Drosophila as a new model organism for the
neurobiology of aggression?

Summary: We report here the effects of several neurobiological determinants on aggressive behaviour in the fruitfly Drosophila melanogaster. This study combines behavioural, transgenic, genetic and pharmacological techniques that are well established in the fruitfly, in the novel context of the neurobiology of aggression. We find that octopamine, dopamine and a region in the Drosophila brain called the mushroom bodies, have profound influence on the expression of aggressive behaviour. Serotonin had no effect. We conclude that Drosophila, with its advanced set of molecular tools and its behavioural richness, has the potential to develop into a new model organism for the study of the neurobiology of aggression.

Introduction

Drosophila is the “jack of all trades” in biology, but has not been studied in the context of the neurobiology of aggression. The fruitfly exhibits aggressive behaviour (Jacobs, 1960) and this behaviour is ethologically well characterized (Dow and von Schilcher, 1975; Jacobs, 1978; Lee and Hall, 2000; Skrzipek et al., 1979). The evolutionary relevance of this aggressive behaviour is also well established (Boake and Hoikkala, 1995; Boake and Konigsberg, 1998; Boake et al., 1998; Dow and von Schilcher, 1975; Hoffmann, 1988, 1989, 1994; Hoffmann and Cacoyianni, 1989; Ringo et al., 1983; Skrzipek et al., 1979; Zamudio et al., 1995). Finally, the ecological circumstances under which Drosophila exhibits territoriality and aggression have been examined in great detail (Hoffmann, 1987, 1988, 1989, 1994; Hoffmann and Cacoyianni, 1989, 1990). Under appropriate conditions, male flies try to occupy a food patch and defend it against other males, even in the laboratory. However, this aggressive behaviour in Drosophila has escaped most neurobiologists’ notice. In this article, we combine ethological, ecological and evolutionary knowledge with molecular, genetic and pharmacological tools to manipulate the aggressive behaviour of Drosophila melanogaster.

To our knowledge, only two genetic factors have been reported to affect aggressive behaviour in Drosophila: The sex-determination hierarchy (SDH) and the beta-alanine pathway. fruitless (fru) and dissatisfaction (dsf) mutants have been described as more aggressive than wildtype controls (Lee and Hall, 2000). Both genes are part of the SDH. Flies carrying mutant alleles of the black (b) gene appear less aggressive, whereas ebony (e) mutants appear more aggressive (Jacobs, 1978). The enzymes encoded by the two genes regulate beta-alanine levels (b flies have reduced, e flies elevated levels).


3D reconstruction of the Drosophila brain. The green structure depicts the mushroom bodies. (Click for a larger image)

It is straightforward to expect genes of the SDH to affect sex-specific behaviours, but the pathways through which they modulate that behaviour are largely unknown. One possibility could be via the regulation of small neuroactive molecules (such as beta-alanine and the biogenic amines) and their receptors. Biogenic amines play a key role in the regulation of aggressive behaviour not only in vertebrates, but also in arthropods (e.g. Edwards and Kravitz, 1997; Heinrich et al., 1999, 2000; Huber et al., 1997a, b; Kravitz, 2000; Schneider et al., 1996; Stevenson et al., 2000). The biogenic amine system in flies is well described (see Monastirioti, 1999). Most serotonin and dopamine mutants in Drosophila are either lethal or affect both serotonin and dopamine, due to their shared synthesis pathways (e.g. Johnson and Hirsh, 1990; Lundell and Hirsh, 1994; Shen et al., 1993; Shen and Hirsh, 1994). However, there are established protocols that are commonly used to manipulate the levels of these amines individually in the adult fly (Neckameyer, 1998; Vaysse et al., 1988). Octopamine null mutants have been generated and characterized (Monastirioti et al., 1996). Interestingly, certain octopamine and dopamine receptors are preferentially expressed in a prominent neuropil in the Drosophila brain called the mushroom bodies (Han et al., 1996, 1998). Thus, all of the prerequisites for a systematic analysis of the neurobiological factors involved in the expression of aggressive behaviour are available: 1) a considerable body of knowledge about the behaviour and its ecological context, 2) circumstantial evidence about possible neurobiological factors involved in regulating the behaviour, and 3) methods for manipulating these factors and for quantifying the behaviour.

In a first attempt to characterize the effects of various candidates for neurobiological factors regulating aggression, we report here the results of a competition experiment. Six male flies competed for a food patch and three mated females. The experimental males have been manipulated either by a classical mutation affecting beta-alanine levels, a P-element mutation affecting octopamine levels, insertion of transgenes affecting synaptic output from the mushroom bodies or by pharmacological treatment affecting serotonin or dopamine levels and then tested for their aggressive behaviour.

Materials and Methods

Flies: Animals were kept on standard cornmeal/molasses medium (see Guo et al., 1996 for recipe) at 25° C and 60% humidity with a 16hr light/8hr dark regime, except where noted. The females in all experiments were mated wildtype Canton S flies.

Mutants: Black1 and ebony1 mutant strains from the laboratory’s 18° C stock collection (provided by S. Benzer in 1970) were kept at 25° C for at least two generations. The M18 P-element octopamine mutant and control stocks (Monastirioti et al., 1996) were kept at 25° C for two generations after arrival.


