Abstract Interorgan signaling events are emerging as key regulators of behavioral plasticity. The foregut and hindgut circuits of the C. elegans enteric nervous system (ENS) control feeding and defecation behavior, respectively. Here, we show that epithelial cells in the midgut integrate feeding state information to control these behavioral outputs by releasing distinct neuropeptidergic signals. In favorable conditions, insulin and noninsulin peptides released from midgut epithelia activate foregut and hindgut enteric neurons, respectively, to sustain normal feeding and defecation behavior. During food scarcity, altered insulin signaling from sensory neurons activates the transcription factor DAF-16/FoxO in midgut epithelia, which blocks both peptidergic signaling axes to the ENS by transcriptionally shutting down the intestinal neuropeptide secretion machinery. Our findings demonstrate that midgut epithelial cells act as integrators relaying internal state information to distinct parts of the ENS to control animal behavior. __________________________________________________________________ Gut epithelium modulates output from distinct enteric circuits by altering secretion of insulin and noninsulin peptides. INTRODUCTION The enteric nervous system (ENS), also referred to as the “second brain,” is an autonomously acting neuronal network present in the digestive tract of all bilaterians, whose primary function is to regulate motility in the alimentary tract ([28]1–[29]4). The mammalian ENS consists of millions of neurons that fall into at least 20 neuron types that regulate gut peristalsis required for digestion, absorption, and egestion of food contents ([30]5–[31]8). Although enteric circuits are able to generate peristaltic behavior independent of central nervous system (CNS) input ([32]4), factors such as nutritional, reproductive, or psychological status of the animal have been shown to influence motility in the gut ([33]9–[34]11). Understanding how internal and external factors are sensed and integrated to coordinate the output from enteric circuits will provide novel insights on gastrointestinal physiology and disease. The ENS in Caenorhabditis elegans has been extensively studied in terms of its anatomy, development, molecular topology, and behavioral output ([35]12–[36]20), which makes it an attractive model for investigating how ENS function is regulated in metazoans. The C. elegans ENS consists of 22 neurons, 20 that are present in the foregut (the pharynx), from here on termed as pharyngeal enteric neurons (PENs), and two that innervate the hindgut, from here on termed as hindgut enteric neurons (HENs) ([37]Fig. 1A). The 20 PENs are further classified into 14 anatomically and molecularly distinct classes forming a densely interconnected network that controls rhythmic pharyngeal pumping ([38]14–[39]16). The two HENs, termed AVL and DVB, are GABAergic neurons that are activated by the neuropeptide NLP-40 released from the intestine, which generates all-or-none action potentials in these neurons to control the rate of defecation ([40]12, [41]18, [42]20, [43]21). The rhythmic activation of the HENs, which form an electrically coupled circuit, results in γ-aminobutyric acid–dependent contraction of enteric muscles leading to expulsion of contents of the gut, a behavior termed as the defecation motor program ([44]12, [45]22, [46]23). Fig. 1. Starvation inhibits ENS function via cell nonautonomous activity of DAF-16/FoxO in the intestine. [47]Fig. 1. [48]Open in a new tab (A) Schematic showing the anatomical locations of PENs and HENs in the alimentary tract of C. elegans; adapted from WormBook ([49]117). (B and C) Pharyngeal pumping rate and frequncy of expulsion on food in fed versus starved (4 hours) wild-type adults and in starvation-induced dauer stage animals. (D and E) Timing of pBoc and Exp events, and expulsion:pBoc ratio on food in fed vs starved (4 hours) wild-type adults. (F) Pharyngeal pumping rate on food in fed versus starved (4 hours) wild-type and daf-16(mu86) adults. (G) Intestine-specific DAF-16 depletion in daf-16(ot853); otSi2[ges-1p::TIR1(F79G)]; daf-2(e1370) dauer stage animals. Animals were treated with either solvent (ethanol) or 100 μM 5-Ph-IAA. DAF-16 was not detected in the midgut epithelial cells after 5-Ph-IAA treatment in 15 of 15 animals. Scale bars, 20 μm. (H) Pharyngeal pumping rate on food after intestine-specific DAF-16 depletion in fed versus starved (4 hours) daf-16(ot853); otSi2[ges-1p::TIR1(F79G)] adults. Animals were treated with either solvent (ethanol) or 100 μM 5-Ph-IAA. (I and J) Frequency of expulsion and expulsion:pBoc ratio on food after intestine-specific DAF-16 depletion in fed versus starved (4 hours) daf-16(ot853); otSi2[ges-1p::TIR1(F79G)] adults treated with 100 μM 5-Ph-IAA. Horizontal line in the middle of data points represents median value of biological replicates in (B), (C), (E), (F), and (H) to (J). Additional horizontal lines represent 25th and 75th percentiles in (B), (F), and (H). *P < 0.05, ***P < 0.001, ****P < 0.0001; ns, not significant. (B and C) Dunn’s multiple comparison test after Kruskal-Wallis test. (E, I and J) Mann-Whitney test. (F and H) Sidak’s multiple comparisons test after two-way analysis of variance (ANOVA). After completion of postembryonic development, C. elegans continuously feeds and defecates, and the rates of both of these behaviors remain highly consistent both within and across individuals ([50]23–[51]26). When an animal is transiently moved away from its food source, there is an immediate reduction in the frequency of foregut and hindgut contractions, which become highly irregular because the animal is no longer actively consuming food ([52]23–[53]25). ENS output in the presence of abundant food can also be altered if an animal has experienced starvation in the recent past ([54]25, [55]27, [56]28), and this effect is more pronounced in the dauer diapause stage, which the animal enters after experiencing acute nutrient deprivation during early larval development ([57]29–[58]31). It remains little understood how different neuronal types in the ENS sense the internal state of an animal to coordinate gut contractility at the systemic level during periods of altered nutritional requirement ([59]10, [60]32, [61]33). Here, we show how distinct neuropeptidergic signals are transmitted from midgut epithelial cells to communicate the internal state of the animal to the foregut and hindgut enteric circuits and show that this interorgan signaling is required to maintain normal ENS output in favorable conditions. Previous work has shown that the neuropeptide NLP-40 released from gut epithelia activates the HENs via G protein–coupled receptor signaling to generate normal defecation behavior ([62]20, [63]21). We find that two neuropeptides of the insulin family, INS-7 and INS-35, are also continuously secreted from midgut epithelial cells, and these peptides are sensed by the insulin receptor on neuromodulatory NSM neurons in the foregut to induce rapid foregut contractions in nutrient replete conditions. When an animal undergoes acute starvation, the absence of insulin signaling from sensory neurons activates the stress-responsive transcription factor DAF-16/FoxO in the midgut epithelial cells ([64]34, [65]35). Using transcriptomic profiling, cell type–specific genetic manipulations and CRISPR-Cas9–generated endogenous reporters, we show that DAF-16/FoxO activation in the midgut epithelia reduces the secretion of neuropeptides from the midgut by transcriptionally silencing not only insulin genes but also genes that constitute the neuropeptide secretion machinery of the midgut epithelial cells. Because of the absence of a neuropeptide relay signaling via the intestine to the two enteric circuits, ENS output is greatly reduced both in terms of foregut and hindgut contractions even after the animal returns to a favorable environment after experiencing starvation. Our findings highlight the role of the intestinal epithelia as an integrator of the metabolic status of the animal and unveil the mechanisms of how the non–cell autonomous activity of a stress-responsive transcription factor modifies enteric behaviors by disrupting gut-to-ENS signaling during periods of nutritional adversity. RESULTS Starvation inhibits output from the ENS via non–cell autonomous activity of DAF-16/FoxO in gut epithelial cells To study how the nutritional state of an organism alters output from the ENS, we measured how the two enteric behaviors in C. elegans, feeding and defecation, are affected by starvation. The rate of feeding, controlled by PENs, is measured as the frequency of foregut contractions, while the rate of defecation, controlled by HENs, is measured as the frequency of expulsion events from the hindgut. The defecation motor program consists of three steps: contraction of posterior body muscles (pBoc), contraction of anterior body muscles (aBoc), and expulsion (Exp) ([66]23). The pBoc step is initiated by a pacemaker in the intestinal epithelial cells ([67]36), which acts upstream of HENs, while the Exp step requires activation of the two HENs that constrict hindgut enteric muscles to release the contents of the gut through the anal opening ([68]20, [69]21). We measured how feeding and defecation behaviors were affected by starvation after the animal was transferred back to food and allowed an acclimation period of 5 min. We used two different starvation paradigms: (i) a 4-hour period of total food deprivation during adulthood that is sufficient to induce risk-taking behavior during food search ([70]37) and (ii) food deprivation combined with animal crowding during early development that induces remodeling into the dauer stage of diapause ([71]29, [72]30). We found that starvation during adulthood reduces the rate of feeding even after the animal is returned to food ([73]Fig. 1B and fig. S1A). In contrast, animals in the dauer stage display no foregut (pharynx) contractions on food, except for a few sporadic pharyngeal pumps, as described previously ([74]Fig. 1B) ([75]29, [76]31). We observe a similar effect of starvation on the rate of defecation. The frequency of pBoc events was unaffected after starvation (fig. S1B), but the rate of expulsions was significantly reduced ([77]Fig. 1, C and D). Half of the pBoc events in starved adults did not result in any expulsion of gut contents ([78]Fig. 1, C and D), which indicates altered communication between the intestinal pacemaker and the HENs. Although these animals underwent a period of food deprivation before the measurement of defecation behavior, their gut is filled with food contents in the 5 min of acclimation before behavioral recordings, which can be observed during the expulsion events. In contrast, dauer stage animals do not initiate any pBoc or Exp events when returned to food ([79]Fig. 1C and fig. S1B), which is likely attributable to the fact that these animals are not actively feeding because their buccal cavity is sealed ([80]38). Starvation during adulthood or in dauer-inducing conditions during postembryonic development both involve activation of the stress-responsive protein DAF-16 ([81]30, [82]39, [83]40), which is the sole C. elegans homolog of mammalian FoxO transcription factors. In nutrient replete conditions, agonist insulin peptides signal via DAF-2, the sole member of the insulin/insulin-like growth factor 1 (IGF-1) receptor (InsR) family in C. elegans, to trigger a phosphorylation cascade that retains DAF-16/FoxO in the cytoplasm in an inactive state (fig. S1C) ([84]35, [85]41–[86]43). When the environment becomes unfavorable, the absence of agonist insulins and presence of antagonistic insulins inhibit signaling via the DAF-2/InsR receptor, and this results in nuclear translocation