The leucokinin (LK) family of neuropeptides has been found widely amongst invertebrates. A member of this family was purified from adults of the fruit fly Drosophila melanogaster. The peptide sequence for Drosophila leucokinin (DLK) was determined as Asn-Ser-Val-Val-Leu-Gly-Lys-Lys-Gln-Arg-Phe-His-Ser-Trp-Gly-amide, making it the longest member of the family characterized to date. Synthetic DLK peptide was shown to act to stimulate fluid secretion in D. melanogaster Malpighian (renal) tubules by approximately threefold, with an EC50 of approximately 10−10 mol l−1, and a secondary effect at approximately 10−7 mol l−1. DLK also acted to elevate intracellular [Ca2+ ] in the Malpighian tubules by approximately threefold, with an EC50 of 10−10 to 10−9 mol l−1. Responses were detected in stellate cells and occasionally in principal cells, although at no concentration tested did [Ca2+ ] in the principal cell increase significantly above background. In stellate cells, DLK produced a biphasic rise in intracellular [Ca2+ ] from resting levels of 80–100 nmol l−1, with a transient peak being followed by a slower rise that peaked at 200–300 nmol l−1 after 3 s, then decayed over approximately 10 s. The wide range of concentrations over which DLK acts suggests the involvement of more than one receptor. The genomic sequence encoding the DLK peptide has been identified, and the gene has been named pp. The gene resides at cytological location 70E3–70F4 of chromosome 3L. The localisation of this first Drosophila LK gene in a genetic model permits a genetic analysis of the locus.

The integration of extracellular signals to produce an appropriate cellular response is central to our understanding of organismal function. In the analysis of such systems, the use of genetic manipulation can potentially prove invaluable, as it allows the effects of mutagenesis of specific components of the pathway to be studied in an otherwise normal organism. For such studies, the fruit fly Drosophila melanogaster provides probably the best balance between genetic power and experimental tractability (Rubin, 1988).

At present, the best D. melanogaster model system for the analysis of neurohormonal control is provided by the renal, or Malpighian, tubules (Dow et al., 1994b, 1998), which regulate salt and water balance by transepithelial secretion. These four simple epithelial tubular structures are composed of precisely determined numbers of cells of multiple types that can be genetically tagged by transposon-based enhancer trapping (Sözen et al., 1997). The main segment of each tubule (Fig. 1) is the chief site of fluid secretion and electrolyte regulation, and it is composed of interspersed Type I (principal) and Type II (stellate and bar-shaped) cells (Sözen et al., 1997).

Fig. 1.

Summary of the transport physiology of the main segment of the Drosophila melanogaster Malpighian tubule. CAP2b, cardiac acceleratory peptide 2b; GC, guanylate cyclase; LK, leucokinin; DNOS, nitric oxide synthase, X, extracellular ligand for cyclic AMP pathway. See text for details.

Fig. 1.

Summary of the transport physiology of the main segment of the Drosophila melanogaster Malpighian tubule. CAP2b, cardiac acceleratory peptide 2b; GC, guanylate cyclase; LK, leucokinin; DNOS, nitric oxide synthase, X, extracellular ligand for cyclic AMP pathway. See text for details.

Fluid production by the main segment of the tissue is energized by a vacuolar H+ -motive ATPase, located in the apical membrane of the large principal cells, where it drives a net secretion of K+via an amiloride-sensitive K+ /H+ exchanger (Davies et al., 1996; Dow et al., 1994b, 1998; O’Donnell et al., 1996; Sözen et al., 1997). Chloride flows through maxi-Cl channels in the stellate cells to balance the charge transfer (O’Donnell et al., 1996, 1998), and water is thought to flow transcellularly via aquaporins in the stellate cells (Dow et al., 1995). A range of organic solutes are actively transported by the main segment principal cells (Sözen et al., 1997).

Our detailed understanding of these transport processes is balanced by a knowledge of the signalling pathways that control them (Fig. 1). The V-ATPase is stimulated by a rise in either cyclic AMP or cyclic GMP level (Dow et al., 1994a,b; O’Donnell et al., 1996); although the extracellular ligand for the former is unknown, the cardiac acceleratory peptide CAP2b is known to act through intracellular Ca2+ in principal cells to stimulate an intrinsic nitric oxide synthase (DNOS) to produce NO and so raise cyclic GMP concentration through the action of NO on a soluble guanylate cyclase (GC; Davies et al., 1995, 1997; Rosay et al., 1997).

Invertebrate leucokinins (LKs) constitute a family of myotropic neuropeptides that are active in all insect species so far studied (Cantera et al., 1992; Chen et al., 1994; Hayes et al., 1994; Holman et al., 1986a,b, 1987a,b; Nässel and Lundquist, 1991; Nässel et al., 1992; O’Donnell et al., 1996; Pannabecker et al., 1993; Schoofs et al., 1992; Veenstra, 1994). Related peptides have been identified in snails (Cox et al., 1997) and shrimp (Nieto et al., 1998), and leucokinin-related immunoreactivity has been described in a nematode (Smart et al., 1993) and an arachnid central nervous system (Schmid and Becherer, 1996). The biological activity of these peptides appears to hinge upon the presence of a C-terminal pentapeptide (Nachman et al., 1995), a feature shared with the vertebrate peptide family of tachykinins. First isolated from Leucophaea maderae (cockroach) on the basis of their ability to induce hindgut contractions, LK peptides can also stimulate ion transport and fluid secretion in insect renal tubules (Coast et al., 1990; Hayes et al., 1989; O’Donnell et al., 1996; Veenstra et al., 1997b), often in a cross-specific manner (Hayes et al., 1989; O’Donnell et al., 1996).

