11
RESEARCH ARTICLE Hobbit regulates intracellular trafficking to drive insulin-dependent growth during Drosophila development Sarah D. Neuman 1,2 and Arash Bashirullah 1,2, * ABSTRACT All animals must coordinate growth rate and timing of maturation to reach the appropriate final size. Here, we describe hobbit, a novel and conserved gene identified in a forward genetic screen for Drosophila animals with small body size. hobbit is highly conserved throughout eukaryotes, but its function remains unknown. We demonstrate that hobbit mutant animals have systemic growth defects because they fail to secrete insulin. Other regulated secretion events also fail in hobbit mutant animals, including mucin-like glueprotein secretion from the larval salivary glands. hobbit mutant salivary glands produce glue-containing secretory granules that are reduced in size. Importantly, secretory granules in hobbit mutant cells lack essential membrane fusion machinery required for exocytosis, including Synaptotagmin 1 and the SNARE SNAP-24. These membrane fusion proteins instead accumulate inside enlarged late endosomes. Surprisingly, however, the Hobbit protein localizes to the endoplasmic reticulum. Our results suggest that Hobbit regulates a novel step in intracellular trafficking of membrane fusion proteins. Our studies also suggest that genetic control of body size, as a measure of insulin secretion, is a sensitive functional readout of the secretory machinery. KEY WORDS: Body size, Growth, Insulin secretion, Regulated exocytosis, Intracellular trafficking, Drosophila INTRODUCTION Body size is exquisitely regulated during metazoan development. In Drosophila, an explosive period of growth occurs during larval development; however, this systemic growth ceases and final body size is specified at the onset of metamorphosis (Nijhout et al., 2014). One of the crucial regulators of body size in Drosophila is the insulin signaling pathway. In this pathway, Drosophila insulin-like peptides (Dilps) are secreted from the insulin-producing cells (IPCs), which are neuroendocrine cells located in the larval central nervous system. The IPCs function in a manner that is analogous to mammalian pancreatic β-cells (Geminard et al., 2006). Dilps then circulate systemically and bind to the insulin receptor (InR) on the membrane of target cells, triggering a downstream signaling cascade that regulates both metabolism and cellular growth (Brogiolo et al., 2001; Rulifson et al., 2002). Disruption of the insulin signaling pathway has a profound effect on Drosophila body size. Ablation of the IPCs (Rulifson et al., 2002), genetic deletion of Dilps (Grönke et al., 2010), or mutation of targets downstream of InR (Böhni et al., 1999; Brogiolo et al., 2001; Murillo-Maldonado et al., 2011) all result in a dramatic reduction in body size. However, despite the importance of the insulin signaling pathway for systemic growth in flies and for disease in humans, the factors regulating insulin secretion remain largely unknown. Insulin secretion in both flies and mammals occurs via regulated exocytosis. This crucial cellular process begins with the biogenesis of cargo protein-containing secretory granules. Cargo proteins targeted to the regulated exocytosis pathway, such as insulin, are translated by ribosomes bound to the rough endoplasmic reticulum (RER) and translocated into the ER lumen co-translationally. These cargo proteins are then transported to ER exit sites (ERES), specialized ER subdomains characterized by a coating of coat protein complex II (COPII), and packaged into COPII-coated vesicles (Brandizzi and Barlowe, 2013). COPII-coated vesicles mediate trafficking of cargo proteins from the ER to the Golgi cisternae; in the Golgi, cargo proteins undergo additional processing and post-translational modification. Cargo proteins are then selectively packaged into secretory granules, specialized organelles that bud from the trans-Golgi Network (TGN) as immature secretory granules (Kögel and Gerdes, 2010). The immature granules undergo a maturation process to refine their content and membrane composition, rendering them competent for exocytosis (Kögel and Gerdes, 2010). Mature secretory granules traffic to the plasma membrane, where they undergo tethering, docking and priming steps, attaching them to the membrane so that cargo proteins can be quickly released to the extracellular environment in a stimulus-dependent manner (Burgoyne and Morgan, 2003). The membrane fusion events leading to extracellular release are mediated by membrane fusion proteins, such as synaptotagmins and soluble N-ethylmaleimide-sensitive factor (NSF) activating protein receptors (SNAREs) (Burgoyne and Morgan, 2003; Hong, 2005). To drive membrane fusion during regulated exocytosis, these proteins must be properly trafficked to secretory granule membranes. Like cargo proteins, secretory granule membrane proteins must first enter the ER. Synaptotagmins contain a single- pass transmembrane domain at their N terminus (Südhof, 2002) that acts as a signal, directing the ribosome to the RER and allowing co-translational translocation of the nascent protein into the ER (Hegde and Keenan, 2011). In contrast, most SNAREs contain a single-pass transmembrane domain at their C terminus and are therefore translated by free ribosomes in the cytosol. SNAREs must then be post-translationally translocated into the ER, a process that requires chaperones to prevent the nascent SNARE from complexing with SNAREs on other organelles and to shelter the hydrophobic transmembrane domain, preventing aggregation (Hegde and Keenan, 2011; Kutay et al., 1995). Once in the ER, secretory granule membrane proteins are thought to follow a similar path to that of cargo proteins: they leave the ER via ERESs, traffic through the Golgi cisternae, and are eventually loaded onto Received 9 November 2017; Accepted 8 May 2018 1 Division of Pharmaceutical Sciences, University of Wisconsin-Madison, Madison, WI 53705-2222, USA. 2 Cellular and Molecular Biology Graduate Program, University of Wisconsin-Madison, Madison, WI 53706, USA. *Author for correspondence ([email protected]) S.D.N., 0000-0001-5731-6426; A.B., 0000-0001-7150-1775 1 © 2018. Published by The Company of Biologists Ltd | Development (2018) 145, dev161356. doi:10.1242/dev.161356 DEVELOPMENT

Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

RESEARCH ARTICLE

Hobbit regulates intracellular trafficking to drive insulin-dependentgrowth during Drosophila developmentSarah D. Neuman1,2 and Arash Bashirullah1,2,*

ABSTRACTAll animals must coordinate growth rate and timing of maturation toreach the appropriate final size. Here, we describe hobbit, a novel andconserved gene identified in a forward genetic screen for Drosophilaanimals with small body size. hobbit is highly conserved throughouteukaryotes, but its function remains unknown. We demonstrate thathobbit mutant animals have systemic growth defects because theyfail to secrete insulin. Other regulated secretion events also fail inhobbit mutant animals, including mucin-like ‘glue’ protein secretionfrom the larval salivary glands. hobbitmutant salivary glands produceglue-containing secretory granules that are reduced in size.Importantly, secretory granules in hobbit mutant cells lack essentialmembrane fusion machinery required for exocytosis, includingSynaptotagmin 1 and the SNARE SNAP-24. These membranefusion proteins instead accumulate inside enlarged late endosomes.Surprisingly, however, the Hobbit protein localizes to the endoplasmicreticulum. Our results suggest that Hobbit regulates a novel step inintracellular trafficking of membrane fusion proteins. Our studies alsosuggest that genetic control of body size, as a measure of insulinsecretion, is a sensitive functional readout of the secretorymachinery.

KEY WORDS: Body size, Growth, Insulin secretion, Regulatedexocytosis, Intracellular trafficking, Drosophila

INTRODUCTIONBody size is exquisitely regulated during metazoan development. InDrosophila, an explosive period of growth occurs during larvaldevelopment; however, this systemic growth ceases and final bodysize is specified at the onset of metamorphosis (Nijhout et al., 2014).One of the crucial regulators of body size in Drosophila is theinsulin signaling pathway. In this pathway, Drosophila insulin-likepeptides (Dilps) are secreted from the insulin-producing cells(IPCs), which are neuroendocrine cells located in the larval centralnervous system. The IPCs function in a manner that is analogous tomammalian pancreatic β-cells (Geminard et al., 2006). Dilps thencirculate systemically and bind to the insulin receptor (InR) on themembrane of target cells, triggering a downstream signalingcascade that regulates both metabolism and cellular growth(Brogiolo et al., 2001; Rulifson et al., 2002). Disruption of theinsulin signaling pathway has a profound effect onDrosophila bodysize. Ablation of the IPCs (Rulifson et al., 2002), genetic deletion ofDilps (Grönke et al., 2010), or mutation of targets downstream of

InR (Böhni et al., 1999; Brogiolo et al., 2001; Murillo-Maldonadoet al., 2011) all result in a dramatic reduction in body size. However,despite the importance of the insulin signaling pathway for systemicgrowth in flies and for disease in humans, the factors regulatinginsulin secretion remain largely unknown.

