Skip to main content

Molecular evolution of neuropeptides in the genus Drosophila

Abstract

Background

Neuropeptides comprise the most diverse group of neuronal signaling molecules. They often occur as multiple sequence-related copies within single precursors (the prepropeptides). These multiple sequence-related copies have not arisen by gene duplication, and it is debated whether they are mutually redundant or serve specific functions. The fully sequenced genomes of 12 Drosophila species provide a unique opportunity to study the molecular evolution of neuropeptides.

Results

We data-mined the 12 Drosophila genomes for homologs of neuropeptide genes identified in Drosophila melanogaster. We then predicted peptide precursors and the neuropeptidome, and biochemically identified about half of the predicted peptides by direct mass spectrometric profiling of neuroendocrine tissue in four species covering main phylogenetic lines of Drosophila. We found that all species have an identical neuropeptidome and peptide hormone complement. Calculation of amino acid distances showed that ortholog peptide copies are highly sequence-conserved between species, whereas the observed sequence variability between peptide copies within single precursors must have occurred prior to the divergence of the Drosophila species.

Conclusion

We provide a first genomic and chemical characterization of fruit fly neuropeptides outside D. melanogaster. Our results suggest that neuropeptides including multiple peptide copies are under stabilizing selection, which suggests that multiple peptide copies are functionally important and not dispensable. The last common ancestor of Drosophila obviously had a set of neuropeptides and peptide hormones identical to that of modern fruit flies. This is remarkable, since drosophilid flies have adapted to very different environments.

Background

Neuropeptides comprise the most diverse group of intercellular signaling molecules in eumetazoan animals and regulate vital physiological processes as hormones, neuromodulators or neurotransmitters. Since neuropeptides are too small to be directly channeled into the regulated secretory pathway, they are post-translationally processed from larger prepropeptides by enzymatic cleavage.

In vertebrates, gene or genome duplications are main events that have led to the diversity of neuropeptides [1–4]. Over time, each prepropeptide gene acquires nucleotide substitutions that - if inside a peptide-coding sequence and not synonymous - will result in altered peptide sequence. If the peptide's function is vital and interference with peptide signaling decreases Darwinian fitness, there will be stabilizing selection on at least that part of the peptide sequence responsible for receptor binding and activation. In consequence, the peptide sequence will be conserved over time [4]. In fact, the sequences of many ortholog neuropeptides, such as oxytocin or somatostatin, have been highly conserved throughout vertebrate phylogeny [4]. However, considerable sequence variation can be found between duplicated peptides of a family, for example, in the growth hormone-releasing factor superfamily [5]. According to a classic model of molecular evolution [6], this is because a duplicated peptide sequence may be able to escape from natural selection and drift neutrally [7] if its original function is maintained by its paralog. In principle, the mutations accumulating in the 'escaped' peptide sequence may then lead to nonfunctionalization, subfunctionalization or neofunctionalization by acquisition of new features such as altered half-life, altered receptor binding kinetics, altered tissue expression patterns (for example, neuropeptides of the NPY family or the POMC prepropeptide [1, 8]) or receptor specificities by peptide-receptor co-evolution [9, 10]. If sub- or neofunctionalized, the new peptide will undergo positive selection for the new function and so become constrained by purifying selection. If the increased amount of peptides resulting from the duplication is beneficial, the duplicated peptide may also immediately increase Darwinian fitness prior to an accumulation of sequence mutations ('more-of-the-same') [10, 11].

A special feature of many neuropeptides that cannot be explained by gene duplication is the occurrence of multiple members of one peptide family within a single prepropeptide. For example, vertebrate prepropeptides encoding, melanocortins, hypocretins, RFamides or tachykinins, contain two to a few members of a single peptide family [12]. In invertebrates, copy numbers can reach even higher numbers. Examples include 37 related peptides from the metamorphosin A precursor of the sea anemone Anthopleura elegantissima [13], 24 different FMRFa-like peptides encoded by the fmrf gene of the cockroach Periplaneta americana [14], 35 FGLamides from the allatostatin precursor of the prawn Macrobrachium rosenbergii [15], up to nine RFamides encoded per flp genes of Caenorhabditis elegans [16], and 35 enterins contained in the enterin precursor of Aplysia [17]. These multiple copies are encoded on the same gene, and often even on the same exon. They most likely have arisen by unequal recombination between nearly identical nucleotide stretches. This has the important consequence that, unlike peptides generated by gene or genome duplication, these copies cannot move to a new genomic location and acquire promoter-driven differential spatial or temporal expression patterns since they are encoded on the same gene, and they cannot be specifically silenced when located on the same exon. Multiple copies are thus equal at birth, at least on the genetic level [18]. Unlike for peptides originating from whole genome duplications, there is also no co-duplicated receptor as a directly available partner for sub- or neofunctionalization. It is therefore difficult to fit them directly into the established models of molecular evolution for duplicated peptide genes [1, 2, 4].

At least two questions arise from this: is the molecular evolution of multiple copy neuropeptides similar to that of duplicated peptides? And more importantly, what is the functional significance of the individual multiple copies contained in given prepropeptides - a long-standing problem in invertebrate neuroendocrinology (see, for example, [19–22]). At one extreme, each peptide copy may have its unique and specific function, receptor or expression pattern. On the other extreme, peptide copies may be functionally redundant if they are co-expressed, co-released and also share an identical effect space [21]. Among others, studies on the effect of multiple co-expressed peptide copies on the neuromuscular junction of Aplysia and Drosophila provide evidence for such a redundancy [22, 23], but differential activities might be found when looking at, for example, different developmental times or target sites. In fact, other studies speak against a functional redundancy, and report differential target-specific effects of multiple copy peptides in insects and molluscs (for example, [19, 24–26]).

To comprehensively investigate whether multiple peptide copies are functionally redundant is extremely difficult by experimental means, especially since peptide copies can show different half-lives in the circulation after release (for example, [27]), or differentially activate the same receptor (for example, [28]). It is also difficult to assess the functional importance of individual copies by genetic means since common techniques target the whole gene. We here have chosen an evolutionary and comparative genomic approach to address the functional significance of multiple peptide copies. This opportunity has recently become possible with the publication of the genomes of 12 different Drosophila species [29]. A standard nomenclature that refers to multiple peptides belonging to the same peptide family located on the same precursor does not exist. Based on [30], we will use the following terminology (see Figure 1): peptide copies aligning at the same position within the precursors of different species will be referred to as orthocopies. Orthocopies do not have to be sequence identical. The different peptide copies within a prepropeptide of a single species are paracopies (that is, not at the same location). The term 'isoform', which has often been used in conjunction with insect neuropeptides, will be avoided because of its differing usage in protein nomenclature.

Figure 1
figure 1

Terminology and amino acid distances. (ai) Peptide copy terminology exemplified by three aligned ASTa prepropeptides from species a1-3. (aii) Processing at dibasic processing sites (indicated in red in (ai)) yields the four neuropeptides ASTa1-4. The C-terminal glycine is further processed to yield the C-terminal amidation. Peptide copies aligning at the same position in the precursor (for example, ASTa1 of species a1-3) will be referred to as orthocopies, which do not have to be sequence-identical. The different copies in a precursor of a single species are paracopies (for example, ASTa1-4 of species a1) = not at the same location. Paracopies may or may not be sequence-identical. (b) Different types of amino acid distances obtained by pairwise comparisons. (bi) The average distance Do between orthocopies is the arithmetic mean of all individual pairwise distances. It does not contain distances between different paracopies. (bii) The average distance between all peptides within a family Df is the arithmetic mean of all individual pairwise distances. It contains all pairwise distances between orthocopies and all paracopies. (biii) The net distance Dnp between paracopies is similar to Df after subtraction of Do. It does not contain the pairwise distances between each set of orthocopies.

We mined the Drosophila genome database [31] for genes encoding homologs of all known D. melanogaster neuropeptide precursor (prepropeptides) encoding neuropeptides up to a size of 50 amino acids. The investigated species belong to the Drosophila and Sophophora subgenera that diverged 40-60 million years ago [32, 33] and contain 97% of the more than 1,000 Drosophila species [34]. We then predicted ortho- and paracopies and analyzed their amino acid sequence variation. This is appropriate since most selection pressure is on the peptide sequence and not on the underlying DNA sequence with its often redundant third codon position. Our reasoning was as follows: if peptides are functionally important and their loss decreases Darwinian fitness, their sequence will be under stabilizing selection and hence their sequence will be conserved in the different species. If peptides have no functional importance and their (functional) loss does not affect fitness, they will be able to escape selection pressure and will accumulate sequence variations during Drosophila radiation. Thus, if peptide copies are functionally unimportant, we expect a high sequence variation between at least some orthocopies that were able to escape from selection pressure since one or several of their fellow paracopies 'do the job' and hence are under stabilizing selection. This in consequence would lead to an increased sequence variation between paracopies. If peptide copies have a functional importance, we expect low sequence variation between all orthocopies due to stabilizing selection. If the different paracopies activate different receptors or induce different receptor conformations that lead to activation of different intracellular signaling pathways, we expect at the same time an increased sequence variation between paracopies due to subfunctionalization. If peptide copies are individually redundant but functionally important along the 'more-of-the-same' concept, we expect low sequence variation between both ortho- and paracopies.