Peptides of the LK family have been shown to raise the Cl shunt conductance in insect renal tubules (Hayes et al., 1989; O’Donnell et al., 1996, 1998; Pannabecker et al., 1993). In D. melanogaster, it has further been shown that LKs act on stellate cells to raise intracellular [Ca2+ ], and thence to raise Cl flux specifically through stellate cells (O’Donnell et al., 1998; Rosay et al., 1997). This, in turn, provides a model, applicable at least to other Diptera, in which stellate cells are conspicuous and possibly to other insects in which such functional specialization may not be apparent (Dow et al., 1998).

Previous studies on the renal response of D. melanogaster to LKs have employed synthetic peptides derived from other orders of insect (Davies et al., 1995; O’Donnell et al., 1996, 1998; Rosay et al., 1997). However, in this study, we describe the isolation and characterization of the first D. melanogaster member of this family, Drosophila leucokinin (DLK). The action of this peptide on tubules is similar to those of non-native peptides in that it raises both fluid secretion rates and intracellular [Ca2+ ]. Interestingly, the two responses occur over distinct concentration ranges, suggesting the possibility of multiple effects.

Flies

Drosophila melanogaster (strain Oregon R) were raised in tubes or bottles on synthetic diet under standard conditions, as described previously (Dow et al., 1994b). Flies transgenic for the aequorin transgene, under control of the UAS promoter (UASaeq) enhancer trap lines that direct GAL4 expression to principal cells (c42), secondary cells (c710) and lower tubules (c507) and a heat-shock/GAL4 construct (hsGAL4) were those described previously (Rosay et al., 1997; Sözen et al., 1997). To produce flies with cell-specific aequorin expression, homozygous UASaeq flies were crossed to flies homozygous for any of the driver insertion lines described, and the adult progeny were used as described below (see Fig. 3).

Peptide purification

Throughout the purification procedure, leucokinin-immunoreactivity was identified by competitive enzyme-linked immunosorbent assay (ELISA), the general procedure for which followed that described elsewhere (Veenstra and Lambrou, 1995), and used an antibody prepared against leucokinin I (Nässel and Lundquist, 1991) and a leucokinin II–glutaraldehyde–thyroglobulin conjugate. This ELISA is more sensitive and less specific for a single leucokinin than the leucokinin IV ELISA used for the isolation of the mosquito leucokinins (Veenstra, 1994).

The Drosophila leucokinin was isolated from 400 g of the Canton S strain (adults 7–14 days old) following essentially the protocol used for the isolation and identification of 13 neuropeptides from the mosquito Aedes aegypti (Veenstra, 1994, 1998; Veenstra et al., 1997a), but with improvements suggested by previous results (Fig. 2). Samples of 20 g of frozen flies were homogenized in 200 ml of Bennett’s mixture [5 % formic acid, 1 % trifluoroacetic acid (TFA), 1 % NaCl and 1 mol l−1 HCl in water; Bennett et al., 1981] in a Waring mixer for 5 min at high speed. Extracts were centrifuged for 20 min at 13 000 g at 4 °C, and the were pellets re-extracted once. The supernatants were loaded on previously activated and equilibrated home-made ‘Super-Sep-Paks’, containing 7 g of preparatory C18 reversed-phase beads (55–105 μm, Millipore Corporation, France), and the was material eluted, collected, lyophilized and stored as described in detail for the mosquito leucokinins (Veenstra, 1994). Preliminary experiments showed that, while some D. melanogaster pigments were poorly retained on such columns, no peptides of interest were lost in this step. After the entire 400 g had been processed, the combined lyophilized material was dissolved in 0.1 % of TFA, reprocessed on a single ‘Super-Sep-Pak’ and the peptide material lyophilized.

Fig. 2.

Procedure for the isolation of Drosophila leucokinin (DLK). (A) Flow chart of isolation procedure. (B) HPLC elution profile of the crude extract. The peak that was collected is marked with an arrow. Abbreviations for chemicals are explained in Materials and methods.

Fig. 2.

Procedure for the isolation of Drosophila leucokinin (DLK). (A) Flow chart of isolation procedure. (B) HPLC elution profile of the crude extract. The peak that was collected is marked with an arrow. Abbreviations for chemicals are explained in Materials and methods.

Four different high-performance liquid chromatography (HPLC) steps were used to purify the Drosophila leucokinin. The first HPLC step used an Econosil C18 reversed-phase column (250 mm×22.5 mm, Altech Associates Inc., Deerfield, IL, USA). The column was equilibrated in 0.1 % heptafluorobutyric acid (HFBA) in water at a flow rate of 10 ml min−1. After injection of the final lyophilisate redissolved in 0.1 % HFBA, the column was run isocratically for 10 min, after which a linear gradient to 16 % CH3CN and 0.1 % HFBA in water over 20 min was started, followed immediately by a linear gradient over 90 min to 52 % CH3CN and 0.1 % HFBA in water. Fractions were collected every 48 s and analysed by ELISA for the presence of leucokinin-immunoreactive peptides. Absorbance was measured at 280 nm.

Fractions containing leucokinin-immunoreactivity were identified and further purified on an Ultremex C1 reversed-phase column (250 mm×4.6 mm; Phenomenex, Torrance, CA, USA). This column was equilibrated with 0.1 % TFA in water and eluted at 1 ml min−1. After injection of the leucokinin-immunoreactive material, the column was eluted isocratically with 0.1 % TFA for 10 min, followed by a linear gradient over 60 min to 13 % CH3CN and 0.1 % TFA in water, and a second linear gradient over 30 min to 32.5 % CH3CN and 0.1 % TFA in water. Fractions were collected every minute and analysed for leucokinin immunoreactivity. Absorbance was measured at 214 and 280 nm.