Insulin secretion in both flies and mammals occurs via regulatedexocytosis. This crucial cellular process begins with the biogenesisof cargo protein-containing secretory granules. Cargo proteinstargeted to the regulated exocytosis pathway, such as insulin, aretranslated by ribosomes bound to the rough endoplasmic reticulum(RER) and translocated into the ER lumen co-translationally. Thesecargo proteins are then transported to ER exit sites (ERES),specialized ER subdomains characterized by a coating of coatprotein complex II (COPII), and packaged into COPII-coatedvesicles (Brandizzi and Barlowe, 2013). COPII-coated vesiclesmediate trafficking of cargo proteins from the ER to the Golgicisternae; in the Golgi, cargo proteins undergo additional processingand post-translational modification. Cargo proteins are thenselectively packaged into secretory granules, specializedorganelles that bud from the trans-Golgi Network (TGN) asimmature secretory granules (Kögel and Gerdes, 2010). Theimmature granules undergo a maturation process to refine theircontent and membrane composition, rendering them competent forexocytosis (Kögel and Gerdes, 2010). Mature secretory granulestraffic to the plasma membrane, where they undergo tethering,docking and priming steps, attaching them to the membrane so thatcargo proteins can be quickly released to the extracellularenvironment in a stimulus-dependent manner (Burgoyne andMorgan, 2003).

The membrane fusion events leading to extracellular release aremediated by membrane fusion proteins, such as synaptotagmins andsoluble N-ethylmaleimide-sensitive factor (NSF) activating proteinreceptors (SNAREs) (Burgoyne and Morgan, 2003; Hong, 2005).To drive membrane fusion during regulated exocytosis, theseproteins must be properly trafficked to secretory granulemembranes. Like cargo proteins, secretory granule membraneproteins must first enter the ER. Synaptotagmins contain a single-pass transmembrane domain at their N terminus (Südhof, 2002) thatacts as a signal, directing the ribosome to the RER and allowingco-translational translocation of the nascent protein into the ER(Hegde and Keenan, 2011). In contrast, most SNAREs contain asingle-pass transmembrane domain at their C terminus and aretherefore translated by free ribosomes in the cytosol. SNAREs mustthen be post-translationally translocated into the ER, a process thatrequires chaperones to prevent the nascent SNARE fromcomplexing with SNAREs on other organelles and to shelter thehydrophobic transmembrane domain, preventing aggregation(Hegde and Keenan, 2011; Kutay et al., 1995). Once in the ER,secretory granule membrane proteins are thought to follow a similarpath to that of cargo proteins: they leave the ER via ERESs, trafficthrough the Golgi cisternae, and are eventually loaded ontoReceived 9 November 2017; Accepted 8 May 2018

1Division of Pharmaceutical Sciences, University of Wisconsin-Madison, Madison,WI 53705-2222, USA. 2Cellular and Molecular Biology Graduate Program,University of Wisconsin-Madison, Madison, WI 53706, USA.

*Author for correspondence ([email protected])

S.D.N., 0000-0001-5731-6426; A.B., 0000-0001-7150-1775

1

© 2018. Published by The Company of Biologists Ltd | Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 2: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

secretory granules produced in the TGN. However, despite theimportance of regulated exocytosis for both normal developmentand disease, many of the trafficking steps during this process remainlargely unknown.Here, we report the identification and characterization of hobbit,

a novel and conserved protein that plays a crucial role in thetrafficking of secretory granule membrane proteins. We haveidentified hobbit in a chemical mutagenesis screen for Drosophilamutant animals with a small body size. hobbit is highly conservedthroughout eukaryotes, but the protein has not been functionallycharacterized.We find that hobbitmutant animals have a small bodysize because they fail to secrete insulin and demonstrate that hobbitis required for maintenance of the major endocrine signalingaxis that regulates insulin release. Furthermore, other regulatedexocytosis events are also disrupted in hobbit mutant animals,including the secretion of mucin-like ‘glue’ proteins from the larvalsalivary glands at the onset of metamorphosis. hobbit mutant cellsproduce secretory granules that contain cargo proteins; however,these granules lack SNAREs and synaptotagmins and therefore arenot competent for release, providing an explanation for whysecretion fails upon loss of hobbit function. We find that the Hobbitprotein itself localizes to the ER, suggesting that Hobbit may play anew role in trafficking of secretory granule membrane proteinsthrough the secretory pathway.

RESULTSA chemical mutagenesis screen for new regulators of bodysizeTo find new regulators of body size in Drosophila, we haveconducted a large-scale ethyl methanesulfonate (EMS) mutagenesisscreen on the third chromosome and identified metamorphosis-specific lethal mutations. We screened over 26,000 mutagenizedchromosomes and identified 8636 lethal mutations; of these, 566were pupal lethal mutations (Wang et al., 2008). Among the 566, weidentified 26 mutant strains in 19 complementation groups thatdisplayed a dramatic small pupa (sP) phenotype.We first focused on

mapping one of the mutants with a dramatic small body size: sP1.Recombination mapping with pairs of dominant markers (Sapiroet al., 2013) followed by complementation tests with chromosomaldeficiencies and known lethal mutations identified sP1 as an alleleof the insulin receptor (InR) (Fig. S1A-C), a gene that has previouslybeen shown to play a crucial role in body size determination (Chenet al., 1996; Fernandez et al., 1995). InRsP1 does not contain anysequence changes within the InR coding sequence; however, InRmRNA transcript levels were significantly reduced in the mutantanimals (Fig. S1D), suggesting that this allele contains a lesiondisrupting critical regulatory sequences. The identification of anovel allele of InR confirms that our screening approach hasuncovered bona fide regulators of systemic growth.

Identification of hobbit, a novel and conserved regulator ofbody sizeWe next mapped the most frequently mutated locus (sP2), whichcontained five alleles. Recombination mapping with dominantmarkers placed this mutation between Roughened (R) and Dichaete(D); however, the mutant animals complemented all chromosomaldeficiencies in this region, indicating that the disrupted gene waslocated in a gap between chromosomal deletions (Fig. S2A).Subsequent Sanger sequencing of candidate genes not covered byavailable deletions revealed lesions in a previously uncharacterizedgene, CG14967 (Fig. S2A). We named this gene hobbit because ofthe dramatic small pupa phenotype displayed by the mutant animals(Fig. 1A and Fig. 2). All five hobbit alleles contained nonsensemutations (Fig. 1C, Fig. S2A), and all exhibited the same smallbody size and lethal phase during metamorphosis (Fig. S2B),indicating that these are likely loss-of-function alleles. Although themajority of hobbit mutant animals arrested during metamorphosis,some animals appeared to die earlier in development, as only 80%of the expected progeny were observed as pupae on the side of vials(hob2/hob3, total n=274). Importantly, hobbit is conserved acrosseukaryotes, with orthologs in fungi, green plants, protists andanimals (Fig. 1B). Additionally, a protein sequence identity plot

Fig. 1. Identification of hobbit, a novel and conservedregulator of systemic growth. (A) hobbit mutant animalsexhibit a dramatic small pupa phenotype. (B) ‘Tree of life’phylogeny showing hobbit orthologs throughout Eukaryota, ingreen plants (green), animals (blue), fungi (red) and someprotists (purple). (C) Protein identity plot showing primarysequence conservation between D. melanogaster hobbit and12 metazoan orthologs (Caenorhabditis elegans, Bombyxmori, Anopheles gambiae, Danio rerio, Xenopus tropicalis,Gallus gallus,Oryctolagus cuniculus,Rattus norvegicus,Musmusculus, Macaca mulatta, Pan troglodytes and Homosapiens); horizontal blue dashed line shows 100% identity.Vertical red dashed lines indicate the position of Drosophilahobbit mutant alleles, which are all nonsense mutations.hob: hobbit.

2

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 3: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

among 13 animal model organisms (from worms to humans)revealed a high degree of primary sequence conservation across thelength of the protein (Fig. 1C). However, no functionallycharacterized domains are predicted beyond a possibletransmembrane domain at the N terminus. Additionally, sequencehomology searches did not reveal any hits outside of hobbit andits orthologs. These results indicate that hobbit encodes a trulynovel protein.Without any obvious molecular signatures or protein domains to

guide our studies, we relied on phenotypic clues to characterize thefunction of hobbit. We began these phenotypic studies by validatingour hobbit mutations using RNAi knockdown and rescueexperiments. Publicly available RNAi lines did not affect hobbitexpression levels; therefore, we generated two new RNAi constructsthat significantly reduced hobbit levels (Fig. S3C). Importantly,ubiquitous hobbit RNAi knockdown phenocopied both the smallbody size and lethality of our mutant animals (Fig. 2). To conduct arescue experiment, we cloned the wild-type fly Hobbit protein andubiquitously overexpressed it in the hobbitmutant background. Thistreatment rescued both the small body size and metamorphosislethality of hobbit mutant animals (Fig. 2). Additionally, we clonedthe human ortholog of hobbit, Hs hob (KIAA0100), andubiquitously overexpressed it in the hobbit mutant background;this treatment also rescued both body size and lethality (Fig. 2),suggesting that the function of hobbit is evolutionarily conserved.

hobbit regulates body size in a non-cell-autonomous mannerAs a first step toward understanding the function of hobbit, wesought to determine why the mutant animals are small. We firstwanted to determine whether hobbit had a cell-autonomous effecton growth. To do this, we generated somatic clones in both mitotic(wing disc) and endocycling (salivary gland) cells and measured the

area of the mutant clones and twin spots. Although hobbit mutanttissues were small (Fig. 3A), the ratio of clone:twin spot area in bothwing disc and salivary gland cells was 1:1 (Fig. 3B), suggesting thathobbit has a non-cell-autonomous effect on growth. We next usedqPCR to determine whether hobbit mRNA expression levels wereregulated in a stage- or tissue-specific manner. We found that hobbitappeared to be ubiquitously expressed across wandering L3 (wL3)tissues (Fig. S3B). Additionally, during development, we observedthat expression levels were highest in early embryos (Fig. S3A),suggesting that hobbit mRNA may be maternally loaded. hobbitthen appears to be expressed at relatively steady levels in wholeanimals, with only a modest increase in expression from early larvalstages to the onset of metamorphosis (Fig. S3A). However, levels ofhobbit do not appear to dictate body size, as ubiquitousoverexpression of hobbit had no effect on final size (Fig. S3D).Altogether, these results indicate that hobbit does not have a cell-autonomous effect on growth and therefore must regulate body sizein a non-cell-autonomous manner.