Schematic representation of Drosophila neuropils (upper panel) and actual spatial expression pattern of the P[GAL4] line mb247 (lower panel). Dorsal is up; meb - median bundle; eb - ellipsoid body; ped - peduncles; ca - calyces; e - esophagus; an - antennal nerve; Kcb - Kenyon cell bodies. Scale bar, 50 µm. (click for a larger image)

Transgenes: Sweeney et al. (1995) developed a method that constitutively blocks synaptic transmission by expressing the catalytic subunit of bacterial tetanus toxin (Cnt-E) in target neurons in the Drosophila brain using the P[GAL4] technique (Brand and Perrimon, 1993). Inspired by the preferential expression of certain dopamine and octopamine receptors in the mushroom bodies (Han et al., 1996, 1998), we used the Cnt-E transgene to block synaptic output from the mushroom bodies (Sweeney et al., 1995). Expression of another transgene, an inactive form of the tetanus toxin light chain (imp-tntQ), controlled for deleterious effects of protein overexpression (Sweeney et al., 1995). The P[GAL4] line mb247 (Schulz et al., 1996) served as a mushroom body specific GAL4 driver (Zars et al., 2000) for both toxins. The trans-heterozygote offspring from the GAL4 driver strain and the two UASGAL4 reporter strains (Cnt-E and imp-tntQ) entered the study.


Immunohistochemical verification of our serotonin treatment. Brain sections from each row were treated identically on the same slide. Brain sections in the left columns are from pCPA treated animals and show a weaker staining than the sections from the 5HTP treated animals in the right column. (click for a larger image)

Pharmacological treatments: Drosophila from the wildtype strain Berlin (wtb) were treated as described by Neckameyer (1998) and Vaysse et al. (1988). Briefly, the animals were fed a sucrose solution containing either 10mg/ml of the serotonin precursor 5HTP (5-hydroxy-tryptophan), or 10mg/ml of the serotonin synthesis inhibitor pCPA (para-chlorophenylalanine) to manipulate serotonin levels. Effectiveness of the treatment was verified qualitatively with standard immunohistochemical techniques using rabbit serotonin antisera (see inset; Buchner et al., 1986, 1988). Alternatively, the animals were treated with 1mg/ml of the dopamine precursor L-DOPA (L-3,4-dihydroxyphenylalanine) or 10mg/ml of the dopamine synthesis inhibitor 3IY (3-iodo-tyrosine) to manipulate dopamine levels. Effectiveness of the treatment was verified by observation of cuticle tanning. A dose of 10mg/ml L-DOPA was lethal, confirming unpublished data from Wendy Neckameyer (Saint Louis University School of Medicine).

Experimental groups: Using the different stocks described above, we arranged six different groups of ‘low’ vs. ‘high’ males, such that the respective amine or the amount of synaptic output from the mushroom bodies was manipulated to produce relative high and low level subgroups.

  1. Wildtype Berlin (wtb)
    Wildtype Berlin flies are randomly assigned to a 'high' or a 'low' group. No difference between the subgroups is expected (negative control).
  2. Serotonin (5ht)
    a) Wildtype Berlin with 10mg/ml 5HTP in sucrose solution. This treatment produces high levels of serotonin (5ht+).
    b) Wildtype Berlin with 10mg/ml pCPA in sucrose solution. This treatment produces low levels of serotonin (5ht-).
  3. Octopamine (oa)
    a) M18 P-element parental stock, from which the jump-out below was generated (red eyed). This strain has normal levels of octopamine (Monastirioti et al., 1996) and will be denoted the ‘high’ subgroup (oa+).
    b) M18 jump-out mutants. As tyramine-beta-hydroxylase (octopamine producing enzyme) null mutants (white eyed), these flies have no detectable octopamine (Monastirioti et al., 1996) and will be denoted the ‘low’ subgroup (oa-).
  4. Dopamine (da)
    a) Wildtype Berlin with 1mg/ml L-DOPA in sucrose solution. This treatment produces high levels of dopamine (da+).
    b) Wildtype Berlin with 10mg/ml 3-iodo-tyrosine in sucrose solution. This treatment produces low levels of dopamine (da-).
  5. beta-alanine (b/e)
    a) ebony mutants with high beta-alanine levels (e).
    b) black mutants with low beta-alanine levels (b).
    This group serves as the positive control, as it is known that e flies are more aggressive than b flies (Jacobs, 1978).
  6. Mushroom bodies (mb)
    a) offspring of P[GAL4] line mb247 with the UAS-IMP-tntQ line. This strain has normal levels of synaptic output from the mushroom bodies and will be referred to as the ‘high’ subgroup (mb+).
    b) offspring of P[GAL4] line mb247 with the UAS-Cnt-E line. This strain has no synaptic output from the mushroom bodies and will be called the ‘low’ subgroup (mb-).

Thus, we arranged four experimental groups and two control groups. For each group, the two subgroups ('high' and 'low') compete against each other in one recording chamber. Each group was tested twice with different animals.