The third HPLC purification step used a Microsorb Phenyl reversed-phase column (250 mm×4.6 mm; Rainin Instrument Company, Inc., Woburn, MA, USA). This column was equilibrated with 13 % CH3CN and 0.1 % TFA in water; after injection of the leucokinin-immunoreactive material, a linear gradient over 10 min to 19.5 % CH3CN and 0.1 % TFA in water was started, followed immediately by a second linear gradient over 60 min to 32.5 % CH3CN and 0.1 % TFA in water, and finally a gradient over 15 min to 45.5 % CH3CN and 0.1 % TFA in water. Fractions were collected at 1 min intervals and analysed by ELISA for leucokinin immunoreactivity. Absorbance was measured at 214 and 280 nm.

The final HPLC step was performed on a Protein Pak I-125 column (300 mm×7.8 mm; Millipore Corporation), used in normal phase mode. The column was equilibrated in 95 % CH3CN and 0.01 % TFA in water at a flow rate of 1.5 ml min−1. After injection of the leucokinin-immunoreactive material, the column was eluted isocratically for 10 min, followed by a gradient over 80 min to 50 % CH3CN and 0.01 % TFA in water. Fractions were collected every 2 min and absorbance was measured at 214 and 280 nm.

The isolated peptide was sequenced on an automated protein sequencer (model 494A, Perkin Elmer Applied Biosystems Procise) by the Unité de Recherche de Biochimie et Structure des Protéines of INRA at Jouy-en-Josas. Drosophila leucokinin was synthesized by the Unité de Biophysique Structurale, Université Bordeaux I on an automated peptide synthesizer (model 341A, Applied Biosystems).

Fluid secretion assay

Fluid secretion experiments were performed on 3-to 7-day-old adults of D. melanogaster as described previously (Dow et al., 1994b). Tubule pairs were placed in 10 μl drops of Schneider’s medium under paraffin oil, and one tubule of the pair was anchored to a metal pin outside the drop. Fluid produced by the remaining tubule generated a small droplet at the urethra under oil. These were collected at 10 min intervals, the diameter measured and their volumes calculated. DLK was serially diluted in Schneider’s medium and added to sets of (usually 10) tubules immediately after the 30 min reading, and fluid production was measured for a further 30 min.

Aequorin measurement of intracellular [Ca2+ ]

The effects of DLK and other peptides on intracellular [Ca2+ ] in principal and stellate cells were measured as described previously (Fig. 3) (O’Donnell et al., 1998; Rosay et al., 1997). Briefly, tubules were dissected from adult (3–10 days old) flies, pooled in groups of 20, and incubated in Schneider’s medium containing 2.5 μmol l−1 coelenterazine for 2–4 h. Luminescence recordings were made with a Berthold–Wallac luminometer. Tubule samples (one per data point) were mock-injected with Schneider’s medium, followed by application of test solutions in Schneider’s medium. Responses were measured over a period of 1 min. At the end of each experiment, the total luminescence of each tubule sample was assessed by injecting 100 mmol l−1 Ca2+ /1 % Triton X-100 to permeabilise the cells and so discharge the remaining aequorin. This allowed luminescence rates to be converted to Ca2+ concentration by reverse integration with a program written in Perl, as described previously (Rosay et al., 1997).

Fig. 3.

Measurement of intracellular [Ca2+ ] by cell-specific expression of aequorin. The P{GAL4} transposon is capable of driving expression of any transgene that has a UAS promoter in a pattern that reflects the genomic context of the P{GAL4} insertion. A collection of P{GAL4} lines that drive expression in particular regions or cell types of Malpighian tubules has been characterised previously. In this experiment, flies from line c42 (homozygous for a P{GAL4} insertion that can drive transgene expression in principal cells of the main segment of the Malpighian tubule) or from line c710 (homozygous for a P{GAL4} insertion that can drive transgene expression in stellate cells) were mated to flies homozygous for a second chromosome insertion of the Aequorea victoria apoaequorin gene under control of the UAS promoter (Rosay et al., 1997). In the progeny of such flies, apoaequorin was expressed in the appropriate cell-specific pattern (Rosay et al., 1997). Tubules were dissected from 3-to 10-day-old adults and incubated with coelenterazine to permit the reconstitution of active aequorin. Ca2+ levels were measured in a luminometer with two automated injection channels, allowing real-time monitoring of light emission at 470 nm (h v470) in response to Drosophila leucokinin (DLK) application. At the end of the experiment, the remaining aequorin (>98 %) was discharged by permeabilisation of the tubules with Triton X-100, allowing the conversion of relative light intensity to real-time [Ca2+ ] by reverse integration as described previously (Rosay et al., 1997).

Fig. 3.

Measurement of intracellular [Ca2+ ] by cell-specific expression of aequorin. The P{GAL4} transposon is capable of driving expression of any transgene that has a UAS promoter in a pattern that reflects the genomic context of the P{GAL4} insertion. A collection of P{GAL4} lines that drive expression in particular regions or cell types of Malpighian tubules has been characterised previously. In this experiment, flies from line c42 (homozygous for a P{GAL4} insertion that can drive transgene expression in principal cells of the main segment of the Malpighian tubule) or from line c710 (homozygous for a P{GAL4} insertion that can drive transgene expression in stellate cells) were mated to flies homozygous for a second chromosome insertion of the Aequorea victoria apoaequorin gene under control of the UAS promoter (Rosay et al., 1997). In the progeny of such flies, apoaequorin was expressed in the appropriate cell-specific pattern (Rosay et al., 1997). Tubules were dissected from 3-to 10-day-old adults and incubated with coelenterazine to permit the reconstitution of active aequorin. Ca2+ levels were measured in a luminometer with two automated injection channels, allowing real-time monitoring of light emission at 470 nm (h v470) in response to Drosophila leucokinin (DLK) application. At the end of the experiment, the remaining aequorin (>98 %) was discharged by permeabilisation of the tubules with Triton X-100, allowing the conversion of relative light intensity to real-time [Ca2+ ] by reverse integration as described previously (Rosay et al., 1997).

