Skip to main content

Tropical parabiotic ants: Highly unusual cuticular substances and low interspecific discrimination

Abstract

Background

Associations between animal species require that at least one of the species recognizes its partner. Parabioses are associations of two ant species which co-inhabit the same nest. Ants usually possess an elaborate nestmate recognition system, which is based on cuticular hydrocarbons and allows them to distinguish nestmates from non-nestmates through quantitative or qualitative differences in the hydrocarbon composition. Hence, living in a parabiotic association probably necessitates changes of the nestmate recognition system in both species, since heterospecific ants have to be accepted as nestmates.

Results

In the present study we report highly unusual cuticular profiles in the parabiotic species Crematogaster modiglianii and Camponotus rufifemur from the tropical rainforest of Borneo. The cuticle of both species is covered by a set of steroids, which are highly unusual surface compounds. They also occur in the Dufour gland of Crematogaster modiglianii in high quantities. The composition of these steroids differed between colonies but was highly similar among the two species of a parabiotic nest. In contrast, hydrocarbon composition of Cr. modiglianii and Ca. rufifemur differed strongly and only overlapped in three regularly occurring and three trace compounds. The hydrocarbon profile of Camponotus rufifemur consisted almost exclusively of methyl-branched alkenes of unusually high chain lengths (up to C49). This species occurred in two sympatric, chemically distinct varieties with almost no hydrocarbons in common. Cr. modiglianii discriminated between these two varieties. It only tolerated workers of the Ca. rufifemur variety it was associated with, but attacked the respective others. However, Cr. modiglianii did not distinguish its own Ca. rufifemur partner from allocolonial Ca. rufifemur workers of the same variety.

Conclusion

We conclude that there is a mutual substance transfer between Cr. modiglianii and Ca. rufifemur. Ca. rufifemur actively or passively acquires cuticular steroids from its Cr. modiglianii partner, while the latter acquires at least two cuticular hydrocarbons from Ca. rufifemur. The cuticular substances of both species are highly unusual regarding both substance classes and chain lengths, which may cause the apparent inability of Cr. modiglianii to discriminate Ca. rufifemur nestmates from allocolonial Ca. rufifemur workers of the same chemical variety.

Background

Associations across different animal taxa require specific adaptations on one or both sides. In particular, recognizing the partner species is a crucial task to any form of association, albeit in host-parasite associations only the latter might need to recognize the partner [1]. Nestmate recognition mechanisms in associating species must therefore go beyond the own species and include the partner species.

In ants, one of the closest and most intriguing interspecific associations is parabioses, where two ant species live together in a common nest. This phenomenon is found in several parts of the world, including Southeast Asia [2] and tropical South America [3]. Parabiotic ants have nestmates not only from their own colony, but also from a completely different species. Their nestmate recognition system therefore needs to include allospecific nestmates. In ants and other social hymenoptera, recognition is based on colony-specific chemical cues on the body surface that are perceived through olfactory or contact chemoreception [4, 5]. Most of them are hydrocarbons [6–8]. Via allogrooming and trophallaxis, the individuals continually take up their nestmates' surface compounds into the postpharyngeal gland (PPG), where they are mixed and redistributed. Through this process, a colony-specific odour is created [5, 9–11]. This colony-specific odour is learned by the colony members and represented as a neuronal template in the nervous system [12]. Nestmates are recognized by comparing the cuticular profile of the encountered individual to the neuronal template (phenotype matching), whereby a mismatch generally results in aggression [5].

Despite this complex nestmate recognition system, a considerable number of insect species manages to be accepted in Hymenoptera colonies, such as Lycaenid larvae, Staphylinidae, Ensifera, and Diptera [13–16] as well as social parasites, such as the parasitic bumblebee Psithyrus [17] and inquiline ant species [1, 6]. In many of these associations, the parasite chemically resembles the host (chemical mimicry) [1, 13–16, 18, 19]. Another possible mechanism to remain incognito is chemical insignificance [1]. Several social parasite species are – like callows – chemically insignificant, i.e. they do not possess an individual surface profile and are hence not recognized as foreign by their hosts [1, 20, 21]. Hydrocarbon profiles of very long chain lengths are difficult to perceive and hence may also promote chemical insignificance [22, 23]. Still, numerous other social parasite species possess distinct profiles that do not resemble their hosts. Since these profiles neither show chemical mimicry nor insignificance, it has been supposed that the host species habituate to the parasites' profiles [24–26].

While the chemical mechanisms of tolerance between species have been studied in associations like social parasitism, little is known about parabiotic associations. It seems likely that parabiotic ants possess a nestmate recognition system that tolerates allospecific nestmates. In the present study we examined the relationship between interspecific tolerance and surface chemistry among the Southeast Asian parabiotic species Crematogaster modiglianii and Camponotus rufifemur. The two species tolerate ants from certain (but not all) foreign parabiotic nests but attack non-parabiotic ant species [2]. We discovered that two morphological varieties of Ca. rufifemur (the 'red' and the 'black' variety, see Methods) also differ in their chemical profiles. This enabled us to study two different levels of chemical similarity – within and between the two varieties. Our research questions were:

(1) Do parabiotic species possess cuticular substances different from related, non-parabiotic species?

(2) Is there evidence for chemical mimicry, i.e., chemical overlap between parabiotic partners?

(3) Do chemical differences within species account for differences in interspecific allocolonial tolerance?

Results

Cuticular substances: Hydrocarbons and other aliphatic components

The cuticular profile of both Camponotus rufifemur and Crematogaster modiglianii highly differed from other, non-parabiotic Camponotus and Crematogaster species [9, 27–29]; unpublished data]. While there were only few aliphatic compounds with a chain length of C20-C33, both species possessed hydrocarbons of very high chain lengths (C35 up to C49, Figure 1) as well as steroids, which have not previously been detected on insect cuticles. The aliphatic profile of Crematogaster modiglianii consisted of hydrocarbons between C33 and C40. Beside n-alkanes and methyl-branched alkanes, more than 68% of its aliphatic cuticular compounds were unsaturated (Figure 1a, Tables 1, 2). Extracts of the body surface and postpharyngeal glands contained the same aliphatic substances in similar quantitative composition.

Figure 1
figure 1

Gas chromatograms of cuticular hydrocarbons of the parabiotic ant species. (a) Crematogaster modiglianii B2, (b) red Camponotus rufifemur R2, (c) black Camponotus rufifemur B4. Graphs were acquired with a GC-FID. Only substances beyond a chain length of 34 are shown since shorter hydrocarbons make up less than 2% of the profile. Numbers refer to table 1. *unknown, irregularly occurring substance. (d) Typical chromatogram of the cuticular steroids of Cr. modiglianii, acquired with GC-MS. Arrows indicate the steroid compounds common to both Cr. modiglianii and Ca. rufifemur. Asterisks indicate the three steroids with highly similar mass spectra used for the second Mantel test. No other steroids were present in the colony shown.

Table 1 Aliphatic cuticular substances found in Crematogaster modiglianii and the two varieties of Camponotus rufifemur
Table 2 Relative quantities of the different aliphatic substance classes in Cr. modiglianii and Ca. rufifemur.

