distribution and modulatory roles of neuropeptides and …352353/... · 2010. 9. 20. · 5...

46
Distribution and modulatory roles of neuropeptides and neurotransmitters in the Drosophila brain Lily Kahsai Tesfai Department of Zoology Stockholm University Stockholm 2010

Upload: others

Post on 31-Jan-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

  • Distribution and modulatory roles of neuropeptides and

    neurotransmitters in the Drosophila brain

    Lily Kahsai Tesfai

    Department of Zoology

    Stockholm University

    Stockholm 2010

  • 2

    © Lily Kahsai Tesfai, Stockholm 2010 Doctoral dissertation 2010 ISBN 978-91-7447-150-2 Printed in Sweden by universitetsservice US-AB, Stockholm, 2010

    The background of the cover is a series of confocal images taken by Lily Kahsai Tesfai. Immunolbeling of 210yGAL4 (red) with serotonin antibody (green) showing from anterior to posterior, the ellipsoid body, fan-shaped body, protocerebral bridge and large neurosecretory cells as well as small interneurons in the adult Drosophila brain.

  • 3

    To

    my family

  • 4

  • 5

    Distribution and modulatory roles of neuropeptides and neurotransmitters in the Drosophila brain

    Doctoral Dissertation 2010

    Lily Kahsai Tesfai Department of Zoology Stockholm University

    106 91 Stockholm

    ABSTRACT The central complex is a prominent neuropil found in the middle of the insect brain. It is considered as a higher center for motor control and information processing. Multiple neuropeptides and neurotransmitters are produced in neurons of the central complex, however, distribution patterns and functional roles of signaling substances in this brain region are poorly known. Thus, this thesis focuses on the distribution of signaling substances and on modulatory roles of neuropeptides in the central complex of Drosophila.

    Immunocytochemistry in combination with GAL4/UAS technique was used to visualize various signaling substances in the central complex. We revealed different central-complex neurons expressing the neuropeptides; Drosophila tachykinin (DTK), short neuropeptide F (sNPF), myoinhibitory peptide (MIP), allatostatin A, proctolin, SIFamide, neuropeptide F and FMRFamide. Subpopulations of DTK, sNPF and MIP-expressing neurons were found to co-localize a marker for acetylcholine. In addition, five metabotropic neurotransmitter receptors were found to be expressed in distinct patterns. Comparison of receptor/ligand distributions revealed a close match in most of the structures studied.

    By using a video-tracking assay, peptidergic modulation of locomotor behavior was studied. Different DTK and sNPF-expressing neurons innervating the central complex were revealed to modulate spatial distribution, number of activity-rest phases and activity levels, suggesting circuit dependent modulation.

    Furthermore, neurosecretory cells in the Drosophila brain that co-express three types of neuropeptides were shown to modulate stress responses to desiccation and starvation.

    In summary, we have studied two different neuropeptides (DTK and sNPF) expressed in interneuronal circuits and neurosecretory cells of the Drosophila brain in more detail. We found that these neuropeptides display multiple actions as neuromodulators and circulating hormones, and that their actions depend on where they are released.

  • 6

    TABLE OF CONTENTS

    LIST OF PAPERS 7

    INTRODUCTION 8

    The central complex of Drosophila melanogaster 8

    Biosynthesis and processing of neuropeptides and neurotransmitters 11

    Neuropeptides and neurotransmitter receptors 12

    Tachykinins related peptides (TKRPs) 13

    Short neuropeptide F (sNPF) 15

    Classical neurotransmitters 16

    Acetylcholine 16 γ-aminobutyric acid 17 Glutamate 18

    Biogenic amine neurotransmitters 18 Serotonin 18 Dopamine 19 Tyramine and octopamine 20

    AIMS OF THE THESIS 21

    METHODS 22

    The GAL4/UAS system 22 Immunocytochemistry and antisera 23 Image analysis 25 Quantification of immunofluorescence 25 Locomotor behavior assay and data analysis 25 Conditional neuronal silencing experiment using UAS-shibire 26 Trikinetics locomotor activity 26 Assays of survival during starvation and desiccation 26 Measurement of water content and loss 27 RESULTS AND DISCUSSION 27 Neuropeptide and neurotransmitter signaling in the central complex: functions and distribution (paper I, II, III) 27 Peptidergic modulation of metabolic stress response (paper IV) 31 SUMMARY AND CONCLUSION 32

    REFERENCES 33

    ACKNOWLEDGEMENTS 44

  • 7

    LIST OF PAPERS

    This thesis is based on the following papers

    Paper I Kahsai L, Martin JR, Winther ÅME. 2010. Neuropeptides in the

    Drosophila central complex in modulation of locomotor behavior.

    J Exp Biol 213(Pt 13):2256-2265.

    Paper II Kahsai L and Winther ÅME. 2010. Chemical neuroanatomy of the

    Drosophila central complex: distribution of multiple neuropeptides

    in relation to neurotransmitters. J Comp Neurol. (In press).

    Paper III Kahsai L, Nässel DR and Winther ÅME. 2010. Distribution of

    metabotropic receptors of serotonin, dopamine, GABA and

    glutamate in the central complex of Drosophila. (Manuscript)

    Paper IV Kahsai L, Kapan N, Dircksen H, Winther ÅME, Nässel DR. 2010.

    Metabolic Stress responses in Drosophila are modulated by brain

    neurosecretory cells that produce multiple neuropeptides. PLoS

    One 5(7):e11480.

    Publication not includes in this thesis

    Dircksen H, Tesfai LK*, Albus C, Nässel DR. 2008. Ion transport peptide splice forms in

    central and peripheral neurons throughout postembryogenesis of Drosophila

    melanogaster. J Comp Neurol 509(1):23-41.

    * Tesfai LK is the same author as Kahsai L

  • 8

    INTRODUCTION

    The nervous system consists of two parts, the central nervous system (CNS) and the

    peripheral nervous system (PNS). In insects, the CNS is composed of the brain and the

    ventral nerve cord that is analogous to the vertebrate spinal chord. The insect brain is

    made of three regions, the protocerebrum, deutocerebrum and tritocerebrum. The

    protocerebrum receives inputs from the compound eyes and other organs, processes

    information and relays signals. It contains centers, or neuropils, where synaptic

    connections are made between axon terminals and dendrites. Neuropil structures such as

    the central complex and the mushroom body are located in the protocerebrum. In

    addition, the protocerebrum contains neurosecretory cells that project to the neurohemal

    organs, the corpora cardiaca and corpora allata. The deutocerebrum receive inputs from

    the antennae and process olfactory information. Neurons in the tritocerebrum controls the

    mouthparts and innervate the digestive canal (Strausfeld, 1976)

    Signal transmission between neurons in the nervous system is mainly performed

    by different chemical substances such as neuropeptides and various types of

    neurotransmitters. Neuropeptides and neurotransmitters are found in the whole animal

    kingdom, starting from one of the simplest levels of organization, the hydra (phylum

    Cnidaria) to one of the most complex systems, the mammals. These signaling substances

    regulate different physiological processes such as locomotion, metabolism, homeostasis,

    reproduction and growth in multicellular organisms.

    The central complex of Drosophila melanogaster

    More than sixty-five years ago Maxwell Power (1943) named a group of structures found

    in the middle of the fruit fly brain as the central complex. The central complex, located

    between the two peduncles of the mushroom body, consists of four interconnected

    neuropil regions: the ellipsoid body (EB), the fan-shaped body (FB), the noduli (NO) and

    the protocerebral bridge (PB). A comprehensive study, based on Golgi silver staining

    analysis, of Hanesch and co-workers (1989) on the neuroarchitecture of the central

    complex in Drosophila has shown at least 50 different neuronal types in this structure

    (Hanesch, 1989).

  • 9

    The doughnut-shaped EB is the most anterior substructure. The FB is the largest

    substructure in the central complex, and is composed of 8 vertical segments and at least 6

    horizontal layers. The two spherical NO are mirror images of each other and lie ventral to

    the FB and posterior to the EB. The PB lies posterior to the FB and is composed of 16

    segments, 8 segments in each brain hemisphere, (Figure 1A). Two paired accessory areas,

    the lateral triangles (LTR) and the ventral bodies (VBO) are closely associated with the

    central complex (Figure 1B). The general design of the central complex, also known as

    central body in other insects, is well conserved. (Homberg, 2008; Loesel et al., 2002).

    The EB in other insects is not circular as in Drosophila but has an arch like structure and

    is called the lower division of the FB (Homberg, 2008)

    Figure 1. (A) The four substructures of the central complex from posterior to anterior,

    protocerebral bridge (pb), the fan-shaped body (fb), the nodule (no) and ellipsoid body

    (eb). (B) Frontal view of the central complex substractures with two paired accessory

    areas, lateral triangle (ltr) and ventral bodies (vbo). Schematic drawing taken from

    (Hanesch, 1989).

    Morphological studies in the fruit fly, Drosophila melanogaster;(Hanesch, 1989; Renn et

    al., 1999; Young and Armstrong, 2010), in the house fly, Musca domestica (Strausfeld,

    1976), and in the locust, Schistocerca gregaria, (Heinze and Homberg, 2008; Homberg,

    1991; Williams, 1975) suggest that the neurons of the central complex can be divided

    into three groups; tangential, columnar and pontine neurons (Fig. 2). Tangential neurons

    typically arborize in a single central-complex substructure, and connect with various

    regions outside the central complex. The majority of the tangential neurons are

    A B

  • 10

    considered to receive input from these different brain regions (Hanesch, 1989; Young and

    Armstrong, 2010). Typical examples of tangential neurons are the ring (R) neurons of the

    EB and the F-neurons of the FB. In dipterans, the R neurons form ring-like arborizations

    around the EB-canal. Columnar neurons interconnect two or more different substructures,

    for example, neurons that connect the FB-EB-NO. Most columnar neurons project to the

    accessory areas or to the NO and are regarded as output pathways (Hanesch, 1989). The

    third group consists of the pontine neurons which are intrinsic neurons of one

    substructure and connect two regions of this substructure, right and left hemisphere or

    dorsal and ventral parts.

    Figure 2. Three examples of different types of central complex neurons. Schematic

    drawing taken from (Hanesch, 1989).

    The central complex receives input from many parts of the brain and is believed to be an

    integration center for motor and sensory functions (Homberg, 2008). Investigations on

    the functional role of the central complex in Drosophila began with behavioral analysis

    of central complex structural mutants. These studies showed the importance of the central

    complex in control of walking behaviors (Strauss and Heisenberg, 1993) and visual flight

    control (Ilius et al., 1994). Later on, using GAL4 directed neuronal disruption or

    combining mutational analysis with GAL4 directed rescue-experiments, other studies

    reported the role of the central complex in coordination of locomotion and in regulation

    of walking activity (Martin et al., 1999; Poeck et al., 2008; Strauss, 2002). Memory

    behaviors such as visual pattern memory (Liu et al., 2006; Pan et al., 2009), spatial

  • 11

    orientation memory (Neuser et al., 2008) and (Wu et al., 2007) have also been associated

    with the central complex. The central complex has also been implicated in courtship

    associated behaviors (Popov et al., 2003; 2004). Furthermore, in locusts and

    grasshoppers, the central complex was shown to play a role in sky compass orientation

    (Heinze and Homberg, 2007; Homberg, 2004; Vitzthum et al., 2002) and in sound

    production (Wenzel et al., 2005).

