a given transcription fac-tor to bind DNA, a nucleosome must be moved or evict-ed, creating a site of “open” or “accessible” chromatin.Several lines of evidence support an important role forchromatin accessibility in transcription factor targetingin the genome. Chief among these are the observationsthat only a small fraction of transcription factor DNA-binding motifs is occupied at a given point in time (Liet al. 2008) and that many sites of transcription factorbinding do not contain a recognizable DNA-binding mo-tif (Kvon et al. 2012). Thus, regulation of chromatin ac-cessibility plays a potentially pivotal role in controllingcell identity by determining where transcription factorscan bind in the genome and hence the sets of genesthat are expressed.If nucleosomes prevent transcription factors from ac-

    cessing DNA, then how do transcription factors come tooccupy their binding sites? Biochemical studies haveidentified a class of transcription factors termed “pioneer”factors that have the unique ability to bind nucleosomalDNA and subsequently enable binding by other tran-scription factors (Zaret and Mango 2016). The prototypepioneer factor is FoxA1, a master regulator of liver devel-opment (Lee et al. 2005). FoxA1 has also been shown toplay an important role in controlling targeting of the estro-gen and androgen receptors in breast and prostate cancercells, respectively (Jozwik and Carroll 2012). More recent-ly, themaster transcription factors of embryonic stem cellidentity—Oct4, Sox2, and Klf4—were shown to have pio-neering activity during induction of pluripotency in in-duced pluripotency stem (iPS) cells (Soufi et al. 2015).While pioneers have the potential to be pivotal regulatorsof gene expression programs, much remains to be learnedabout their function. For example, it is not clear why theyexhibit pioneering activity in only a subset of the cells inwhich they are expressed. Pioneers also may not be theonly factors that control chromatin accessibility. Othertranscription factors can work together to compete nucle-osomes off DNA, consistent with earlier in vitro work ontranscription factor binding to nucleosomal templates(Taylor et al. 1991).In addition to spatial control, gene expression patterns

    are also temporally regulated in development. For exam-ple, in a variety of animals, neural stem cells producedaughter cells with distinct identities at different timesof development to create the vast diversity of neuronsand glia found in the nervous system (Kohwi and Doe2013). InDrosophila embryos, an intrinsic cascade of tran-scription factor expression specifies the distinct temporalidentities of neural stem cell progeny (Isshiki et al. 2001).A similarmechanismusing a different transcription factorcascade diversifies neural identities in theDrosophila lar-val brain (Erclik et al. 2017). In contrast to stem cell line-ages, coordinating the timing of gene expression acrossfields of cells, such as a tissue, often involves the use of se-creted signals. For example, thyroid hormone controls theinitiation and progression of metamorphosis in frogs (Shi2013), whereas the sex hormones control the developmentof secondary sex traits during adolescence in mammals(Romeo 2003).

    In Drosophila and other insects, developmental timingis controlled by the steroid hormone ecdysone (Ashburner1990; Thummel 2002). Secreted by the prothoracic glandat stereotypical stages of development, ecdysone travelsthrough the hemolymph to reach target tissues, where itbinds to its receptor, the ecdysone receptor (EcR) (King-Jones andThummel 2005). Like other nuclear hormone re-ceptors, EcR is a transcription factor that differentially reg-ulates gene expression in the presence and absence ofligand. Studies initially performed in the larval salivarygland revealed that, upon binding ecdysone, EcR activatestranscription of a set of early genes, many of which aretranscription factors (Ashburner 1990). The early geneproducts thenworkwithEcR to activate a set of late genes,which encode the proteins that mediate the physiologicalresponse to hormone signaling (e.g., the glue proteinsmade by the salivary gland that adhere the pupa to a sub-strate during metamorphosis). Transcriptional profilingfrom a diverse collection of cell lines showed that the re-sponse to ecdysone is both widespread and highly celltype-specific (Stoiber et al. 2016). Mapping of hormone-re-sponsive enhancers in cultured cells recently revealed thattissue-specific responses to ecdysone are influenced bymotif content in DNA regulatory elements (Shlyuevaet al. 2014). Despite these efforts, the precise mechanismsthrough which ecdysone signaling controls temporal-spe-cific gene expression inDrosophila remain elusive.To ask how spatial and temporal information are inte-

    grated by regulatory DNA during specification of tissueidentities, we recently performed open chromatin profil-ing at two stages of Drosophila appendage development(McKay and Lieb 2013). In flies, the distinct identity ofeach appendage is determined by the expression of differ-entmaster transcription factors with different DNA-bind-ing domains. For example, leg identity is determined bythe homeodomain transcription factor Distalless and thezinc finger transcription factor Sp1 (Estella and Mann2010). In contrast, dorsal appendage identities, includingthe wing and haltere, are specified by vestigial and itsTEA domain-containing DNA-binding partner, scalloped(Halder et al. 1998). Despite the differences in mastertranscription factor identities between these tissues andcontrary to our expectations, we found that the open chro-matin profiles in wings, legs, and halteres are nearly thesame, with the exception of the master regulator locithemselves, which exhibit differential accessibility be-tween the appendages (McKayandLieb2013).The similar-ity in appendageopenchromatinprofiles indicates that themaster transcription factors are not the sole determinantsof chromatin accessibility, if they do so at all. This leavesthe question of which factors are responsible for control-ling chromatin accessibility in the appendages.One clue to the potential identity of these factors came

    from comparisons of open chromatin profiles between ap-pendages at different stages of development. We foundthat the different adult appendages shared very similaropen chromatin profiles, similar to our findings from anearlier stage of appendage development, the third instarimaginal discs. However, open chromatin profiles of theadult appendages were markedly different from those of

    Regulation of temporal gene expression

    GENES & DEVELOPMENT 863

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/http://www.cshlpress.com

  • the imaginal discs. This indicates that a coordinatechange in chromatin accessibility occurs during append-age development and also suggests that passage throughdevelopmental time has a greater impact than cell lineageon chromatin accessibility. Because the appendages arenot in physical contact with each other inside developingflies, we reasoned that a systemic signal, such as ecdy-sone, contributes to control of chromatin accessibility.

    Here, we examined themechanisms controlling tempo-ral gene regulation in Drosophila. Using a time course ofwing development that encompasses the transition be-tween larval and pupal stages, we used RNA sequencing(RNA-seq) to show that gene expression is temporally dy-namic as wings differentiate and undergo the complexmorphogenetic events that create the adult appendage.We then carried out open chromatin profiling and trans-genic reporter analysis to show that these changes ingene expression are accompanied by genome-wide chang-es in the accessibility of temporal-specific transcriptionalenhancers. Finally, we used ChIP-seq (chromatin immu-noprecipitation [ChIP] combined with high-throughputsequencing) and loss-of-function analyses to show thatthe ecdysone-induced transcription factor E93 is requiredto drive the normal sequence of chromatin accessibilitychanges. Importantly, E93 is required for not only increas-ing the accessibility of late-acting enhancers but alsodecreasing the accessibility of early-acting enhancers.

    Together, these findings demonstrate that E93 specifiestemporal identity by directly regulating accessibility oftemporal-specific transcriptional enhancers. More broad-ly, this work helps to explain how hormone signalingcan influence tissue-specific gene expression programsto drive development forward in time.