The Camponotus rufifemur surface profile mainly contained compounds beyond C38, beside traces of lighter components. The two morphological varieties exhibit almost completely different surface profiles. The only substances in common were trace n-alkanes between C27 and C30 and C37-9-ene (Table 1). The red variety exhibited a highly unusual cuticular profile, 98% of the hydrocarbon quantities being methyl-branched alkenes. The main compounds, 27-MeC39-14-ene and 27-MeC39-16-ene, accounted for 88.7% of the total hydrocarbons. The other different methyl-branched alkenes were similar in respect to the positions of the methyl group and the double bond (Table 3). Chain lengths ranged from C38 to C41, with trace compounds between C24 and C37 (Figure 1b, Tables 1,2).

Table 3 Diagnostic ions of the methyl-branched alkenes in the red Ca. rufifemur variety.

The profile of the black Ca. rufifemur variety consisted of even larger molecules, with 92.8% of the surface compounds between C44 and C49 (Table 1). At least 80% of the compounds were unsaturated (Table 2). Methyl-branched alkenes were also present, albeit not as abundant as in the red Ca. rufifemur variety. Minor compounds included n-alkanes, methyl-branched alkanes and aldehydes (Table 2). In both Ca. rufifemur varieties, PPG and surface extracts contained the same aliphatic compounds in similar relative quantities.

Cuticular substances: Steroids

Besides aliphatic compounds, the surface profile of both ant species contained up to 24 components with a basic steroid structure (as inferred from mass spectra and diagnostic ions). Their mass spectra indicate a close chemical interrelatedness of the compounds. Due to the high substance quantities necessary for NMR analysis, their spatial molecular structure has not yet been resolved but is under investigation. Crematogaster modiglianii possessed high amounts of steroids on the body surface (2.59 ± 0.58 μg/worker, n = 11 colonies, mean and SE) which by far exceeded the hydrocarbons (0.48 ± 0.05 μg/worker, n = 11 colonies, mean and SE). In contrast, postpharyngeal gland extracts only contained minor amounts of steroids but high quantities of hydrocarbons. High steroid amounts of the same quantitative composition were also found in the Dufour gland, in separate alitrunk and gaster cuticular extracts and, albeit in lower amounts, in head cuticular extracts. They also occurred in cuticular extractions of living ants with SPME fibres, thus confirming that their presence in hexane extracts was not an artefact of concomitantly extracted glands. Altogether, Cr. modiglianii extracts contained 24 different steroid components with an abundance higher than 0.1% in at least one colony (percent of total steroid abundance). Their retention indices ranged between 20.38 and 25.77. Six of the 24 steroids were found in all Cr. modiglianii colonies in similar relative compositions. An additional eleven steroids were abundant in certain colonies but absent in others. The remaining seven steroids were irregularly found and never occurred in relative abundances higher than 1% (percent of total steroid abundance). Camponotus rufifemur extracts (both varieties) contained up to eight different steroids, all of which also occurred in Cr. modiglianii. The absolute steroid quantities in Ca. rufifemur were lower than the hydrocarbon quantities (black variety: 0.66 ± 0.22 μg steroids/worker and 1.79 ± 0.29 μg hydrocarbons/worker, n = 3 colonies; red variety: 0.41 ± 0.14 μg steroids/worker and 9.71 ± 3.79 μg hydrocarbons/worker, n = 4 colonies, mean and SE given).

Chemical overlap among the parabiotic species

Six hydrocarbons were shared between both parabiotic species. The red Ca. rufifemur variety shared three hydrocarbons with Cr. modiglianii. These were the two methyl-branched alkenes, 27-MeC39-14-ene and 27-MeC39-16-ene, which are the main constituents of the red Ca. rufifemur surface profile, and its saturated derivative, 13-MeC39 (Table 1). All three are absent in the black Ca. rufifemur variety. Cr. modiglianii colonies living with the red Ca. rufifemur variety (henceforth, 'red' Cr. modiglianii) exhibited significantly more 27-MeC39-14-ene and 27-MeC39-16-ene than those associated with the black variety (henceforth, 'black' Cr. modiglianii) (Mann-Whitney W = 30, p = 0.0043; N1 = 5, N2 = 6 colonies, Figure 2). The quantities of 13-MeC39 were not compared since they could not be separated from other methyl-branched C39 alkanes in Cr. modiglianii (Table 1). Traces of three other hydrocarbons common in Cr. modiglianii were detected in the black Ca. rufifemur variety (C35:1, C35, C37-9-ene, Table 1). Albeit the associated Cr. modiglianii possessed slightly more C37-9-ene than those living with the red variety, no significant differences were found.

Figure 2
figure 2

Relative abundance of 27-MeC39-14-ene and 27-MeC39-16-ene in Cr. modiglianii workers living with the red vs. Cr. modiglianii workers living with the black Ca. rufifemur variety. Median, quartiles, range, and outliers (i.e. all data points deviating from the box by more than 1.5 times the interquartile range) are shown in the present and the following figures. The number of analyzed colonies is given above each plot. ** highly significant (p = 0.0043) according to U test.

Eight of the steroids common in Cr. modiglianii were also frequently found in Ca. rufifemur (inclusion criterion: median abundance > 0% in 11 colonies of both species; Figure 1d). Their relative abundances varied between parabiotic nests but were significantly correlated among the two species within a nest (Mantel test: r = 0.49, p = 0.041, N = 11; Bray-Curtis distances: 0.13 ± 0.08 (Ca. rufifemur), 0.43 ± 0.31 (Cr. modiglianii); mean and s.d.). A second Mantel test considered only three steroids with very similar mass spectra, which were present in all extracts (retention indices: 21.92, 22.24, 24.47; marked with asterisks in Figure 1d). This test yielded a highly significant correlation of steroid abundance among the two species of each parabiotic nest (r = 0.620, p < 0.001, N = 11; Bray-Curtis distances: 0.06 ± 0.04 (Ca. rufifemur), 0.13 ± 0.07 (Cr. modiglianii)).

Differences in allocolonial tolerance

Chemical differences between the two Ca. rufifemur varieties accounted for much of the variance in interspecific confrontations. In general, Cr. modiglianii workers tolerated only allocolonial Ca. rufifemur workers of the variety they were associated with. The focal Crematogaster modiglianii colony, which lived together with the red Ca. rufifemur variety, showed high aggression towards dead workers of the black Ca. rufifemur variety but not towards those of the red one (Figure 3). The generalized linear model (GLM) for total aggression explained 65.6% of the total deviance and yielded a highly significant effect of the Camponotus variety (58.7% explained deviance, Table 4). The remaining deviance could in part be attributed to differences between Camponotus colonies (p = 0.04), whereas the difference between intracolonial and allocolonial Camponotus was not significant (p = 0.12, Table 4). The non-parabiotic Camponotus (Tanaemyrmex) arrogans was attacked to a similar degree as the black Ca. rufifemur variety (Figure 3). When the analysis focused on the proportion of strong aggression only, the results were similar, with slightly stronger effects. Cr. modiglianii very rarely climbed onto the Ca. rufifemur bodies in this experimental series ('mounting behaviour').

