Skip to main content
Advertisement
  • Loading metrics

Genetic Dissection of Aversive Associative Olfactory Learning and Memory in Drosophila Larvae

Abstract

Memory formation is a highly complex and dynamic process. It consists of different phases, which depend on various neuronal and molecular mechanisms. In adult Drosophila it was shown that memory formation after aversive Pavlovian conditioning includes—besides other forms—a labile short-term component that consolidates within hours to a longer-lasting memory. Accordingly, memory formation requires the timely controlled action of different neuronal circuits, neurotransmitters, neuromodulators and molecules that were initially identified by classical forward genetic approaches. Compared to adult Drosophila, memory formation was only sporadically analyzed at its larval stage. Here we deconstruct the larval mnemonic organization after aversive olfactory conditioning. We show that after odor-high salt conditioning larvae form two parallel memory phases; a short lasting component that depends on cyclic adenosine 3’5’-monophosphate (cAMP) signaling and synapsin gene function. In addition, we show for the first time for Drosophila larvae an anesthesia resistant component, which relies on radish and bruchpilot gene function, protein kinase C activity, requires presynaptic output of mushroom body Kenyon cells and dopamine function. Given the numerical simplicity of the larval nervous system this work offers a unique prospect for studying memory formation of defined specifications, at full-brain scope with single-cell, and single-synapse resolution.

Author Summary

Learning and memory helps organisms to predict and adapt to events in their environment. Gained experience leaves traces of memory in the nervous system. Yet, memory formation in vertebrates and invertebrates is a highly complex and dynamic process that consists of different phases, which depend on various neuronal and molecular mechanisms. To understand which changes occur in a brain when it learns, we applied a reductionist approach. Instead of studying complex cases, we analyzed learning and memory in Drosophila larvae that have a simple brain that is genetically and behaviorally accessible and consists of only about 10,000 neurons. Drosophila larvae are able to learn to associate an odor with punishing high salt concentrations. It is therefore possible to correlate changes in larval behavior with molecular events in identifiable neurons after classical olfactory conditioning. We show that under these circumstances larvae form two parallel memory phases; a short lasting component (lSTM) that is molecularly conserved throughout the animal kingdom as it depends on the classical cAMP pathway. In parallel they establish a larval anesthesia resistant memory (lARM) that relies on a different molecular signal. lARM has not been described in larvae before.

Introduction

Experience leaves traces of memory in the nervous system. This assists organisms to predict and adapt to events in their environment. Both invertebrates and vertebrates possess a variety of different learning mechanisms [1, 2]. Associative learning, for instance, enables animals to draw on past experience to predict the occurrence of food, predators or social partners [3]. Several studies in vertebrates and invertebrates have revealed that associative memories consist of distinct phases, which differ in duration and time of expression. Throughout the animal kingdom, a labile, short-term memory can be distinguished from a robust, long-term memory [46]. Long-term memory—in contrast to short-term memory—is resistant to anesthetic disruption and depends on consolidation processes including de novo protein synthesis [4, 69].

Genetic studies in adult Drosophila following olfactory classical conditioning using electric shock as a negative reinforcer have identified distinct temporal memory phases—short-term memory (STM), middle-term memory (MTM), long-term memory (LTM) and a so-called anesthesia-resistant memory (ARM) [10, 11]. STM and MTM are both considered to be unconsolidated whereas ARM and LTM are consolidated forms of memory. The main property of STM and MTM is a dependency on the cyclic adenosine 3’5’-monophosphate (cAMP) pathway [12] as exemplified by early studies of rutabaga (rut) encoded type I Ca2+-dependent adenylyl cyclase (AC1) [13, 14] and dunce (dnc) encoded type 4 cAMP-specific phosphodiesterase (PDE4) [1517]. Consolidated LTM and ARM are assumed to be represented by separate molecular pathways [18]. In contrast to ARM formation, LTM requires cAMP response element-binding protein (CREB) dependent transcription and de-novo protein synthesis [10, 19, 20]. Nevertheless, ARM is resistant to anesthetic agents [21], which cause retrograde amnesia in both invertebrates and vertebrates [6, 8, 21, 22]. Furthermore ARM formation requires the activity of the radish gene [23, 24]. Taken together, in adult Drosophila classical conditioning following odor-electric shock reinforcement establishes at least four sequential and/or parallel memory phases (but see also [25] for a further subdivision of ARM). However, there is growing evidence that things are unlikely to be as straightforward as originally envisaged. For example, changing parameters of the training regime, such as feeding state, age of flies, timing of the stimuli and the reinforcing stimulus affects distinct aspects of memory formation and in the most extreme case leads to a mechanistically different type of memory being formed [2629].

Based on the above described, well-established genetic interventions that have functional implications for adult Drosophila we have analyzed memory formation at the larval stage. Although Drosophila larvae are able to form olfactory and visual memories [3041], only a few studies have described larval memory formation in more detail. Larval olfactory memory also consists of different phases [32, 36, 40, 42, 43]. However, some of the studies identified only a short-lasting memory [32, 42], while others studies came to the conclusion that the larval memory consists of both, a short-lasting and a long-lasting component [36, 40, 43]. Furthermore, genetic dissection of the larval memory linked memory formation to the cAMP pathway [32, 36, 40, 42, 43]. However, two of these studies have shown in addition, that rsh1 mutants and turnip (tur) mutants, which are reduced in protein kinase C (PKC) activity, showed an impairment in larval memory [36, 40]. Recapitulating the appearance of sequential and/or parallel memory phases in larvae is rather difficult, since these molecular processes were suggested to be independent of cAMP signaling.

Here we have deconstructed the larval mnemonic organization after odor-high salt conditioning. Therefore we adapted paradigms from adult Drosophila, which allowed us to identify different components of larval memory. We applied (i) a cold shock in order to identify an anesthesia resistant form and (ii) blocked protein synthesis in order to distinguish protein synthesis independent from the protein synthesis dependent forms.

We have shown that depending on the training regime Drosophila larvae are capable of forming distinct memory phases. Following odor-high salt training we identify three different specifications. We describe for the first time an anesthesia resistant memory in larvae (lARM) that it is not affected by cold shock treatment and is evident for up to four hours after training. The component (we use this term here as we were not able to distinguish between the acquisition, consolidation and retrieval of lARM) relies on radish and bruchpilot gene function, as well as presynaptic output of mushroom body Kenyon cells (MB KCs) and dopaminergic signaling. Furthermore, it utilizes the PKC pathway in contrast to traditional cAMP signaling. Second, we describe a short lasting component (evident for up to 20 minutes after one cycle training) that depends on traditional cAMP signaling and synapsin gene function. Third, we identify a CREB dependent component that requires a spaced training protocol, which is composed of five cycles of conditioning spaced by rest intervals of 15 minutes.

Results

Drosophila larvae establish an aversive olfactory memory that lasts several hours

Third instar Drosophila larvae are able to learn to associate an odor with punishing high salt concentrations [39, 44]. Thus we utilized a well-established and standardized two odor reciprocal olfactory conditioning paradigm with 1.5M sodium chloride (NaCl) as negative reinforcement and tested memory persistence by assaying larvae at increasing times after training (Fig 1A). Please note that the standardized paradigm consists of three training trials (Fig 1B). Significant aversive olfactory memory was evident up to four hours after training (Fig 1C). However, the memory exhibited a gradual decay as the time interval increased and was no longer statistically significant after five hours (Fig 1C). The result is supported by nonlinear regression analysis, which describes the retention curve of odor-high salt memory through an exponential decay function (Fig 1C). This suggests that the initially formed odor-high salt memory gradually decays over time.

thumbnail
Fig 1. Aversive olfactory memory after odor-high salt conditioning lasts up to four hours

A: Schematic drawing of the used two odor reciprocal training paradigm. During training, thirty larvae receive the odor n-amylacetate (AM) paired with an aversive reinforcer (high salt concentration) while benzaldehyde (BA) was presented alone (AMSalt / BAPure) (Group 1). Group 2 receives the reverse contingency (AMPure / BASalt). The training was three times repeated. During test both odors are presented on opposite sides. After 5 minutes the number of larvae on each odor side is counted for both reciprocally trained groups and a performance index (PI) is calculated that quantifies associative olfactory memory. B: Flowchart that summarizes the details of the behavioral paradigm in an alternative way. This representation is used throughout the manuscript. Note, for simplification the reciprocally trained group is not shown. C: Larval aversive olfactory memory using three training repetitions was tested in wild type larvae at different time points after conditioning ranging from 0–300 minutes. The aversive memory is stable up to four hours (One sample t test, p<0.05 for t = 0-250min; p>0.05 for t = 300min). The memory decay was fitted into an exponential decay function (nonlinear regression analysis, R2 = 0.257, τ = -145.9). Memory performance significantly different from random distribution (p<0.05) is indicated in black, random distribution (p≥0.05) in light grey. Sample size is n = 16 for each group. All data are given as means ± s.e.m.

https://doi.org/10.1371/journal.pgen.1006378.g001

Aversive olfactory learning and memory is independent of de-novo protein synthesis and resistant to cold shock

Our data show that larvae can associate odors with high salt punishment and that the learning dependent change in behavior lasts several hours. In adult Drosophila two types of longer-lasting memories were described, called ARM and LTM. Besides being resistant to anesthetic disruption, ARM is apparently independent of protein synthesis [29]. Yet, LTM formation requires de novo protein synthesis [10, 19]. In order to test if the memory is dependent on de novo protein synthesis, we fed larvae the translation-inhibitor cycloheximide (CXM) 20 hours before the experiment [10]. Then odor-high salt memory was tested immediately or 60 minutes after three cycle standard training (Fig 2A). Performance was unaffected by CXM treatment (Fig 2A and S1A Fig.), suggesting that the formed memory is independent of de novo protein synthesis. This conclusion is further supported by two additional findings. First, the deleterious effect of blocking protein synthesis using CXM became apparent by constantly feeding CXM over a longer period of time. CXM treated larvae did neither pupate nor eclose in contrast to both control groups (S1B Fig.). Second, the transcription factor cAMP response element-binding protein (CREB) is universally required for LTM, and it has been reported that a dominant-negative dCreb2b repressor transgene driven by a heat-shock promoter (hs-dCreb2b) reduces LTM formation in a heat-shock dependent manner [19, 20]. Expression of dCreb2b via OK107-Gal4 specifically in the larval MB Kenyon cells did not change odor-high salt memory tested immediately or 60 minutes after training when compared to both genetic controls (Fig 2B and S1C Fig). Yet, adult Drosophila are only capable of forming LTM following a spaced training protocol composed of at least five cycles of conditioning separated by inter-trial intervals of 15 minutes [10, 11, 25]. Therefore we established a spaced training paradigm for larval odor-high salt conditioning (S1D Fig.; five training cycles, 15 minutes inter-trial interval). Spaced training induced a learning dependent change of the behavior of two genetic control groups, but not in the behavior of transgenic larvae expressing dCreb2b via OK107-Gal4 specifically in the larval MB Kenyon cells (S1D Fig.). Thus, the obtained results suggest that the established type of odor-high salt memory is paradigm dependent. However, the prominent component established following three cycle standard training is independent of protein synthesis—and therefore by a general criteria of memory formation not LTM.

thumbnail
Fig 2. Drosophila larvae establish an anesthesia resistant type of memory (lARM)

Training and different treatment protocols are shown at the top of each panel. A: Effect of cycloheximide (CXM) on larval aversive olfactory memory of wild type larvae tested directly and 60 minutes after three cycle standard training. Application of CXM 20 hours before training did neither affect aversive olfactory memory at 0 minutes nor at 60 minutes after training (Two way ANOVA, p = 0.313). B: Expression of a dominant-negative dCREB2-b repressor transgene (dCREB2-b) in MB KCs via OK107-Gal4 did not affect larval aversive olfactory memory tested 0 and 60 minutes after three cycle standard training (Two way ANOVA, p = 0.134 between the experimental group OK107-Gal4/UAS-dCREB2-b and both control groups). C: Effect of cold shock application on the retention of larval aversive olfactory memory. Directly after training wild type larvae received a one minute cold shock at 4°C. The memory was tested 10, 60, 120 and 180 minutes after three cycle standard training. Cold shock application did not reduce aversive olfactory memory at any time point (Two way ANOVA, p = 0.403). D: Effect of cold shock treatment on larval aversive olfactory memory tested 60 minutes after three cycle standard training. Cold shock treatment was applied 0, 10, 20 or 40 minutes after conditioning. Cold shock applied at different time points did not reduce aversive olfactory memory of wild type larvae tested 60 minutes after three cycle standard training (Two way ANOVA, p = 0.149). E: Cold shock application administered directly after odor-6 mM quinine training did not reduce aversive olfactory memory in wild type larvae tested 10 minutes after three cycle standard training (Unpaired t test, p = 0.610). F: Conditioning with 2.0M fructose reward in wild type larvae led to an appetitive olfactory memory, which is partially sensitive to cold shock treatment (Unpaired t test, p = 0.031). Yet, appetitive olfactory memory is not completely vanished (One sample t test, p = 0.026). Sample size is n = 16 for each group if not indicated otherwise. In Fig 2E and 2F differences between groups are depicted above the respective box plots, at which ns indicates p≥0.05 and * p<0.05. Grey boxes indicate a memory performance above chance level (p<0.05). Small circles indicate outliers. In Fig 2A–2D differences between groups are depicted below the symbols, at which ns indicates p≥0.05. Memory performance significantly different from random distribution (p<0.05) is indicated in black. The data in Fig 2A–2D are shown as means ± s.e.m. The data in Fig 2E and 2F are shown as box plots.

https://doi.org/10.1371/journal.pgen.1006378.g002

Next, we tested whether odor-high salt memory following three cycle standard training is resistant to anesthesia. We established a cold shock treatment protocol. We trained larvae as described before but put them directly into cold water (4°C) for one minute after training. Larvae were then transferred onto a room temperature agar plate to recover and memory was tested after different retention times. As shown in Fig 2C (see also S2C Fig.) applying a cold shock treatment did not disrupt odor-high salt memory tested 10, 60, 120 and 180 minutes after training (10 minutes is necessary for recovery from the cold). Even applying a stronger cold shock of 5 minutes, which completely paralyzed larvae, did not affect odor-high salt memory (S2B Fig.). We also tested whether cold shock treatment applied 0, 10, 20 or 40 minutes after training disrupted 60 minutes memory. Again, no significant defect was revealed (Fig 2D and S2D Fig). To test if larval memory following three cycle standard training is in general resistant to cold shock treatment we additionally used 6mM quinine as a negative reinforcer [34, 45] and 2.0M fructose as an appetitive reinforcer (Fig 2E and 2F). For both stimuli the established memory was resistant to cold shock treatment. Please note that in case of fructose reinforcement the obtained memory was partially reduced. Implications for larval appetitive olfactory learning and memory are later discussed. All in all, our results show for the first time that larvae independent of the applied reinforcer are able to form a type of anesthesia resistant memory.

