• Keine Ergebnisse gefunden

A Command Chemical Triggers an Innate Behavior by Sequential Activation of Multiple Peptidergic Ensembles

N/A
N/A
Protected

Academic year: 2022

Aktie "A Command Chemical Triggers an Innate Behavior by Sequential Activation of Multiple Peptidergic Ensembles"

Copied!
13
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

A Command Chemical Triggers an Innate Behavior by Sequential Activation

of Multiple Peptidergic Ensembles

Young-Joon Kim,1Dusˇan Zˇitnˇan,2,3 C. Giovanni Galizia,1,4Kook-Ho Cho,1 and Michael E. Adams1,2,*

1Department of Entomology

2Department of Cell Biology & Neuroscience 5429 Boyce Hall

University of California, Riverside Riverside, California 92521

3Institute of Zoology Slovak Academy of Sciences Du´bravska´ cesta 9

84206 Bratislava Slovakia

Summary

Background:At the end of each molt, insects shed their old cuticle by performing the ecdysis sequence, an innate behavior consisting of three steps: pre-ecdysis, ecdysis, and postecdysis. Blood-borne ecdysis-trigger- ing hormone (ETH) activates the behavioral sequence through direct actions on the central nervous system.

Results:To elucidate neural substrates underlying the ecdysis sequence, we identified neurons expressing ETH receptors (ETHRs) inDrosophila. Distinct ensem- bles of ETHR neurons express numerous neuropeptides including kinin, FMRFamides, eclosion hormone (EH), crustacean cardioactive peptide (CCAP), myoinhibitory peptides (MIP), and bursicon. Real-time imaging of in- tracellular calcium dynamics revealed sequential activa- tion of these ensembles after ETH action. Specifically, FMRFamide neurons are activated during pre-ecdysis;

EH, CCAP, and CCAP/MIP neurons are active prior to and during ecdysis; and activity of CCAP/MIP/bursicon neurons coincides with postecdysis. Targeted ablation of specific ETHR ensembles produces behavioral defi- cits consistent with their proposed roles in the behav- ioral sequence.

Conclusions: Our findings offer novel insights into how a command chemical orchestrates an innate be- havior by stepwise recruitment of central peptidergic ensembles.

Introduction

Elucidation of mechanisms underlying assembly, orga- nization, and performance of behaviors is central to our understanding of brain function. Instinctive or cen- trally coordinated innate behaviors, also known as fixed action patterns, provide favorable models for behavioral analysis at the molecular and cellular levels. How such animal behaviors are organized and initiated were ques- tions brought into clearer focus over the past 60 years by

ethologists [1], who defined the concepts of ‘‘innate releasing mechanisms’’ and the chaining of central net- works that together organize sequential behaviors.

These concepts are now evolving rapidly with the advent of molecular biology and the use of transgenic animals as models[2, 3]. We are interested in developing behavioral paradigms amenable to analysis with molec- ular genetics, neural imaging, and biological chemistry.

One behavioral model worthy of attention is insect ecdysis.

Insects pass through multiple developmental stages during their life history, and each transition requires molting and ecdysis, whereby a new exoskeleton (cuti- cle) is produced and the old is shed. Removal of the old cuticle is achieved by performance of a fixed action pattern known as the ecdysis behavioral sequence. The sequence is driven by blood-borne ecdysis-triggering hormones (ETHs) released from peripherally distributed endocrine Inka cells [4–8]. In Drosophila, genetic null mutant flies (eth2) carrying a microdeletion at theeth locus show behavioral deficits and lethality at the first ecdysis; injection of ETH rescuesethmutant animals [7]. This suggests that peptides play important roles in orchestrating animal behavior.

Recent identification of theDrosophilaETH receptor gene[9, 10]provided a basis for the present analysis of central nervous system (CNS) mechanisms involved in initiation and regulation of the ecdysis behavioral se- quence. In the present work, we identify primary ETH target neurons in the CNS and find that they are orga- nized as peptidergic ensembles producing kinin, FMRFamides, eclosion hormone (EH), crustacean cardi- oactive peptide (CCAP), myoinhibitory peptides (MIPs), and bursicon. Using real-time calcium imaging, we show that activation of each peptidergic ensemble coin- cides with initiation and performance of specific behav- ioral subunits. Phenotypes of knockout flies bearing ablations of specific classes of ETH receptor (ETHR) neurons verify their roles in regulation of ecdysis and postecdysis behaviors. Our study demonstrates that ETH directly and sequentially activates multiple pepti- dergic networks in the CNS, recruiting consecutive phases of the ecdysis behavioral sequence. To our knowledge, this is the first description of an innate be- havior programmed by multiple peptidergic ensembles within the CNS.

Results

Analysis of the Pupal Ecdysis Behavioral Sequence InDrosophila, pupal ecdysis is preceded by pupariation, whereby the prepupa contracts its body into a barrel shape to form the puparium composed of the old larval cuticle. The underlying new pupal cuticle then separates from the puparium during pupal ecdysisw12 hr later.

The stereotypic nature of pupal ecdysis and reliable de- velopmental markers make it a favorable model for the behavioral analysis and neural imaging.

*Correspondence:michael.adams@ucr.edu

4Present address: Lehrstuhl fu¨r Neurobiologie, Universita¨t Kon- stanz, D-78457 Konstanz, Germany.

Konstanzer Online-Publikations-System (KOPS) - URN: http://nbn-resolving.de/urn:nbn:de:bsz:352-opus-121623 URL: http://kops.ub.uni-konstanz.de/volltexte/2010/12162

(2)

Pupal ecdysis consists of three centrally patterned behavioral subunits performed sequentially: pre-ecdy- sis (w10 min), ecdysis (w5 min), and postecdysis (w60–70 min) (Figure 1A)[11]. We examined the behav- ioral sequence through the semitransparent puparium (‘‘puparium-intact’’) as described previously [11, 12], but found that the puparium obscures and places con- straints on some movements. This made it particularly difficult to discriminate differences in abdominal swing- ing movements during ecdysis and postecdysis (see below). Therefore, we utilized a ‘‘puparium-free’’ prepa- ration by surgically removing the puparium immediately after pre-ecdysis onset[13]. The improved visibility and room for movement in this preparation allowed for a more complete analysis of natural and ETH-induced behavior. The following description of the natural pupal ecdysis sequence resulted from comparison of behav- iors observed in both puparium-intact and puparium- free prepupae (Figure 1A).

Pre-ecdysis

About 5 min after in vivo ETH release (Figure 1B), pre- ecdysis commences with the abrupt appearance of an air bubble at the posterior end of the prepupa (time zero). Pre-ecdysis involves anteriorly directed rolling contractions along the lateral edges of the abdomen, alternating on the left and right sides of the animal (Fig- ure 1A, Pre-ecdysis;Movie S1in theSupplemental Data available online). These contractions move the air bub- ble anteriorly to separate pupal cuticle from the pupar- ium [12]. This behavior is completed within w10 min and is followed by ecdysis behavior (Figure 1A). Pre- ecdysis behavior is the same in puparium-intact and puparium-free animals.

Ecdysis

In higher Diptera including Drosophila, the incipient adult head develops within the prepupal thorax. During pupal ecdysis, head eversion results from lateral swing- ing movements of the abdomen occurring along with Figure 1. Analysis of theDrosophila Pupal Ecdysis Sequence and Its Relationship to ETH Release

(A) Timelines and detailed depiction of se- quentially executed pre-ecdysis, ecdysis, and postecdysis behaviors. All time points are relative to the onset of pre-ecdysis. Ecdy- sis behaviors were observed in puparium- free and puparium-intact preparations (see Resultsfor details). The color-coded bars in- dicate each behavioral phase. Pre-ecdysis is characterized by anteriorly directed rolling contractions alternating along lateral edges of the abdomen. Ecdysis swinging move- ments occur along with anteriorly directed abdominal peristaltic contractions. Head eversion occurs shortly after the onset of ec- dysis (yellow arrow). Postecdysis is charac- terized by two behavioral subroutines: poste- riorly directed abdominal swinging and stretch-compression movements. Arrows in- dicate sites of abdominal contractions.