Figure 3
figure 3

Total aggression of Crematogaster modiglianii (colony R0) against different Camponotus colonies and species. Data are given as proportions in relation to the total number of interactions. Each plot represents 10 replicates.

Table 4 GLM for total aggression of Cr. modiglianii towards dead Ca. rufifemur workers from different colonies.

In the arena confrontations, Cr. modiglianii was significantly more aggressive towards Ca. rufifemur from the respective other variety. The parameter 'within/across variety' explained 18.2% of the total deviance, followed by 'variety combination' (13.0% explained deviance), while 'intra-/allocolonial' did not explain a significant part of the deviance (Table 5). Cr. modiglianii workers frequently climbed on Ca. rufifemur bodies and walked around on them for up to one minute. This 'mounting behaviour' represented on average 18.6% of all interactions (Figure 4). The workers (especially in one of the two colonies) mounted Ca. rufifemur workers of their 'own' variety in significantly higher proportions (GLM for both colonies: Fdf = 1 = 6.85, p = 0.011) but did not otherwise differentiate between intracolonial and allocolonial Ca. rufifemur workers (Fdf = 1 = 0.14, p = 0.71).

Figure 4
figure 4

Total aggression (a, b) and mounting behaviour (c, d) of Cr. modiglianii towards dead Ca. rufifemur from different colonies in arena assays. Data are given as proportions in relation to the total number of interactions. Each plot represents 10–13 replicates. (a), (c) Cr. Modiglianii B4, (b), (d) Cr. modiglianii R2.

Table 5 GLM for total aggression of Cr. modiglianii towards dead Ca. rufifemur from different colonies.

In order to examine whether the differentiation between the colour varieties occurred in colonies in situ as well, we re-analyzed previous behavioural experiments reported in [2]. Allocolonial aggression of Cr. modiglianii towards Ca. rufifemur was highly variable in this dataset, and we confirmed a high impact of the two chemical varieties on allocolonial aggression. The variable 'within/across varieties' (colonies A and B: black variety, colony C: red variety) explained 60.1% of the total variance of the data and was a clearly more powerful predictor than the differentiation between intra- vs. allocolonial combination (0.03% deviance explained, Table 6). 'Red' Cr. modiglianii colonies only attacked black Ca. rufifemur intruders and vice versa (Figure 5a). The highly significant impact of 'variety combination' (Table 6), however, showed that red Cr. modiglianii was more aggressive towards black Ca. rufifemur than black Cr. modiglianii towards red Ca. rufifemur.

Figure 5
figure 5

Total aggression in allocolonial confrontations between parabiotic partners assays, pooled for variety combinations. Data are from [2], given as proportions in relation to the total number of interactions. The numbers above each plot indicate the overall number of replicates and the number of colony combinations tested. P values are given according to GLMs with binomial error distribution. (a) Cr. modiglianii towards Ca. rufifemur, (b) Ca. rufifemur towards Cr. modiglianii.

Table 6 GLM for total aggression in interspecific live confrontations

In confrontations of Ca. rufifemur towards allocolonial Cr. modiglianii, Menzel et al. [2] had found low levels of aggression albeit they were higher than against intracolonial Cr. modiglianii. Similar to above, Ca. rufifemur workers were more aggressive towards Cr. modiglianii from the respective other variety (Figure 5b, Table 6).

Discussion

Unusual features of the cuticular profiles in parabiotic ants

To our knowledge, steroids have not been found in surface extracts of other ant species up to now, and to our knowledge have been found on insect cuticles only in one halictid bee [30]. However, various Crematogaster species are known to have highly efficient poisons [31, 32]. The genus Crematogaster has evolved a peculiar system of venom production which involves a cooperation of Dufour and poison gland. In several species the venom consists of precursors from the Dufour gland which are derivatized by enzymes from the poison gland [33, 34]. Crematogaster poisons – from Dufour and poison glands, but also from hypertrophied metapleural glands – belong to such different chemical classes as cyclohexan derivatives, crematofuranes (cembranoid diterpenes), coumarin derivatives, alkylphenols, alkylresorcinols, salicylic acids, resorcylic acids, and polyfunctionalized long-chain derivatives [33, 35–38]. Since extracts of Cr. modiglianii Dufour glands contained the same steroid composition as the body surface (but no other compounds), they are probably produced in this gland and then distributed onto the body surface. In Cr. modiglianii, steroid synthesis did not depend on biosynthetic precursors acquired from food. In two colonies kept in the laboratory for 15 and 6 months, respectively, the steroid profile did not change despite of an artificial diet of cockroaches, honey solution and Bhatkar diet (F.M. pers. obs.). Moreover, in one forest colony, the steroid profile remained relatively constant over three years, corroborating that the steroid composition is rather genetically determined than dependent on environmental factors.

It is notable that 98% of the entire hydrocarbon profile of the red Ca. rufifemur (and ≥ 13% of the black Ca. rufifemur hydrocarbon profile) were methyl-branched alkenes. This substance class seems to be generally very rare in insects and has been detected only in several Diptera and one Noctuid moth as pheromones [39–41]. Among ants, they have been found in traces in the ponerine ant Pachycondyla villosa and in two Leptothorax species [42, 43], but in higher abundances only in Nothomyrmecia macrops surface profiles, which is probably the most primitive existent ant species [44]. That they make up almost the entire hydrocarbon profile is therefore highly unusual. Another unusual feature in both parabiotic species is the high hydrocarbon chain lengths. Although common in this study (Table 1), hydrocarbons beyond C37 have not been found in non-parabiotic Camponotus and Crematogaster species [9, 27, 28]; unpublished data. Other studies report small concentrations of heavier hydrocarbons in other ant genera, but always accompanied by high amounts of lighter ones [45, 46]. It is possible that extremely long-chain hydrocarbons are difficult to perceive by receptors and thus promote interspecific tolerance [23, 47]. In one case, we observed that a non-parabiotic Cr. modiglianii colony was initially very aggressive against (black) Ca. rufifemur workers but treated them amicably (and had hence become habituated) after less than 24 h of exposure. Unsaturation in these long-chain hydrocarbons might be necessary to maintain a minimum fluidity of the cuticular profile [47].

Chemical overlap among parabiotic partners

Given the high allocolonial tolerance between parabiotic partners, the hydrocarbon overlap of the two species is surprisingly small. While the red Ca. rufifemur variety shared two compounds with its partner, the black variety only shared three trace compounds with Cr. modiglianii but otherwise possessed a completely different hydrocarbon profile. We tentatively suppose that Cr. modiglianii acquires 27-MeC39-14-ene 27-MeC39-16-ene from its red Ca. rufifemur partner although Ca. rufifemur generally tolerates Cr. modiglianii workers, including those lacking these substances [2]. In a Cr. modiglianii colony kept in the laboratory without its previous red Camponotus partner, the compound disappeared from the profile after eight months of separation (F.M. pers. obs.). It is possible that the other hydrocarbons of the red Ca. rufifemur are acquired by Cr. modiglianii as well but remain beyond detectability due to their low abundances. The hydrocarbons of the black Ca. rufifemur, in contrast, were never found on Cr. modiglianii surface extracts. This is probably due to their high chain lengths, which makes the cuticular profile more solid and do not allow chemical transfer [47]. In the light of the low overall hydrocarbon overlap among the two parabiotic ant species, chemical camouflage, a mechanism often found in social parasites [13–15], must be dismissed as an explanation for mutual tolerance. However, the existence of only few substances common to both species might be a sufficient signal for tolerating the partner [48].