The radish gene is necessary for larval anesthesia resistant learning and memory

It was shown in adult Drosophila that the radish (rsh) gene plays a pivotal role for the formation of ARM [11, 24]. Hereinafter we therefore focused on rsh gene function. We first analyzed the memory performance of rsh mutant larvae following three cycle standard training immediately after training or after 60 minutes (Fig 3A). In both cases rsh1 mutants showed no significant performance (Fig 3A). To ascertain whether this effect is due to the mutation in the radish gene we performed a rescue experiment (Fig 3B). We tested rsh1 mutants that harbor a wild type rsh transgene, hs-rsh, that allows to induce ubiquitous expression of rsh following heat shock [23]. Non-induced larvae showed a lack of anesthesia resistant learning and/or memory, similar to larvae that carry only the rsh1 mutation. Yet, ubiquitous expression of rsh shortly before the experiment rescued the phenotype (Fig 3B). However, at a reduced level as compared to wild type controls (Fig 3B). Yet, task-relevant sensory-motor abilities of rsh1 larvae are defective in responding to the odor benzaldeyhde (BA) (S3B and S3D Fig). To clearly show that the impairment for rsh1 mutants is based on a loss of the ability to associate odor with high salt concentrations, we performed additional experiments. We used a one odor reciprocal paradigm (S3C Fig.) [46]. Here BA presentation is replaced by paraffin oil that does not provide any olfactory information for the larva. Again rsh1 larvae showed no anesthesia resistant learning and/or memory (S3C Fig.). In summary, we thus conclude that the behavioral phenotype is due to the fact that the mutation in the rsh gene prevents larvae from establishing, consolidating and/or recalling anesthesia resistant memory. Please note that our experiments did not allow to distinguish between the three different processes.

thumbnail
Fig 3. Odor-high salt learning and/or memory depends on radish gene function

Training and temperature protocols are shown at the top of each panel. A: In contrast to wild type larvae, aversive olfactory learning and/or memory is impaired in rsh1 mutants tested 0 and 60 minutes after three cycle standard training (One sample t test, p = 0.95 tested at 0 minutes, p = 0.31 tested at 60 minutes). B: Rescue experiment of rsh1 learning and/or memory phenotype using a hs-rsh transgene. The transgene was induced via a heat-shock of 35°C for 5.5 hours (continuous line). The control group was kept at 22°C for 5.5 hours (dashed line). Without heat-shock experimental animals (rsh1, w,rsh1;hs-rsh) tested 0 minutes after three cycle standard training distributed randomly (One sample t test, p = 0.09 for rsh1, p = 0.52 for w,rsh1;hs-rsh). Yet, both genetic controls showed an aversive memory (One sample t test, p<0.0001 for wild type, Wilcoxon signed rank test, p = 0.0001 for w1118). After heat-shock application only rsh1 mutants distributed randomly (One sample t test, p = 0.25). Yet, ubiquitous induction of rsh expression partially rescues the learning and/or memory phenotype (One sample t test, p = 0.0009, unpaired t test, p = 0.01 comparing w,rsh1;hs-rsh and w1118). Both control groups showed a memory performance above chance level (One sample t test, p<0.0001 for wild type, p<0.0001 for w1118). Differences between groups are depicted above the respective box plots, at which ns indicates p≥0.05 and * p<0.05. Grey boxes show memory performance above chance level (p<0.05), whereas white boxes indicate random distribution (p≥0.05). Small circles indicate outliers. Sample size is n = 16 for each group if not indicated otherwise.

https://doi.org/10.1371/journal.pgen.1006378.g003

Bruchpilot gene function is necessary at the presynaptic terminals of mushroom body Kenyon cells for anesthesia resistant learning and memory

Next we analyzed if intrinsic MB KCs are required for anesthesia resistant learning and/or memory following three cycle standard training due to its conserved role in larval and adult olfactory memory formation [32, 42, 47, 48]. Expression of the temperature-sensitive dominant negative form of dynamin shibirets1 (UAS-shits1) [48, 49] via the OK107-Gal4 in all KCs allows to block synaptic KC output at a restrictive temperature of 35°C due to impaired vesicle recycling (Fig 4A). In contrast to both genetic control groups, OK107-Gal4/UAS-shits1 larvae showed no anesthesia resistant learning and/or memory (Fig 4A). Yet significant difference was only detectable between the UAS-shits1/+ control and OK107-Gal4 /UAS-shits1 (Fig 4A). Control experiments revealed no gross defects in task-relevant sensory-motor abilities (S4A Fig.). In addition UAS-mCD8::GFP expression driven by OK107-Gal4 verified MB specificity in all KCs besides a limited expression in the ventral nerve cord and brain hemispheres (Fig 4D) [48]. Repetition of the experiment with a second mushroom body specific driver H24-Gal4 [48] verified the results obtained for OK107-Gal4 (S4C and S4D Fig). Thus, we conclude that KC output is necessary for anesthesia resistant learning and/or memory.

thumbnail
Fig 4. Presynaptic output of mushroom body Kenyon cells is necessary for odor-high salt learning and/or memory

Training and temperature shift protocols are shown at the top of each panel. A: Effect of the blockade of presynaptic output of MB KCs via UAS-shits1 using OK107-Gal4 driver line on odor-high salt learning and/or memory using three cycle standard training. Larvae were raised at the permissive temperature (19°C) and shifted to restrictive temperature during training and testing. In contrast to both genetic controls that show lARM (One sample t test, p = 0.01 for OK107-Gal4/+, p = 0.0003 for UAS-shits1/+), shits1 expression driven via OK107-Gal4 leads to a complete loss of odor-high salt learning and/or memory tested immediately after three cycle standard training (One sample t test, p = 0.64). Note, only the UAS-shits1/+ control but not the OK107-Gal4/+ control was significantly different from the experimental group UAS-shits1/OK107-Gal4 (Tukey post hoc test, p = 0.018 and p = 0.22, respectively). B: The presynaptic protein Synapsin is not involved in lARM formation. syn97 loss-of-function mutants showed odor-high salt learning and memory (One sample t test, p<0.0001) and behaved non-distinguishable from the genetic control group (Unpaired t test, p = 0.96). C: In contrast to both genetic controls (OK107-Gal4/+ and UAS-brp-RNAiB3C8), knockdown of the presynaptic protein brp in the KCs by driving UAS-brp-RNAiB3C8 via OK107-Gal4 abolishes odor-high salt learning and/or memory (One sample t test, p = 0.56 for OK107-Gal4/UAS-brp-RNAiB3C8, p<0.0001 for OK107-Gal4/+, p<0.0001 for UAS-brp-RNAiB3C8). D: Shows a frontal view projection (left) of a OK107-Gal4/UAS-mCD8::GFP larval brain labeling the entire set of MB KCs (anti-GFP in green and anti-FasII, anti-ChAT neuropil staining in magenta). Additional staining is detectable in the ventral nerve cord (vnc) and neurons that express Drosophila insulin-like proteins (dilp). On the upper right panel a zoom in of the MB is shown. On the lower right panel only the GFP channel is depicted. Scale bars: left panels 50μm, right panels 25μm. E: Shows frontal view projections of a wild type brain (left panel) or the syn97 loss-of-function mutant (right panel) stained with anti-synapsin (magenta). As reported, anti-synapsin was only detected in the wild type brain but completely absent in the syn97 loss-of-function mutant. Scale bars: 50μm. F: Shows a single section of a frontal view of brain hemispheres of a wild type (left panel) and an experimental larva (right panel) using anti-nc82 to recognize Brp. In contrast to wild type larvae, no anti-nc82 staining was detectable in the MB (shown at the peduncle and the medial lobe level by arrows). Scale bars: 25μm. Differences between groups are depicted above the respective box plots, ns indicates p≥0.05. Different lowercase letters indicate statistical significant differences from each other at level p<0.05. Grey boxes show memory performance above chance level (p<0.05), white boxes indicate random distribution (p≥0.05). Small circles indicate outliers. Sample size is n = 16 for each group if not indicated otherwise

https://doi.org/10.1371/journal.pgen.1006378.g004

In adult Drosophila two presynaptic determinants, Synapsin (Syn) and Bruchpilot (Brp), play a pivotal role in controlling the release of KC vesicles. The evolutionary conserved phosphoprotein Syn is responsible for building a reserve pool of vesicles necessary to maintain vesicle release under high action potential frequencies [5053]. Adult syn97 mutants showed a defect in aversive olfactory memory that is independent of ARM formation [54, 55]. The active zone protein Brp, which is a homolog to the ELKS/CAST protein family, is an essential component of the presynaptic dense bodies regulating the release probability of synaptic vesicles [5658]. The presence of Brp in presynaptic terminals of KCs of adults was suggested to be necessary for establishing ARM [59]. To investigate if both proteins are required for anesthesia resistant learning and/or memory following three cycle standard training, we tested a syn deficient mutant syn97 and brp specific RNAi knockdown in all MB KCs via OK107-Gal4 (Fig 4B and 4C). Gene activity of syn was not required for anesthesia resistant learning and/or memory (Fig 4B). The performance of syn97 mutants was statistically indistinguishable from wild type larvae that served as a genetic control (Fig 4B). Lack of the Syn protein in syn97 was verified using a Syn specific antibody (Fig 4E) [47]. In contrast Brp function was necessary for anesthesia resistant learning and/or memory (Fig 4C). It was completely absent in OK107-Gal4;UAS-brp-RNAiB3C8 larvae (Fig 4C). Cell specific knockdown of brp in MB KCs was verified by antibody staining (Fig 4F). In addition, brp RNAi knockdown did not reveal gross defects in task-relevant sensory-motor abilities (S4B Fig.). Consequently, we suggest presynaptic activity of the active zone protein Brp in MB KCs is necessary to establish, consolidate and/or retrieve lARM. Please note that our experiments did not allow to distinguish between the three different processes.