    Biosynthesis and processing of neuropeptides and neurotransmitters

    Synthesis of neuropeptides occurs within the ribosomes in the neuronal cell body. Gene

    transcription in the nucleus leads to the synthesis of a precursor protein, a pre-propeptide,

    in the rough endoplasmic reticulum. The pre-propeptide is then transported to the Golgi

    apparatus where it is further modified and packaged into large dense-core vesicles (100-

    200 nm in diameter). The mature peptide in vesicles is then transported along the

    microtubules by motor proteins towards the pre-synaptic terminal and vesicles are stored

    in preparation for release. If an action potential reaches the pre-synaptic terminal, voltage

    gated calcium channels are opened resulting in an influx of calcium ions. The rise of

    cytosolic calcium triggers fusion between the vesicle membrane and the plasma

    membrane of the pre-synaptic neuron and causes release of neuropeptides. The

    neuropeptide then binds to a receptor on the post-synaptic neuron and causes a post-

    synaptic effect which can either be an excitatory or inhibitory effect, but more common a

    modulatory effect in the CNS (Nässel, 2002). Neuropeptides can function as

    neuromodulators when released at a short distance, synaptically or non-synaptically, or

    they can also act as neurohormones when released at a longer distance into the blood

    stream (Nässel, 2009; O'Shea and Schaffer, 1985).

    Until now, in Drosophila about 42 confirmed neuropeptide precursor genes have

    been identified (Hewes and Taghert, 2001; Nässel, 2002; Nässel and Winther, 2010).

    Neuropeptides in insects are found expressed in interneurons, neurosecretory neurons,

    and in motor and sensory neurons, suggesting multifunctional roles of peptides. (See

    review by Nässel, 2002; Nässel and Winther, 2010).

  • 12

    Neuropeptides are sometimes found co-localized with classical neurotransmitters

    within the same neuron termination. Classical neurotransmitters include acetylcholine,

    glutamate, γ-aminobutyric acid (GABA) and biogenic amines (such as serotonin,

    dopamine, octopamine, tyramine and histamine). In contrast to neuropeptides that are

    synthesized in the cell bodies, neurotransmitters are synthesized and packaged into small

    synaptic vesicles (50 nm in diameter) within the pre-synaptic terminal. One or more

    biosynthetic enzymes are required for production of the mature neurotransmitter.

    Moreover, removal of a neurotransmitter from the synaptic cleft (to terminate a signal) is

    accomplished by uptake into the pre-synaptic terminal or into glial cells via vesicular

    transporters whereas neuropeptides are removed from the synaptic cleft by enzymatic

    degradation or diffusion (Nässel, 2002).

    Neuropeptides and neurotransmitter receptors

    Signal transduction most commonly occurs through two types of receptors; G-protein

    coupled receptors (GPCRs) also known as metabotropic receptors and ligand-gated ion

    channel receptors (LGICs). GPCRs are characterized by seven transmembrane (7TM)

    protein domains that are connected by three extracellular and three intracellular loops.

    The extracellular N-terminus domain can be glycosylated whereas the intracellular C-

    terminus domain can be phosphorylated. GPCRs are directly coupled to heterotrimeric

    GTP-binding (G) proteins that are composed of three subunits α, β, and γ. G proteins

    exist in two functional states, in an inactive state when the α subunit is bound to GDP and

    an active state when the α subunit is bound to GTP. Activation of a GPCR leads to

    binding of the α subunit to GTP, the α-GTP then disassociates from the βγ subunits.

    Depending on the signal transduction pathway, the Gα-GTP complex activates or inhibits

    one of the effector enzymes; adenylyl cyclase, guanylyl cyclase, phospholipace C or

    calcium channels; to generate second messenger such as cAMP, cGMP inositol 1,4,5-

    triphosphate (IP3) and diacylglycerol and calcium respectively (Lodish et al., 1999).

    LGICs are transmembrane proteins usually made of five subunits with an

    extracellular ligand-binding site and an ion channel that selectively transmits small ions

    such as sodium, potassium, calcium or chlorine. LGICs mediate fast post-synaptic

    effects, usually lasting less than a millisecond. GPCRs on the other hand signal slower

  • 13

    post-synaptic transmission lasting from seconds to hours (Lodish et al., 1999). If a

    receptor is continually stimulated by a signaling substance, the post-synaptic neuron can

    decrease its responsiveness for that ligand; this reaction is known as receptor

    desensitization.

    GPCRs in eukaryotes represent the largest group of cell surface receptors. In

    Drosophila out of 13600 genes about 163 genes are coding for GPCRs which is about 2%

    of the total genome. In the nematode Caenorhabditis elegans, GPCRs constitute about

    6% of the total genome (Brody and Cravchik, 2000; Hauser et al., 2006; Hewes and

    Taghert, 2001). Based on sequence similarity, GPCRs are classified into four families.

    These are: (1) The rhodopsin receptor family that are activated by a wide range of ligands

    such as light, odorants, neurotransmitters, neuromodulators and neuropeptides. (2)

    Secretin receptor family that are mostly receptors for numerous peptide hormones. (3)

    Metabotropic glutamate receptor family, which include receptors for GABA, glutamate

    and calcium ions. (4) Atypical metabotropic receptor family (Brody and Cravchik, 2000;

    Hauser et al., 2006; Hewes and Taghert, 2001).

    Tachykinin-related peptides

    Tachykinin-like peptides are found in both vertebrates and invertebrates (Hökfelt et al.,

    2000; Nässel, 2002). Both vertebrate and invertebrate tachykinins are suggested to be

    multifunctional peptides involved in modulatory processes in the central and the

    peripheral nervous system (Nässel, 2002; Nässel and Winther, 2010). The invertebrate

    tachykinins are divided into two groups, one group share an identical C-terminus amino

    acid sequence (FXGLMamide) as the vertebrate tachykinins. Invertebrate tachykinins of

    this group have only been isolated from the salivary gland of two octopus species,

    Eledone aldovrandi (Anastasi and Erspamer, 1962) and Octopus vulgaris (Kanda et al.,

    2003), and from one mosquito, Aedes aegypti (Champagne and Ribeiro, 1994). The

    second group, known as the tachykinin-related peptides (TKRPs), is only found in

    invertebrates and contains the amino acid sequence (FX1GX2Ramide) at the C- terminus.

    TKRPs have been isolated from the brain and gut of various invertebrate species

    (reviewed in Nässel, 2002).

  • 14

    In Drosophila, the TKRP encoding gene, (dtk) is composed of 289 amino acids

    and encodes six different TKRPs (DTK-1-6) (Siviter et al., 2000). The dtk gene is

    localized on the third chromosome and contains four exons. DTK-1-4 are encoded on

    exon 2 while DTK-5 and DTK-6 (also designated TAP-5) are on exon 3 (Figure 3). Two

    metabotropic tachykinin receptors have been identified in Drosophila; DTKR

    (Drosophila TK receptor) (Li et al., 1991) and NKD (neurokinin receptor from

    Drosophila) (Monnier et al., 1992). Recent report revealed that NKD is activated by

    DTK-6 (Poels et al., 2009) whereas DTKR is suggested to be sensitive to DTK-1-5 (Birse

    et al., 2006). Using immunocytochemistry and in situ hybridization techniques,

    tachykinins in Drosophila have been detected in the optic lobes, antennal lobes, the

    central complex, in the ventral ganglion and in endocrine cells of the midgut (Siviter et

    al., 2000; Winther et al., 2003).

    Figure 3. Organization of the Drosophila tachykinin gene (Taken from Siviter et al,

    2000)

    Studies have shown a possible role of TKRPs as cotransmmiters and neuromodulators

    (see Nässel, 2002). For instance, TKRPs in Drosophila (Ignell et al., 2009) and

    cockroaches (Nässel, 2002) have been found colocalized with the inhibitory

    neurotransmitter GABA in local interneurons of the antennal lobe indicating the action of

    the peptide as cotransmitter and/or neuromodulator in GABA signalling. Moreover,

    TKRPs in crayfish were shown to modulate photoreceptors via presynaptic GABA

    inhibition (Glantz et al., 2000). Several in vitro studies on TKRPs have also revealed

    myostimilatory action of the peptide on gut muscles in Drosophila, regulation of AKH

    release and induction of secretion in Malphigian tubules of moths and locusts (Nässel et

    al., 1995). See also reviews (Van Loy, 2009). However, recently in vivo studies

  • 15

    employing tachykinin deficient flies have shown that tachykinins have a role in

    modulating locomotion (Winther et al., 2006) and olfactory processing (Ignell et al.,

    2009; Winther et al., 2006; Winther and Ignell, 2010)

    Short neuropeptide F (sNPF)

    Peptides in invertebrates designated neuropeptide F are structurally related to the

    vertebrate neuropeptide Y (NPY) (Brown et al., 1999; De Jong-Brink et al., 2001). The

    name NPF and NPY refers to the amino acid phenylalanine (F) and tyrosine (Y) at the C-

    terminal end, respectively. In Drosophila, distantly related or totally unrelated NPF-like

    peptides known as short NPFs (sNPF) have been identified (De Loof et al., 2001; Nässel,

    2002; Vanden Broeck, 2001). The sNPF gene in Drosophila encodes a precursor protein

    which is predicted to consist of four different peptides (sNPF1-4) (Hewes and Taghert,

    2001; Vanden Broeck, 2001). sNPF-1 and sNPF-2 contain RLRFa at their C- terminal

    whereas sNPF-3 and sNPF-4 have RLRWa at the C-terminal end. So far, only one

    metabotropic sNPF receptor (sNPF-R) has been identified (Feng et al., 2003; Mertens et

    al., 2002; Reale et al., 2004).

    sNPF is found widely expressed in numerous small interneurons in the CNS, in

    the mushroom body, in the central complex, in chemosensory neurons of the antennal

    lobe as well as in neurosecretory cells and in certain clock neurons(Johard et al., 2008;

    Johard et al., 2009; Nässel et al., 2008). Studies have shown a role of sNPF in regulation

    of feeding and growth via the insulin pathway (Lee et al., 2009; Lee et al., 2008; Lee et

    al., 2004). In addition, a possible neuroendocrine function of sNPF has been reported

    after the identification of sNPF1-3 in the hemolymph of Drosophila (Garczynski et al.,

    2006) as well as sNPF immunoreactivity in gut endocrine cells (Veenstra et al., 2008).

  • 16

    Classical neurotransmitters

    Acetylcholine

    Acetylcholine is the principal excitatory neurotransmitters in the insect CNS. It is

    synthesized from the compounds choline and acetyl-CoA by the enzyme choline acetyl

    transferase (ChAT). Immunoreactivity against ChAT in Drosophila has revealed an

    abundant and widespread distribution in the nervous system including the calyces of the

    mushroom bodies, antennal lobes, optic lobes, suboesophageal ganglia and in the central

    complex (Buchner et al., 1986; Yasuyama and Salvaterra, 1999). By employing

    Drosophila mutants for the enzyme that degrades acetylcholine (acetylcholinesterase) the

    importance of acetylcholine in numerous behaviors, such as optomotor behavior and

    courtship, was shown (Greenspan et al., 1980).