The steroid components, in contrast, showed high congruence among both species. We found that the relative composition of eight steroid compounds differs between colonies but is very similar among the two species of a parabiotic nest. Since it is highly improbable that Ca. rufifemur is able to synthetically copy the steroid profile of each respective partner colony, this result suggests that Ca. rufifemur acquires steroids from Cr. modiglianii. Notably, only a certain set of steroids is transferred to Camponotus, while others, despite of high abundance in Cr. modiglianii, were almost or completely absent from the Ca. rufifemur profile.

Possible transfer mechanisms

Two mechanisms seem possible for the observed transfer of chemical cues, namely trophallaxis and direct physical contact. Via trophallaxis, individual ants exchange not only food but also the PPG content, i.e. hydrocarbons relevant for nestmate recognition [49]. The PPG of Cr. modiglianii indeed contained steroids, albeit in much lower concentrations than on the body surface, thus making trophallaxis a possible pathway for chemical transfer. Interspecific trophallaxis has been observed between the two parabiotic species (F.M. and A. Endler, pers. obs.) and also shown via stained food only fed to Cr. modiglianii (F.M., pers. obs.).

Another possible transfer mechanism is direct physical contact. We frequently observed that Cr. modiglianii climbed on living or dead Ca. rufifemur individuals (workers and alates). The latter sometimes tried to shake them off but did not show aggression. Though almost never observed in the field, this 'mounting behaviour' could be easily induced in the laboratory by keeping the two species separate for one or two days. Mounting may therefore represent another possible mechanism for transfer of surface chemicals.

Partner recognition is not colony-specific

The red and the black variety of Camponotus rufifemur are chemically distinct and – apart from trace compounds – do not share any hydrocarbons. The two dominant surface components of the red variety (substance #52, Table 1) are present in Crematogaster modiglianii colonies associated with this Ca. rufifemur variety but almost completely absent from those living with the black variety. Their abundance thus allows separating 'red' from 'black' Cr. modiglianii albeit the remaining surface profile is similar. The existence of two chemical Ca. rufifemur varieties accounts for most of the aggression variance in allocolonial encounters between the two species. Cr. modiglianii usually tolerated living or dead Ca. rufifemur workers of the same variety as their parabiotic partner but fiercely attacked those of the respective other variety (Figures 3, 4, 5, Tables 4, 5, 6). An analogous pattern was found in Ca. rufifemur. Despite of generally low aggression levels, black Ca. rufifemur workers were significantly more aggressive towards 'red' Cr. modiglianii workers than towards allocolonial 'black' Cr. modiglianii (Figure 5b). However, we did not detect a corresponding difference in the red Ca. rufifemur.

While much of the interspecific aggression can be explained by chemical differences, however, the low interspecific aggression within chemical varieties is still surprising. Rather than recognizing heterospecific nestmates, the two species seemingly recognize only the chemical variety of their partner and do not discriminate within these varieties. Nestmate recognition rather depends on volatile substances than on substances only perceivable through antennal contact [50]. Due to their low volatility [47], very long-chain hydrocarbons are less detectable than short-chain molecules. Thus, olfactory receptors may additionally absorb traces of lighter hydrocarbons, thereby blurring inter-colony profile differences and hampering inter-colony discrimination [23]. The role of the steroids in the nestmate discrimination process is still unclear and under investigation.

The high interspecific tolerance strongly contrasts with the South American parabioses of Crematogaster limata and the ponerine ant Odontomachus mayi, where the ants never tolerated heterospecific workers from foreign parabioses [3]. In these associations, very low chemical overlap was found (no substance data given), suggesting that both species habituated to each other's colony-specific profiles. The associated Chilean species Camponotus morosus and Solenopsis gayi also showed distinct hydrocarbon profiles [51]. In contrast to non-associated colonies, however, associated C. morosus had acquired small amounts of the S. gayi hydrocarbons. In both of these species, only individuals from associated colonies were tolerant towards allocolonial allospecifics [51], indicating that the acquisition of allospecific hydrocarbons promoted mutual tolerance.

Conclusion

In this study we document the cuticular chemistry of the parabiotically associated ant species Camponotus rufifemur and Crematogaster modiglianii. In contrast to neotropical parabioses, these ant species did not show heterospecific nestmate recognition. In our experiments, Cr. modiglianii did not discriminate its partner Ca. rufifemur colony from other Ca. rufifemur colonies of the same chemical variety (nor vice versa). Rather, Cr. modiglianii distinguished only between the two Ca. rufifemur varieties, accepting the familiar one but attacking the respective other. This reduced discrimination of heterospecific nestmates may be caused by two unusual properties of the cuticular surface: Transfer of Ca. rufifemur hydrocarbons to the Cr. modiglianii profile (in one of the Ca. rufifemur varieties only), and the generally high chain hydrocarbon lengths in the two parabiotic species. As hypothesized elsewhere [23], extremely long-chain hydrocarbons may be difficult to perceive by receptors and hence promote chemical insignificance (sensu [1]). It is currently investigated whether the cuticular steroids unique to these species play a role in nestmate or partner recognition.

Methods

Study site and ants

The studies were conducted at Danum Valley Conservation Area from September to November in the years 2004 and 2007. Danum Valley represents one of the major remaining patches of tropical lowland rainforest in Sabah (Malaysian Borneo). The site has a typical equatorial rainforest climate with a mean annual temperature of 26.9°C and a yearly rainfall of 2700 mm. We studied parabiotic associations of Camponotus (Myrmotarsus) rufifemur Emery 1900 and Crematogaster (Paracrema) modiglianii Emery 1900. Their nests are commonly found in hollow, living tree trunks in the rainforest. Extracts of one parabiotic nest from the Kuala Belalong Field Studies Center (Brunei) were analyzed in addition.

Camponotus rufifemur occurs in two sympatric morphological varieties that have not previously been described (although Emery [52] notes that specimen from Sarawak are darker in colour than those from Sumatra). While one variety (henceforth, 'red' variety) has a reddish alitrunk and light red-brown legs, the other one (henceforth, 'black' variety) possesses a black alitrunk and dark red-brown legs. The area between the frontal carinae of soldiers is dull in the red but shining in the black variety. Although the ratio head width/scape length (in frontal view) tends to be higher in the large soldier caste of the black variety than in that of the red variety, no significant morphometric differences were found. In the following, we will refer to the varieties as 'red' and 'black' Ca. rufifemur. In order to allow a differentiation, the respective associated Cr. modiglianii will be called 'red' and 'black' Cr. modiglianii although we did not find morphological distinctions within this species. Voucher specimen of Cr. modiglianii and both Ca. rufifemur varieties are deposited at the Department of Zoology III, University of Würzburg and at the Forest Resarch Center in Sepilok, Sabah (Malaysia).