Anesthesia resistant learning and memory is independent of the cAMP/protein kinase A pathway but requires protein kinase C activity

Molecular studies in several model organisms–including Drosophila—elucidate cAMP as crucial second messenger in memory formation. A proposed model for the molecular mechanism underlying olfactory memory formation is shown in Fig 5A. An association between the odorant and the reinforcement signals elicits an activation of type I Ca2+-dependent AC encoded by the rut gene via calcium/calmodulin and G-protein stimulation [13, 14, 60]. This synergistic activation of AC produces an increase in intracellular cAMP concentration [12]. Intolerable cAMP concentration is prevented through the activity of a type 4 cAMP-specific PDE encoded by the dnc gene [1214] cAMP for its part activates PKA [61]. The activation of PKA leads either to the phosphorylation of a variety of downstream targets (e.g. Synapsin, Na+ and K+ channels) for forming a short-lasting memory [47, 6264] or the phosphorylation of CREB forming a long-lasting memory [19, 20, 65].

thumbnail
Fig 5. Odor-high salt learning and/or memory following three cycle standard training is independent of the cAMP/PKA pathway

Training protocols are shown at the top of each panel. A: Working hypothesis that links the classical cAMP/PKA signaling pathway to the association of the unconditioned (US, high salt) and conditioned stimulus (CS, odor). Joined presentation of US and CS activates AC through simultaneous activation of Ca2+/CaM and G-protein stimulation. This results in the activation of the cAMP/PKA pathway. The activation of PKA leads either to the phosphorylation of a variety of downstream targets to change neuronal signaling on a shorter time scale or to phosphorylate CREB to form long-lasting memory. An arrowhead indicates stimulatory effects, whereas inhibitory effects are represented as a bar. GPCR: G-protein coupled receptor, VDCC: voltage-dependent calcium channel, ACh: acetylcholine, DA: dopamine, AC: adenylate cyclase, CaM: calcium/calmodulin-dependent protein kinase, cAMP: cyclic AMP, PDE: phosphodiesterase, PKA: protein kinase A, CREB: cyclic AMP response element-binding protein. B: rut1, rut2080 and dnc1 mutant larvae showed lARM comparable to wild type controls (One way ANOVA, p = 0.14). C: DCO encodes for the major catalytic subunit of PKA in Drosophila. Adults covering heterozygously the alleles DCOB3/+ and DCOH2/+ show a 50% reduction of PKA activity. DCOB3/+ and DCOH2/+ heterozygous mutant larvae showed normal lARM (One sample t test, p = 0.001 and p = 0.004), they performed at the same level as wild type larvae that served as a genetic control (One way ANOVA, p = 0.84). D: Epidermal growth factor receptor (EGFR) signaling to a Ras/Neurofibromatosis type I (NFI) pathway was reported to activate PKA. Expression of a dominant-negative isoform of EGFR (EGFRDN) in the MB KCs via OK107-Gal4 does not affect lARM (One sample t test, p = 0.0017), as the experimental group performed at the same level as both genetic controls (Kruskal-Wallis, p = 0.14). E: Working hypothesis for an alternative signaling pathway that allows larvae to form lARM. GPCR activation activates PLC that binds to downstream target elements, which stimulate PKC. Activation of typical forms of PKC needs also the binding of intracellular Ca2+. An arrowhead indicates stimulatory effects, whereas inhibitory effects are represented as a bar. GPCR: G-protein coupled receptor, DA: dopamine, PLC: phospholipase C, PKC: protein kinase C, DAG: diacylglycerol, IP3: inositoltriphosphat. F: Suppression of PKC activity in MB KCs by expressing an inhibitory pseudosubstrate of PKC (PKCi) under the control of OK107-Gal4 leads to a decrease in odor-high salt learning and/or memory (Kruskal-Wallis, p = 0.0001). Differences between groups are depicted above the respective box plots, at which ns indicates p≥0.05. Different lowercase letters indicate statistical significant differences at level p<0.05. Grey boxes show memory performance above chance level (p<0.05), whereas white boxes indicate random distribution (p≥0.05). Small circles indicate outliers. Sample size is n = 16 for each group if not indicated otherwise.

https://doi.org/10.1371/journal.pgen.1006378.g005

To uncover the molecular pathways responsible for anesthesia resistant learning and/or memory following three cycle standard training we tested larvae carrying three classical learning mutations having deficits in cAMP signaling: rutabaga1, rutabaga2080 and dunce1 [14, 17, 24] (Fig 5B). All three mutants showed lARM that was indistinguishable from wild type controls (Fig 5B). These results indicate that the formation, consolidation and retrieval of lARM is independent of the cAMP/PKA signaling pathway. This conclusion is further supported by two additional findings. First, hypomorphic alleles of the DCO gene locus (DCOB3 and DCOH2), which encodes the major catalytic subunit of the cAMP-dependent PKA (PKAc) showed normal lARM similar to genetic controls (Fig 5C). In adults these heterozygous DCOB3/+ and DCOH2/+ mutants show a 50% reduction of PKA activity and suppress age-related memory impairment [27, 66]. Second, epidermal growth factor receptor (EGFR) signaling to a Ras/Neurofibromatosis type I (NFI) pathway was suggested to act via a Rut-AC independent AC to activate PKA function [67]. Notably pan neuronal expression of a dominant-negative isoform of EGFR (EGRFDN) impairs olfactory memory formation of Drosophila larvae after bidirectional conditioning [68]. Yet, expression of EGFRDN in all KCs using the OK107-Gal4 did not affect lARM (Fig 5D). Please note that our results do not exclude a potential contribution for these genes at later time points after three cycle standard training.

Therefore, biochemical pathways independent of cAMP/PKA signaling cascades have to be involved in lARM tested directly after three cycle standard training. PKC signaling (Fig 5E) may serve this function as tur mutants that have a reduced PKC activity are impaired in olfactory learning in adult Drosophila [69]. Furthermore, expression of a truncated constitutively active isoform of PKC (PKCζ) rescues the memory defects of rsh1 mutants [70]. In fact, transgenic larvae expressing a specific peptide inhibitor of PKC (PKCi) in all KCs using OK107-Gal4 showed strongly reduced anesthesia resistant learning and/or memory in contrast to both genetic controls (Fig 5F). Control experiments revealed no gross defects in task-relevant sensory-motor abilities (S5 Fig.). Thus, we conclude that the formation, consolidation and retrieval lARM tested directly after training is independent of the cAMP/PKA pathway and may instead require PKC signaling in KCs.

Dopamine signaling is necessary for lARM formation

The current model for associative learning in Drosophila states that during training the unconditioned punishing stimulus is mediated by a specific set of dopaminergic neurons onto MB KCs via G-protein receptor signaling [42, 7174]. In Drosophila the dopamine D1-like receptor family that includes two different dopamine receptors, called dDA1 and DAMB, was reported to be necessary for larval and adult learning [75, 76]. Generally, activation of D1-like receptors was shown to be linked with cAMP/PKA-signaling via Gαs signaling (Fig 5A) [77]. Yet more recently it was reported that D1-like receptors also activate phospholipase C (PLC) via the Gαo signaling, which leads to an activation of PKC (Fig 5E) [77]. Thus we were wondering if lARM formation depends on dopaminergic signaling. In line with prior results, we found that mutants for both receptor genes dumb2 (for dDA1) and damb (for DAMB) show a defect in anesthesia resistant learning and/or memory following three cycles standard training (Fig 6A, S6A Fig.) [72]. In addition, fumin (fmn) mutant larvae that have a mutation in the dopamine transporter (dDAT) gene [78]—and thus have enhanced DA levels in adults [79]–also show an impairment in anesthesia resistant learning and/or memory (Fig 6B) (for further details see S6B and S6C Fig). Finally, acute oral administration of methylphenidate (“Ritalin”, MPH) rescued the behavioral phenotype of rsh1 mutant larvae in a dose-dependent manner (Fig 6C and S6D Fig). MPH application in Drosophila, similar to its function in humans, increases DA levels by inhibiting the dopamine transporter (dDAT) that mediates dopamine reuptake from the synaptic cleft [80]. In adults it was shown that oral administration of MPH rescues deficits in optomotor responses of rsh1 mutants [81]. Summarizing, three different experiments suggest that dopaminergic signaling is necessary to establish, consolidate and/or retrieve lARM. Please note that–although the effect of DA is likely specific for the establishment of the memory—our experiments did not allow to distinguish between the three different processes.

thumbnail
Fig 6. Dopaminergic signaling is necessary for odor-high salt learning and/or memory following three cycle standard training

Training and methylphenidate treatment protocols are shown at the top of each panel. A: Odor-high salt learning and/or memory was reduced in damb as well as in dumb2 dopamine (DA) receptor mutants compared to their respective controls (Unpaired t test, p = 0.002 for damb, Mann-Whitney test, p = 0.001 for dumb2). B: An increase in DA signaling through mutating the dopamine transporter (DAT) in fumin (fmn) mutant larvae leads to reduction of odor-high salt learning and/or memory compared to control larvae (Unpaired t test, p = 0.015). C: Effect of methylphenidate (MPH) treatment prior of three cycle standard training on odor-high salt learning and/or memory. Larvae were fed MPH for one hour in order to impair DAT function. MPH application was done at a concentration of 0.0 (control), 0.5 and 2.0mM. An acute increase in dopaminergic signaling through a reversible blockage of DAT via MPH application leads to a restoration of odor-high salt learning and/or memory in rsh1 mutants (One sample t test, p = 0.04 for 0.5 mM, p = 0.003 for 2.0 mM). Significant differences between wild type and rsh1 mutants was seen for 0.0 mM and 0.5 mM, but not for 2 mM MPH (Bonferroni post hoc pairwise comparisons, p<0.0001, p = 0.015 and p = 0.412, respectively). Sample size is n = 16 for each group if not indicated otherwise. In Fig 6A and 6B differences between groups are depicted above the respective box plots, at which * indicates p<0.05. Grey boxes show memory performance above chance level (p<0.05). Small circles indicate outliers. In Fig 6C differences between groups are depicted below the symbols, at which ns indicates p≥0.05 and * p<0.05. Memory performance significantly different from random distribution (p<0.05) is indicated in black, random distribution (p≥0.05) in light grey. The data are shown as means ± s.e.m.

https://doi.org/10.1371/journal.pgen.1006378.g006

One cycle odor-high salt conditioning establishes two distinct memory phases in Drosophila larvae

Conditioning Drosophila larvae via three cycle standard training takes about 45 minutes. Yet, two studies on larval aversive olfactory learning suggest that short lasting memory phases exist that are only detectable up to 20 or 50 minutes after training onset [36, 42]. These results could mean that three cycle standard training–routinely used in our previous experiments–based on its temporal dimension does not allow to identify short lasting memories. Therefore we established a one cycle training paradigm that only takes about 10 minutes for conditioning (Fig 7). Significant aversive olfactory memory was evident 0, 10, 20 and 60 minutes after training (Fig 7A and S7A Fig.). To our surprise, one cycle training increased aversive memory compared to three cycle training (both groups were tested immediately after training) (S7B Fig.).

thumbnail
Fig 7. Aversive olfactory memory after one cycle odor-high salt conditioning consist of two different components.

Training and cold shock treatment protocols are shown at the top of each panel. A: Aversive olfactory memory following one cycle training was tested at different time points after conditioning. 60 minutes after training the memory was still detectable (One sample t test, p<0.0001 for each group). B: Cold shock application immediately after one cycle odor-high salt training reduced aversive olfactory learning and/or memory. Yet, the effect was only detectable in comparison to control animals when tested 10 minutes after one cycle training (Bonferroni post hoc pairwise comparisons p<0.05). No difference was seen between both groups 20 and 60 minutes after one cycle training (Bonferroni post hoc pairwise comparisons p>0.05). C: rsh1 mutants compared to wild type control animals showed reduced or completely impaired aversive olfactory learning and/or memory after one cycle training (p<0.0001, for both). When tested directly after one cycle training rsh1 mutants showed aversive olfactory memory, however, at a lower level than wild type larvae (Unpaired t test, p<0.0001). D: rut2080 and dnc1 mutants compared to wild type control animals showed reduced aversive olfactory learning and/or memory tested immediately after one cycle training (Bonferroni post hoc pairwise comparisons p<0.0001 for both). No difference was detectable for both mutants when tested 20 minutes after conditioning. Furthermore no differences in the memory performance for the mutants were observable between 0 and 20 minutes (p<0.0001 for each comparison). E: syn97 loss-of-function mutants showed reduced aversive olfactory learning and/or memory tested immediately after a single-training cycle (Unpaired t test, p<0.0001). However, the mutation in syn97 did not completely abolish aversive olfactory learning and/or memory (One sample t test, p<0.0001). F: Compared to both genetic controls (OK107-Gal4/+ and UAS-brp-RNAiB3C8), knockdown of the presynaptic protein brp in MB KCs by driving UAS-brp-RNAiB3C8 via OK107-Gal4 reduced aversive olfactory learning and/or memory tested immediately after one cycle training (Kruskal-Wallis, p = 0.0001). In contrast to three cycle standard training, abolishment of aversive olfactory learning and/or memory tested immediately after one cycle training was only partially (One sample t test, p<0.0001). Sample size is n = 16 for each group if not indicated otherwise. In Fig 7B–7D differences between groups are depicted below the data; ns indicated p≥0.05 and * p<0.05. Memory performance significant different from random distribution (p<0.05) is indicated in black, whereas random distribution (p≥0.05) was shown in light grey. The data are shown as means ± s.e.m. In Fig 7E and 7F differences between groups are depicted above the respective box plots; * indicates p<0.05. Different lowercase letters indicated statistical significant differences from each other at level p<0.05. Grey boxes indicate a memory performance above chance level (p<0.05). Small circles indicate outliers.

https://doi.org/10.1371/journal.pgen.1006378.g007

Next, we tested whether odor-high salt memory following one cycle training is resistant to anesthesia. As shown in Fig 7B (see also S7D Fig.) applying a cold shock treatment did partially disrupt odor-high salt memory tested 10 minutes after training (10 minutes is necessary for recovery from the cold). In contrast, memory tested 20 or 60 minutes after one cycle training was cold shock resistant (Fig 7B). Based on these results we conclude that—independent of the number of training trials—odor-high salt conditioning leads to the formation of lARM. However, at the same time a second short lasting memory is established that can only be detected for up to 30 minutes after training onset. Therefore the short lasting memory can only be analyzed after one cycle training but not in longer lasting protocols using two or three training cycles (S7C Fig.). This conclusion is further supported by two additional findings. First, genetic interference with rsh and brp gene function, both involved in the formation, consolidation and/or retrieval of lARM, tested immediately after one cycle training only partially impaired the performance of experimental larvae (Fig 7C and 7F). These results are different than the ones obtained following three cycle standard training (Fig 3 and Fig 4C) since they imply the presence of a second, lARM independent, memory phase. Second, rut2080, dnc1 and, syn97 mutants tested immediately following one cycle training performed on a lower level than genetic controls (Fig 7D and 7E). Again, the results are different compared to the ones obtained following three cycle standard training (Fig 4B and Fig 5B) and suggest that the formation, consolidation and/or retrieval of a larval short lasting memory (lSTM) under these circumstance depends on cAMP signaling.