    Acetylcholine receptors are divided into ionotrpic nicotinic acetylcholine

    receptors (nAChRs) and metabotropic muscarinic acetylcholine receptors (mAChRs).

    nAChRs are activated by nicotine and blocked by α-bungarotoxin, a snake poison,

    whereas mAChR is activated by the mushroom alkaloid muscarine and blocked by

    atropin. Five mAChR subtypes (M1-M5) has been identified in mammals (Eglen et al.,

    2001). M1, M3 and M5 are coupled to the stimulatory G protein (Gq) thus, stimulating

    phospholipase C whereas M2 and M4 receptors are coupled to the inhibitory G protein,

    (Gi), inhibiting adenylyl cyclase. In Drosophila, only one mAchR gene (Dm1) with

    similar paharmacological profile to the mammalian M3 and M1 subtypes has been

    identified (Shapiro et al., 1989). mAchR in Drosophila is expressed in the whole nervous

    system with strong expression in the antennal lobe glomeruli (Blake et al., 1993). The

    Drosophila genome predicts 10 genes that express nAChRs (Sattelle et al., 2005). Of

    these seven genes are encoding α subunits; Dα1 or ALS (alpha-like subunit), Dα2 or

    SAD (second alpha-like subunit), Dα3-Dα7 and the remaining three genes are coding for

    non α subunits, Dβ1-Dβ3. Using immunocytochemistry and in situ hybridization Dα1-

    Dα4 and Dβ1 and Dβ2 have been shown to be expressed only in the nervous system

    (Kopczynski et al., 1998; Lansdell and Millar, 2000). The expression pattern for the rest

    of nAChR subunits is not yet described.

  • 17

    γ-aminobutyric acid

    γ-Aminobutyric acid (GABA) is synthesized from glutamate through the biosynthetic

    enzyme glutamic acid decarboxyalse (GAD). GABA is the major inhibitory

    neurotransmitter in both vertebrates and insects. Immunocytochemical studies using

    antibodies against GABA and GABA signaling components, such as GAD and vesicular

    GABA transporter (vGAT), have shown GABA producing neurons in the nervous

    system. GABA is expressed in major neuropil regions such as the anntenal lobe, the optic

    lobe, the mushroom body calyces and in some substructures in the central complex.

    (Enell et al., 2007; Fei et al., 2010; Hamasaka et al., 2005; Kolodziejczyk et al., 2008;

    Okada et al., 2009).

    In Drosophila, GABA activates two types of GABA receptors known as

    ionotropic GABAA receptors and metabotropic GABAB receptors. The GABAA receptors

    are ligand gated chloride channels that are blocked by picrotoxin. The GABAA receptors

    can be formed by three subunits, the RDL (for resistance to dieldrin) (Buckingham et al.,

    2005; French-Constant et al., 1993), LCCH3 (ligand-gated chloride channel homologue

    3) (Zhang et al., 1995) and Grd (GABA and glycine-like receptor of Drosophila) (Harvey

    et al., 1994). Immunocytochemical studies have localized RDL mostly in neuropil

    structures such as the optic lobes, the antennal lobes, the mushroom bodies and in the

    central complex in the adult Drosophila brain (Enell et al., 2007; Harrison et al., 1996).

    LCCH3 was found mostly distributed in neuronal cell bodies in the adult nervous system

    (Harrison et al., 1996). The distribution of Grd has not yet been mapped in Drosophila.

    RDL has been shown to be significant for olfactory learning (Liu et al., 2009; Liu et al.,

    2007).

    Three types of GABAB receptors exist in Drosophila, GABABR1-3 (Mezler et al.,

    2001). GABABR1 and GABABR2 hetrodimerize to mediate their synaptic transmission

    via Go/Gi protein (Kaupmann et al., 1998; Mezler et al., 2001). Based on in situ

    hybridization in the Drosophila embryo, GABABR1 and GABABR2 have been shown to

    have similar expression patterns whereas the GABABR3 displayed a unique distribution

    pattern and different pharmacology (Mezler et al., 2001). The GABABR protein is found

    distributed in the adult nervous system, most abundantly in the mushroom body calyces,

    central complex, optic lobe and antennal lobe (Enell et al., 2007; Hamasaka et al., 2005;

  • 18

    Kolodziejczyk et al., 2008; Okada et al., 2009; Root et al., 2008). The distribution of

    GABABR3 is not known in Drosophila. GABA signaling via GABAB receptors has been

    shown to be involved in a wide variety of behaviors such as olfaction (Olsen and Wilson,

    2008; Root et al., 2008; Wilson and Laurent, 2005), locomotion and ethanol tolerance

    (Dzitoyeva et al., 2003), circadian behavior (Hamasaka et al., 2005) and in development

    (Dzitoyeva et al., 2005).

    Glutamate

    Glutamate is known as one of the principal excitatory neurotransmitters in the vertebrate

    brain and in insect motoneurons. Several studies have concentrated on its role in the

    insect neuromuscular junction (Jan and Jan, 1976). However, more recent studies using

    the vesicular glutamate transporter (vGluT) as a marker in Drosophila have localized

    glutamate in sensory neurons, interneurons as well as in motoneurons in Drosophila

    (Daniels et al., 2008; Mahr and Aberle, 2006).

    Drosophila express both ionotropic and metabotropic glutamate receptors. Until

    now, out of the 30 predicted ionotropic receptor types encoded in the Drosophila genome

    only 9 have been cloned (Littleton and Ganetzky, 2000). Six of the nine cloned receptors

    are expressed in NMJ (Qin et al., 2005) and the rest of the NMDA-like receptors are

    found distributed in the Drosophila brain (Xia et al., 2005). The only functional

    metabotropic receptor known in Drosophila is DmGluRA (Drosophila melanogaster

    glutamate receptor A) (Parmentier et al., 1996). DmGluRA is expressed in most neuropil

    regions except the lobes of the mushroom bodies (Devaud et al., 2008). In addition, it is

    expressed in the neuromuscular junction where it modulates excitatory neurotransmission

    (Bogdanik et al., 2004) and in certain clock circuits regulating circadian locomotor

    behavior (Hamasaka et al., 2007).

    Biogenic amine neurotransmitters

    Serotonin

    Serotonin (5-HT) is synthesized from the amino acid tryptophan by two enzymes:

    tryptophan hydroxylase (TRH) and dopa decarboxylase (Ddc). By using an antibody

    against 5-HT, the distribution of 5-HT in the developing and adult Drosophila has been

  • 19

    mapped (Nässel, 1988; Valles and White, 1988). 5-HT in Drosophila has been reported

    to play an important role in aggression (Alekseyenko et al., 2010; Dierick and Greenspan,

    2007; Johnson et al., 2009) and in learning and memory (Sitaraman et al., 2008).

    Both ionotropic and metabotropic 5-HT receptors exist in mammals. However in

    insects, only metabotropic 5-HT receptors have been identified (Tierney, 2001). In

    Drosophila four types of 5-HT receptors that exhibit sequence similarity to the

    mammalian 5-HT receptors are known; these are 5-HT1A, 5-HT1B (Saudou et al., 1992),

    5-HT7 (Witz et al., 1990) and 5-HT2 receptors (Colas et al., 1995). 5-HT1A and 5-HT1B

    belong to the same family of receptors and inhibit adenylate cyclase while stimulating

    phospholipase C (Saudou et al., 1992) whereas 5-HT7 activates adenylate cyclase (Witz

    et al., 1990). The transduction mechanism of the fourth receptor, 5-HT2, is not clearly

    investigated, however, it is proposed that it is similar to the mammalian 5-HT2 receptor

    that activates phospholipase C (Nichols, 2007). The distribution of 5-HT1A and 5-HT1B

    in the brain of Drosophila has been partially mapped (Yuan et al., 2006; Yuan et al.,

    2005). 5-HT1A is expressed in the mushroom body where it was shown to have a role in

    sleep (Yuan et al., 2006). 5-HT1B is also expressed in the mushroom body and is

    additionally found in certain clock neurons (Yuan et al., 2005). By using 5-HT1B mutant

    flies a function of 5-HT1B in circadian behavior has been reported (Yuan et al., 2005). 5-

    HT2 is expressed mainly in the EB of the central complex and in certain neurons in the

    protocerebrum, and its role in circadian behavior has been shown (Nichols, 2007). In

    addition 5-HT1A and 5-HT2 has been implicated in modulating aggressive behavior in

    Drosophila (Johnson et al., 2009).

    Dopamine

    Dopamine is synthesized from the amino acid tyrosine into its precursor molecule L-

    DOPA by the rate limiting enzyme tyrosine hydroxylase (TH), there after L-DOPA is

    converted into dopamine by the enzyme dopa decarboxyalse (Ddc). Based on

    immunocytochemical studies in the nervous system of Drosophila and in two other flies,

    the distribution pattern of dopamine was found to match with that of TH (Friggi-Grelin et

    al., 2003; Nässel and Elekes, 1992). In the Drosophila brain, using antibodies to TH and

    GAL4 expression analysis, about 600 dopamine expressing neurons, organized in 15 cell

  • 20

    cluster have been reported to innervate all major neuropil structures (Friggi-Grelin et al.,

    2003; Mao and Davis, 2009). Dopamine is involved in behaviors such as locomotion

    (Pendleton et al., 2002), courtship (Liu et al., 2008; Neckameyer, 1998) sleep and arousal

    (Andretic et al., 2005; Foltenyi et al., 2007; Kume et al., 2005).

    Three types of metabotropic dopamine receptors are known in Drosophila. Two

    of these, DopR (also known as dDA1) (Kim et al., 2003) and DopR2 (also known as

    DAMB) (Han et al., 1996) are orthologs of the mammalian D1 receptors. The third one,

    D2R (also known as Dop2R) is the ortholog of the mammalian D2 receptors (Hearn et

    al., 2002). The D1-like receptors are coupled to the stimulatory G protein (Gαs) and the

    D2 like receptors are coupled to the inhibitory G protein (Gαi/o) thereby activating and

    inhibiting adenylate cyclase, respectively. In addition a novel GPCR that is activated by

    both dopamine and ecdysteroids has also been proposed as a fourth likely dopamine

    receptor (Srivastava et al., 2005).