Recording chambers: Aggression was studied in cylindrical cages similar to those used by Hoffmann (1987): 100mm Petri-Dishes, top and bottom separated by a 40mm high spacer (i.e. a cylindrical chamber of 100mm diameter and 40mm height). The bottom of the chamber was filled with 2% Agar to moisturize the chamber. Flies were introduced by gentle aspiration through a small hole in the spacer. A food patch (10mm diameter, 12mm high) was positioned in the centre of the chamber, containing a mixture of minced 2% agar, apple juice, syrup and a live yeast suspension (after Reif, 1998), filled to level with the rim of the containing vial. The chamber was placed in a Styrofoam box (used to ship biochemical reagents on dry-ice; outer measures: 275x275mm, height: 250mm, inner measures: 215x215mm, height: 125mm) to standardize lighting conditions and to shield the chambers from movements by the experimenters. Two Styrofoam boxes with one chamber each were arranged underneath video cameras, focused on the food patch in a darkened room at 25°C. Ring-shaped neon-lights (Osram L32W21C, power supply Philips BRC406) on top of the boxes provided homogenous illumination throughout the experiment.

Table1: Experimental time-course.

 

Day1

Day2

Day3

Day4

Day5

Day6

Day7

Day8

Day9

Day10

Day11

Put in vials

5ht

wtb

oa

da

wtb

b/e

mb

5ht

da

oa

b/e

mb

         

Mark

       

5ht

wtb

oa

da

wtb

b/e

mb

5ht

da

oa

b/e

mb

 

Record

         

5ht

wtb

oa

da

wtb

b/e

mb

5ht

da

oa

b/e

mb

Table 1: Experimental time-course. Two groups were treated in separate vials but in parallel each experimental day. Each group was treated in two replicates, starting with different flies on different days. For abbreviations see Materials and Methods.

Experimental time course: The stocks were treated completely in parallel (see Table 1). A 5% sucrose solution (in Drosophila ringer) with or without added treatment was pipetted onto 5 pieces of filter paper snugly fitting in cylindrical (12x40mm) vials before transferring newly ecclosed (0-24 hours) male flies into the vials. The flies were transferred into new vials with new solution and new filter paper on a daily basis for 5 days. Each group was treated in two replicates, starting with new flies on different days (see Table 1). On the 5th day, 4-6 flies per subgroup were briefly immobilised on a cold plate and marked on the thorax with either green or white acrylic paint (one small dot). At 08.00hrs (1 hour after lights-on) on the 6th day, the animals of the two groups treated in parallel were transferred into the recording chambers (3 mated but otherwise untreated Canton S females and 6 males; 3 males of each paired subgroup) and placed underneath the video cameras with all conditions identical to those during the recording time, except the VCRs were turned off. Continuing the parallel treatment of two groups per day, two video set-ups were used simultaneously (‘left’ and ‘right’). After an acclimatisation period of 2 hours, the VCRs were set to record. For each group, we recorded four hours of fly behaviour, once in each location (yielding the two replicates for each group), resulting in 12 video tapes (see Table 2). Data from both replicates were pooled. Since each group was measured twice with 6 (3+3) experimental animals (males) for each recording, the total number of observed males was 6 animals x 2 replicates x 6 groups = 72. Recording of the experiments was randomised across days.


Table 2: Colour codes and recording dates.

 

No.

Left

No.

Right

Day6

#1

5ht+:green / 5ht-: white

#2

wtb

Day7

#3

oa+: green / oa-: white

#4

da+:green / da-:white

Day8

#5

wtb

#6

e: green / b: white

Day9

#7

mb-: green / mb+: white

#8

5ht+: green / 5ht-: white

Day10

#9

da+: green / da-: white

#10

oa+: green / oa-: white

Day11

#11

e: green / b: white

#12

mb-: green / mb+: white

Table 2: Colour Codes and recording dates. Each group was measured twice, once under each camera with different flies. Each of the 12 experiments was saved on individually numbered, 4h video tapes. This table was used to break the code after the behavioural scoring had been done blindly. For abbreviations see Materials and Methods.

Behavioural scoring: Only male-male interactions were counted. Mated females lose their receptivity to male advances and the males cease courting quickly, refraining from courting for a number of hours (courtship conditioning; e.g. Greenspan and Ferveur, 2000). Little courtship behaviour was thus observed after the acclimatisation period.
Behavioural scoring was done blindly before the colour codes on the flies' thoraces were decoded into ‘high’ and ‘low’. An interaction between two males was classified as either aggressive or non-aggressive similarly to Hoffmann (1987). Briefly, we classified encounters that contained the previously described boxing, head-butting, lunging, wrestling, tussling, charging and chasing behaviours (Dow and von Schilcher, 1975; Hoffmann, 1987, 1988, 1989, 1994; Hoffmann and Cacoyianni, 1989, 1990; Jacobs, 1978; Skrzipek et al., 1979) as aggressive. Encounters that only contained approach, leg contact, wing vibration or wing flapping were classified as non-aggressive. If the encounter was classified as aggressive, it was straightforward to discern the aggressor as one animal attacking and/or chasing the other. Non-aggressive encounters can usually not be classified directionally. Thus, with 3 ‘high‘ and 3 ‘low‘ animals in the recording chamber, any interaction between them falls into 7 categories:

1: high attacks high              aggressive encounter (1aggr)
2: high attacks low              aggressive encounter (2aggr)
3: high/high                         non-aggressive encounter (3nonaggr)
4: high/low                          non-aggressive encounter (4nonaggr)
5: low/low                          non-aggressive encounter (5nonaggr)
6: low attacks high              aggressive encounter (6aggr)
7: low attacks low               aggressive encounter (7aggr)

This design thus yielded 7 values, one value for each of the respective interaction categories, giving each of the 6 groups a characteristic aggression-profile (Fig. 1A).