of unphosphorylated DAF-16/FoxO, where it activates the transcription of stress-responsive genes (fig. S1C) ([87]35, [88]42, [89]43). We tested whether starvation affects enteric functions in animals that lack the DAF-16/FoxO transcription factor. We found that daf-16 null adults do not reduce their rate of feeding when placed on food after the same period of starvation that reduces foregut contractions in wild-type animals ([90]Fig. 1F). Similarly, in contrast to wild-type adults ([91]Fig. 1, C to E), a 4-hour starvation period did not significantly alter the rate of defecation in adults that lack daf-16 (fig. S1, D to F). A previous study has shown that in the dauer diapause stage, the removal of DAF-16/FoxO from the midgut epithelial cells can initiate foregut contractions in conditions that completely silence ENS output ([92]40), which suggests a non–cell autonomous role of this transcription factor in regulating ENS function. By visualizing an endogenously green fluorescent protein (GFP)–tagged DAF-16/FoxO protein, we found that this transcription factor localizes to the nuclei of midgut epithelial cells in starvation conditions that reduce the rates of feeding and defecation during adulthood (fig. S1G). To test whether removing DAF-16/FoxO only from the midgut epithelial cells can suppress the effects of starvation on enteric behaviors in adults, we performed spatially controlled depletion of DAF-16/FoxO protein using the AID2 technology, an improved version of the auxin-inducible degron (AID) system ([93]44–[94]46). These experiments were performed in a genetic background where daf-16 is endogenously tagged with AID* ([95]40), and tissue-specific heterologous expression of TIR1(F79G) was used for spatially controlled depletion of DAF-16/FoxO in the presence of 5-phenyl-indole-3-acetic acid (5-Ph-IAA), the synthetic auxin ligand for TIR1(F79G) ([96]44–[97]46). We find that AID2-mediated depletion of DAF-16/FoxO from all tissues using a ubiquitously expressed TIR1(F79G) transgene results in 99.8% of animals with the dauer-constitutive daf-2(e1370) mutation failing to arrest in the dauer diapause stage at 25°C, thus reconstituting the dauer-defective phenotype of daf-16 null alleles (fig. S1H) ([98]35, [99]42, [100]43). In contrast, if DAF-16/FoxO is depleted from only the midgut epithelial cells using a TIR1(F79G) transgene driven by the midgut-specific ges-1 promoter, then 98% of daf-2(e1370) animals arrest as SDS-resistant dauers at 25°C ([101]Fig. 1G and fig. S1I), which supports previous evidence that animals can attain dauer morphology even after depletion of endogenous DAF-16/FoxO protein from their intestine ([102]40). Using this intestine-specific AID2 system, we found that the effect of starvation on feeding rate can be completely suppressed if DAF-16/FoxO is selectively depleted from midgut epithelia of adult stage animals ([103]Fig. 1H and fig. S2A). Similarly, starvation does not affect defecation behavior in animals that have DAF-16/FoxO depleted only from their midgut epithelial cells ([104]Fig. 1, I and J, and fig. S2B). Unlike the effect of starvation on defecation behavior in wild-type animals ([105]Fig. 1, D and E), DAF-16/FoxO removal specifically from midgut epithelia resulted in the expulsion of gut contents immediately after most pBoc events even in starved animals ([106]Fig. 1J and fig. S2B). This establishes a non–cell autonomous role of DAF-16/FoxO to modify output from PENs and HENs to regulate the rates of feeding and defecation, respectively, after a period of nutrient deprivation during adulthood. Evidence for two signaling axes from midgut epithelial cells to the ENS Our previous work has shown that DAF-16/FoxO also functions cell autonomously in the PENs to silence foregut contractions during acute starvation ([107]40). This suggests that when food is abundant, insulin peptides released from midgut epithelial cells might act on the PENs via interorgan signaling to prevent DAF-16/FoxO activation and generate normal feeding behavior, while during food deprivation, DAF-16/FoxO activity in the midgut likely alters the release of these gut-derived insulin peptides to activate DAF-16/FoxO in PENs via “FoxO-to-FoxO” signaling ([108]47). To test whether enteric neurons actively need insulin signaling to maintain normal ENS output in well-fed conditions, we inhibited DAF-2/InsR specifically in all enteric neurons by expressing a dominant negative form of the DAF-2 receptor using a driver that is expressed in both PENs and HENs (fig. S3A) ([109]40, [110]48). This dominant negative form of the protein, termed DAF-2(DN), has the intracellular kinase domain replaced by a fluorescent protein (fig. S3A), which allows DAF-2(DN) to bind insulin peptides on the extracellular side but does not allow the signal to be transmitted to its intracellular downstream effectors. Insulin/IGF-1 receptors need to be phosphorylated by their dimerization partner to achieve activation ([111]49), but the lack of a kinase domain in DAF-2(DN) results in inactivation of the wild-type DAF-2/InsR molecule it dimerizes with and thus producing a dominant negative effect. We found that expression of DAF-2(DN) in all neurons of the ENS (HENs and PENs) is sufficient to reduce the animal’s rate of feeding even in nutrient-rich conditions, but it has no effect on defecation behavior ([112]Fig. 2A and fig. S3, B to E). In terms of change in the rate of feeding, the effect of inhibiting DAF-2/InsR signaling in PENs of well-fed animals is identical to the effect of starvation on wild-type animals ([113]Fig. 2A). When animals expressing DAF-2(DN) in PENs were subjected to starvation, it does not further reduce their rate of feeding ([114]Fig. 2A), which indicates that starvation in wild-type animals reduces feeding rate likely via inhibition of DAF-2/InsR signaling in PENs. In contrast, each pBoc event is almost always followed by an expulsion of gut contents in animals that express DAF-2(DN) in HENs (fig. S3, D and E), which suggests that the communication between the intestinal defecation pacemaker and the HENs does not require active insulin signaling in the hindgut-innervating neurons. Hence, we conclude that the control of PENs and HENs by DAF-16/FoxO activation in the midgut involves independent signaling axes. Fig. 2. DAF-16/FoxO silences transcription of several metabolic pathway genes in the gut upon acute starvation. [115]Fig. 2. [116]Open in a new tab (A) Pharyngeal pumping rate on food in fed versus starved (4 hours) wild-type and otIs913[pha-4prom2::daf-2(DN)] adults. Horizontal line in the middle of data points and additional horizontal lines represent median of biological replicates, and 25th and 75th percentiles, respectively. *P < 0.05, ***P < 0.001; ns, not significant. Sidak’s multiple comparisons test after two-way ANOVA. (B) Schematic of RNA-seq strategy to determine gene expression changes mediated by intestinal DAF-16 in dauer stage animals. (C and D) Principal components analysis (PCA) plot and heatmap of sample-to-sample distance matrix for all samples used in RNA-seq analysis. (E) Volcano plots for expression changes due to auxin alone or auxin in the presence of intestinal TIR1(F79G). Dashed vertical lines indicate the fold change cutoff of 1.5-fold. (F) Venn diagram for the number of differentially expressed genes for each of the four comparisons. (G) Heatmap of 1205 genes that are differentially expressed after depletion of intestinal DAF-16 in dauers. (H) Overlap of genes regulated by intestinal DAF-16 in dauers (this study) with all genes that undergo transcriptional changes during dauer formation as previously reported ([117]50). (I) Overlap of genes regulated by intestinal DAF-16 in dauers (this study) with known class I and class II transcriptional targets of DAF-16 as previously reported ([118]52). (J) Venn diagram showing overlap of genes regulated by intestinal DAF-16 in dauers (this study) with all genes that are differentially expressed in elt-2; elt-7 double mutant animals as previously reported ([119]53). Identification of DAF-16/FoxO–induced gene expression changes in the intestine during acute starvation We sought to identify these gut-to-ENS interorgan signaling axes by profiling DAF-16/FoxO–dependent gene expression changes in the intestine during nutrient deprivation. For this comparison, we chose animals arrested in the dauer diapause stage because: (i) the dauer stage is a more robust and stable state of nutritional stress response compared to starvation during adulthood ([120]39); (ii) the reduction in feeding rate is much stronger in the dauer stage, i.e., a complete silencing of feeding behavior ([121]Fig. 1B); and (iii) DAF-16/FoxO removal selectively from the dauer intestine using AID2 results in a strong reversal of the silenced state of the pharynx without affecting dauer morphology ([122]Fig. 1G, figs. S1I and S3, F and G). Because the AID2 system requires two components to achieve degradation of AID*-tagged DAF-16/FoxO, TIR1(F79G), and its synthetic auxin ligand 5-Ph-IAA, we included three controls in our analysis: only TIR1(F79G) but no auxin, only auxin but no TIR1(F79G), and neither TIR1(F79G) nor auxin ([123]Fig. 2B). Since the AID2 system has not yet been extensively used for whole transcriptome level comparisons in animals, the use of these controls allows to estimate the gene expression changes due to TIR1(F79G) alone or auxin alone ([124]Fig. 2B). We find that at the whole transcriptome level, TIR1(F79G) alone or auxin alone do not produce noticeable gene expression changes ([125]Fig. 2, C and D). All five biological replicates of the experimental condition cluster separately from all the replicates of the three control conditions ([126]Fig. 2, C and D). The synthetic auxin ligand, 5-Ph-IAA, alters gene expression only when acting together with TIR1(F79G), but not alone ([127]Fig. 2E). On estimation of the number of differentially expressed genes for each comparison, we found that auxin alone or TIR1(F79G) alone altered the expression of less than 50 genes, while when both of the AID2 components were acting together, the expression of ~1000 genes was affected ([128]Fig. 2F). To find the genes that are affected by the removal of DAF-16/FoxO from the intestine, we selected the 1205 transcripts that were differentially expressed in the experimental condition in comparison to either the no TIR1(F79G) control or the no auxin control ([129]Fig. 2, F and G). Among these 1205 genes, 933 were up-regulated and 272 were down-regulated ([130]Fig. 2G and data S1). We compared the genes identified in our transcriptomic analysis to another study that has profiled the transcriptomes of the dauer and predauer L2D stages ([131]50). In the L2D stage, the animal is uncommitted to dauer formation and can resume reproductive growth if conditions become favorable ([132]51). We found that genes that are up-regulated during dauer formation are significantly enriched among genes that are down-regulated after DAF-16/FoxO removal from the dauer intestine (152 of 272, P = 2.53 × 10^−69) ([133]Fig. 2H). Similarly, genes that are down-regulated during dauer formation are significantly enriched among genes that are up-regulated after the intestine-specific removal of DAF-16/FoxO (382 of 933, P = 5.46 × 10^−110) ([134]Fig. 2H), which suggests that depleting DAF-16/FoxO only from the intestine can reverse several of the gene expression changes associated with dauer formation. The transcriptional targets of DAF-16/FoxO are classified into class I and class II, with class I targets being up-regulated by this transcription factor, and class II targets being down-regulated ([135]52). We found that genes up-regulated after DAF-16/FoxO removal from the dauer intestine have a significant enrichment of class