Cyclic nucleotide assays

Intracellular cyclic AMP and cyclic GMP concentrations were measured by radioimmunoassay (Amersham Biotrak Amerlex M kits) as described previously (Davies et al., 1995). Briefly, tubules were dissected in Schneider’s medium and exposed to the desired concentration of DLK for 10 min in the presence of the non-specific phosphodiesterase inhibitor, isobutylmethyl xanthine (IBMX), before lysis and assay by competitive immunoassay according to the manufacturer’s instructions.

Identification of the gene encoding DLK

Because the DLK peptide sequence was too short to permit authoritative database searching, a putative propeptide sequence was constructed by adding a second glycine at the C terminus, to act as an amidation signal, and flanking the resulting sequence with lysine-arginine dibasic cleavage signals. The resulting sequence was used for low-stringency TBLASTN searches of the Genbank and Berkeley Drosophila Genome Project (BDGP).

Expression of the putative gene, identified by the searches, was verified by reverse-transcription polymerase chain reaction (RT–PCR) using primers that matched the genomic fraction, and head mRNA as the template. This procedure has been described elsewhere (Davies et al., 1977).

Peptide purification

After the first HPLC step, a leucokinin-immunoreactive peak was detected in fractions 113, 114 and 115, corresponding to a retention time of 89.6–92.0 min. The immunoreactive material eluted from the C1 column with a retention time of 36–38 min, and subsequently on the phenyl column it was recovered between 31 and 33 min. The final HPLC column yielded one major peak at 63.4 min containing all leucokinin-immunoreactive material and three minor contaminants which were well separated from the Drosophila leucokinin. Sequence analysis yielded unambiguously the following sequence, Asn-Ser-Val-Val-Leu-Gly-Lys-Lys-Gln-Arg-Phe-His-Ser-Trp-Gly and showed that a total of approximately 600 pmol had been purified.

As all leucokinins identified so far are C-terminally amidated, and the leucokinin I ELISA used did not recognize a synthetic non-amidated analogue of Aedes leucokinin II and thus appears to recognize only C-terminally amidated peptides, the Drosophila leucokinin was synthesized with a C-terminal amide. The synthetic Drosophila leucokinin had the same retention time on HPLC as the isolated peptide and thus confirms the presence of the C-terminal amide in the natural peptide, since oligopeptide analogues differing only in the presence or absence of a C-terminal amide differ significantly (approximately 2 min) in retention time under the conditions used. Hence, the structure of Drosophila leucokinin is Asn-Ser-Val-Val-Leu-Gly-Lys-Lys-Gln-Arg-Phe-His-Ser-Trp-Gly-amide (this sequence has been deposited in the SwissProt database with the accession number P81829).

Fluid secretion assay

The DLK peptide produced a clear stimulation of fluid secretion, with a time course similar to that seen for other members of the leucokinin family (Fig. 4A). Typically, the tubule responded maximally in less than the sampling interval, and the increase in fluid secretion rate was sustained over many minutes. The extent of the increase in fluid secretion (approximately 2–3 nl min−1 ) is comparable with that observed previously in D. melanogaster tubules for non-native leucokinins. The dose–response curve (Fig. 4B) shows that half-maximal effect is seen between 10−11 to 10−10 mol l−1, and shows significant additional stimulation above 10−7 mol l−1.

Fig. 4.

Effects of Drosophila leucokinin (DLK) on rates of fluid secretion by Malpighian tubules. (A) Typical experiment, showing the response of tubule fluid secretion rate to DLK, added to a final concentration of 10−9 mol l−1 at 30 min. (B) Dose–response curve for the effect of DLK on fluid secretion rate. Data are presented as means ± S.E.M., N>10.

Fig. 4.

Effects of Drosophila leucokinin (DLK) on rates of fluid secretion by Malpighian tubules. (A) Typical experiment, showing the response of tubule fluid secretion rate to DLK, added to a final concentration of 10−9 mol l−1 at 30 min. (B) Dose–response curve for the effect of DLK on fluid secretion rate. Data are presented as means ± S.E.M., N>10.

Aequorin measurement of intracellular [Ca2+ ]

DLK produced a rapid elevation of intracellular [Ca2+ ] in stellate cells (Fig. 5), as we have reported for other members of the leucokinin family. However, the authentic DLK peptide elicited a rapid transient response in the first time-point (100 ms after injection), followed by a slower rise in [Ca2+ ] that reached a higher peak 3 s after injection, then decayed rapidly within 10–20 s (Fig. 5). The time course of the slower peak was not sensitive to the concentration of DLK (Fig. 5). The approximately 200 nmol l−1 increase in cytosolic [Ca2+ ] above a resting level of 70 nmol l−1 (Fig. 5) is consistent with values we have previously reported for non-native peptides (O’Donnell et al., 1998), and with the known EC50 for Drosophila calmodulin, of approximately 250 nmol l−1.

Fig. 5.

Drosophila leucokinin (DLK) stimulates an increase in intracellular [Ca2+ ] in stellate cells. Typical intracellular [Ca2+ ] traces for stellate (left-hand panels) and principal (right-hand panels) cells, for resting cells and after mock injection of vehicle (M) or of DLK at 10−6 mol l−1 (top panels), 10−7 mol l−1 (middle panels) or 10−8 mol l−1 final concentration (bottom panels).

Fig. 5.

Drosophila leucokinin (DLK) stimulates an increase in intracellular [Ca2+ ] in stellate cells. Typical intracellular [Ca2+ ] traces for stellate (left-hand panels) and principal (right-hand panels) cells, for resting cells and after mock injection of vehicle (M) or of DLK at 10−6 mol l−1 (top panels), 10−7 mol l−1 (middle panels) or 10−8 mol l−1 final concentration (bottom panels).