Data analysis: A log-linear analysis (delta=0.005, criterion for convergence=0.0005, max. iterations 500) was performed over the 6x7 table of observed behavioural frequencies to determine the effect of the treatments on the distribution of behavioural classes. To normalize for the total number of encounters, two derived parameters were computed from the raw data. The first is the likelihood that an individual of one subgroup will attack during an encounter (attack probability, P[A]). It is calculated as the fraction of all encounters in that group involving a 'high' (or 'low', respectively) animal, where such an animal was the aggressor:

, i.e.:

 and

 .

Thus, P[A] describes the probability that a given individual will act aggressively against any other individual it encounters. The second derived parameter assesses the representation of each subgroup in the total number of encounters (encounter probability, P[E]). It is calculated analogously to the first parameter as the fraction of all encounters in a group, where an animal of a specific subgroup (i.e. ‘high’ or ‘low’) participated:

, i.e.:

 and

.

Thus, P[E] describes the probability that an individual of one subgroup will be a participant in an encounter.

While the P[A] can be said to describe the level of aggression of a certain subgroup, the P[E] can be perceived as a control measure for the overall number of interactions in that subgroup, as influenced by e.g. general activity, visual acuity etc. After the data transformation, the resulting probabilities were tested against random distribution using c2 tests.

Results

We performed two 4 hour experiments with 4 experimental and 2 control groups in each experiment. In all, 48 hours of video tape were analysed containing 9881 encounters (an average of 3.4 enc./min or 137.2 enc./male). The two 4 hour experiments were pooled for each group, yielding one 7-score aggression profile for each group (Fig. 1A). A log-linear analysis over the 6 groups and the 7 behavioural classes yields a p<0.0001 (Pearson c²=6479.426, d.f.=30), suggesting the various treatments were effective in changing the proportions of the different classes of encounters in each group.

Fig. 1: Raw and derived data from all six groups. A: Raw behavioural scores. Two different graphs are depicting the same data in order to facilitate the interpretation of the complex data structure obtained from our experiments. A1: multiple bars graph. A2: single bar graph. See Methods for details on behavioural classification. B: Derived probabilities. B1: The probability of attacking. For each subgroup (high, low) the fraction of encounters where a member of that subgroup was the aggressor is calculated from the total number of subgroup encounters. B2: The probability of an encounter. For each subgroup (high, low) the fraction of encounters (irrespective of classification) in which a member of that subgroup participated is calculated from the total number of encounters.
1aggr: high attacks high aggressive encounter;  2aggr: high attacks low aggressive encounter; 3nonaggr: high/high non-aggressive encounter; 4nonaggr: high/low non-aggressive encounter; 5nonaggr: low/low non-aggressive encounter; 6aggr: low attacks high aggressive encounter; 7aggr: low attacks low aggressive encounter.

The raw data (Fig. 1A), reveal that the 2 control groups behaved according to our expectations. The wtb negative control shows a uniform distribution of aggressive encounters, whereas the beta-alanine positive control is skewed towards the mutants with high levels of beta-alanine (Fig. 1A1).

The clearest effects among experimental groups were obtained from the octopamine mutants and the mb group. Both octopamine null mutants (oa-) and animals with inhibited mushroom bodies (mb-) are virtually non-aggressive (Fig. 1A). In the representation in Fig 1A2, the octopamine group seems similar to the wildtype control except for the missing values for 6aggr and 7aggr. However, while the oa+ animals appear to show a wildtype level of aggression, the mb+ animals show elevated levels of aggression compared to all other groups (Fig. 1A).

It also appears that our serotonin treatment had little effect on aggression (Fig. 1A).

The dopamine treatment appears to be somewhat effective in decreasing the number of aggressive encounters in animals with high levels of dopamine, while the animals with low levels of dopamine seem to have similar, if not slightly higher numbers of aggressive encounters than the wildtype controls. Obviously, the number of non-aggressive encounters in the dopamine treated animals is strongly elevated (Fig. 1A). Interestingly, the two subgroups show inverted profiles for intra- and inter-subgroups aggression (i.e. 1aggr/2aggr and 6aggr/7aggr).
The total number of encounters also varies considerably between the different treatments (Fig. 1A2).

Table 3: Chi ²-Statistics for derived probabilities

 

Prob. of Attack P[A]

Prob. of Encounter P[E]

 

Chi² (Yates; df: 1)

p-Value (Yates)

Chi² (Yates; df: 1)

p-Value (Yates)

WT Berlin

0.01

p=0.92

1.85

p=0.17

Serotonin

0.31

p=0.58

13.11

p<0.0003

Octopamine

92.33

p<0.0001

403.71

p<0.0001

Dopamine

36.62

p<0.0001

1109.17

p<0.0001

beta-alanine

2080.64

p<0.0001

177.13

p<0.0001

Mushroom-bodies

3061.61

p<0.0001

315.84

p<0.0001

Table 3: Chi²-Statistics for derived probabilities P[A] and P[E].