Preparation of extracts

Extracts were prepared from both body surface and postpharyngeal glands (PPGs). For body rinses, 10 to 90 ants were killed by freezing and immersed in hexane for ten minutes. Extracts from single individuals contained quantities too low for reliable substance identification. Eleven parabiotic nests were sampled with one to eight (mean: 3.5) replicates per colony and species (ten from Danum Valley, one from Kuala Belalong). PPG extracts were obtained from three to four freshly dissected PPGs per sample dissolved in hexane. Octadecane (n-C18) was used as internal standard in most samples. Cuticular substances were additionally obtained from living Cr. modiglianii workers brought into the laboratory in Würzburg with solid-phase microextraction (SPME). A SPME fibre (Supelco) coated with a 100 μm polydimethylsiloxan film was rubbed on the ant for 3 min and then directly injected into a ThermoQuest Trace GC.

Chemical analysis

Substances were identified by coupled capillary gas chromatography-mass spectrometry (GC-MS) with a Hewlett Packard 6890 series gas chromatograph coupled to a HP 5973 Mass Selective Detector. The GC was equipped with a J&W Scientific DB-5 fused silica capillary column (30 m × 0.25 mm ID; df = 0.25 μm). Temperature was kept at 60°C for 2 min then increased by 60°C/min up to 200°C and subsequently by 4°C/min to 320°C, where it remained constant for 10 min. Helium was used as carrier gas with a constant flow of 1 ml/min. A split/splitless injector was installed at 250°C in the splitless mode for 30 s. The electron impact mass spectra (EI-MS) were recorded with an ionisation voltage of 70 eV, a source temperature of 230°C and an interface temperature of 325°C. For analysis of hydrocarbons beyond C41, we used a DB-1 HT column (30 m × 0.25 mm ID; df = 0,25 μm). Temperature was raised from 60°C by 5°C/min up to 350°C and then kept constant for 10 min. The interface had a temperature of 350°C. All other settings were as above. The software MSD ChemStation (Version A.03.00) for Windows was used for data acquisition. We restricted the analyses to substances with a retention time beyond that of C19 since compounds with shorter chain length are likely to be too volatile to be relevant for nestmate recognition [5, 6]. Substances present in less than 50% of the samples are given in Table 1 (marked with *) but were disregarded from further analysis.

For quantification of steroid-like compounds and aliphatics shorter than C33, we used ion counts from the GC-MS data and analysed both substance classes separately. Heavier hydrocarbons (beyond C33) were quantified using a high-resolution ThermoQuest Trace GC-FID with H2 as carrier gas in order to achieve a better separation of the substances. We used a nonpolar capillary column [DB1 (J&W Scientific, Folsom, CA), 20 m × 0.18 mm, 0.18 μm film thickness] and the first temperature program given above (split closed for 30 s for extracts and for 2 min when using SPME fibers). The split/splitless injector port was kept at 260°C and the flame ionization detector (FID) at 340°C. Peak areas were computed with Chrom-Card 1.19 (CE Instruments, Milan, Italy). Mean absolute substance quantities were estimated by comparing substance peak areas with that of the internal standard (acquired with GC-FID) and dividing by the number of extracted individuals.

Profile similarities between the two partner species were analyzed for eleven parabiotic nests (including one from Kuala Belalong Field Studies Center). The average proportions of the steroid components per colony and species were calculated. The distances between colonies were calculated for each species separately using Bray-Curtis index of similarity and then compared between species using a Mantel test (1000 permutations).

Identification of cuticular hydrocarbons

Alkanes, methyl-branched alkanes and alkenes were characterized using diagnostic ions and retention indices calculated using Kovats' method [53]. Unsaturated methyl-branched hydrocarbons were hydrated under a H2 atmosphere using Palladium on activated carbon as catalyst to determine the position of the methyl group. The position of the double bond in methyl-branched and n-alkenes was determined using DMDS derivatization following [54]. For methyl-branched alkenes, DMDS derivatization was insufficient for substance characterization since the position of the double bond relative to the methyl group remained unresolved and left two possible structures. Therefore, we cleaved the molecules in two parts at the position of the double bond via ozonisation. We diluted the sample in approx. 3 ml hexane, applied a constant flow of O3 (300 mg/h) for ten minutes from a glass pipette (EO3G Ozone Generator, Easelec Technology Inc.) and directly injected the sample into the GC-MS. Ozonisation succeeded for substance 52 but not for the substances 37, 43, 56, and 61 (surface compounds of the red Ca. rufifemur, Table 1). However, it is highly probable that all methyl-branched alkenes are produced via the same biosynthetic pathway. We therefore tentatively inferred the position of the double bond from the structure of substances 52 (Table 1) and possibilities left from the DMDS results, which had succeeded for all of the above substances. Double bond positions in alkenes with chain lengths higher than C41 as well as in dienes and trienes could not be determined due to their low abundance and/or their high chain length, which resulted in derivatives which could not be detected using GC-MS. Aldehydes were identified by comparing their mass spectra to a commercial library (Wiley 275) and therefore remain tentative.

For the substances of the black Ca. rufifemur profile beyond C44, retention indices were calculated based on the retention times of an n-alkane standard (C21 to C40), C47 and C49, and therefore remain preliminary. These substances were identified based on mass spectra and hydrated samples. Unsaturation was further confirmed via fractionation using a SiOH column treated with AgNO3. However, their characterization remains preliminary since the DMDS derivatized substances could not be detected using GC-MS.

Behavioural experiments

We studied the reaction of Crematogaster modiglianii towards dead Ca. rufifemur workers from different colonies in Borneo. The reverse situation (Ca. rufifemur towards Cr. modiglianii) was not studied in this paper since Ca. rufifemur shows little discrimination between different Cr. modiglianii workers [2]. A Cr. modiglianii colony (R0) had been collected in the forest circa one week prior to the experiments and was kept together with its red Ca. rufifemur partner in its original nest (a small tree trunk) in an open plastic box. The dead ants were placed onto the nest trunk with forceps such that several ants could interact with it simultaneously. During three minutes, each observed interaction was classified as peaceful (antennating), weakly (open mandibles) or strongly aggressive (biting or locking mandibles). An additional behaviour classified as peaceful was 'mounting', where the smaller Crematogaster (body length approx. 2–3 mm) climbed onto the Camponotus body (body length 5–13 mm). Continued interactions were recorded again after 10 s (the same behavioural classification as used in [2]).