    Tyramine and octopamine

    Tyramine, a precursor to octopamine is synthesized by the enzyme tyrosine

    decarboxyalse (TDC) from the amino acid tyrosine. Octopamine, an end product in the

    synthesis of tyramine is synthesized from tyramine by the enzyme tyramine β-

    hydroxylase (TBH). Immunocytochemical data has revealed about 100 octopamine

    expressing neurons in the Drosophila brain (Busch et al., 2009; Monastirioti, 1999;

    Sinakevitch and Strausfeld, 2006). Subsets of these neurons were shown to innervate the

    central complex (Busch et al., 2009; Sinakevitch and Strausfeld, 2006). Octopamine in

    Drosophila has been shown to be involved in several behaviors such as locomotion

    (Yellman et al., 1997), ovulation and egg laying (Cole et al., 2005; Monastirioti, 2003),

    ethanol tolerance (Scholz et al., 2000), appetitive reinforcement (Schwaerzel et al., 2003),

    cocaine sensitization (Hardie et al., 2007) and aggression (Hoyer et al., 2008). Tyramine

    has recently been implicated in locomotor activity (Saraswati et al., 2004) and cocaine

    sensitization (Hardie et al., 2007).

    Two GPCRs that are activated by both octopamine and tyramine has been cloned

    in Drosophila. One receptor that has a preference to tyramine over octopamine is known

    as tyramine/octopamine receptor (Octyr) (Arakawa et al., 1990). The second receptor

  • 21

    known as OAMB (octopamine receptor in the mushroom body) is activated by both

    octopamine and tyramine but preferentially binds octopamine (Han et al., 1998). Both

    receptors stimulate the production of intracellular calcium signal (Robb et al., 1994),

    however Octyr and OAMB seem to exhibit opposite action towards adenylate cyclase,

    OAMB being the activator and Octyr the inhibitor (Han et al., 1998; Saudou et al., 1990).

    OAMB is found widely distributed in the mushroom body and in the central complex

    (Han et al., 1998) in Drosophila. Reports have also revealed the expression of OAMB in

    the thoracic and abdominal ganglion and in the reproductive regions (Lee et al., 2003). A

    mutation in the OAMB gene in Drosophila resulted in female sterility due to defective

    ovulation and egg laying (Lee et al., 2003).

    AIMS OF THE THESIS

    The main objective of this thesis is to understand the roles of brain neuropeptides by

    using the central complex of Drosophila melanogaster as a model to study interneuronal

    peptidergic circuits. Moreover, as an example of a different type of peptide expressing

    cell we studied the roles of neuropeptides in brain neurosecretory cells.

    Immunocytochemical studies have shown that various types of neuropeptides are

    expressed in the central complex of different insects. However, the neuromodulatory

    roles of these neuropeptides in the central complex are not known. Thus, we investigated

    functional roles of two neuropeptides, DTK and sNPF in the central complex of

    Drosophila. The central complex has been shown to be involved in the regulation of

    locomotion and therefore, we examined the functions of DTK and sNPF in aspects of

    locomotor behavior in Drosophila (paper I). In paper II, the objective was to provide a

    detailed map of the distribution of neuropeptides in relation to other signaling substances,

    as well as to describe the chemical neuroanatomy of the central complex of Drosophila.

    Since signaling substances mediate their action through specific receptors, I wanted to

    explore the correlation between neurotransmitters and GPCRs in the central complex

    (paper III). In our final project (paper IV), I was interested in studying additional

  • 22

    functions of neuropeptides in the nervous system. Thus, we examined the hormonal roles

    of DTK and sNPF in metabolic stress responses in Drosophila.

    METHODS

    The GAL4/UAS system

    In all four studies we used the GAL4-UAS technique (developed by (Brand, 1993 #819).

    This technique was used in combination with immunocytochemistry, RNA interference

    (RNAi) and the temperature sensitive allele of shibire (shits1). The GAL4/UAS technique

    involves two transgenic fly lines. One fly line express a yeast Gal4 gene, encoding for the

    transcriptional activator protein GAL4, cloned downstream of an endogenous Drosophila

    promoter. The second fly line express a yeast-specific regulatory sequence known as

    upstream activating sequence (UAS) fused to a gene of interest. When these two

    transgenic flies are crossed, the GAL4 protein (activator protein) binds to the UAS (target

    sequence) and induces transcription of the gene of interest in the progeny of the cross.

    We exploited the GAL4/UAS technique to drive the expression of green

    fluorescent protein in order to visualize specific neurons and then label these neurons

    with different antibodies. We also used the GAL4 to direct the expression of UAS-RNAi

    into specific cells and thereby degrading the mRNA of the target gene. The RNAi

    pathway in the cell is first triggered by the presence of double stranded RNA (dsRNA)

    whose sequence is complementary to the target mRNA. The dsRNA is cleaved by the

    enzyme DICER into small interfering RNA (siRNA) (21-23 nucleotides). The siRNAs

    are then assembled into multiprotein complexes known as RNA-induced silencing

    complexes (RISCs). The siRNAs binds to the complementary target mRNA and directs

    gene silencing. The catalytic component of the RISC complex cleave the target mRNA

    and prevent translation. The GAL4/UAS technique was also used to drive the expression

    of a UAS-shits1 transgene in specific neurons to block synaptic transmission (Kitamoto,

    2001). Shits1 encodes a temperature-sensitive mutation of dynamin that is essential for

    synaptic vesicle recycling. Dynamin is inactivated at a restrictive temperature (>29°C)

  • 23

    resulting in a rapid and reversible inhibition of synaptic transmission in the specific

    neurons where GAL4 is expressed.

    Immunocytochemistry and antisera

    Adult Drosophila heads aged 4-6 days old were dissected in 0.01M phosphate -buffered

    saline with 0.5% Triton X-100, pH 7.2 (PBS-Tx) and fixed in ice-cold 4%

    paraformaldehyde in 0.1 M sodium phosphate buffer pH 7.4 (PB) for 4 hours. Whole-

    mounted or cryostat sectioned adult Drosophila heads were used for

    immunocytochemistry. Whole mount tissues were incubated in primary antisera for 72

    hrs while sections were incubated between 24-48 hrs at 4°C. See Table 1 for primary

    antisera and their references used in all four papers. For detection of primary antisera

    Cy3-tagged goat anti-rabbit or goat anti-mouse antisera or Cy2-tagged goat anti-mouse or

    goat anti rabbit antisera (Jackson Immuno Reasearch) were used. For each experiment at

    least 5-10 adult brains were analyzed.

  • 24

    Antibody Host Dilution Immunogen Source/characterization

    Ast-A Rabbit polyclonal 1:5000 APSGAQRLYGFGLa (Vitzthum et al., 1996a)

    FMRFa Rabbit polyclonal 1:1000 GAQATTTQDGSVEQDQPPG (Chin et al., 1990)

    MIP Rabbit polyclonal 1:4000 GWQDLQGGWa (Predel et al., 2001)

    Proctolin Rabbit polyclonal 1:1000 RYLPT (Taylor et al., 2004)

    SIFa Rabbit polyclonal 1:4000 AYRKPPFNGSIFa (Terhzaz et al., 2007)

    sNPF Rabbit polyclonal 1:4000 DPSLPQMRRTAYDDLLEREL (Johard et al., 2008)

    ITP Rabbit polyclonal 1:1000 GGGDEEEKFNQ (Ring et al., 1998)

    TK Rabbit polyclonal 1:2000 APSGFLGVRa (Winther and Nassel, 2001)

    ChAT Mouse monoclonal 1:500 Choline acetyltransferase

    (recombinant)

    University of Iowa Hybridoma

    Bank, Iowa City, IA

    TH Mouse monoclonal 1:250 Full sequence of rat TH Immunostar, Hudson, WI

    5-HT Mouse monoclonal 1:50 5-HT DAKO, Glostrup, Denmark

    DopR Mouse polyclonal 1:200 (Kim et al., 2003)

    GABABR2 Rabbit polyclonal 1:16000 CLNDDIVRLSAPPVRREMPS (Hamasaka et al., 2005)

    DmGluRA Mouse monoclonal 1:10 (Eroglu et al., 2002; Panneels et

    al., 2003)

    nc-82 Mouse monoclonal 1:10 Drosophila head homogenate University of Iowa Hybridoma

    Bank, Iowa City, IA

    Table 1. Primary antisera used for immunocytochemistry (Paper I-IV)

  • 25

    Image analysis

    Specimens were imaged on a Zeiss laser scanning microscope (LSM 510 META, Zeiss)

    based on Axiovert S100 microscope (Jena, Germany). An argon2/488nm and HeNe

    543nm laser were used for scanning inages. The images were obtained at an optical

    section thickness of 0.5-2 µm and 20x, 40x and 63x oil immersion objective lenses were

    used.The images were processed with Zeiss LSM software and edited for contrast and

    brightness in Adobe Photoshop CS3 Extended version 10.0.

    Quantification of immunofluorescence

    Immunocytochemistry and imaging of specimens were carried out as described above.

    Immunofluorescence was quantified in single optical sections in set regions of interest

    (ROI) using ImageJ.v142. (http://rsb.info.nih.gov/ij). The data were analyzed using Prism

    v4.0 (GraphPad, CA, USA). Three-five brains per genotype and experiment were

    measured.

    Locomotor behavior assay and data analysis

    For paper I, spontaneous walking activity of female flies aged 3-4 days was quantified

    using a video-tracking paradigm described by Martin (2004). Single flies were recorded

    simultaneously at 5Hz sampling rate using the EthoVision software (Noldus, The

    Netherlands) for 7 hours. Experiments were performed at 25°C in 60% relative humidity.

    A total of 20-24 female flies were analyzed for each genotype. Analyses of data were

    done according to Martin (2004). Four parameters were extracted from EthoVision

    software: frequency of passages through a virtual zone located at the centre of the arena

    (2 cm in diameter), total distance moved (mm), mean walking speed (mm/sec) and

    number of episodes of activity and inactivity (start/stop). A fly was considered to be

    active if it had an activity level of or above 4 mm/sec and considered inactive if it had an

    activity level below 2 mm/sec. In order to account for differences in overall activity, the

    frequency of passages in the central zone was normalized to the duration of movements

    (sec). Statistical analysis were made using analysis of variance (ANOVA) and Tukey’s

    post hoc test; Statistica version 4 (StatSoft, Inc.).

  • 26

    Conditional neuronal silencing experiment using UAS-shibire

    Adult female flies aged 3-4 days were analyzed for total distance moved (mm), start/stop,

    mean walking speed (mm/sec) and frequency in zone/sec using the video-tracking

    paradigm as described above. To silence specific neuronal circuits UAS-shits1 was used at

    restrictive temperature (30°C) (Kitamoto, 2001). Prior to monitoring the locomotor

    activity, control flies and flies expressing shits1 were pre-warmed at 30°C for 1 hour and

    then video-recorded at 30°C for 5 hours. All flies for this experiment were raised at 19 °C

    during the entire development. A total of 24 female flies were analyzed for each

    genotype.

    Trikinetics locomotor activity

    In paper IV, we measured the activity of starved flies. Male flies aged 4-6 days were

    analyzed using Trikinetics activity monitor (Trikinetics, Brandeis CA USA). Single flies were placed in individual 5 mm monitoring glass tubes filled in one end with 2 cm of

    0.5% aqueous agarose. Tubes were placed in Trikinetic monitoring racks in an incubator

    at 25o C with a light-dark cycle of 12:12h. All monitoring of activity started 3h after onset

    of starvation (recordings started at 5.5h after lights on) and activity data (crossings of an

    infrared beam) were collected by the Trikintics computer software every 15 minutes.