II DAF-16 targets (304 of 933, P = 4.53 × 10^−96) ([136]Fig. 2I). Similarly, genes that were down-regulated after DAF-16/FoxO removal from the dauer intestine have a significant enrichment of class I DAF-16 targets (114 of 272 = 8.86 × 10^−50) ([137]Fig. 2I), which demonstrates that a large proportion of the genes identified in our study are direct transcriptional targets of DAF-16/FoxO. We performed pathway enrichment analysis on genes that are up-regulated after gut-specific DAF-16/FoxO removal in dauers and found that these genes are involved in several metabolic functions of the gut, such as synthesis and processing of fatty acids and vitamins, and function of organelles such as lysosomes and peroxisomes (fig. S3H). To identify if DAF-16/FoxO cooperates with other transcription factors in the gut to alter the expression of genes related to intestinal function, we scanned the upstream promoter regions of these genes for enrichment of any known transcription factor binding sites. Genes that are up-regulated after DAF-16/FoxO depletion from the dauer intestine show an enrichment for the binding sites of GATA and nuclear hormone receptor (NHR) family transcription factors in their upstream sequences (data S2). Because the GATA transcription factors ELT-2 and ELT-7 and several members of the NHR family of transcription factors are required for the expression of metabolism-related genes in intestinal epithelial cells ([138]53–[139]56), we speculated that DAF-16/FoxO shuts down several functional processes of the gut during acute nutritional deprivation by antagonizing the ability of these gut-specific transcription factors to activate specific target genes. To test this, we compared the genes that we identified to be affected by intestine-specific DAF-16/FoxO depletion with genes that are differentially expressed in animals that lack ELT-2 and ELT-7, GATA transcription factors that contribute to the differentiation program of the C. elegans intestine ([140]Fig. 2J) ([141]53). We found that ~30% of genes (273 of 933, P = 8.23 × 10^−136) that are up-regulated after DAF-16/FoxO depletion from the dauer intestine are down-regulated in animals that lack both ELT-2 and ELT-7 ([142]Fig. 2J), suggesting that DAF-16/FoxO induces global down-regulation of several intestinal differentiation genes by antagonizing the activity of gut-specific GATA transcription factors during acute starvation. For genes that are down-regulated after DAF-16/FoxO depletion from the dauer intestine, their upstream sequences show the expected enrichment of the DAF-16 binding element (DBE) “GTAAACA,” which is similar to the binding sites of other forkhead box (FOX) family of transcription factors (data S2) ([143]52, [144]57). In addition, we see enrichment of “TGCACTT,” the binding site for DAF-12/VDR, which is an NHR family transcription factor that cooperates with DAF-16/FoxO to remodel tissues during dauer formation (data S2) ([145]40, [146]58). We also found enrichment of binding sites for GATA transcription factors in the upstream sequences of genes that are down-regulated after gut-specific DAF-16/FoxO depletion (data S2). A small but statistically significant proportion of these genes (25 of 272, P = 0.0045) are also down-regulated in animals that lack ELT-2 and ELT-7, and many of these genes (e.g., mtl-1, ftn-1, gst-38, hsp-12.3 and hsp-12.6) are involved in responding to stressful conditions ([147]Fig. 2J). This suggests that DAF-16/FoxO cooperates not only with another dauer-associated transcription factor DAF-12/VDR but possibly also with GATA transcription factors such as ELT-2 to up-regulate stress responsive genes in the intestine during acute nutritional deprivation, as previously shown in the context of life span regulation for some of the same genes (mtl-1 and hsp-12.6) ([148]59). DAF-16/FoxO shuts down the expression of the gut-derived insulin peptide INS-7 to silence foregut contractions during acute starvation Because the primary goal of our transcriptomic analysis was to identify changes in expression of secreted factors from the gut that act via the DAF-2/InsR receptor in PENs, we focused on the expression change of insulin family of genes after intestine-specific DAF-16/FoxO removal during starvation. The C. elegans genome encodes 39 members of the insulin family of peptides that show a complex and dynamic pattern of expression ([149]41, [150]60–[151]62). We found that the gut-specific removal of DAF-16/FoxO during acute starvation up-regulates transcript levels of ins-7 and ins-33 and down-regulates the transcript levels of ins-1, ins-18, and ins-24 ([152]Fig. 3A). Compared to the reported changes of insulin gene expression during dauer formation in the presence of DAF-16/FoxO ([153]50), we found that they are affected in the opposite direction, i.e., ins-1, ins-18, and ins-24 are up-regulated, while ins-7 and ins-33 are down-regulated in dauers vs predauer L2d ([154]Fig. 3B), which demonstrates that the expression change of these insulins during acute starvation is mediated via DAF-16/FoxO activity in the intestine. Fig. 3. Reversing the intestinal DAF-16/FoxO–induced changes in the expression of insulin peptide genes mildly reverses the silenced state of the dauer pharynx. [155]Fig. 3. [156]Open in a new tab (A) Change in expression levels of insulin family genes that are differentially expressed after intestine-specific DAF-16 depletion in dauers. (B) Change in expression levels of insulin family genes shown in (A) from the previously reported dauer versus L2d dataset ([157]50). (C) Expression of the endogenously tagged ins-1 reporter allele syb5452[ins-1::SL2::gfp::his-44] in fed adult and starvation-induced dauer stage animals. (D) Pharyngeal pumping rate on food in starvation-induced dauer stage animals with single or combination of null mutations in ins-1, ins-18, and ins-24 genes. (E) Pharyngeal pumping rate on food in starvation-induced dauer stage animals with defective neuropeptide secretion from the intestine, i.e., aex-1(sa9) and aex-5(sa23) mutants. (F) Expression of the endogenously tagged ins-7 reporter allele syb5424[ins-7::SL2::gfp::his-44] in fed adult and starvation-induced dauer stage animals. (G) Pharyngeal pumping rate on food in starvation-induced dauer stage animals that constitutively express ins-7 in the intestine (otEx8240 and otEx8241 lines). Horizontal line in the middle of data points represents median value of biological replicates in (D), (E), and (G). ***P < 0.001, ****P < 0.0001; ns, not significant. (D, E, and G) Dunn’s multiple comparison test after Kruskal-Wallis test. Scale bars, 20 μm (C and F). Because members of the insulin peptide family can act as agonists or antagonists of DAF-2/InsR receptor to produce a combinatorial effect on insulin/IGF-1 signaling at the systemic level ([158]35, [159]62–[160]64), there are two possible scenarios for how insulin peptides made in the gut can affect ENS output: (i) agonist insulins are released from the gut in favorable conditions that act on DAF-2/InsR in the PENs to prevent DAF-16/FoxO activation and generate normal feeding behavior in the presence of abundant food, and (ii) antagonist insulins are released from the gut in unfavorable conditions that inhibit DAF-2/InsR signaling in the PENs resulting in DAF-16/FoxO activation, which inhibits ENS output during periods of nutritional adversity (fig. S4A). DAF-16/FoxO activation in the intestine during starvation conditions is expected to shut down the expression of the agonist insulins and increase the expression of the antagonist insulins (fig. S4A). Because the expression of ins-1, ins-18, and ins-24 are up-regulated by intestinal DAF-16 during acute starvation ([161]Fig. 3, A and B), we speculated that active release of these insulins in the dauer stage might be required to completely silence foregut contractions. We generated CRISPR-based endogenous expression reporters for these three insulin genes, and unexpectedly, none of these are expressed in the intestine in well-fed or starved conditions ([162]Fig. 3C and fig. S4, B and C). All of these insulins are expressed in head neurons, while ins-1 and ins-18 are also expressed in tail neurons ([163]Fig. 3C and fig. S4, B and C), which suggests that intestinal DAF-16/FoxO activation in the gut epithelia non–cell autonomously increases the expression of ins-1, ins-18, and ins-24 in the nervous system during acute starvation. We generated CRISPR-based null alleles for all of these insulin genes and found that the deletion of ins-1, ins-18, and ins-24, either individually or in combination, do not reverse the silenced state of the dauer pharynx ([164]Fig. 3D), suggesting that up-regulation of these insulin peptides is not required to silence foregut contractions during acute starvation. We found that a large number of noninsulin neuropeptide genes are transcriptionally down-regulated when DAF-16/FoxO is depleted from the dauer intestine (fig. S4D). This is consistent with a previous study that showed that the entire family of FMRFamide-like neuropeptide (flp) genes is transcriptionally up-regulated during dauer formation ([165]50), which suggests that some of these neuropeptides might be involved in silencing ENS output during starvation. We tested whether blocking pathways related to neuropeptide processing or release affects the silenced state of the dauer pharynx. We found that disruption of unc-31/CAPS, a calcium-dependent activator protein required for the release of neuropeptides in dense-core vesicles ([166]65, [167]66), sbt-1/SCG5, a chaperone for proprotein convertases that cleave propeptides ([168]67), or trap-1/SSR1, an ER membrane protein required for biogenesis of antagonistic insulin peptides ([169]68), did not induce foregut contractions in dauer stage animals (fig. S4E). Because unc-31/CAPS, sbt-1/SCG5, and trap-1/SSR1 are involved in the processing and release of neuropeptides from the nervous system but the signal that silences ENS output is regulated by DAF-16/FoxO in the gut, we also manipulated the genes that are involved in secretion of neuropeptides from the intestinal epithelial cells. Disrupting the functions of aex-1/UNC13D or aex-5/PCSK5, constituents of the neuropeptide secretion machinery of the gut ([170]21, [171]69), did not generate contractions in the silenced dauer pharynx ([172]Fig. 3E), which indicates that active secretion of neuropeptides from the gut epithelial cells is likely not required to silence PEN activity during acute starvation. Next, we focused on ins-7 and ins-33, insulin genes that are down-regulated by intestinal DAF-16/FoxO in dauer stage animals ([173]Fig. 3, A and B). We generated CRISPR-based endogenous expression reporters for both these genes and found that ins-7 is only expressed in the anterior region of the midgut of well-fed adults, while ins-33 is expressed in hypodermal cells, but not in the gut ([174]Fig. 3F and fig. S4F). During acute starvation, the expression of both ins-7 and ins-33 are reduced in the respective tissues ([175]Fig. 3F and fig. S4F). We found that constitutive expression of ins-7 from gut epithelia produces a mild but significant reversal of the silenced dauer pharynx, while the constitutive expression of ins-33 produces no effect ([176]Fig. 3G and fig. S4G). This indicates that release of the INS-7 peptide from gut epithelia can induce foregut contractions in starvation conditions that completely silence ENS output. The role of INS-7 in regulating foregut contractions in nutrient replete conditions is described in a subsequent section. DAF-16/FoxO shuts down the neuropeptide secretion machinery of the gut during acute starvation In addition to altering the expression of neuropeptides, we found that the depletion of DAF-16/FoxO from the dauer intestine also up-regulates the