Our study assumes that neuropeptides are expressed and processed as predicted in silico from the genome. This is not given per se, since neuropeptides can undergo differential splicing and post-translational processing. To biochemically underpin our assumption in a manageable amount of time, direct MALDI-TOF (matrix-assisted laser desorption ionization-time of flight) mass spectrometric peptide profiling lends itself as a fast and reliable method. We therefore directly profiled the major neuropeptide release sites of four species covering the main Drosophila lineages. In D. melanogaster, these sites contain about 50% of all biochemically identified neuropeptides and the majority of peptide hormones [35–37].

Our data provide a first genomic prediction of neuropeptides and prepropeptides, and the first chemical neuropeptide characterizations for the newly sequenced Drosophila species.

The results suggest that both the peptidome and the peptide hormone complement are conserved throughout Drosophila, and that the degree of sequence variation corresponds well with the pharmacological efficacy of the peptides. This provides molecular evidence for a general functional importance of multiple paracopies.

Results

Genomics and peptide prediction

We mined the genomes of the 11 newly sequenced Drosophila species for homologs of the D. melanogaster peptide precursor genes Akh (CG1171), Ast (CG13633), Ast-C (CG149199), capa (CG15520), Ccap (CG4910), Crz (CG3302), Dh (CG8348), Dh31 (CG13094),ETH (CG18105), Fmrf (CG2346), hug (CG6371), IFa (CG33527), Leucokinin (CG13480), Mip (CG6456), Dms (CG6440), npf (CG10342), Nplp1 (CG3441), Pdf (CG6496), Proct (CG7105), Dsk (CG18090), sNPF (CG13968), and Dtk (CG14734). We then predicted the encoded neuropeptides; an overview of their numbers is given in Table 1. With the exception of the FMRFa-like peptides (see below), the analyzed genes code for the same number of neuropeptides in each species (43 in total, plus 10-17 FMRFa-like peptides). The translated coding sequences for the prepropeptides and predicted peptides are given as Additional data files 1 and 2.

Table 1 Peptide genes and encoded peptides

Mass spectrometric characterization

In Drosophila larvae, the main neurohemal organs that store and release peptide hormones are the ring gland, and the thoracic and abdominal perisympathetic organs. The epitracheal cells (Inka cells) are endocrine glands along the trachea. These tissues represent a rich source of neuropeptides: their peptidome contains about half of all known D. melanogaster neuropeptides [35, 36]. To biochemically assess whether the neuropeptides are expressed and processed as predicted, we directly profiled these neurohemal organs in D. sechellia, D. pseudoobscura, D. mojavensis and D. virilis. These species cover main phylogenetic lines within Drosophila. Obtained masses in the range of 850-2,500 Da were matched to the theoretical masses of predicted peptides (Table 2). This - and the observed tissue distribution - revealed that the peptidome of the investigated peptide release sites is identical in all species, at least in the mass range up to 2.5 kDa. In other words, all fruit flies appear to store the same set of (ortholog) peptides as D. melanogaster in the respective neurohemal release sites [35, 36].

Table 2 Amino acid sequences and mono-isotopic masses of detected peptides

Direct mass spectrometric profiling of the ring gland

The ring gland contained the adipokinetic hormone (AKH; pQLTFSPDWa), the AKH processing intermediate pQLTFSPDWGK, myosuppressin (MS), corazonin, corazonin3-11, the pyrokinins CAPA-PK2-15 and hugin (HUG)-PK, and the CAPA precursor peptide B (CPPB) (Figure 2 and Additional data file 3). As in D. melanogaster, the mass peak at 974.6 Da indicates the presence of SPSLRLRFa in D. sechellia, D. pseudoobscura, and D. virilis. The origin of SPSLRLRFa in these species is ambiguous, since it could represent short neuropeptide F (sNPF)-14-11 or its sequence-identical paralog sNPF-212-19. The finding of mass peaks at 974.6 and 992.6 in ring gland profiles of D. mojavensis - corresponding to sNPF4-11 and the aberrant D. mojavensis sNPF-212-19 SPSMRLRFa - indicates that, in fact, both sNPF-14-11 and sNPF-212-19 occur in the ring gland of Drosophila species. In D. sechellia, a mass peak corresponding to the full sNPF-1 was found in one preparation.

Figure 2
figure 2

Direct peptide profiling of the ring gland of different Drosophila species. (ai-di) Mass range 900-1,600 Da. The protonated mass of AKH is not visible, but the Na+ and K+ adducts are prominent. (aii-dii) Mass range 2,050-2,250 Da. Only one mass peak corresponding to CPPB is visible.

Direct mass spectrometric profiling of neurohemal release sites in the ventral ganglion

The neurohemal organs of the ventral ganglion are the thoracic and abdominal perisympathetic organs. In Drosophila and other flies, these organs persist during the larval stages but are subsequently reduced during pupal metamorphosis. In the adult fly, the innervating peptidergic neurites supply a neurohemal zone directly below the dorsal neural sheath [38, 39]. Since we did not succeed to specifically dissect the tiny larval perisympathetic organs, we directly profiled adult dorsal neural sheath preparations that were carefully cleaned of attached nervous tissue (n = 5-9 for each species). As in D. melanogaster [35], preparations from thoracic portions of the dorsal neural sheath contained the FMRFa-like peptides of the FMRF-prepropeptide (Figure 3ai-di). Preparations from abdominal portions contained the CAPA peptides CAPA-PVK-1 and -2, CAPA-PK and CPPB (Figure 3aii-dii). Occasionally, mass peaks corresponding to CAPA peptides were found in thoracic preparations, and FMRFa-like peptides in abdominal preparations. This corresponds to the variable extent of overlap of the more posterior CAPA neuron projections with more anterior FMRFa-like peptide neuron projections. Concomitantly, mass spectra from intermediate portions of the dorsal neural sheath consistently showed both CAPA and FMRFa-like peptide peaks.

Figure 3
figure 3

Direct peptide profiling of the dorsal neural sheath of different Drosophila species. (ai-di) Thoracic portion containing FMRFa-like peptides. Note that peak intensity corresponds with isocopy number in the FMRFa prepropeptide. Small peaks corresponding to CAPA peptides from overlapping Va neurites are visible in D. virilis and D. mojavensis. (aii-dii) Abdominal portion containing CAPA peptides. Peaks corresponding to drosokinin and IPNa in D. pseudoobscura and D. sechellia represent contaminations with ganglionic neurites or the segmental nerve.

In each species, the masses of all predicted FMRFa-like peptides of the FMRF-prepropeptide could be detected (Table 2) with the exception of FMRFa-1. This peptide invariantly has the carboxy-terminal sequence FMHFa in the investigated species, and thereby lacks the easily protonated Arg that makes FMRFa-1 difficult to detect in peptide mixtures by the MALDI process [35, 40]. In many FMRFa-like peptide-containing spectra, a mass peak around 2 kDa was prominent. In each species, this mass peak matched the theoretical mass of the respective extended form of FMRFa-5 (FMRFa-5ext), which would result from prohormone cleavage of FMRF-4 and FMRF-6 without internal cleavage of the single Arg cleavage site of FMRF-5 (Additional data file 4). An extended form of FMRFa-5 had not been described from D. melanogaster. We therefore reviewed our old data from D. melanogaster larvae [36]. In many spectra, we found a distinct mass peak at 2,003.0 Da, which matches the theoretical mass of FMRF-5ext of D. melanogaster but was previously overlooked. The consistent occurrence of prominent mass peaks corresponding to the theoretical mass of FMRF-5ext in the different Drosophila species is unlikely to have occurred by chance, and therefore indicates a new processing product of the Drosophila FMRFa precursor. It is unclear whether FMRF-5ext is released as a peptide hormone, or only represents a processing intermediate.

Besides CAPA- and FMRFa-like peptides, mass peaks corresponding to leucokinin and IPNa were occasionally detected in dorsal neural sheath preparations (Figure 3). Leucokinin and IPNa are dominant peptides in ventral ganglion preparations [35] and likely represent a contamination of the dorsal neural sheath by adhering peptidergic neurites.

Direct mass spectrometric profiling of the peritracheal cells

The larval peritracheal cells are located at stereotypic locations near the primary branchings of trachea from the main trunk [41]. As in D. melanogaster, spectra obtained with the laser beam directed at these branching sites consistently showed mass peaks corresponding to ecdysis-triggering hormone (ETH)-1 and -2 in all species (Figure 4). The mass of ETH-1 was detected in 8 out of 15 preparations in D. virilis, in 9 out of 11 preparations in D. mojavensis, in 12 out of 13 preparations in D. pseudoobscura, and 4 out of 6 preparations in D. sechellia. Equivalent numbers for ETH-2 were 11/15, 6/11, 7/13 and 6/6.

Figure 4
figure 4

Direct peptide profiling of tracheal preparations containing the peritracheal cells of different Drosophila species. Peaks corresponding to the [M+H]+ or [M+Na]+ adducts of the two ETHs are visible besides the typical and possibly non-peptidergic tracheal peaks [22].