    Data were collected for 40 h and activity records from a minimum of 64 individual flies

    for each genotype were pooled to obtain average activity levels. The data were analyzed

    using Microsoft Excel Assays of survival during starvation and desiccation

    Starvation experiments were performed according to the protocol of Lee and Park (2004).

    In brief, male flies, aged 4-6 days, were anesthetized on ice and then placed individually

    in 2 ml cotton-capped glass vials containing 500 µl of 0.5% aqueous agarose. All vials

    were placed in an incubator with 12:12 light:dark conditions at 25 °C. The vials were

    checked for dead flies every 12 hours until no living flies were left. For desiccation

    experiments the same protocol was followed, except that the vials were empty and the

    vials were checked every hour until there were no living flies.

  • 27

    Measurement of water content and loss

    Total water content of desiccated and non-desiccated male flies was determined. Flies

    were weighed before and after drying the flies. Two groups of flies of each genotype

    (experimental and controls) were weighed: one group with fed flies (0 hour desiccation)

    and the second group after 16 hours of desiccation without food and water. Groups of 5

    male flies were weighed (on a Sartorius, Göttingen, Germany) after anesthetizing them

    on ice (wet weight), and then dried at 60°C for 24 hours. Dried flies were weighed after

    reaching room temperature (dry weight). Thereafter the water content of each genotype

    was calculated by subtracting the wet weight from the dry weight.Since we had to use

    ded dry weight it was necessary to use separate flies for 0h and 16h.

    RESULTS AND DISCUSSION

    Neuropeptide and neurotransmitter signaling in the central complex: functions and

    distribution (paper I, II, III)

    In paper I, we investigated the roles of two neuropeptides, DTK and sNPF, that were

    found to be abundantly expressed in the FB. By using immunocytochemistry in

    combination with different FB-GAL4 lines, we first identified three types of TK-

    expressing neurons: columnar, tangential and pontine neurons. In addition, we identified

    sNPF-expressing tangential neurons that project to the FB. Then, using RNAi, we down-

    regulated the levels of DTK and sNPF in these neurons and measured the locomotor

    activity of the flies using the video-tracking locomotor behavior assay. We revealed the

    involvement of DTK-expressing tangential and columnar neurons in modulation of

    spatial orientation and activity-rest phases. On the other hand, sNPF-expressing

    tangential neurons appeared to regulate in the fine-tuning of mean walking speed and

    distance walked.

    Furthermore, by employing UAS-shits1 to block synaptic transmission in these

    neurons we confirmed that tangential and columnar DTK-expressing neurons are

    involved in modulation of spatial distribution and activity phases. Although shi has been

    reported to perturb behaviors when expressed in peptidergic neurons, it is not known if or

  • 28

    how shi-dependant vesicle endocytosis affects peptide signaling (Wülbeck et al., 2009).

    Certainly it will block the effect of any co-localized classical transmitter. In paper II, we

    were interested in investigating the relation between neuropeptides and other

    neurotransmitters in the central complex of Drosophila since studies in mammals (see

    review by (Burnstock, 2004; Hökfelt et al., 2000) and insects (see review by (Nässel,

    2002) have suggested that neuropeptides mainly act as co-transmitters. Therefore, we

    compared the distribution patterns of DTK, in relation to different neurotransmitters.

    Interestingly we revealed a population of DTK expressing neurons co-localizing with a

    marker for the excitatory neurotransmitter acetylcholine. In paper I, we observed

    locomotor phenotypes when shi was expressed in putative peptidergic neurons. Even

    though, we at present do not know if these DTK/acetylcholine producing neurons are

    involved in modulation of locomotor behavior, we may speculate that they do. Based on

    their location in the superior median protocerebrum in the brain, comparing the labeling

    patterns in Paper I and II, this is not unlikely. Thus, the locomotor phenotypes induced by

    shi might be accomplished by interfering with the recycling of synaptic vesicles

    containing acetylcholine.

    When using the same GAL4 that was used to knock-down sNPF levels by RNAi,

    to drive the expression of UAS-shits1, flies displayed different locomotor phenotypes than

    when down regulating the peptide level. One possibility is that shi is interfering with

    receptor internalization rather than with peptide release. This has recently been suggested

    to occur when expressing shits1 in PDF expressing clock neurons in Drosophila (Wülbeck

    et al., 2009). Unfortunately, the distribution of sNPF receptor is not known,

    In paper II, we examined the distribution patterns of neuropeptides in the central

    complex. By combining enhancer trap central complex-GAL4 drivers and

    immunocytochemistry, we described the distribution patterns of eight different

    neuropeptides in the FB. These were: DTK, sNPF, MIP, proctolin, FMRFa, allatostatin

    A, SIFa and NPF. Three of these peptides, MIP, dFMRFa and SIFa were for the first time

    described in the central complex of Drosophila. Each of the peptides displayed distinct

    distribution patterns in different layers of the FB, suggesting different functional roles of

    these peptides in the central complex. For example, DTK and sNPF are found abundantly

  • 29

    expressed in almost all layers of the FB whereas others such as proctolin, FMRFa,

    allatostatin A and SIFa were found to have more restricted distributions.

    In paper II we used a NPF-GAL4 (Wen et al., 2005) to visualize the expression of

    this peptide in the central complex. Double labeling, combining NPF antibody and the

    NPF-GAL4, has shown that the GAL4 drives expression in several additional layers of

    the FB and in the EB, regions that are not labeled by the antibody (Krashes et al., 2009).

    This additional labeling has been suggested to be post-synaptic areas and that the

    antibody recognizes pre-synaptic varicosities only (Krashes et al., 2009). Here, we

    confirmed this using the neuronal markers, neuronal synaptobrevin (nSyb)-GFP (Ito et

    al., 1998) and Drosophila Down syndrome adhesion molecule (Dscam)-GFP (Wang et

    al., 2004) (Fig. 4). UAS-nSyb-GFP is expressed in pre-synaptic sites, where axon

    terminals are localized, whereas (Dscam)-GFP is used to visualize post-synaptic sites (Ito

    et al., 1998; Wang et al., 2004).

  • 30

    Figure 4. Pre-synaptic and post-synaptic expression of dNPF-GAL4. Immunolabeling of

    dNPF-GAL4-UAS-nSyb and dNPF-GAL4-UAS-Dscam with a synaptic marker nc82

    antibody on whole-mounted adult Drosophila brains. (A) Strong GFP expression of

    dNPF-GAL4-UAS-nSyb reveals pre-synaptic sites in a dorsal layer of the FB (arrow).

    Weak labeling could also been seen in central layers. (arrowhead) (B) Labeling of dNPF-

    GAL4-UAS-Dscam shows post-synaptic regions in central layers of the FB (arrow),

    weak labeling was observed in a dorsal layer (arrowhead). Images are projections of

    confocal sections. D, dorsal; C, central; V, ventral. Scale bars 20 µm. The dNPF-GAL4

    fly line was kindly provided by Scott Waddell, University of Massachusetts Medical

    School, UAS-nSyb-gfp and UAS-Dscam-gfp were obtained from the Bloomington

    Drosophila Stockcenter.

    Furthermore, in paper II, using markers to demonstrate signaling with acetylcholine,

    serotonin, dopamine, tyramine and octopamine in the central complex we have shown the

    distribution of these transmitters in the central complex. In paper III, we investigated the

    relation between some of these neurotransmitters and their metabotropic receptors. Thus

    we choose to explore the distribution pattern of five GPCRs, two serotonin (5-HT)

    receptors, 5-HT1B and 5HT-7; dopamine receptor, DopR; GABA receptor, GABABR

  • 31

    and glutamate receptor, DmGluRA. We found all receptors expressed in the EB. DopR,

    GABAB and DmGluRA were found expressed in the FB. GABABR and DmGluRA were

    localized to the PB. Strong expression of DopR and DmGluRA was detected in the NO.

    We found all GPCRs to match well with their ligands in most of the central complex

    structures.

    Peptidergic modulation of metabolic stress response (paper IV)

    We identified five pairs of large neurosecretory cells, designated ipc1 and ipc2a

    neurons that co-express three neuropeptides, DTK, sNPF and ITP, in the protocerebrum

    of the adult Drosophila brain. ITP was first identified by Audsley and colleagues as a

    peptide stimulating ion and water reabsorption in the locust (Audsley et al., 1992).

    Recently, its distribution in the larval, pupal and adult Drosophila brain has been

    described (Dircksen et al., 2008). This study revealed 4 pairs of ITP-immunoreactive

    neurons in the larval brain and 7 pairs in the adult brain, designated as ipc1-4.

    Tachykinin-related peptides in moths and locusts have been suggested to stimulate

    secretion in Malphigian tubules (Johard et al., 2003; Skaer et al., 2002) and AKH release

    in locusts (Nässel et al., 1995). They have also been associated with feeding in locusts

    (Lange, 2001; Winther and Nassel, 2001). Moreover, the DTK receptor DTKR, is

    expressed in cells of the Malpighian tubules (Birse et al., 2006). sNPF is known to be

    involved in regulation of feeding and growth via insulin-like peptides in Drosophila (Lee

    et al., 2009; Lee et al., 2008; Lee et al., 2004). Hence, based on the previous findings of

    the roles of these three neuropeptides, we were interested in examining if DTK and/or

    sNPF are involved in water reabsorption and/or starvation in Drosophila. We

    demonstrate signaling with also wanted to determine whether sNPF and DTK are linked

    to the AKH/insulin pathways since we found axonal terminations containing these

    peptides in neurohemal organs adjacent to AKH and insulin producing cells. To answer

    these questions, we down-regulated the levels of DTK and sNPF in ipc1 and 2a by using

    RNAi and monitored fly survival under desiccation and starvation. Both desiccated and

    starved flies with decreased amount of peptides in ipc1 and 2a exhibited shorter life span

  • 32

    than control flies. Next, by comparing whole body water content in desiccated flies to

    none-desiccated flies, we revealed decreased water retention in DTK-knockdown flies.

    Studies have revealed that ablation of AKH cells in starved flies leads to increased

    locomotor activity in search of food (Isabel et al., 2005; Lee and Park, 2004). Thus, we

    monitored locomotor activity of DTK-knockdown flies under starvation using Trikinetics

    locomotor activity monitors. The transgenic and the controls flies displayed normal

    locomotor activity until the experimental flies started to perish after 18h of starvation,

    suggesting that DTK might not directly affect AKH in Drosophila. The effects of DTK

    and sNPF knockdown in the ipc-1 and 2 neurons produced stress phenotypes opposite to

    those seen after knockdown of insulin production/signaling. Since sNPF has been shown

    to stimulate insulin production (Lee et al., 2008; Lee et al., 2004) we can suggest that the

    ipc-1 and 2 neurons do not regulate insulin signaling. The overall findings indicate that

    the circuits of DTK, sNPF and ITP expressing neurons (ipc1 and ipc2a) are not directly

    linked to AKH or insulin signaling but target other pathways to regulate responses to

    nutritional stress (desiccation and starvation). Perhaps, mapping receptor distribution of

    DTK, sNPF and ITP in the future would reveal the targets for these hormonal peptides. In

    summary, we have localized clusters of cells expressing three neuropeptides with axonal

    terminations in neurohemal organs, and shown that two of these peptides regulate

    nutritional-stress response in Drosophila.