We next began to investigate known causes of non-cell-autonomous growth defects in Drosophila. Developmental timingabnormalities leading to a shortened larval growth phase andprecocious entry into metamorphosis can cause reduced body sizethrough non-cell-autonomous mechanisms (King-Jones et al.,2005). Larvae with this defect develop normally until the thirdlarval instar, but then prematurely enter metamorphosis. We used amarker, called Sgs3-GFP, which is induced in the larval salivaryglands at the mid-third larval instar transition, to synchronizeanimals at this developmental time point and then measure theduration of the final stage of the third larval instar. All controlanimals entered metamorphosis within 24 h of Sgs3-GFPexpression (Fig. 3D, solid line). hobbit mutant animals inducedexpression of Sgs3-GFP, indicating that the animals reached the

Fig. 2. hobbit mutant animals can be rescued by fly andhuman proteins. Size quantification and lethal phase analysisfor control, ubiquitous hobbit-RNAi [UAS-Dcr/+; tub>/UAS-hob(i) 1 and 2], hobbit mutants (hob2/hob3), hobbit mutantcontrols for rescue experiments (act>,hob2/hob3, hob2/UAS-hob,hob3 and UAS-Hs-hob, hob2/hob3), fly hobbit mutantrescue (act>,hob2/UAS-hob,hob3), and human ortholog rescue(UAS-Hs-hob,hob2/act>,hob3). Ubiquitous expression ofhobbit-RNAi reduces body size and results in lethality duringmetamorphosis. Controls for rescue experiments do notsignificantly alter body size or lethality compared with hob2/hob3 mutant animals. Expression of fly hobbit or the humanortholog (Hs-hob) both rescue body size and lethality. Body sizequantified by pupa volume is expressed as a percentagerelative to wild type (100%). Data are shown as mean±s.e.m.Volume quantification and lethal phase analysis wereperformed on the same animals; n=100 animals per genotype.> in genotypes denotes GAL4; for example, ‘act>’ is shorthandfor act-GAL4. A, adult; PA, pharate adult; P/iP, pupa/incompletepupa; PP, prepupa.

3

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 4: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

mid-third instar transition. Importantly, hobbit mutant animals didnot pupariate precociously; instead, hobbit mutant animals weredevelopmentally delayed following the mid-third instar transition,with some animals taking >40 h to pupariate after induction of Sgs3-GFP (Fig. 3D, dashed line). To analyze this developmental delayphenotype further, we also measured timing throughout the larvalgrowth phase [from L1 to puparium formation (PF)]. We found thathobbitmutant animals were developmentally delayed and developedasynchronously throughout larval development (Fig. 3C; comparethe slopes of the solid and dashed lines). Taken together, theseresults indicate that hobbit regulates growth non-cell-autonomously,but a precocious developmental timing defect is not responsible forthe small body size of hobbit mutant animals.

Insulin secretion fails in hobbit mutant animalsThe insulin signaling pathway plays a central role in the regulationof Drosophila body size. Interestingly, hobbit mutant animalsexhibited a dominant genetic interaction with our mutant allele ofInR; double heterozygous animals were significantly smaller thaneither single heterozygote (Fig. S4A), suggesting that hobbit mayplay a role in the insulin pathway. Like loss of insulin itself, hobbithas a non-cell-autonomous effect on growth, leading us to testwhether hobbit mutant animals had defects in insulin secretion.Consistent with previous reports (Géminard et al., 2009), control

(fed) animals secreted insulin and therefore exhibited little stainingfor Dilp2 (Ilp2 – FlyBase) or Dilp5 (Ilp5 – FlyBase) in the IPCs;however, starved animals retained insulin, resulting in bright Dilp2/5 signal that returned to steady-state levels within 2 h of re-feeding(Fig. 4A, Fig. S5A). In contrast, hobbit mutant animals exhibitedbright Dilp2 and Dilp5 signal in fed, starved, and re-fed states(Fig. 4B, Fig. S5B), indicating that these animals do not properlysecrete insulin and suggesting that the insulin secretion defect mayunderlie the small body size of hobbit mutant animals.

To confirm this observed insulin secretion defect, we analyzedinsulin/insulin-like growth factor (IGF) signaling (IIS) responses inperipheral tissues. When insulin is secreted, it binds to InR,initiating a downstream kinase signaling cascade that includes theactivation of phosphatidylinositol 3-kinase (PI3K), a lipid kinasethat phosphorylates phosphatidylinositol 4,5-bisphosphate (PIP2)to generate phosphatidylinositol 3,4,5-trisphosphate (PIP3) (Brittonet al., 2002). We used a ubiquitously expressed GFP-tagged PIP3reporter (tGPH) (Britton et al., 2002) to determine plasmamembrane levels of PIP3 in the fat body of control and hobbitmutant animals, and saw a reduction in plasma membrane signal inhobbit mutant fat body cells (Fig. S4B). This result indicates thathobbit mutant tissues contain less PIP3, indicative of reduced IISsignaling. Additionally, activation of IIS inhibits activity of thetranscription factor foxo; therefore, reduced IIS should result in

Fig. 3. hobbit has a non-cell-autonomous effect on growth. (A) Wing imaginal discs dissected from wandering L3 (wL3) larvae of control and hobbit mutant(hob1/hob2) animals show that tissues are small in hobbit mutant animals. Nuclei stained with DAPI (white). (B) hob2 mutant Flp/FRT clones (marked byloss of GFP and outlined in white) are the same size as paired twin spots (marked by 2× GFP and outlined in red) in both wing discs and salivary glands.Quantification of the ratio of clone:twin spot area (in arbitrary units, a.u.) confirms a 1:1 ratio between clone area:twin spot area. Data are shown as mean±s.d.n=10 clones/twin spots per tissue. (C) Analysis of developmental timing throughout larval development in control and hobbit mutant animals. Animals weresynchronized at 0-4 h after hatching from the embryo (0-4 h after L1), then allowed to age until the onset of metamorphosis (puparium formation). hobbitmutantanimals (dashed line) are delayed and asynchronous compared with control animals (solid line). y-axis shows percentage of animals pupariated, x-axisshows developmental time point in h after L1. Three replicates of n>50 animals (control total n=172, hob2/hob3 total n=214) were analyzed; data are shownas mean±s.d. (D) Analysis of developmental timing during the third larval instar in control and hobbit mutant animals. Animals were synchronized at 0-4 hafter the start of the mid-third larval instar transition (0-4 h after Sgs3-GFP expression), then allowed to age until puparium formation. All control animalspupariate within 24 h of Sgs3-GFP expression (solid line); in contrast, hobbit mutant animals take considerably longer to pupariate (dashed line). y-axis showspercentage of animals pupariated, x-axis shows developmental time point in h after Sgs3-GFP expression. Three replicates of n>40 animals (control total n=146,hob2/hob3 total n=137) were analyzed; data are shown as mean±s.d. Scale bars: 50 µm.