    Results

    Gene expression is temporally dynamic in pupal wings

    To examine the mechanisms underlying temporal regula-tion of gene expression, we focused on the early stages ofDrosophila pupal wing development. By the end of larvaldevelopment (Fig. 1A), the wing disc consists of ∼50,000cells, cell fates along the proximal–distal axis have beenpatterned, and precursors of adult structures such aswing veins and sensory organs are being specified (Cohen1993). During the next 2 d of pupal development (Fig. 1B,C), cell fates continue to bemore finely determined, whilethewing undergoes a final round of cell division (Guo et al.2016). This time is also characterized by dramatic mor-phological changes at both the tissue and cellular levels:Changes in cell shape drive eversion of the wing pouch,and changes in cell adhesion allow the apposed dorsaland ventral surfaces of the wing epithelium to form theupper and lower layers of the wing blade. Cytoskeletal

    Figure 1. Gene expression is temporally dynamic in pupalwings. (A–C ) Immunostaining ofwings from three developmental time points.DAPI (top row) and phospho-tyrosine (bottom row) label nuclei and cell membranes, respectively. (D) MA plots of RNA-seq signal in an-notated genes for consecutive time points. Differentially expressed genes are colored red. The top two GO terms for differentially ex-pressed genes are indicated with P-values in parentheses. Bars, 50 µm.

    Uyehara et al.

    864 GENES & DEVELOPMENT

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/http://www.cshlpress.com

  • changes also result in extrusion of the cell membrane toproduce a single cuticular hair (trichome) from eachwing blade epithelial cell (Fig. 1C, bottom row; Adleret al. 2000). Not surprisingly, these developmental chang-es are associated with widespread changes in gene expres-sion. To quantify these changes, we performed RNA-seqonwing discs dissected fromwandering third instar larvae(Fig. 1A, “L3”) and wings dissected from flies 24 h (Fig. 1B,“24 h”) and 44 h (Fig. 1C, “44 h”) after puparium forma-tion (Guo et al. 2016). Pairwise comparisons betweensuccessive time points revealed thousands of genesboth increasing and decreasing between each time point(edgeR false discovery rate [FDR]

  • dynamic open chromatin sites (78%–89%) is located distalto gene promoters (Supplemental Fig. S2D). Finally, plotsof the average FAIRE signal in temporally dynamic openchromatin indicate that many temporally dynamic openchromatin sites are used transiently in development. Forexample, sites closing between L3 and 24 h tend to stayclosed, and sites opening between 24 h and 44 h tend tobe closed at L3 (i.e., prior to 24 h) (Fig. 2C). Thus, the dy-namic gene expression exhibited by early pupal wings co-incides with dynamic changes in chromatin accessibility.

    Temporally dynamic open chromatin sites correspondto temporal-specific transcriptional enhancers

    Open chromatin sites are highly correlated with function-al DNA regulatory element activity (McKay and Lieb2013). Our findings above suggest that temporally dynam-ic open chromatin sitesmay be transiently used promoter-distal enhancers in pupal wings. To test this directly, wecloned open chromatin sites from three genes for use intransgenic reporter assays. These sites were chosenbecause they exhibit temporally dynamic accessibility,and the neighboring genes are required for proper wing de-velopment. Candidate enhancers were cloned into report-er constructs and integrated into the genome as singlecopies via ΦC31-mediated site-specific recombination.Altogether, we cloned six temporally dynamic open chro-matin sites. Each of these six sites corresponds to a tempo-rally regulated transcriptional enhancer. We discuss eachof them in turn.

    We first examined two candidate enhancers from thetnc locus. Asmentioned above, tnc encodes an extracellu-lar matrix protein involved in cell adhesion. Consistentwith a role for tnc in mediating adhesion between thedorsal and ventral surfaces of the wing pouch, RNAi-mediated knockdown of tnc results in defects in wingmorphology (Fraichard et al. 2010). We cloned two tempo-rally dynamic open chromatin sites located ∼40 kb up-stream of the tnc promoter. Our FAIRE-seq data showthat these sites increase in accessibility between the L3

    and 24-h time points and subsequently decrease in acces-sibility between 24 h and 44 h (Fig. 3A; Supplemental Figs.S3A, S4A). While neither reporter is active in L3 wingdiscs, there is activity in 24-h wings. We termed thesethe tncblade (blade) and tncwv (wing vein) enhancers.tncblade is active most strongly in the interveins betweenthe first and second and between the fourth and fifth lon-gitudinal veins and in cells near the proximal posteriormargin. It is also active at lower levels in the intervein be-tween the third and fourth longitudinal veins (Fig. 3A).tncwv is activemost strongly near the first, fifth, and sixthlongitudinal veins and at lower levels in the third longitu-dinal vein (Supplemental Fig. S4A). Thus, tncblade andtncwv are active in complementary domains of the 24-hpupal wing. Since tnc is expressed nearly ubiquitously atthis stage of wing development, these open chromatinsites likely correspond to bona fide transcriptional en-hancers that interpret different spatial inputs.

    We next examined two candidate enhancers from thenubbin (nub) locus, which encodes a transcription factorrequired for proximal-distal axis and vein developmentin wings (Cifuentes and Garcia-Bellido 1997). The clonedcandidate enhancers are located∼5 and 6.5 kb upstream ofthe nub promoter. Our FAIRE-seq data show that thesesites progressively increase in accessibility between theL3 and 44-h time points (Fig. 3B; Supplemental Figs.S3B, S4B). Immunofluorescence experiments show thata reporter carrying the distal site is not active in L3 wingdiscs. By 44 h, it shows strong activity in the L3 and L5wing veins and weaker activity in the L2 and L4 veins(Fig. 3B). We thus designated this as the nubvein enhancer.Consistent with the nubvein activity pattern, hypomor-phic nub alleles show defects in wing vein development(Cifuentes and Garcia-Bellido 1997). Immunostaining ofthe more proximal site, which we named nubmargin,shows reporter activity near the wing margin and theposterior cross-vein of 44-h wings (Supplemental Fig.S4B), again consistent with defects observed in nub hypo-morphic alleles (Cifuentes and Garcia-Bellido 1997).Thus, temporally dynamic open chromatin sites identify

    Figure 3. Temporally dynamic open chromatin cor-responds to temporal-specific enhancer activity. (Toprow) Browser shots of FAIRE-seq signal from the tnc(A), nub (B), and broad (C ) loci, with cloned regionsindicated by gray boxes, and depicted enhancers indi-cated by green boxes. (Middle and bottom rows)Immunostaining of reporter activity in wings at theindicated early and late time points. Enhancer activi-ty is in green. Bars, 50 µm. Additional time points areshown in Supplemental Figure S3.

    Uyehara et al.

    866 GENES & DEVELOPMENT

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/http://www.cshlpress.com

  • functionally relevant enhancers with temporal-specificactivity.Last, we examined two candidate enhancers from the

    broad (br) locus, which encodes a family of transcriptionfactors active in third instar and early prepupal tissues, in-cluding the wing (Kiss et al. 1988; Guo et al. 2016). Usingour FAIRE-seq data, we identified open chromatin sites atthe br locus that are accessible in L3 wing discs but sub-sequently decrease in accessibility by 24 h and 44 h (Fig.3C; Supplemental Figs. S3C, S4C). These candidate en-hancers were cloned upstream of GAL4 to allow for flex-ibility in the reporters used. Crossing these GAL4 driversto flies containing UAS-GFP revealed that both openchromatin sites are transcriptional enhancers active inL3 wing discs. The brdisc enhancer is located ∼40 kb up-stream of the br promoter. Similar to Br protein, brdisc

    is active nearly ubiquitously in wing imaginal disc epithe-lial cells, with higher levels along the anterior–posteriorand dorsal–ventral boundaries in the wing pouch (Fig.3C). We next sought to determine whether the decreasein accessibility of brdisc between L3 and 24 h coincideswith a decrease in enhancer activity. Since there are a lim-ited number of cell divisions in pupal wings, GFP signalcan persist even after an enhancer turns off. Therefore,we used a destabilized GFP reporter, reasoning that in-creased GFP degradation may make the reporter moresensitive to the enhancer’s activity state even if GAL4persists. Consistent with the timing of brdisc closing, wefound that it shuts off between L3 and 24 h (Fig. 3C).Thus, the timing of brdisc closing coincides with the tim-ing of it turning off. We identified a second br enhancer,which we termed brade (Supplemental Fig. S4C). This en-hancer is located ∼30 kb upstream of the br promoter andis active in L3 wing discs in a pattern similar to the ade-pithelial cells located in the notum of the wing. Likethe brdisc enhancer, there is no sign of brade reporter activ-ity in the wing blade by 24 h, consistent with expecta-tions, since the adepithelial cells remain in the notumto form the indirect flightmuscles. Notably, br is requiredfor proper differentiation of these cells into adult muscles(Lovato et al. 2005). Together, these findings supportthe premise that temporally dynamic open chromatinsites correspond to temporal-specific transcriptional en-hancers and that genes use different DNA regulatory ele-ments to control their expression at different stages ofdevelopment.