Discussion

Drosophila larvae are able to establish an anesthesia resistant form of memory

Memory formation and consolidation usually describes a chronological order, parallel existence or completion of distinct short-, intermediate- and/or long-lasting memory phases. For example, in honeybees, in Aplysia, and also in mammals two longer-lasting memory phases can be distinguished based on their dependence on de novo protein synthesis [8285]. In adult Drosophila classical odor-electric shock conditioning establishes two co-existing and interacting forms of memory—ARM and LTM—that are encoded by separate molecular pathways [18].

Seen in this light, memory formation in Drosophila larvae established via classical odor-high salt conditioning seems to follow a similar logic. It consist of lSTM and lARM (Fig 8A) (for a spaced training protocol see also S1D Fig.). Aversive olfactory lSTM was already described in two larval studies using different negative reinforcers (electric shock and quinine) and different training protocols (differential and absolute conditioning) [36, 42]. Our results introduce for the first time lARM that was also evident directly after conditioning but lasts longer than lSTM (Fig 8A). lARM was established following different training protocols that varied in the number of applied training cycles (S7C Fig.) and the type of negative or appetitive reinforcer (Fig 2C–2F). Thus, lSTM and lARM likely constitute general aspects of memory formation in Drosophila larvae that are separated on the molecular level.

thumbnail
Fig 8. A molecular working hypothesis for lARM formation

A: Memory formation in Drosophila larvae occurs through at least two different components, which are genetically and functionally distinct. First, larval short-term memory (lSTM, light blue) occurs immediately after training, but decays within 20 minutes. Second, larval anesthesia-resistant memory (lARM, light magenta) also appears immediately after training, but lasts for a longer period of time. In contrast to lSTM lARM is resistant to anesthetic disruption. At any given time interval after training the observed memory is the summed output of both components. B: Memory formation in Drosophila larvae after classical olfactory conditioning depends on protein kinase A (PKA) and protein kinase C (PKC) function, which are involved in two different signaling pathways. During conditioning MB KCs receive an olfactory stimulus via cholinergic projection neurons (conditioned stimulus CS) and a punishing stimulus from dopaminergic neurons (unconditioned stimulus US). In KCs binding of dopamine leads to a dissociation of a G protein subunit from G-protein coupled receptors (GPCR). The CS is perceived by KCs via acetylcholine receptors (AChR). Its activation leads to an opening of a voltage-dependent calcium channel (VDCC) and leads to an intracellular Ca2+ influx. This mechanism is thought to be shared between both signaling pathways. In the classical pathway (left side, molecular elements tested in this study are highlighted in light blue), coincident stimulation of CS and US leads to an activation of type I adenylyl cyclase (AC) via Ca2+/Calmodulin and dopamine dependent G protein (Gαs) signaling, respectively. Active AC catalyzes the intracellular cAMP production, which is negatively regulated through a phosphodiesterase (PDE) to maintain cAMP concentrations at a tolerable level. cAMP serves as a regulatory signal for PKA, which phosphorylates different substrates like Synapsin (Syn) or CREB in order to induce short or long-lasting presynaptic plasticity. PKC signaling (right side, molecular elements tested in this study are highlighted in light magenta) has been shown in different species as an integral pathway for memory formation. Dopamine receptors were reported to couple to Gαq to regulate phospholipase C (PLC). Activation of PLC increases intracellular inositol triphosphate (IP3) and diacylglycerol (DAG) levels. Whereas IP3 stimulates the release of Ca2+ from the endoplasmic reticulum, DAG is a physiological activator of PKC. The downstream elements of PKC are not well described. We suggest that lSTM formation depends on the classical cAMP/PKA pathway, whereas early lARM formation depends on PKC signaling rather than on cAMP/PKA signaling. This assumption is based on the analysis of nine different genes (light purple: lARM phenotype, light blue: lSTM phenotype). In addition, we suggest that PKC can be linked with Radish (Rsh) and Bruchpilot (Brp) as direct or indirect downstream partners. This is so far hypothetical. Yet, a structural analysis on Rsh reported that it has several PKC phosphorylation sites. Ultimately, regulation of Brp via PKC signaling would change the organization of the active zone to provide a molecular substrate for presynaptic plasticity. DA: dopamine, Ach: Acetylcholine, CaM: Calmodulin, RP: reserve pool, RRP: readily releasable pool.

https://doi.org/10.1371/journal.pgen.1006378.g008

Molecular pathway underlying aversive learning and memory in Drosophila larvae

Memory formation depends on the action of distinct molecular pathways that strengthen or weaken synaptic contacts of defined sets of neurons (reviewed in [1, 73, 8688]). The cAMP/PKA pathway is conserved throughout the animal kingdom and plays a key role in regulating synaptic plasticity. Amongst other examples it was shown to be crucial for sensitization and synaptic facilitation in Aplysia [1, 86], associative olfactory learning in adult Drosophila and honeybees [85, 88], long-term associative memory and long-term potentiation in mammals [8992].

For Drosophila larvae two studies by Honjo et al. [42] and Khurana et al. [36] suggest that aversive lSTM depends on intact cAMP signaling. In detail, they showed an impaired memory for rut and dnc mutants following absolute odor-bitter quinine conditioning [42] and following differential odor-electric shock conditioning [36]. Thus, both studies support our interpretation of our results. We argue that odor-high salt training established a cAMP dependent lSTM due to the observed phenotypes of rut, dnc and syn mutant larvae (Fig 7D and 7E). The current molecular model is summarized in Fig 8B. Yet, it has to be mentioned that all studies on aversive lSTM in Drosophila larvae did not clearly distinguish between the acquisition, consolidation and retrieval of memory. Thus, future work has to relate the observed genetic functions to these specific processes.

In contrast, lARM formation utilizes a different molecular pathway. Based on different experiments, we have ascertained, that lARM formation, consolidation and retrieval is independent of cAMP signaling itself (Fig 5B), PKA function (Fig 5C), upstream and downstream targets of PKA (Figs 5D and 4B) and de-novo protein synthesis (Fig 2A and 2B) (but see also for spaced conditioning S1D Fig.). Instead we find that lARM formation, consolidation and/or retrieval depends on rsh gene function (Fig 3), brp gene function (Fig 4C), dopaminergic signaling (Fig 6) and requires presynaptic signaling of MB KCs (Fig 4A and 4C and S4C–S4E Fig.).

Interestingly, studies on adult Drosophila show that rsh and brp gene function, as well as dopaminergic signaling and presynaptic MB KC output are also necessary for adult ARM formation [2325, 59]. Thus, although a direct comparison of larval and adult ARM is somehow limited due to several variables (differences in CS, US, training protocols, test intervals, developmental stages, and coexisting memories), both forms share some genetic aspects. This is remarkable as adult ARM and lARM use different neuronal substrates. The larval MB is completely reconstructed during metamorphosis and the initial formation of adult ARM requires a set of MB α/β KCs that is born after larval life during puparium formation [25, 93, 94].

In addition, we have elicited the necessity of PKC signaling for lARM formation in MB KCs (Fig 5C). The involvement of the PKC pathway for memory formation is also conserved throughout the animal kingdom. For example, it has been shown that PKC signaling is an integral component in memory formation in Aplysia [9598], long-term potentiation and contextual fear conditioning in mammals [99101] and associative learning in honeybees [102]. In Drosophila it was shown that PKC induced phosphorylation cascade is involved in LTM as well as in ARM formation [103]. Although the exact signaling cascade involved in ARM formation in Drosophila still remains unclear, we established a working hypothesis for the underlying genetic pathway forming lARM based on our findings and on prior studies in different model organisms (Fig 8B). Thereby we do not take into account findings from Horiuchi et al. [66] and Scheunemann et al. [104] in adult Drosophila. These studies show that PKA mutants have increased ARM [66] and that dnc sensitive cAMP signaling supports ARM [104]. Thus both studies directly link PKA signaling with ARM formation.

Working hypothesis for lARM formation

It was shown that KCs act on MB output neurons to trigger a conditioned response after training [105, 106]. Work from different insects suggests that the presynaptic output of an odor activated KCs is strengthened if it receives at the same time a dopaminergic, punishment representing signal. Our results support these models as they show that lARM formation requires accurate dopaminergic signaling (Fig 6) and presynaptic output of MB KCs (Fig 4A and 4C). Yet, for lARM formation dopamine receptor function seems to be linked with PKC pathway activation (Fig 5). Indeed, in honeybees, adult Drosophila and vertebrates it was shown that dopamine receptors can be coupled to Gαq proteins and activate the PKC pathway via PLC and IP3/DAG signaling [107, 108]. As potential downstream targets of PKC we suggest radish and bruchpilot. Interference with the function of both genes impairs lARM (Figs 3 and 4C). The radish gene encodes a functionally unknown protein that has many potential phosphorylation sites for PKA and PKC [23]. Thus considerable intersection between the proteins Rsh and PKC signaling pathway can be forecasted. Whether this is also the case for the bruchpilot gene that encodes for a member of the active zone complex remains unknown. The detailed analysis of the molecular interactions has to be a focus of future approaches. Therefore, our working hypothesis can be used to define educated guesses. For instance, it is not clear how the coincidence of the odor stimulus and the punishing stimulus are encoded molecularly. The same is true for ARM formation in adult Drosophila. Based on our working hypothesis we can speculate that PKC may directly serve as a coincidence detector via a US dependent DAG signal and CS dependent Ca2+ activation.

Appetitive learning and memory in Drosophila larva

Do our findings in general apply to learning and memory in Drosophila larvae? To this the most comprehensive set of data can be found on sugar reward learning. Drosophila larva are able to form positive associations between an odor and a number of sugars that differ in their nutritional value [31, 32, 47, 109111]. Using high concentrations of fructose as a reinforcer in a three cycle differential training paradigm (comparable to the one we used for high salt learning and fructose learning (Figs 1 and 2F)) Michels et al. [47, 111] found that learning and/or memory in syn97 mutant larvae is reduced to ∼50% of wild type levels. Thus, half of the memory seen directly after conditioning seems to depend on the cAMP-PKA-synapsin pathway. Our results in turn suggest that the residual memory seen in syn97 mutant larvae is likely lARM (Fig 2F). Thus, aversive and appetitive olfactory learning and memory share general molecular aspects. Yet, the precise ratio of the cAMP-dependent and independent components rely on the specificities of the used odor-reinforcer pairings. Two additional findings support this conclusion. First, Kleber et al. [112] recently showed that memory scores in syn97 mutant larvae are only lower than in wild type animals when more salient, higher concentrations of odor or fructose reward are used. Usage of low odor or sugar concentrations does not give rise to a cAMP-PKA-synapsin dependent learning and memory phenotype. Second, Honjo et al. [32] showed that learning and/or memory following absolute one cycle conditioning using sucrose sugar reward is completely impaired in rut1, rut2080 and dnc1 mutants. Thus, for this particular odor-reinforcer pairing only the cAMP pathway seems to be important. Therefore, a basic understanding of the molecular pathways involved in larval memory formation is emerging. Further studies, however, will be necessary in order to understand how Drosophila larvae make use of the different molecular pathways with respect to a specific CS/US pairing.

Materials and Methods

Fly stocks and maintenance

Fly strains were reared on standard Drosophila medium at 25°C or 18°C in constant darkness or with a 14/10 hr light/dark cycle. For behavioral analysis we used rut1, rut2080, dnc1, rsh1 [14, 17, 24] (kindly provided by T. Preat), DCOB3, DCOH2 [27] (kindly provided by M. Saitoe), fumin [78] (kindly provided by M. Heisenberg) and syn97 [54, 111] (kindly provided by B. Gerber) mutants. All lines were outcrossed over several generations with wild type CantonS that was used as a genetic control. In addition, we used the two dopamine receptor mutants damb and dumb2 and their genetic controls [71]. Note, that in contrast to earlier studies the damb mutant was outcrossed to CantonS over several generations. To rescue the rsh dependent phenotype by artificial ubiquitous rsh expression we used rsh1; hs-rsh larvae [23] (kindly provided by T. Zars). To express Gal4 in all larval KCs we used the driver line OK107 [48, 113] (DGRC no.: 106098). UAS-shits was used to acutely block synaptic output from KCs [49] (BDSC No.: 7068). In addition, we used the four effector lines UAS-EGFRDN [68] (kindly provided by T. Roeder), UAS-RNAiB3C8 [59] (kindly provided by H. Tanimoto), UAS-PKCi [114] (kindly provided by B. Brembs) and UAS-Creb2-b [115] (kindly provided by S. Waddell).