expression of several genes that encode for components of the neuropeptide secretion machinery of the gut ([177]Fig. 4A). Among these genes, the strongest up-regulation is observed for aex-1/UNC13D, a SNARE regulator ([178]21, [179]69), while modest up-regulation is found for aex-4/SNAP25, a t-SNARE protein ([180]21), and aex-5/PCSK5, a proprotein convertase ([181]Fig. 4A) ([182]21, [183]69). All of these aex genes are significantly down-regulated during dauer formation ([184]Fig. 4B) ([185]50). We generated CRISPR-based endogenous reporters for these aex genes to profile their expression pattern in well-fed adults, starvation-induced dauers, and during the L3 larval stage, which is the developmental equivalent of the dauer stage in the reproductive phase of growth. For aex-1/UNC13D, we observe strong expression in all intestinal epithelial cells in adults and L3 larvae grown in nutrient replete conditions ([186]Fig. 4C). In well-fed adults, we also observe aex-1/UNC13D expression in uterine cells and dim expression in the hypodermis ([187]Fig. 4C). In dauers, we observe a strong reduction in aex-1/UNC13D expression in the intestinal epithelium ([188]Fig. 4, C and D). For both aex-4/SNAP25 and aex-5/PCSK5, we observe strong expression in the epithelial cells of the midgut during adulthood and L3 larval stage ([189]Fig. 4E and fig. S5A). In well-fed adults, aex-4/SNAP25 is also expressed in uterine cells, while aex-5/PCSK5 is also expressed in the body wall muscles of animals ([190]Fig. 4E and fig. S5A). In the dauer stage, both aex-4/SNAP25 and aex-5/PCSK5 undergo a strong reduction in expression in the intestinal epithelial cells ([191]Fig. 4, E and F, and fig. S5, A and B). To determine whether the reduction in expression of aex genes during starvation is mediated by DAF-16/FoxO, we quantified the expression of these genes in a daf-16 null background in both fed and starved conditions. We found that the expression of neuropeptide secretion machinery genes aex-1 and aex-4 and the gut-derived insulin peptide ins-7 is strongly down-regulated in midgut epithelial cells when adult animals were subjected to prolonged starvation, and this reduction in expression was suppressed in the absence of DAF-16/FoxO ([192]Fig. 4, G and H, and fig. S5C). In summary, acute starvation strongly down-regulates the expression of components of the intestinal neuropeptide secretion machinery in a DAF-16/FoxO–dependent manner. Fig. 4. DAF-16/FoxO shuts down the intestinal neuropeptide secretion machinery genes during prolonged starvation. [193]Fig. 4. [194]Open in a new tab (A) Change in expression levels of aex-1, aex-4, and aex-5 genes after intestine-specific DAF-16 depletion in dauers. (B) Change in expression levels of aex-1, aex-4, and aex-5 genes from the previously reported dauer versus L2d dataset ([195]50). (C and D) Expression of the endogenously tagged aex-1 reporter allele aex-1(ot1543[aex-1::SL2::gfp::his-44]) in fed adult and L3 larva and in starvation-induced dauer stage animals. Quantification of aex-1 expression in midgut epithelial cells is shown in (D). (E and F) Expression of the endogenously tagged aex-4 reporter allele aex-4(ot1530[aex-4::SL2::gfp::his-44]) in fed adult and L3 larva and in starvation-induced dauer stage animals. Quantification of aex-4 expression in midgut epithelial cells is shown in (F). (G) Expression of the endogenously tagged aex-1 reporter allele aex-1(ot1543[aex-1::SL2::gfp::his-44]) in the midgut epithelial cells of fed versus starved (24 hours) adults. (H) Expression of the endogenously tagged aex-4 reporter allele aex-4(ot1530[aex-4::SL2::gfp::his-44]) in the midgut epithelial cells of fed versus starved (24 hours) adults. Horizontal line in the middle of data points and additional horizontal lines represent median of biological replicates and 25th and 75th percentiles, respectively, in (D) and (F) to (H). *P < 0.05, ****P < 0.0001; ns, not significant. (D and F) Dunn’s multiple comparison test after Kruskal-Wallis test. (G and H) Sidak’s multiple comparisons test after two-way ANOVA. Scale bars, 20 μm (C and E) and 5 μm (G and H). From existing chromatin immunoprecipitation sequencing (ChIP-seq) datasets, we found that the upstream genomic regions of aex-1/4/5 genes contain DAF-16/FoxO binding peaks (fig. S6, A to C). These regions also contain binding peaks for ELT-2, a GATA transcription factor that is the master regulator of intestinal differentiation (fig. S6, A to C) ([196]54). This suggests that ELT-2 might be required for the expression of aex-1/4/5 genes for normal secretory functions of the intestinal epithelia in favorable conditions, while DAF-16/FoxO activation antagonizes the activity of this gut-specific GATA transcription factor to silence the expression of aex-1/4/5 genes during nutritional adversity. Because multiple components of the neuropeptide secretion machinery of the intestine are transcriptionally silenced during starvation, we generated a secretion reporter to directly test whether the release of neuropeptides from gut epithelia is hindered in adverse conditions. We generated a transgene with the insulin peptide INS-1 translationally fused to a TagRFP fluorescent protein and expressed it specifically in intestinal epithelial cells ([197]Fig. 5A). From the same construct, a GFP-tagged histone H2B protein is synthesized after an SL2 trans-splicing event, which reports whether transcription of the transgene is affected in any condition ([198]Fig. 5A). In well-fed adults and L3 larvae, the INS-1::TagRFP peptide is synthesized only in gut epithelia, but most of this protein is secreted from the basolateral side of these cells, and the tagged INS-1 peptide accumulates in the coelomocytes, macrophage-like scavenger cells in the body coelom ([199]Fig. 5A). In the dauer diapause stage, we observe that instead of being released from the basolateral side of gut epithelia, the INS-1::TagRFP peptide is now released into the lumen of the gut from the apical side of intestinal cells ([200]Fig. 5, A and B), a phenomenon that has been reported previously, but whose physiological relevance was unknown ([201]70). In addition, we observe strong accumulation of the tagged INS-1 peptide in the cytoplasm of gut epithelial cells in dauers ([202]Fig. 5, A and B), suggesting that the secretion of this neuropeptide is severely reduced during acute starvation conditions. We tested whether this neuropeptide secretion block in the dauer intestine is a direct consequence of DAF-16/FoxO activation. After gut-specific depletion of DAF-16/FoxO in dauers, we observe a robust increase in the secretion of INS-1::TagRFP from the basolateral side of intestinal cells with reliably stronger accumulation in the coelomocytes of the animal ([203]Fig. 5, C and D). These findings suggest that DAF-16/FoxO cell autonomously silences the secretion of neuropeptides from the intestine into the body coelom during acute starvation. Fig. 5. DAF-16/FoxO shuts down the release of insulin peptides from the midgut epithelium during prolonged starvation. [204]Fig. 5. [205]Open in a new tab (A and B) Secretion of INS-1::TagRFP from the intestine into coelomocytes in fed adult and L3 larva and in starvation-induced dauer stage animals expressing otEx8059[ges-1p::ins-1::tagRFP::SL2::gfp::his-44]. Quantification of INS-1::TagRFP intensity in gut lumen and gut cytoplasm and the ratio of INS-1::TagRFP intensity in coelomocytes: Gut cytoplasm is shown in (B). (C and D) Secretion of INS-1::TagRFP from the intestine into coelomocytes in daf-16(ot853); otSi2[ges-1p::TIR1(F79G)]; daf-2(e1370); otEx8220[ges-1p::ins-1::tagRFP] dauer stage animals after intestine-specific DAF-16 depletion. Animals were treated with either solvent (ethanol) or 100 μM 5-Ph-IAA. Ratio of INS-1::TagRFP intensity in coelomocytes: Gut lumen is shown in (D). (E and F) Secretion of INS-1::TagRFP from the intestine into coelomocytes in fed versus starved (24 hours) adults expressing otIs904[ges-1p::ins-1::tagRFP::SL2::gfp::his-44]. Ratio of INS-1::TagRFP intensity in coelomocytes: Gut cytoplasm is shown in (F). (G and H) Secretion of INS-1::TagRFP from the intestine into coelomocytes in fed wild-type or aex-1(ot1357) adults expressing otIs904[ges-1p::ins-1::tagRFP::SL2::gfp::his-44]. Ratio of INS-1::TagRFP intensity in coelomocytes: Gut cytoplasm is shown in (H). Horizontal line in the middle of data points and additional horizontal lines represent median of biological replicates and 25th and 75th percentiles, respectively in (B), (D), (F), and (H). *P < 0.05, ***P < 0.001, ****P < 0.0001; ns, not significant. (B) Dunn’s multiple comparison test after Kruskal-Wallis test. (D, F, and H) Mann-Whitney test. Scale bars, 20 μm (A, C, E, and G). Next, we tested whether neuropeptide secretion from gut epithelial cells is blocked only during the dauer diapause state or also when adult animals are subjected to acute starvation. We found that extended food deprivation in adults strongly reduces the basolateral secretion of the tagged INS-1 peptide from the gut, which is apparent from a much reduced INS-1::TagRFP accumulation in the coelomocytes ([206]Fig. 5, E and F). The lower insulin signal in the coelomocytes of starved adults is not due to reduced expression of this construct in the gut because the intensity of GFP-tagged H2B in the intestinal nuclei remains unchanged during starvation (fig. S7A). Since aex-1/UNC13D shows a strong DAF-16/FoxO–induced silencing during prolonged starvation ([207]Fig. 4G), which is associated with a neuropeptide secretion block from the intestine ([208]Fig. 5, E and F), we asked whether aex-1/UNC13D depletion also inhibits neuropeptide release from the gut of adults. We generated a CRISPR-based deletion of the entire aex-1/UNC13D protein coding region and found much reduced secretion of the tagged INS-1 peptide from the intestinal epithelial cells into the coelomocytes in aex-1/UNC13D null animals even in nutrient rich conditions ([209]Fig. 5, G and H). These results collectively demonstrate a role of aex-1/UNC13D in neuropeptide release from gut epithelia in well-fed conditions, and DAF-16/FoxO–mediated silencing of aex-1/UNC13D strongly reduces the secretion of neuropeptides from the gut during acute starvation conditions. Active secretion of INS-7 and INS-35 from gut epithelia is essential for normal foregut contractility in favorable conditions Because starvation-induced DAF-16/FoxO activity in the gut reduces intestinal neuropeptide secretion and the output from PENs in terms of foregut contractions, we speculated that active neuropeptide release from the gut is required to generate normal feeding behavior in favorable conditions. We found that animals with previously described loss-of-function mutations in the intestinal secretion machinery genes ([210]21, [211]69), aex-1/UNC13D and aex-5/PCSK5, have a significantly reduced rate of foregut contractions compared to wild type even in nutrient replete conditions ([212]Fig. 6A). This phenotype was also observed in animals with the CRISPR-generated null allele of aex-1/UNC13D ([213]Fig. 6B). Because aex-1/UNC13D and aex-5/PCSK5 mutants were initially identified in a genetic screen for animals with defective defecation behavior ([214]23), it can be postulated that these animals are feeding less because of their lower rate of expulsion from the hindgut. Although previous studies have observed no correlation between the rates of foregut and hindgut contractions ([215]71), we tested whether a defecation defective strain that does not affect neuropeptide secretion from the midgut can affect the rate of feeding. For this assay, we used a loss-of-function allele of nlp-40, a neuropeptide that is rhythmically released from the intestinal epithelia to activate the HENs and initiate expulsion of