In contrast, DLK had no effect on intracellular [Ca2+ ] in principal cells, as delineated by line c42 (Figs 5, 6). Principal cell responses were seen occasionally, with a similar time course to those observed in stellate cells; however, these sporadic results did not constitute a statistically significant response (Fig. 6).

Fig. 6.

Effects of Drosophila leucokinin (DLK) on intracellular [Ca2+ ]. Data are derived from traces for stellate cells (see left-hand panel in Fig. 5) and principal cells (see right-hand panel in Fig. 5). For each cell type and concentration, mean Ca2+ concentrations, throughout the experiment, are shown as initial background levels (white columns), maximum response to mock injection (light grey columns), second background reading (dark grey columns) and maximum response to DLK (black columns). A ‘clean’ experiment is therefore one in which the background and mock readings (the first three columns in a series) are all very similar. Data are presented as means ± S.E.M. (N>4); significant responses to DLK, compared with mock injection, are marked with an asterisk (Student’s t-test for matched samples, two-tailed, taking critical level for P=0.05).

Fig. 6.

Effects of Drosophila leucokinin (DLK) on intracellular [Ca2+ ]. Data are derived from traces for stellate cells (see left-hand panel in Fig. 5) and principal cells (see right-hand panel in Fig. 5). For each cell type and concentration, mean Ca2+ concentrations, throughout the experiment, are shown as initial background levels (white columns), maximum response to mock injection (light grey columns), second background reading (dark grey columns) and maximum response to DLK (black columns). A ‘clean’ experiment is therefore one in which the background and mock readings (the first three columns in a series) are all very similar. Data are presented as means ± S.E.M. (N>4); significant responses to DLK, compared with mock injection, are marked with an asterisk (Student’s t-test for matched samples, two-tailed, taking critical level for P=0.05).

Cyclic nucleotide assays

By contrast with its effects on intracellular [Ca2+ ], DLK did not affect cyclic AMP levels in tubules (Fig. 7). This is consistent with our previous data for non-native peptides (Davies et al., 1995). DLK elicited a small reduction in cyclic GMP levels in tubules (Fig. 7). This might imply some cross-talk between the two signalling pathways or the existence of multiple DLK receptor subtypes; however, the significance of this is not clear.

Fig. 7.

Drosophila leucokinin (DLK) has a minimal effect on tubule cyclic nucleotide levels. Tubule cyclic AMP and cyclic GMP levels were measured by radioimmunoassay, as described in the text. Treatments are: 10−8, Drosophila leucokinin to 10−8 mol l−1 final concentration; 10−7, Drosophila leucokinin to 10−7 mol l−1 final concentration; IBMX, control with isobutylmethyl xanthine alone; CAP2b, positive control for cyclic GMP assay, with the cardiac acceleratory peptide CAP2b added to 10−7 mol l−1 final concentration. (There is no known agonist that raises cyclic AMP levels in Drosophila melanogaster Malpighian tubules.) Data are presented as means ± S.E.M. (N=3 independent samples, each pooled from 20 tubules). Where error bars are not shown, they are too small to reproduce.

Fig. 7.

Drosophila leucokinin (DLK) has a minimal effect on tubule cyclic nucleotide levels. Tubule cyclic AMP and cyclic GMP levels were measured by radioimmunoassay, as described in the text. Treatments are: 10−8, Drosophila leucokinin to 10−8 mol l−1 final concentration; 10−7, Drosophila leucokinin to 10−7 mol l−1 final concentration; IBMX, control with isobutylmethyl xanthine alone; CAP2b, positive control for cyclic GMP assay, with the cardiac acceleratory peptide CAP2b added to 10−7 mol l−1 final concentration. (There is no known agonist that raises cyclic AMP levels in Drosophila melanogaster Malpighian tubules.) Data are presented as means ± S.E.M. (N=3 independent samples, each pooled from 20 tubules). Where error bars are not shown, they are too small to reproduce.

Identification of the gene encoding DLK

Using the putative propeptide sequence, no hits were obtained on the 80 000 expressed sequence tags generated by the Berkeley Drosophila Genome Project, implying that DLK is a low-abundance transcript. However, a recent genomic sequence (Genbank Accession AC006496), corresponding to the first-pass sequence of Bacterial Artificial Chromosome (BAC) clone BACR48007, contains a DNA sequence that encodes the putative propeptide completely (Fig. 8). Although this preliminary sequence is too short to permit the full identification of the open reading frame (and thus of any related peptides encoded by the same gene), it is sufficient to localise the gene to 70E3–70F4 on chromosome 3L. Expression of the putative gene was verified by RT–PCR off head mRNA template. The resulting products were sequenced on both strands, and the sequence deposited in the GenBank data bank, with the accession number AF192342.

Fig. 8.

Genomic sequence encoding Drosophila leucokinin (DLK). Top rows show the genomic DNA sequence corresponding to the reverse complement of bases 44013–44132 of Genbank accession number AC006496, while the lower rows show the corresponding amino acid sequence. Putative dibasic flanking cleavage signals are shown in bold type, while the C-terminal amidation signal is shown in italic type. Amino acids are centred on their codons.

Fig. 8.

Genomic sequence encoding Drosophila leucokinin (DLK). Top rows show the genomic DNA sequence corresponding to the reverse complement of bases 44013–44132 of Genbank accession number AC006496, while the lower rows show the corresponding amino acid sequence. Putative dibasic flanking cleavage signals are shown in bold type, while the C-terminal amidation signal is shown in italic type. Amino acids are centred on their codons.

The gene encoding the DLK peptide was named pp. Although there are no obvious candidate loci that might correspond to this gene, the proximity of several P-element insertion stocks means that mutagenesis of the locus should prove straightforward.