Aversive olfactory conditioning

Experiments were conducted on assay plates (85mm diameter, Cat. No.: 82.1472, Sarstedt, Nümbrecht) filled with a thin layer of 2.5% agarose containing either pure agarose (Sigma Aldrich Cat. No.: A5093, CAS No.: 9012-36-6) or agarose plus reinforcer. We used 1.5M and 2.0M sodium chloride (Sigma Aldrich Cat. No.: S7653, CAS No.: 7647-14-5) [44], 2.0M fructose (Sigma Aldrich Cat. No.: 47740, CAS No.: 57-48-7) [110, 116] and 6mM quinine (quinine hemisulfate salt monohydrate, Sigma Aldrich Cat.No.: Q1250, CAS No.: 207671-44-1) [34, 45]. As olfactory stimuli, we used 10 μl amyl acetate (AM, Fluka Cat. No.: 46022; CAS No.: 628-63-7; diluted 1:250 in paraffin oil, Fluka Cat. No.: 76235, CAS No.: 8012-95-1) and benzaldehyde (BA, undiluted; Fluka Cat. No.: 12010, CAS No.: 100-52-7). Odorants were loaded into custom-made Teflon containers (4.5-mm diameter) with perforated lids [109]. Learning ability was tested by exposing a first group of 30 animals to AM while crawling on agarose medium containing in addition sodium chloride as a negative reinforcer. After 5 min, larvae were transferred to a fresh Petri dish in which they were allowed to crawl on pure agarose medium for 5 min while being exposed to BA (AM+/BA). A second group of larvae received the reciprocal training (AM/BA+). If not stated otherwise, three training cycles are used. Depending on the memory retention larvae were transferred onto another agarose plate prior to training and kept for the indicated time before testing the memory. To increase humidity tap water was added. Memory is tested by transferring larvae onto test plates on which AM and BA were presented on opposite sides. For fructose reinforcement the test plates contains pure agarose whereas for sodium chloride and quinine reinforcement agarose plates containing the respective reinforcer are used. After 5 min, individuals were counted as located on the AM side (#AM), the BA side (#BA), or in a 1 cm neutral zone. By subtracting the number of larvae on the BA side from the number of larvae on the AM side, and dividing by the total number of counted individuals (#TOTAL), we determined a preference index for each training group:

  1. (1a) PREFAM+/BA = (# AM—# BA) / # TOTAL
  2. (1b) PREFAM/BA+ = (# AM—# BA) / # TOTAL

To measure specifically the effect of associative learning that is of the odor-reinforcement contingency, we then calculated the associative performance index (PI) as the difference in preference between the reciprocally trained larvae:

  1. (2) PI = (PREFAM+/BA−PREFAM/BA+) / 2

Negative PIs thus represent aversive associative learning, whereas positive PIs indicate appetitive associative learning. Division by 2 ensures scores are bound within (-1; 1).

Heat-shock treatment

Heat shocks were applied for six hours. Therefore, food vials with six days old larvae were transferred into an incubator at 35°C. We heat-shocked the w,rhs1;hs-rsh transgenic flies, rsh1 mutants and as controls wild type and w1118 larvae. Afterwards each group received an aversive olfactory training regime at room temperature.

Cold-shock treatment

For cold-shock experiments larvae were incubated in ice tap water (4°C) for one minute. Larvae were allowed to recover by transferring them onto agarose plates. They started moving within one minute and were kept on the agarose plates at 23°C until testing.

Cycloheximide treatment

To test if aversive olfactory learning is dependent on de novo protein synthesis, wild type larvae were fed 35 mM cycloheximide (+CXM; Sigma Aldrich Cat. No.: C7698; CAS No.: 66-81-9) in 5% sucrose (w/v), 5% sucrose alone (-CXM) or tab water (-CXM, -SUC) for 20 hours before the experiment [10]. Therefore, 300 ml of solution was added into food vials. Before the experiment larvae were gently transferred to an empty Petri dish and washed with tap water before training and testing.

MPH feeding

Methylphenidate (MPH; Sigma Aldrich Cat. No.: M2892; CAS No.: 298-59-9), was orally administered to the larvae, as it was shown that oral consumption of MPH is sufficient to inhibit the Drosophila dopamine transporter [80]. MPH was diluted in tap water with a concentration of 2 mM, stored in the refrigerator and used within 7 days. MPH was applied for one hour to a group of 30 larvae in Petri dishes with an inner diameter of 35 mm (Greiner). Afterwards larvae were gently transferred to an empty Petri dish and washed with tap water before training and testing.

Acutely blocking synaptic output with shibirets

To acutely block synaptic output we used UAS-shits1 [49]. Immediately before the experiment, larvae were incubated for 2 min in a water-bath at 37°C. The behavioral experiments were then performed as described before, at restrictive temperature of about 35°C in a custom made chamber placed within a fume hood. Control experiments were performed with incubation at room temperature and at permissive temperature of about 23°C.

Statistical methods

All statistical analyses and visualizations were done with GraphPad Prism 5.0. Groups that did not violate the assumption of normal distribution (Shapiro–Wilk test) and homogeneity of variance (Bartlett's test) were analyzed with parametric statistics: unpaired t-test (comparison between two groups) or Oneway ANOVA followed by planned pairwise comparisons between the relevant groups with a Tukey honestly significant difference HSD post hoc test (comparisons between groups larger than two). Experiments with data that were significantly different from the assumptions above were analyzed with non-parametric tests, such as Mann–Whitney test (comparison between two groups) or Kruskal–Wallis test followed by Dunn's multiple pairwise comparison (comparisons between groups larger than two). To compare single genotypes against chance level, we used One sample t test or Wilcoxon signed rank test. For statistical test concerned with factors equal two or more, two way ANOVA was applied followed by the planned pairwise multiple comparisons (Bonferroni). The significance level of statistical tests was set to 0.05. Figure alignments were done with Adobe Photoshop. Data were presented as box plots, 50% of the values of a given genotype being located within the boxes, whiskers represent the entire set of data. Outliers are indicated as open circles. The median performance index was either indicated as a bold line and the mean as a cross within the box plot or symbol expressed as means ± s.e.m. Unless stated otherwise, all olfactory conditioning experiments are n = 16.

Immunostaining

Third instar larvae were put on ice and dissected in phosphate-buffered saline (PBS) [71, 117, 118]. Brains were fixed in 3.6% formaldehyde (Merck, Darmstadt) in PBS for 30 min. After eight times rinsing in PBT (PBS with 3% Triton-X 100, Sigma-Aldrich, St. Louis, MO), brains were blocked with 5% normal goat serum (Vector Laboratories, Burlingame, CA) in PBT for 2 hours and then incubated for two days with primary antibodies at 4°C. Before applying the secondary antibodies for two days at 4°C, brains were washed eight times with PBT. After secondary antibody incubation, brains were washed eight times with PBT and mounted in Vectashield (Vector Laboratories, Burlingame, CA) between two cover slips and stored at 4°C in darkness. Images were taken with a Zeiss LSM 510M confocal microscope with x25 or x40 glycerol objectives. The resulting image stacks were projected and analyzed with Image-J (National Institutes of Health, Bethesda, Maryland, http://imagej.nih.gov/ij) software. Contrast and brightness adjustment as well as rotation and organization of images were performed in Photoshop (Adobe Systems Inc., San Jose, CA).

Antibodies

To analyze the expression pattern of OK107-Gal4 rabbit anti-GFP antibody (A6455, Molecular Probes; 1:1000) and two different mouse antibodies for staining the cholinergic neuropil (ChAT4B1; DSHB, Iowa City, IA, 1:150) and axonal tracts (1d4 anti-Fasciclin 2; DSHB, Iowa City, IA; 1:50) were applied [71, 117]. A specific antibody for the Synapsin protein was used to verify the mutation syn97 (monoclonal mouse anti-syn, 3C11; DSHB, Iowa City, IA, 1:10) [54]. To analyze if the expression level of Bruchpilot is specifically reduced in the MB KCs by driving UAS-RNAiB3C8 via OK107-Gal4, monoclonal mouse anti-nc82 was used (nc82, DSHB, Iowa City, IA, 1:10) [58]. As secondary antibodies goat anti-rabbit IgG Alexa Fluor 488 (A11008, Molecular Probes, 1:200) and goat anti-mouse IgG Alexa Fluor 647 (A21235, Molecular Probes, 1:200) were used.

Supporting Information

S1 Fig. Aversive olfactory learning and/or memory is independent of de-novo protein synthesis.

Training and different treatment protocols are shown at the top of each panel. A: Cycloheximide (CXM) treatment applied before training did not reduce aversive olfactory learning and/or memory of wild type larvae. For all three groups aversive olfactory learning and/or memory tested immediately after three cycle standard training was significantly different from random distribution (One sample t test, p<0.0001, respectively) and not significantly different from each other (One way ANOVA, p = 0.33). For all three groups aversive olfactory learning and/or memory tested 60 minutes after three cycle standard training was significantly different from random distribution (One sample t test, p<0.0001, p = 0.0004, and p<0.0001, respectively) and not significantly different from each other (Kruskal-Wallis, p = 0.29). B: CXM treatment prevented wild type larvae from pupation (after 5 days) and eclosion (after 10 days) after metamorphosis (red line). Control groups (blue lines) that were raised on standard food or on a sucrose diet showed no effect. Results are shown as means and s.e.m. For each group 10 repetitions were done. A significant number of surviving animals is indicated in black (p<0.05), whereas a non-significant number of pupae or flies is marked in light grey (p≥0.05).C: Aversive olfactory learning and/or memory was not affected when interfering with CREB function of mushroom body Kenyon cells (MB KCs) using UAS-dCREb2-b and OK107-Gal4. Experimental OK107-Gal4/UAS-dCREB2b larvae showed learning and/or memory (One sample t test, p<0.0001, respectively) comparable to two genetic controls (Kruskal-Wallis, p = 0.15) when tested immediately after three cycle standard training. Experimental OK107-Gal4/UAS-dCREB2b larvae showed learning and memory tested 60 minutes after three cycle training (One sample t test, p = 0.0004, p = 0.006 and p = 0.002, respectively) comparable to two genetic controls (One way ANOVA, p = 0.64). D: We trained larvae using a spaced training protocol consisting of five training cycles with 15 minutes rest intervals in between. Aversive olfactory learning and/or memory is completely abolished when interfering with CREB function of MB KCs using UAS-dCREb2-b/OK107-Gal4 larvae (One sample t test, p = 0.180). Both genetic controls showed aversive olfactory learning and memory indistinguishable from each other (Tukey post hoc test, p = 0.916). Sample size is n = 16 for each group if not indicated otherwise. In S1A,S1C,S1D Fig. differences between groups are depicted above the respective box plots; ns indicates p≥0.05. Different lowercase letters indicate statistical significant differences from each other at level p<0.05. Grey boxes show memory performance above chance level (p<0.05), whereas white boxes indicate random distribution (p≥0.05). Small circles indicate outliers in S1B Fig.

https://doi.org/10.1371/journal.pgen.1006378.s001

(TIF)

S2 Fig. Drosophila larvae establish an anesthesia resistant type of memory following odor-high salt conditioning.

Training and different treatment protocols are shown at the top of each panel. A: Wild type larvae were trained using the standard three cycle odor-high salt conditioning paradigm. Three different groups were tested: An experimental group received a cold shock (one minute in a 4°C ice water bath) directly after conditioning. A second group received the same cold shock treatment before the conditioning phase. A third group was not cold shocked. Memory was tested after a short recovering phase of 10 minutes after conditioning. Cold shock treatment before or after three cycle standard training did not affect learning and memory of the larvae (One sample t test, p<0.0001, p = 0.0002, p<0.0001 and p<0.0001, respectively). All three groups performed on the same level (One way ANOVA, p = 0.68). B: In addition, we also tested if a harsh cold shock treatment of 5 minutes at 4°C that completely paralyses the larvae affected the performance of the animals. Also under these conditions experimental larvae showed learning and memory that was resistant to cold shock anesthesia (Two way ANOVA, p = 0.07 comparing duration of the cold shock, p = 0.76 comparing if cold shock treatment was applied or not). C: Aversive olfactory learning and memory tested at 10, 60, 120 and 180 minutes after conditioning. Experimental groups received a cold shock directly after three cycle standard training. In all four cases cold shock treated larvae behaved on a comparable level as control groups (Unpaired t test, p = 0.4, p = 0.5, p = 0.68 and p = 0.16, respectively). D: Aversive olfactory memory was tested 60 minutes after three cycle standard training; experimental groups received a cold shock 0, 10, 20 or 40 minutes after conditioning. In all four cases cold shock treated larvae behaved as control groups (Unpaired t test, p = 0.5, p = 0.88, p = 0.71 and p = 0.79, respectively). Sample size is n = 16 for each group if not indicated otherwise. In S2A,S2C,S2D Fig. differences between groups are depicted above the respective box plots; ns indicates p≥0.05. Grey boxes indicate a performance above chance level (p<0.05). In S2B Fig. differences between groups are depicted below the symbols; ns indicates p≥0.05. A performance significant different from random distribution (p<0.05) is indicated in black. The data are shown as means ± s.e.m.

https://doi.org/10.1371/journal.pgen.1006378.s002

(TIF)

S3 Fig. Sensory acuity tests for larvae with impaired radish gene function.