Peptide copy numbers

Alignment of the prepropeptide sequences showed that the peptide families of each Drosophila species consist of an identical set and number of ortholog neuropeptide copies, with the exception of FMRFa-like peptides (Table 1; Additional data files 1 and 2). For example, in all species the crustacean cardioactive peptide (CCAP) precursor contains one CCAP, and the allatostatin A (ASTa) precursor contains 4 ASTa peptides. The FMRFa precursor, however, encodes 10 FMRFa-like peptides in D. mojavensis and D. virilis, 11 FMRFa-like peptides in D. ananassae, 12 FMRFa-like peptides in D. willistoni, 14 FMRFa-like peptides in D. erecta, 17 FMRFa-like peptides in D. grimshawi, and 13 FMRFa-like peptides in all other species. The fmrf gene contains 2 exons, of which exon II codes for the whole FMRFa prepropeptide. The differences in peptide-coding sequences can thus not be explained by exon duplication. Higher numbers of tandem repeats exist for FMRFa-2 (DPKQDMRFa; for example, 5 copies in D. melanogaster, 7 copies in D. grimshawi) in all species but D. mojavensis and D. virilis. This may suggest that mispairing of template versus replicating nucleotide sequences coding for this peptide has resulted in insertions/deletions during Drosophila evolution and has caused the high number of FMRFa-like peptide copies.

Two prepropeptides contain neuropeptides that are usually not grouped into the same peptide family: the CAPA prepropeptide contains two periviscerokinins and one pyrokinin, and the neuropeptide-like precursor (NPLP)1 prepropeptide contains one MTYamide, one IPNamide and one non-amidated peptide. The CAPA pyrokinin and the NPLP1 peptides have therefore been treated as single copy peptides (but see Discussion).

Peptide-coding sequences are more conserved than spacer sequences

If the neuropeptide sequences are subjected to stabilizing selection due to their signaling function, it is reasonable to assume that the peptide-coding parts of the prepropeptides are more conserved than the spacers (the parts separating the bioactive peptides), which by existing evidence do not act as signaling molecules in insects. In other words, the sequence similarity between ortholog neuropeptide parts of the prepropeptides is likely to be higher than the sequence similarity of ortholog spacer parts. To test this hypothesis, it is not sufficient to simply calculate amino acid identities, since substitutions of amino acids do not occur randomly but are correlated with their physico-chemical characteristics [42]. We thus calculated the overall average amino acid distance Dso for each set of orthologs (Figure 5) based on the Jones-Thornton-Taylor (JTT) matrix [43] as a more appropriate measure of sequence variation (see Material and methods). The raw values are listed in Additional data file 5. The median Dso between peptide orthologs was 0.041, and thus significantly lower than the calculated 0.408 for the spacers (Figure 5; Mann-Whitney, two-tailed p < 0.0001, U = 211.5), although the sequence of several spacers was quite conserved (for example, in the CCAP or CAPA prepropeptides). In contrast to the peptides (p < 0.01), the spacer distances followed a Poisson distribution.

Figure 5
figure 5

Plot of the average distance between orthocopies and ortholog spacers. Each data point represents the average amino acid distance Dso between orthocopies or ortholog spacer regions. With the exception of FMRFa-7, the peptide orthocopy distances have values below 0.3 and do not follow a Poisson distribution as is seen for the spacers.

A closer look at the data (Additional data file 5) shows that high Dso values only occur in multiple copy peptide families. For example, neuropeptide F (NPF) and MTYa show the highest Dso for single copy families (0.134 and 0.093, respectively). The respective maximum values for multiple copy families are 0.748 for FMRFa-7, 0.601 for sulfakinin (SK)-0, 0.267 for ETH-2, 0.208 for FMRFa-7, and 0.205 for CAPA-PVK-2. Yet, orthocopy sets without sequence variation or with low Dso values occur not only in single copy, but also in each multiple copy peptide family: CAPA-PVK-1 (0.042), ETH-1 (0.025), SK1- (0), ASTa-3 and -4 (0), sNPF-1 (0), myoinhibiting peptide (MIP)-3 and -4 (0), Drosophila tachykinin (DTK)-3 (0.02) and FMRFa-2 and -6 (0).

The average distance between all peptides in a family is higher for families with multiple paracopies

To test for differences in the sequence variability between single and multiple copy peptide families, we computed the average amino acid distance Daf for each amino acid position between all paracopies within a peptide family (Figure 6a, d) and then calculated the mean (Figure 6b). For single copy peptides, we calculated the corresponding average amino acid distance Dao for the respective orthologs (Figure 6c). The results in Figure 6 show that the mean Daf between paracopies of multiple copy peptide families is typically higher than the Dao observed between the single copy peptides. Due to a large standard variation, these differences are only significant for amino acid positions 5 and 7 from the carboxyl terminus (paired t-test, p < 0.05). This reflects the spread of sequence variation in multiple copy peptide families. For most amino acid positions there are families that show no variation, and, at the same time, families with considerable sequence variation. The high mean Dap at position 1 from the carboxyl terminus mostly originates from the sNPFs, which end either RFa (sNPF-1 and -2) or RWa (sNPF-3 and -4). There is no clear tendency that the sequence variation increases from the carboxyl to the amino terminus; a correlation between Daf and copy number is not discernible (Figure 6d).

Figure 6
figure 6

Plot of the average distance between all paracopies in a family. Each data point represents the average amino acid distance Daf between all paracopies of a peptide family for each amino acid position throughout the species as outlined in (a). (b-d) The mean ± standard deviation of the data (b) for single copy peptides (c) and multiple copy peptide families (d). Paracopy number is color-coded in (d): black, 2; blue, 3; red, 4; green, 5; purple, 6; and brown, 10-17. The asterisks indicate a significant difference between multiple and single copy peptides.

Orthologs of single and multiple copy peptide families are equally sequence-conserved

The distance Daf contains both the sequence variation between individual orthologs (inter-ortholog variation) as well as between individual paracopies (inter-paracopy variation; Figure 1). To test the contribution of the inter-ortholog variation to Daf, we calculated the average amino acid distance Dao for each amino acid position for each set of orthocopies individually (Figure 7). A comparison of Figures 7c and 6b shows that the mean Dao for the ten carboxy-terminal amino acids is considerably smaller than the mean Daf and not significantly different between single and multiple copy peptides. This region likely contains the active core of the peptides, which typically consists of the last five to seven carboxy-terminal amino acids and the amidation signal (for example, [44–47]). Somewhat higher Dao values occurred for more amino-terminal amino acids. This shows that the orthologs are strongly sequence-conserved throughout Drosophila, irrespective of whether they belong to a single or multiple copy peptide family - with the exception of FMRFa-7 and SK-0 (see below).

Figure 7
figure 7

Plot of the average distance between orthocopies for each amino acid position. Each data point represents the average amino acid distance Dao between the orthocopies for each amino acid position throughout the species as outlined in (a). (b) The Dao for multiple copy peptide families. (c) The mean Dao ± standard deviation for single (black) and multiple (red) copy peptide families (see Figure 6c). The different shapes code for paracopy numbers: filled square, 1; filled triangle, 2; inverted filled triangle, 3; filled diamond, 4; filled circle, 5; open square, 6; open triangle, 7; open triangle, 8; open diamond, 9; open circle, 10; cross, 11; plus sign, 12; asterisk, 13.

Sequence variation mostly originates from sequence variation between paracopies

We next calculated the average net amino acid distances (Figure 1) between paracopies Danp for multiple copy peptide families; results are shown in Figure 8. The mean Danp was higher than the mean Dao of single (Figure 8b) or multiple copy (compare Figure 7c) peptides throughout amino acid position 1-11. This was again only significant for positions 5 and 7 due to the high standard variations (paired t-test, p < 0.05). As for Daf, the high variation of Danp reflects the spread of the degree in sequence variation between multiple copy peptide families: given positions were variable in some peptide families, but imvariantly occupied by the same amino acid in others. The high Danp of 1.69 at position 1 from the carboxyl terminus is again caused by the carboxy-terminal difference RFa and RWa between the sNPFs. When omitting the RWamides sNPF-3 and -4 - which could not be biochemically detected yet - this value drops to 0.42. With this value, it seems that the amino acids at positions 1-3 and 8 from the carboxyl terminus are the most conserved amino acids between the paracopies of each multiple copy peptide family.

Figure 8
figure 8

Plot of the net distance between paracopies for each amino acid position. (a) Each data point represents the average net amino acid distance Danp ± standard deviation between the paracopies for each amino acid position throughout the species. (b) The mean ± standard deviation of the data for multiple copy peptides compared to the mean Dao ± standard deviation of single copy peptides (see Figure 6c). The asterisks indicate a significant difference between multiple and single copy peptides.