Peptide purification

In our previous peptide isolation work on the mosquito Aedes aegypti, we noticed that the difference in selectivity between the C18 and phenyl reversed-phase columns was very limited, although the resolving power of the first preparative HPLC column was remarkable. From this, we reasoned that it might be possible to eliminate the HPLC step using HFBA on the phenyl column by using HFBA as the pairing ion for the preparative column and simultaneously decreasing the fraction size for this column. Because absorbance during the first HPLC purification step at 214 nm is off-scale, this had the added advantage of eliminating the problems associated with working with HFBA, which is usually too impure to obtain good absorbance recordings at 214 nm. Although it is impossible to judge whether the efficacy of these improvements will be valid for other purifications from the same batch of material, we obtained excellent purification from 400 g of whole Drosophila melanogaster in just four HPLC purification steps.

The leucokinin peptide family

The DLK peptide is clearly a member of the invertebrate leucokinin family, sharing both the canonical -Phe-X-X-Trp-Gly-NH2 C terminus and broader similarity with the rest of the family (Table 1). Of the known members, it is most similar in both sequence and overall length to Aedes leucokinin I (Veenstra et al., 1997b) although, at 15 residues, DLK is the longest leucokinin identified to date.

Table 1.

Comparison of the Drosophila leucokinin (DLK) sequence with known leucokinin members

Comparison of the Drosophila leucokinin (DLK) sequence with known leucokinin members
Comparison of the Drosophila leucokinin (DLK) sequence with known leucokinin members

Tubule effects of DLK

Although originally identified in a hindgut motility assay, interest in leucokinins has focused on their diuretic actions on Malpighian tubules. In the D. melanogaster system, it proved possible to measure (for the first time in any insect) an increase in intracellular [Ca2+ ] in stellate cells in response to a leucokinin (Rosay et al., 1997). Here, we have expanded these results using an authentic peptide. DLK is the most potent stimulant of fluid secretion yet identified in D. melanogaster; it produces a rapid, sustained increase in the rate of fluid secretion to 2–4 nl min−1, with an EC50 of less than 0.1 nmol l−1 (Fig. 4). This concentration makes it likely that DLK is an authentic ligand for the tubule receptor, because most neuropeptide receptors have affinities in the range 10−9 to 10−11 mol l−1 (Watson and Arkinstall, 1994).

The effects of DLK on intracellular [Ca2+ ] are consistent with this model. As expected from previous studies (O’Donnell et al., 1996; Rosay et al., 1997), the primary sites of intracellular Ca2+ response to DLK stimulation in D. melanogaster tubules are stellate cells, where a rapid (100 ms) and transient Ca2+ ‘spike’ is followed by a slower Ca2+ ‘wave’ that peaks after 3 s (Fig. 5). Since increased cytosolic Ca2+ levels return to baseline within approximately 20 s of DLK application while the effects on fluid secretion persist for at least 30 min (Fig. 4), it is necessary to invoke a long-lived downstream mediator of Ca2+ signalling.

Although the fluid secretion and [Ca2+ ] data paint a persuasive picture confirming that DLK acts to stimulate fluid production through a selective action on intracellular [Ca2+ ] in the stellate cells, there is scope for further complexity. DLK elicits significant elevations of the secretion response over a very wide concentration range (10−12 to 10−4 mol l−1 ; Fig. 4). The maximum fluid secretion rates observed at concentrations of 10−7, 10−6, 10−5 and 10−4 mol l−1 are significantly higher than those observed at 10−10, 10−9 or 10−8 mol l−1 (Fig. 4). This bimodal response is reminiscent of the position in Aedes aegypti tubules, where the collapse of the apical potential (and by inference the increase in Cl shunt conductance) occurs at a concentration higher than that required to stimulate fluid production (Veenstra et al., 1997b). Furthermore, A. aegypti leucokinins I and III, but not leucokinin II, have the capacity to increase the fluid secretion rate of isolated mosquito tubules (Veenstra et al., 1997b). Taken together, these lines of evidence suggest that more than one class of receptor may be involved in LK signalling in Malpighian tubules. It is possible that leucokinins modulate not only stellate cell Cl conductance but also processes that might include the remodelling of intercellular junctions (Pannabecker et al., 1993) or the recruitment of plasma membrane water channels (Dow et al., 1995). Given the possibility that, as in A. aegypti, further related peptides might exist in D. melanogaster, it is possible that the responses elicited by DLK at high concentration are through low-affinity interaction with a receptor for a related peptide.

A common feature of vertebrate and invertebrate neuropeptide families is that individual family members tend to arise through processing of a common propeptide. For example, the three LKs identified in A. aegypti are encoded by a single cDNA (Veenstra, 1994; Veenstra et al., 1997b). We are accordingly investigating the gene that encodes DLK, based on the chromosomal localisation at 70E3–70F4. As well as providing information on other neuropeptides encoded by the same gene, the reverse genetic tools unique to D. melanogaster will provide a unique opportunity to dissect the functions of individual neuropeptides.

We are most grateful to Ms Gayle McEwen for her help in the experiments and to Mrs I. Durant for technical assistance. We thank Mme Laurence Costes and Mr William Cazenave for technical assistance, Mr S. Geoffre for peptide synthesis and Mr J. C. Huet and J. C. Pernollet for peptide sequencing. This work was supported by grants from the Wellcome Trust and the BBSRC to J.A.T.D. and S.A.D., and from CNRS and Région Aquitaine to the French group.