With significant effects of our treatments on the distribution of the behaviours within each group, we can process the data in order to determine the effect of our treatments on the propensity of the animals to become aggressive. The fraction of all encounters involving a ‘subgroup’ animal, where such an animal was the aggressor is calculated (Fig. 1B1; P[A] see Materials and Methods). The P[A] allows us to estimate the effects of the treatments on aggression. On P[A] values, Chi² tests can be computed to test the null hypothesis that our treatments had no effect on the probability of being aggressive. Table 3 summarizes the Chi² results for all six groups. The statistics confirm the effects already visible in the raw data (Fig 1A): the two control groups (wtb and b/e) were consistent with our expectations. The obvious effect of octopamine null mutants being completely non-aggressive is corroborated by our statistical analysis, as are the extreme effects of expressing active and inactive, respectively, tetanus toxin in the flies' mushroom bodies (Fig. 1B1). The serotonin treatment had no significant effect on the probability of the flies becoming aggressive during an encounter, despite the fact that we could verify the effectiveness of the treatment immunohistochemically (see inset in Material and Methods). The group in which the dopamine levels were manipulated shows a moderate, but statistically reliable, effect of high dopamine levels leading to a higher probability to attack in an encounter.

Despite most of our treatments having a record of influencing aggression in other animals, the possibility exists that the different treatments may have altered the number of aggressive encounters indirectly by altering the total number of encounters via other factors such as general activity or visual acuity, etc. A likely candidate variable where such effects could be detected is the distribution of encounters over the subgroups, P[E]. For instance, if the treatment rendered the animals of one subgroup inactive, the P[E] of this subgroup should be smaller than that of the other subgroup. If the obtained aggression scores were but a reflection of asymmetric P[E]s, they should follow the pattern of P[E] asymmetry. Fig. 1B2 depicts the distribution of encounters over the two subgroups, independently of  encounter classification. Again, c² statistics are performed and summarized in Table 2. All treatments led to a significant asymmetry in P[E] between subgroups, with the exception of the negative wtb controls. However, the pattern of asymmetry does not seem to match the pattern of asymmetry in the level of aggression (see Discussion).

Discussion

Most importantly for this first study of the effects of various treatments on aggression in Drosophila, the animals in the control groups behaved exactly as expected: no differences were detected among the subgroups of the wtb negative control and previously published higher aggression levels in the ebony (high beta-alanine) than in the black (low beta-alanine) flies (Jacobs, 1978) could be reproduced. These findings corroborate our pilot studies in which we repeatedly observed the same pattern (unpubl. data).

Octopamine null mutants exhibit strongly reduced aggression, as do flies with low levels of synaptic output from their mushroom bodies. Interestingly, certain types of octopamine and dopamine receptors are preferentially expressed in the mushroom bodies of wildtype flies (Han et al., 1996, 1998). It is tempting to interpret this phenocopy of the octopamine mutants as resulting from Kenyon cells being the major regulators of octopamine- (and/or dopamine-) mediated aggression. Recently, temperature sensitive shibirets1 constructs have been developed to conditionally block synaptic transmission (e.g. Dubnau et al., 2001; Kitamoto, 2001; McGuire et al., 2001; Waddell et al., 2000). Unfortunately, at the time of our experiments, the shibirets1 constructs were not yet available. Future experiments definitely should include shibirets1 constructs to replicate our mb- results, examine the high levels of aggression in the mb+ flies and look for other brain areas involved in aggression. Replication of our results using the shibirets1 constructs would also eliminate the possible explanation that the expression of tetanus-toxin anywhere in the fly’s brain abolishes aggressive behaviour and solve the problem of UAS promoter leakiness. The octopamine result is conspicuous in another respect: it is consistent with studies in crickets, where depletion of octopamine and dopamine decreases aggressiveness (Stevenson et al., 2000), but it contrasts with studies in crustaceans where high octopamine levels tend to bias behaviour towards submissiveness (Antonsen and Paul, 1997; Heinrich et al., 2000; Huber et al., 1997a).

The high aggression observed in the mb+ animals is difficult to interpret. In principle, the inactive toxin should not have any effect on the secretion of neurotransmitter at the synapse. More likely is an insertion effect of the P-element containing the imp-tntQ transgene. In that case it would be extremely interesting to characterize the genetic environment within which the P-element lies in order to find the gene responsible for such aggressiveness. One may argue that high aggressiveness by flies of one subgroup may produce low aggression in the respective other subgroup. In the case of the mb group, this is unlikely, because there still should be at least some aggression between mb- animals, even if mb+ animals attacked every other male they encountered. Moreover, mb- animals seemed unaffected by the repeated attacks from mb+ males and kept coming back to the patch soon after an mb+ male chased it off the patch (this is the reason for the high 2aggr value in Fig. 1). However, mb- animals were never observed to be the aggressor. It thus seems more likely that the high frequency of attacks by mb+ males is due to a combination of high levels of aggression due to insertion effects of the imp-tntQ transgene and returning mb- males repeatedly eliciting aggressive behaviours in the mb+ males.