(B) Timing of ETH release during pupal ecdy- sis. ETH release was detected as a loss of EGFP fluorescence in Inka cells expressing chimeric ETH-EGFP (seeResults). The time course of ETH-EGFP depletion is shown in the left panel; each horizontal bar indicates the time window of fluorescence reduction in each cell. The right panel shows real-time traces of ETH-EGFP fluorescence from three representative Inka cells, which are color coded in the left panel (red, green, blue colors). In both panels, arrows and arrow- heads indicate onset and completion of fluo- rescence reduction, respectively. The gray dotted lines indicate pre-ecdysis onset (time zero). Note that a small surge of fluorescence near time zero is an artifact caused by internal movements associated with onset of pre- ecdysis.

(C) All Inka cells are depleted of ETH by the end of ecdysis sequence. Inka cells were dis- sected and examined before ecdysis (2–3 hr before behavioral onset; n = 29 from five pre- pupae) or after ecdysis (30–50 min after be- havioral onset; n = 34 from seven pupae).

ETH release was detected as loss of fluores- cence as described in (B).

(3)

anteriorly directed peristaltic contractions (Figure 1A;

Movie S1). In puparium-intact preparations, head ever- sion occursw1 min (1.160.08 min, n = 5) after the onset of ecdysis swinging and is completed withinw5 s. After completion of head eversion, ecdysis contractions con- tinue forw15 min, facilitating expansion of wing pads and legs to their final size. The frequency of ecdysis swinging (w5 swings/min) decreases markedly after head eversion (w2 swings/min; n = 5). In puparium- free animals (Figure 1A), head eversion occurs sooner (0.360.52 min after ecdysis onset, n = 7), and the dura- tion of ecdysis behavior lasts onlyw5 min. Later in ecdy- sis, anteriorly directed swinging contractions are often interrupted by posteriorly directed ones, indicating a transition to postecdysis.

Postecdysis

This behavior consists of two behavioral subroutines:

postecdysis swinging and stretch-compression move- ments of the abdomen (Figure 1A). Postecdysis swing- ing occurs along with posteriorly directed peristaltic contractions and alternates with longitudinal move- ments of the abdomen, referred to as ‘‘stretch-compres- sion.’’ The frequency and intensity of postecdysis con- tractions wane gradually until they are detected mainly in the anterior part of the abdomen; they cease w100 min after pre-ecdysis onset. Postecdysis behav- ior concludes with compression of the pupa at the pos- terior end of the puparium.

ETH Release Coincides with Initiation of the Ecdysis Sequence

To confirm the role of ETH in initiation of the pupal ecdy- sis sequence, we monitored its release from endocrine Inka cells in vivo by using time-lapse EGFP fluorescence imaging in pharate pupae (prepupae) carrying the chimeric transgene 2eth3-egfp. In this transgenic fly, EGFP is expressed as part of a fusion protein with the ETH propeptide precursor, and loss of EGFP fluores- cence indicates ETH release[7]. Because pharate pupae generally are immobile and Inka cells are located imme- diately below the semitransparent puparium, in situ imaging of Inka cell in intact pharate pupae is feasible.

We monitored two to three Inka cells simultaneously in each experiment.

Depletion of ETH-EGFP occurs in about 50% of mon- itored Inka cells (n = 13/24) shortly before pre-ecdysis onset (time zero). The time course of secretory activity for each Inka cell was variable (Figure 1B). The mean value (6 standard deviation) for the timing of ETH re- lease was 4.56 2.95 min prior to pre-ecdysis onset, and the duration of ETH secretion was 4.463.10 min (n = 13). On the other hand, 40% of Inka cells (n = 9/24) showed no sign of secretory activity by pre-ecdysis on- set. After initiation of pre-ecdysis contractions, it was usually impossible to continue monitoring loss of EGFP fluorescence as a result of movement artifacts.

All Inka cells are depleted of ETH-EGFP by the end of ecdysis sequence (Figure 2C).

Injection of ETH Induces the Ecdysis Sequence Because ETH release coincides with onset of pupal pre- ecdysis (see previous section), we wanted to determine whether ETH injection would trigger the pupal ecdysis sequence. TheDrosophilageneethencodes a precursor producing one copy each of two peptides, ETH1 and ETH2, which share similar structure and biological activ- ity[14]. In vivo experiments were carried out primarily in puparium-free preparations.

Injection of ETH1 alone (0.4, 4, or 20 pmol) into pharate pupae (w1–2 hr prior to natural ecdysis) induced within 1–3 min strong pre-ecdysis contractions followed by ec- dysis and postecdysis contractions sequentially (n = 23;

Figure 2); saline-injected controls showed no responses (n = 4, not shown). ETH-induced pre-ecdysis showed a strong dose dependence (Figure 2), with higher doses inducing shorter pre-ecdysis duration and higher fre- quency of contractions. Similar but somewhat less pro- nounced dose-dependent effects were observed during ecdysis behavior, whereas the frequency of postecdysis contractions showed little or no dose dependence dur- ing the first 10 min of behavior (Figure 2).

Injection of ETH2 was less efficacious for induction of the behavioral sequence compared to ETH1. ETH2 (0.4 pmol) generated prolonged pre-ecdysis behavior lasting toR50 min, but no ecdysis behavior. In contrast, injection of the same dose of ETH1 (0.4 pmol) produced the complete behavioral sequence consisting of pre-ec- dysis, ecdysis, and postecdysis. Higher doses of ETH2 Figure 2. Comparison of Timelines for Natu- ral and ETH-Induced Ecdysis Behavioral Sequences

Different amounts of ETH1 alone, ETH2 alone, or a cocktail of ETH1 and ETH2 induced pre- cocious behaviors in puparium-free pharate pupae. With the wet weight of pupae (w1.3 mg) taken into account, 0.4, 4, and 20 pmol ETH injections are assumed to result in con- centrations ofw300 nM, 3mM, and 15mM in vivo, respectively. Error bars represent stan- dard deviation (SD).

(4)

(20 pmol) generated a behavioral sequence comparable to that induced by 4 pmol ETH1 in terms of pre-ecdysis duration and frequency of contractions (Figure 2).

We also examined behaviors after injecting a cocktail of ETH1 and ETH2 (0.4 pmol of each peptide). Because the two peptides are processed from the same precur- sor, it is likely that these peptides are coreleased under natural conditions. The duration of the behavioral se- quence induced by injection of the cocktail was similar to the naturally occurring sequence or one induced by 0.4 pmol ETH1 alone (Figure 2). We estimate that a 0.4 pmol ETH injection into a prepupaw10 hr after pupar- ium formation (wet weight = 1.360.3 mg; n = 7) results in a concentration ofw300 nM in vivo.

ETH Receptors Are Expressed in Diverse Ensembles of Peptidergic Neurons

ETH acts directly on the CNS to initiate the ecdysis be- havioral sequence in moths and flies via unknown sig- naling pathways within the CNS[15, 16]. A starting point for elucidation of these downstream signaling pathways is identification of primary neuronal targets of ETH. The ETH receptor gene (CG5911), first identified inDrosoph- ila, encodes two G protein-coupled receptors, ETHR-A and ETHR-B, via alternative splicing [9, 10]. We em- ployed in situ hybridization for identification of central neurons expressing ETHR-A and ETHR-B by using DNA probes specific for each receptor subtype.

We located ETHR-A and ETHR-B transcripts in mutu- ally exclusive populations of neurons distributed throughout the CNS (Figure S1), suggesting that two subtypes of ETH receptors likely mediate different func- tions. Further analysis revealed that most ETHR-A neu- rons are peptidergic. Neurons expressing ETHR-B have not been identified thus far.

Multiple ensembles of ETHR-A neurons are classified according to specific neuropeptides they express. Pep- tides expressed in these ensembles were identified by using GAL4 transgenes under control of neuropeptide promoters to driveUAS-GFPorUAS-GCaMPexpres- sion (GAL4::GFP or GAL4::GCaMP). Expression of neu- ropeptides in these cells was confirmed by combining immunohistochemical staining and in situ hybridization.

The first ETHR-A ensemble comprises six pairs of lat- eral abdominal neurons producing kinin, also known as drosokinin (Figures 3A and 3A0)[17]. These cells project axons posteriorly along the lateral edge of the neuropile and then turn anteriorly along the midline of ventral nerve cord, where they arborize and form possible central release sites. Axons of these cells also exit the CNS through nerve roots, suggesting peripheral kinin release.

The second ETHR-A ensemble contains three pairs of ventrolateral FMRFamide neurosecretory cells (Tv1–3 or T1–3) in the thoracic neuromeres TN1–3 (Figures 3B and 3B0). These cells project axons into the dorsomedial neurohemal organs (NHOs) specialized for peptide re- lease into the hemolymph[18, 19].