A: Schematic representation of chemosensory acuity tests. Olfactory perception is analyzed by putting 30 larvae in the middle of a Petri dish with either an amyl acetate (AM) or a benzaldehyde (BA) containing odor container on one side and an empty container (EC) on the other side. After 5 minutes larvae are counted to calculate an olfactory preference index. For gustatory acuity tests, 30 larvae are put in the middle of a Petri dish that contained pure agarose on one side and agarose plus a high salt concentration on the other side. After 5 minutes larvae were counted to calculate a gustatory preference index. B: Naive olfactory and gustatory acuity tests for rsh1 mutant and wild type control larvae. Olfactory preference for AM of rsh1 mutant larvae were not different from the one of wild type controls (Unpaired t test, p = 0.30). rsh1 mutant larvae, however, did not show any significant preference for BA (One sample t test, p = 0.23). Gustatory avoidance for high-salt concentration of rsh1 mutant larvae is not different from the one of wild type controls (Unpaired t test, p = 0.84). C: Due to the fact that rsh1 mutants showed an impaired BA preference, we applied a one odor paradigm. In contrast to three cycle standard training, larvae received only AM and instead of BA paraffin oil (no odor information) during odor-high salt conditioning (one odor paradigm). In line with the results for two odor conditioning rsh1 mutant larvae behaved significantly different compared to wild type control larvae (Unpaired t test, p = 0.0001) and showed no learning and/or memory (One sample t test, p = 0.26). The training protocol is shown at the top of the panel. D: Naive olfactory and gustatory acuity tests of experimental and control larvae used to rescue rsh gene function. Olfactory preference for AM for rsh1 and w,rsh1;hs-rsh were comparable to their controls at both temperatures (Mann-Whitney test, p = 0.24, p = 0.82, p = 0.22 and p = 0.10, respectively). However, again rsh1 larvae showed significant differences in their BA preference compared to controls (Unpaired t test, p<0.0001, p = 0.49, p<0.0001 and p = 0.68, respectively). High-salt avoidance of rsh1 and w,rsh1;hs-rsh larvae was indistinguishable from the behavior of controls (Mann-Whitney test, p = 0.06, p = 0.36, p = 0.11 and p = 0.34, respectively). Sample size is n = 16 for each group if not indicated otherwise. Differences between groups are depicted above the respective box plots; ns indicates p≥0.05 and * p<0.05. Grey boxes indicate a memory performance above chance level (p<0.05), whereas white boxes indicate a memory performance at chance level (p≥0.05). Small circles indicate outliers.

https://doi.org/10.1371/journal.pgen.1006378.s003

(TIF)

S4 Fig. Presynaptic output of mushroom body Kenyon cells is not necessary for naïve behaviors towards olfactory and gustatory stimuli.

A: Sensory acuity tests when interfering with neuronal output of mushroom body Kenyon cells (MB KCs) using UAS-shits1 and OK107-Gal4. Experimental (OK107-Gal4 /UAS-shits1) and control larvae (OK107-Gal4 /+ and UAS-shits1) showed no difference in their naïve responses to AM, BA and high salt (for AM: One way ANOVA, p = 0.25, for BA: Kruskal-Wallis, p = 0.09 and for high salt: One way ANOVA, p = 0.08). B: Sensory acuity tests when knocking down brp in the MB KCs via UAS-brp-RNAiB3C8 and OK107-Gal4. Experimental (OK107-Gal4/UAS-brp-RNAiB3C8) and control larvae (OK107-Gal4/+, UAS-brp-RNAiB3C8/+) showed no difference in their naïve responses to AM, BA and high salt (Kruskal-Wallis, p = 0.54, p = 0.27 and p = 0.68, respectively). C: Blockade of presynaptic output of MB KCs via UAS-shits1 using another driver line H24 completely impaired aversive olfactory learning and/or memory. Larvae were raised at the permissive temperature (19°C) and shifted to restrictive temperature before and during three cycle standard training and testing. In contrast to both genetic controls aversive olfactory learning and/or memory tested immediately after three cycle standard training was completely abolished in H24-Gal4/UAS-shits1 larvae (One sample t test, p<0.0001 for both control groups and p = 0.853 for H24-Gal4/UAS-shits1). D: Sensory acuity tests when interfering with neuronal output of mushroom body Kenyon cells (MB KCs) using UAS-shits1 and H24-Gal4. Experimental (H24-Gal4/UAS-shits1) and control larvae (H24-Gal4/+, UAS-shits1/+) showed no difference in their naïve responses to AM, BA and high salt (One way ANOVA, p = 0.517, p = 0.184 and p = 0.753, respectively). E: Shows a frontal view projection (left) of a H24-Gal4/UAS-mCD8::GFP larval hemispheres labeling the entire set of MB KCs (anti-GFP in green and anti-FasII, anti-ChAT neuropil staining in magenta). The observed staining is nearly specific for the larval MB. Below a zoom in of the MB is shown. Further below only the GFP channel is depicted. Scale bars: upper panel 50μm, middle and lower panel 25μm. Sample size is n = 16 for each group if not indicated otherwise. Differences between groups are depicted above and below the respective box plots; ns indicates p≥0.05. Grey boxes indicate a memory performance above chance level (p<0.05). Small circles indicate outliers.

https://doi.org/10.1371/journal.pgen.1006378.s004

(TIF)

S5 Fig. Sensory acuity tests of larvae with impaired PKC function.

A: Sensory acuity tests when suppressing PKC activity in the MB KCs using an inhibitory pseudo substrate of PKC (PKCi). For more details see S3A Fig. The inhibition of PKC did not alter the perception of AM, BA or high salt. Experimental and control groups were indistinguishable from each other (One way ANOVA, p = 0.99, p = 0.40 and p = 0.50, respectively). Sample size is n = 16 for each group if not indicated otherwise. Differences between groups are depicted above or below the respective box plots, ns indicates p≥0.05. Grey boxes indicate a memory performance above chance level (p<0.05). Small circles indicate outliers.

https://doi.org/10.1371/journal.pgen.1006378.s005

(TIF)

S6 Fig. Sensory acuity test for larvae with impaired dopaminergic signaling.

Training and methylphenidate (MPH) treatment protocols are shown at the top of each panel. A: Sensory acuity tests of the dopamine (DA) receptor mutants dumb2 and damb. Both receptors mutants perceived AM, BA and high salt stimuli comparable to controls (for AM: unpaired t test, p = 0.25 and p = 0.83, for BA: Mann-Whitney test, p = 0.07 and unpaired t test, p = 0.41, for high salt: unpaired t test, p = 0.57 and p = 0.28). B: Sensory acuity tests of the DAT mutant fumin (fmn). fmn mutant larvae showed no difference in their naïve responses to AM and high salt (Unpaired t test, p = 0.10 and p = 0.64, respectively). However, the naïve response to BA was impaired (Unpaired t test, p = 0.06). C: In line with the results for two odor conditioning, fmn mutant larvae using a one odor learning paradigm (to omit BA as a sensory stimulus) behaved significantly different compared to wild type control larvae (paired t test, p = 0.0003). They showed no aversive olfactory learning and/or memory (One sample t test, p = 0.06). D: Aversive olfactory memory in rsh1 and wild type control larvae after methylphenidate (MPH) treatment using different concentrations. Memory was tested directly after three cycle standard training. Aversive olfactory learning and memory was indistinguishable from random distribution for the rsh1 mutant without MPH application (One sample t test, p = 0.94). After application of 0.5 mM MPH rsh1 mutant larvae showed reduced learning and/or memory compared to wild type controls (One sample t test, p = 0.03, unpaired t test, p = 0.01). After application of 2.0 mM MPH rsh1 mutant larvae performed as wild type controls (Unpaired t test, p = 0.11). Sample size is n = 16 for each group if not indicated otherwise. Differences between groups are depicted above or below the respective box plots; ns indicates p≥0.05 and * p<0.05. Grey boxes indicate a memory performance above chance level (p<0.05), whereas white boxes indicate a memory performance at chance level (p≥0.05). Small circles indicate outliers.

https://doi.org/10.1371/journal.pgen.1006378.s006

(TIF)

S7 Fig. Aversive olfactory high salt reinforced learning and memory using only one training trial.

Training and cold shock treatment protocols are shown at the top of each panel. A: Aversive olfactory learning and/or memory using a one cycle training protocol was tested at four different retention times after conditioning (0, 10, 20 and 60 minutes). Statistical significant differences were revealed between the groups (One way ANOVA, p<0.0001). The performance indices measured immediately and 10 minutes after one cycle training were indistinguishable from each other (Tukey post hoc test, p = 0.945). Alike the performance indices measured 20 and 60 minutes after once cycle training showed no statistical significance differences (Tukey post hoc test, p = 0.874). However the performance indices measured 0 and 10 minutes after conditioning where at a higher level than the ones measured 20 and 60 minutes after conditioning when analyzed with Tukey post hoc test (p<0.0001 for 0 and 20, p<0.0001 for 0 and 60, p = 0.0001 for 10 and 20 and p<0.0001 for 10 and 60). B: Aversive olfactory learning and memory established with increasing training cycles (One cycle, two cycles and three cycles) revealed significant differences between one cycle training and two or three cycle training (One way ANOVA, p = 0.003; Tukey post hoc test p = 0.037, p = 0.033, respectively). For two and three cycle training no difference was detected (p = 0.705). C: Cold shock treatment directly after training partially impaired aversive olfactory learning and/or memory when tested 10 minutes after one cycle training. No difference was detected between cold shock treated and control larvae tested 10 minutes after two and three cycle training (Unpaired t-test, p = 0.0001 for one training cycle, p = 0.623 for two training cycles and p = 0.396 for three training cycles). D: Aversive olfactory learning and memory of cold shock treated larvae and control groups tested 10, 20 and 60 minutes after one cycle training. Only cold shocked treated larvae tested 10 minutes after one cycle training showed a significant reduction compared to controls that perceived no cold shock. No effect was seen between cold shocked and control groups tested 20 and 60 minutes after one cycle training (Unpaired t test, p = 0.0001 for 10 minutes, p = 0.934 for 20 minutes and p = 0.681 for 60 minutes). E: Aversive olfactory learning and/or memory of rsh1 mutants measured immediately and 20 minutes after one cycle training was reduced compared to wild type larvae (Unpaired t test, p<0.0001 for both). Furthermore, rsh1 mutant revealed a complete loss of aversive olfactory learning and/or memory only when measured 20 minutes after one cycle training but not when measured immediately after one cycle training (One sample t-test, p = 0.383). F: Aversive olfactory learning and/or memory of rut2080 and dnc1 mutants was reduced compared to wild type larvae when measured directly after one cycle training (One way ANOVA, p<0.0001; Tukey post hoc test, p<0.0001 for both compared to wild type control and p = 0.902 compared to each other). For both mutants learning and memory was not completely abolished (One sample t test, p<0.0001 for both). Aversive olfactory learning and memory of both mutants tested 20 minutes after one cycle training was indistinguishable from the one of wild type larvae (One way ANOVA, p = 0.106).

https://doi.org/10.1371/journal.pgen.1006378.s007

(TIF)

Acknowledgments

We thank Lyubov Pankevych and Margarete Ehrenfried for fly care and maintenance. Furthermore we thank Claudius Neumann, Tilman Triphan, Wolf Hütteroth, and Veit-Johannes Widmann for proof reading the manuscript. We would like to express special thanks to Scott Waddell for his critical comments on this work. The anatomical analyses were supported by the Bioimaging Facility of the University of Konstanz.

Author Contributions

  1. Conceptualization: AW AST.
  2. Data curation: AW MA LB KB CP JS MS AST.
  3. Formal analysis: AW MA LB KB CP JS MS AST.
  4. Funding acquisition: AST.
  5. Investigation: AW MA LB KB CP JS.
  6. Methodology: AW AST.
  7. Project administration: AW AST.
  8. Resources: MS.
  9. Supervision: AW AST.
  10. Validation: AW MA LB KB CP JS MS AST.
  11. Visualization: AW MA LB KB CP JS MS AST.
  12. Writing – original draft: AW AST.
  13. Writing – review & editing: AW AST.