Sequence variation is not related to receptor number

The majority of G protein-coupled peptide receptors of D. melanogaster have been deorphanized, with some still uncharacterized to date [48]. From the obtainable literature [48–51], we compiled the number of characterized D. melanogaster G protein-coupled receptors per peptide family. Although these numbers may be subjected to future changes, to date there are either only one or two receptors known for the paracopies of each peptide family. The occurrence of two receptors for some peptide family opens the possibility for receptor-ligand coevolution and subsequent sub- and neofunctionalization of paracopies. To test for this, we plotted the Danp between multiple copy peptide families with one known receptor against those with two known receptors. Although there are differences (Figure 9), they are neither statistically significant nor do they follow an obvious pattern. This result speaks against a sub- or neofunctionalization of paracopies during the evolution of Drosophila.

Figure 9
figure 9

Plot of the net distance between paracopies for each amino acid position. Each data point represents the average net amino acid distance Danp ± standard deviation between the paracopies for each amino acid position throughout the species for peptide families with one (open black squares) or two (closed red triangles) known receptors.

Discussion

We datamined the 11 new Drosophila genomes for homologs of the 22 described prepropeptide genes of D. melanogaster encoding neuropeptides up to a length of 50 amino acids [52, 53]. From these data, we were able to predict 53-60 neuropeptides for each species. These peptides are known or are likely to signal via G protein-coupled receptors [48]. Larger protein hormones (>50 amino acids) have not been included, because they are expected to have a smaller proportion of residues that are important for their pharmacological efficacy (see, for example, [54]), which makes it difficult to directly compare their sequence variability to that of the smaller neuropeptides.

Accuracy of peptide predictions

The obtained mass fingerprints of the neurohemal organs and endocrine cells were identical: in each species, the obtained masses corresponded to the respective orthologs in D. melanogaster neurohemal organs or peritracheal cells [35, 36]. Vice versa, for each peptide characterized in the neurohemal organs or peritracheal cells of D. melanogaster, there was a mass present that corresponded to the respective ortholog in the other species. Similar to the use of peptide fingerprints in proteomics, the exactly matching tissue fingerprints chemically identify the underlying peptides and precursor products with high probability. All fingerprint masses matched the respective theoretical masses calculated for the in silico predicted peptides. In conclusion, the mass spectrometric profiling supports our in silico prediction of the neuropeptidome.

The peptidome is evolutionarily conserved throughout the genus Drosophila

The finding of identical peptide hormone complements in the mass range of 800-2,500 Da in main Drosophila phylogenetic lineages suggests that the peptidome of the major neurohemal organs and the peritracheal cells has been evolutionary stable for at least 40-60 million years since the divergence of the Drosophila species from their last common ancestor. Obviously, all Drosophila species share the same peptidergic hormonal communication possibilities. Even more, our genomic comparisons suggest that the whole peptidome is highly conserved throughout the genus Drosophila. We observed no loss of peptide precursors, or individual peptides as suggested to have occurred, for example, between flies and mosquitoes [55]. Thus, the number of peptide copies in each precursor was identical throughout the species with exception for the FMRFamides, which most likely duplicated by unequal recombination. These recombination events must have occurred independently from each other, since multiple repeats of FMRFa-2 coding sequences are present both in the Sophophora subgroup and the Hawaiian species of the Drosophila subgroup (D. grimshawi), but are lacking in the other Drosophila subgroup species D. virilis and D. mojavensis. Unequal recombination is also the likely mechanism behind the duplication of most other multiple copy peptides, but for them recombination must have occurred prior to Drosophila speciation.

The high conservancy of the peptidome is remarkable and unexpected, since drosophilid flies have undergone several radiations and have adapted to a variety of environments with, for example, a very different supply of water, such as sea shores, forests and deserts [34, 56]. In contrast, the genome of the tenebrionid beetle Tribolium castaneum shows a gene expansion for putative diuretic peptides [57]. This has been interpreted as an adaptation to dry conditions in tenebrionids, a beetle family that thrives in deserts and other very dry places. It is, however, unclear whether this is a special tenebrionid or a common beetle feature. At least for fruit flies, our data show that the adaptation to different environments is not paralleled by changes in the number or increased sequence variability of diuretic hormones or other neuropeptides.

Neuropeptide sequences are subjected to stabilizing selection

In our analysis, spacer sequences showed significantly higher amino acid distances than peptide sequences. This suggests that Drosophila neuropeptides are subjected to stabilizing selection or evolutionary constraint to a much larger extent than spacer sequences. This is further supported by the non-random distribution of peptide distances not observed for spacers. This finding may not be unexpected, but is shown here for the first time on a neuropeptidome level.

In Drosophila, there is a higher proportion of highly constrained codons in essential genes than in any other dispensability class [58]. As hypothesized at the outset, stabilizing selection and the resulting sequence conservation may thus indicate functional importance of neuropeptides, signaling molecules for which single amino acid exchanges can result in drastically altered receptor efficacy, binding or effect (for example, [28, 59]). If this hypothesis is correct, then the observed low inter-orthocopy distances (Dao, Dso) indicate that the multiple peptide copies are functionally important and not individually dispensable.

The observed higher amino acid distances that reflect a considerable sequence variation for the spacers do not allow us to conclude that structural features of the spacers are unimportant for proper peptide processing and packaging into secretory vesicles. They speak, however, against a general signaling function of the spacer regions ('associated peptides') at the receptor binding site, where single amino acid changes can already result in altered efficacy, effect or specificity (for example, [28, 59]). Nevertheless, this conclusion needs proper physiological testing. Several spacer regions are quite conserved throughout the Drosophila species (for example, in the CAPA and CCAP precursor), and a FMRFa-spacer-derived peptide has been shown to modulate the activity of FMRFa at the receptor in Lymnea [60].

Peptide copies are unlikely to have undergone a phase of neutral mutation

The comparably high Danp distances show that there is sequence-variation between paracopies (inter-paracopy variation). Assuming that paracopies at some point have arisen from a common ancestor, we have hypothesized at the outset that newly duplicated paracopies can escape selection pressure and may be allowed to drift neutrally. However, the small Dao distances between orthocopies (inter-orthocopy variation) do not support this scenario. A significant difference in sequence variation between the individual sets of orthocopies was not observed between single and multiple copy peptide families, and inter-orthocopy distances were small compared to the distances found between spacer regions. This suggests that: the inter-paracopy variation originates from a time before divergence of the Drosophila taxa; and paracopies have never fully escaped selection pressure and have never experienced a phase of neutral mutation. Hence, the classic theory for duplicated genes may only apply in a limited sense for paracopies.

At the outset, we reasoned that peptide copies following the 'more-of-the-same' concept will show low sequence variation between paracopies. For paracopies with differential activities, we expected at the same time an increased sequence variation between paracopies. Since the inter-paracopy distances Danp were similar for all multiple copy peptide families and did not correlate with receptor number, it is not possible to draw conclusions from our data in this respect. For Drosophila, there are also no data regarding different half-lives during peptide degradation, or induction of ligand-selective receptor conformation and activity [61] as has been demonstrated for locust and cockroach AKHs [27, 28].

The CAPA and NPLP1 prepropeptides contain neuropeptides that are usually not grouped into the same peptide family. We have therefore treated the CAPA pyrokinin and the NPLP1 peptides as single copy peptides. However, some sequence similarities can be found between CAPA pyrokinins and periviscerokinins, and between the amino-terminal stretches of the NPLP1 peptides [62]. It has also been shown that the CAPA pyrokinin specifically activates a G protein-coupled receptor (CG9918) that is evolutionarily related to the other Drosophila pyrokinin receptors, but is not activated by CAPA periviscerokinins and the HUG-pyrokinin at physiological concentrations [63]; data on NPLP1 peptide receptors are not yet available. Thus, it is possible that at least the CAPA peptides are, in fact, an example of multiple copies that have sub- or neofunctionalized by acquiring sequence variation: the CAPA prepropeptide appears to date back at least to the origin of insects, since it contains a few periviscerokinins plus one highly sequence-conserved pyrokinin in all insect taxa investigated so far [64]. If this is the case, this sub- or neofunctionalization must have occurred a long time before the radiation of Drosophila. While this justifies the classification of at least the CAPA pyrokinin as a single copy peptide in this study, it emphasizes the need for further comparisons similar to that reported here for Drosophila on larger phylogenetic units spanning longer evolutionary time frames. Such studies will soon become possible with the increasing number of fully sequenced insect genomes.

The calculated distances correlate with pharmacological efficacy

Although the inter-orthocopy distance Dso was, in general, very low throughout the peptides, we observed differences in Dso within the CAPA-PVK, ETH, FMRFamide and tachykinin peptide families. Can these differences be correlated to differential activities? A comparison of the calculated Dso values with the available pharmacological data shows that there is at least a correlation to the reported efficacies: the paracopies with lower amino acid distance typically are the ones with higher receptor or pharmacological activity.