4

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 5: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

increased expression of foxo target genes, such as InR and thetranslational regulator 4E-BP (Thor – FlyBase) (Jünger et al., 2003;Puig et al., 2003). Using qPCR, we observed significantupregulation of both InR and 4E-BP in hobbit mutant wholeanimals (Fig. S4C). Taken together, these results indicate that IIS isreduced in hobbit mutant peripheral tissues, as would be expectedwith a defect in insulin secretion.We next wanted to determine which tissues require hobbit function

for insulin secretion. IPC-specific hobbit RNAi knockdownsignificantly reduced body size (Fig. 4C), suggesting that hobbit isrequired in the IPCs for insulin secretion. Consistent with this result,pan-neuronal RNAi knockdown of hobbit using elav-GAL4 alsoreduced animal body size (Fig. S4D). However, IPC-specific

expression of hobbit in the mutant background did not rescue bodysize (Fig. 4C,D) or insulin secretion (Fig. 4E, Fig. S5C), suggestingthat additional tissues may require hobbit function for insulin release.Previous studies have shown that theDrosophila fat body functions asa nutrient sensor, releasing secreted signals that circulate systemicallyand promote insulin release in response to dietary inputs (Delanoueet al., 2016; Géminard et al., 2009; Rajan and Perrimon, 2012),prompting us to test whether hobbitwas also required in the fat body.Fat body-specific RNAi knockdown of hobbit significantly reducedbody size (Fig. 4C). In contrast, fat body-specific expression of hobbitin mutant animals did not rescue body size (Fig. 4C,D) or insulinsecretion (Fig. 4E, Fig. S5C). However, expression of hobbit in boththe IPCs and fat body significantly rescued the small body size of

Fig. 4. hobbit is required for insulin secretion.(A) Immunofluorescence staining for Drosophila insulin-likepeptide Dilp2 (red) in IPCs (marked with GFP in green) ofcontrol animals shows that Dilp2 is secreted under fedconditions, retained under starved conditions, and secretedwithin 2 h of re-feeding. (B) Dilp2 staining in hobbit mutantIPCs in fed, starved, and re-fed states shows that Dilp2 is notsecreted under any condition. (C) IPC-specific hobbit RNAiknockdown (with Dilp2-GAL4) or fat body-specific hobbitRNAi knockdown (with Cg-GAL4) reduces body size.However, IPC- or fat body-specific (with ppl-GAL4 orCg-GAL4) hobbit expression does not rescue hobbit mutantsmall pupae, whereas expression in both IPCs and fat bodydoes rescue size. None of these treatments rescues hobbitmutant lethality. Size was quantified by pupa volume andexpressed relative to wild type (100%); data are shown asmean±s.e.m. Volume and lethal phase data shown for hob2/UAS-hob,hob3 are the same as that shown in Fig. 2 but areincluded here for comparison. Volume measurements andlethal phase analysis were performed on the same animals;n=100 per genotype. ***P<0.0001 calculated by two-tailedt-test. (D) Representative image of body size rescue in hobbitmutant animals when hobbit is expressed in the IPCs only,fat body only, or IPCs and fat body. (E) Dilp2 (red) is notsecreted in fed hobbit mutant animals upon expression ofhobbit in the IPCs only (with Dilp2-GAL4) or fat body only(with ppl-GAL4). (F) Dilp2 (red) secretion is rescued in hobbitmutant animals when hobbit is expressed in both the IPCs(marked with GFP in green) and fat body (with Cg-GAL4).IPC images in all panels representative of n≥20 analyzedper condition/genotype. > in genotypes denotes GAL4; forexample, ‘Dilp2>’ is shorthand for Dilp2-GAL4. A, adult; FB,fat body; n.s., not significant; PA, pharate adult; P/iP, pupa/incomplete pupa; PP, prepupa. Scale bars: 50 µm.

5

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 6: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

hobbit mutant animals and restored the ability to secrete insulin(Fig. 4C,D,F, Fig. S5D). Other tissues, including the midgut (Songet al., 2017), prothoracic gland (Colombani et al., 2005) and muscles(Demontis et al., 2014; Langerak et al., 2018) also secrete signalsthat influence systemic growth and regulation; however, RNAiknockdown of hobbit in these tissues had no significant effect onbody size (Fig. S4D), suggesting that hobbit is primarily required inthe IPCs and fat body to regulate systemic growth. Taken together,these results indicate that hobbit regulates body size by maintainingthe major endocrine axis required for insulin secretion.

Secretion of mucin-like glue proteins is disrupted in hobbitmutant salivary glandsAlthough hobbit expression in both the IPCs and fat body rescuedbody size, it did not rescue lethality (Fig. 4C), suggesting that hobbithas essential functions in other tissues. Because the fat body andIPCs both utilize regulated exocytosis, we investigated whetherother regulated secretion events were disrupted in hobbit mutantanimals. First, hobbit mutant pupae have a soft cuticle, suggestingthat hobbit function may be required for deposition of chitin.Another conspicuous regulated exocytosis event during Drosophiladevelopment is the secretion of mucin-like ‘glue’ proteins from thelarval salivary glands at the onset of metamorphosis (Biyashevaet al., 2001). Glue proteins are synthesized in the acinar cells of thesalivary glands during the final day of larval development, secretedinto the lumen via regulated exocytosis just prior to pupariumformation, and expelled onto the surface of the animal at pupariumformation, allowing it to adhere to a solid surface duringmetamorphosis (Biyasheva et al., 2001; Kang et al., 2017; Roussoet al., 2016; Tran et al., 2015). Interestingly, glue secretion, assayedby imaging GFP-tagged glue protein (Sgs3-GFP), failed in hobbitmutant animals (Fig. S6A), and this effect was cell-autonomous, ashobbit mutant Flp/FRT somatic clones also failed to secrete glue(Fig. 5A). Taken together, these results demonstrate that hobbitplays a crucial, cell-autonomous role in regulated exocytosis.

Secretory granules in hobbit mutant tissues lack membraneproteins required for maturation and exocytosisOur next goal was to identify the mechanism by which hobbitregulates secretion. We first analyzed the morphology of the glue-containing secretory granules in hobbit mutant salivary glands todetermine whether there were any defects. We found that thesegranules were significantly smaller compared with controls(Fig. 5B, Fig. S6B), suggesting that the granules do not matureproperly, as glue granules are known in increase in size prior torelease (Niemeyer and Schwarz, 2000). Exocytosis and granule-granule fusion during maturation are mediated by the presence andactivity of membrane fusion proteins, such as SNAREs andsynaptotagmins, which are present on both secretory granules andtarget membranes. Strikingly, secretory granules in hobbit mutantsalivary glands do not contain these membrane fusion proteins.Salivary gland-specific expression of GFP-tagged Synaptotagmin 1(Syt1-GFP) showed that this protein was present around secretorygranules in control glands but absent from granule membranes inhobbit mutant glands (Fig. 5C). Immunofluorescence stainingwith an anti-Syt1 antibody confirmed these results (Fig. S6C).Furthermore, immunofluorescence staining for the SNARE proteinSNAP-24 also showed a similar mislocalization in hobbit mutantglands (Fig. S6D). Because secretory granules in hobbit mutantcells do not contain membrane fusion proteins, these granules arenot competent to undergo maturation or exocytosis, explaining whyregulated exocytosis fails in hobbit mutant cells.

hobbit is required for normal trafficking of secretory granulemembrane proteinsWe next wanted to determine what happens to secretory granulemembrane proteins in hobbit mutant cells. We observed that theseproteins accumulated in large structures that were distinct fromsecretory granules (Fig. 6A, Fig. S6C,D). This result suggests thatgranule membrane proteins may aberrantly accumulate in one ormore specific subcellular compartments. To determine where Syt1and SNAP-24 accumulate, we used a collection of endogenouslyregulated, EYFP-tagged Rab proteins (Dunst et al., 2015), becauseRab proteins are markers for many subcellular compartmentsand organelles. We first compared Rab protein localization andmorphology in control and hobbit mutant salivary glands andobserved striking defects with Rab7; hobbit mutant salivary glandscontained numerous enlarged Rab7-positive late endosomes

Fig. 5. Secretory granules in hobbit mutant cells are not competent forexocytosis. (A) Flp/FRT hob2 clone, marked by loss of RFP-nls (gray)and outlined in white, shows that Sgs3-GFP glue proteins (red) are notsecreted in hobbit mutant salivary gland cells dissected from prepupalanimals. (B) Sgs3-GFP glue granules in hobbit mutant salivary glands aresignificantly smaller than those in controls. Graph shows quantification bygranule area; data are shown as mean±s.d. n≥60 granules per genotype.***P<0.001 calculated by two-tailed t-test. (C) Synaptotagmin-1-GFP(Syt1-GFP, in green, expression driven in salivary glands by hs-GAL4), isloaded onto glue granule membranes in control glands but not in hobbit mutantglands. Glue protein shown in red. All images were acquired from live, unfixedtissue. Scale bars: 50 µm (A); 5 µm (B,C).