    A temporal cascade of ecdysone-induced transcriptionfactors is expressed in pupal wings

    The above findings suggest that temporal changes in geneexpression are driven by temporal changes in the accessi-bility of transcriptional enhancers. We next sought toidentify factors that could be involved in controlling theaccessibility of these enhancers. We reasoned that ecdy-sone signaling may be involved, since it controls develop-mental transitions in insects (Thummel 2001; King-Jonesand Thummel 2005), and our previous work suggestedthat an extrinsic signal may coordinate temporal changesin chromatin accessibility between the appendages

    (McKay and Lieb 2013). We performed RNA-seq at sixtime points in pupal wings (Guo et al. 2016). Consistentwith the Ashburner model of ecdysone signaling (Fig.4A; Ashburner 1990), we observed a clear temporal cas-cade of ecdysone-induced transcription factor expressionsuch that each time point in early pupal wing develop-ment can be defined by a distinct combination of thesetranscription factors (Fig. 4B; Supplemental Table S2).Moreover, the timing of each factor’s expression coincideswith the timing of its requirement inDrosophila develop-ment. For example, br is required for the transition fromlarval to prepupal stages (Kiss et al. 1988), and we foundthat it is expressed specifically at the L3 time point inwings. Likewise, ftz-f1 is required for the transition fromprepupal to pupal stages (Broadus et al. 1999), and wefound that it is expressed specifically at the 6-h time pointin wings. Finally, the transcription factor E93 is expressedat the 18-h and 24-h time points, when it is required forbract development in pupal legs (Mou et al. 2012). Thus,a temporal cascade of ecdysone-induced transcription fac-tors occurs in early pupal wings.If ecdysone-induced transcription factors control chro-

    matin accessibility, one may expect to find their DNA-

    Figure 4. A temporal cascade of ecdysone-induced transcriptionfactors in pupal wings. (A) Diagram of the Ashburner model of ec-dysone signaling. (B) Heat map of gene expression values for se-lected ecdysone-induced genes across six stages of wingdevelopment, plotted as a fraction of the maximum expressionvalue. Blue shows high expression, and gray shows low expres-sion. (C ) Heat maps of DNA-binding sitemotif enrichment in dy-namic FAIRE peaks for selected transcription factors. (D) DAPIstain of L3 wing discs (top) and bright-field images of 96- h wings(bottom) from wild-type (left) and E93 mutants (right). Bars: top,75 µm; bottom, 500 µm.

    Regulation of temporal gene expression

    GENES & DEVELOPMENT 867

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/http://www.cshlpress.com

  • binding motifs to be overrepresented in temporally dy-namic open chromatin sites. To ask this question, welooked for enrichment of known DNA-binding motifs(Zhu et al. 2011) for a set of ecdysone-induced transcrip-tion factors (Supplemental Table S3) in temporally dy-namic FAIRE peaks relative to temporally static FAIREpeaks. We observed significant enrichment (P < 0.05) formultiple motifs in FAIRE peaks that open or close be-tween successive time points (Fig. 4C). In contrast, wedid not find any enrichment in temporally dynamic peaksfor themotif of Scalloped (Sd), theDNA-binding partner ofthe wing master transcription factor Vestigial (Halderet al. 1998). We also looked for enrichment of the motiffor GAGA factor (GAF), a transcription factor often associ-ated with transcriptional enhancers and open chromatinsites (Fuda et al. 2015). We found that the GAF motifwas enriched in both dynamic and static FAIRE peaks,thus causing no relative motif enrichment and suggestingthat GAF is not responsible for the temporal dynamics.Together, these findings are consistent with a role for ec-dysone-induced transcription factors in regulating tempo-rally dynamic open chromatin sites.

    Themotif for the ecdysone-induced transcription factorE93 was strongly overrepresented in open chromatin sitesthat close between L3 and 24 h as well as those that openbetween 24 h and 44 h (Fig. 4C). We therefore sought todetermine whether E93 plays a role in wing development.E93 encodes a pipsqueak domain-containing transcriptionfactor that was first identified as an ecdysone targetrequired for autophagy of the larval salivary gland (Baeh-recke and Thummel 1995; Lee et al. 2000). More recently,E93was shown to act as a competence factor for temporal-specific gene regulation in the pupal leg (Mou et al. 2012).Our RNA-seq data show that E93 is transcribed at highlevels in pupal wings at the 18- and 24-h time points(Fig. 4B). To ask whether E93 is required for normalwing development, we compared the morphology ofwild-type and E93 mutant wings. At the L3 stage, wild-type and E93 mutant wings are indistinguishable (Fig.4D), consistent with expectations, since E93 is not ex-pressed at this time. At 24 h, when E93 is expressed athigh levels, E93mutantwings display defects in cell adhe-sion between the dorsal and ventral surfaces of the wingepithelium (data not shown). At 96 h, following the periodof E93 expression, E93 mutant wings are dramaticallysmaller than wild-type wings, with significant defects invein development (Fig. 4D). Thus, E93 is essential forproper wing development, and the appearance of defectsin E93mutants is commensuratewith the timing of its ex-pression in pupal wings.

    ChIP-seq reveals that E93 binds open chromatin sitesin pupal wings

    The above findings reveal that the temporal changes ingene expression that occur during pupal wing develop-ment coincide with temporal changes in the accessibilityof thousandsof open chromatin sites,manyofwhich couldbe transcriptional enhancers. We next sought to examinethe role of E93 in this process. As a first step,we performed

    ChIP-seq to identify sites in the genome to which E93binds. We used an E93 protein trap fly strain generatedby the Drosophila Gene Disruption Project (Venkenet al. 2011). In this strain, the endogenous E93 gene has atransposon inserted within an intron that is shared by allannotated E93 isoforms. The transposon carries an “artifi-cial exon” cassette with the coding sequence for GFP andother epitope tags in the same reading frame as E93,flanked by splice acceptor and donor sites. Upon tran-scription and translation, an E93 fusion protein is ex-pressed (referred to here as E93GFSTF) that can beimmunoprecipitatedwith antibodies toGFP. Importantly,the E93GFSTF chromosome complements a deletion en-compassing the E93 locus [Df(3R)93FX2], demonstratingthat the fusion protein is functional. Supporting this func-tionality, immunostaining of E93GFSTF flies shows clearfusion protein expression in 24-h and 44-h wings (Supple-mental Fig. S5).