The third class of ETHR-A neurons comprises the eclosion hormone (EH)-producing VM neurons in the brain, which project one axonal branch anteriorly into the neurohemal ring gland and a second posteriorly along the dorsal midline of the entire ventral nerve cord (Figures 3C, 3C0, 3D and 3D0)[20].

The fourth ETHR-A ensemble is composed of paired dorsolateral neurons producing CCAP, MIPs, and bursi- con in subesophageal, thoracic, and abdominal neuro- meres (SN1–3, TN1–3, AN1–7, respectively) (Figures 3E and 3E0). These cells are likely homologs of moth neu- rons 27/704, on the basis of their anatomy, peptide coexpression profile, and functional roles during pupal ecdysis (seeResults). We henceforth refer to these neu- rons asDrosophilaneurons 27/704 and subdivide them on the basis of peptide coexpression. In AN1–4, CCAP is colocalized with MIPs (Figures 3F, 3F0, and 3F00) and the heterodimeric peptide hormone bursicon (composed of burs and pburs subunits;Figure 3G). In TN2–3 and AN5–

9, CCAP is colocalized with burs, but pburs is not expressed in these neurons[21, 22]. Finally, CCAP is co- localized with MIPs in large paired neurons of AN8,9 (Figures 3F, 3F0, and 3F00), but ETHR-A expression has not been confirmed in these cells. The presence of MIP mRNA in abdominal neurons 27/704 was further confirmed by in situ hybridization (Figure S2).

Ca2+Imaging of Primary ETH Targets in Transgenic Flies

Having shown that ETH receptors occur in diverse en- sembles of peptidergic neurons, we asked whether these cells are activated by ETH and whether this activ- ity coincides with specific behavioral steps of the ecdy- sis sequence. We monitored calcium dynamics in each group of ETHR-A neurons by driving expression of the GFP-based Ca2+sensor, GCaMP[23, 24], in genetically defined sets of neurons with the binary GAL4/UAS sys- tem [25]. Ca2+ elevation induces a conformational change of GCaMP, increasing its GFP fluorescence [23, 24]. Using optical imaging of GFP fluorescence, we monitored [Ca2+]i dynamics of ETHR neurons and associated these events with each behavioral phase in- duced by ETH.

An abundance of evidence indicates that the ecdysis behavioral sequence in insects is centrally patterned.

In particular, the onset and duration of each behavior in the sequence (pre-ecdysis I, pre-ecdysis II, ecdysis) is the same whether observed in vivo or as fictive behav- ior recorded from the isolated CNS in vitro[5, 26–28]. On the basis of this evidence, we associated [Ca2+]idynam- ics of ETHR-A neuron ensembles of the isolated CNS with behaviors observed in puparium-free preparations.

FMRFamide Neurons and Their Neurohemal Endings Become Active Early in Pre-ecdysis

We monitored [Ca2+]ilevels in ETHR-A/FMRFamide Tv neurons by preparing transgenic flies doubly homozy- gous for FMRFa-GAL4 and UAS-GCaMP. Prior to ETH1 exposure (4–6 hr prior to ecdysis), Tv cell bodies and neurohemal endings in the dorsomedial NHO (inset pictures inFigures 4A and 4B) exhibit low levels of basal GCaMP fluorescence.

Exposure of the CNS to ETH1 (600 nM) elicits robust increases in calcium-associated fluorescence in cell bodies and axon terminals of all Tv neurons (Figure 4A).

At this concentration of ETH1, calcium dynamics typi- cally are characterized by transient, spike-shaped fluc- tuations superimposed upon a slow upward shift of the baseline, beginning 7.963.49 min (n = 9) after expo- sure to the peptide. This response lastsw10–15 min,

(5)

after which weaker spike-like fluctuations continue with- out baseline changes until the end of recordings (w40 min). We estimate that a concentration of 600 nM ETH1 results from a dose ofw0.4 pmol of the peptide in vivo (seeFigure 2). Thus the major calcium response of Tv neurons coincides with the early phase of pre-ec- dysis, and weaker activity persists through ecdysis and postecdysis (Figures 4A and 4C). On the other hand, ETH2 alone (600 nM) generated Ca2+ responses after a longer delay comparable to one following exposure to 60 nM ETH1 (Figure 4C; see below). The longer delay of Ca2+responses after ETH2 fits with our observations of in vivo behavior, where ETH2 is a less potent agonist than ETH1 (Figure 2). The cocktail of ETH1 and ETH2 (600 nM each) evokes Ca2+dynamics after a delay sim- ilar to that induced by ETH1 alone (Figure 4C).

Overall, [Ca2+]idynamics observed in Tv neurons are synchronized (Figure 4A). In many preparations, Tv neu- rons from the same neuromere appear to be strongly coupled, given that they produce precisely synchro- nized Ca2+dynamics (see traces from Tv3L and Tv3R

inFigure 4A). Transient Ca2+signals are obvious in the terminal processes of Tv neurons in NHO, the release sites of FMRFamides.

Lower concentrations of ETH1 elicit calcium dynamics after a somewhat longer delay. Interestingly, calcium dy- namics are obvious first in neurohemal endings of the NHO, followed by fluctuations in cell bodies (Figures 4B and 4C). This was particularly evident at 60 nM ETH1, where a robust calcium response in the NHO was accom- panied by only a weak response in the Tv2 cell body (Fig- ure 4B). No calcium responses are observed in Tvs ex- posed to 6 nM ETH1 (Figure 4C). In our experiments, peakDF/F reached 9.3465.13% in response to 600 nM ETH1 over a baseline noise level of less than 0.7%.

EH Neurons Reach Peak Activity at Ecdysis

VM neurons producing eclosion hormone (EH) have been implicated as primary ETH targets during fly and moth ecdysis[8, 29–31]. We demonstrate here expres- sion of ETHR-A in VM neurons, confirming that they are primary ETH targets. To determine whether ETH Figure 3. ETH Receptor Subtype A Expression in Multiple Peptidergic CNS Ensembles ofDrosophilaPharate Pupae

(A–E) Double labeling for ETHR-A and peptides in neuronal ensembles. Peptidergic neurons were labeled by using either peptide-specific an- tibodies or GFP/GCaMP targeted by the binary GAL4/UAS system (GAL4::GFP or GAL4::GCaMP). (A and A0) ETHR-A-specific in situ hybridiza- tion labeled six pairs of lateral abdominal neurons ([A], arrows) producing kinin ([A0], C127::GFP). (B and B0) ETHR-A is detected in three pairs of thoracic ventrolateral neurosecretory cells ([B], Tv1–3, arrows) producing FMRFamides ([B0], FMRFa::GCaMP). (C and D) ETHR-A expression in a pair of brain EH cells (EH::GCaMP, arrows) of pharate pupae (C and C0) and adults (D and D0). (E and E0) Intense ETHR-A expression (E) in pep- tidergic neurons coexpressing CCAP, MIPs, and/or bursicon ([E0], CCAP::GCaMP, red) in SN, TN, and AN1–7. These neurons are likely homologs of moth neurons 27/704 (seeResults). Arrows indicate position of FMRFamides neurons described above (green). Note that most neurons 27/704 (CCAP::GCaMP) express ETHR-A; arrowheads in (E) point to the few ETHR-A cells that do not show CCAP::GCaMP staining.

(F) Myoinhibitory peptides (MIPs, green, [F and F0]) are coexpressed in neurons 27/704 (CCAP::GCaMP) in AN1–4 and AN8,9 (arrows, red, [F and F00]).

(G) Schematic diagram showing peptidergic ensembles producing ETHR-A. Note that neurons 27/704 are subdivided into three subgroups on the basis of cotransmitter expression: CCAP/MIPs/bursicon in AN1–4; CCAP/MIPs in AN8,9; and CCAP in SNs, TNs, and AN5–7.

(6)

elicits activity in EH neurons, we prepared transgenic flies doubly homozygous for EHup-GAL4 and UAS- GCaMP, which show GCaMP fluorescence only in these cells (Figure 5A;Movie S2).