Our serotonin treatment has no significant effect on aggression, despite the fact that we could verify the effectiveness of the treatment immunohistochemically (see inset in Material and Methods). Also, Vaysse et al. (1988) observed effects on learning and memory after identical treatment, indicating that this pharmacological manipulation of serotonin levels in principle can have behavioural effects. Moreover, we observed a noticeable increase in activity in the 5ht- flies, a subjective impression that is corroborated by the significantly increased P[E] of this subgroup (Fig. 1B2). Nevertheless, the possibility remains that the observed difference in serotonin immunoreactivity was not high enough to generate significant differences in aggression though high enough to affect other behaviours. The lack of serotonergic effect on aggression was also repeatedly observed in our pilot studies (unpubl. data). Lee and Hall (2001) have reported that the pattern of serotonergic cells in the Drosophila brain is unaltered in the more aggressive fru mutants, confirming the idea that serotonin is not crucial for regulating aggressive behaviour in the fly. The serotonin results presented here are also consistent with data in crickets, where serotonin depletion appears to have no effect (Stevenson et al., 2000); they contrast with data in crustaceans, where injections of serotonin increase the level of aggressive behaviour (Edwards and Kravitz, 1997; Huber et al., 1997a, b; Kravitz, 2000). Our serotonin data thus parallel our octopamine data in conforming with insect data but contrasting with observations in crustaceans. Perhaps aminergic control of aggression functions fundamentally differently in those two arthropod groups?

Our dopamine treatment had complex effects. The absolute number of non-aggressive encounters appears elevated compared to the wildtype controls (Fig. 1A), reducing overall aggression probabilities (Fig. 1B1; P[A]). Also, while the raw data indicate higher aggression scores in the animals with low dopamine (Fig. 1A1), the P[A] is higher in animals with high dopamine levels (Fig. 1B1). Taking the number of encounters that each subgroup experiences (Fig. 1B2, P[E]) into account, it seems as if the higher raw scores for the ‘low’ dopamine animals is generated by the higher P[E] in this subgroup. Once that factor is accounted for (Fig. 1B1), the perceived difference between raw and derived data disappears.

A general point of concern is side effects of our treatments. Both e and b flies exhibit varying degrees of visual impairment (unpubl. data and Heisenberg, 1971, 1972; Hovemann et al., 1998; Jacobs, 1978), with e showing more severe defects than b flies (unpubl. data and Jacobs, 1978). Without screening pigments (i.e. white-), the M18 octopamine jump-out mutants are expected to have severely impaired vision compared with the control strain still carrying the P-element. Also, the extent to which the treatments may affect general activity is largely unknown (but see Martin et al., 1998). One may assume that a subgroup’s P[E] should reflect overall activity. Not surprisingly, the more visually impaired e and oa- flies have lower P[E]s than the b and oa+ subgroups, respectively (Fig. 1B2). However, the probability to attack seems entirely unaffected by this measure of general activity, as the relations are reversed. Moreover, both the dopamine and the mushroom body groups show a higher probability to attack in the respective ‘high’ subgroup (Fig. 1B1), but their P[E]s are inverted with respect to their P[A]s (Fig. 1B2). Thus, while both vision and general activity may influence aggression, those factors seem to have only marginal effects compared to the determinants studied here.

Of course, this study is only a beginning. We did not address encounter duration, behavioural composition or opponent identity/recognition, let alone potential mechanisms as to how the identified factors might exert their effects. However, one can conclude that our method reproduced published data (the e/b group) and yielded new insights into the neurobiological determinants of aggression in Drosophila melanogaster. Serotonin appears to have no effect, while dopamine, octopamine and the mushroom bodies could be linked to the promotion of aggressive behaviour. We hope to inspire others to exploit Drosophila’s numerous technical advantages for the neurobiology of aggression.

Acknowledgements: We are very grateful to Martin Heisenberg for providing labspace, flies, equipment, intellectual support and comments on the manuscript, to Dieter Dudacek for immunohistochemistry, to Troy Zars for providing the P[GAL4] driver and UAS toxin lines, to Maria Monastirioti for providing the M18 stocks, to Wendy Neckameyer for most helpful discussions on the use of dopamine pharmacology, to Jay Hirsh for advice on serotonin and dopamine mutants, to Marcus Reif for helpful support on food-patch medium, to Reinhard Wolf for supplying ‘Apfelmost’ for the food-patch medium, to David Pettigrew, Randall Hayes, Gregg Phares, Gabi Putz, Robin Hiesinger, Sean McGuire and Douglas Armstrong for discussion and comments on the manuscript and to Hans Kaderaschabek and his team for designing and producing our special equipment. Our special gratitude goes out to Edward Kravitz. This study would probably never have been initiated if it wasn’t for his curiosity and enthusiastic support. We owe much to his constant feedback and his participation in our very constructive discussions.