gut contents ([216]20). Unlike the aex-1/UNC13D null mutant, the depletion of nlp-40 does not affect the release of other neuropeptides from the intestinal cells, although both of these strains have similarly strong defects in expulsion behavior (fig. S7, B and C). We found that animals lacking NLP-40 show the same frequency of foregut contractions as wild-type animals (fig. S7D), indicating that inhibiting defecation behavior without disrupting the neuropeptide secretion machinery of the gut does not affect the output from PENs. Fig. 6. Active intestinal secretion of INS-7 and INS-35 is essential for normal pharyngeal pumping behavior in favorable conditions. [217]Fig. 6. [218]Open in a new tab (A) Pharyngeal pumping rate on food in fed wild-type, aex-1(sa9), and aex-5(sa23) adults. (B) Pharyngeal pumping rate on food in fed wild-type and aex-1(ot1357) adults. (C) DAF-16 depletion from enteric neurons of the foregut in daf-16(ot853); otIs908[pha-4prom2::TIR1(F79G)] adults. Animals were treated with either solvent (ethanol) or 100 μM 5-Ph-IAA. The anterior and posterior bulbs of the pharynx are shown. Nonpharyngeal neurons where otIs908 is expressed are labeled. (D) Pharyngeal pumping rate on food after DAF-16 depletion from enteric neurons of the foregut in fed daf-16(ot853); otIs908[pha-4prom2::TIR1(F79G)] adults in wild-type or aex-1(ot1464) background. Animals were treated with either solvent (ethanol) or 100 μM 5-Ph-IAA. (E) Expression of the endogenously tagged ins-35 reporter allele syb6236[ins-35::SL2::gfp::his-44] in fed adult and starvation-induced dauer stage animals. (F) Pharyngeal pumping rate on food in fed wild-type, ins-7(ot1427), ins-35(ot1443), and ins-7(ot1427); ins-35(ot1443) adults. (G) Pharyngeal pumping rate on food in starvation-induced dauer stage animals that constitutively coexpress ins-7 and ins-35 in the intestine (otEx8222, otEx8223, and otEx8224). Horizontal line in the middle of data points and additional horizontal lines represent median of biological replicates and 25th and 75th percentiles, respectively in (A), (B), (D), (F), and (G). *P < 0.05, **P < 0.01, ***P < 0.001, and ****P < 0.0001; ns, not significant. (A, F and G) Dunn’s multiple comparison test after Kruskal-Wallis test. (B) Mann-Whitney test. (D) Sidak’s multiple comparisons test after two-way ANOVA. Scale bars, 20 μm (C and E). Because the intestine-to-pharynx interorgan signal is likely an insulin peptide that acts via the DAF-2/InsR receptor to maintain DAF-16/FoxO in an inactive state in PENs to generate normal foregut contractions in nutrient-rich conditions ([219]Fig. 2A), we hypothesized that removing DAF-16/FoxO specifically from the ENS should reverse the reduced feeding rate in animals that have defective neuropeptide secretion from the midgut. To experimentally test this hypothesis, we expressed TIR1(F79G) specifically in the ENS of an animal that has its endogenous daf-16/FoxO locus tagged with AID* ([220]Fig. 6C). In the presence of the auxin ligand, AID*-tagged endogenous DAF-16/FoxO protein is specifically depleted from the ENS of the animal ([221]Fig. 6C). We found that ENS-specific DAF-16/FoxO depletion does not affect the rate of feeding in wild-type animals, but it significantly increases the rate of foregut contractions in the intestinal neuropeptide secretion defective aex-1/UNC13D null animals ([222]Fig. 6D). After depleting DAF-16/FoxO specifically from the ENS of aex-1/UNC13D null animals, their feeding rates become similar to that of wild-type animals grown in the abundance of food ([223]Fig. 6D). This increase in the rate of foregut contractions is not observed in aex-1/UNC13D null animals that are treated with the same dose of auxin in the absence of TIR1(F79G) (fig. S7E), ruling out the possibility that suppression of this phenotype is due to treatment with the synthetic auxin ligand. This suggests that a feeding defect due to the absence of neuropeptide secretion from midgut epithelial cells can be completely suppressed by the removal of a transcription factor from the PENs, indicating the presence of a midgut-to-ENS communication that is modulated during nutritional adversity. To identify the molecular nature of the interorgan signal from the gut, we interrogated a recent single-cell RNA sequencing (RNA-seq) dataset that captured gene expression from all major cell types in C. elegans ([224]72). We found that only two insulin genes, ins-7 and ins-35, are specifically enriched in the intestinal cluster (fig. S8A). We confirmed using a CRISPR-based endogenous expression reporter that ins-7 is only expressed in the anterior region of the midgut of well-fed adults, and its expression is strongly reduced during acute starvation ([225]Fig. 3F). We also generated a CRISPR-based reporter for ins-35 and found that this insulin peptide is only expressed in all of the midgut epithelial cells of the animal in well-fed conditions and in no other tissue type ([226]Fig. 6E). In the dauer diapause stage, ins-35 expression is restricted to the anterior cells of intestinal epithelia ([227]Fig. 6E), but at the functional level, INS-35 secretion from the basolateral side will be much reduced due to the neuropeptide secretion block in acute starvation conditions ([228]Fig. 5, A and B). To determine whether these intestinally produced insulin peptides constitute the gut-to-pharynx interorgan signal, we measured the rate of feeding in animals with loss-of-function mutations in the ins-7 and ins-35 genes. We found that the rate of foregut contractions in the abundance of food was significantly lower in both the ins-7 and ins-35 mutant strains in comparison to wild-type animals (fig. S8, B and C). However, because the ins-7 loss-of-function allele ok1573 also removes a part of the neighboring gene htas-1, and the ins-35 loss-of-function allele ok3297 only removes the last exon of ins-35, we generated CRISPR-based null alleles for both ins-7 and ins-35 by precisely deleting the entire coding region of these insulin peptide genes. We found that the rate of foregut contractions in nutrient replete conditions is significantly reduced in animals with a full locus deletion of either ins-7 or ins-35 ([229]Fig. 6F). If both these insulin peptides are removed simultaneously, then we do not observe an additive effect ([230]Fig. 6F). These results indicate that INS-7 and INS-35 released from the gut epithelial cells are required for normal feeding behavior in favorable conditions. We found that similar to ins-7 ([231]Fig. 3G), the constitutive expression of ins-35 in midgut epithelial cells results in a mild but significant reversal of the silenced state of the dauer pharynx (fig. S8D). However, the constitutive expression of ins-7 and ins-35 simultaneously in the intestine of dauers generates a much stronger output from the PENs, and we observe higher than 20 contractions per minute in the top quartile of these animals, which are still arrested in the dauer diapause stage ([232]Fig. 6G). Because the feeding behavior was recorded 5 min after these dauer stage animals were transferred to food, it could be argued that ectopic expression of ins-7 and ins-35 triggers a faster exit from the dauer stage when these animals encounter food. We consider this unlikely because: (i) these insulin peptides do not affect the dauer exit process ([233]63), and (ii) molecular and behavioral changes associated with dauer exit are observed at least an hour after the dauer stage animals encounter food ([234]73, [235]74). To rule out the possibility of a faster dauer exit in animals coexpressing ins-7 and ins-35 from the gut, we measured their feeding behavior in the daf-7(e1372) dauer-constitutive genetic background. In the presence of the daf-7(e1372) mutation, animals constitutively arrest in the dauer stage at 25°C independent of the presence of food ([236]75). We found that in daf-7(e1372) dauers, the constitutive coexpression of ins-7 and ins-35 from the intestine is sufficient to induce rapid foregut contractions (fig. S8E). These findings indicate that the gut-derived insulin peptides, INS-7 and INS-35, not only are required for normal feeding behavior in favorable environments but also are sufficient to induce foregut contractions in harsh environmental conditions that completely silence output from the PENs. The serotonergic NSM neurons sense insulin signals from the gut to generate normal feeding behavior We sought to identify the site of action of the gut-derived insulin peptides in the pharyngeal ENS. Previous studies have found that the bilaterally symmetric MC neuron pair is the pacemaker of the pharynx because their ablation strongly reduces the rate of foregut contractions ([237]24, [238]26). To test whether pacemaker activity of the MC neurons requires active DAF-2/InsR signaling, we generated a strain that expresses DAF-2(DN) only in the MC neurons in the entire ENS ([239]Fig. 7A). The inhibition of DAF-2/InsR in MC neurons did not affect the rate of foregut contractions ([240]Fig. 7B), suggesting that the MC neuron pair does not require insulin signals to generate normal feeding output in favorable conditions. Fig. 7. NSM senses INS-7 and INS-35 from the gut to enhance pharyngeal nervous system output during the fed state. [241]Fig. 7. [242]Open in a new tab (A) Expression of otIs937[ceh-19prom2::daf-2(DN)::eBFP2::SL2::tagRFP] in MC(L/R) neurons. (B) Pharyngeal pumping rate on food in fed wild-type and otIs937[ceh-19prom2::daf-2(DN)::eBFP2::SL2::tagRFP] adults. (C) Expression of otIs942[tph-1prom5::daf-2(DN)::eBFP2::SL2::tagRFP] in NSM(L/R) neurons. (D) Pharyngeal pumping rate on food in fed wild-type and otIs942[tph-1prom5::daf-2(DN)::eBFP2::SL2::tagRFP] adults. (E) Pharyngeal pumping rate on food in fed versus starved (4 hours) wild-type and otIs942[tph-1prom5::daf-2(DN)::eBFP2::SL2::tagRFP] adults. (F) Pharyngeal pumping rate on food in fed otIs942[tph-1prom5::daf-2(DN)::eBFP2::SL2::tagRFP] expressing adults in wild-type or ins-7(ot1427); ins-35(ot1443) background. (G) Pharyngeal pumping rate on food in fed versus starved (4 hours) wild-type and tph-1(ot1545) adults. (H) Pharyngeal pumping rate on food in starvation-induced dauer animals that constitutively coexpress ins-7 and ins-35 in the intestine (otEx8222) in wild-type or tph-1(ot1545) background. (I and J) DAF-16 protein in NSM neuron of fed versus starved (4 hours) adults and in starvation-induced dauer daf-16(ot853); otIs625[cat-1::SL2::mCherry::his-44] animals. Yellow asterisks show the nuclei of neighboring cells in (I). Nucleus-to-cytoplasm ratio of DAF-16 protein is shown in (J). (K and L) Secretion of NLP-40::TagRFP from the intestine into coelomocytes in fed versus starved (8 hours) adults expressing otIs927[ges-1p::nlp-40::tagRFP::SL2::gfp::his-44]. Horizontal line in the middle of data points represents median value of biological replicates in (B), (D to H), (J), and (L). Additional horizontal lines represent 25th and 75th percentiles in (B), (D to G), (J), and (L). *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001; ns, not significant. (B, D and L) Mann-Whitney test. (E to H) Sidak’s multiple comparisons test after two-way ANOVA. (J) Dunn’s multiple comparison test after Kruskal-Wallis test. Scale bar, 20 μm (A, C, and K) and 1 μm (I). Next, we targeted the serotonergic NSM neuron pair in the ENS, which has a known neuromodulatory role in increasing the rate of feeding in the presence of food ([243]76). To inhibit DAF-2/InsR signaling in the NSM neurons, we expressed DAF-2(DN) using a cis-regulatory element of the tph-1/TPH gene that drives expression exclusively in the NSM neurons in the entire nervous system ([244]Fig. 7C) ([245]77). We found that inhibition of DAF-2/InsR signaling only in the NSM neurons is sufficient to reduce the rate of feeding in