EH neurons are highly sensitive to ETH1, exhibiting robust [Ca2+]idynamics upon exposure to concentra- tions as low as 6 nM (Figures 5A and 5C). No detectable fluorescence responses are observed after exposure to 0.6 nM ETH1 over a period of 50–60 min (n = 6). The latency to Ca2+responses is inversely proportional to the concentration of ETH1; higher ETH1 concentrations evoke faster responses (Figures 5A and 5C). The cock- tail of ETH1 and ETH2 (600 nM each) elicited Ca2+re- sponses after aw10–15 min delay (Figure 5C).

Close examination of these ETH-evoked fluorescence responses reveals two components distinguished by slow and fast kinetics. The slow component is character- ized by a gradual increase in baseline levels of Ca2+fol- lowed by a decrease over 20–30 min, whereas the fast component is composed of transient, spike-like activity (Figure 5A). Fast components have durations ranging from 5–20 s (Figure 5B). PeakDF/F responses are quite variable, even among a group of neurons exposed to the same ETH1 concentration. We could not detect any significant concentration dependence in peakDF/F.

Distinct Subsets of Neurons 27/704 Are Active during Different Phases of the Ecdysis Sequence We analyzed ETH-evoked Ca2+ signals of neurons 27/704 in transgenic flies carrying CCAP-GAL4 and UAS-GCaMP. Use of the CCAP promoter to drive GCaMP expression resulted in a reporter pattern identi- cal to that described previously (Figure 3E0)[11]. Upon exposure to 600 nM ETH1, distinct subsets of neurons 27/704 exhibited reproducible, stereotypic Ca2+ re- sponses in terms of peak intensity, latency, and termi- nation of Ca2+dynamics (Figure 6;Movie S3). According to the magnitude of peak fluorescence intensity (peak DF/F), neurons 27/704 fall into three major groups:

strong responders (R5 peak DF/F), weak responders (%5 peakDF/F), and nonresponders (Figure 6C;Fig- ure S3). The strong-responder group includes neurons 27/704 in AN1–4 (CCAP/MIPs/bursicon), AN8,9 (CCAP/

MIPs), and TN3 (CCAP). Weak responders are neurons 27/704 in SN2–3, TN1–2, and AN7 producing CCAP only. Neurons in the brain, SN1, and AN5,6 showed no re- producible Ca2+dynamics in response to 600 nM ETH1.

In response to ETH1, neurons 27/704 in TN3 and AN8,9 become active within 10–15 min, whereas neu- rons 27/704 in AN1–4 are activated after a 15–25 min de- lay (Figure 6). Neurons in TN3 and AN8,9 are therefore Figure 4. ETH-Evoked Ca2+Dynamics in Thoracic Ventrolateral Neurons or Tv1–3 Expressing ETHR-A and FMRFamide

(A) Six hundred nanomolar ETH1 induces robust and coordinated Ca2+responses in all Tv cell bodies expressing FMRFamides after a short delay (4–6 min). Inset video images show Tv neurons (yellow lined boxes) from which Ca2+responses were recorded. Note that bilaterally paired Tv neurons (Tv 3L,3R) exhibit highly synchronized Ca2+signaling.

(B) Activation of dorsal neurohemal organs (NHO) innervated by Tv neurons occurs before that of the cell bodies.

(C) Latency to onset of Ca2+responses in Tv neurons is dependent on ETH concentration. The behavioral timeline shown in a top axis depicts the effect of 0.4 pmol ETH1 injected into puparium-free pharate pupae (seeFigure 2).

(7)

activated just prior to ecdysis onset, indicating their possible roles in initiation and maintenance of ecdysis behavior. In addition, Ca2+ dynamics observed in AN8,9 neurons terminated early in postecdysis, support- ing this interpretation. On the other hand, Ca2+dynamics of neurons in AN1–4 begin during ecdysis and increase in intensity during the entire postecdysis period, sug- gesting their roles in these events (Figure 6). The cocktail of ETH1 and ETH2 (600 nM each) evoked Ca2+dynamics similar to those induced by ETH1 alone (Figure 6C).

Two groups of neurons 27/704 in abdominal neuro- meres (AN1–4 versus AN8,9) exhibit differences in sensi- tivity to ETH and in their patterns of [Ca2+]idynamics. We found that 6–60 nM ETH1 activates neurons in AN1–4 (n = 4), whereas higher concentrations of ETH1 (R600 nM) are required to activate neurons in AN8,9 (n = 9). In addi- tion, neurons in AN8,9 generate transient (1–2 min) Ca2+

spikes over a 15–20 min period after ETH1 activation (Figures 6A and 6C), whereas neurons in AN1-4 generally produce slower, more persistent Ca2+dynamics (Figures 6B and 6C). These differences among subgroups of neu- rons 27/704 suggest their different functional roles dur- ing the ecdysis sequence.

Targeted Ablations of Specific ETHR Neurons Have Behavioral Consequences

To evaluate behavioral roles of specific ETHR neurons, we investigated phenotypes of the pupal ecdysis se- quence in transgenic flies bearing targeted ablations of ETHR neurons, including Tv FMRFamide neurons, EH neurons, and CCAP neurons (27/704 homologs). In con- trol flies carrying UAS-reaper and UAS-GCaMP, but lacking theGAL4driver, pupal ecdysis was executed as in wild-type flies: pre-ecdysis (0–10 min), ecdysis

(10–23 min), and postecdysis (23–100 min) (Figure 7).

Given that we used puparium-intact animals, the dura- tion of ecdysis behavior may have been overestimated.

Transgenic flies bearing targeted ablations of Tv FMRFamide neurons (FMRF-KO) were generated by crossing females doubly homozygous forFMRFa-GAL4, UAS-GCaMP with homozygous UAS-reaper males.

Pupal ecdysis of FMRFa-KO flies is very similar to that of control flies. Because FMRFa-GAL4 drives expression of GAL4 only in three pairs of thoracic Tv neurons and one pair of unidentified neurons in SN, FMRFa-KO flies lost only Tv neurons and not other FMRFamide neurons in the CNS (n = 7; Figure S4).

FMRFa-KO flies complete pupal ecdysis without any detectable abnormality, except that pre-ecdysis con- tractions appear weaker than in control flies.

We then analyzed pupal ecdysis of VM neuron knock- out flies (EH-KO). Behavioral analysis showed that, al- though they complete pupal ecdysis without any severe defects or lethality, ecdysis onset is delayedw4 min (Figure 7). As a result of this delay, EH-KO flies show lon- ger pre-ecdysis than control flies (Figure 7). Additional parameters governing pre-ecdysis, ecdysis, and post- ecdysis are indistinguishable between EH-KO and control flies.

Finally, we generated CCAP-KO flies in order to exam- ine the functional roles of neurons 27/704 (CCAP neu- rons) in pupal ecdysis. As expected, CCAP-KO flies (n = 7/7) failed to initiate ecdysis contractions and could not complete head eversion. Instead, they show pro- longed pre-ecdysis contractions forw25 min, followed by weak random contractions of the abdomen (different from ecdysis and postecdysis contractions of control flies) for the next 50 min (Figure 7).

Figure 5. ETH-Evoked Ca2+ Dynamics in Neurons Producing ETHR-A and EH (A) ETH1 evokes robust Ca2+responses in VM neurons producing EH. The inset video image shows EH neurons in the brain (arrows).

(B) The time course of fast Ca2+oscillations were monitored with high frequency sam- pling rate (45 hz).

(C) The latency of EH neuron activation by ETH is concentration dependent. Note that EH neurons are more sensitive to ETH1 than Tv neurons (see Figure 4). The behavioral timeline shown in a top axis depicts the effect of 0.4 pmol ETH1 injected into puparium-free pharate pupae (seeFigure 2).

(8)

Discussion

ETH Is a Command Chemical for an Innate Behavior We have described orchestration of an innate behavior, theDrosophilapupal ecdysis sequence, by the endo- crine peptide ETH. ETH release coincides with onset of

behavior, and injection of ETH triggers the complete behavioral sequence, consistent with its role in ecdysis activation previously established in the mothsManduca sexta[4, 5, 28]andBombyx mori[32, 33]and inDrosoph- ilalarvae and adults[7, 14, 31]. Absence of ETH causes lethal ecdysis deficiency, a phenotype that is rescued by Figure 6. ETH-Evoked Ca2+Dynamics in Neurons 27/704 Producing CCAP, MIPs, and/or Bursicon

(A and B) Representative intracellular Ca2+signals of neurons 27/704 (CCAP::GCaMP) in AN7–9 (A) and AN1–3 (B) upon exposure to 600 nM ETH1. (A1and B1) Video images show cell bodies and processes (NP; yellow lined boxes) of neurons where Ca2+responses were recorded.