    We performed ChIP-seq for E93 on dissected 24-h pupalwings (Fig. 5A). Peak calling with MACS2 identified 8477significantly bound sites genome-wide.De novomotif dis-covery analysis (Bailey 2011) identified a sequence en-riched in E93 ChIP peaks that is very similar to an E93motif derived from a bacterial one-hybrid screen for Dro-sophila transcription factors (Fig. 5B; Zhu et al. 2011), sup-porting the quality of the data. Overall, E93 bindingcorresponds well with 24-h open chromatin in pupalwings (Fig. 5C,D); 50% of E93 ChIP peaks are containedwithin the top 6225 24-h FAIRE peaks, and 96% of E93ChIP peaks are contained within all 24-h FAIRE peaks(Fig. 5C). While there is good correspondence betweenE93 binding and open chromatin, not all 24-h FAIRE peaksare bound by E93 (Supplemental Fig. S6A), demonstratingthat the E93 ChIP signal is specific and not simply an in-direct consequence of open chromatin. This is further sup-ported by differences in the distribution of E93 ChIP and24-h FAIRE peak locations across the genome: E93 bindspreferentially to promoter-distal sites in the genome,whereas FAIRE peaks overlap proximal promoter regionswith greater frequency (Supplemental Fig. S6B). We nextexamined the relationship between E93 binding and tem-porally dynamic open chromatin in pupal wings. Fifty-onepercent of FAIRE peaks that change in accessibility be-tween L3 and 24 h are directly bound by E93 at 24h. Likewise, 51% of FAIRE peaks that change in accessi-bility between 24 h and 44 h are directly bound by E93at 24 h (Fig. 5E). In contrast, only 14% of temporally dy-namic FAIRE peaks in early embryos (McKay and Lieb2013) are bound by E93 in 24-h pupal wings. Thus, E93directly binds a significant majority of open chromatinsites that change accessibility (opening or closing) be-tween L3 and 44 h.

    E93 binding is required for temporally dynamic openchromatin changes

    The high degree of overlap between E93 binding and tem-porally dynamic open chromatin sites suggests that E93may play a direct role in controlling chromatin accessibil-ity during pupal wing development. To test this

    Uyehara et al.

    868 GENES & DEVELOPMENT

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/http://www.cshlpress.com

  • hypothesis, we performed FAIRE-seq in E93mutantwingsat L3, 24 h, and 44 h (Fig. 6A). In L3 wings, we observedvery few changes in open chromatin between wild-typeand E93 mutants (Fig. 6B), consistent with expectations,since E93 is not yet expressed at this time. In contrast,we observed thousands of changes in open chromatin be-tween wild-type and E93 mutant wings at 24 h and 44h. For example, 1508 FAIRE peaks out of the top 7699peaks from each pair of data sets are more open in wild-type wings than in E93 mutants at 24 h (Fig. 6B), demon-strating that E93 is required for accessibility at these sites.Surprisingly, 659 FAIRE peaks are more open in E93 mu-tant wings than in wild type at 24 h, indicating that E93is not required to promote accessibility at these sites; in-stead, it is required for the opposite: promoting nucleo-some occupancy. Thus, loss of E93 results in not onlythe loss of accessibility at thousands of sites in the ge-nome but also the inappropriate presence of accessiblechromatin at hundreds of additional sites.We next asked whether sites that depend on E93 for

    proper chromatin accessibility correspond to temporallydynamic FAIRE peaks. Indeed, 70% of E93-dependentFAIRE peaks are temporally dynamic between L3 and 24h in wild-type wings (Fig. 6C). This includes 53% of sitesthat normally open between L3 and 24 h in wild-typewings but fail to open in E93 mutants and 27% of sitesthat normally close between L3 and 24 h in wild-type

    wings but fail to close in E93 mutants (Fig. 6D; Supple-mental Fig. S7A). In contrast, only 4% of sites that changein accessibility during embryogenesis overlap an E93-de-pendent FAIRE peak (Fisher’s exact test P-value < 2.2 ×10−16) (Supplemental Fig. S7C). FAIRE peaks that do notchange in accessibility between L3 and 24 h in wild-typewings exhibit no change in accessibility in E93 mutants(Fig. 6D, “static” peaks). We obtained similar results forthe 24- to 44-h interval. The lower number of E93-depen-dent FAIRE peaks that overlap a dynamic FAIRE peak atthis later time interval (Fig. 6C; Supplemental Fig. S7D)is possibly due to a persistent failure in chromatin acces-sibility from the earlier time point, such as sites that failto open in E93 mutants at 24 h and stay closed at 44 h(e.g., highlighted region in Fig. 6A). Similarly, these indi-rect effects likely explain the increase in the fraction ofstatic peaks that exhibit E93 dependence during thistime interval (Fig. 6D). Importantly, we found that 50%of FAIRE peaks that are dependent on E93 for accessibilityat 24 h are directly bound by E93 (Fig. 6E). Together, thesefindings demonstrate that E93 controls temporal progres-sion of development by directly and indirectly regulatingthe accessibility of thousands of sites in the genome. Inthe absence of E93, nearly half of the expected open chro-matin changes fail to occur. One consequence of this fail-ure is that E93 mutants exhibit a heterochronic openchromatin defect: Open chromatin profiles of E93mutant

    Figure 5. E93 binds temporally dynamic open chromatin. (A) Browser shot from the fringe locus showing FAIRE-seq and E93 ChIP-seqsignals (Z-score) from pupal wings. (B) Position weight matrices comparing the E93 motif discovered in ChIP peaks with the known E93motif. (C ) Cumulative distribution plot of E93 ChIP peak overlap with 24-h FAIRE peaks (red line) relative to randomly shuffled FAIREpeaks (gray line). (D) Heat maps plotting E93 ChIP-seq and FAIRE-seq signals (Z-score) in E93 ChIP peaks from 24-h pupal wings. (E)Stacked bar plots showing the fraction of temporally dynamic FAIRE peaks (opening and closing) that overlap an E93 ChIP peak. (∗) Over-lap P-value

  • wings at 24 h are as similar to those of wild-type wings atL3 as they are to those of wild-type wings at 24 h (Supple-mental Fig. S8).

    E93 controls temporal-specific enhancer activity throughthree distinct mechanisms

    The results described above suggest that the developmen-tal defects observed in E93 mutants are due to a failure tomake temporally required changes in the accessibility ofDNA regulatory elements genome-wide. To directly testthis hypothesis, we examined the consequences of E93loss-of-function on the activity of the temporally dynamictranscriptional enhancers that we identified above. ChIP-seq shows that the nubvein enhancer is directly bound byE93 at 24 h, and FAIRE-seq in wild-type wings showsthat nubvein progressively opens after the L3 stage (Fig.7). The timing of this accessibility coincides with increas-ing nubvein enhancer activity inwing veins (Fig. 4). FAIRE-seq from E93 mutant wings reveals that this enhancer isdependent on E93 for its accessibility: It fails to open at24 h and remains closed at 44 h in the absence of E93(Fig. 7). Using the GAL4-UAS system to drive an E93

    RNAi construct specifically in the posterior compartmentof the wing with En-GAL4, we observed a strong loss ofnubvein activity upon E93 knockdown specifically in theregions where the RNAi was expressed (Fig. 7). The en-hancer remains active in only a few cells in the proximalwing after E93 knockdown, and most RNAi-expressingcells show complete loss of GFP. Thus, the failure toopen the nubvein enhancer in E93 mutant flies correlateswith a failure to activate the enhancer in transgenic re-porter assays.

    We next examined the brdisc enhancer, which is openand active in L3 wings but closed and inactive in 24-hand 44-h wings (Fig. 3). ChIP-seq reveals that E93 isdirectly bound to this enhancer, and FAIRE-seq showsthat it remains persistently open at 24 h and 44 h in E93mutant wings (Fig. 7). Consistent with this persistent ac-cessibility, brdisc is expressed in E93 mutants at 24 h,when it would normally be off in wild-type wings (Fig.7). The brdisc pattern in 24-h E93 mutant wings is nearlyubiquitous, similar to its pattern earlier in L3 wings.Thus, E93 is required to close this enhancer after the L3stage, and failure to do so results in its aberrant expressionat later developmental stages.