For the ETHs, the EC50 of the paracopy with the lowest sequence variation (ETH-1) is nine times more potent in heterologous receptor assays than the more sequence-variable ETH-2 [65, 66]. Two groups have characterized the FMRFa receptor of D. melanogaster in heterologous expression systems. Both found that FMRFa-6 (PDNFMRFa) has the highest potency to activate the FMRFa receptor, whereas FMRFa-7 has no activity at all [67, 68]. Meeusen and colleagues [68] report similar EC50 values for FMRFa-2 to -5. All were only slightly less potent than FMRFa-6. Cazzamali and Grimmelikhuijzen [67] found the following order: FMRFa-6 > FMRFa-2 > FMRFa-3,-5,-8 > FMRFa-1 > FMRFa-4. This pharmacological ranking correlates well with the calculated Dso distances (in brackets): FMRFa-6 = FMRFa-2 (0) < FMRFa-5 (0.04) < FMRFa-8 (0.064) < FMRFa-4 (0.11) < FMRFa-1 (0.127) < FMRFa-3 (0.172) << FMRFa-7 (0.748).

For the DTKs, our data predict the following ranking of pharmacological activity: DTK-3 > DTK-1 > DTK-6 > DTK4 = DTK-5 > DTK-2. This again corresponds quite well with the efficacy of DTKs on DTKR - one of the two DTK receptors known - in HEK-293 cells [69]: DTK-1 > DTK-3 = DTK-6 > DTK-4 > DTK-5 > DTK-2. For CAPA-PVKs, the copy with the lower sequence variation (CAPA-PVK-1) is more effective in inducing calcium responses and fluid transport in the Malpighian tubules [70] and about 1.5-times more potent in receptor assays than the more sequence variable CAPA-PVK-2 [65, 71]. In other words, the more potent peptides were typically the more sequence-conserved. This might extend well beyond Drosophila. For example, CAPA-PVK-2 orthologs are much more sequence-variable in their carboxyl terminus than CAPA-PVK-1 orthologs not only in Drosophila, but also in other flies [72, 73]. At the same time, the housefly Musca domestica CAPA-PVK-2 shows a ten-times diminished efficacy in fluid secretion assays on housefly Malpighian tubules compared to M. domestica CAPA-PVK-1 [45]. The degree of sequence variation appears not to be linked with peptide position along the precursor.

In contrast to the CAPA-PVKs, ETHs, FMRFamides and DTKs, the SK-1 and -2, MIP and ASTa copies all showed a consistently low inter-orthocopy distance. The pharmacological profiles of SK and MIP copies on their respective receptors have not been characterized so far, but such data exist for the two Drosophila ASTa receptors (DARs) expressed in Chinese hamster ovary (CHO) cells [74, 75]. The data suggest that DAR-1 has a lower sensitivity for ASTa-4 than for ASTa-1 to -3, whereas DAR-2 is more sensitive to ASTa-3 to -4. It has, however, to be kept in mind that not all peptide receptors have been deorphanized to date, and that signaling properties and specificities of receptors may be changed by modifying proteins such as RAMP or RGS in native cells. It is also possible that ligand-selective receptor conformations may exist [61], and the ligand properties and activated intracellular pathways in vivo may be different to those in heterologous expression systems. This may explain why - in contrast to the data from heterologous expression systems - all FMRFamides had a similar dose-response effect at the neuromuscular junction [22]. Only FMRFa-7 (SAPQDFVRSa) was inactive in all systems. Clearly, further data, especially from bioassays, will be needed to confirm the observed correlations between sequence variation and efficacy.

The evolution of neuropeptides and their receptors is linked, and neuropeptide receptors are under evolutionary pressure to maintain a high affinity to the authentic ligands [9, 59]. The finding that the more sequence-variable neuropeptides typically had a lower pharmacological efficacy does not speak for the occurrence of fast adaptive structural changes of G protein-coupled receptors to maintain a high ligand affinity to sequence-variable peptides during the evolution of Drosophila. Does that mean that peptide copies with high amino acid distances are functionally unimportant? The by far highest distances were found for FMRFa-7 (0.748) and SK-0 (0.601). The carboxyl terminus of FMRFa-7 (FVRSa) is highly deviated and is the likely cause for its lack of receptor activation and bioactivity [22, 67] (and see above). The high Dso value of FMRFa-7 is around the mean value found for spacer regions, which suggests that FMRFa-7 has escaped selection pressure to a considerable amount. The high Dso value for SK-0 correlates with its inactivity at the DSK-R1 receptor [76] and its lack of bioactivity at physiological concentrations below 1 μM [77]. Unlike SK-1 and -2, SK-0 has, furthermore, not been found biochemically so far [35, 37]. Hugin-gamma, a predicted but obviously not processed peptide [78] that seems to be missing from the genome of D. persimilis [55], shows a Dso of 0.259. Nevertheless, the synthetic D. melanogaster HUG-gamma is still able to activate the receptor [65, 79]. We propose from this as a rough estimate that the peptides with an amino acid distance above 0.6 are nonfunctionalized. Distances below 0.3 and the non-Gaussian distribution of all other peptide copies suggest that they are under stabilizing selection that prevents nonfunctionalization by random or deleterious mutations.

Conclusion

Taken together, our data provide evidence that the peptidome and the neuropeptide hormone complement has been conserved during the evolution of Drosophila, and shows that multiple peptide copies with biological activity are under stabilizing selection. Sequence conservation largely correlates with pharmacological activity. While all this suggest that multiple peptide copies are functionally important, it remains unclear why paracopies are under stabilizing selection.

It has to be stressed that our data are based on only a relatively small number of data points. This was unavoidable by the simple fact that further multiple copy neuropeptide families in Drosophila have not been identified. Consequently, our conclusions will need further validation and we hope that our work will provoke further studies on new data from the rapidly increasing number of genome projects. Our study emphasizes the value of these genome projects, and stresses the need for more comprehensive structure-activity studies and pharmacological characterization of peptides both in receptor and bioassays.

Materials and methods

Flies

D. virilis was obtained from a colony in Ulm. Genome library strains of D. mojavensis,D. pseudoobscura, and D. sechellia were obtained from the Tucson Drosophila Stock Center. Flies were kept at standard conditions - a light:day cycle of 12 h:12 h and either 18°C or 25°C. D. virilis and D. sechellia were raised on standard cornmeal medium, D. mojavensis and D. pseudoobscura on standard banana-Opuntia medium.

Database searches

Peptide precursor genes were identified by tblastn homology searches against the respective D. melanogaster coding sequences using the PAM30 matrix of the Drosophila species BLAST site [80]. The coding sequences were identified and translated with GENSCAN [81] and compared with the GLEAN-predicted sequences in the databank. Amino acid sequences were aligned with the ClustalW algorithm implemented in MEGA3.1 [82] and plotted using GeneDoc [83]. Signal peptides were predicted by SignalP 3.0 [84]. Mono-isotopic masses of the predicted bioactive neuropeptides were calculated with Data Explorer 4.0 software (Applied Biosystems, Darmstadt, Germany).

Peptide prediction

We predicted the processed bioactive peptides based on cleavage site consensus sequences [85] and comparison with the chemically characterized processing products from D. melanogaster [35–37]. Mono-isotopic masses were calculated for all peptides and listed for each species. Peptide designations were inferred from prepropeptide alignment with the ortholog D. melanogaster sequence, for example, the myoinhibiting peptide encoded on the Mip orthologs that aligned with D. melanogaster MIP-3 was also named MIP-3. This allows easy identification and reference to ortholog peptides. In the fmrf-precursor of D. melanogaster, several paracopies are sequence-identical and named either FMRFa-2 or -3. These paracopies and their orthologs were designated according to their position on the gene as FMRFa-2', FMRFa-2", and so on.

Calculations of sequence variation/amino acid distances

The parts of the prepropeptides between the signal peptide and the bioactive peptide sequences flanked by the mono- or dibasic cleavage sites were assigned as spacers. For distance calculations of peptides, sequences were aligned from their carboxyl terminus. Gaps that occurred due to variable peptide copy length were deleted pairwise. Spacers were aligned by the ClustalW algorithm prior to distance calculation. Average distances were calculated as absolute values in MEGA3.1 for pairwise comparisons as outlined in Figure 1 by iterative procedures under a maximum likelihood formulation using the JTT matrix [43, 82]. The JTT matrix was calculated from data of the Swiss-Prot protein sequence database and gives a measure of the probability that a given amino acid i is being replaced by residue j per occurrence of j [43]. Variable mutation rates among sites were assumed. Since the peptide sequences are too short to reliably estimate gamma parameters, we adopted a gamma distance with α = 2.4, which is very close to the true distance for sequence divergence under the JTT model [86]. Data were processed and plotted using Microsoft Excel and GraphPad Prism 4.0 (GraphPad Software, San Diego, CA, USA).

Sample preparation

The central nervous system was dissected free from all surrounding tissue in standard Drosophila saline. Ring glands of L3 larvae (selected after size: D. virilis >3 mm; D. mojavensis, D. sechellia >2 mm; D. pseudoobscura >2.5 mm) were punched out using pulled glass capillaries and spotted directly onto the MALDI target and left to dry. For isolation of the thoracic and abdominal neurohemal sites, the thoracic or abdominal part of the ventral ganglion of adults was cut out and the lateral parts were removed using fine scissors. The dorsal neural sheath was then isolated and freed from cells using tungsten micro-needles. The isolated sheaths were transferred to the MALDI target using pulled glass capillaries and left to dry. This method results in clean spectra from the neurohemal endings [35, 36]. For direct profiling of the peritracheal cells, the main branches of the trachea from L3 larvae were dissected free from other tissue and transferred directly onto the MALDI target using fine insect needles. The peritracheal cells were targeted by directing the laser beam to the obtuse angle between the main trachea and the diverging first order trachea.