(A2and B2) Following exposure to ETH, cell bodies and neural processes (NP) showed robust Ca2+response after characteristic delays. (A3

and B3) Time-lapse video images captured in the course of Ca2+responses after ETH1 application. Timing of video-image recording (A2a–f

and B2a–f) is indicated by vertical arrows in (A2) and (B2) (faint red). White arrows in (A3) and (B3) indicate elevation of fluorescence signals.

(C) In the left panel, a schematic diagram summarizes Ca2+dynamics in neurons 27/704 (CCAP::GCaMP) in the prepupal CNS. Upon exposure to ETH, each subtype of neurons 27/704 exhibited stereotypic Ca2+dynamics, as measured by peak fluorescence response, latency to onset, or termination of Ca2+activity. The middle and right panels show the latency to onset (red) or termination (black) of Ca2+responses induced by ETH1 alone (600 nM) or by a cocktail of ETH1 and ETH2 (600 nM each). The behavioral timeline shown in a top axis depicts the effect of ETH1 (0.4 pmol) or the cocktail of ETH1 and ETH2 (0.4 pmol each) injected into puparium-free pharate pupae (seeFigure 2).

(9)

ETH injection[7]. ETH therefore functions as a ‘‘com- mand chemical’’ to orchestrate an innate behavior.

ETH Activates Behavior by Recruitment of Multiple Peptidergic Ensembles in the CNS

We identified primary CNS targets of ETH by using ETHR-specific in situ hybridization. ETHR-A occurs in multiple classes of peptidergic neurons producing EH, CCAP/MIPs/bursicon, FMRFamides, or kinin. The fol- lowing sections examine in detail the likely roles of pep- tidergic ETHR-A ensembles in the stepwise recruitment of the ecdysis behavioral sequence.

VM Neurons

We showed expression of ETHR-A in VM neurons, which release EH. In response to ETH, VM neurons become active prior to ecdysis behavior and reach peak levels of activity during ecdysis. These results provide further support for a previously described positive-feedback signaling pathway between VM neurons and Inka cells [8, 31, 34]. This feedback is thought to ensure depletion of ETH from Inka cells.

These findings are striking because independent evi- dence indicates that homologous VM neurons in the mothManducaare direct targets of ETH and that their secretory products regulate ecdysis behaviors down- stream of ETH. For example, isolated EH neurons of Manducarespond to direct action of ETH with increased excitability and spike broadening [30]. In response to ETH action, these neurons release EH, causing cGMP elevation and increased excitability in CCAP-containing neurons 27/704 of the thoracic and abdominal ganglia [35]. CCAP and MIPs, cotransmitters produced by neu- rons 704, are implicated in eliciting ecdysis behavior (unpublished data).

A homologous role for EH in activation ofDrosophila 27/704 neurons has not been clearly demonstrated [31, 36]. For example, no cGMP elevation is observed in these neurons during the natural ecdysis sequence [29, 31, 37]. This lack of cGMP elevation suggests that CCAP neurons are not directly targeted by EH inDro- sophila. Nevertheless, EH-knockout flies exhibit a delay in ecdysis initiation, suggesting that EH may modulate excitability in 27/704 cells indirectly through release of

additional factors within the CNS. We therefore propose that activation of EH neurons by ETH serves two pur- poses: (1) release of EH into the hemocoel functions as part of a positive-feedback pathway to ensure ETH de- pletion from Inka cells[8]; (2) release of EH within the CNS synergizes direct ETH actions on different subsets of neurons 27/704 producing CCAP, MIPs, and bursi- con, perhaps indirectly through release of downstream signals within the CNS.

Neurons 27/704

We have shown that neurons 27/704 expressing ETHR-A respond to ETH with unique patterns of Ca2+dynamics.

These neurons are subdivided by pattern of transmitter expression: CCAP/MIPs/bursicon in AN1–4; CCAP/MIPs in AN8,9; and CCAP in SN1–3, TN1–3, and AN5–7 (Figures 3E, 3F, and 3G). We determined temporal pat- terns of Ca2+dynamics in each neuronal subset relevant to the behaviors observed. On the basis of these tempo- ral patterns, we propose that direct action of ETH on neurons 27/704 in TN3 and in AN8,9 induces initiation and execution of ecdysis contractions and head ever- sion. In support of this, we show that ablation of CCAP neurons abolishes ecdysis contractions and head ever- sion ([11]and this study). Our parallel study inManduca showed that neurons 704 expressing ETHR-A and their peptide cotransmitters, CCAP and MIPs, are implicated in control of the ecdysis motor pattern, supporting the homologous function of 27/704 neurons inDrosophila (unpublished data).

Neurons 27/704 in AN1–4 produce CCAP, MIPs, and bursicon, and therefore a cocktail of these peptides is likely released within the CNS and into the hemolymph during postecdysis. We suggest that centrally released peptides control postecdysis movements, whereas blood-borne CCAP/MIPs regulate heart beat and blood pressure for cuticle expansion and bursicon controls sclerotization of expanded new cuticle. Bursicon was recently identified as a heterodimeric peptide hormone regulating cuticle plasticization, sclerotization, and mel- anization[21, 22, 38].

FMRFamide Neurons

The Drosophila FMRFamide gene (FMRFa) encodes multiple FMRFamide-related neuropeptides, which are Figure 7. Targeted Ablations of Subsets of ETH Receptor Neurons Result in Various Degrees of Behavioral Defects

The ecdysis behavioral sequence observed in the puparium-intact preparation of control insects is indistinguishable from that of wild-type W1118. Comparison of control and knockout flies carrying ablations of specific subsets of ETHR neurons revealed important behavioral differences.

CCAP-KO flies show prolonged pre-ecdysis and completely lack ecdysis and postecdysis motor pattern. EH-KO flies showw4 min delay in ecdysis onset. The ecdysis sequence of FMRFa-KO flies is very similar to that of control flies. Note that duration of ecdysis (asterisk) was over- estimated because the puparium-intact preparation provides somewhat compromised discern between ecdysis and early phase of postecdysis.

Error bars represent SD.

(10)

expressed in many different cell types, including neuro- endocrine cells, interneurons, and perhaps motoneu- rons[18, 19, 39, 40]. Among these diverse FMRFamide- producing neurons, ETHR-A expression is confined to three pairs of thoracic neurosecretory neurons, Tv1–3.

Results of the present study show that the Tv neurons are activated early in pre-ecdysis and that they remain active during ecdysis and postecdysis. However, FMRFa-KO flies show no differences in timing of the ec- dysis behavioral sequence. Because FMRFamides en- hance twitch tension of larval body-wall muscles through synaptic modulation at the neuromuscular junction[41], blood-borne FMRFamides released from Tv neurons likely facilitates pre-ecdysis, ecdysis, and postecdysis contractions. Thus the role of Tv neurons as primary ETH targets may be enhancement of muscle contraction during the behaviors. Further work to substantiate this is in progress.

Kinin Neurons

We report here expression of ETHR-A in kinin neurons of abdominal neuromeres ofDrosophila(Figure 3A).Dro- sophilakinin is known to be involved in water balance [17, 42, 43], but its central functions have not been de- scribed or considered. Expression of ETHR-A in kinin neurons appears to be a conserved mechanism in fly and moth, because we also found that the Manduca ETHR-A is expressed in abdominal neurosecretory cells (L3,4), which produce kinins and diuretic hormones (DHs) (unpublished data). We further found that the isolated ManducaCNS generates the fictive pre-ecdysis motor pattern upon exposure to a cocktail of kinin and DHs.

These findings suggest that ETH activates L3,4neurons inManducato release kinins and DHs centrally, which initiate and execute pre-ecdysis. On the basis of the conservation betweenDrosophilaandManducain spa- tial expression pattern of ETHR, we propose that ETH initiates pre-ecdysis behavior indirectly via central re- lease of kininDrosophila. Functional roles for ETHR-A kinin neurons in theDrosophilaecdysis behavioral se- quence will be published elsewhere.

Relationship of Ca2+Signaling and Secretory Activity of Peptidergic Neurons

We detected robust Ca2+signaling in multiple peptider- gic ETHR ensembles in response to ETH application.

Although our imaging results alone do not provide direct evidence for peptide secretion from these ensembles, several observations make this likely. In Drosophila, ETH release from Inka cells and subsequent reduction of CCAP-immunoreactivity (-IR) in descending axons was detected immediately after pupal ecdysis [11].