    Figure 6. E93 binding is required for temporally dynamic open chromatin changes. (A) Browser shot showing FAIRE-seq signal fromwild-type and E93 mutant wings. E93 ChIP-seq signal from wild-type 24-h wings is shown in black. The nubvein and nubmargin enhancers areshown in green. (B) MA plots of FAIRE-seq signal in the top 7699 FAIRE peaks from each wild-type and E93mutant wing data set. Differ-entially accessible peaks are colored red. (C ) Stacked bar plots of the fraction of E93-dependent FAIRE peaks that overlap a temporallydynamic FAIRE peak. (D) Line plots of the average FAIRE-seq signal in FAIRE peaks that close, open, or remain unchanged between con-secutive time points. The percentage of FAIRE peaks in each category that are E93-dependent is shown. Solid lines showwild-type FAIRE-seq signal. Dashed lines show E93mutant FAIRE-seq signal. (E) Stacked bar plot showing the fraction of E93-dependent FAIRE peaks thatoverlap an E93 ChIP peak.

    Uyehara et al.

    870 GENES & DEVELOPMENT

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/lookup/suppl/doi:10.1101/gad.298182.117/-/DC1http://genesdev.cshlp.org/http://www.cshlpress.com

  • Finally, we examined the tncblade enhancer, which isopen and active in 24-h wings (Fig. 3). ChIP-seq showsthat E93 is directly bound to tncblade (Fig. 7). However, de-spite this binding, tncblade does not significantly change inaccessibility in E93 mutant wings, demonstrating thatE93 is not required for promoting the accessibility ofthis enhancer. Nevertheless, RNAi-mediated knockdownof E93 in the anterior compartment of the wing using Ci-GAL4 results in loss of tncblade activity in RNAi-express-ing cells. Thus, while the tncblade enhancer is not depen-dent on E93 for its accessibility, it is still dependent onE93 for its activity. Importantly, the mutual dependenceof the nubvein and tncblade enhancers on E93 for transcrip-tional activity, combined with the specific dependence ofnubvein on E93 for accessibility, suggests that distinct bio-chemical mechanisms underlie E93 function at theseenhancers.

    Discussion

    The mechanisms controlling transcription factor target-ing in the genome are incompletely understood, particu-larly in the context of animal development. Here, we

    show that the hormone-induced transcription factor E93plays a direct role in controlling temporal changes in chro-matin accessibility in the developing Drosophila wing.Together with our previous findings, this work supportsa model in which two axes of information regulate en-hancer activity in developing appendages: Temporal infor-mation is provided by hormone-induced transcriptionfactors that regulate accessibility of transcriptional en-hancers, and spatial information is provided by the ap-pendage master transcription factors that differentiallyregulate the activity of these enhancers.

    Transcription factor targeting and temporal generegulation

    The importance of master transcription factors in specify-ing spatial identity during development suggests that theymay control where other transcription factors bind in thegenome. One prediction of this model is that tissueswhose identities are determined by different master tran-scription factors would exhibit different genome-wideDNA-binding profiles. However, we found recently thatthe Drosophila appendages (wings, legs, and halteres),

    Figure 7. E93 controls temporal-specific enhancer activity through three distinct modalities. (A) Browser shots of FAIRE-seq and ChIP-seq signal fromwild-type andE93mutantwings at the indicated loci. (B,C ) Immunostaining of reporter activity for each indicated enhanc-er. (B) Reporter activity in control or wild-type wings. (C ) Reporter activity (green) in wings expressing E93 RNAi under control of Ci-GAL4 (tncblade) or En-GAL4 (nubvein) or in E93 mutant wings (brdisc). The dotted lines indicate the boundary between RNAi-expressingand RNAi-nonexpressing cells. Arrows indicate loss of reporter activity. Bars, 50 µm.

    Regulation of temporal gene expression

    GENES & DEVELOPMENT 871

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/http://www.cshlpress.com

  • which use different transcription factors to determinetheir identities, share nearly identical open chromatinprofiles. Moreover, these shared open chromatin profileschange coordinately over developmental time. There aretwo possible explanations for these findings. Either (1)different transcription factors produce the same openchromatin profiles in different appendages or (2) transcrip-tion factors shared by each appendage control open chro-matin profiles instead of the master transcription factorsof appendage identity. We favor the second model forseveral reasons. Since the appendage master transcriptionfactors possess different DNA-binding domains with dis-tinct DNA-binding specificities, it is unlikely for themto bind the same sites in the genome. Supporting thisexpectation, ChIP for Scalloped and Homothorax, twotranscription factors important for appendage identity,shows clear tissue-specific binding in both the wing andeye–antennal imaginal discs (Slattery et al. 2013). Wealso prefer the secondmodel because it provides a relative-ly straightforward mechanism for the observed temporalchanges in open chromatin: By changing the expressionof the shared temporal transcription factor over time,the open chromatin profiles that it controls would changeas well. In contrast, expression of appendage master tran-scription factors is relatively stable over time, making itunlikely for them to be sufficient for temporal changesin open chromatin.

    We propose that control of chromatin accessibility inthe appendages is mediated at least in part by transcrip-tion factors downstream fromecdysone signaling. Accord-ing to this model, a systemic pulse of ecdysone initiates atemporal cascade of hormone-induced transcription fac-tor expression in each of the appendages. We thus referto these as “temporal” transcription factors. Temporaltranscription factors can directly regulate the accessibilityof transcriptional enhancers by opening or closing them,thereby conferring temporal specificity to their activityand driving development forward in time. Master tran-scription factors then bind accessible enhancers depend-ing on their DNA-binding preferences (or other means ofbinding DNA) and differentially regulate the activity ofthese enhancers to control spatial patterns of gene expres-sion, thus shaping the unique identities of individualappendages.

    Our experiments with E93 provide direct support forthis model. In wild-type wings, thousands of changes inopen chromatin occur after the large pulse of ecdysonethat triggers the end of larval development. In E93 mu-tants, ∼40% of these open chromatin changes fail to oc-cur. Importantly, nearly three-quarters of sites thatdepend on E93 for accessibility correspond to temporallydynamic sites in wild-type wings. Thus, chromatin acces-sibility is not grossly defective across the genome; instead,defects occur specifically in sites that change in accessi-bility over time. This finding, combined with the largefraction of temporally dynamic sites that depend on E93for accessibility, indicates that E93 controls a genome-wide shift in the availability of temporal-specific tran-scriptional enhancers. Supporting this hypothesis, weshow that temporal-specific enhancers depend on E93

    for both accessibility and activity. Since we propose thatthe response to ecdysone is shared across the appendages,we predict that similar defects occur in appendages be-sides the wing. It remains to be seen whether other ecdy-sone-induced transcription factors besides E93 controlaccessibility of enhancers at different developmentaltimes. It also remains to be seen how the temporal tran-scription factors work with the appendage master tran-scription factors to control appendage-specific enhanceractivity.

    Mechanisms of temporal transcription factor function

    Our findings suggest that E93 controls temporal-specificgene expression through three different modalities thatpotentially rely on three distinct biochemical activities.The enrichment of E93 motifs and binding of E93 to tem-porally dynamic sites indicate that it contributes to thisregulation directly. We propose that these combined ac-tivities drive development forward in time by turning offearly-acting enhancers and simultaneously turning onlate-acting enhancers.

    First, as in the case of the tncblade enhancer, E93 appearsto function as a conventional activator. In the absence ofE93, tncblade fails to express at high levels, but the acces-sibility of the enhancer does not measurably change. Thissuggests that binding of E93 to tncblade is required to re-cruit an essential coactivator. Importantly, this findingdemonstrates that E93 is not solely a regulator of chroma-tin accessibility. E93 binds many open chromatin sites inthe genome without regulating their accessibility andthus may regulate the temporal-specific activity of manyother enhancers. In addition, since the tncblade enhanceropens between L3 and 24 h even in the absence of E93(Fig. 7A), theremust be other factors that control its acces-sibility, perhaps, for example, transcription factors in-duced by ecdysone earlier in the temporal cascade.