References

  1. 1. Kandel ER, Dudai Y, Mayford MR. The molecular and systems biology of memory. Cell. 2014;157(1):163–86. pmid:24679534
  2. 2. Menzel R. The honeybee as a model for understanding the basis of cognition. Nature reviews Neuroscience. 2012;13(11):758–68. pmid:23080415
  3. 3. Pavlov IP. Conditioned Reflexes.: Oxford University Press.; 1927.
  4. 4. Davis HP, Squire LR. Protein synthesis and memory: a review. Psychological bulletin. 1984;96(3):518–59. pmid:6096908
  5. 5. Emptage NJ, Carew TJ. Long-term synaptic facilitation in the absence of short-term facilitation in Aplysia neurons. Science. 1993;262(5131):253–6. pmid:8211146
  6. 6. McGaugh JL. Memory—a century of consolidation. Science. 2000;287(5451):248–51. pmid:10634773
  7. 7. Dudai Y. Consolidation: fragility on the road to the engram. Neuron. 1996;17(3):367–70. pmid:8816699
  8. 8. Duncan CP. The retroactive effect of electroshock on learning. Journal of comparative and physiological psychology. 1949;42(1):32–44. pmid:18111554
  9. 9. Lechner HA, Squire LR, Byrne JH. 100 years of consolidation—remembering Muller and Pilzecker. Learn Mem. 1999;6(2):77–87. pmid:10327233
  10. 10. Tully T, Preat T, Boynton SC, Del Vecchio M. Genetic dissection of consolidated memory in Drosophila. Cell. 1994;79(1):35–47. pmid:7923375
  11. 11. Tully T, Boynton S, Brandes C, Dura JM, Mihalek R, Preat T, et al. Genetic dissection of memory formation in Drosophila melanogaster. Cold Spring Harbor symposia on quantitative biology. 1990;55:203–11. pmid:2132815
  12. 12. Tomchik SM, Davis RL. Dynamics of learning-related cAMP signaling and stimulus integration in the Drosophila olfactory pathway. Neuron. 2009;64(4):510–21. pmid:19945393
  13. 13. Levin LR, Han PL, Hwang PM, Feinstein PG, Davis RL, Reed RR. The Drosophila learning and memory gene rutabaga encodes a Ca2+/Calmodulin-responsive adenylyl cyclase. Cell. 1992;68(3):479–89. pmid:1739965
  14. 14. Livingstone MS, Sziber PP, Quinn WG. Loss of calcium/calmodulin responsiveness in adenylate cyclase of rutabaga, a Drosophila learning mutant. Cell. 1984;37(1):205–15. pmid:6327051
  15. 15. Byers D, Davis RL, Kiger JA, Jr. Defect in cyclic AMP phosphodiesterase due to the dunce mutation of learning in Drosophila melanogaster. Nature. 1981;289(5793):79–81. pmid:6256649
  16. 16. Chen CN, Denome S, Davis RL. Molecular analysis of cDNA clones and the corresponding genomic coding sequences of the Drosophila dunce+ gene, the structural gene for cAMP phosphodiesterase. Proceedings of the National Academy of Sciences of the United States of America. 1986;83(24):9313–7. PubMed Central PMCID: PMC387128. pmid:3025834
  17. 17. Dudai Y, Jan YN, Byers D, Quinn WG, Benzer S. dunce, a mutant of Drosophila deficient in learning. Proceedings of the National Academy of Sciences of the United States of America. 1976;73(5):1684–8. PubMed Central PMCID: PMC430364. pmid:818641
  18. 18. Isabel G, Pascual A, Preat T. Exclusive consolidated memory phases in Drosophila. Science. 2004;304(5673):1024–7. pmid:15143285
  19. 19. Perazzona B, Isabel G, Preat T, Davis RL. The role of cAMP response element-binding protein in Drosophila long-term memory. J Neurosci. 2004;24(40):8823–8. pmid:15470148
  20. 20. Yin JC, Wallach JS, Del Vecchio M, Wilder EL, Zhou H, Quinn WG, et al. Induction of a dominant negative CREB transgene specifically blocks long-term memory in Drosophila. Cell. 1994;79(1):49–58. pmid:7923376
  21. 21. Quinn WG, Dudai Y. Memory phases in Drosophila. Nature. 1976;262(5569):576–7. pmid:822344
  22. 22. Erber J. Retrograde amnesia in honeybees (Apis mellifera carnica). Journal of comparative and physiological psychology. 1976;90(1):41–6. pmid:987070
  23. 23. Folkers E, Waddell S, Quinn WG. The Drosophila radish gene encodes a protein required for anesthesia-resistant memory. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(46):17496–500. PubMed Central PMCID: PMC1634833. pmid:17088531
  24. 24. Folkers E, Drain P, Quinn WG. Radish, a Drosophila mutant deficient in consolidated memory. Proceedings of the National Academy of Sciences of the United States of America. 1993;90(17):8123–7. PubMed Central PMCID: PMC47300. pmid:8367473
  25. 25. Bouzaiane E, Trannoy S, Scheunemann L, Placais PY, Preat T. Two independent mushroom body output circuits retrieve the six discrete components of Drosophila aversive memory. Cell reports. 2015;11(8):1280–92. pmid:25981036
  26. 26. Hirano Y, Masuda T, Naganos S, Matsuno M, Ueno K, Miyashita T, et al. Fasting launches CRTC to facilitate long-term memory formation in Drosophila. Science. 2013;339(6118):443–6. pmid:23349290
  27. 27. Yamazaki D, Horiuchi J, Nakagami Y, Nagano S, Tamura T, Saitoe M. The Drosophila DCO mutation suppresses age-related memory impairment without affecting lifespan. Nature neuroscience. 2007;10(4):478–84. pmid:17322874
  28. 28. Galili DS, Ludke A, Galizia CG, Szyszka P, Tanimoto H. Olfactory trace conditioning in Drosophila. J Neurosci. 2011;31(20):7240–8. pmid:21593308
  29. 29. Krashes MJ, Waddell S. Rapid consolidation to a radish and protein synthesis-dependent long-term memory after single-session appetitive olfactory conditioning in Drosophila. J Neurosci. 2008;28(12):3103–13. PubMed Central PMCID: PMC2516741. pmid:18354013
  30. 30. Hendel T, Michels B, Neuser K, Schipanski A, Kaun K, Sokolowski MB, et al. The carrot, not the stick: appetitive rather than aversive gustatory stimuli support associative olfactory learning in individually assayed Drosophila larvae. Journal of comparative physiology A, Neuroethology, sensory, neural, and behavioral physiology. 2005;191(3):265–79. pmid:15657743
  31. 31. Rohwedder A, Pfitzenmaier JE, Ramsperger N, Apostolopoulou AA, Widmann A, Thum AS. Nutritional value-dependent and nutritional value-independent effects on Drosophila melanogaster larval behavior. Chem Senses. 2012;37(8):711–21. pmid:22695795
  32. 32. Honjo K, Furukubo-Tokunaga K. Induction of cAMP response element-binding protein-dependent medium-term memory by appetitive gustatory reinforcement in Drosophila larvae. J Neurosci. 2005;25(35):7905–13. pmid:16135747
  33. 33. Neuser K, Husse J, Stock P, Gerber B. Appetitive olfactory learning in larvae: effects of repetition, reward strength, age, gender, assay type and memory span. Animal Behaviour. 2005;69(4):891–8.
  34. 34. Apostolopoulou AA, Mazija L, Wust A, Thum AS. The neuronal and molecular basis of quinine-dependent bitter taste signaling in Drosophila larvae. Frontiers in behavioral neuroscience. 2014;8:6. PubMed Central PMCID: PMC3902218. pmid:24478653
  35. 35. Eschbach C, Cano C, Haberkern H, Schraut K, Guan C, Triphan T, et al. Associative learning between odorants and mechanosensory punishment in larval Drosophila. The Journal of experimental biology. 2011;214(Pt 23):3897–905. pmid:22071180
  36. 36. Khurana S, Abu Baker MB, Siddiqi O. Odour avoidance learning in the larva of Drosophila melanogaster. Journal of biosciences. 2009;34(4):621–31. pmid:19920347
  37. 37. Pauls D, Pfitzenmaier JER, Krebs-Wheaton R, Selcho M, Stocker RF, Thum AS. Electric Shock-Induced Associative Olfactory Learning in Drosophila Larvae. Chem Senses. 2010;35(4):335–46. pmid:20212010
  38. 38. Schleyer M, Saumweber T, Nahrendorf W, Fischer B, von Alpen D, Pauls D, et al. A behavior-based circuit model of how outcome expectations organize learned behavior in larval Drosophila. Learn Mem. 2011;18(10):639–53. pmid:21946956
  39. 39. Gerber B, Hendel T. Outcome expectations drive learned behaviour in larval Drosophila. Proceedings Biological sciences / The Royal Society. 2006;273(1604):2965–8. PubMed Central PMCID: PMC1639518.
  40. 40. Aceves-Pina EO, Quinn WG. Learning in Normal and Mutant Drosophila Larvae. Science. 1979;206:93–6. pmid:17812455
  41. 41. Khurana S, Robinson BG, Wang Z, Shropshire WC, Zhong AC, Garcia LE, et al. Olfactory conditioning in the third instar larvae of Drosophila melanogaster using heat shock reinforcement. Behavior genetics. 2012;42(1):151–61. pmid:21833772
  42. 42. Honjo K, Furukubo-Tokunaga K. Distinctive neuronal networks and biochemical pathways for appetitive and aversive memory in Drosophila larvae. J Neurosci. 2009;29(3):852–62. pmid:19158309
  43. 43. Tully T, Cambiazo V, Kruse L. Memory through metamorphosis in normal and mutant Drosophila. J Neurosci. 1994;14(1):68–74. pmid:8283252
  44. 44. Niewalda T, Singhal N, Fiala A, Saumweber T, Wegener S, Gerber B. Salt processing in larval Drosophila: choice, feeding, and learning shift from appetitive to aversive in a concentration-dependent way. Chem Senses. 2008;33(8):685–92. PubMed Central PMCID: PMC2565773. pmid:18640967
  45. 45. El-Keredy A, Schleyer M, Konig C, Ekim A, Gerber B. Behavioural analyses of quinine processing in choice, feeding and learning of larval Drosophila. Plos One. 2012;7(7):e40525. PubMed Central PMCID: PMC3393658. pmid:22802964
  46. 46. Saumweber T, Husse J, Gerber B. Innate attractiveness and associative learnability of odors can be dissociated in larval Drosophila. Chem Senses. 2011;36(3):223–35. PubMed Central PMCID: PMC3038274. pmid:21227902
  47. 47. Michels B, Chen YC, Saumweber T, Mishra D, Tanimoto H, Schmid B, et al. Cellular site and molecular mode of synapsin action in associative learning. Learn Mem. 2011;18(5):332–44. pmid:21518740
  48. 48. Pauls D, Selcho M, Gendre N, Stocker RF, Thum AS. Drosophila larvae establish appetitive olfactory memories via mushroom body neurons of embryonic origin. J Neurosci. 2010;30(32):10655–66. pmid:20702697
  49. 49. Kitamoto T. Conditional modification of behavior in Drosophila by targeted expression of a temperature-sensitive shibire allele in defined neurons. Journal of neurobiology. 2001;47(2):81–92. pmid:11291099
  50. 50. Akbergenova Y, Bykhovskaia M. Synapsin maintains the reserve vesicle pool and spatial segregation of the recycling pool in Drosophila presynaptic boutons. Brain research. 2007;1178:52–64. pmid:17904536
  51. 51. Akbergenova Y, Bykhovskaia M. Synapsin regulates vesicle organization and activity-dependent recycling at Drosophila motor boutons. Neuroscience. 2010;170(2):441–52. PubMed Central PMCID: PMC4706455. pmid:20638447
  52. 52. Li L, Chin LS, Shupliakov O, Brodin L, Sihra TS, Hvalby O, et al. Impairment of synaptic vesicle clustering and of synaptic transmission, and increased seizure propensity, in synapsin I-deficient mice. Proceedings of the National Academy of Sciences of the United States of America. 1995;92(20):9235–9. PubMed Central PMCID: PMC40959. pmid:7568108
  53. 53. Pieribone VA, Shupliakov O, Brodin L, Hilfiker-Rothenfluh S, Czernik AJ, Greengard P. Distinct pools of synaptic vesicles in neurotransmitter release. Nature. 1995;375(6531):493–7. pmid:7777058
  54. 54. Godenschwege TA, Reisch D, Diegelmann S, Eberle K, Funk N, Heisenberg M, et al. Flies lacking all synapsins are unexpectedly healthy but are impaired in complex behaviour. The European journal of neuroscience. 2004;20(3):611–22. pmid:15255973
  55. 55. Knapek S, Gerber B, Tanimoto H. Synapsin is selectively required for anesthesia-sensitive memory. Learn Mem. 2010;17(2):76–9. pmid:20154352
  56. 56. Fouquet W, Owald D, Wichmann C, Mertel S, Depner H, Dyba M, et al. Maturation of active zone assembly by Drosophila Bruchpilot. The Journal of cell biology. 2009;186(1):129–45. PubMed Central PMCID: PMC2712991. pmid:19596851
  57. 57. Kittel RJ, Wichmann C, Rasse TM, Fouquet W, Schmidt M, Schmid A, et al. Bruchpilot promotes active zone assembly, Ca2+ channel clustering, and vesicle release. Science. 2006;312(5776):1051–4. pmid:16614170
  58. 58. Wagh DA, Rasse TM, Asan E, Hofbauer A, Schwenkert I, Durrbeck H, et al. Bruchpilot, a protein with homology to ELKS/CAST, is required for structural integrity and function of synaptic active zones in Drosophila. Neuron. 2006;49(6):833–44. pmid:16543132
  59. 59. Knapek S, Sigrist S, Tanimoto H. Bruchpilot, a synaptic active zone protein for anesthesia-resistant memory. J Neurosci. 2011;31(9):3453–8. pmid:21368057
  60. 60. Dudai Y, Corfas G, Hazvi S. What is the possible contribution of Ca2+-stimulated adenylate cyclase to acquisition, consolidation and retention of an associative olfactory memory in Drosophila. Journal of comparative physiology A, Sensory, neural, and behavioral physiology. 1988;162(1):101–9. pmid:3127581