To remove salts, a small droplet of ice-cold water was added onto the dried tissues, and aspirated off after about 1 s. Small nanoliter volumes of matrix (saturated solution of re-crystallized α-cyano-4-hydroxycinnamic acid in 30% MeOH/30% EtOH/0.1% trifluoroacetic acid for neurohemal organs, or 60% acetonitrile/0.1% trifluoroacetic acid for peritracheal cells were added to the samples using a manual oocyte injector (Drummond Scientific, Broomall, PA, USA). In prior tests MeOH/EtOH/H2O (30:30:40) resulted in improved mass spectra compared to 60% MeOH as previously described for D. melanogaster [35, 36].

MALDI-TOF mass spectrometry

MALDI-TOF mass spectra were acquired in positive ion mode on a Voyager DE RP mass spectrometer (Applied Biosystems, Darmstadt, Germany) equipped with a pulsed nitrogen laser emitting at 337 nm. Samples were analyzed in reflectron mode using a delayed extraction time of 400 nsec and an accelerating voltage of 20 kV. To suppress matrix signals, the low mass gate was set to 850 Da. Laser power was adjusted to provide optimal signal-to-noise ratios. Data were analyzed using Data Explorer 4.0 software (Applied Biosystems), with a mass tolerance of 0.5 Da.

Additional data files

The following additional data are available. Additional data file 1 is a Fasta list containing the prepropeptide sequences as identified by BLAST searches in the 12 Drosophila genomes. Additional data file 2 is a Fasta list of the peptide sequences predicted from the identified prepropeptides. Additional data file 3 is a figure showing the frequency of occurrence of peptides in mass spectrometric profiles of the ring gland. Additional data file 4 is a figure containing mass spectrograms on the processing of FMRFa-5ext from the FMRFa-prepropeptide. Additional data file 5 is a table listing the overall average amino acid distances Dso.

Abbreviations

AKH:

adipokinetic hormone

AST:

allatostatin

CCAP:

crustacean cardioactive peptide

CPPB:

CAPA precursor peptide B

DAR:

Drosophila allatostatin receptor

DTK:

Drosophila tachykinin

ETH:

ecdysis-triggering hormone

HUG:

hugin

JTT:

Jones-Thornton-Taylor

MALDI-TOF:

matrix-assisted laser desorption ionization-time of flight

MIP:

myoinhibiting peptide

MS:

myosuppressin

NPF:

neuropeptide F

NPLP:

neuropeptide-like precursor

SK:

sulfakinin

sNPF:

short neuropeptide F.

References

  1. Conlon JM, Larhammar D: The evolution of neuroendocrine peptides. Gen Comp Endocrinol. 2005, 142: 53-59.

    Article  PubMed  CAS  Google Scholar 

  2. Niall HD: The evolution of peptide hormones. Annu Rev Physiol. 1982, 44: 615-624.

    Article  PubMed  CAS  Google Scholar 

  3. Danielson PB, Dores RM: Molecular evolution of the opioid/orphanin gene family. Gen Comp Endocrinol. 1999, 113: 169-186.

    Article  PubMed  CAS  Google Scholar 

  4. Holmgren S, Jensen J: Evolution of vertebrate neuropeptides. Brain Res Bull. 2001, 55: 723-735.

    Article  PubMed  CAS  Google Scholar 

  5. Hoyle CHV: Neuropeptide families: evolutionary perspectives. Regul Pept. 1998, 73: 1-33.

    Article  PubMed  CAS  Google Scholar 

  6. Ohno S: Evolution by Gene Duplication. 1970, Berlin: Springer Verlag

    Book  Google Scholar 

  7. Kimura M: Recent development of the neutral theory viewed from the Wrightian tradition of theoretical population genetics. Proc Natl Acad Sci USA. 1991, 88: 5969-5973.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  8. De Souza FSJ, Bumaschny VF, Low MJ, Rubinstein M: Subfunctionalization of expression and peptide domains following the ancient duplication of the proopiomelanocortin gene in teleost fishes. Mol Biol Evol. 2005, 22: 2417-2427.

    Article  PubMed  CAS  Google Scholar 

  9. Darlison MG, Richter D: Multiple genes for neuropeptides and their receptors: co-evolution and physiology. Trends Neurosci. 1999, 22: 81-88.

    Article  PubMed  CAS  Google Scholar 

  10. Taylor JS, Raes J: Small-scale gene duplications. The Evolution of the Genome. Edited by: Gregory TR. 2005, Amsterdam: Elsevier, 289-327.

    Chapter  Google Scholar 

  11. Thornton J: New genes, new functions: gene family evolution and phylogenetics. Evolutionary Genetics. Edited by: Fox CW, Wolf JB. 2006, New York: Oxford University Press, 157-172.

    Google Scholar 

  12. Kastin AJ: Handbook of Biologically Active Peptides. 2006, Amsterdam: Elsevier

    Google Scholar 

  13. Leviev I, Grimmelikhuijzen CJP: Molecular cloning of a preprohormone from sea anemones containing numerous copies of a metamorphosis-inducing neuropeptide: a likely role for dipeptidyl aminopeptidase in neuropeptide precursor processing. Proc Natl Acad Sci USA. 1995, 92: 11647-11651.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  14. Predel R, Neupert S, Wicher D, Gundel M, Roth S, Derst C: Unique accumulation of neuropeptides in an insect: FMRFamide-related peptides in the cockroach, Periplaneta americana. Eur J Neurosci. 2004, 20: 1499-1513.

    Article  PubMed  CAS  Google Scholar 

  15. Yin GL, Yang JS, Cao JX, Yang WJ: Molecular cloning and characterization of FGLamide allatostatin gene from the prawn, Macrobrachium rosenbergii. Peptides. 2006, 27: 1241-1250.

    Article  PubMed  CAS  Google Scholar 

  16. Husson SJ, Mertens I, Janssen T, Lindemans M, Schoofs L: Neuropeptidergic signaling in the nematode Caenorhabditis elegans. Prog Neurobiol. 2007, 82: 33-55.

    Article  PubMed  CAS  Google Scholar 

  17. Furukawa Y, Makamaru K, Wakayama H, Fujisawa Y, Minakata H, Ohta S, Morishita F, Matsushima O, Li L, Romanova E, Sweedler JV, Park JH, Romero A, Cropper EC, Dembrow NC, Jing J, Weiss KR, Vilim FS: The enterins: a novel family of neuropeptides isolated from the enteric nervous system and CNS of Aplysia. J Neurosci. 2001, 21: 8247-8261.

    PubMed  CAS  Google Scholar 

  18. Lynch M, Katju V: The altered evolutionary trajectories of gene duplicates. Trends Genet. 2004, 20: 544-549.

    Article  PubMed  CAS  Google Scholar 

  19. Lange AB, Bendena WG, Tobe SS: The effect of thirteen Dip-allatostatins on myogenic and induced contractions of the cockroach (Diploptera punctata) hindgut. J Insect Physiol. 1995, 41: 581-588.

    Article  CAS  Google Scholar 

  20. Predel R, Eckert M: Neurosecretion: peptidergic systems in insects. Naturwissenschaften. 2000, 87: 343-350.

    Article  PubMed  CAS  Google Scholar 

  21. Brezina V, Weiss KR: Analyzing the functional consequences of transmitter complexity. Trends Neurosci. 1997, 20: 538-543.

    Article  PubMed  CAS  Google Scholar 

  22. Hewes RS, Snowdeal EC, Saitoe M, Taghert PH: Functional redundancy of FMRFamide-related peptides at the Drosophila larval neuromuscular junction. J Neurosci. 1998, 18: 7138-7151.

    PubMed  CAS  Google Scholar 

  23. Brezina V, Bank B, Cropper EC, Rosen S, Vilim FS, Kupfermann I, Weiss KR: Nine members of the myomodulin family of peptide cotransmitters at the B16-ARC neuromuscular junction of Aplysia. J Neurophysiol. 1995, 74: 54-72.

    PubMed  CAS  Google Scholar 

  24. Duve H, Elia AJ, Orchard I, Johnsen AH, Thorpe A: The effects of CalliFMRFamides and other FMRFamide-related neuropeptides on the activity of the heart of the blowfly Calliphora vomitoria. J Insect Physiol. 1993, 39: 31-40.

    Article  CAS  Google Scholar 

  25. Lehman HK, Greenberg MJ: The actions of FMRFamide-like peptides on visceral and somatic muscles of the snail Helix aspersa. J Exp Biol. 1987, 131: 55-68.

    PubMed  CAS  Google Scholar 

  26. Tobe SS, Zhang JR, Bowser PRF, Donly BC, Bendena WG: Biological activities of the allatostatin family of peptides in the cockroach, Diploptera punctata, and potential interactions with receptors. J Insect Physiol. 2000, 46: 231-242.