References:

  • Antonsen, B. L. and Paul, D. H. (1997). Serotonin and octopamine elicit stereotypical agonistic behaviors in the squat lobster Munida quadrispina (Anomura, Galatheidae). J Comp Physiol [A] 181, 501-510.
  • Boake, C. R. B. and Hoikkala, A. (1995). Courtship Behavior and Mating Success of Wild-Caught Drosophila-Silvestris Males. Anim. Behav. 49, 1303-1313.
  • Boake, C. R. B. and Konigsberg, L. (1998). Inheritance of male courtship behavior, aggressive success, and body size in Drosophila silvestris. Evolution 52, 1487-1492.
  • Boake, C. R. B., Price, D. K. and Andreadis, D. K. (1998). Inheritance of behavioural differences between two interfertile, sympatric species, Drosophila silvestris and D- heteroneura. Heredity 80, 642-650.
  • Brand, A. H. and Perrimon, N. (1993). Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118, 401-415.
  • Buchner, E., Bader, R., Buchner, S., Cox, J., Emson, P. C., Flory, E., Heizmann, C. W., Hemm, S., Hofbauer, A. and Oertel, W. H. (1988). Cell-specific immuno-probes for the brain of normal and mutant Drosophila melanogaster. I. Wildtype visual system. Cell Tissue Res. 253, 357-370.
  • Buchner, E., Buchner, S., Crawford, G., Mason, W. T., Salvaterra, P. M. and Sattelle, D. B. (1986). Choline Acetyltransferase-Like Immunoreactivity in the Brain of Drosophila-Melanogaster. Cell Tissue Res. 246, 57-62.
  • Dow, M. A. and von Schilcher, F. (1975). Aggression and mating success in Drosophila melanogaster. Nature 254, 511-512.
  • Dubnau, J., Grady, L., Kitamoto, T. and Tully, T. (2001). Disruption of neurotransmission in Drosophila mushroom body blocks retrieval but not acquisition of memory. Nature 411, 476-480.
  • Edwards, D. H. and Kravitz, E. A. (1997). Serotonin, social status and aggression. Curr Opin Neurobiol 7, 812-819.
  • Greenspan, R. J. and Ferveur J. F. (2000). Courtship in Drosophila. Annu Rev Genet 34, 205-232.
  • Guo, A., Liu, L., Xia, S.-Z., Feng, C.-H., Wolf, R. and Heisenberg, M. (1996). Conditioned visual flight orientation in Drosophila; Dependence on age, practice and diet. Learn. Mem. 3, 49-59.
  • Han, K. A., Millar, N. S. and Davis, R. L. (1998). A novel octopamine receptor with preferential expression in Drosophila mushroom bodies. J Neurosci 18, 3650-3658.
  • Han, K. A., Millar, N. S., Grotewiel, M. S. and Davis, R. L. (1996). DAMB, a novel dopamine receptor expressed specifically in Drosophila mushroom bodies. Neuron 16, 1127-1135.
  • Heinrich, R., Braunig, P., Walter, I., Schneider, H. and Kravitz, E. A. (2000). Aminergic neuron systems of lobsters: morphology and electrophysiology of octopamine-containing neurosecretory cells. J Comp Physiol [A] 186, 617-629.
  • Heinrich, R., Cromarty, S. I., Horner, M., Edwards, D. H. and Kravitz, E. A. (1999). Autoinhibition of serotonin cells: an intrinsic regulatory mechanism sensitive to the pattern of usage of the cells. Proc Natl Acad Sci U S A 96, 2473-2478.
  • Heisenberg, M. (1971). Separation of receptor and lamina potentials in the electroretinogram of normal and mutant Drosophila. J. Exp. Biol., 85-100.
  • Heisenberg, M. (1972). Comparative behavioral studies on two visual mutants of Drosophila. J. Comp. Physiol. 80, 119-136.
  • Hoffmann, A. A. (1987). A Laboratory Study of Male Territoriality in the Sibling Species Drosophila-Melanogaster and Drosophila-Simulans. Anim. Behav. 35, 807-818.
  • Hoffmann, A. A. (1988). Heritable Variation for Territorial Success in 2 Drosophila- Melanogaster Populations. Anim. Behav. 36, 1180-1189.
  • Hoffmann, A. A. (1989). Geographic variation in the territorial success of Drosophila melanogaster males. Behav Genet 19, 241-255.
  • Hoffmann, A. A. (1994). Genetic Analysis of territoriality in Drosophila melanogaster. In Quantitative genetic studies of behavioural evolution, (ed. Boake, C.R.B.), pp. 188-205. Chicago: Chicago University press.
  • Hoffmann, A. A. and Cacoyianni, Z. (1989). Selection for Territoriality in Drosophila-Melanogaster - Correlated Responses in Mating Success and Other Fitness Components. Anim. Behav. 38, 23-34.
  • Hoffmann, A. A. and Cacoyianni, Z. (1990). Territoriality in Drosophila-Melanogaster as a Conditional Strategy. Anim. Behav. 40, 526-537.
  • Hovemann, B. T., Ryseck, R. P., Walldorf, U., Stortkuhl, K. F., Dietzel, I. D. and Dessen, E. (1998). The Drosophila ebony gene is closely related to microbial peptide synthetases and shows specific cuticle and nervous system expression. Gene 221, 1-9.
  • Huber, R., Orzeszyna, M., Pokorny, N. and Kravitz, E. A. (1997a). Biogenic amines and aggression: experimental approaches in crustaceans. Brain Behav Evol 50 Suppl 1, 60-68.
  • Huber, R., Smith, K., Delago, A., Isaksson, K. and Kravitz, E. A. (1997b). Serotonin and aggressive motivation in crustaceans: altering the decision to retreat. Proc Natl Acad Sci U S A 94, 5939-5942.