  61. 61. Gervasi N, Tchenio P, Preat T. PKA dynamics in a Drosophila learning center: coincidence detection by rutabaga adenylyl cyclase and spatial regulation by dunce phosphodiesterase. Neuron. 2010;65(4):516–29. pmid:20188656
  62. 62. Bruggemann A, Pardo LA, Stuhmer W, Pongs O. Ether-a-go-go encodes a voltage-gated channel permeable to K+ and Ca2+ and modulated by cAMP. Nature. 1993;365(6445):445–8. pmid:7692301
  63. 63. Gordon D, Moskowitz H, Zlotkin E. Sodium channel polypeptides in central nervous systems of various insects identified with site directed antibodies. Biochimica et biophysica acta. 1990;1026(1):80–6. pmid:2165810
  64. 64. Zhou Y, Wang J, Wen H, Kucherovsky O, Levitan IB. Modulation of Drosophila slowpoke calcium-dependent potassium channel activity by bound protein kinase a catalytic subunit. J Neurosci. 2002;22(10):3855–63. pmid:12019304
  65. 65. Yin JC, Del Vecchio M, Zhou H, Tully T. CREB as a memory modulator: induced expression of a dCREB2 activator isoform enhances long-term memory in Drosophila. Cell. 1995;81(1):107–15. pmid:7720066
  66. 66. Horiuchi J, Yamazaki D, Naganos S, Aigaki T, Saitoe M. Protein kinase A inhibits a consolidated form of memory in Drosophila. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(52):20976–81. PubMed Central PMCID: PMC2634933. pmid:19075226
  67. 67. Hannan F, Ho I, Tong JJ, Zhu Y, Nurnberg P, Zhong Y. Effect of neurofibromatosis type I mutations on a novel pathway for adenylyl cyclase activation requiring neurofibromin and Ras. Human molecular genetics. 2006;15(7):1087–98. PubMed Central PMCID: PMC1866217. pmid:16513807
  68. 68. Rahn T, Leippe M, Roeder T, Fedders H. EGFR signaling in the brain is necessary for olfactory learning in Drosophila larvae. Learn Mem. 2013;20(4):194–200. pmid:23512935
  69. 69. Choi KW, Smith RF, Buratowski RM, Quinn WG. Deficient protein kinase C activity in turnip, a Drosophila learning mutant. The Journal of biological chemistry. 1991;266(24):15999–606. pmid:1874743
  70. 70. Drier EA, Tello MK, Cowan M, Wu P, Blace N, Sacktor TC, et al. Memory enhancement and formation by atypical PKM activity in Drosophila melanogaster. Nature neuroscience. 2002;5(4):316–24. pmid:11914720
  71. 71. Selcho M, Pauls D, Han KA, Stocker RF, Thum AS. The role of dopamine in Drosophila larval classical olfactory conditioning. Plos One. 2009;4(6):e5897. PubMed Central PMCID: PMC2690826. pmid:19521527
  72. 72. Waddell S. Reinforcement signalling in Drosophila; dopamine does it all after all. Current opinion in neurobiology. 2013;23(3):324–9. pmid:23391527
  73. 73. Heisenberg M. Mushroom body memoir: from maps to models. Nature reviews Neuroscience. 2003;4(4):266–75. pmid:12671643
  74. 74. Busto GU, Cervantes-Sandoval I, Davis RL. Olfactory learning in Drosophila. Physiology. 2010;25(6):338–46. PubMed Central PMCID: PMC3380424. pmid:21186278
  75. 75. Berry JA, Cervantes-Sandoval I, Nicholas EP, Davis RL. Dopamine is required for learning and forgetting in Drosophila. Neuron. 2012;74(3):530–42. pmid:22578504
  76. 76. Kim YC, Lee HG, Han KA. D1 dopamine receptor dDA1 is required in the mushroom body neurons for aversive and appetitive learning in Drosophila. J Neurosci. 2007;27(29):7640–7. pmid:17634358
  77. 77. Neve KA, Seamans JK, Trantham-Davidson H. Dopamine receptor signaling. J Recept Signal Transduct Res. 2004;24(3):165–205. pmid:15521361
  78. 78. Kume K, Kume S, Park SK, Hirsh J, Jackson FR. Dopamine is a regulator of arousal in the fruit fly. J Neurosci. 2005;25(32):7377–84. pmid:16093388
  79. 79. Ueno T, Kume K. Functional characterization of dopamine transporter in vivo using Drosophila melanogaster behavioral assays. Frontiers in behavioral neuroscience. 2014;8:303. PubMed Central PMCID: PMC4153294. pmid:25232310
  80. 80. Berglund EC, Makos MA, Keighron JD, Phan N, Heien ML, Ewing AG. Oral administration of methylphenidate blocks the effect of cocaine on uptake at the Drosophila dopamine transporter. ACS chemical neuroscience. 2013;4(4):566–74. PubMed Central PMCID: PMC3629741. pmid:23402315
  81. 81. van Swinderen B, Brembs B. Attention-like deficit and hyperactivity in a Drosophila memory mutant. J Neurosci. 2010;30(3):1003–14. pmid:20089909
  82. 82. Bailey CH, Giustetto M, Zhu H, Chen M, Kandel ER. A novel function for serotonin-mediated short-term facilitation in aplysia: conversion of a transient, cell-wide homosynaptic hebbian plasticity into a persistent, protein synthesis-independent synapse-specific enhancement. Proceedings of the National Academy of Sciences of the United States of America. 2000;97(21):11581–6. PubMed Central PMCID: PMC17243. pmid:11027355
  83. 83. Fischer A, Sananbenesi F, Schrick C, Spiess J, Radulovic J. Distinct roles of hippocampal de novo protein synthesis and actin rearrangement in extinction of contextual fear. J Neurosci. 2004;24(8):1962–6. pmid:14985438
  84. 84. Lattal KM, Abel T. Different requirements for protein synthesis in acquisition and extinction of spatial preferences and context-evoked fear. J Neurosci. 2001;21(15):5773–80. pmid:11466449
  85. 85. Muller U. Learning in honeybees: from molecules to behaviour. Zoology (Jena). 2002;105(4):313–20.
  86. 86. Kandel ER. The molecular biology of memory storage: a dialog between genes and synapses. Biosci Rep. 2001;21(5):565–611. pmid:12168768
  87. 87. Dudai Y. The neurobiology of consolidations, or, how stable is the engram? Annual review of psychology. 2004;55:51–86. pmid:14744210
  88. 88. Davis RL. Olfactory memory formation in Drosophila: from molecular to systems neuroscience. Annual review of neuroscience. 2005;28:275–302. pmid:16022597
  89. 89. Nguyen PV, Woo NH. Regulation of hippocampal synaptic plasticity by cyclic AMP-dependent protein kinases. Progress in neurobiology. 2003;71(6):401–37. pmid:15013227
  90. 90. Abel T, Nguyen PV, Barad M, Deuel TA, Kandel ER, Bourtchouladze R. Genetic demonstration of a role for PKA in the late phase of LTP and in hippocampus-based long-term memory. Cell. 1997;88(5):615–26. pmid:9054501
  91. 91. Huang YY, Martin KC, Kandel ER. Both protein kinase A and mitogen-activated protein kinase are required in the amygdala for the macromolecular synthesis-dependent late phase of long-term potentiation. J Neurosci. 2000;20(17):6317–25. pmid:10964936
  92. 92. Selcher JC, Weeber EJ, Varga AW, Sweatt JD, Swank M. Protein kinase signal transduction cascades in mammalian associative conditioning. The Neuroscientist: a review journal bringing neurobiology, neurology and psychiatry. 2002;8(2):122–31.
  93. 93. Technau G, Heisenberg M. Neural reorganization during metamorphosis of the corpora pedunculata in Drosophila melanogaster. Nature. 1982;295(5848):405–7. pmid:6799834
  94. 94. Lee T, Lee A, Luo L. Development of the Drosophila mushroom bodies: sequential generation of three distinct types of neurons from a neuroblast. Development. 1999;126(18):4065–76. pmid:10457015
  95. 95. Manseau F, Sossin WS, Castellucci VF. Long-term changes in excitability induced by protein kinase C activation in Aplysia sensory neurons. Journal of neurophysiology. 1998;79(3):1210–8. pmid:9497402
  96. 96. Sacktor TC, Kruger KE, Schwartz JH. Activation of protein kinase C by serotonin: biochemical evidence that it participates in the mechanisms underlying facilitation in Aplysia. J Physiol (Paris). 1988;83(3):224–31.
  97. 97. Sossin WS, Sacktor TC, Schwartz JH. Persistent activation of protein kinase C during the development of long-term facilitation in Aplysia. Learn Mem. 1994;1(3):189–202. pmid:10467596
  98. 98. Sugita S, Goldsmith JR, Baxter DA, Byrne JH. Involvement of protein kinase C in serotonin-induced spike broadening and synaptic facilitation in sensorimotor connections of Aplysia. Journal of neurophysiology. 1992;68(2):643–51. pmid:1527580
  99. 99. Atkins CM, Selcher JC, Petraitis JJ, Trzaskos JM, Sweatt JD. The MAPK cascade is required for mammalian associative learning. Nature neuroscience. 1998;1(7):602–9. pmid:10196568
  100. 100. MacDonald JF, Kotecha SA, Lu WY, Jackson MF. Convergence of PKC-dependent kinase signal cascades on NMDA receptors. Curr Drug Targets. 2001;2(3):299–312. pmid:11554554
  101. 101. Sacktor TC. PKMzeta, LTP maintenance, and the dynamic molecular biology of memory storage. Progress in brain research. 2008;169:27–40. pmid:18394466
  102. 102. Grunbaum L, Muller U. Induction of a specific olfactory memory leads to a long-lasting activation of protein kinase C in the antennal lobe of the honeybee. J Neurosci. 1998;18(11):4384–92. pmid:9592115
  103. 103. Zhang J, Little CJ, Tremmel DM, Yin JC, Wesley CS. Notch-inducible hyperphosphorylated CREB and its ultradian oscillation in long-term memory formation. J Neurosci. 2013;33(31):12825–34. PubMed Central PMCID: PMC3728690. pmid:23904617
  104. 104. Scheunemann L, Jost E, Richlitzki A, Day JP, Sebastian S, Thum AS, et al. Consolidated and labile odor memory are separately encoded within the Drosophila brain. J Neurosci. 2012;32(48):17163–71. pmid:23197709
  105. 105. Aso Y, Sitaraman D, Ichinose T, Kaun KR, Vogt K, Belliart-Guerin G, et al. Mushroom body output neurons encode valence and guide memory-based action selection in Drosophila. eLife. 2014;3.
  106. 106. Barnstedt O, Owald D, Felsenberg J, Brain R, Moszynski JP, Talbot CB, et al. Memory-Relevant Mushroom Body Output Synapses Are Cholinergic. Neuron. 2016;89(6):1237–47. PubMed Central PMCID: PMC4819445. pmid:26948892
  107. 107. Reale V, Hannan F, Hall LM, Evans PD. Agonist-specific coupling of a cloned Drosophila melanogaster D1-like dopamine receptor to multiple second messenger pathways by synthetic agonists. J Neurosci. 1997;17(17):6545–53. pmid:9254667
  108. 108. Beggs KT, Tyndall JD, Mercer AR. Honey bee dopamine and octopamine receptors linked to intracellular calcium signaling have a close phylogenetic and pharmacological relationship. Plos One. 2011;6(11):e26809. PubMed Central PMCID: PMC3214027. pmid:22096499
  109. 109. Scherer S, Stocker RF, Gerber B. Olfactory learning in individually assayed Drosophila larvae. Learn Mem. 2003;10(3):217–25. PubMed Central PMCID: PMC202312. pmid:12773586
  110. 110. Rohwedder A, Wenz NL, Stehle B, Huser A, Yamagata N, Zlatic M, et al. Four Individually Identified Paired Dopamine Neurons Signal Reward in Larval Drosophila. Current biology: CB. 2016.
  111. 111. Michels B, Diegelmann S, Tanimoto H, Schwenkert I, Buchner E, Gerber B. A role for Synapsin in associative learning: the Drosophila larva as a study case. Learn Mem. 2005;12(3):224–31. PubMed Central PMCID: PMC1142449. pmid:15930500
  112. 112. Kleber J, Chen YC, Michels B, Saumweber T, Schleyer M, Kahne T, et al. Synapsin is required to "boost" memory strength for highly salient events. Learn Mem. 2016;23(1):9–20. PubMed Central PMCID: PMC4749839. pmid:26670182
  113. 113. Connolly JB, Roberts IJ, Armstrong JD, Kaiser K, Forte M, Tully T, et al. Associative learning disrupted by impaired Gs signaling in Drosophila mushroom bodies. Science. 1996;274(5295):2104–7. pmid:8953046
  114. 114. Broughton SJ, Kane NS, Arthur B, Yoder M, Greenspan RJ, Robichon A. Endogenously inhibited protein kinase C in transgenic Drosophila embryonic neuroblasts down regulates the outgrowth of type I and II processes of cultured mature neurons. Journal of cellular biochemistry. 1996;60(4):584–99. pmid:8707897
  115. 115. Yu D, Akalal DB, Davis RL. Drosophila alpha/beta mushroom body neurons form a branch-specific, long-term cellular memory trace after spaced olfactory conditioning. Neuron. 2006;52(5):845–55. PubMed Central PMCID: PMC1779901. pmid:17145505
  116. 116. Apostolopoulou AA, Widmann A, Rohwedder A, Pfitzenmaier JE, Thum AS. Appetitive associative olfactory learning in Drosophila larvae. Journal of visualized experiments: JoVE. 2013;(72).
  117. 117. Selcho M, Pauls D, Huser A, Stocker RF, Thum AS. Characterization of the octopaminergic and tyraminergic neurons in the central brain of Drosophila larvae. J Comp Neurol. 2014.
  118. 118. Thum AS, Leisibach B, Gendre N, Selcho M, Stocker RF. Diversity, variability, and suboesophageal connectivity of antennal lobe neurons in D. melanogaster larvae. J Comp Neurol. 2011;519(17):3415–32. pmid:21800296