6

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 7: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

(Fig. 6B). Furthermore, secretory granule membrane proteinsaccumulated inside the enlarged late endosomes (Fig. 6C). Lateendosomes are essential sorting sites, traditionally thought to directproteins to either recycling or degradative pathways (Maxfield andMcGraw, 2004). These results indicate that trafficking of secretorygranule membrane proteins is severely disrupted in hobbit mutantsecretory cells.To begin to understand how hobbit regulates trafficking of

secretory granule membrane proteins, we looked at where theHobbit protein localized within the cell. To do this, we generateda GFP-tagged hobbit overexpression construct. Ubiquitousoverexpression of this Hobbit-GFP construct rescued hobbitmutant animals (Fig. S7A), indicating that the constructproduced functional Hobbit protein. Given that Syt1 andSNAP-24 are trapped inside Rab7-positive endosomes in hobbitmutant cells, we first tested whether Hobbit-GFP colocalizedwith Rab7. However, we did not observe any significant overlapbetween Hobbit-GFP and Rab7 (Fig. 6D). We next tested thepossibility that Hobbit localizes to secretory granules to facilitateloading of secretory granule membrane proteins. Instead, weobserved that Hobbit-GFP was present between secretorygranules and did not appear to colocalize with SNAP-24(Fig. 7B), indicating that Hobbit is not enriched on secretorygranule membranes. Given that secretory granules are generatedin the Golgi body, we also tested whether Hobbit-GFPcolocalized with a pan-Golgi marker, but we did not observeany colocalization between Hobbit-GFP and Golgi-RFP(Fig. S7B). In contrast, we found that Hobbit-GFP stronglycolocalized with the ER marker KDEL-RFP (Fig. 7A),suggesting that Hobbit localizes to the ER. The presence ofmany secretory granules dramatically condenses the cytoplasm;therefore, to confirm that Hobbit-GFP localizes to the ER, weanalyzed Hobbit-GFP localization in salivary glands at pupariumformation, the stage immediately following the completion ofglue secretion. Hobbit-GFP did not appear to colocalize with acytoplasmic/soluble mCherry at puparium formation (Fig. 7C),nor did KDEL-RFP appear to colocalize with a cytoplasmic/soluble GFP (Fig. 7D). However, Hobbit-GFP and KDEL-RFPstrongly colocalized at puparium formation (Fig. 7E, Fig. S7C),demonstrating that Hobbit-GFP is strongly enriched within theER. Taken together, our results indicate that hobbit is a novel andconserved regulator of intracellular trafficking; furthermore,hobbit appears to play a specific role in trafficking secretorygranule membrane proteins through the secretory pathway.

DISCUSSIONRegulated exocytosis is a crucial process for both systemic signalingand homeostatic mechanisms in multicellular organisms. Secretorygranule membrane proteins, such as SNAREs and synaptotagmins,are indispensable for regulated exocytosis; without these proteins,secretory granules cannot mature properly, nor can they fuse withthe plasma membrane. Therefore, proper trafficking and loading ofthese proteins onto secretory granules is an essential step in theproduction of mature granules that are competent for release. In thisstudy, we have identified that hobbit, a novel and conserved protein,is required for proper trafficking of secretory granule membraneproteins during regulated exocytosis.

We first identified hobbit in a screen for Drosophila mutantanimals that arrest during metamorphosis with a small pupaphenotype. Other mutant animals with reduced body size havebeen identified in flies; however, most of these mutant alleles areviable and specifically disrupt insulin secretion or insulin signaling(Böhni et al., 1999; Brogiolo et al., 2001; Chen et al., 1996; Rajanand Perrimon, 2012). Additionally, ablation of the IPCs results inviable animals that are reduced in size (Rulifson et al., 2002),demonstrating that insulin secretion is not essential fordevelopment. As a result, screening efforts to identify animalswith reduced body size have primarily focused on mutations thatresult in viable animals. In contrast, our characterization of hobbitsuggests that lethal mutations with a small body size represent anew, uncharacterized class of systemic growth regulators. Mutantalleles in this class, such as hobbit, are required for insulin secretionfrom the IPCs but are also required for secretory events in othertissues. These crucial functions in other tissues underlie the lethalityobserved in the mutant animals. Further analysis and additionalscreening for these hobbit-like mutations, including those from oursmall pupa mutant collection, may reveal other important regulatorsof insulin secretion and intracellular trafficking.

Despite the striking evolutionary conservation of the Hobbitprotein, it does not contain any functionally characterized domainsor obvious signatures that provide clues to its molecular function.Our identity plot analysis reveals several highly conserved regionswithin the Hobbit protein; however, these conserved regions do notexhibit homology with any other proteins outside of hobbitorthologs. Databases and prediction programs, such as Pfam, doannotate potential protein domains within Hobbit; however, thelocation of these domains within the protein sequence does notcorrespond with the regions of high sequence conservation that wehave identified. Furthermore, these automated prediction algorithms

Fig. 6. Trafficking of membrane fusion proteins on secretory granules is disrupted in hobbit mutant cells. (A) Syt1-GFP (green) localizes in largestructures that are distinct from glue granules (red) in hobbitmutant cells.UAS-Syt1-GFPwas expressed in salivary glands using hs-GAL4. Images were acquiredfrom live, unfixed tissue. (B) Rab7-positive late endosomes (red) are dramatically enlarged in hobbit mutant salivary glands. Rab7-EYFP contains anEYFP tag at the endogenous Rab7 locus (Dunst et al., 2015). Images were acquired from live, unfixed tissue. (C) Rab7-positive endosomes (red) containaccumulations of the SNARE protein SNAP-24 (green) in hobbit mutant glands. (D) Hobbit-GFP (green, expression driven by Sgs3-GAL4) does not colocalizewith Rab7 (red). Scale bars: 5 µm.

7

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 8: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

annotate domains containing sequence motifs, such as ‘GFWDK’,that are not present in the Hobbit protein, suggesting that thedomains have been incorrectly assigned. Some bioinformaticprograms predict a transmembrane domain at the N terminus ofthe Hobbit protein. If present, a transmembrane domain wouldlikely play a crucial role in hobbit localization and function.However, detailed structure-function studies will be required tovalidate and assign a function to each of the uncharacterizedconserved regions within the Hobbit protein. These studies willreveal new insights into the molecular function of hobbit and mayalso identify new sequence motifs that are essential for theregulation of intracellular trafficking.

Our results demonstrate that hobbit plays a crucial role inregulated exocytosis by regulating trafficking of secretory granulemembrane proteins. Secretory granules in hobbit mutant cells lackSNAREs and synaptotagmins, membrane fusion proteins that arerequired for exocytosis. These secretory granule membrane proteinsaccumulate inside enlarged late endosomes in hobbit mutant cells.However, how hobbit regulates trafficking of SNAREs andsynaptotagmins remains unclear. Integral membrane proteins,such as synaptotagmins, are synthesized via co-translationaltranslocation into the ER. These proteins are still produced inhobbit mutant cells, suggesting that hobbit may function post-ERentry, such as in ER-to-Golgi traffic. However, disruption of ER exitdisrupts all secretory traffic (Ward et al., 2001) and would thereforebe expected to inhibit trafficking of secretory granule cargoproteins, but this process is unaffected in hobbit mutant cells.Therefore, our data could suggest that there is a selective pathwayrequired for ER-to-Golgi transport of secretory granule membraneproteins. We do not, however, observe accumulation of secretorygranule membrane proteins in the ER; instead, these proteinsaccumulate inside Rab7-positive late endosomes. A similaraccumulation phenotype has been described in mutant alleles ofPI4KII, a lipid kinase that appears to play an important role inretrograde trafficking (Burgess et al., 2012). Retrograde traffickingrecycles many proteins from endosomes to the TGN (Bonifacinoand Hurley, 2008), although it is not known whether secretorygranule membrane proteins undergo such recycling. A retrogradetrafficking function would be consistent with the phenotypes weobserve in hobbit mutant cells. However, given that knownregulators of retrograde trafficking, including PI4KII, localize toendosomes (Burgess et al., 2012), it is unclear how hobbit, an ER-localized protein, could regulate this process. Our data raises thepossibility that a new trafficking route from the ER, to endosomes,to secretory granules may be used for loading of secretory granulemembrane proteins. Direct sorting of cargo proteins fromendosomes to secretory granules has been described (Topalidouet al., 2016), raising the possibility that a similar trafficking routecould be used for secretory granule membrane proteins. Althoughfurther studies will be required to resolve these questions, it appearsthat hobbit does not function in any of the characterized traffickingpathways that are known to play a role in regulated exocytosis.

In conclusion, here we have described hobbit, a novel and highlyconserved regulator of intracellular trafficking and regulatedexocytosis. Our studies of hobbit mutant phenotypes reveal thatthis protein regulates a new trafficking step that is required to delivermembrane proteins to secretory granules, rendering them competentfor release. Additionally, our data suggests that hobbit, and othermetamorphosis-lethal mutants with growth defects, define a newclass of body size regulators that drive growth by playing crucialroles in intracellular trafficking and secretion of insulin. Theseresults also highlight the power of unbiased chemical mutagenesisscreens to uncover new genes and new biology.