Furthermore, ETH-induced ecdysis in flies led to a major decrease in EH-IR after aw10–20 min delay[29]. This corresponds to the timing of ETH-evoked Ca2+signals in EH neurons (Figure 5). In addition to ETH-induced Ca2+ responses in cell bodies, we observed robust Ca2+ signals in axonal processes of EH neurons (not shown), CCAP neurons (Figure 6A), and dorsomedial NHO innervated by Tv neurons producing FMRFamides (Figure 4B). Because intracellular Ca2+is a critical sec- ond messenger in the regulation of secretory events, it is likely that Ca2+dynamics in axons and NHO reflect se- cretory activity in these structures. Finally, in moths, considerable decreases in EH-IR, CCAP-IR, and MIP-

IR also have been observed during and/or after larval ec- dyses[8, 32, 35, 44, 45]. These observations together suggest that calcium dynamics described here in ETHR neurons are indicative of peptide secretion.

Timing of Peptidergic Ensemble Activation

We showed that ETH elicits intracellular Ca2+signals in five distinct peptidergic ensembles. Although all of these neurons express the same ETHR subtype (ETHR-A; Figure 3), their patterns of Ca2+ dynamics were quite different (Figures 4–7). Each cell type produc- ing FMRFamide, EH, CCAP, CCAP/MIPs, or CCAP/

MIPs/bursicon is activated by ETH after a specific and reproducible delay. Delays in the cellular activation of ETHR neurons may underlie sequential recruitment of multiple neural ensembles and behaviors they control.

Why do we observe differential delays in the re- sponses of neuronal ETH targets? When the ETHR is expressed heterologously in mammalian cell lines, ETH-dependent activation of the receptor is rapid and generates strong [Ca2+]i dynamics within 5–20 s [10].

We propose that ETH activates inhibitory inputs, which suppress activity and release of neuropeptides from ETHR-A neurons during specific time periods.

Evidence supporting this proposal comes from sev- eral sources. It has been shown in both fly (Drosophila) and moth (Manduca) that descending inputs from ce- phalic ganglia and thoracic ganglia delay ecdysis onset.

For example, during theDrosophilaeclosion sequence (adult ecdysis), neck ligation during pre-eclosion causes an immediate transition to ecdysis[29]. Normally, pre- eclosion behavior proceeds forw1 hr prior to the transi- tion to eclosion. Neck ligation may therefore eliminate descending inhibition from cephalic ganglia. In ETH- injectedManduca, activation of ecdysis circuitry is al- ready complete shortly after pre-ecdysis onset, well be- fore initiation of ecdysis behavior some 30 min later. The delay (30 min) between activation of ecdysis circuits and onset of ecdysis behavior is controlled by descending inhibition from cephalic and thoracic ganglia[28, 45, 46].

Intriguingly, preliminary experiments indicate that treat- ment of the Drosophila CNS with the Ca2+ channel blocker, omega-Aga-IVA[47], accelerates activation of EH and CCAP neurons after ETH exposure (Y.-J.K. and M.E.A., unpublished data). Furthermore, this Ca2+channel blocker even allows ETH to induce strong Ca2+responses from nonresponder 27/704 neurons (e.g., neurons in abdominal neuromere 5–8). These observations could be explained by disinhibition of these neurons through block of inhibitory-transmitter release by omega-Aga-IVA.

The identities of descending inhibitors have not yet been determined. Neurons expressing ETHR-B are pos- sible candidates. This hypothesis could be tested by an- alyzing ETH-evoked Ca2+responses of ETHR-A neurons in flies bearing ETHR-B mutant background. Creation of such mutant lines is in progress.

Alternative hypotheses for the delays in peptidergic ensemble activation are certainly possible. For example, circulating ETH levels may be dynamic, starting at low levels and becoming elevated later in the behavioral se- quence. This has been shown inManduca[5, 48]. If pep- tidergic ensembles have differing sensitivities to ETH, those with lower sensitivity might be activated later. Fur- ther work is needed to test these hypotheses.

(11)

A Model for Peptidergic Regulation ofDrosophila Pupal Ecdysis

In Drosophila, pupal ecdysis is accomplished by se- quential recruitment of three major behavioral units:

pre-ecdysis (0–10 min), ecdysis (10–15 min), and post- ecdysis (15–100 min;Figure 1A). Each behavioral unit is programmed in the CNS and sequentially activated by direct actions of ETH, which is synthesized and released from peripheral endocrine Inka cells. Around 4–5 min before pre-ecdysis onset, a sizeable portion (w50%) of Inka cells initiates secretion of ETH into the hemolymph (Figure 1B), whereas the remaining portion completes secretion after onset of pre-ecdysis. Appear- ance of ETH in the hemolymph activates ETHR-A in neurons expressing neuropeptides including kinin, FMRFamides (Tv1–3), EH, or CCAP, MIPs, and bursicon (Figure 3), but they are not released until descending inhibition is removed at key times during the ecdysis sequence. Upon activation of ETHR, the central release of kinin initiates pre-ecdysis contractions, whereas Tv neurons secrete FMRFamides to enhance neuromuscu- lar transmission. ETH activates neurons producing EH, CCAP, CCAP/MIPs, and CCAP/MIPs/bursicon at differ- ent times (Figure 8). EH cells in the brain and neurons producing CCAP in TN3 and CCAP/MIPs in AN8,9 be- come active w10–13 min after pre-ecdysis initiation (Figures 5, 6, and 8). EH participates in timing the activa- tion of ecdysis neurons, whereas CCAP and MIPs from TN3 and AN8,9 control initiation and execution of the ec- dysis motor program (Figure 8). At the end of ecdysis (25 min after pre-ecdysis onset), neurons in AN1–4 secrete a cocktail of CCAP, MIPs, and bursicon, which likely reg- ulate postecdysis contractions and processes associ- ated with cuticle expansion, hardening, and tanning (Figure 8).

Conclusions

The behavioral repertoire of animals is replete with innate or instinctive motor sequences[1]. The ecdysis sequence in insects is composed of a series of motor patterns that occurs over a relatively prolonged time scale. Of particular interest is the initiation and regula- tion of this behavioral sequence by an endogenous chemical, ecdysis-triggering hormone. In the present study, we mapped central ETH receptor neurons, and discovered that they comprise multiple peptidergic en- sembles, which are recruited sequentially to generate each phase of the ecdysis sequence. Ensemble-specific knockout analysis supports this interpretation.

Each step of the ecdysis sequence—pre-ecdysis, ecdysis, postecdysis—is driven by a central pattern generator (CPG) within the CNS in the absence of sen- sory input. It is known that amines and peptides can modulate and reconfigure neuronal circuits comprising CPGs so as to elicit a variety of motor patterns[49]. It seems likely that the multiple peptidergic ensembles we describe here as targets for ETH may be involved in configuring and activating CPGs underlying each step of the ecdysis sequence.

Processes in the brain that govern behaviors over lon- ger time frames such as sleep, mood, sexual activities, and even learning and memory could be associated with coordinated release of neuromodulators such as peptides [50]. Further work on activation of central

peptidergic ensembles in the CNS may shed light on mechanisms underlying release of a variety of behaviors.

Experimental Procedures Fly Strains

CCAP-GAL4flies[11]were obtained from J. Ewer (Cornell Univer- sity, Ithaca, New York).FMRFa-GAL4[51]andC127-GAL4,UAS- GFP[52]flies were provided by P. Taghert (Washington University, St. Louis).UAS-GCaMP[24]was obtained from R. Axel (Columbia University, New York). Other flies were obtained from the Blooming- ton Stock Center (Bloomington, Indiana):EHup-GAL4(stock number 6310),UAS-reaper(5824),UAS-CD8m-GFP(5137), and W1118(5905).

In Situ Hybridization and Immunohistochemistry

For in situ hybridization, we prepared dioxigen-labeled single- stranded DNA probes specific for ETHR-A (+636 to +1261; Genbank accession number, AY220741) and ETHR-B (+483 to +1386;

AY220742). We prepared flies expressing GFP or a GFP derivative (GCaMP) in peptidergic neurons and used them for in situ hybridiza- tion and GFP immunohistochemical staining. Further details are pro- vided in the Supplemental Data.