    Article  PubMed  CAS  Google Scholar 

  27. Oudejans RCH, Vroemen SF, Jansen RFR, Horst Van der DJ: Locust adipokinetic hormones: carrier-independent transport and differential inactivation at physiological concentrations during rest and flight. Proc Natl Acad Sci USA. 1996, 93: 8654-8659.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  28. Wicher D, Agricola HJ, Söhler S, Gundel M, Heinemann SH, Wollweber L, Stengl M, Derst C: Differential receptor activation by cockroach adipokinetic hormones produces differential effects on ion currents, neuronal activity, and locomotion. J Neurophysiol. 2006, 95: 2314-2325.

    Article  PubMed  CAS  Google Scholar 

  29. Drosophila 12 Genomes Consortium: Evolution of genes and genomes on the Drosophila phylogeny. Nature. 2007, 450: 203-218.

  30. Patterson C: Homology in classical and molecular biology. Mol Biol Evol. 1988, 5: 603-625.

    PubMed  CAS  Google Scholar 

  31. Gilbert GG: DroSpeGe: rapid access database for new Drosophila species genomes. Nucleic Acids Res. 2007, 35: D480-D485.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  32. Russo CAM, Takezaki N, Nei M: Molecular phylogeny and divergence times of Drosophilid species. Mol Biol Evol. 1995, 12: 391-404.

    PubMed  CAS  Google Scholar 

  33. Tamura K, Subramanian S, Kumar S: Temporal patterns of fruit fly (Drosophila) evolution revealed by mutation clocks. Mol Biol Evol. 2004, 21: 36-44.

    Article  PubMed  CAS  Google Scholar 

  34. Ashburner M, Golic KG, Hawley RS: Drosophila - a Laboratory Handbook. 2005, Cold Spring Harbor: Cold Spring Harbor Laboratory Press

    Google Scholar 

  35. Predel R, Wegener C, Russell WK, Tichy SE, Russell DH, Nachman RJ: Peptidomics of CNS-associated neurohemal systems of adult Drosophila melanogaster : a mass spectrometric survey of peptides from individual flies. J Comp Neurol. 2004, 474: 379-392.

    Article  PubMed  CAS  Google Scholar 

  36. Wegener C, Reinl T, Jänsch L, Predel R: Direct mass spectrometric peptide profiling and fragmentation of larval peptide hormone release sites in Drosophila melanogaster reveals tagma-specific peptide expression and differential processing. J Neurochem. 2006, 96: 1362-1374.

    Article  PubMed  CAS  Google Scholar 

  37. Baggerman G, Boonen K, Verleyen P, De Loof A, Schoofs L: Peptidomic analysis of the larval Drosophila melanogaster central nervous system by two-dimensional capillary liquid chromatography quadrupole time-of-flight mass spectrometry. J Mass Spectrom. 2005, 40: 250-260.

    Article  PubMed  CAS  Google Scholar 

  38. Nässel DR, Ohlsson LG, Cantera R: Metamorphosis of identified neurons innervating thoracic neurohemal organs in the blowfly: transformation of cholecystokininlike immunoreactive neurons. J Comp Neurol. 1988, 267: 343-356.

    Article  PubMed  Google Scholar 

  39. Santos JG, Pollák E, Rexer KH, Molnár L, Wegener C: Morphology and metamorphosis of the peptidergic Va neurons and the median nerve system of the fruit fly, Drosophila melanogaster. Cell Tissue Res. 2006, 326: 187-199.

    Article  PubMed  CAS  Google Scholar 

  40. Krause E, Wenschuh H, Jungblut H: The dominance of arginine-containing peptides in MALDI-derived tryptic mass fingerprints of proteins. Anal Chem. 1999, 71: 4160-4165.

    Article  PubMed  CAS  Google Scholar 

  41. O'Brien MA, Taghert PH: A peritracheal neuropeptide system in insects: release of myomodulin-like peptides at ecdysis. J Exp Biol. 1998, 201: 193-209.

    PubMed  Google Scholar 

  42. Dayhoff MO, Schwartz RM, Orcut BC: A model of evolutionary change in proteins. Atlas of Protein Sequence and Structure. Edited by: Dayhoff MO, Schwartz RM, Orcutt BC. 1978, Silver Spring, MD: National Biomedical Research Foundation, 345-352.

    Google Scholar 

  43. Jones DT, Taylor WR, Thornton JM: The rapid generation of mutation data matrices from protein sequences. Comput Appl Biosci. 1992, 8: 275-282.

    PubMed  CAS  Google Scholar 

  44. Nachman RJ, Holman GM, Cook BJ: Active fragments and analogs of the insect neuropeptide leucopyrokinin: structure-function studies. Biochem Biophys Res Commun. 1986, 137: 936-942.

    Article  PubMed  CAS  Google Scholar 

  45. Nachman RJ, Coast GM: Structure-activity relationships for in vitro diuretic activity of CAP2b in the housefly. Peptides. 2007, 28: 57-61.

    Article  PubMed  CAS  Google Scholar 

  46. Wells C, Aparicio K, Salmon A, Zadel A, Fuse M: Structure-activity relationship of ETH during ecdysis in the tobacco hornworm, Manduca sexta. Peptides. 2006, 27: 698-709.

    Article  PubMed  CAS  Google Scholar 

  47. Taneja-Bageshwar S, Strey A, Zubrzak P, Pietrantonio PV, Nachman RJ: Comparative structure-activity analysis of insect kinin core analogs on recombinant kinin receptors from southern cattle tick Boophilus microplus (Acari: Ixodidae) and mosquito (Aedes aegypti) (Diptera: Culicidae). Arch Insect Biochem Physiol. 2006, 62: 128-140.

    Article  PubMed  CAS  Google Scholar 

  48. Hauser F, Williamson M, Cazzamali G, Grimmelikhuijzen CJP: Identifying neuropeptide and protein hormone receptors in Drosophila melanogaster by exploiting genomic data. Brief Funct Genomic Proteomic. 2006, 4: 321-330.

    Article  PubMed  CAS  Google Scholar 

  49. Jörgensen LM, Hauser F, Cazzamali G, Williamson M, Grimmelikhuijzen CJP: Molecular identification of the first SIFamide receptor. Biochem Biophys Res Commun. 2006, 340: 696-701.

    Article  PubMed  Google Scholar 

  50. Hyun S, Lee Y, Hong ST, Bang S, Paik D, Kang J, Shin J, Lee J, Jeon K, Hwang S, Bae E, Kim J: Drosophila GPCR han is a receptor for the circadian clock neuropeptide PDF. Neuron. 2005, 48: 267-278.

    Article  PubMed  CAS  Google Scholar 

  51. Mertens I, Vandingenen A, Johnson EC, Shafer OT, Li W, Trigg JS, De Loof A, Schoofs L, Taghert PH: PDF receptor signaling in Drosophila contributes to both circadian and geotactic behaviors. Neuron. 2005, 48: 213-219.

    Article  PubMed  CAS  Google Scholar 

  52. Broeck Vanden J: Neuropeptides and their precursors in the fruitfly, Drosophila melanogaster. Peptides. 2001, 22: 241-254.

    Article  Google Scholar 

  53. Taghert PH, Veenstra JA: Drosophila neuropeptide signaling. Adv Genet. 2003, 49: 1-65.

    Article  PubMed  CAS  Google Scholar 

  54. Conlon JM: Molecular evolution of insulin in non-mammalian vertebrates. Am Zool. 2000, 40: 200-212.

    CAS  Google Scholar 

  55. Bader R, Wegener C, Pankratz MJ: Comparative neuroanatomy and genomics of hugin and pheromone biosynthesis activating neuropeptide (PBAN). Fly. 2007, 1: 228-231.

    Article  PubMed  Google Scholar 

  56. Markow TA, O'Grady PM: Drosophila biology in the genomic age. Genetics. 2007, 177: 1269-1276.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  57. Li B, Predel R, Neupert S, Hauser F, Tanaka Y, Cazzamali G, Williamson M, Arakane Y, Verleyen P, Schoofs L, Schachtner J, Grimmelikhuijzen CJP, Park Y: Genomics, transcriptomics, and peptidomics of neuropeptides and protein hormones in the red flour beetle Tribolium castaneum. Genome Res. 2008, 18: 113-122.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  58. Larracuente AM, Sackton TB, Greenberg AJ, Wong A, Singh ND, Sturgill D, Zhang Y, Oliver B, Clark AG: Evolution of protein-coding genes in Drosophila. Trends Genetics. 2008, 24: 114-123.

    Article  CAS  Google Scholar 

  59. Cho HJ, Acharjee S, Moon MJ, Oh DY, Vaudry H, Kwon HB, Seong JY: Molecular evolution of neuropeptide receptors with regard to maintaining high affinity to their authentic ligands. Gen Comp Endocrinol. 2007, 153: 98-107.

    Article  PubMed  CAS  Google Scholar 

  60. Brezden BL, Yeoman MS, Gardner DR, Benjamin PR: FMRFamide-activated Ca2+ channels in Lymnaea heart cells are modulated by 'SEEPLY', a neuropeptide encoded on the same gene. J Neurophysiol. 1999, 81: 1818-1826.