    SUMMARY AND CONCLUSION

    Data presented in this thesis has shown multiple actions of neuropeptides by

    exploring the roles of two types of peptide expressing neurons; interneurons of the central

    complex and neurosecretory cells in the Drosophila brain. The peptides DTK and sNPF,

    expressed in interneurons and in neurosecretory cells, were shown to modulate different

    aspects of locomotor behavior and metabolic stress responses, respectively.

    This study has also explored the co-localization of neuropeptides with other

    neuropeptides and signaling substances. It is interesting to note that out of two classical

    neurotransmitters and four biogenic amines examined, only a marker for acetylcholine

    was found co-localized with DTK, sNPF and MIP in neuronal cell bodies innervating the

    central complex. Coexpression of dFMRFa and a GABA marker was also detected in a

  • 33

    pair of neurons that may innervate the EB. These findings suggests that some GABAergic

    and cholinergic neurons utilize neuropeptides as co-transmitters Certainly, we cannot

    exclude the possibility that some of the GAL4 lines used to identify transmitter

    expression may not be complete in their expression, thus additional co-localization of

    neuropeptides and the signaling substances examined may be present. Co-localization of

    two different neuropeptides in neurons of the insect CNS is not uncommon (see Nässel,

    2006; Nässel, 2010), but here we for the first time demonstrated the co-expression of

    three different neuropeptides, DTK, sNPF and ITP, in specific neurosecretory cells (in

    total five neurosecretory cells in two clusters). Additionally, we studied the distribution

    of metabotropic neurotransmitter receptors in relation to their ligands to explore possible

    sites of transmitter action in the central complex. It is apparent that different serotonin

    and dopamine receptor types display differential distributions in central complex circuits,

    suggesting that each of these monoamines may activate different sets of follower neurons

    whose GPCRs couple to different transduction systems. For glutamate and GABA we

    only explored metabotropic receptors and thus presence of ionotropic receptors for these

    neurotransmitters will further increase signaling complexity. This kind of receptor

    distribution studies can serve to guide future exploration of the functional roles of these

    signaling substances, just as the neuropeptide expression in the central complex guided us

    to investigate peptidergic modulation of locomotor function.

    REFERENCES

    Alekseyenko OV, Lee C, Kravitz EA. 2010. Targeted manipulation of serotonergic neurotransmission affects the escalation of aggression in adult male Drosophila melanogaster. PLoS One 5(5):e10806.

    Anastasi A, Erspamer V. 1962. Occurrence and some properties of eledoisin in extracts of posterior salivary glands of Eledone. Br J Pharmacol Chemother 19:326-336.

    Andretic R, van Swinderen B, Greenspan RJ. 2005. Dopaminergic modulation of arousal in Drosophila. Curr Biol 15(13):1165-1175.

    Arakawa S, Gocayne JD, McCombie WR, Urquhart DA, Hall LM, Fraser CM, Venter JC. 1990. Cloning, localization, and permanent expression of a Drosophila octopamine receptor. Neuron 4(3):343-354.

    Audsley N, McIntosh C, Phillips JE. 1992. Isolation of a neuropeptide from locust corpus cardiacum which influences ileal transport. J Exp Biol 173:261-274.

  • 34

    Birse RT, Johnson EC, Taghert PH, Nässel DR. 2006. Widely distributed Drosophila G-protein-coupled receptor (CG7887) is activated by endogenous tachykinin-related peptides. J Neurobiol 66(1):33-46.

    Blake AD, Anthony NM, Chen HH, Harrison JB, Nathanson NM, Sattelle DB. 1993. Drosophila nervous system muscarinic acetylcholine receptor: transient functional expression and localization by immunocytochemistry. Mol Pharmacol 44(4):716-724.

    Bogdanik L, Mohrmann R, Ramaekers A, Bockaert J, Grau Y, Broadie K, Parmentier ML. 2004. The Drosophila metabotropic glutamate receptor DmGluRA regulates activity-dependent synaptic facilitation and fine synaptic morphology. J Neurosci 24(41):9105-9116.

    Brody T, Cravchik A. 2000. Drosophila melanogaster G protein-coupled receptors. J Cell Biol 150(2):F83-88.

    Brown MR, Crim JW, Arata RC, Cai HN, Chun C, Shen P. 1999. Identification of a Drosophila brain-gut peptide related to the neuropeptide Y family. Peptides 20(9):1035-1042.

    Buchner E, Buchner S, Crawford G, Mason T, Salvaterra PM, Satelle DB. 1986. Choline acetyltransferase-like immunoreactivity in the brain of Drosophila melanogaster. Cell Tissue Res 246:57-62.

    Buckingham SD, Biggin PC, Sattelle BM, Brown LA, Sattelle DB. 2005. Insect GABA Receptors: Splicing, Editing, and Targeting by Antiparasitics and Insecticides. Molecular Pharmacology 68(4):942-951.

    Burnstock G. 2004. Cotransmission. Curr Opin Pharmacol 4(1):47-52. Busch S, Selcho M, Ito K, Tanimoto H. 2009. A map of octopaminergic neurons in the

    Drosophila brain. J Comp Neurol 513(6):643-667. Champagne DE, Ribeiro JM. 1994. Sialokinin I and II: vasodilatory tachykinins from the

    yellow fever mosquito Aedes aegypti. Proc Natl Acad Sci U S A 91(1):138-142. Chin AC, Reynolds ER, Scheller RH. 1990. Organization and expression of the

    Drosophila FMRFamide-related prohormone gene. DNA Cell Biol 9(4):263-271. Colas JF, Launay JM, Kellermann O, Rosay P, Maroteaux L. 1995. Drosophila 5-HT2

    serotonin receptor: coexpression with fushi-tarazu during segmentation. Proc Natl Acad Sci U S A 92(12):5441-5445.

    Cole SH, Carney GE, McClung CA, Willard SS, Taylor BJ, Hirsh J. 2005. Two functional but noncomplementing Drosophila tyrosine decarboxylase genes: distinct roles for neural tyramine and octopamine in female fertility. J Biol Chem 280(15):14948-14955.

    Daniels RW, Gelfand MV, Collins CA, DiAntonio A. 2008. Visualizing glutamatergic cell bodies and synapses in Drosophila larval and adult CNS. J Comp Neurol 508(1):131-152.

    De Jong-Brink M, ter Maat A, Tensen CP. 2001. NPY in invertebrates: molecular answers to altered functions during evolution. Peptides 22(3):309-315.

    De Loof A, Baggerman G, Breuer M, Claeys I, Cerstiaens A, Clynen E, Janssen T, Schoofs L, Vanden Broeck J. 2001. Gonadotropins in insects: an overview. Arch Insect Biochem Physiol 47(3):129-138.

  • 35

    Devaud JM, Clouet-Redt C, Bockaert J, Grau Y, Parmentier ML. 2008. Widespread brain distribution of the Drosophila metabotropic glutamate receptor. Neuroreport 19(3):367-371.

    Dierick HA, Greenspan RJ. 2007. Serotonin and neuropeptide F have opposite modulatory effects on fly aggression. Nat Genet 39(5):678-682.

    Dircksen H, Tesfai LK, Albus C, Nassel DR. 2008. Ion transport peptide splice forms in central and peripheral neurons throughout postembryogenesis of Drosophila melanogaster. J Comp Neurol 509(1):23-41.

    Dzitoyeva S, Dimitrijevic N, Manev H. 2003. Gamma-aminobutyric acid B receptor 1 mediates behavior-impairing actions of alcohol in Drosophila: adult RNA interference and pharmacological evidence. Proc Natl Acad Sci U S A 100(9):5485-5490.

    Dzitoyeva S, Gutnov A, Imbesi M, Dimitrijevic N, Manev H. 2005. Developmental role of GABAB(1) receptors in Drosophila. Brain Res Dev Brain Res 158(1-2):111-114.

    Eglen RM, Choppin A, Watson N. 2001. Therapeutic opportunities from muscarinic receptor research. Trends Pharmacol Sci 22(8):409-414.

    Enell L, Hamasaka Y, Kolodziejczyk A, Nässel DR. 2007. gamma-Aminobutyric acid (GABA) signaling components in Drosophila: immunocytochemical localization of GABA(B) receptors in relation to the GABA(A) receptor subunit RDL and a vesicular GABA transporter. J Comp Neurol 505(1):18-31.

    Eroglu C, Cronet P, Panneels V, Beaufils P, Sinning I. 2002. Functional reconstitution of purified metabotropic glutamate receptor expressed in the fly eye. EMBO Rep 3(5):491-496.

    Fei H, Chow DM, Chen A, Romero-Calderon R, Ong WS, Ackerson LC, Maidment NT, Simpson JH, Frye MA, Krantz DE. 2010. Mutation of the Drosophila vesicular GABA transporter disrupts visual figure detection. J Exp Biol 213(Pt 10):1717-1730.

    Feng G, Reale V, Chatwin H, Kennedy K, Venard R, Ericsson C, Yu K, Evans PD, Hall LM. 2003. Functional characterization of a neuropeptide F-like receptor from Drosophila melanogaster. Eur J Neurosci 18(2):227-238.

    Foltenyi K, Andretic R, Newport JW, Greenspan RJ. 2007. Neurohormonal and neuromodulatory control of sleep in Drosophila. Cold Spring Harb Symp Quant Biol 72:565-571.

    French-Constant RH, Rocheleau TA, Steichen JC, Chalmers AE. 1993. A point mutation in a Drosophila GABA receptor confers insecticide resistance. Nature 363(6428):449-451.

    Friggi-Grelin F, Coulom H, Meller M, Gomez D, Hirsh J, Birman S. 2003. Targeted gene expression in Drosophila dopaminergic cells using regulatory sequences from tyrosine hydroxylase. J Neurobiol 54(4):618-627.

    Garczynski SF, Brown MR, Crim JW. 2006. Structural studies of Drosophila short neuropeptide F: Occurrence and receptor binding activity. Peptides 27(3):575-582.

    Glantz RM, Miller CS, Nassel DR. 2000. Tachykinin-related peptide and GABA-mediated presynaptic inhibition of crayfish photoreceptors. J Neurosci 20(5):1780-1790.

  • 36

    Greenspan RJ, Finn JA, Jr., Hall JC. 1980. Acetylcholinesterase mutants in Drosophila and their effects on the structure and function of the central nervous system. J Comp Neurol 189(4):741-774.

    Hamasaka Y, Rieger D, Parmentier ML, Grau Y, Helfrich-Forster C, Nassel DR. 2007. Glutamate and its metabotropic receptor in Drosophila clock neuron circuits. J Comp Neurol 505(1):32-45.

    Hamasaka Y, Wegener C, Nässel DR. 2005. GABA modulates Drosophila circadian clock neurons via GABAB receptors and decreases in calcium. J Neurobiol 65(3):225-240.