Behavioral Analysis

The puparium-free preparation was used for analysis of the natural or peptide-induced pupal ecdysis sequence in W1118flies, and the puarium-intact preparation was used for analysis of knockout (KO) Figure 8. Summary of ETH1-Evoked Ca2+Responses from Pepti- dergic ETHR Neurons and a Model for Peptidergic Regulation of the Ecdysis Sequence

(A) Multiple peptidergic neurons expressing ETHR-A are activated sequentially by ETH (600 nM), regulating consecutive phases of pre-ecdysis, ecdysis, and postecdysis behaviors. The behavioral timeline shown in a bottom axis depicts the effect of 0.4 pmol ETH1 injected into puparium-free pharate pupae (seeFigure 2).

(B) A model for peptidergic regulation of the ecdysis behavioral sequence (seeDiscussion).

(12)

flies. KO flies were prepared by crossing lines doubly homozygous forUAS-GCaMPandGAL4driver (those used for in vitro Ca2+imag- ing; female) to homozygous UAS-rpr flies (male). UAS-GCaMP, UAS-rprflies were used as controls. Further details are provided in theSupplemental Datasection.

ETH Release

For in vivo detection of ETH release, pharate pupae of transgenic 2eth3-egfp flies [7] were imaged under a stereomicroscope (SMZ1500, Nikon) equipped with an epi-fluorescence attachment for GFP and imaging capabilities. Further details on ETH release are provided in theSupplemental Datasection.

In Vitro Ca2+Imaging

The in vitro Ca2+imaging experiments were done in the pharate pupae of doubly homozygous flies carrying variousGAL4drivers and UAS-GCaMP responder: for FMRFa, UAS-GCaMP;FMRFa- GAL4/FMRFa-GAL4; for EH, UAS-GCaMP;EHup-GAL4/EHup- GAL4; and for CCAP,UAS-GCaMP;CCAP-GAL4/CCAP-GAL4. We used pharate pupae about 4–6 hr before onset of pupal ecdysis.

The CNS was dissected under ice-chilled fly saline, embedded in 80 ul of 1.5% low melting temperature agarose solution (Sigma type VII-A), and solidified for 30 min in 10–15!C humidified chamber.

We discarded CNS preparations showing spontaneous activities.

For detailed accounts on the imaging setup and protocol, see the Supplemental Datasection.

Supplemental Data

Supplemental Data include Supplemental Experimental Procedures and four figures and are available with this article online at:http://

www.current-biology.com/cgi/content/full/16/14/1395/DC1/.

Acknowledgments

We thank Yoonseong Park for sage advice and critical reading of the manuscript. We thank John Ewer, Jing Wang, S. Robinow, and Paul Taghert for reagents and Ronald Nachman for amino acid analyses.

This work was supported by a grant from the National Institutes of Health (GM 67310) to M.E.A.

Received: June 7, 2006 Accepted: June 8, 2006 Published: July 24, 2006

References

1. Lorenz, K.Z. (1981). The Foundations of Ethology (New York:

Springer-Verlag).

2. Manoli, D.S., Foss, M., Villella, A., Taylor, B.J., Hall, J.C., and Baker, B.S. (2005). Male-specific fruitless specifies the neural substrates of Drosophila courtship behaviour. Nature 436, 395–400.

3. Stockinger, P., Kvitsiani, D., Rotkopf, S., Tirian, L., and Dickson, B.J. (2005). Neural circuitry that governs Drosophila male court- ship behavior. Cell121, 795–807.

4. Zitnan, D., Kingan, T.G., Hermesman, J.L., and Adams, M.E.

(1996). Identification of ecdysis-triggering hormone from an epitracheal endocrine system. Science271, 88–91.

5. Zitnan, D., Ross, L.S., Zitnanova, I., Hermesman, J.L., Gill, S.S., and Adams, M.E. (1999). Steroid induction of a peptide hormone gene leads to orchestration of a defined behavioral sequence.

Neuron23, 523–535.

6. Zitnan, D., Zitnanova, I., Spalovska, I., Takac, P., Park, Y., and Adams, M.E. (2003). Conservation of ecdysis-triggering hor- mone signalling in insects. J. Exp. Biol.206, 1275–1289.

7. Park, Y., Filippov, V., Gill, S.S., and Adams, M.E. (2002). Deletion of the ecdysis-triggering hormone gene leads to lethal ecdysis deficiency. Development129, 493–503.

8. Ewer, J., Gammie, S.C., and Truman, J.W. (1997). Control of insect ecdysis by a positive-feedback endocrine system: Roles of eclosion hormone and ecdysis triggering hormone. J. Exp.

Biol.200, 869–881.

9. Iversen, A., Cazzamali, G., Williamson, M., Hauser, F., and Grim- melikhuijzen, C.J. (2002). Molecular identification of the first

insect ecdysis triggering hormone receptors. Biochem. Bio- phys. Res. Commun.299, 924–931.

10. Park, Y., Kim, Y.J., Dupriez, V., and Adams, M.E. (2003). Two subtypes of ecdysis-triggering hormone receptor in Drosophila melanogaster. J. Biol. Chem.278, 17710–17715.

11. Park, J.H., Schroeder, A.J., Helfrich-Forster, C., Jackson, F.R., and Ewer, J. (2003). Targeted ablation of CCAP neuropeptide- containing neurons of Drosophila causes specific defects in execution and circadian timing of ecdysis behavior. Develop- ment130, 2645–2656.

12. Chadfield, C.G., and Sparrow, J.C. (1984). Pupation in drosoph- ila melanogaster and the effect of the Lethalcryptocephal muta- tion. Dev. Genet.5, 103–114.

13. Zdarek, J., and Friedman, S. (1986). Pupal ecdysis in flies: Mech- anisms of evagination of the head and expansion of the thoracic appendages. J. Insect Physiol.32, 917–924.

14. Park, Y., Zitnan, D., Gill, S.S., and Adams, M.E. (1999). Molecular cloning and biological activity of ecdysis-triggering hormones in Drosophila melanogaster. FEBS Lett.463, 133–138.

15. Zitnan, D., and Adams, M.E. (2004). Neuroendocrine regulation of insect ecdysis. In Comprehensive Molecular Insect Science, Volume 3, I.G. Lawrence, I. Kostas, and S.G. Sarjeet, eds. (Am- sterdam: Elsevier), pp. 1–60.

16. Ewer, J., and Reynolds, S. (2002). Neuropeptide control of molt- ing in insects. In Hormones, Brain and Behavior, D.W. Pfaff, A.P.

Arnold, S.E. Fahrbach, A.M. Etgen, and R.T. Rubin, eds. (San Diego: Academic), pp. 1–92.

17. Terhzaz, S., O’Connell, F.C., Pollock, V.P., Kean, L., Davies, S.A., Veenstra, J.A., and Dow, J.A. (1999). Isolation and characteriza- tion of a leucokinin-like peptide of Drosophila melanogaster.

J. Exp. Biol.202, 3667–3676.

18. Schneider, L.E., Sun, E.T., Garland, D.J., and Taghert, P.H.

(1993). An immunocytochemical study of the FMRFamide neuro- peptide gene products in Drosophila. J. Comp. Neurol.337, 446–

460.

19. White, K., Hurteau, T., and Punsal, P. (1986). Neuropeptide- FMRFamide-like immunoreactivity in Drosophila: Development and distribution. J. Comp. Neurol.247, 430–438.

20. Horodyski, F.M., Ewer, J., Riddiford, L.M., and Truman, J.W.

(1993). Isolation, characterization and expression of the eclosion hormone gene of Drosophila melanogaster. Eur. J. Biochem.

215, 221–228.

21. Dewey, E.M., McNabb, S.L., Ewer, J., Kuo, G.R., Takanishi, C.L., Truman, J.W., and Honegger, H.W. (2004). Identification of the gene encoding bursicon, an insect neuropeptide responsible for cuticle sclerotization and wing spreading. Curr. Biol.14, 1208–1213.

22. Luo, C.W., Dewey, E.M., Sudo, S., Ewer, J., Hsu, S.Y., Honegger, H.W., and Hsueh, A.J. (2005). Bursicon, the insect cuticle-hard- ening hormone, is a heterodimeric cystine knot protein that ac- tivates G protein-coupled receptor LGR2. Proc. Natl. Acad. Sci.

USA102, 2820–2825.

23. Nakai, J., Ohkura, M., and Imoto, K. (2001). A high signal- to-noise Ca(2+) probe composed of a single green fluorescent protein. Nat. Biotechnol.19, 137–141.