nutrient replete conditions ([246]Fig. 7D). NSM-specific DAF-2(DN) expression also strongly suppresses the effect of starvation on the rate of feeding ([247]Fig. 7E), suggesting a role of insulin signaling in NSM neurons to mediate an internal state-dependent change in feeding behavior. Compared to wild-type animals, the reduction in feeding rate due to NSM-specific DAF-2/InsR inhibition is similar to that in animals that lack INS-7 and INS-35 ([248]Fig. 7F). To confirm that the gut derived insulins, INS-7 and INS-35, act via DAF-2/InsR in NSM to regulate the rate of foregut contractions in favorable environments, we performed an epistasis experiment. We found that inhibiting DAF-2/InsR in the NSM neurons of animals that lack INS-7 and INS-35 results in no further decrease in their rate of feeding in nutrient rich conditions ([249]Fig. 7F). These results indicate that the neuromodulatory NSM neurons in the ENS sense the insulin peptides released from the midgut epithelial cells to generate rapid foregut contractions when food is abundant. Since the NSM neurons increase the rate of foregut contractions in a serotonin (5-HT)–dependent manner ([250]78), we asked whether 5-HT plays a role in altering the rate of feeding in starved animals. Previous studies have shown that animals lacking tph-1/TPH, the tryptophan hydroxylase enzyme essential for neuronal 5-HT biosynthesis, are unable to increase foregut contractions when they encounter food ([251]79). Using a CRISPR-generated null allele of tph-1/TPH, we show that animals lacking the ability to synthesize 5-HT in neurons have a much lower rate of feeding than wild-type animals ([252]Fig. 7G). Unlike wild-type animals, starvation does not reduce the rate of foregut contractions in tph-1/TPH null animals ([253]Fig. 7G), which supports the role of serotonergic signaling in modulating feeding behavior in the context of nutritional status of the animal. We also found that constitutive coexpression of the insulin peptides ins-7 and ins-35 from the intestine cannot reverse the silenced state of the dauer pharynx in a tph-1/TPH null background ([254]Fig. 7H), which indicates that the gut-derived insulin peptides require neuronal serotonergic signaling to induce foregut contractions. Since inhibiting DAF-2/InsR signaling in the serotonergic NSM neurons mimics the effect of starvation on feeding behavior ([255]Fig. 7E), we asked whether reduced DAF-2/InsR signaling during starvation results in nuclear translocation of DAF-16/FoxO in the NSM neurons (fig. S1C). Using a strain where mCherry-tagged histone H2B marks the nuclei of all monoaminergic neurons, we show that endogenous DAF-16/FoxO protein translationally fused to mNeonGreen predominantly localizes to the cytoplasm of NSM neurons in well-fed adults ([256]Fig. 7, I and J). In contrast, DAF-16/FoxO translocates to the nucleus of NSM neurons in starved adults or in animals arrested in the dauer diapause state ([257]Fig. 7, I and J). These results collectively indicate that reduced insulin signaling during nutrient deprivation triggers DAF-16/FoxO activation in the serotonergic NSM neurons of the foregut enteric circuit to inhibit feeding behavior. Intestinal DAF-16/FoxO controls the second neuropeptidergic signaling axis to HENs Previous analysis has shown that the intestinally expressed AEX-1/4/5 peptidergic release machinery acts in the same pathway as NLP-40 to activate AEX-2, the NLP-40 receptor in HENs to control defecation behavior ([258]20, [259]21). The DAF-16/FoxO–dependent down-regulation of aex-1/4/5 under starvation conditions, described above, would therefore be expected to affect NLP-40 release from intestinal cells. To test this prediction, we generated a strain that expresses the NLP-40 neuropeptide fused with TagRFP specifically in the midgut epithelial cells. In well-fed adults, NLP-40::TagRFP synthesized in the intestine is secreted from the basolateral side of these cells into the body coelom, which is visualized by the accumulation of this neuropeptide in the coelomocytes ([260]Fig. 7K). The secretion of NLP-40 from the midgut epithelium is indeed strongly reduced after prolonged starvation in adults ([261]Fig. 7, K and L). Our findings collectively show that DAF-16/FoxO shuts down the release of both insulin and noninsulin peptidergic signals from the midgut epithelial cells to enteric neurons in the foregut and hindgut circuits during periods of nutrient scarcity. DISCUSSION How the internal state of an animal affects CNS functions has received extensive attention ([262]32, [263]80), but how internal states affect the ENS remains much less studied ([264]81). Here, we have shown that the epithelial cells in the C. elegans midgut coordinate the function of two branches of the ENS by controlling the release of distinct peptidergic signals ([265]Fig. 8). These two distinct signaling axes—insulins from the midgut to the PENs and NLP-40 from the midgut to HENs—are coordinated via DAF-16/FoxO transcriptionally regulating key nodes of the peptidergic secretion machinery ([266]Fig. 8). On the basis of available ChIP-seq data for DAF-16/FoxO and the master regulator of intestinal differentiation, ELT-2 (fig. S6, A to C), we propose that ELT-2 normally activates aex-1/4/5 and that this activation is antagonized by DAF-16/FoxO under food-deprived conditions. Fig. 8. Model for regulation of both feeding and defecation behavior by neuropeptides released from midgut epithelial cells in favorable or starved conditions. [267]Fig. 8. [268]Open in a new tab On sensing food-related cues in the environment, C. elegans chemosensory neurons release insulin peptides to modulate DAF-16/FoxO activity in the intestine ([269]35), which, as we have shown here, regulates the release of other insulins to control pharyngeal pumping and of a noninsulin neuropeptide to control enteric muscle contractions. We hypothesize that the relay of insulin signals from sensory neurons to midgut back to enteric neurons permits midgut epithelia cells to serve as an integrator of various types of nutritional signals both from the external environment and from internal nutritional signals from digested food within the alimentary tract. The nutrient sensing target of rapamycin complex 1 (TORC1) and TORC2 pathways have been shown to regulate DAF-16/FoxO activity specifically in the intestine ([270]82–[271]84), and the AAK-2/AMPK kinase directly phosphorylates DAF-16/FoxO to increase its activity during periods of extended starvation ([272]85). Reduced TORC2 signaling in the gut in adverse environments inhibits the expression of insulin peptides in chemosensory neurons, which suggests a role of the intestine as an amplifier of a starvation signal since it can further reduce the release of agonist insulin peptides from the nervous system via a feed-forward loop ([273]86). Hence, intestinal DAF-16/FoxO may integrate signals from multiple stress response pathways to modify nervous system function during periods of extended nutritional adversity. It has been well documented that peptides of the insulin family can cross the blood-brain barrier to affect CNS function in mammals ([274]87). Research in invertebrate models has shed light on the mechanistic bases by which the release of specific insulin peptides is actively regulated from non-neuronal tissues in the gut to modulate behaviors during different internal states of the animal. Neuropeptides released from intestinal epithelia have been shown to affect sensory behaviors such as thermotaxis and chemotaxis during prolonged starvation ([275]37, [276]88). Altered insulin signaling from the gut during periods of food scarcity can also modify secretion of neuropeptides from specific sensory neurons ([277]89). Our findings show that in addition to the transcriptional regulation of individual insulin genes during starvation conditions, a remodeling of the entire secretion machinery of the gut occurs during acute nutritional deprivation, indicating a broader role of the gut epithelia as a secretory organ that controls animal behavior. This expands on the previously described “FoxO-to-FoxO” signaling mechanism by demonstrating that intestinal DAF-16/FoxO modifies the function of other organs not only by orchestrating insulin-mediated cross-tissue FoxO activation ([278]47) but also via coordinating other interorgan signaling axes that do not directly involve insulin signaling. An internal state-induced global block of all gut-to-brain signaling during adverse environments would affect the release not only of neuropeptides but also of yolk proteins, metabolites, and other signaling molecules that are released from the gut in secretory vesicles to modulate several aspects of the animal’s physiology ([279]90). We found that the pacemaker neurons in the ENS of the C. elegans foregut are not receptive to the gut-to-brain insulin signaling axis, which agrees with the previously proposed autonomous nature of the MC neurons ([280]24, [281]26). Instead, the neuromodulatory NSM neurons receive insulin cues from the midgut epithelial cells to modify the output from the foregut ENS circuit depending on the nutritional status of the animal. We hypothesize that activation of DAF-16/FoxO in NSM during starvation conditions alters 5-HT release from these neurons. This is supported by a previous study that found an altered response of the NSM neurons to food stimuli in animals that have experienced extended food deprivation ([282]91). Altered 5-HT release from NSM neurons may modulate the rate of foregut contractions by acting on known serotonergic receptors that are expressed in both neurons and muscles of the foregut, as described previously ([283]76, [284]78, [285]92). For the noninsulin peptidergic signal NLP-40 from the midgut that regulates defecation behavior by activating HENs ([286]20), our results demonstrate a physiological regulation of the release of this neuropeptide during nutrient scarcity. A reduced rate of expulsion from the hindgut of a starved animal would increase the residence time of food in its intestinal lumen, which will likely support increased nutrient absorption, a phenomenon that has been observed when Drosophila is exposed to low nutrient conditions ([287]9). We observe different feeding responses in starved adults and dauer diapause arrested larvae on encountering food. This difference is likely due to the activity of an additional stress-responsive transcription factor, DAF-12/VDR, during dauer remodeling ([288]30), which cell autonomously shuts down pacemaker activity in the ENS of the foregut in dauers ([289]40). Reconstituting the gut-to-pharynx insulin signaling during the dauer diapause stage activates this silenced circuit likely via the release of 5-HT from NSM neurons ([290]Fig. 7H), which is supported by a previous study that induced foregut contractions in dauer stage animals by exposing them to exogenous 5-HT ([291]93). While the inhibitory effect on foregut contractions is unambiguous in the starvation-induced dauer diapause stage, the effect of prolonged starvation on the feeding rate of adults is dependent on the conditions in which the behavior is recorded. Initial studies reported an increase in foregut contractions in starved animals, but this behavior was observed in the absence of food ([292]27, [293]94). A subsequent study found an increase in the rate of feeding immediately after starved adults were returned to food ([294]28). This is consistent with strong calcium activity in the NSM neurons after a starved animal encounters food and this activity lasts for less than 5 min after food exposure ([295]91). These findings indicate a heightened sensitivity of the animal to food-derived cues after an extended period of starvation. Instead, if a starved animal is allowed to habituate on