    Han KA, Millar NS, Davis RL. 1998. A novel octopamine receptor with preferential expression in Drosophila mushroom bodies. J Neurosci 18(10):3650-3658.

    Han KA, Millar NS, Grotewiel MS, Davis RL. 1996. DAMB, a novel dopamine receptor expressed specifically in Drosophila mushroom bodies. Neuron 16(6):1127-1135.

    Hanesch UF, K.-F, Heisenberg, M. 1989. Neuronal architecture of the central complex in Drosophila melanogaster. Cell Tissue Res 257:343-366.

    Hardie SL, Zhang JX, Hirsh J. 2007. Trace amines differentially regulate adult locomotor activity, cocaine sensitivity, and female fertility in Drosophila melanogaster. Dev Neurobiol 67(10):1396-1405.

    Harrison JB, Chen HH, Sattelle E, Barker PJ, Huskisson NS, Rauh JJ, Bai D, Sattelle DB. 1996. Immunocytochemical mapping of a C-terminus anti-peptide antibody to the GABA receptor subunit, RDL in the nervous system in Drosophila melanogaster. Cell Tissue Res 284(2):269-278.

    Harvey RJ, Schmitt B, Hermans-Borgmeyer I, Gundelfinger ED, Betz H, Darlison MG. 1994. Sequence of a Drosophila ligand-gated ion-channel polypeptide with an unusual amino-terminal extracellular domain. J Neurochem 62(6):2480-2483.

    Hauser F, Cazzamali G, Williamson M, Blenau W, Grimmelikhuijzen CJ. 2006. A review of neurohormone GPCRs present in the fruitfly Drosophila melanogaster and the honey bee Apis mellifera. Prog Neurobiol 80(1):1-19.

    Hearn MG, Ren Y, McBride EW, Reveillaud I, Beinborn M, Kopin AS. 2002. A Drosophila dopamine 2-like receptor: Molecular characterization and identification of multiple alternatively spliced variants. Proc Natl Acad Sci U S A 99(22):14554-14559.

    Heinze S, Homberg U. 2007. Maplike representation of celestial E-vector orientations in the brain of an insect. Science 315(5814):995-997.

    Heinze S, Homberg U. 2008. Neuroarchitecture of the Central Complex of the Desert Locust: Intrinsic and Columnar Neurons. Journal of Comparative Neurology 511(4):454-478.

    Hewes RS, Taghert PH. 2001. Neuropeptides and neuropeptide receptors in the Drosophila melanogaster genome. Genome Res 11(6):1126-1142.

    Hofbauer A, Ebel T, Waltenspiel B, Oswald P, Chen YC, Halder P, Biskup S, Lewandrowski U, Winkler C, Sickmann A, Buchner S, Buchner E. 2009. The Wuerzburg hybridoma library against Drosophila brain. J Neurogenet 23(1-2):78-91.

    Homberg U. 1991. Neuroarchitecture of the Central Complex in the Brain of the Locust Schistocerca-Gregaria and S-Americana as Revealed by Serotonin Immunocytochemistry. Journal of Comparative Neurology 303(2):245-254.

  • 37

    Homberg U. 2004. In search of the sky compass in the insect brain. Naturwissenschaften 91(5):199-208.

    Homberg U. 2008. Evolution of the central complex in the arthropod brain with respect to the visual system. Arthropod Structure & Development 37(5):347-362.

    Hoyer SC, Eckart A, Herrel A, Zars T, Fischer SA, Hardie SL, Heisenberg M. 2008. Octopamine in male aggression of Drosophila. Curr Biol 18(3):159-167.

    Hökfelt T, Broberger C, Xu ZQ, Sergeyev V, Ubink R, Diez M. 2000. Neuropeptides--an overview. Neuropharmacology 39(8):1337-1356.

    Ignell R, Root CM, Birse RT, Wang JW, Nässel DR, Winther AM. 2009. Presynaptic peptidergic modulation of olfactory receptor neurons in Drosophila. Proc Natl Acad Sci U S A 106(31):13070-13075.

    Ilius M, Wolf R, Heisenberg M. 1994. Central Complex of Drosophila-Melanogaster Is Involved in-Flight Control - Studies on Mutants and Mosaics of the Gene Ellipsoid Body Open. Journal of neurogenetics 9(3):189-206.

    Isabel G, Martin JR, Chidami S, Veenstra JA, Rosay P. 2005. AKH-producing neuroendocrine cell ablation decreases trehalose and induces behavioral changes in Drosophila. Am J Physiol Regul Integr Comp Physiol 288(2):R531-538.

    Ito K, Suzuki K, Estes P, Ramaswami M, Yamamoto D, Strausfeld NJ. 1998. The organization of extrinsic neurons and their implications in the functional roles of the mushroom bodies in Drosophila melanogaster Meigen. Learning & memory (Cold Spring Harbor, NY 5(1-2):52-77.

    Jan LY, Jan YN. 1976. L-glutamate as an excitatory transmitter at the Drosophila larval neuromuscular junction. J Physiol 262(1):215-236.

    Johard HA, Coast GM, Mordue W, Nässel DR. 2003. Diuretic action of the peptide locustatachykinin I: cellular localisation and effects on fluid secretion in Malpighian tubules of locusts. Peptides 24(10):1571-1579.

    Johard HA, Enell LE, Gustafsson E, Trifilieff P, Veenstra JA, Nässel DR. 2008. Intrinsic neurons of Drosophila mushroom bodies express short neuropeptide F: relations to extrinsic neurons expressing different neurotransmitters. J Comp Neurol 507(4):1479-1496.

    Johard HA, Yoishii T, Dircksen H, Cusumano P, Rouyer F, Helfrich-Forster C, Nässel DR. 2009. Peptidergic clock neurons in Drosophila: ion transport peptide and short neuropeptide F in subsets of dorsal and ventral lateral neurons. J Comp Neurol 516(1):59-73.

    Johnson O, Becnel J, Nichols CD. 2009. Serotonin 5-HT(2) and 5-HT(1A)-like receptors differentially modulate aggressive behaviors in Drosophila melanogaster. Neuroscience 158(4):1292-1300.

    Kanda A, Iwakoshi-Ukena E, Takuwa-Kuroda K, Minakata H. 2003. Isolation and characterization of novel tachykinins from the posterior salivary gland of the common octopus Octopus vulgaris. Peptides 24(1):35-43.

    Kaupmann K, Malitschek B, Schuler V, Heid J, Froestl W, Beck P, Mosbacher J, Bischoff S, Kulik A, Shigemoto R, Karschin A, Bettler B. 1998. GABA(B)-receptor subtypes assemble into functional heteromeric complexes. Nature 396(6712):683-687.

  • 38

    Kim YC, Lee HG, Seong CS, Han KA. 2003. Expression of a D1 dopamine receptor dDA1/DmDOP1 in the central nervous system of Drosophila melanogaster. Gene Expression Patterns 3(2):237-245.

    Kitamoto T. 2001. Conditional modification of behavior in Drosophila by targeted expression of a temperature-sensitive shibire allele in defined neurons. J Neurobiol 47(2):81-92.

    Kolodziejczyk A, Sun X, Meinertzhagen IA, Nässel DR. 2008. Glutamate, GABA and acetylcholine signaling components in the lamina of the Drosophila visual system. PLoS One 3(5):e2110.

    Kopczynski CC, Noordermeer JN, Serano TL, Chen WY, Pendleton JD, Lewis S, Goodman CS, Rubin GM. 1998. A high throughput screen to identify secreted and transmembrane proteins involved in Drosophila embryogenesis. Proc Natl Acad Sci U S A 95(17):9973-9978.

    Krashes MJ, DasGupta S, Vreede A, White B, Armstrong JD, Waddell S. 2009. A neural circuit mechanism integrating motivational state with memory expression in Drosophila. Cell 139(2):416-427.

    Kume K, Kume S, Park SK, Hirsh J, Jackson FR. 2005. Dopamine is a regulator of arousal in the fruit fly. J Neurosci 25(32):7377-7384.

    Lange AB. 2001. Feeding state influences the content of FMRFamide- and tachykinin-related peptides in endocrine-like cells of the midgut of Locusta migratoria. Peptides 22(2):229-234.

    Lansdell SJ, Millar NS. 2000. Cloning and heterologous expression of Dalpha4, a Drosophila neuronal nicotinic acetylcholine receptor subunit: identification of an alternative exon influencing the efficiency of subunit assembly. Neuropharmacology 39(13):2604-2614.

    Lee G, Park JH. 2004. Hemolymph sugar homeostasis and starvation-induced hyperactivity affected by genetic manipulations of the adipokinetic hormone-encoding gene in Drosophila melanogaster. Genetics 167(1):311-323.

    Lee HG, Seong CS, Kim YC, Davis RL, Han KA. 2003. Octopamine receptor OAMB is required for ovulation in Drosophila melanogaster. Dev Biol 264(1):179-190.

    Lee KS, Hong SH, Kim AK, Ju SK, Kwon OY, Yu K. 2009. Processed short neuropeptide F peptides regulate growth through the ERK-insulin pathway in Drosophila melanogaster. FEBS Lett 583(15):2573-2577.

    Lee KS, Kwon OY, Lee JH, Kwon K, Min KJ, Jung SA, Kim AK, You KH, Tatar M, Yu K. 2008. Drosophila short neuropeptide F signalling regulates growth by ERK-mediated insulin signalling. Nat Cell Biol 10(4):468-475.

    Lee KS, You KH, Choo JK, Han YM, Yu K. 2004. Drosophila short neuropeptide F regulates food intake and body size. J Biol Chem 279(49):50781-50789.

    Li XJ, Wolfgang W, Wu YN, North RA, Forte M. 1991. Cloning, heterologous expression and developmental regulation of a Drosophila receptor for tachykinin-like peptides. EMBO J 10(11):3221-3229.

    Littleton JT, Ganetzky B. 2000. Ion channels and synaptic organization: analysis of the Drosophila genome. Neuron 26(1):35-43.

    Liu G, Seiler H, Wen A, Zars T, Ito K, Wolf R, Heisenberg M, Liu L. 2006. Distinct memory traces for two visual features in the Drosophila brain. Nature 439(7076):551-556.

  • 39

    Liu T, Dartevelle L, Yuan C, Wei H, Wang Y, Ferveur JF, Guo A. 2008. Increased dopamine level enhances male-male courtship in Drosophila. J Neurosci 28(21):5539-5546.

    Liu X, Buchanan ME, Han KA, Davis RL. 2009. The GABAA receptor RDL suppresses the conditioned stimulus pathway for olfactory learning. J Neurosci 29(5):1573-1579.

    Liu X, Krause WC, Davis RL. 2007. GABAA Receptor RDL Inhibits Drosophila Olfactory Associative Learning. Neuron 56(6):1090-1102.

    Lodish H, Berk A, Zipursky SL, Matsudaira P, Baltimore D, Darnell JE. 1999. Molecular Cell Biology.