    Second, as in the case of the nubvein enhancer, E93 is re-quired to promote chromatin accessibility. In this capaci-ty, E93 may function as a pioneer transcription factor toopen previously inaccessible chromatin. Alternatively,E93 may combine with other transcription factors, suchas the wing master transcription factors, to compete nu-cleosomes off DNA. Testing the ability of E93 to bind nu-cleosomal DNA will help to discriminate between thesetwo alternatives. In either case, we propose that this func-tion of E93 is necessary to activate late-acting enhancersacross the genome. Since only half of E93-dependent en-hancers are directly bound by E93 at 24 h (Fig. 6E), it isalso possible that E93 regulates the expression of othertranscription factors that control chromatin accessibility.Alternatively, if E93 uses a “hit and run” mechanism toopen these enhancers, our ChIP time point may havebeen too late to capture E93 binding at these sites.

    Finally, as in the case of the brdisc enhancer, E93 is re-quired to decrease chromatin accessibility. We proposethat this function of E93 is necessary to inactivate early-acting enhancers across the genome. Current models ofgene regulation do not adequately explain how sites ofopen chromatin are rendered inaccessible, but the ability

    Uyehara et al.

    872 GENES & DEVELOPMENT

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/http://www.cshlpress.com

  • to turn off early-acting enhancers is clearly an importantrequirement in developmental gene regulation. It mayalso be an important contributor to diseases such as can-cer, which exhibits widespread changes in chromatin ac-cessibility relative to matched normal cells (Stergachiset al. 2013). Thus, this role of E93 may represent a newfunctional class of transcription factor (“reverse pioneer”)or conventional transcriptional repressor activity. Addi-tional work is required to decipher the underlying mech-anisms. Notably, recent work on the temporal dynamicsof iPS cell reprogramming suggest a similar role forOct4, Sox2, and Klf4 in closing open chromatin to inacti-vate somatic enhancers (Chronis et al. 2017).

    Materials and methods

    Drosophila culture and genetics

    Flies were grown at 25°C under standard culture conditions. Thegenotype of the wild-type strain was w1118/yw, hs-FLP. Late wan-dering third instar larvaewere used for the L3 stage. White prepu-pae were used as the 0-h time point for pupal staging. Thefollowing genotypes were also used: UAS-E93 RNAi (ViennaDrosophila Resource Center, no. 104390), E93 protein trap (Dro-sophilaGene Disruption Project, BloomingtonDrosophila StockCenter [BDSC], no. 43675),UAS-nls-GFP (chromosome 2; BDSC,no. 4775), UAS-nls-GFP (chromosome 3; BDSC, no. 4776), E934

    (gift of Craig Woodard), Df(3R)93Fx2 (gift of Eric Baehrecke), andUAS-destabilized-GFP (gift of Brian McCabe).

    Sample preparation for high-throughput sequencing

    A minimum of 40 wings were dissected from staged female fliesin 1× PBS and transferred to ice for subsequent processing. RNAwas prepared as described previously (Guo et al. 2016), and theKAPA stranded mRNA-seq kit was used for library construction.FAIRE-seq was performed as described previously (McKay andLieb 2013), and the Rubicon Thruplex DNA-seq kit was usedfor library construction. ChIP-seq was performed on 24-h ± 1-hmanually dissected wings (n = 717 Rep1; n = 1280 Rep2) frombothmale and female E93 protein trap flies.Wings were dissectedin 1× PBS and kept on ice. Batches of wings from 20 pupae werefixed in 4% paraformaldehyde, 50 mM HEPES, 100 mM NaCl,1 mMEDTA, and 0.5 mMEGTA for 20 min at room temperaturefollowed by quenchingwith 125mMglycine in 1× PBS and 0.01%Triton. Fixedwingswere dounce-homogenized in 10mMHEPES,10 mM EDTA, 0.5 mM EGTA, 0.25% Triton, and 1 mM PMSF.Nuclei were pelleted at 4500g for 20 min and resuspended in 10mM HEPES, 200 mM NaCl, 1 mM EDTA, 0.5 mM EGTA,0.01% Triton, and 1 mM PMSF. After nutating for 10 min at 4°C, nuclei were pelleted again and resuspended in 140 mMNaCl, 10 mM HEPES, 1 mM EDTA, 0.5 mM EGTA, 1 mMPMSF, and 0.1% SDS followed by sonication on ice with a Bran-son sonifier until the average chromatin fragment size was 200base pairs. The soluble chromatin fraction was used for ChIP.Briefly, extracts were precleared with protein A Dynabeads for 2h at 4°C, and cleared extracts were incubated with 5 µg of rabbitanti-GFP antibody (Abcam, ab290) overnight at 4°C. Bead pull-down was performed for 3 h the following day. Antibody–beadcomplexes were washed successively and then eluted with 1%SDS, 250 mM NaCl, 10 mM Tris, and 1 mM EDTA. Sampleswere treatedwithRNaseA and proteinaseK and heated overnightat 65°C to reverse cross-links, and purifiedDNAwas recovered byphenol-chloroform/ethanol precipitation. Rubicon Thruplex

    DNA-seq kit was used for library construction. All sampleswere sequenced on an Illumina HiSeq 2500 at the University ofNorth Carolina High-Throughput Sequencing Facility. Addition-al details are available on request.

    Sequencing data analysis

    Sequencing reads were aligned to the dm3 reference genome.RNA-seq analysis was performed as described previously (Guoet al. 2016). We defined differentially expressed genes as thosehaving a logCPM >2 in at least one sample and changing by atleast twofold between pairwise time points. GO analysis was per-formed using DAVID (Database for Annotation, Visualization,and Integrated Discovery) (Huang da et al. 2009) and REViGO (re-duce and visualize GO) (Supek et al. 2011). FAIRE-seq analysisand peak calling were performed as described previously (McKayand Lieb 2013; Schulz et al. 2015). FAIRE-seq and ChIP-seq datawere visualized using Integrative Genomics Viewer (Robinsonet al. 2011).Z-scoreswere calculated using themean and standarddeviation per chromosome arm. To focus on a high-confidenceset of peaks, we chose a MACS2 −log10-adjusted P-value of 40from the 44-h wild-type data set and selected an equivalent num-ber of peaks (n = 7699) from the remaining data sets. edgeR (Rob-inson et al. 2010) was used for differential peak calling, asdescribed previously (Schulz et al. 2015). Briefly, BedTools (Quin-lan and Hall 2010) was used to calculate read depth for each set ofpeaks. FAIRE peaks with an FDR

  • with existing GFP-marked GAL4 drivers. Each nub and tnc en-hancer was integrated into the attP2 site on chromosome 3. Inte-gration of each reporter into its respective attP sitewas confirmedby PCR. Immunostaining and confocal imaging were performedas described previously (McKay and Lieb 2013). The followingantibodies were used: rabbit anti-GFP (1:1000; Abcam, ab290),mouse anti-phospho-tyrosine (1:1000; Fisher Scientific, clone4G10), and rabbit anti-E93 (1:2500; this study). Polyclonal anti-bodies to E93 were raised in rabbits using amino acid sequences271–520 of the E93-PA isoform, which is present in all annotatedE93 isoforms. In some cases, 30-h wings were used in figure imag-es due to their ease of mounting relative to 24-h wings. In all cas-es, reporter analysis was also conducted at 24 h, and no significantdifferences in reporter pattern were observed between 24 h and30 h. All primer sequences and vectors are available on request.

    Acknowledgements

    Transgenic fly stocks were obtained from the ViennaDrosophilaResource Center (VDRC; http://www.vdrc.at). Stocks obtainedfrom the Bloomington Drosophila Stock Center (National Insti-tutes of Health P40OD018537) were used in this study. Thiswork was supported by start-up funds provided by The Uni-versity of North Carolina at Chapel Hill to D.J.M. C.M.U. wassupported in part by National Institutes of Health grantT32GM007092.