    PubMed  CAS  Google Scholar 

  61. Kenakin T: Ligand-selective receptor conformations revisited: the promise and the problem. Trends Pharmacol Sci. 2003, 24: 346-354.

    Article  PubMed  CAS  Google Scholar 

  62. Verleyen P, Baggerman G, Wiehart U, Schoeters E, van Lommel A, De Loof A, Schoofs L: Expression of a novel neuropeptide, NVGTLARDFQLPIPNamide, in the larval and adult brain of Drosophila melanogaster. J Neurochem. 2004, 88: 311-319.

    Article  PubMed  CAS  Google Scholar 

  63. Cazzamali G, Torp M, Hauser F, Williamson M, Grimmelikhuijzen CJP: The Drosophila gene CG9918 codes for a pyrokinin-1 receptor. Biochem Biophys Res Commun. 2005, 335: 14-19.

    Article  PubMed  CAS  Google Scholar 

  64. Predel R, Wegener C: Biology of the CAPA peptides in insects. Cell Mol Life Sci. 2006, 63: 2477-2490.

    Article  PubMed  CAS  Google Scholar 

  65. Park Y, Kim YJ, Adams ME: Identification of G protein-coupled receptors for Drosophila PRXamide peptides, CCAP, corazonin, and AKH supports a theory of ligand-receptor coevolution. Proc Natl Acad Sci USA. 2002, 99: 11423-11428.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  66. Iversen A, Cazzamali G, Williamson M, Hauser F, Grimmelikhuijzen CJP: Molecular identification of the first insect ecdysis triggering hormone receptors. Biochem Biophys Res Commun. 2002, 299: 924-931.

    Article  PubMed  CAS  Google Scholar 

  67. Cazzamali G, Grimmelikhuijzen CJP: Molecular cloning and functional expression of the first insect FMRFamide receptor. Proc Natl Acad Sci USA. 2002, 99: 12073-12078.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  68. Meeusen T, Mertens I, Clynen E, Baggerman G, Nichols R, Nachman RJ, Huybrechts R, De Loof A, Schoofs L: Identification in Drosophila melanogaster of the invertebrate G protein-coupled FMRFamide receptor. Proc Natl Acad Sci USA. 2002, 99: 15363-15368.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  70. Kean L, Cazenave W, Costes L, Broderick KE, Graham S, Pollock VP, Davies SA, Veenstra JA, Dow JA: Two nitridergic peptides are encoded by the gene capability in Drosophila melanogaster. Am J Physiol Regul Integr Comp Physiol. 2002, 282: R1297-R1307.

    Article  PubMed  CAS  Google Scholar 

  71. Iversen A, Cazzamali G, Williamson M, Hauser F, Grimmelikhuijzen CJP: Molecular cloning and functional expression of a Drosophila receptor for the neuropeptides capa-1 and -2. Biochem Biophys Res Commun. 2002, 299: 628-633.

    Article  PubMed  CAS  Google Scholar 

  72. Nachman RJ, Russell WK, Coast GM, Russell DH, Predel R: Mass spectrometric assignment of Leu/Ile in neuropeptides from single neurohemal organ preparations of insects. Peptides. 2005, 26: 2151-2156.

    Article  PubMed  CAS  Google Scholar 

  73. Nachman RJ, Russell WK, Coast GM, Russell DH, Miller JA, Predel R: Identification of PVK/CAP2b neuropeptides from single neurohemal organs of the stable fly and horn fly via MALDI-TOF/TOF tandem mass spectrometry. Peptides. 2006, 27: 521-526.

    Article  PubMed  CAS  Google Scholar 

  74. Larsen MJ, Burton KJ, Zantello MR, Smith VG, Lowery DL, Kubiak TM: Type A allatostatins from Drosophila melanogaster and Diploptera punctata activate two Drosophila allatostatin receptors, DAR-1 and DAR-2, expressed in CHO cells. Biochem Biophys Res Commun. 2001, 286: 895-901.

    Article  PubMed  CAS  Google Scholar 

  75. Lenz C, Williamson M, Hansen GN, Grimmelikhuijzen CJP: Identification of four Drosophila allatostatins as the cognate ligands for the Drosophila orphan receptor DAR-2. Biochem Biophys Res Commun. 2001, 286: 1117-1122.

    Article  PubMed  CAS  Google Scholar 

  76. Kubiak TM, Larsen MJ, Burton KJ, Bannow CA, Martin RA, Zantello MR, Lowery DE: Cloning and functional expression of the first Drosophila melanogaster sulfakinin receptor DSK-R1. Biochem Biophys Res Commun. 2002, 291: 313-320.

    Article  PubMed  CAS  Google Scholar 

  77. Palmer GC, Tran T, Duttlinger A, Nichols R: The drosulfakinin 0 (DSK 0) peptide encoded in the conserved Dsk gene affects adult Drosophila melanogaster crop contractions. J Insect Physiol. 2007, 53: 1125-1133.

    Article  PubMed  CAS  Google Scholar 

  78. Neupert S, Johard HAD, Nässel DR, Predel R: Single-cell peptidomics of Drosophila melanogaster neurons identified by GAL4-driven fluorescence. Anal Chem. 2007, 79: 3690-3694.

    Article  PubMed  CAS  Google Scholar 

  79. Rosenkilde C, Cazzamali G, Williamson M, Hauser F, Söndergaard L, DeLotto R, Grimmelikhuijzen CJP: Molecular cloning, functional expression, and gene silencing of two Drosphila receptors for the Drosophila neuropeptide pyrokinin-2. Biochem Biophys Res Commun. 2003, 309: 485-494.

    Article  PubMed  CAS  Google Scholar 

  80. Drosophila Species Genomes BLAST. [http://insects.eugenes.org/species/blast/]

  81. The New GENSCAN Web Server at MIT. [http://genes.mit.edu/GENSCAN.html]

  82. Kumar S, Tamura K, Nei M: MEGA3: integrated software for molecular evolutionary genetics analysis and sequence alignment. Brief Bioinform. 2004, 5: 150-163.

    Article  PubMed  CAS  Google Scholar 

  83. Nicholas KB, Nicholas HB, Deerfield DW: GeneDoc: analysis and visualization of genetic variation. EMBNEW.News. 1997, 4: 14-

    Google Scholar 

  84. SignalP 3.0 Server. [http://www.cbs.dtu.dk/services/SignalP/]

  85. Veenstra JA: Mono- and dibasic proteolytic cleavage sites in insect neuroendocrine peptide precursors. Arch Insect Biochem Physiol. 2000, 43: 49-63.

    Article  PubMed  CAS  Google Scholar 

  86. Zhang J, Nei M: Accuracies of ancestral amino acid sequences inferred by the parsimony, likelihood, and distance methods. J Mol Evol. 1997, 44: S139-S146.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

We thank Arndt von Haeseler (Vienna), Dick R Nässel (Stockholm), Reinhard Predel (Jena), Klaus Reinhard (Sheffield) and Steffen Roth (Bergen) for stimulating discussions, help with the statistics and critical reading of the manuscript; two unknown referees for valuable comments; Fritz-Olaf Lehmann, Uwe Rose (Ulm) and the Tucson Drosophila Stock Center for the kind gift of flies; the Botanical Garden Marburg for supply of Opuntia leaves; J Kahnt and Rolf K Thauer (Max-Planck-Institute of Terrestrial Microbiology, Marburg) for the use of their mass spectrometer; Ruth Hyland and Renate Renkawitz-Pohl for fly housing; and Uwe Homberg for general support. Funded by the Deutsche Forschungsgemeinschaft DFG (WE 2652/2) and the Fonds der Chemischen Industrie (Sachkosten-Zuschuss für den Hochschullehrer-Nachwuchs to CW).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Wegener.

Additional information

Authors' contributions

AG and CW carried out the databank searches and direct peptide profiling, analyzed the mass data, and drafted the manuscript. AG reared the flies. CW carried out sequence alignments and calculated the distances. Both authors read and approved the final manuscript.

Electronic supplementary material

13059_2008_1993_MOESM1_ESM.doc

Additional data file 1: Prepropeptide sequences as identified by BLAST searches in the 12 Drosophila genomes. (DOC 136 KB)

Additional data file 2: Peptide sequences predicted from the identified prepropeptides. (FAS 20 KB)

13059_2008_1993_MOESM3_ESM.eps

Additional data file 3: Frequency of occurrence of peptides in mass spectrometric profiles of the ring gland. (EPS 3 MB)

Additional data file 4: Mass spectrograms on the processing of FMRFa-5ext from the FMRFa-prepropeptide. (EPS 16 MB)

Additional data file 5: Overall average amino acid distances Dso. (DOC 38 KB)

Authors’ original submitted files for images

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wegener, C., Gorbashov, A. Molecular evolution of neuropeptides in the genus Drosophila. Genome Biol 9, R131 (2008). https://doi.org/10.1186/gb-2008-9-8-r131

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/gb-2008-9-8-r131

Keywords