MATERIALS AND METHODSFly stocks, food and developmental stagingThe following stocks were obtained from the Bloomington DrosophilaStock Center: w1118, Df(3R)BSC677, Df(3R)ED6058, Df(3R)BSC678,Df(3R)Exel6186, InRE19, InR93Dj−4, InR05545, tGPH, UAS-Dcr-2,tub-GAL4, act-GAL4, elav-GAL4, Myo1A-GAL4, repo-GAL4, Mhc-GAL4,hs-flp, neoFRT80B, Ubi-GFP 61EF, Sgs3-GFP, Dilp2-GAL4, UAS-GFP,UAS-mCherry, ppl-GAL4, Cg-GAL4, Ubi-mRFP-nls 3L, UAS-Syt1-GFP,hs-GAL4, Sgs3-GAL4, Rab7-EYFP, UAS-Golgi-RFP, UAS-KDEL-RFP.Sgs3-DsREDwas kindly provided by A. Andres (University of Nevada, LasVegas, NV, USA), and phm-GAL4 was kindly provided by Carl Thummel

Fig. 7. Hobbit localizes to the endoplasmic reticulum. (A) Hobbit-GFP(green) strongly colocalizes with the endoplasmic reticulum (ER) markerKDEL-RFP (red) in wandering L3 salivary glands. Expression of Hobbit-GFPand KDEL-RFP was driven by Sgs3-GAL4; images were acquired from live,unfixed tissue. (B) Hobbit-GFP (expressed using Sgs3-GAL4; green) does notappear to colocalize with the secretory granule membrane protein SNAP-24(red). (C) Hobbit-GFP (green) does not colocalize with cytoplasmic mCherry(red) in glands at puparium formation (0 h PF). Expression was driven bySgs3-GAL4; images were acquired from live, unfixed tissue. (D) KDEL-RFP (red)does not colocalize with cytoplasmic GFP (green) in glands at 0 h PF.Expression was driven bySgs3-GAL4; images were acquired from live, unfixedtissue. (E) Hobbit-GFP, in green, strongly colocalizes with KDEL-RFP (red) inglands at 0 h PF. Expression was driven bySgs3-GAL4; images were acquiredfrom live, unfixed tissue. Scale bars: 50 µm (A); 5 µm (B-E).

8

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 9: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

(University of Utah, UT, USA). The five hobbit mutations and InRsP1 weregenerated in an EMS mutagenesis screen (Wang et al., 2008); the methodsused to map these mutations are described elsewhere (Sapiro et al., 2013).Experiments using hs-GAL4 were carried out without heat-shock treatment,as hs-GAL4 ‘leaks’ constitutively in the salivary glands. All experimentswere performed in temperature-controlled incubators at 25°C in uncrowdedbottles or vials. For insulin secretion and tGPH assays, larvae were picked at0-4 h after hatching from the embryo (0-4 h after L1) and grown on foodcontaining 34 g inactivated yeast, 83 g corn flour, 60 g sucrose and 10 g agarper liter. Forty-eight hours later, at the start of the third larval instar, theselarvae were starved on agar plates containing PBS/1% sucrose for 24 h.Larvaewere then re-fed for the appropriate time on the food described above.To generate somatic clones in the salivary glands, embryos were heat-shocked at 37°C for 30 min at 0-4 h after egg lay (AEL); for wing discs, thesame heat-shock treatment was given to larvae at 48-52 h AEL. Fordevelopmental timing from L1, larvae containing the Sgs3-GFP transgenewere collected at 0-4 after hatching from the embryo (0-4 h after L1) andallowed to age at 25°C until expression of Sgs3-GFP was detected. Animalswere then scored for puparium formation beginning 12 h after induction ofSgs3-GFP. For developmental experiments from Sgs3-GFP, larvaecontaining the Sgs3-GFP transgene were collected at 88 h AEL and scoredfor the absence of Sgs3-GFP expression. These larvae were then checkedevery 4 h for induction of Sgs3-GFP; once expressed, larvae were allowed toage at 25°C on grape agar plates with yeast paste until puparium formation.

Phylogenetic tree and identity plot analysishobbit orthologs were identified using the HMMER web server (Finn et al.,2015) with the CG14967 protein sequence as the query sequence.Significant hits (P≤0.0001) were submitted to phyloT (http://phylot.biobyte.de/) to generate a phylogenetic tree based on NCBI taxonomyinformation. We used Interactive Tree of Life (Letunic and Bork, 2016) tovisualize, edit and annotate the final tree. To assay amino acid conservation,we used ClustalOmega (Sievers et al., 2011) to generate a protein alignmentbetween fly hobbit and its orthologs in C. elegans, B. mori, A. gambiae,D. rerio, X. tropicalis, G. gallus, O. cuniculus, R. norvegicus,M. musculus,M. mulatta, P. troglodytes and H. sapiens. The resulting alignment wasvisualized in JalView (Waterhouse et al., 2009), and identity/similarityscores at each amino acid position were extracted to generate the finalidentity plot.

Molecular cloning and transgenic animal generationTo generate the UAS-hobbit transgene, the full-length hobbit codingsequence was amplified from w1118 cDNA using the following primers: 5′-AAAAGGTACCATGATGCTACAGCTACTTCTATTCTGCCTGG-3′ and5′-AAAAAAGCGGCCGCTCAATCATTTCCCGATCTCTTTCCG-3′.The resulting ∼6.9 kb product was digested with KpnI and NotI, ligated intothe Gateway entry vector pENTR 1A (Invitrogen), and sequence verified.LR Clonase (Invitrogen) was used to recombine the entry clone into thepTW (Drosophila Genomics Resource Center) destination vector.Successful recombination was confirmed by sequencing. The resultingUAS-hobbit plasmid was then injected into w1118 flies using standardmethods (Genetic Services). A similar strategy was used to generate theUAS-KIAA0100 transgene. The KIAA0100 coding sequence was amplifiedfrom oligo(dT)-primed qPCR human reference cDNA (Clontech) using thefollowing primers: 5′-CCCCCCTTTAAAACCATGCCTCTGTTCTTCT-CCGC-3′ and 5′-AAAAAAGCGGCCGCTTTCATTTGCGCCTGCCAA-A-3′. The resulting ∼6.2 kb product was digested with DraI and NotI, andentry and destination clones were generated as described above. KIAA0100has many predicted isoforms; therefore, only the full-length productmatching the NCBI consensus coding sequence (CCDS) was used.Sequence-verified plasmids were injected into w1118 flies using standardprotocols (Genetic Services). To generate UAS-hobbit-GFP, we firstamplified the GFP coding sequence from pTGW (Drosophila GenomicsResource Center) using the following primers: 5′-AAAAAAGGTACCA-GTGAGCAAGGGCGAGGAGCT-3′ and 5′-AAAAAATCTAGACTTG-TACAGCTCGTCCATGC-3′. The resulting PCR product was digested withKpnI and XbaI and ligated into pBID-UASC-G (Wang et al., 2012). Theplasmid was sequence verified, and the hobbit entry clone was recombined

into this destination vector as described above. Sequence-verified plasmidswere injected into VK00027 flies for phiC31-mediated site-directedintegration (Rainbow Transgenic Flies). To generate the UAS-hobbit-RNAi lines, we followed the strategy outlined by the Transgenic RNAiProject (TRiP) (Ni et al., 2011). The primers for hobbit-RNAi 1 were: 5′-CTAGCAGTATGATGCTACAGCTACTTCTATAGTTATATTCAAGCA-TATAGAAGTAGCTGTAGCATCATGCG-3′ and 5′-AATTCGCATGAT-GCTACAGCTACTTCTATATGCTTGAATATAACTATAGAAGTAGCT-GTAGCATCATAC-3′. The primers for hobbit-RNAi 2 were: 5′-CTAG-CAGTATGCAGCGAATTGTTGTTAAATAGTTATATTCAAGCATAT-TTAACAACAATTCGCTGCATGCG-3′ and 5′-AATTCGCATGCAG-CGAATTGTTGTTAAATATGCTTGAATATAACTATTTAACAACAA-TTCGCTGCATACTG-3′. The primers were annealed, phosphorylated,and ligated into pVALIUM20 (TRiP) that had been linearized by digestionwith NheI and EcoRI. The resulting plasmids were sequence verified andinjected for phiC31-mediated site-directed integration into attP2-containingflies (Rainbow Transgenic Flies). (Note: A newly generated RNAi constructfrom TRiP contains the same shRNA hairpin as hobbit-RNAi 2 inserted onchromosome 2; hobbit-RNAi 2 is inserted on chromosome 3.)

Pupa volume quantification and lethal phase analysisAll light microscope images of pupae were captured on an Olympus SXZ16stereomicroscope coupled to an Olympus DP72 digital camera with DP2-BSW software. All pupae pictured together were originally captured in thesame image but were individually rotated and aligned post-acquisition toimprove image aesthetics. To quantify body size, we used Photoshop CS6 tomeasure the length and width of each pupa, and calculated pupa volumeusing a published formula (Delanoue et al., 2016). No statistical methodswere used to determine sample size. Animals were grown at controlleddensity and on a controlled diet to avoid any nutritional effects on body size.After image capture, the pupae were transferred to a grape agar plate andaged at 25°C for one week for lethal phase analysis. Lethal phases weredetermined using Bainbridge and Bownes staging criteria (Bainbridge andBownes, 1981).