    References

    Adler PN, Liu J, Charlton J. 2000. Cell size and themorphogenesisof wing hairs in Drosophila. Genesis 28: 82–91.

    Akalin A, Franke V, Vlahovicek K,MasonCE, Schubeler D. 2015.Genomation: a toolkit to summarize, annotate and visualizegenomic intervals. Bioinformatics 31: 1127–1129.

    AshburnerM. 1990. Puffs, genes, and hormones revisited.Cell 61:1–3.

    Baehrecke EH, Thummel CS. 1995. The Drosophila E93 genefrom the 93F early puff displays stage- and tissue-specific reg-ulation by 20-hydroxyecdysone. Dev Biol 171: 85–97.

    Bailey TL. 2011. DREME: motif discovery in transcription factorChIP-seq data. Bioinformatics 27: 1653–1659.

    Bailey TL, Johnson J, Grant CE, Noble WS. 2015. The MEMEsuite. Nucleic Acids Res 43: W39–W49.

    Broadus J, McCabe JR, Endrizzi B, Thummel CS, Woodard CT.1999. The Drosophila β FTZ-F1 orphan nuclear receptor pro-vides competence for stage-specific responses to the steroidhormone ecdysone. Mol Cell 3: 143–149.

    Chronis C, Fiziev P, Papp B, Butz S, Bonora G, Sabri S, Ernst J,Plath K. 2017. Cooperative binding of transcription factors or-chestrates reprogramming. Cell 168: 442–459 e420.

    Cifuentes FJ, Garcia-BellidoA. 1997. Proximo-distal specificationin the wing disc of Drosophila by the nubbin gene. Proc NatlAcad Sci 94: 11405–11410.

    Cohen SM. 1993. Imaginal disc development. In The develop-ment ofDrosophila melanogaster (ed. Bate M, Martinez AriasA), pp. 747–841. Cold Spring Harbor Laboratory Press, ColdSpring Harbor, NY.

    Estella C, Mann RS. 2010. Non-redundant selector and growth-promoting functions of two sister genes, buttonhead andSp1, inDrosophila leg development. PLoS Genet 6: e1001001.

    Fakhouri TH, Stevenson J, Chisholm AD, Mango SE. 2010. Dy-namic chromatin organization during foregut development

    mediated by the organ selector gene PHA-4/FoxA. PLoSGenet6: e1001060.

    Fraichard S, Bouge AL, Kendall T, Chauvel I, Bouhin H, BunchTA. 2010. Tenectin is a novel αPS2βPS integrin ligand requiredfor wing morphogenesis and male genital looping in Droso-phila. Dev Biol 340: 504–517.

    FudaNJ, GuertinMJ, Sharma S, DankoCG,Martins AL, Siepel A,Lis JT. 2015. GAGA factormaintains nucleosome-free regionsand has a role inRNApolymerase II recruitment to promoters.PLoS Genet 11: e1005108.

    Giresi PG, Kim J, McDaniell RM, Iyer VR, Lieb JD. 2007. FAIRE(Formaldehyde-Assisted Isolation of Regulatory Elements)isolates active regulatory elements from human chromatin.Genome Res 17: 877–885.

    Guo Y, Flegel K, Kumar J, McKayDJ, Buttitta LA. 2016. Ecdysonesignaling induces two phases of cell cycle exit in Drosophilacells. Biol Open 5: 1648–1661.

    Halder G, Polaczyk P, Kraus ME, Hudson A, Kim J, Laughon A,Carroll S. 1998. The Vestigial and Scalloped proteins act to-gether to directly regulate wing-specific gene expression inDrosophila. Genes Dev 12: 3900–3909.

    Han C, Jan LY, Jan YN. 2011. Enhancer-driven membrane mark-ers for analysis of nonautonomous mechanisms reveal neu-ron-glia interactions in Drosophila. Proc Natl Acad Sci 108:9673–9678.

    Huang daW, ShermanBT, Lempicki RA. 2009. Systematic and in-tegrative analysis of large gene lists using DAVID bioinfor-matics resources. Nat Protoc 4: 44–57.

    Isshiki T, Pearson B, Holbrook S, Doe CQ. 2001.Drosophila neu-roblasts sequentially express transcription factors which spec-ify the temporal identity of their neuronal progeny. Cell 106:511–521.

    Jiang L, Pearson JC, Crews ST. 2010. Diversemodes ofDrosophilatracheal fusion cell transcriptional regulation.Mech Dev 127:265–280.

    Jozwik KM, Carroll JS. 2012. Pioneer factors in hormone-depen-dent cancers. Nat Rev Cancer 12: 381–385.

    King-Jones K, Thummel CS. 2005. Nuclear receptors—a perspec-tive from Drosophila. Nat Rev Genet 6: 311–323.

    Kiss I, Beaton AH, Tardiff J, FristromD, Fristrom JW. 1988. Inter-actions and developmental effects of mutations in the Broad-Complex ofDrosophilamelanogaster.Genetics 118: 247–259.

    Kohwi M, Doe CQ. 2013. Temporal fate specification and neuralprogenitor competence during development. Nat Rev Neuro-sci 14: 823–838.

    Kvon EZ, Stampfel G, Yanez-Cuna JO, Dickson BJ, Stark A. 2012.HOT regions function as patterned developmental enhancersand have a distinct cis-regulatory signature. Genes Dev 26:908–913.

    LawrenceM,GentlemanR,CareyV. 2009. rtracklayer: anR pack-age for interfacing with genome browsers. Bioinformatics 25:1841–1842.

    Lawrence M, Huber W, Pages H, Aboyoun P, Carlson M, Gentle-man R, Morgan MT, Carey VJ. 2013. Software for computingand annotating genomic ranges. PLoS Comput Biol 9:e1003118.

    LeeCY,Wendel DP, Reid P, LamG, Thummel CS, Baehrecke EH.2000. E93 directs steroid-triggered programmed cell death inDrosophila. Mol Cell 6: 433–443.

    Lee CS, Friedman JR, Fulmer JT, Kaestner KH. 2005. The initia-tion of liver development is dependent on Foxa transcriptionfactors. Nature 435: 944–947.

    Li XY, MacArthur S, Bourgon R, Nix D, Pollard DA, Iyer VN,Hechmer A, Simirenko L, Stapleton M, Luengo HendriksCL, et al. 2008. Transcription factors bind thousands of active

    Uyehara et al.

    874 GENES & DEVELOPMENT

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://www.vdrc.athttp://www.vdrc.athttp://www.vdrc.athttp://www.vdrc.athttp://www.vdrc.athttp://genesdev.cshlp.org/http://www.cshlpress.com

  • and inactive regions in the Drosophila blastoderm. PLoS Biol6: e27.

    Lovato TL, Benjamin AR, Cripps RM. 2005. Transcription ofmyocyte enhancer factor-2 in adult Drosophila myoblasts isinduced by the steroid hormone ecdysone. Dev Biol 288:612–621.

    Mann RS, Carroll SB. 2002. Molecular mechanisms of selectorgene function and evolution. Curr Opin Genet Dev 12:592–600.

    McKay DJ, Lieb JD. 2013. A common set of DNA regulatory ele-ments shapes Drosophila appendages. Dev Cell 27: 306–318.

    McLeay RC, Bailey TL. 2010. Motif enrichment analysis: a uni-fied framework and an evaluation on ChIP data. BMC Bioin-formatics 11: 165.

    Mou X, Duncan DM, Baehrecke EH, Duncan I. 2012. Control oftarget gene specificity during metamorphosis by the steroidresponse gene E93. Proc Natl Acad Sci 109: 2949–2954.