Bennett
,
H. P. J.
,
Browne
,
C. A.
and
Solomon
,
S.
(
1981
).
Purification of two major forms of rat pituitary corticotropin using only reversed-phase liquid chromatography
.
Biochemistry
20
,
4530
4538
.
Cantera
,
R.
,
Hansson
,
B. S.
,
Hallberg
,
E.
and
Nässel
,
D. R.
(
1992
).
Postembryonic development of leucokinin I-immunoreactive neurons innervating a neurohemal organ in the turnip moth Agrotis segetum
.
Cell Tissue Res
.
269
,
65
77
.
Chen
,
Y. T.
,
Veenstra
,
J. A.
,
Davis
,
N. T.
and
Hagedorn
,
H. H.
(
1994
).
A comparative study of leucokinin-immunoreactive neurons in insects
.
Cell Tissue Res
.
276
,
69
83
.
Coast
,
G. M.
,
Holman
,
G. M.
and
Nachman
,
R. J.
(
1990
).
The diuretic activity of a series of cephalomyotropic neuropeptides, the achetakinins, on isolated Malpighian tubules of the house cricket, Acheta domesticus
.
J. Insect Physiol
.
36
,
481
488
.
Cox
,
K. J.
,
Tensen
,
C. P.
,
Van der Schors
,
R. C.
,
Li
,
K. W.
,
van Heerikhuizen
,
H.
,
Vreugdenhil
,
E.
,
Geraerts
,
W. P.
and
Burke
,
J. F.
(
1997
).
Cloning, characterization and expression of a G-protein-coupled receptor from Lymnaea stagnalis and identification of a leucokinin-like peptide, PSFHSWSamide, as its endogenous ligand
.
J. Neurosci
.
17
,
1197
1205
.
Davies
,
S. A.
,
Huesmann
,
G. R.
,
Maddrell
,
S. H. P.
,
O’Donnell
,
M. J.
,
Skaer
,
N. J. V.
,
Dow
,
J. A. T.
and
Tublitz
,
N. J.
(
1995
).
CAP2b, a cardioacceleratory peptide, is present in Drosophila and stimulates tubule fluid secretion via cGMP
.
Am. J. Physiol
.
269
,
R1321
R1326
.
Davies
,
S. A.
,
Kelly
,
D. C.
,
Goodwin
,
S. F.
,
Wang
,
S.-Z.
,
Kaiser
,
K.
and
Dow
,
J. A. T.
(
1996
).
Analysis and inactivation of vha55, the gene encoding the V-ATPase B-subunit in Drosophila melanogaster, reveals a larval lethal phenotype
.
J. Biol. Chem
.
271
,
30677
30684
.
Davies
,
S. A.
,
Stewart
,
E. J.
,
Heusmann
,
G. R.
,
Skaer
,
N. J. V.
,
Maddrell
,
S. H. P.
,
Tublitz
,
N. J.
and
Dow
,
J. A. T.
(
1997
).
Neuropeptide stimulation of the nitric oxide signalling pathway in Drosophila melanogaster Malpighian tubules
.
Am. J. Physiol
.
273
,
R823
R827
.
Dow
,
J. A. T.
,
Davies
,
S. A.
and
Sözen
,
M. A.
(
1998
).
Fluid secretion by the Drosophila melanogaster Malpighian tubule
.
Am. Zool
.
38
,
450
460
.
Dow
,
J. A. T.
,
Kelly
,
D. C.
,
Davies
,
S. A.
,
Maddrell
,
S. H. P.
and
Brown
,
D.
(
1995
).
A novel member of the Major Intrinsic Protein family in Drosophila – are aquaporins involved in insect Malpighian (renal) tubule fluid secretion?
J. Physiol., Lond
.
489
,
P110
P111
.
Dow
,
J. A. T.
,
Maddrell
,
S. H. P.
,
Davies
,
S.-A.
,
Skaer
,
N. J. V.
and
Kaiser
,
K.
(
1994a
).
A novel role for the nitric oxide/cyclic GMP signalling pathway: the control of fluid secretion in Drosophila
.
Am. J. Physiol
.
266
,
R1716
R1719
.
Dow
,
J. A. T.
,
Maddrell
,
S. H. P.
,
Görtz
,
A.
,
Skaer
,
N. V.
,
Brogan
,
S.
and
Kaiser
,
K.
(
1994b
).
The Malpighian tubules of Drosophila melanogaster: a novel phenotype for studies of fluid secretion and its control
.
J. Exp. Biol
.
197
,
421
428
.
Hayes
,
T. K.
,
Holman
,
G. M.
,
Pannabecker
,
T. L.
,
Wright
,
M. S.
,
Strey
,
A. A.
,
Nachman
,
R. J.
,
Hoel
,
D. F.
,
Olson
,
J. K.
and
Beyenbach
,
K. W.
(
1994
).
Culekinin depolarizing peptide: a mosquito leucokinin-like peptide that influences insect Malpighian tubule ion transport
.
Regul. Peptides
52
,
235
248
.
Hayes
,
T. K.
,
Pannabecker
,
T. L.
,
Hinckley
,
D. J.
,
Holman
,
G. M.
,
Nachman
,
R. J.
,
Petzel
,
D. H.
and
Beyenbach
,
K. W.
(
1989
).
Leucokinins, a new family of ion transport stimulators and inhibitors in insect Malpighian tubules
.
Life Sci
.
44
,
1259
1266
.
Holman
,
G. M.
,
Cook
,
B. J.
and
Nachman
,
R. J.
(
1986a
).
Isolation, primary structure and synthesis of two neuropeptides from Leucophaea maderae – members of a new family of cephalomyotropins
.
Comp. Biochem. Physiol
.
84C
,
205
211
.
Holman
,
G. M.
,
Cook
,
B. J.
and
Nachman
,
R. J.
(
1986b
).
Primary structure and synthesis of two additional neuropeptides from Leucophaea maderae – members of a new family of cephalomyotropins
.
Comp. Biochem. Physiol
.
84C
,
271
276
.
Holman
,
G. M.
,
Cook
,
B. J.
and
Nachman
,
R. J.
(
1987a
).
Isolation, primary structure and synthesis of Leukokinin V and Leukokinin VI – myotropic peptides of Leucophaea maderae