Quantitative real-time PCR (qPCR)qPCR was performed as previously described (Ihry et al., 2012). Total RNAwas extracted from whole animals or tissues from the appropriatedevelopmental stage and genotype using the RNeasy Plus Mini Kit(Qiagen), and cDNAwas synthesized from 100-400 ng of total RNA usingSuperScript III First-Strand Synthesis System (Invitrogen). For wL3 pairedtissue samples, the wing discs, salivary glands, central nervous system, andthe rest of the carcass were dissected from the same pool of animals to enablea direct comparison of expression levels across tissues. qPCR wasperformed on a Roche LightCycler 480 using LightCycler 480 SYBRGreen I Master Mix (Roche). In all experiments, samples were runsimultaneously in biological triplicate, and amplification efficiencies werecalculated for each primer pair. The Relative Expression Software Tool(REST) (Pfaffl et al., 2002) was used to calculate relative expression. RESTcalculates standard error by using a confidence interval centered on themedian, which generates error bars that reflect asymmetric tendencies in thedata. Primer sequences for rp49 (RpL32 – FlyBase) (Ihry et al., 2012) and4E-BP (Demontis and Perrimon, 2010) were published previously. Newprimers were designed using ApE for hobbit or FlyPrimerBank (Hu et al.,2013) for InR. hob F: 5′-TTTGGTGAAGAGGTTTCGGGTC-3′; hob R:5′-TCTTTGTTGATGCGAATGTCACG-3′. InR F: 5′-GAAGTGGAGA-CGACGGGTAAA-3′; InR R: 5′-TCGCGCTGTTGTCGATTGTT-3′.

Immunofluorescence staining and confocal microscopyTissues were dissected from the appropriate animals, fixed for 30 min inPBS with 0.1% Triton X-100 (PBST) and 4% formaldehyde, blockedovernight in PBST/4% bovine serum albumin (BSA) at 4°C, and stainedusing the appropriate primary and secondary antibodies diluted in PBST/4%BSA. Primary antibodies used were rat anti-Dilp2 (1:200; Géminard et al.,2009), rabbit anti-Dilp5 (1:800; Géminard et al., 2009) (both a gift fromP. Leopold, Institute of Biology Valrose, France), rabbit anti-SNAP-24(1:200; Niemeyer and Schwarz, 2000) (gift from T. Schwarz, HarvardMedical School, MA, USA), mouse anti-Syt1 (1:50; Developmental Studies

9

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 10: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

Hybridoma Bank, 3H2 2D7), mouse anti-Rab7 (1:50; DevelopmentalStudies Hybridoma Bank, Rab7). The Rab7 antibody was generated andvalidated previously (Riedel et al., 2016). Secondary antibodies were used at1:200 and included anti-rat Alexa Fluor 568 (Invitrogen, A11077), anti-rabbit Alexa Fluor 633 (Invitrogen, A21071), anti-rabbit Cy3 (JacksonImmunoResearch Laboratories, 711-165-152), anti-mouse Cy3 (JacksonImmunoResearch Laboratories, 715-165-150). DAPI was used to labelnuclei (1:1000; Invitrogen). Stained tissues were mounted in Vectashield(Vector Laboratories). Images were captured on an Olympus FluoViewFV1000 confocal microscope and optimized with FV10-ASW software. Forlive tissue imaging, tissues were dissected and mounted in PBS and imagedfor no longer than 5 min after dissection. IPC images were acquired from az-series comprising 40 slices at a 0.59 µm step size. Identical laser settingsand optimization parameters were used for all samples; images shownas sum-intensity projections generated by FV10-ASW software.Quantification of clone:twin spot area ratios and glue granule areas wasperformed using ImageJ. Although true blinding was not possible given theexperimental conditions and genotypes analyzed, whenever possible, resultswere independently confirmed by another lab member who was blinded tothe genotype or experimental condition of the samples being analyzed.

AcknowledgementsStocks obtained from the Bloomington Drosophila Stock Center (NIHP40OD018537) and reagents obtained from the Drosophila Genomics ResourceCenter (NIH 2P40OD010949) were used in this study. We are grateful to JessicaSmoko for technical assistance and Yunsik Kang for helpful discussions.

Competing interestsThe authors declare no competing or financial interests.

Author contributionsConceptualization: S.D.N., A.B.; Methodology: S.D.N., A.B.; Validation: S.D.N.;Formal analysis: S.D.N., A.B.; Investigation: S.D.N.; Data curation: S.D.N.; Writing -original draft: S.D.N.; Writing - review & editing: S.D.N., A.B.; Visualization: S.D.N.,A.B.; Supervision: A.B.; Project administration: A.B.; Funding acquisition: A.B.

FundingThis work was supported in part by the National Institutes of Health (GM123204 toA.B.) and the National Science Foundation (GRFP DGE-1256259 to S.D.N.).Deposited in PMC for release after 12 months.

Supplementary informationSupplementary information available online athttp://dev.biologists.org/lookup/doi/10.1242/dev.161356.supplemental

ReferencesBainbridge, B. S. P. and Bownes, M. (1981). Staging the metamorphosis ofDrosophila melanogaster. Development 66, 57-80.

Biyasheva, A., Do, T.-V., Lu, Y., Vaskova, M. and Andres, A. J. (2001). Gluesecretion in the Drosophila salivary gland: a model for steroid-regulatedexocytosis. Dev. Biol. 231, 234-251.

Bohni, R., Riesgo-Escovar, J., Oldham, S., Brogiolo, W., Stocker, H., Andruss,B. F., Beckingham, K. and Hafen, E. (1999). Autonomous control of cell andorgan size by CHICO, a Drosophila homolog of vertebrate IRS1-4. Cell 97,865-875.

Bonifacino, J. S. and Hurley, J. H. (2008). Retromer. Curr. Opin. Cell Biol. 20,427-436.

Brandizzi, F. and Barlowe, C. (2013). Organization of the ER–Golgi interface formembrane traffic control. Nat. Rev. Mol. Cell Biol. 14, 382-392.

Britton, J. S., Lockwood, W. K., Li, L., Cohen, S. M. and Edgar, B. A. (2002).Drosophila’s insulin/PI3-kinase pathway coordinates cellular metabolism withnutritional conditions. Dev. Cell 2, 239-249.

Brogiolo, W., Stocker, H., Ikeya, T., Rintelen, F., Fernandez, R. and Hafen, E.(2001). An evolutionarily conserved function of the Drosophila insulin receptor andinsulin-like peptides in growth control. Curr. Biol. 11, 213-221.

Burgess, J., Del Bel, L. M., Ma, C.-I. J., Barylko, B., Polevoy, G., Rollins, J.,Albanesi, J. P., Kramer, H. and Brill, J. A. (2012). Type II phosphatidylinositol4-kinase regulates trafficking of secretory granule proteins in Drosophila. J. CellSci. 125, 3040-3050.

Burgoyne, R. D. and Morgan, A. (2003). Secretory granule exocytosis. Physiol.Rev. 83, 581-632.

Chen, C., Jack, J. and Garofalo, R. S. (1996). The Drosophila insulin receptor isrequired for normal growth. Endocrinology 137, 846-856.

Colombani, J., Bianchini, L., Layalle, S., Pondeville, E., Dauphin-Villemant, C.,Antoniewski, C., Carre, C., Noselli, S. and Leopold, P. (2005). Antagonisticactions of ecdysone and insulins determine final size in Drosophila. Science 310,667-670.

Delanoue, R., Meschi, E., Agrawal, N., Mauri, A., Tsatskis, Y., McNeill, H. andLeopold, P. (2016). Drosophila insulin release is triggered by adipose Stuntedligand to brain Methuselah receptor. Science 353, 1553-1556.

Demontis, F. and Perrimon, N. (2010). FOXO/4E-BP signaling in Drosophilamuscles regulates organism-wide proteostasis during aging. Cell 143, 813-825.

Demontis, F., Patel, V. K., Swindell, W. R. and Perrimon, N. (2014). Intertissuecontrol of the nucleolus via a myokine-dependent longevity pathway. Cell Rep. 7,1481-1494.

Dunst, S., Kazimiers, T., von Zadow, F., Jambor, H., Sagner, A., Brankatschk,B., Mahmoud, A., Spannl, S., Tomancak, P., Eaton, S. et al. (2015).Endogenously tagged rab proteins: a resource to study membrane trafficking inDrosophila. Dev. Cell 33, 351-365.

Fernandez, R., Tabarini, D., Azpiazu, N., Frasch, M. and Schlessinger, J. (1995).The Drosophila insulin receptor homolog: a gene essential for embryonicdevelopment encodes two receptor isoforms with different signaling potential.EMBO J. 14, 3373-3384.

Finn, R. D., Clements, J., Arndt, W., Miller, B. L., Wheeler, T. J., Schreiber, F.,Bateman, A. and Eddy, S. R. (2015). HMMERweb server: 2015 Update. NucleicAcids Res. 43, W30-W38.