The aggressiveness of two other Cr. modiglianii colonies was estimated in arena confrontations. The workers had been collected in the forest one day prior to the tests and were kept in a plastic box among nestmates (but separate from the partner species) over night. Five Cr. modiglianii individuals were placed into a fluon-covered plastic cylinder (Ø 7.5 cm, height 5 cm) on top of a paper sheet floor. After 1 min to calm down, a dead Camponotus specimen was introduced. For the following 100 s we recorded the behaviour of the ants as above. Each living or dead ant was used for one assay only. In all of the above assays, we performed ten replicates per treatment.

From each replicate we calculated the proportions of all aggressive versus all non-aggressive interactions. Both strong and total (including weak) aggression were analyzed using generalized linear models (GLM) with quasibinomial error distribution and logit link function. In order to determine whether confrontations within and across chemical varieties differ, we used the according explanatory variable 'within/across variety' with two factor levels (which collapsed to 'Camponotus variety' in the first dataset). The variable 'variety combination' (with the factor levels 'black→black', 'black→red', 'red→red', and 'red→black') was nested in the former one. Further explanatory variables were 'colony combination' (nested in 'variety combination'), which collapsed to 'Camponotus colony' in the first dataset, and 'intra-/allocolonial'. Due to their nested structure, no interactions between the variables were possible. The impact of each variable was determined by likelihood ratio tests (F tests). We also re-analyzed data from [2] in a similar way, where we included the number of workers present in the experimental arena as explanatory variables (see [2] for details on the experimental setup). Since the statistical results for total aggression and for strong aggression only were similar, only the former will be reported in the results section. All computations were performed in R Version 2.5.1 [55].

References

  1. Lenoir A, D'Ettorre P, Errard C, Hefetz A: Chemical Ecology and Social Parasitism in Ants. Annual Review of Entomology. 2001, 46: 573-599. 10.1146/annurev.ento.46.1.573.

    Article  CAS  PubMed  Google Scholar 

  2. Menzel F, Linsenmair KE, Blüthgen N: Selective interspecific tolerance in tropical Crematogaster-Camponotus associations. Animal Behaviour. 2008, 75: 837-846. 10.1016/j.anbehav.2007.07.005.

    Article  Google Scholar 

  3. Orivel J, Errard C, Dejean A: Ant gardens: Interspecific recognition in parabiotic ant species. Behavioral Ecology and Sociobiology. 1997, 40: 87-93. 10.1007/s002650050319.

    Article  Google Scholar 

  4. Hölldobler B, Wilson EO: The Ants. 1990, Berlin, Heidelberg: Springer Verlag

    Book  Google Scholar 

  5. Vander Meer RK, Morel L: Nestmate Recognition in Ants. Pheromone communication in social insects. Ants, wasps, bees, and termites. Edited by: Vander Meer RK, Breed MD, Espelie KE, Winston ML. 1998, Boulder: Westview Press, 79-103.

    Google Scholar 

  6. Lenoir A, Malosse C, Yamaoka R: Chemical mimicry between parasitic ants of the genus Formicoxenus and their host Myrmica (Hymenoptera, Formicidae). Biochemical Systematics and Ecology. 1997, 25: 379-389. 10.1016/S0305-1978(97)00025-2.

    Article  CAS  Google Scholar 

  7. Lahav S, Soroker V, Hefetz A: Direct behavioral evidence for hydrocarbons as ant recognition discriminators. Naturwissenschaften. 1999, 86: 246-249. 10.1007/s001140050609.

    Article  CAS  Google Scholar 

  8. Wagner D, Tissot M, Cuevas W, Gordon DM: Harvester ants utilize cuticular hydrocarbons in nestmate recognition. Journal of Chemical Ecology. 2000, 26: 2245-2257. 10.1023/A:1005529224856.

    Article  CAS  Google Scholar 

  9. Boulay R, Hefetz A, Soroker V, Lenoir A: Camponotus fellah colony integration: worker individuality necessitates frequent hydrocarbon exchanges. Animal Behaviour. 2000, 59: 1127-1133. 10.1006/anbe.2000.1408.

    Article  PubMed  Google Scholar 

  10. Boulay R, Katzav-Gozansky T, Hefetz A, Lenoir A: Odour convergence and tolerance between nestmates through trophallaxis and grooming in the ant Camponotus fellah (Dalla Torre). Insectes Sociaux. 2004, 51: 55-61. 10.1007/s00040-003-0706-0.

    Article  Google Scholar 

  11. Soroker V, Vienne C, Hefetz A, Nowbahari E: The Postpharyngeal Gland as a "Gestalt" Organ for Nestmate Recognition in the Ant Cataglyphis niger. Naturwissenschaften. 1994, 81: 510-513.

    CAS  Google Scholar 

  12. Errard C, Hefetz A, Jaisson P: Social discrimination tuning in ants: template formation and chemical similarity. Behavioral Ecology and Sociobiology. 2006, 59: 353-363. 10.1007/s00265-005-0058-z.

    Article  Google Scholar 

  13. Akino T: Chemical camouflage by myrmecophilous beetles Zyras comes (Coleoptera: Staphylinidae) and Diaritiger fossulatus (Coleoptera: Pselaphidae) to be integrated into the nest of Lasius fuliginosus (Hymenoptera: Formicidae). Chemoecology. 2002, 12: 83-89. 10.1007/s00049-002-8330-4.

    Article  CAS  Google Scholar 

  14. Akino T, Knapp JJ, Thomas JA, Elmes GW: Chemical mimicry and host specificity in the butterfly Maculinea rebeli, a social parasite of Myrmica ant colonies. Proceedings of the Royal Society of London, Series B. 1999, 266: 1419-1426. 10.1098/rspb.1999.0796.

    Article  CAS  Google Scholar 

  15. Akino T, Mochizuki R, Morimoto M, Yamaoka R: Chemical camouflage of myrmecophilous cricket Myrmecophilus. sp to be integrated with several ant species. Japanese Journal of Applied Entomology and Zoology. 1996, 40: 39-46.

    Article  CAS  Google Scholar 

  16. Howard RW, Stanley-Samuelson DW, Akre RD: Biosynthesis and chemical mimicry of cuticular hydrocarbons from an obligate predator, Microdon albicomatus Novak (Diptera: Syrphidae) and its ant prey, Myrmica incompleta Provancheri (Hymenoptera: Formicidae). Journal of the Kansas Entomological Society. 1990, 63: 437-443.

    Google Scholar 

  17. Dronnet S, Simon X, Verhaeghe J-C, Rasmont P, Errard C: Bumblebee inquilinism in Bombus (Fernaeldaepsithyrus) sylvestris (Hymenoptera, Apidae): behavioural and chemical analyses of host-parasite interactions. Apidologie. 2005, 36: 59-70. 10.1051/apido:2004070.

    Article  Google Scholar 

  18. Elgar MA, Allan RA: Predatory spider mimics acquire colony-specific cuticular hydrocarbons from their ant model prey. Naturwissenschaften. 2004, 91: 143-147. 10.1007/s00114-004-0507-y.