    Mullen AC, Orlando DA, Newman JJ, Loven J, Kumar RM, Bilo-deau S, Reddy J, Guenther MG, DeKoter RP, Young RA. 2011.Master transcription factors determine cell-type-specific re-sponses to TGF-β signaling. Cell 147: 565–576.

    PhamTH,Minderjahn J, Schmidl C, Hoffmeister H, SchmidhoferS, ChenW, LangstG, BennerC, RehliM. 2013.Mechanisms ofin vivo binding site selection of the hematopoietic mastertranscription factor PU.1. Nucleic Acids Res 41: 6391–6402.

    Quinlan AR, Hall IM. 2010. BedTools: a flexible suite of utilitiesfor comparing genomic features. Bioinformatics 26: 841–842.

    Ramirez F, Dundar F, Diehl S, Gruning BA,Manke T. 2014. deep-Tools: a flexible platform for exploring deep-sequencing data.Nucleic Acids Res 42: W187–W191.

    Robinson MD, McCarthy DJ, Smyth GK. 2010. edgeR: a Biocon-ductor package for differential expression analysis of digitalgene expression data. Bioinformatics 26: 139–140.

    Robinson JT, Thorvaldsdottir H, Winckler W, GuttmanM, Land-er ES, Getz G, Mesirov JP. 2011. Integrative genomics viewer.Nat Biotechnol 29: 24–26.

    Romeo RD. 2003. Puberty: a period of both organizational andactivational effects of steroid hormones on neurobehaviouraldevelopment. J Neuroendocrinol 15: 1185–1192.

    Schulz KN, Bondra ER, Moshe A, Villalta JE, Lieb JD, Kaplan T,McKay DJ, Harrison MM. 2015. Zelda is differentially re-quired for chromatin accessibility, transcription factor bind-ing, and gene expression in the early Drosophila embryo.Genome Res 25: 1715–1726.

    Shi YB. 2013. Unliganded thyroid hormone receptor regulatesmetamorphic timing via the recruitment of histone deacety-lase complexes. Curr Top Dev Biol 105: 275–297.

    ShlyuevaD, Stelzer C, GerlachD, Yanez-Cuna JO, RathM, BorynLM, Arnold CD, Stark A. 2014. Hormone-responsive enhanc-er-activity maps reveal predictive motifs, indirect repression,and targeting of closed chromatin. Mol Cell 54: 180–192.

    Slattery M, Voutev R, Ma L, Negre N, White KP, Mann RS. 2013.Divergent transcriptional regulatory logic at the intersectionof tissue growth and developmental patterning. PLoS Genet9: e1003753.

    Sobala LF, Adler PN. 2016. The gene expression program for theformation of wing cuticle in Drosophila. PLoS Genet 12:e1006100.

    Soufi A, GarciaMF, Jaroszewicz A, OsmanN, Pellegrini M, ZaretKS. 2015. Pioneer transcription factors target partial DNAmotifs on nucleosomes to initiate reprogramming. Cell 161:555–568.

    Stergachis AB, Neph S, Reynolds A, Humbert R, Miller B, PaigeSL, Vernot B, Cheng JB, Thurman RE, Sandstrom R, et al.2013. Developmental fate and cellular maturity encoded inhuman regulatory DNA landscapes. Cell 154: 888–903.

    Stoiber M, Celniker S, Cherbas L, Brown B, Cherbas P. 2016.Diverse hormone response networks in 41 independent Dro-sophila cell lines. G3 (Bethesda) 6: 683–694.

    Supek F, Bosnjak M, Skunca N, Smuc T. 2011. REViGO summa-rizes and visualizes long lists of gene ontology terms. PLoSOne 6: e21800.

    Taylor IC,Workman JL, Schuetz TJ, Kingston RE. 1991. Facilitat-ed binding ofGAL4 andheat shock factor to nucleosomal tem-plates: differential function of DNA-binding domains. GenesDev 5: 1285–1298.

    Thummel CS. 2001. Molecular mechanisms of developmentaltiming in C. elegans and Drosophila. Dev Cell 1: 453–465.

    Thummel CS. 2002. Ecdysone-regulated puff genes 2000. InsectBiochem Mol Biol 32: 113–120.

    Venken KJ, Schulze KL, Haelterman NA, Pan H, He Y, Evans-Holm M, Carlson JW, Levis RW, Spradling AC, Hoskins RA,et al. 2011. MiMIC: a highly versatile transposon insertion re-source for engineering Drosophila melanogaster genes. NatMethods 8: 737–743.

    Zaret KS,Mango SE. 2016. Pioneer transcription factors, chroma-tin dynamics, and cell fate control. Curr Opin Genet Dev 37:76–81.

    ZhuLJ,ChristensenRG,KazemianM,HullCJ, EnuamehMS,Bas-ciottaMD, Brasefield JA, ZhuC, Asriyan Y, LapointeDS, et al.2011. FlyFactorSurvey: a database ofDrosophila transcriptionfactor binding specificities determinedusing the bacterial one-hybrid system.Nucleic Acids Res 39: D111–D117.

    Regulation of temporal gene expression

    GENES & DEVELOPMENT 875

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/http://www.cshlpress.com

  • 10.1101/gad.298182.117Access the most recent version at doi: 2017 31: 862-875 originally published online May 23, 2017Genes Dev.

    Christopher M. Uyehara, Spencer L. Nystrom, Matthew J. Niederhuber, et al. regulation of chromatin accessibilityHormone-dependent control of developmental timing through

    Material

    Supplemental

    http://genesdev.cshlp.org/content/suppl/2017/05/23/gad.298182.117.DC1

    References

    http://genesdev.cshlp.org/content/31/9/862.full.html#related-urls

    Articles cited in:

    http://genesdev.cshlp.org/content/31/9/862.full.html#ref-list-1This article cites 57 articles, 10 of which can be accessed free at:

    Related Content

    Genes Dev. May 1, 2017 31: 847-848Sophia A. Praggastis and Carl S. ThummelRight time, right place: the temporal regulation of developmental gene expression

    Open Access

    Open Access option.Genes & DevelopmentFreely available online through the

    License

    Commons Creative

    .http://creativecommons.org/licenses/by/4.0/License (Attribution 4.0 International), as described at

    , is available under a Creative CommonsGenes & DevelopmentThis article, published in

    ServiceEmail Alerting

    click here.right corner of the article orReceive free email alerts when new articles cite this article - sign up in the box at the top

    http://genesdev.cshlp.org/subscriptionsgo to: Genes & Development To subscribe to

    © 2017 Uyehara et al.; Published by Cold Spring Harbor Laboratory Press

    Cold Spring Harbor Laboratory Press on July 27, 2017 - Published by genesdev.cshlp.orgDownloaded from

    http://genesdev.cshlp.org/lookup/doi/10.1101/gad.298182.117http://genesdev.cshlp.org/content/suppl/2017/05/23/gad.298182.117.DC1http://genesdev.cshlp.org/content/31/9/862.full.html#ref-list-1http://genesdev.cshlp.org/content/31/9/862.full.html#related-urlshttp://genesdev.cshlp.org/content/genesdev/31/9/847.full.htmlhttp://creativecommons.org/licenses/by/4.0/http://genesdev.cshlp.org/cgi/alerts/ctalert?alertType=citedby&addAlert=cited_by&saveAlert=no&cited_by_criteria_resid=genesdev;31/9/862&return_type=article&return_url=http://genesdev.cshlp.org/content/31/9/862.full.pdfhttp://genesdev.cshlp.org/cgi/adclick/?ad=41173&adclick=true&url=https%3A%2F%2Fwww.exiqon.com%2Fbiofluids-guide%3Futm_source%3DCSHL%26utm_medium%3Dbanner%26utm_content%3DGenDev-2017-07%26utm_campaign%3DBiofluidshttp://genesdev.cshlp.org/subscriptionshttp://genesdev.cshlp.org/http://www.cshlpress.com