  • Jacobs, M. E. (1960). Influence of light on mating of Drosophila melanogaster. Ecology 41, 182-188.
  • Jacobs, M. E. (1978). Influence of beta-alanine on mating and territorialism in Drosophila melanogaster. Behav Genet 8, 487-502.
  • Johnson, W. A. and Hirsh, J. (1990). Binding of a Drosophila POU-domain protein to a sequence element regulating gene expression in specific dopaminergic neurons. Nature 343, 467-470.
  • Kitamoto, T. (2001). Conditional modification of behavior in Drosophila by targeted expression of a temperature-sensitive shibire allele in defined neurons. J Neurobiol 47, 81-92.
  • Kravitz, E. A. (2000). Serotonin and aggression: insights gained from a lobster model system and speculations on the role of amine neurons in a complex behavior. J Comp Physiol [A] 186, 221-238.
  • Lee, G. and Hall, J. C. (2000). A newly uncovered phenotype associated with the fruitless gene of Drosophila melanogaster: aggression-like head interactions between mutant males. Behav Genet 30, 263-275.
  • Lee, G. and Hall, J. C. (2001). Abnormalities of male-specific FRU protein and serotonin expression in the CNS of fruitless mutants in Drosophila. J Neurosci 21, 513-526.
  • Lundell, M. J. and Hirsh, J. (1994). Temporal and spatial development of serotonin and dopamine neurons in the Drosophila CNS. Dev Biol 165, 385-396.
  • Martin, J. R., Ernst, R. and Heisenberg, M. (1998). Mushroom bodies suppress locomotor activity in Drosophila melanogaster. Learn Mem 5, 179-191.
  • McGuire, S. E., Le, P. T. and Davis, R. L. (2001). The role of Drosophila mushroom body signaling in olfactory memory. Science 293, 1330-1333.
  • Monastirioti, M. (1999). Biogenic amine systems in the fruit fly Drosophila melanogaster. Microsc Res Tech 45, 106-121.
  • Monastirioti, M., Linn, C. E., Jr. and White, K. (1996). Characterization of Drosophila tyramine beta-hydroxylase gene and isolation of mutant flies lacking octopamine. J Neurosci 16, 3900-3911.
  • Neckameyer, W. S. (1998). Dopamine modulates female sexual receptivity in Drosophila melanogaster. J Neurogenet 12, 101-114.
  • Reif, M. (1998). The evolution of learning. PhD thesis, University of Würzburg, pp. 131.
  • Ringo, J., Kananen, M. K. and Wood, D. (1983). Aggression and Mating Success in 3 Species of Drosophila. Z Tierpsych - J Comp Ethol 61, 341-350.
  • Schneider, H., Budhiraja, P., Walter, I., Beltz, B. S., Peckol, E. and Kravitz, E. A. (1996). Developmental expression of the octopamine phenotype in lobsters, Homarus americanus. J Comp Neurol 371, 3-14.
  • Schulz, R. A., Chromey, C., Lu, M. F., Zhao, B. and Olson, E. N. (1996). Expression of the D-MEF2 transcription in the Drosophila brain suggests a role in neuronal cell differentiation. Oncogene 12, 1827-1831.
  • Shen, J., Beall, C. J. and Hirsh, J. (1993). Tissue-specific alternative splicing of the Drosophila dopa decarboxylase gene is affected by heat shock. Mol Cell Biol 13, 4549-4555.
  • Shen, J. and Hirsh, J. (1994). Cis-regulatory sequences responsible for alternative splicing of the Drosophila dopa decarboxylase gene. Mol Cell Biol 14, 7385-7393.
  • Skrzipek, K. H., Kroner, B. and Hager, H. (1979). Inter-Male Agression in Drosophila-Melanogaster - Laboratory Study. Z Tierpsych - J Comp Ethol 49, 87-103.
  • Stevenson, P. A., Hofmann, H. A., Schoch, K. and Schildberger, K. (2000). The fight and flight responses of crickets depleted of biogenic amines. J Neurobiol 43, 107-20.
  • Sweeney, S. T., Broadie, K., Keane, J., Niemann, H. and O'Kane, C. J. (1995). Targeted expression of tetanus toxin light chain in Drosophila specifically eliminates synaptic transmission and causes behavioral defects. Neuron 14, 341-351.
  • Vaysse, G., Galissie, M. and Corbiere, M. (1988). Induced variation of serotonin in Drosophila melanogaster and its relation to learning performance. J Comp Psychol 102, 225-229.
  • Waddell, S., Armstrong, J. D., Kitamoto, T., Kaiser, K. and Quinn, W. G. (2000). The amnesiac gene product is expressed in two neurons in the Drosophila brain that are critical for memory. Cell 103, 805-813.
  • Zamudio, K. R., Huey, R. B. and Crill, W. D. (1995). Bigger isn’t always better - body-size, developmental and parental temperature and male territorial success in Drosophila melanogaster. Anim Behav 49, 671-677.
  • Zars, T., Fischer, M., Schulz, R. and Heisenberg, M. (2000). Localization of a short-term memory in Drosophila. Science 288, 672-675.

Authors: Andrea Baier*, Britta Wittek* and Björn Brembs
* These authors contributed equally to this work.

Full reference: Baier A.; Wittek B. and Brembs B. (2002): Drosophila as a new model organism for the neurobiology of aggression? J. Exp. Biol. 205 (9): 1233–1240. (PDF)

 
homelearningevolutionmetabiology