    Article  CAS  PubMed  Google Scholar 

  19. Howard RW, Perez-Lachaud G, Lachaud JP: Cuticular hydrocarbons of Kapala sulcifacies (Hymenoptera: Eucharitidae) and its host, the ponerine ant Ectatomma ruidum (Hymenoptera: Formicidae). Annals of the Entomological Society of America. 2001, 94: 707-716. 10.1603/0013-8746(2001)094[0707:CHOKSH]2.0.CO;2.

    Article  CAS  Google Scholar 

  20. D'Ettorre P, Mondy N, Lenoir A, Errard C: Blending in with the crowd: social parasites integrate into their host colonies using a flexible chemical signature. Proceedings of the Royal Society of London, Series B. 2002, 269: 1911-1918. 10.1098/rspb.2002.2110.

    Article  Google Scholar 

  21. Jeral JM, Breed MD, Hibbard BE: Thief ants have reduced quantities of cuticular compounds in a ponerine ant, Ectatomma ruidum. Physiological Entomology. 1997, 22: 207-211. 10.1111/j.1365-3032.1997.tb01160.x.

    Article  CAS  Google Scholar 

  22. Akino T: Cuticular hydrocarbons of Formica truncorum (Hymenoptera: Formicidae): Description of new very long chained hydrocarbon components. Applied Entomology and Zoology. 2006, 41: 667-677. 10.1303/aez.2006.667.

    Article  CAS  Google Scholar 

  23. Lambardi D, Dani FR, Turillazzi S, Boomsma JJ: Chemical mimicry in an incipient leaf-cutting ant social parasite. Behavioral Ecology and Sociobiology. 2007, 61: 843-851. 10.1007/s00265-006-0313-y.

    Article  Google Scholar 

  24. Errard C, Regla JI, Chambris A: Interspecific recognition in chilean parabiotic ant species. Presented at 3rd Jornadas Chile – Francia di Etologia e Sociobiologia. 1996, Universidad Mayor, Santiago, Chile

    Google Scholar 

  25. Liu ZB, Yamane S, Yamamoto H, Wang QC: Nestmate discrimination and cuticular profiles of a temporary parasitic ant Lasius sp. and its host L. fuliginosus (Hymenoptera, Formicidae). Journal of Ethology. 2000, 18: 69-73. 10.1007/s101640070002.

    Article  Google Scholar 

  26. Liu ZB, Bagneres AG, Yamane S, Wang QC, Kojima J: Cuticular hydrocarbons in workers of the slave-making ant Polyergus samurai and its slave, Formica japonica (Hymenoptera: Formicidae). Entomological Science. 2003, 6: 125-133. 10.1046/j.1343-8786.2003.00014.x.

    Article  Google Scholar 

  27. Boulay R, Katzav-Gozansky T, Vander Meer RK, Hefetz A: Colony insularity through queen control on worker social motivation in ants. Proceedings of the Royal Society of London, Series B. 2003, 270: 971-977. 10.1098/rspb.2002.2325.

    Article  Google Scholar 

  28. Endler A, Liebig J, Schmitt T, Parker JE, Jones GR, Schreier P, Hölldobler B: Surface hydrocarbons of queen eggs regulate worker reproduction in a social insect. Proceedings of the National Academy of Sciences of the United States of America. 2004, 101: 2945-2950. 10.1073/pnas.0308447101.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  29. Meskali M, Bonavitacougourdan A, Provost E, Bagneres AG, Dusticier G, Clement JL: Mechanism underlying cuticular hydrocarbon homogeneity in the ant Camponotus vagus (Scop) (Hymenopters, Formicidae) – role of postpharyngeal glands. Journal of Chemical Ecology. 1995, 21: 1127-1148. 10.1007/BF02228316.

    Article  CAS  PubMed  Google Scholar 

  30. Ayasse M, Engels W, Lübke G, Taghizadeh T, Francke W: Mating expenditures reduced via female sex pheromone modulation in the primitively eucosial halictine bee, Lasioglossum (Evylaeus) malachurum (Hymenoptera: Halictidae). Behavioral Ecology and Sociobiology. 1999, 45: 95-106. 10.1007/s002650050543.

    Article  Google Scholar 

  31. Heredia A, De Bizeau JC, Quinet Y: Toxicity of the venom in three neotropical Crematogaster ants (Formicidae: Myrmicinae). Chemoecology. 2005, 15: 235-242. 10.1007/s00049-005-0317-5.

    Article  Google Scholar 

  32. Marlier JF, Quinet Y, de Bizeau JC: Defensive behaviour and biological activities of the abdominal secretion in the ant Crematogaster scutellaris (Hymenoptera: Myrmicinae). Behavioural Processes. 2004, 67: 427-440. 10.1016/j.beproc.2004.07.003.

    Article  CAS  PubMed  Google Scholar 

  33. Leclerq S, Braekman JC, Kaisin M, Daloze D: Venom constituents of three species of Crematogaster ants from Papua New Guinea. Journal of natural products. 1997, 60: 1143-1147. 10.1021/np970202h.

    Article  Google Scholar 

  34. Pasteels JM, Daloze D, Boeve JL: Aldehydic contact poisons and alarm pheromone of the ant Crematogaster scutellaris (Hymenoptera: Myrmicinae) – enzyme mediated production from acetate precursors. Journal of Chemical Ecology. 1989, 15: 1573-1561. 10.1007/BF01012379.

    Article  Google Scholar 

  35. Daloze D, de Biseau JC, Leclerq S, Braekman J-C, Quinet Y, Pasteels JM: (13E,15E,18Z,20Z)-1-Hydroxypentacosa-13,15,18,20-tetraen-11-yn-4-one 1 acetate, from the venom of a Brazilian Crematogaster ant. Tetrahedron Letters. 1998, 39: 4671-4672. 10.1016/S0040-4039(98)00875-2.

    Article  CAS  Google Scholar 

  36. Jones TH, Brunner SR, Edwards AA, Davidson DW, Snelling RR: 6-Alkylsalicylic acids and 6-alkylresorcylic acids from ants in the genus Crematogaster from Brunei. Journal of Chemical Ecology. 2005, 31: 407-417. 10.1007/s10886-005-1349-6.

    Article  CAS  PubMed  Google Scholar 

  37. Laurent P, Hamdani A, Braekman JC, Daloze D, Isbell LA, de Biseau JC, Pasteels JM: New 1-alk(en)yl-1,3,5-trihydroxycyclohexanes from the Dufour gland of the African ant Crematogaster nigriceps. Tetrahedron Letters. 2003, 44: 1383-1386. 10.1016/S0040-4039(02)02870-8.

    Article  CAS  Google Scholar 

  38. Leclerq S, de Biseau JC, Braekman JC, Daloze D, Quinet Y, Luhmer M, Sundin A, Pasteels JM: Furanocembranoid diterpenes as defensive compounds in the dufour gland of the ant Crematogaster brevispinosa rochai. Tetrahedron. 2000, 56: 2037-2042. 10.1016/S0040-4020(00)00113-7.

    Article  Google Scholar 

  39. Carlson DA, Mramba F, Sutton BD, Bernier UR, Geden CJ, Mori K: Sex pheromone of the tsetse species, Glossina austeni : isolation and identification of natural hydrocarbons, and bioassay of synthesized compounds. Medical and veterinary entomology. 2005, 19: 470-479. 10.1111/j.1365-2915.2005.00596.x.