food for 5 min, then its rate of feeding is consistently lower than that in the fed state, which we have validated in blinded experiments (fig. S1A). Two recent studies that quantified feeding behavior over longer timescales, one using automated pharyngeal pumping measurements and another that quantified feeding rate independent of foregut contractions, have not observed an increase in the rate of feeding after starved animals are returned to food ([296]25, [297]95). This emphasizes the importance of separating the immediate response to the introduction of a new stimulus from the persistent effect of a change in the internal state of the animal in studies that measure behavioral output from autonomous circuits. The NSM neurons in the foregut enteric circuit can modulate behavioral outputs by altering overall brain state, which involves volumetric release of 5-HT to modify neural activity across multiple circuits via extra-synaptic communication ([298]78, [299]96). Axonal arborization patterns of serotonergic neurons in the mammalian brain suggest that 5-HT release from individual neuron types can potentially have widespread effects on brain activity ([300]97). In the context of enteric circuits, the role of neuronally released 5-HT in enhancing gut contractions is evolutionarily conserved in the Drosophila and mammalian gut ([301]98, [302]99), but how these regulatory mechanisms are influenced during different physiological states of the animal need further exploration. Altered 5-HT signaling has been linked to functional gastrointestinal disorders in several clinical studies, but the internal or external factors that lead to 5-HT dysregulation in these patients remain poorly understood ([303]100). Understanding how impairment of insulin signaling affects the function of serotonergic neurons in the ENS might shed new light on this unexplored aspect of gastrointestinal physiology. Limitations of the study In this study, we have used two types of starvation paradigms, (i) adult starvation on solid media and (ii) food deprivation and crowding during larval development that induces dauer formation. It is possible that additional pathways are involved in responding to other forms of starvation, such as bacterial dilution or food deprivation in liquid culture. We have elucidated the role of insulin and noninsulin peptidergic signaling between the gut epithelial cells and the enteric circuits in the foregut and hindgut of C. elegans, but it is possible that other forms of interorgan signaling involving small molecules (e.g., metabolites, short peptides, etc.) transported in secretory vesicles also contribute to communicating the starvation status of the animal from the intestinal epithelia to enteric neurons. Last, we have identified three components of the intestinal neuropeptide secretion machinery, the aex-1/4/5 genes, which are down-regulated by DAF-16/FoxO during starvation. It is likely that many other uncharacterized genes identified in our transcriptomic analysis are also essential components of the secretion machinery of the intestine that have not yet been functionally implicated in processing and release of signaling molecules from the gut. MATERIALS AND METHODS Maintenance of C. elegans strains All C. elegans strains used in this study and their genotypes are listed in data S3. All strains were maintained at 20°C on nematode growth medium (NGM) plates seeded with Escherichia coli (OP50 strain), unless otherwise specified. Strains with temperature-sensitive daf-2(e1370) and daf-7(e1372) alleles were maintained at 15°C. All strains were grown in optimum conditions in the abundance of food for at least three generations before using them in any experiment. All experiments in this study were performed using animals that were grown at 25°C for their entire developmental period (from unhatched embryos to adulthood) to be consistent with the temperature used for starvation-induced dauer formation assays. DNA constructs and microinjections All plasmids used in this study were generated via NEBuilder HiFi DNA assembly [New England Biolabs (NEB)] using the manufacturer’s protocol. All plasmid sequences were validated using Sanger (Azenta) and/or Oxford Nanopore (Plasmidsaurus) sequencing. The pha-4prom2 sequence in pSS35 [pha-4prom2::TIR1(F79G)::mTur2::tbb-2 3′UTR] and pSS36 [pha-4prom2::daf-2(DN)::ebfp2::tbb-2 3′UTR] is a 2393–base pair (bp) fragment upstream of isoform b of the pha-4 gene (−2387 to +6 position with respect to the start codon of pha-4 isoform b), which drives expression in all 22 enteric neurons (20 pharyngeal neurons and the two hindgut-innervating neurons, AVL and DVB) and also in RIS and PVT neurons ([304]48). The TIR1(F79G) sequence in pSS35 was generated by introducing the F79G [Phe(TTC) to Gly(GGA)] mutation in the Arabidopsis thaliana TIR1 sequence, as described previously ([305]45). The daf-2(DN)::ebfp2 sequence in pSS36 is from a previous study that replaced the tyrosine kinase and carboxyl-terminal domains of the cDNA of daf-2 isoform a (corresponding to amino acids 1246 to 1846) with the blue fluorescent protein ebfp2 sequence ([306]40). The tbb-2 3′UTR sequence is a 156-bp sequence immediately downstream of the stop codon of the tbb-2 gene. The ceh-19prom2 sequence of pSS43 [ceh-19prom2::daf-2(DN)::eBFP2::SL2::tagRFP-T::tbb-2 3′UTR] is a 1522-bp fragment upstream of isoform b of the ceh-19 gene (−1516 to +6 position with respect to the start codon of ceh-19 isoform b). The tph-1prom5 sequence of pSS44 [tph-1prom5::daf-2(DN)::eBFP2::SL2::tagRFP-T::tbb-2 3′UTR] is a previously described 232-bp fragment upstream of isoform a of the tph-1 gene (−232 to −1 position with respect to the start codon of tph-1 isoform a) ([307]77). The pSS29 [ges-1p::ins-1::tagRFP-T::SL2::gfp::his-44::tbb-2 3′UTR] and pSS34 [ges-1p::ins-1::tagRFP-T::SL2::ebfp2::his-44::tbb-2 3′UTR] constructs were generated by cloning the entire genomic locus (start codon to stop codon) of the ins-1 gene downstream of a 1999-bp ges-1 promoter fragment (−1999 to −1 position with respect to the start codon of ges-1). The pSS31 [ges-1p::nlp-40::tagRFP-T::SL2::gfp::his-44::tbb-2 3′UTR] construct was generated by cloning the cDNA of isoform a of the nlp-40 gene downstream of the ges-1 promoter fragment described above. pZW1 [ges-1p::ins-7::tbb-2 3′UTR], pZW3 [ges-1p::ins-33::tbb-2 3′UTR], and pZW24 [ges-1p::ins-35::tbb-2 3′UTR] were generated by cloning the entire genomic locus (start codon to stop codon) of ins-7 isoform b, ins-33, and ins-35, respectively, downstream of the ges-1 promoter fragment described above. Before microinjection, all plasmids were linearized using a single cutter restriction enzyme that cuts in the plasmid backbone. Linearized plasmids were injected at the concentrations listed in data S3 into both gonadal arms of young adult animals. F1 progeny was picked on the basis of the expression of the co-injection marker, and animals that transmitted the array to F2 progeny were used to generate transgenic lines. At least two independent lines (obtained from different injected P0 animals) for each injection type were used in subsequent experiments. CRISPR-Cas9 genome editing CRISPR-Cas9 genome editing was performed using a modified version of a previously described protocol ([308]101). All CRISPR injection mixes constituted of Streptococcus pyogenes Cas9 nuclease [250 ng/μl; Integrated DNA Technologies (IDT)], trans-activating CRISPR RNA (tracrRNA; 100 ng/μl; IDT), all crRNAs [(combined); 56 ng/μl; IDT], and each single-stranded oligodeoxynucleotide (ssODN; 100 ng/μl) repair template (IDT). First, tracrRNA and all crRNAs were incubated at 95°C for 5 min followed by at 10°C for 5 min. Cas9 was added to this mix, pipetted several times, and incubated at 25°C for 10 min. All ssODN and plasmids were subsequently added to this mix, and the volume was filled to 20 μl using nuclease-free water. The mixture was centrifuged at 17,900g for 2 min, and the supernatant was used for microinjections. For all CRISPR edits on chromosome II, the pRF4 [rol-6p::rol-6(su1006)] plasmid was co-injected at a concentration of 60 ng/μl. For CRISPR edits on all other chromosomes, dpy-10 co-CRISPR was performed using crRNA and ssODN listed in data S4. For all CRISPR edits, F1 animals were picked on the basis of roller phenotype and were genotyped by polymerase chain reaction (PCR) amplification of the edited locus, followed by Sanger (Azenta) or Oxford Nanopore (Plasmidsaurus) sequencing. For confirmed heterozygous edits, nonroller progeny was singled from F1 plates and genotyped by PCR to obtain homozygous lines. The genomic sequences for all genes edited in this study were obtained from WormBase ([309]102). The otSi2[ges-1p::TIR1(F79G)::mRuby::unc-54 3′UTR] single-copy transgene was generated using a previously described technique by introducing the F79G [Phe(TTC) to Gly(GGA)] mutation in the TIR1 sequence of ieSi61[ges-1p::TIR1::mRuby::unc-54 3′UTR] using a crRNA and an ssODN listed in data S4 ([310]45). The ins-18(ot1326), ins-18(ot1328), ins-1(ot1360), ins-1(ot1363), aex-1(ot1357), aex-1(ot1464), daf-16(ot1705), ins-7(ot1427), ins-35(ot1443), and tph-1(ot1545) alleles were obtained by deleting the entire coding sequence (start codon to stop codon) of the corresponding gene using two crRNAs and an ssODN listed in data S4. The otDf2 allele was generated using two crRNAs and an ssODN listed in data S4 to delete a 25.5-kb cluster of insulin genes, many of which (ins-24, ins-28, ins-29, and ins-30) are up-regulated in dauers ([311]50). The aex-1(ot1543[aex-1::SL2::gfp::his-44]), aex-4(ot1530[aex-4::SL2::gfp::his-44]), and aex-5(ot1532[aex-5::SL2::gfp::his-44]) alleles were generated using a crRNA (listed in data S4) and a long single-stranded repair template to insert the SL2::gfp::his-44 sequence immediately after the stop codon of the corresponding gene. For aex-1(ot1543[aex-1::SL2::gfp::his-44]), the SL2::gfp::his-44 sequence was inserted after the stop codon of isoform a of aex-1. The long single-stranded repair templates were generated using a previously described method ([312]101). The SL2::gfp::his-44 sequence was PCR amplified with primers containing 35-bp homology sequence corresponding to each end of the genomic locus to be edited. One of the two primers used for PCR amplification had a 5′ phosphate modification. The PCR product (10 μg) was digested with lambda exonuclease (NEB) to digest the DNA strand with 5′ phosphorylation. The unphosphorylated strand of DNA was purified using the Monarch PCR & DNA Cleanup Kit (NEB) and was used as a repair template at a final concentration of 100 ng/μl in the CRISPR injection mix. Suny Biotech generated the ins-1(syb5452[ins-1::SL2::gfp::his-44]), ins-18(syb5462[ins-18::SL2::gfp::his-44]), ins-24(syb5447[ins-24::SL2::gfp::his-44]), ins-7(syb5424[ins-7::SL2::gfp::his-44]), ins-33(syb5561[ins-33::SL2::gfp::his-44]), and ins-35(syb6236[ins-35::SL2::gfp::his-44]) alleles by inserting the SL2::gfp::his-44 sequence immediately after the stop codon of the corresponding gene. FLInt-mediated transgene integration The otIs904, otIs908, otIs911, otIs912, otIs913, otIs927, otIs937, and otIs942 transgenes were inserted at the intergenic oxTi553[eft-3p::tdTomato::H2B::unc-54 3′UTR + Cbr-unc-119(+)] V. locus using the fluorescent landmark interference (FLInt) method ([313]103). The FLInt injection mixes were prepared similar to CRISPR mixes, i.e., first, tracrRNA (final concentration of 100 ng/μl) and a crRNA (final concentration: 56 ng/μl) targeting the tdTomato locus (listed in data S4) were incubated at 95°C for 5 min, followed by at 10°C for 5 min. Streptococcus pyogenes Cas9 nuclease (final concentration: 250 ng/μl) was added to this mix, pipetted several times, and incubated at 25°C for 10 min. After this, all plasmids along with GeneRuler 1 kb plus DNA ladder (Thermo Fisher Scientific) were added to the mix at the concentrations listed in data S3. The volume was filled to 20 μl using nuclease-free water and centrifuged at 17,900g for 2 min, and the supernatant was used for microinjections. All injected animals were maintained at room temperature, and F1 progeny expressing the co-injection marker was singled onto plates. F2 progeny was singled from F1 plates that had at least 75% of the animals expressing the co-injection marker. F2 plates that transmitted the co-injection marker to 100% of the progeny were used to generate lines. 