    Löesel R, Nassel DR, Strausfeld NJ. 2002. Common design in a unique midline neuropil in the brains of arthropods. Arthropod Struct Dev 31(1):77-91.

    Mahr A, Aberle H. 2006. The expression pattern of the Drosophila vesicular glutamate transporter: a marker protein for motoneurons and glutamatergic centers in the brain. Gene Expr Patterns 6(3):299-309.

    Mao Z, Davis RL. 2009. Eight different types of dopaminergic neurons innervate the Drosophila mushroom body neuropil: anatomical and physiological heterogeneity. Front Neural Circuits 3:5.

    Martin JR. 2004. A portrait of locomotor behaviour in Drosophila determined by a video-tracking paradigm. Behav Processes 67(2):207-219.

    Martin JR, Ernst R, Heisenberg M. 1999. Temporal pattern of locomotor activity in Drosophila melanogaster. Journal of comparative physiology 184(1):73-84.

    Mertens I, Meeusen T, Huybrechts R, De Loof A, Schoofs L. 2002. Characterization of the short neuropeptide F receptor from Drosophila melanogaster. Biochem Biophys Res Commun 297(5):1140-1148.

    Mezler M, Müller T, Raming K. 2001. Cloning and functional expression of GABAB receptors from Drosophila. European Journal of Neuroscience 13(3):477-486.

    Monastirioti M. 1999. Biogenic amine systems in the fruit fly Drosophila melanogaster. Microscopy Research and Technique 45(2):106-121.

    Monastirioti M. 2003. Distinct octopamine cell population residing in the CNS abdominal ganglion controls ovulation in Drosophila melanogaster. Dev Biol 264(1):38-49.

    Monnier D, Colas JF, Rosay P, Hen R, Borrelli E, Maroteaux L. 1992. NKD, a developmentally regulated tachykinin receptor in Drosophila. J Biol Chem 267(2):1298-1302.

    Neckameyer WS. 1998. Dopamine modulates female sexual receptivity in Drosophila melanogaster. J Neurogenet 12(2):101-114.

    Neuser K, Triphan T, Mronz M, Poeck B, Strauss R. 2008. Analysis of a spatial orientation memory in Drosophila. Nature.

    Nichols CD. 2007. 5-HT2 receptors in Drosophila are expressed in the brain and modulate aspects of circadian behaviors. Dev Neurobiol 67(6):752-763.

    Nässel D. 2009. Neuropeptide signaling near and far: how localized and timed is the action of neuropeptides in brain circuits? Invertebrate Neuroscience 9(2):57-75.

    Nässel DR. 1988. Serotonin and serotonin-immunoreactive neurons in the nervous system of insects. Prog Neurobiol 30(1):1-85.

  • 40

    Nässel DR. 2002. Neuropeptides in the nervous system of Drosophila and other insects: multiple roles as neuromodulators and neurohormones. Prog Neurobiol 68(1):1-84.

    Nässel DR, Elekes K. 1992. Aminergic neurons in the brain of blowflies and Drosophila: dopamine- and tyrosine hydroxylase-immunoreactive neurons and their relationship with putative histaminergic neurons. Cell Tissue Res 267(1):147-167.

    Nässel DR, Enell LE, Santos JG, Wegener C, Johard HA. 2008. A large population of diverse neurons in the Drosophila central nervous system expresses short neuropeptide F, suggesting multiple distributed peptide functions. BMC Neurosci 9:90.

    Nässel DR, Passier PC, Elekes K, Dircksen H, Vullings HG, Cantera R. 1995. Evidence that locustatachykinin I is involved in release of adipokinetic hormone from locust corpora cardiaca. Regul Pept 57(3):297-310.

    Nässel DR, Winther ÅME. 2010. Drosophila neuropeptides in regulation of physiology and behavior. Progress in Neurobiology 92(1):42-104.

    O'Shea M, Schaffer M. 1985. Neuropeptide function: the invertebrate contribution. Annu Rev Neurosci 8:171-198.

    Okada R, Awasaki T, Ito K. 2009. Gamma-aminobutyric acid (GABA)-mediated neural connections in the Drosophila antennal lobe. J Comp Neurol 514(1):74-91.

    Olsen SR, Wilson RI. 2008. Lateral presynaptic inhibition mediates gain control in an olfactory circuit. Nature 452(7190):956-960.

    Pan Y, Zhou Y, Guo C, Gong H, Gong Z, Liu L. 2009. Differential roles of the fan-shaped body and the ellipsoid body in Drosophila visual pattern memory. Learning & memory (Cold Spring Harbor, NY 16(5):289-295.

    Panneels V, Eroglu C, Cronet P, Sinning I. 2003. Pharmacological characterization and immunoaffinity purification of metabotropic glutamate receptor from Drosophila overexpressed in Sf9 cells. Protein Expr Purif 30(2):275-282.

    Parmentier ML, Pin JP, Bockaert J, Grau Y. 1996. Cloning and functional expression of a Drosophila metabotropic glutamate receptor expressed in the embryonic CNS. J Neurosci 16(21):6687-6694.

    Pendleton RG, Rasheed A, Sardina T, Tully T, Hillman R. 2002. Effects of tyrosine hydroxylase mutants on locomotor activity in Drosophila: a study in functional genomics. Behav Genet 32(2):89-94.

    Poeck B, Triphan T, Neuser K, Strauss R. 2008. Locomotor control by the central complex in Drosophila-An analysis of the tay bridge mutant. Dev Neurobiol 68(8):1046-1058.

    Poels J, Birse RT, Nachman RJ, Fichna J, Janecka A, Vanden Broeck J, Nassel DR. 2009. Characterization and distribution of NKD, a receptor for Drosophila tachykinin-related peptide 6. Peptides 30(3):545-556.

    Popov AV, Peresleni AI, Ozerskii PV, Shchekanov EE, Savvateeva-Popova EV. 2003. [Role of the protocerebral bridge in the central complex of Drosophila melanogaster brain in the control of courtship behavior and sound production of males]. Zh Evol Biokhim Fiziol 39(6):530-539.

    Popov AV, Peresleni AI, Ozerskii PV, Shchekanov EE, Savvateeva-Popova EV. 2004. [The fan-shaped and ellipsoid bodies of the brain central complex are involved in the control of courtship behavior and communicative sound production in

  • 41

    Drosophila melanogaster males]. Ross Fiziol Zh Im I M Sechenova 90(4):385-399.

    Power ME. 1943. The brain of the Drosophila melanogaster. J Morphol :(72):517-559. Predel R, Rapus J, Eckert M. 2001. Myoinhibitory neuropeptides in the American

    cockroach. Peptides 22(2):199-208. Qin G, Schwarz T, Kittel RJ, Schmid A, Rasse TM, Kappei D, Ponimaskin E, Heckmann

    M, Sigrist SJ. 2005. Four different subunits are essential for expressing the synaptic glutamate receptor at neuromuscular junctions of Drosophila. J Neurosci 25(12):3209-3218.

    Reale V, Chatwin HM, Evans PD. 2004. The activation of G-protein gated inwardly rectifying K+ channels by a cloned Drosophila melanogaster neuropeptide F-like receptor. Eur J Neurosci 19(3):570-576.

    Renn SC, Armstrong JD, Yang M, Wang Z, An X, Kaiser K, Taghert PH. 1999. Genetic analysis of the Drosophila ellipsoid body neuropil: organization and development of the central complex. J Neurobiol 41(2):189-207.

    Ring M, Meredith J, Wiens C, Macins A, Brock HW, Phillips JE, Theilmann DA. 1998. Expression of Schistocerca gregaria ion transport peptide (ITP) and its homologue (ITP-L) in a baculovirus/insect cell system. Insect Biochem Mol Biol 28(1):51-58.

    Robb S, Cheek TR, Hannan FL, Hall LM, Midgley JM, Evans PD. 1994. Agonist-specific coupling of a cloned Drosophila octopamine/tyramine receptor to multiple second messenger systems. EMBO J 13(6):1325-1330.

    Root CM, Masuyama K, Green DS, Enell LE, Nässel DR, Lee C-H, Wang JW. 2008. A Presynaptic Gain Control Mechanism Fine-Tunes Olfactory Behavior. Neuron 59(2):311-321.

    Saraswati S, Fox LE, Soll DR, Wu CF. 2004. Tyramine and octopamine have opposite effects on the locomotion of Drosophila larvae. J Neurobiol 58(4):425-441.

    Sattelle DB, Jones AK, Sattelle BM, Matsuda K, Reenan R, Biggin PC. 2005. Edit, cut and paste in the nicotinic acetylcholine receptor gene family of Drosophila melanogaster. Bioessays 27(4):366-376.

    Saudou F, Amlaiky N, Plassat JL, Borrelli E, Hen R. 1990. Cloning and characterization of a Drosophila tyramine receptor. EMBO J 9(11):3611-3617.

    Saudou F, Boschert U, Amlaiky N, Plassat JL, Hen R. 1992. A family of Drosophila serotonin receptors with distinct intracellular signalling properties and expression patterns. EMBO J 11(1):7-17.

    Scholz H, Ramond J, Singh CM, Heberlein U. 2000. Functional ethanol tolerance in Drosophila. Neuron 28(1):261-271.

    Schwaerzel M, Monastirioti M, Scholz H, Friggi-Grelin F, Birman S, Heisenberg M. 2003. Dopamine and octopamine differentiate between aversive and appetitive olfactory memories in Drosophila. J Neurosci 23(33):10495-10502.

    Shapiro RA, Wakimoto BT, Subers EM, Nathanson NM. 1989. Characterization and functional expression in mammalian cells of genomic and cDNA clones encoding a Drosophila muscarinic acetylcholine receptor. Proc Natl Acad Sci U S A 86(22):9039-9043.

    Sinakevitch I, Strausfeld NJ. 2006. Comparison of octopamine-like immunoreactivity in the brains of the fruit fly and blow fly. J Comp Neurol 494(3):460-475.

  • 42

    Sitaraman D, Zars M, Laferriere H, Chen YC, Sable-Smith A, Kitamoto T, Rottinghaus GE, Zars T. 2008. Serotonin is necessary for place memory in Drosophila. Proc Natl Acad Sci U S A 105(14):5579-5584.

    Siviter RJ, Coast GM, Winther AM, Nachman RJ, Taylor CA, Shirras AD, Coates D, Isaac RE, Nassel DR. 2000. Expression and functional characterization of a Drosophila neuropeptide precursor with homology to mammalian preprotachykinin A. J Biol Chem 275(30):23273-23280.

    Skaer NJ, Nassel DR, Maddrell SH, Tublitz NJ. 2002. Neurochemical fine tuning of a peripheral tissue: peptidergic and aminergic regulation of fluid secretion by Malpighian tubules in the tobacco hawkmoth M. sexta. J Exp Biol 205(Pt 13):1869-1880.

    Srivastava DP, Yu EJ, Kennedy K, Chatwin H, Reale V, Hamon M, Smith T, Evans PD. 2005. Rapid, nongenomic responses to ecdysteroids and catecholamines mediated by a novel Drosophila G-protein-coupled receptor. J Neurosci 25(26):6145-6155.

    Strausfe