    Article  CAS  PubMed  Google Scholar 

  40. Etges WJ, Jackson LL: Epicuticular hydrocarbon variation in Drosophila mojavensis cluster species. Journal of Chemical Ecology. 2001, 27: 2125-2149. 10.1023/A:1012203222876.

    Article  CAS  PubMed  Google Scholar 

  41. Francke W, Plass E, Zimmermann N, Tietgen H, Tolasch T, Franke S, Subchev M, Toshova T, Pickett JA, Wadhams LJ, Woodcock CM: Major sex pheromone component of female herald moth Scoliopteryx libatrix is the novel branched alkene (6Z, 13)-methylheneicosene. Journal of Chemical Ecology. 2000, 26: 1135-1149. 10.1023/A:1005475725742.

    Article  CAS  Google Scholar 

  42. Tentschert J, Bestmann HJ, Heinze J: Cuticular compounds of workers and queens in two Leptothorax ant species – a comparison of results obtained by solvent extraction, solid sampling, and SPME. Chemoecology. 2002, 12: 15-21. 10.1007/s00049-002-8322-4.

    Article  CAS  Google Scholar 

  43. Lucas C, Fresneau D, Kolmer K, Heinze J, Delabie JHC, Pho DB: A multidisciplinary approach to discriminating different taxa in the species complex Pachycondyla villosa (Formicidae). Biological Journal of the Linnean Society. 2002, 75: 249-259. 10.1111/j.1095-8312.2002.tb01425.x.

    Article  Google Scholar 

  44. Brown WV, Jaisson P, Taylor RW, Lacey MJ: Novel internally branched, internal alkenes as major components of the cuticular hydrocarbons of the primitive Australian ant Nothomyrmecia macrops Clark (Hymenoptera: Formicidae). Journal of Chemical Ecology. 1990, 16: 2623-2635. 10.1007/BF00988074.

    Article  CAS  PubMed  Google Scholar 

  45. Dalecky A, Renucci M, Tirard A, Debout G, Roux M, Kjellberg F, Provost E: Changes in composition of cuticular biochemicals of the facultatively polygynous ant Petalomyrmex phylax during range expansion in Cameroon with respect to social, spatial and genetic variation. Molecular Ecology. 2007, 16: 3778-3791. 10.1111/j.1365-294X.2007.03429.x.

    Article  PubMed  Google Scholar 

  46. Lucas C, Pho DB, Fresneau D, Jallon JM: Hydrocarbon circulation and colonial signature in Pachycondyla villosa. Journal of Insect Physiology. 2004, 50: 595-607. 10.1016/j.jinsphys.2004.04.006.

    Article  CAS  PubMed  Google Scholar 

  47. Gibbs A, Pomonis JG: Physical properties of insect cuticular hydrocarbons: The effects of chain lengths, methyl branching and unsaturation. Comparative Biochemistry and Physiology. 1995, 112B: 243-249.

    Article  CAS  Google Scholar 

  48. D'Ettorre P, Heinze J, Schulz C, Francke W, Ayasse M: Does she smell like a queen? Chemoreception of a cuticular hydrocarbon signal in the ant Pachycondyla inversa. Journal of Experimental Biology. 2004, 207: 1085-1091. 10.1242/jeb.00865.

    Article  PubMed  Google Scholar 

  49. Lenoir A, Fresneau D, Errard C, Hefetz A: Individuality and colonial identity in ants: the emergence of a social representation concept. Information Processing in Social Insects. Edited by: Detrain C, Pasteels JM, Deneubourg JL. 1999, Basel: Birkhäuser, 219-237.

    Chapter  Google Scholar 

  50. Brandstaetter AS, Endler A, Kleineidam CJ: Nestmate recognition in ants is possible without tactile interaction. Naturwissenschaften. 2008, 95: 601-608. 10.1007/s00114-008-0360-5.

    Article  CAS  PubMed  Google Scholar 

  51. Errard C, Regla JI, Hefetz A: Interspecific recognition in Chilean parabiotic ant species. Insectes Sociaux. 2003, 50: 268-273. 10.1007/s00040-003-0677-1.

    Article  Google Scholar 

  52. Emery C: Formiche raccolte da Elio Modigliani in Sumatra, Engano e Mentawei. Annali del Museo Civico di Storia Naturale di Genova, Serie 2. 1900, 20: 661-722.

    Google Scholar 

  53. Carlson DA, Bernier UR, Sutton BD: Elution patterns from capillary GC for methyl-branched alkanes. Journal of Chemical Ecology. 1998, 24: 1845-1865. 10.1023/A:1022311701355.

    Article  CAS  Google Scholar 

  54. Dunkelblum E, Tan SH, Silk PJ: Double-bond location in monounsaturated fatty acids by dimethyl disulfide derivatization and mass spectrometry: Application to analysis of fatty acids in pheromone glands of four Lepidoptera. Journal of Chemical Ecology. 1985, 11: 265-277. 10.1007/BF01411414.

    Article  CAS  PubMed  Google Scholar 

  55. R Development Core Team: A language and environment for statistical computing. 2007, Vienna, R Foundation for Statistical Computing, [http://www.R-project.org]

    Google Scholar 

Download references

Acknowledgements

We thank Annett Endler and Heike Feldhaar for support regarding chemical analyses. Brigitte Fiala kindly provided samples from Brunei for further analysis. We are grateful to the Danum Valley Management Committee (DVMC) and Malaysian Economic Planning Unit (EPU) for research permission, the Royal Society's South East Asia Rainforest Research Programme (SEARRP) and the staff at Danum Valley Field Center, especially Bernadus 'Mike' Bala Ola for logistic support and help in the field. Dr. Arthur Chung (Forest Research Center, Sepilok) kindly supported our work as local collaborator. Finally, we would like to thank three anonymous reviewers for helpful comments on the manuscript. This research was supported by the Sonderforschungsbereich 'Mechanisms and Evolution of Arthropod Behaviour' (SFB-554) of the German Research Foundation (DFG) and complies with current laws of Malaysia and Germany. F.M. was supported by a doctoral fellowship provided by the German National Academic Foundation (Studienstiftung des deutschen Volkes) and the Bayerische Graduiertenförderung.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Schmitt.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

NB conceived the study and designed the experimental setup. FM collected the samples, performed the quantitative analysis of the cuticular substances, the behavioural experiments and the statistical analyses and wrote the manuscript. TS contributed significantly to the concept and the design of the study and identified the cuticular substances. Both NB and TS contributed to the preparation of the manuscript. All authors read and approved the final manuscript.

Authors’ original submitted files for images

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Menzel, F., Blüthgen, N. & Schmitt, T. Tropical parabiotic ants: Highly unusual cuticular substances and low interspecific discrimination. Front Zool 5, 16 (2008). https://0-doi-org.brum.beds.ac.uk/10.1186/1742-9994-5-16

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/1742-9994-5-16

Keywords