5-Ph-IAA treatment for AID2 For all AID2 experiments except the RNA-seq protocol, the animals were treated with the desired concentration of 5-Ph-IAA (BioAcademia) as described previously with some modifications ([314]104). A 100 μl of freshly prepared solution of 5-Ph-IAA in 25% ethanol (v/v) was added on top of 60-mm NGM plates seeded with OP50 bacteria. For control plates, an equal volume of 25% ethanol (v/v) was added such that the final ethanol concentration on the plate is 0.25% (v/v). All plates were stored in the dark for the entire duration of the experiment. Age-synchronized embryos obtained via the alkaline bleaching method were transferred to the plates 1 day after addition of 5-Ph-IAA, and the animals were grown at 25°C until they reached the desired developmental stage for the experiment. Dauer formation assays A previously described protocol was used to obtain starvation-induced dauers ([315]105). Eight gravid adult worms were transferred to NGM plates seeded with OP50 bacteria at 25°C. After 6 to 8 days when all animals on the plates were starved due to lack of bacterial food, animals were collected in 1% SDS and incubated for 20 min with gentle agitation to kill all non-dauer stages. Animals were centrifuges at 1150g for 2 min and washed twice with M9 buffer to remove all SDS. At this point, the tube contained living dauer stage animals and the carcasses of non-dauer stages. Animals were pipetted on a fresh NGM plate seeded with OP50 bacteria and live dauer animals were used for experiments within 40 min of removing the 1% SDS solution. Dauers for strains with the daf-2(e1370) or daf-7(e1372) mutation were generated by transferring age-synchronized embryos obtained via the alkaline bleaching method to NGM plates seeded with OP50 bacteria. Animals were maintained at 25°C for 3 days, after which all animals carrying the daf-2(e1370) or daf-7(e1372) mutation were arrested in the dauer stage. For dauer formation assays in the daf-2(e1370) genetic background after tissue-specific depletion of DAF-16 using AID2, age-synchronized embryos obtained via the alkaline bleaching method were maintained at 25°C for 3 days, after which the proportion of animals arrested in the dauer stage was estimated on the basis of dauer morphology and resistance to 1% SDS. The experiments were performed on at least three independent days. Adult starvation assays To starve adult animals, age-synchronized embryos obtained via the alkaline bleaching method were transferred to NGM plates seeded with OP50 bacteria at 25°C. After 48 hours, adult stage animals were washed off the plate using M9 buffer and centrifuged at 300g for 1 min. Animals were then washed three times with M9 buffer to get rid of all bacteria. For the first two washes, the supernatant was removed after centrifugation at 300g for 1 min. For the final wash, animals were allowed to settle down by gravity for 5 min after which the supernatant was removed, and animals were pipetted on unseeded NGM plates for the desired amount of time. At the end of the starvation period at 25°C, animals were transferred to NGM plates seeded with OP50 bacteria and were used for experiments within 15 min of transferring them on food. Pharyngeal pumping assays All pharyngeal pumping assays were performed on freely moving animals. The required number of animals was transferred to an NGM plate seeded with a uniform thin layer of OP50 bacteria. Animals were allowed to settle down for 5 min, and the movement of the grinder of the pharynx was recorded using a hand-held tally counter by observing the animals under a Nikon Eclipse E400 upright microscope equipped with differential interference contrast (DIC) optics. For age-synchronized adults, the number of grinder movements in a 20-s period was recorded using a 20× air objective lens of the Nikon Eclipse E400 microscope and was multiplied by three to obtain pharyngeal pumps per minute. For dauer stage animals, the number of grinder movements in a 1-min period was recorded using a 50× air objective lens of the Nikon Eclipse E400 microscope. The rate of pharyngeal pumping was recorded from at least eight animals on each day and at least on two independent days. Defecation assays All defecation assays were performed on freely moving animals. The required number of animals was transferred to an NGM plate seeded with a uniform thin layer of OP50 bacteria. Animals were allowed to settle down for 5 min, after which the posterior end of animals is observed using a 20× air objective lens of a Nikon Eclipse E400 upright microscope equipped with DIC optics. Each animal was observed for a 6 min period, during which the timings of pBoc and Exp events (that release food contents from the anal opening) were recorded using the Behavioral Observation Research Interactive Software software version 8.23 ([316]106). Defecation behavior was recorded from at least five animals on each day and at least on two independent days. RNA isolation and library preparation OH14654: daf-16(ot853[daf-16::mNG::AID]); daf-2(e1370) and OH17582: daf-16(ot853[daf-16::mNG::AID]); otSi2[ges-1p::TIR1(F79G)::mRuby::unc-54 3’ UTR]; daf-2(e1370) strains were used for this experiment. Before growing animals, 250 μl of a freshly prepared solution of 5-Ph-IAA in 25% ethanol (v/v) was uniformly spread on top of each unseeded 100-mm NGM plate to obtain a final 5-Ph-IAA concentration of 100 μM on the plates. For control plates, an equal volume of 25% ethanol (v/v) was added such that the final ethanol concentration on the plate is 0.2% (v/v). All plates were stored in the dark for the entire duration of the experiment. After 1 day of adding 5-Ph-IAA to the plates, a uniform thin layer of OP50 bacteria was spread on top of each plate and incubated at 25°C for an additional day. Subsequently, ~8000 age-synchronized embryos obtained via the alkaline bleaching method were transferred to each plate and on three plates per condition. All plates were maintained at 25°C for 3 days after which all animals were arrested in the dauer stage due to presence of the daf-2(e1370) mutation. Any animals that were not arrested in the dauer stage (less than 10 animals per plate) were picked and removed using a worm pick. A total of ~24,000 animals from three 100-mm NGM plates for each condition, which constituted a biological replicate, were collected in M9 buffer. Animals were washed three times using M9 buffer with centrifugation at 1300g for 2 min after each wash to remove all bacteria. A 200 μl of TRIzol reagent (Thermo Fisher Scientific) was added to the worm pellet, and samples were homogenized using a Kontes Pellet Pestle Motor (Thermo Fisher Scientific) for 30 s. Samples were immediately frozen on dry ice for 5 min, and the homogenization-freezing cycles were repeated for a total of five times. All homogenized samples were stored at −80°C. After five biological replicates were collected for each condition, total RNA was isolated from all samples using the RNeasy Micro Kit (QIAGEN) using the manufacturer’s protocol. A 100 ng of total RNA per sample was used for library preparation using the Universal RNA-seq with NuQuant kit (Tecan Genomics) using the manufacturer’s protocol with custom AnyDeplete (IC0149S) to deplete ribosomal RNA. All 20 libraries (4 conditions × 5 biological replicates) were pooled at a final concentration of 2 nM and were multiplexed for 75-bp single-end sequencing on a NextSeq 550 system (Illumina). RNA-seq data analysis and visualization All sequencing analyses were performed on the Galaxy web platform ([317]107). The quality control of raw RNA-seq reads was performed using FastQC. Adapter sequences were trimmed using Trimmomatic ([318]108). The trimmed reads were aligned to the C. elegans genome (WS284 release) using STAR ([319]109). Read counts were assigned to genes using featureCounts ([320]110). The differential expression of genes across the four conditions was determined using DESeq2 ([321]111). The principal components analysis (PCA) and sample-to-sample distance matrix plots were generated using DESeq2. The heatmap for differentially expressed genes was generated on the Galaxy web platform using the heatmap.2 function from the ggplot2 package on R. The overlap between different gene lists was identified using Venny 2.1.0. Proportioned Venn diagrams were generated using Venn Diagram Plotter (Pacific Northwest National Laboratory). Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis and enrichment of transcription factor binding motifs in upstream sequences of differentially expressed genes were performed using ShinyGO ([322]112–[323]114). ChIP-seq peaks for DAF-16 and ELT-2 were visualized using JBrowse2 on Wormbase ([324]102). Microscopy and image quantification Animals were mounted on a 5% agarose pad on top of a glass slide and were anesthetized using 50 mM sodium azide. Strains expressing daf-16(ot971[daf-16::gfp]) or daf-16(ot853[daf-16::mNG::AID]) were anesthetized using 4 mM levamisole (instead of sodium azide). Animals expressing daf-16(ot853), daf-16(ot971), otEx8059, otIs937, and otIs942 were imaged on an inverted Zeiss LSM 980 laser scanning confocal microscope using a 40× water immersion objective lens. All other strains were imaged on a Zeiss Imager Z2 upright microscope equipped with Colibri 7 light-emitting diodes (Zeiss) using a 40× oil immersion objective lens. A full z-stack was acquired for all animals to cover their entire body width. All images were processed and quantified on the Fiji platform on ImageJ ([325]115). The mean background subtracted signal intensity from the two coelomocytes in the anterior region of the body was quantified for INS-1::TagRFP and NLP-40::TagRFP secretion assays. For the same assays, the signal intensity from the intestinal lumen and intestinal cytoplasm were calculated as the mean background subtracted intensity from two different regions of the intestinal lumen or cytoplasm, respectively. For quantification of signal intensity from the intestinal nuclei, the mean background subtracted intensity was calculated from two intestinal nuclei, one each from the first two rows of the intestinal cells. Statistical analysis Data points for all biological replicates and sample sizes (N) are displayed in the figures. Statistical tests used in each figure and the corresponding P values are listed in the figure legends. P < 0.05 was considered to be statistically significant. To calculate the P value for overlap between two gene lists, hypergeometric tests were performed on R. All other statistical tests and plotting were performed on GraphPad Prism 10. For each type of quantitative assay described in this study, sample sizes were estimated by performing a pilot experiment using eight animals per condition to determine the mean and SD. Using these values, the required sample size was estimated using G*Power for a type I error rate of 0.05 and a statistical power of 0.95 ([326]116). If the estimated sample size per condition was less than 35, the experiment was performed again using the estimated total number of animals. For all subsequent assays of the same type, the same sample size per condition was used. Acknowledgments