Geminard, C., Arquier, N., Layalle, S., Bourouis, M., Slaidina, M., Delanoue, R.,Bjordal, M., Ohanna, M., Ma, M., Colombani, J. et al. (2006). Control ofmetabolism and growth through insulin-like peptides in Drosophila. Diabetes55, S5-S8.

Geminard, C., Rulifson, E. J. and Leopold, P. (2009). Remote control of insulinsecretion by fat cells in Drosophila. Cell Metab. 10, 199-207.

Gronke, S., Clarke, D.-F., Broughton, S., Andrews, T. D. and Partridge, L. (2010).Molecular evolution and functional characterization of Drosophila insulin-likepeptides. PLoS Genet. 6, e1000857.

Hegde, R. S. and Keenan, R. J. (2011). Tail-anchored membrane protein insertioninto the endoplasmic reticulum. Nat. Rev. Mol. Cell Biol. 12, 787-798.

Hong, W. (2005). SNAREs and traffic. Biochim. Biophys. Acta 1744, 120-144.Hu, Y., Sopko, R., Foos, M., Kelley, C., Flockhart, I., Ammeux, N., Wang, X.,

Perkins, L., Perrimon, N. and Mohr, S. E. (2013). FlyPrimerBank: an onlinedatabase for drosophila melanogaster gene expression analysis and knockdownevaluation of RNAi reagents. G3 (Bethesda) 3, 1607-1616.

Ihry, R. J., Sapiro, A. L., Bashirullah, A., Akagi, K. and Takai, M. (2012).Translational control by the DEAD box RNA helicase belle regulates ecdysone-triggered transcriptional cascades. PLoS Genet. 8, e1003085.

Junger, M. A., Rintelen, F., Stocker, H., Wasserman, J. D., Vegh, M.,Radimerski, T., Greenberg, M. E. and Hafen, E. (2003). The DrosophilaForkhead transcription factor FOXO mediates the reduction in cell numberassociated with reduced insulin signaling. J. Biol. 2, 20.

Kang, Y., Neuman, S. D. and Bashirullah, A. (2017). Tango7 regulates corticalactivity of caspases during reaper-triggered changes in tissue elasticity. Nat.Commun. 8, 603.

King-Jones, K., Charles, J.-P., Lam, G. and Thummel, C. S. (2005). Theecdysone-induced DHR4 orphan nuclear receptor coordinates growth andmaturation in Drosophila. Cell 121, 773-784.

Kogel, T. and Gerdes, H.-H. (2010). Maturation of secretory granules. ResultsProbl. Cell Differ. 50, 1-20.

Kutay, U., Ahnert-Hilger, G., Hartmann, E., Wiedenmann, B. andRapoport, T. A.(1995). Transport route for synaptobrevin via a novel pathway of insertion into theendoplasmic reticulum membrane. EMBO J. 14, 217-223.

Langerak, S., Kim, M.-J., Lamberg, H., Godinez, M., Main, M., Winslow, L.,O’Connor, M. B. and Zhu, C. C. (2018). The Drosophila TGF-beta/Activin-likeligands Dawdle and Myoglianin appear to modulate adult lifespan throughregulation of 26S proteasome function in adult muscle. Biol. Open 7, bio.029454.

Letunic, I. and Bork, P. (2016). Interactive tree of life (iTOL) v3: an online tool for thedisplay and annotation of phylogenetic and other trees. Nucleic Acids Res. 44,W242-W245.

Maxfield, F. R. and McGraw, T. E. (2004). Endocytic recycling. Nat. Rev. Mol. CellBiol. 5, 121-132.

Murillo-Maldonado, J. M., Sanchez-Chavez, G., Salgado, L. M., Salceda, R. andRiesgo-Escovar, J. R. (2011). Drosophila insulin pathway mutants affect visualphysiology and brain function besides growth, lipid, and carbohydratemetabolism. Diabetes 60, 1632-1636.

Ni, J.-Q., Zhou, R., Czech, B., Liu, L.-P., Holderbaum, L., Yang-Zhou, D., Shim,H.-S., Handler, D., Karpowicz, P., Binari, R. et al. (2011). A genome-scaleshRNA resource for transgenic RNAi in Drosophila. Nat. Methods 8, 405-407.

Niemeyer, B. A. and Schwarz, T. L. (2000). SNAP-24, a Drosophila SNAP-25homologue on granule membranes, is a putative mediator of secretion andgranule-granule fusion in salivary glands. J. Cell Sci. 113, 4055-4064.

Nijhout, H. F., Riddiford, L. M., Mirth, C., Shingleton, A. W., Suzuki, Y. andCallier, V. (2014). The developmental control of size in insects.Wiley Interdiscip.Rev. Dev. Biol. 3, 113-134.

10

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT

Page 11: Hobbit regulates intracellular trafficking to drive insulin-dependent … · hobbit regulates body size in a non-cell-autonomous manner As a first step toward understanding the function

Pfaffl, M. W., Horgan, G. W. and Dempfle, L. (2002). Relative expression softwaretool (REST) for group-wise comparison and statistical analysis of relativeexpression results in real-time PCR. Nucleic Acids Res. 30, e36.

Puig, O., Marr, M. T., Ruhf, M. L. and Tjian, R. (2003). Control of cell number byDrosophila FOXO: downstream and feedback regulation of the insulin receptorpathway. Genes Dev. 17, 2006-2020.

Rajan, A. and Perrimon, N. (2012). Drosophila Cytokine Unpaired 2 RegulatesPhysiological Homeostasis by Remotely Controlling Insulin Secretion. Cell 151,123-137.

Riedel, F., Gillingham, A. K., Rosa-Ferreira, C., Galindo, A. and Munro, S.(2016). An antibody toolkit for the study of membrane traffic in Drosophilamelanogaster. Biol. Open 84, bio.018937.

Rousso, T., Schejter, E. D. and Shilo, B.-Z. (2016). Orchestrated content releasefrom Drosophila glue-protein vesicles by a contractile actomyosin network. Nat.Cell Biol. 18, 181-190.

Rulifson, E. J., Kim, S. K. and Nusse, R. (2002). Ablation of insulin-producingneurons in flies: growth and diabetic phenotypes. Science 296, 1118-1120.

Sapiro, A. L., Ihry, R. J., Buhr, D. L., Konieczko, K. M., Ives, S. M., Engstrom,A. K., Wleklinski, N. P., Kopish, K. J. and Bashirullah, A. (2013). Rapidrecombination mapping for high-throughput genetic screens in Drosophila. G3(Bethesda) 3, 2313-2319.

Sievers, F., Wilm, A., Dineen, D., Gibson, T. J., Karplus, K., Li, W., Lopez, R.,McWilliam, H., Remmert, M., Soding, J. et al. (2011). Fast, scalable generationof high-quality protein multiple sequence alignments using Clustal Omega. Mol.Syst. Biol. 7, 539.

Song, W., Cheng, D., Hong, S., Sappe, B., Hu, Y., Wei, N., Zhu, C., O’Connor,M. B., Pissios, P. and Perrimon, N. (2017). Midgut-derived activin regulatesglucagon-like action in the fat body and glycemic control.Cell Metab. 25, 386-399.

Sudhof, T. C. (2002). Synaptotagmins: why so many? J. Biol. Chem. 277,7629-7632.

Topalidou, I., Cattin-Ortola, J., Pappas, A. L., Cooper, K., Merrihew, G. E.,MacCoss, M. J. and Ailion, M. (2016). The EARP complex and its interactorEIPR-1 are required for cargo sorting to dense-core vesicles. PLoS Genet. 12,e1006074.

Tran, D. T., Masedunskas, A., Weigert, R. and Hagen, K. G. and Ten, (2015).Arp2/3-mediated F-actin formation controls regulated exocytosis in vivo. Nat.Commun. 6, 10098.

Wang, L., Evans, J., Andrews, H. K., Beckstead, R. B., Thummel, C. S. andBashirullah, A. (2008). A genetic screen identifies new regulators of steroid-triggered programmed cell death in Drosophila. Genetics 180, 269-281.

Wang, J.-W., Beck, E. S. and McCabe, B. D. (2012). A modular toolset forrecombination transgenesis and neurogenetic analysis of Drosophila. PLoS ONE7, e42102.

Ward, T. H., Polishchuk, R. S., Caplan, S., Hirschberg, K. and Lippincott-Schwartz, J. (2001). Maintenance of Golgi structure and function depends on theintegrity of ER export. J. Cell Biol. 155, 557-570.

Waterhouse, A. M., Procter, J. B., Martin, D. M. A., Clamp, M. and Barton, G. J.(2009). Jalview Version 2-A multiple sequence alignment editor and analysisworkbench. Bioinformatics 25, 1189-1191.

11

RESEARCH ARTICLE Development (2018) 145, dev161356. doi:10.1242/dev.161356

DEVELO

PM

ENT