24. Wang, J.W., Wong, A.M., Flores, J., Vosshall, L.B., and Axel, R.

(2003). Two-photon calcium imaging reveals an odor-evoked map of activity in the fly brain. Cell112, 271–282.

25. Brand, A.H., and Perrimon, N. (1993). Targeted gene expression as a means of altering cell fates and generating dominant pheno- types. Development118, 401–415.

26. Truman, J.W. (1978). Hormone action on the insect nervous sys- tem part 1 hormonal release of stereotyped motor programs from the isolated nervous system of the cecropia silk moth.

J. Exp. Biol.74, 151–174.

27. Weeks, J.C., and Truman, J.W. (1984). Neural organization of peptide-activated ecdysis behaviors during the metamorphosis of manduca-sexta 1 conservation of the peristalsis motor pat- tern at the larval-pupal transformation. J. Comp. Physiol. [A]

155, 407–422.

28. Zitnan, D., and Adams, M.E. (2000). Excitatory and inhibitory roles of central ganglia in initiation of the insect ecdysis behav- ioural sequence. J. Exp. Biol.203, 1329–1340.

(13)

29. Baker, J.D., McNabb, S.L., and Truman, J.W. (1999). The hor- monal coordination of behavior and physiology at adult ecdysis in Drosophila melanogaster. J. Exp. Biol.202, 3037–3048.

30. Gammie, S.C., and Truman, J.W. (1999). Eclosion hormone pro- vides a link between ecdysis-triggering hormone and crusta- cean cardioactive peptide in the neuroendocrine cascade that controls ecdysis behavior. J. Exp. Biol.202, 343–352.

31. Clark, A.C., del Campo, M.L., and Ewer, J. (2004). Neuroendo- crine control of larval ecdysis behavior in Drosophila: Complex regulation by partially redundant neuropeptides. J. Neurosci.

24, 4283–4292.

32. Zitnan, D., Hollar, L., Spalovska, I., Takac, P., Zitnanova, I., Gill, S.S., and Adams, M.E. (2002). Molecular cloning and function of ecdysis-triggering hormones in the silkworm Bombyx mori.

J. Exp. Biol.205, 3459–3473.

33. Adams, M.E., and Zitnan, D. (1997). Identification of ecdysis-trig- gering hormone in the silkworm Bombyx mori. Biochem. Bio- phys. Res. Commun.230, 188–191.

34. Kingan, T.G., Gray, W., Zitnan, D., and Adams, M.E. (1997). Reg- ulation of ecdysis-triggering hormone release by eclosion hor- mone. J. Exp. Biol.200, 3245–3256.

35. Gammie, S.C., and Truman, J.W. (1997). Neuropeptide hierar- chies and the activation of sequential motor behaviors in the hawkmoth, Manduca sexta. J. Neurosci.17, 4389–4397.

36. McNabb, S.L., Baker, J.D., Agapite, J., Steller, H., Riddiford, L.M., and Truman, J.W. (1997). Disruption of a behavioral se- quence by targeted death of peptidergic neurons in Drosophila.

Neuron19, 813–823.

37. Ewer, J., and Truman, J.W. (1996). Increases in cyclic 30, 50-gua- nosine monophosphate (cGMP) occur at ecdysis in an evolu- tionarily conserved crustacean cardioactive peptide-immunore- active insect neuronal network. J. Comp. Neurol.370, 330–341.

38. Mendive, F.M., Van Loy, T., Claeysen, S., Poels, J., Williamson, M., Hauser, F., Grimmelikhuijzen, C.J., Vassart, G., and Vanden Broeck, J. (2005). Drosophila molting neurohormone bursicon is a heterodimer and the natural agonist of the orphan receptor DLGR2. FEBS Lett.579, 2171–2176.

39. Nambu, J.R., Murphy-Erdosh, C., Andrews, P.C., Feistner, G.J., and Scheller, R.H. (1988). Isolation and characterization of a Dro- sophila neuropeptide gene. Neuron1, 55–61.

40. Schneider, L.E., and Taghert, P.H. (1988). Isolation and charac- terization of a Drosophila gene that encodes multiple neuropep- tides related to Phe-Met-Arg-Phe-NH2 (FMRFamide). Proc. Natl.

Acad. Sci. USA85, 1993–1997.

41. Hewes, R.S., Snowdeal, E.C., 3rd, Saitoe, M., and Taghert, P.H.

(1998). Functional redundancy of FMRFamide-related peptides at the Drosophila larval neuromuscular junction. J. Neurosci.

18, 7138–7151.

42. Holman, G.M., Hachman, R.J., Schoofs, L., Hayes, T.K., Wright, M.S., and DeLoof, A. (1989). The Leucophaea maderae hindgut preparation: A rapid and sensitive bioassay tool for the isolation of insect myotropins of other insect species. Insect Biochem.

21, 107–121.

43. Hayes, T.K., Pannabecker, T.L., Hinckley, D.J., Holman, G.M., Nachman, R.J., Petzel, D.H., and Beyenbach, K.W. (1989). Leu- cokinins, a new family of ion transport stimulators and inhibitors in insect Malpighian tubules. Life Sci.44, 1259–1266.

44. Davis, N.T., Blackburn, M.B., Golubeva, E.G., and Hildebrand, J.G. (2003). Localization of myoinhibitory peptide immunoreac- tivity in Manduca sexta and Bombyx mori, with indications that the peptide has a role in molting and ecdysis. J. Exp. Biol.

206, 1449–1460.

45. Novicki, A., and Weeks, J.C. (1996). The initiation of pre-ecdysis and ecdysis behaviors in larval Manduca sexta: The roles of the brain, terminal ganglion and eclosion hormone. J. Exp. Biol.199, 1757–1769.

46. Fuse, M., and Truman, J.W. (2002). Modulation of ecdysis in the moth Manduca sexta: The roles of the suboesophageal and tho- racic ganglia. J. Exp. Biol.205, 1047–1058.

47. Adams, M.E. (2004). Agatoxins: Ion channel specific toxins from the American funnel web spider, Agelenopsis aperta. Toxicon 43, 509–525.

48. Kim, Y.J., Spalovska-Valachova, I., Cho, K.H., Zitnanova, I., Park, Y., Adams, M.E., and Zitnan, D. (2004). Corazonin receptor

signaling in ecdysis initiation. Proc. Natl. Acad. Sci. USA101, 6704–6709.

49. Marder, E., and Thirumalai, V. (2002). Cellular, synaptic and net- work effects of neuromodulation. Neural Netw.15, 479–493.

50. Bullock, T.H., Bennett, M.V., Johnston, D., Josephson, R., Marder, E., and Fields, R.D. (2005). Neuroscience. The neuron doctrine, redux. Science310, 791–793.

51. Suster, M.L., Martin, J.R., Sung, C., and Robinow, S. (2003). Tar- geted expression of tetanus toxin reveals sets of neurons involved in larval locomotion in Drosophila. J. Neurobiol.55, 233–246.

52. Hewes, R.S., Park, D., Gauthier, S.A., Schaefer, A.M., and Taghert, P.H. (2003). The bHLH protein Dimmed controls neuro- endocrine cell differentiation in Drosophila. Development130, 1771–1781.

Referenzen

ÄHNLICHE DOKUMENTE

Adult human neural crest-derived stem cells from the inferior turbinate (ITSCs) are able to efficiently differentiate into glutamatergic neurons.. A: Schematic view of the

(2013), principal neurons in the PrS can be classified into 3 major classes, generally conform to neurons of the periarchicortex like the entorhinal cortex and less resembling

When different individuals from the same mouse line were compared, varying total numbers of Lsi1 or Lsi2 mGFP-positive GCs (or pyramidal neurons) did not affect the fractions

(A) Mean responses of VS neurons (N = 4) to motion in the preferred direction when recorded in bridge mode (gray) and when recorded in bridge mode before and after pattern motion

To determine the preferred rotation axes of the different VS neurons when stimulated with naturalistic optic flow, we cal- culated for each neuron the coherence rate between the

1. The responses of all three mechanosensory cell types, innervating the ventral area of the skin, were investigated. The list of response features were extended

Figure 6.7A shows the ratio between the information rate estimate of the neural network and the linear method as a function of the number of neurons in the population.. B) Ratio

Specifically, we characterize how distinct types of adaptation currents affect (i) spike rates, interspike interval variability and phase response properties of single neurons,