.
Comp. Biochem. Physiol
.
88C
,
27
30
.
Holman
,
G. M.
,
Cook
,
B. J.
and
Nachman
,
R. J.
(
1987b
).
Isolation, primary structure and synthesis of Leukokinin VII and Leukokinin VIII – the final members of this new family of cephalomyotropic peptides isolated from head extracts of Leucophaea maderae
.
Comp. Biochem. Physiol
.
88C
,
31
34
.
Nachman
,
R. J.
,
Coast
,
G. M.
,
Holman
,
G. M.
and
Beier
,
R. C.
(
1995
).
Diuretic activity of C-terminal group analogs of the insect kinins in Acheta domesticus
.
Peptides
16
,
809
813
.
Nässel
,
D. R.
and
Lundquist
,
C. T.
(
1991
).
Insect tachykinin-like peptide: distribution of leucokinin immunoreactive neurons in the cockroach and blowfly brains
.
Neurosci. Lett
.
130
,
225
228
.
Nässel
,
D. R.
,
Lundquist
,
C. T.
and
Brodin
,
E.
(
1992
).
Diversity in tachykinin-like peptides in the insect brain
.
Acta Biol. Hung
.
43
,
175
188
.
Nieto
,
J.
,
Veelaert
,
D.
,
Derua
,
R.
,
Waelkens
,
E.
,
Cerstiaens
,
A.
,
Coast
,
G.
,
Devreese
,
B.
,
Van Beeumen
,
J.
,
Calderon
,
J.
,
De Loof
,
A.
and
Schoofs
,
L.
(
1998
).
Identification of one tachykinin- and two kinin-related peptides in the brain of the white shrimp, Penaeus vannamei
.
Biochem. Biophys. Res. Commun
.
248
,
406
411
.
O’Donnell
,
M. J.
,
Dow
,
J. A. T.
,
Huesmann
,
G. R.
,
Tublitz
,
N. J.
and
Maddrell
,
S. H. P.
(
1996
).
Separate control of anion and cation transport in Malpighian tubules of Drosophila melanogaster
.
J. Exp. Biol
.
199
,
1163
1175
.
O’Donnell
,
M. J.
,
Rheault
,
M. R.
,
Davies
,
S. A.
,
Rosay
,
P.
,
Harvey
,
B. J.
,
Maddrell
,
S. H. P.
,
Kaiser
,
K.
and
Dow
,
J. A. T.
(
1998
).
Hormonally-controlled chloride movement across Drosophila tubules is via ion channels in stellate cells
.
Am. J. Physiol
.
43
,
R1039
R1049
.
Pannabecker
,
T. L.
,
Hayes
,
T. K.
and
Beyenbach
,
K. W.
(
1993
).
Regulation of epithelial shunt conductance by the peptide leucokinin
.
J. Membr. Biol
.
132
,
63
76
.
Rosay
,
P.
,
Davies
,
S. A.
,
Yu
,
Y.
,
Sözen
,
M. A.
,
Kaiser
,
K.
and
Dow
,
J. A. T.
(
1997
).
Cell-type specific monitoring of intracellular calcium in Drosophila using an aequorin transgene
.
J. Cell Sci
.
110
,
1683
1692
.
Rubin
,
G. M.
(
1988
).
Drosophila melanogaster as an experimental organism
.
Science
240
,
1453
1459
.
Schmid
,
A.
and
Becherer
,
C.
(
1996
).
Leucokinin-like immunoreactive neurones in the central nervous system of the spider Cupiennius salei
.
Cell Tissue Res
.
284
,
143
152
.
Schoofs
,
L.
,
Holman
,
G. M.
,
Proost
,
P.
,
Vandamme
,
J.
,
Hayes
,
T. K.
and
De Loof
,
A.
(
1992
).
Locustakinin, a novel myotropic peptide from Locusta migratoria; isolation, primary structure and synthesis
.
Regul. Peptides
37
,
49
57
.
Smart
,
D.
,
Johnston
,
C. F.
,
Shaw
,
C.
,
Halton
,
D. W.
and
Buchanan
,
K. D.
(
1993
).
Use of specific antisera for the localisation and quantitation of leucokinin immunoreactivity in the nematode, Ascaris suum
.
Comp. Biochem. Physiol
.
106C
,
517
522
.
Sözen
,
M. A.
,
Armstrong
,
J. D.
,
Yang
,
M. Y.
,
Kaiser
,
K.
and
Dow
,
J. A. T.
(
1997
).
Functional domains are specified to single-cell resolution in a Drosophila epithelium
.
Proc. Natl. Acad. Sci. USA
94
,
5207
5212
.
Veenstra
,
J. A.
(
1994
).
Isolation and identification of three leucokinins from the mosquito Aedes aegypti
.
Biochem. Biophys. Res. Commun
.
202
,
715
719
.
Veenstra
,
J. A.
(
1998
).
Isolation and identifcation of three RFamide-immunoreactive peptides from the mosquito Aedes aegypti
.
Peptides
20
,
31
38
.
Veenstra
,
J. A.
and
Lambrou
,
G.
(
1995
).
Isolation of a novel RFamide peptide from the midgut of the American cockroach, Periplaneta americana
.
Biochem. Biophys. Res. Commun
.
213
,
519
524
.
Veenstra
,
J. A.
,
Noriega
,
F. G.
,
Graf
,
R.
and
Feyereisen
,
R.
(
1997a
).
Identification of three allatostatins and their cDNA from the mosquito Aedes aegypti
.
Peptides
18
,
937
942
.
Veenstra
,
J. A.
,
Pattillo
,
J. M.
and
Petzel
,
D. H.
(
1997b
).
A single cDNA encodes all three Aedes leucokinins, which stimulate both fluid secretion by the Malpighian tubules and hindgut contractions
.
J. Biol. Chem
.
272
,
10402
10407
.
Watson
,
S.
and
Arkinstall
,
S.
(
1994
).
The G-protein Linked Receptor Facts Book
.
London
:
Academic Press
.