V-ATPases are phylogenetically widespread, highly conserved, multisubunit proton pumps. Originally characterised in endomembranes, they have been found to energise transport across plasma membranes in a range of animal cells and particularly in certain epithelia. While yeast is the model of choice for the rapid generation and identification of V-ATPase mutants, it does not allow their analysis in a plasma membrane context. For such purposes, Drosophila melanogaster is a uniquely suitable model. Accordingly, we have cloned and characterised genes encoding several V-ATPase subunits in D. melanogaster and, using P-element technology, we have succeeded in generating multiple new alleles. Reporter gene constructs reveal ubiquitous expression, but at particularly high levels in those epithelia thought to be energised by V-ATPases, and several of the alleles have lethal recessive phenotypes characterised by epithelial dysfunction. These results, while providing the first gene knockouts of V-ATPases in animals, also illustrate the general utility of D. melanogaster as a model for the genetic analysis of ion transport and its control in epithelia.

Ion-transporting ATPases fall into three main classes (Pedersen and Carafoli, 1987a,b): F-type, multisubunit H+-motive pumps found on membranes of mitochondria, chloroplasts and bacteria, and normally driven in reverse as ATP synthases; P-type, cation pumps with single catalytic subunits (and frequently with accessory subunits) found on plasma membranes and some endomembranes, and typified by the Na+/K+-ATPase; and V-type, or ‘vacuolar’ ATPases, which transport H+ across eukaryotic endomembranes and which, more recently, have been found to energise ion transport across plasma membranes (reviewed in Gluck, 1992).

Insects are conspicuous exemplars of higher organisms in which epithelial function is energised by a proton, rather than a Na+, gradient (Harvey, 1992b; Wieczorek and Harvey, 1995). The evidence assembled is based on the failure to purify a primary electrogenic K+-ATPase from any of the many insect epithelia which displayed active transport of K+; on the demonstration that K+-transporting activity was concentrated in the apical membranes of the goblet cavity of larval lepidopteran midgut (Dow et al. 1984; Dow and Peacock, 1989); on the biochemical purification of this membrane (Harvey et al. 1983a); on the demonstration that the major ATPase of this membrane was in fact an H+-ATPase (Wieczorek et al. 1991); and on the subsequent demonstration that apical plasma membranes of all those epithelia in which the classical ‘K+ pump’ had been invoked contained V-ATPases, by immunocytochemistry, by direct protein biochemistry or by enzymology (Klein et al. 1991; Wieczorek, 1992). These results led to the development of the ‘Wieczorek model’ for insect epithelia, that active transport of alkali metal cations (i.e. Na+ or K+) is the product of a primary apical plasma membrane H+-motive V-ATPase that drives one or more alkali metal/H+ exchangers (Wieczorek, 1992). This view is now accepted as orthodoxy, because insect epithelia are found to be exquisitely sensitive to both bafilomycin (a specific H+ V-ATPase inhibitor) and amiloride (a blocker of Na+ channels or exchangers) (Bertram et al. 1991; Dow et al. 1994a; Maddrell and O’Donnell, 1992). Additionally, interference with the alkali metal transport pathway (by amiloride application, Ba2+ blockade of basal K+ entry or depletion of K+ from the bathing medium) unmasks the apical H+-ATPase by inducing an apical acidification (Bertram et al. 1991; Dow, 1992; Dow et al. 1994a; Maddrell and O’Donnell, 1992; Weltens et al. 1992). The concept that epithelia can be energised by a proton-motive, rather than a Na+-motive, force can be extended to several vertebrate epithelia, such as the kidney (Gluck and Nelson, 1992) and frog skin (Harvey, 1992a).

While this model is widely accepted, some problems remain. The exchanger remains enigmatic, largely due to the failure to clone any insect gene with homology to the classical vertebrate Na+/H+ exchanger. There may be a single promiscuous exchanger that can accept either Na+ or K+, or there may be separate exchangers for each metal ion species; the latter possibility would allow a straightforward way of explaining the apparent differences in alkali metal ion specificity of different insect epithelia. The Wieczorek model proposes that, in midgut, the exchanger may be electrophoretic (Azuma et al. 1995), and this may directly explain the remarkably high pH values encountered in larval lepidopteran gut lumen (Dow, 1984, 1992); however, there is evidence in other insect epithelia that any exchanger is electroneutral (van Kerkhove, 1994), and the existence of any exchanger has been challenged even in midgut (Küppers and Bunse, 1996).

A heretical alternative explanation, based on the insensitivity of electrical measurements of the trichogen sensilla of insects to amiloride or harmaline, is that there is no exchanger and that the apical V-ATPase is primarily an H+-ATPase, but with the additional ability to transport alkali metal cations (Küppers and Bunse, 1996). Given that the intracellular pH is 7, and that intracellular [K+] is approximately 100 mmol l−1, even if the pump were 105:1 selective in favour of H+ over K+, under normal conditions the two ion species would be transported at nearly equal rates. Classically, the assignment of both F-ATP synthases and V-ATPases as purely H+-motive would argue against so radical a model; however, recent evidence suggests that, at least in some species, both F-ATP synthases and V-ATPases may have significant alkali metal transporting ability (Takase et al. 1994), and that the ion-specificity of the ATPase may be switched by mutagenesis of the proteolipid subunit (Zhang and Fillingame, 1995). If this were true, then, by selective targetting of appropriate subunits, an organism might have both a proton-motive endosomal V-ATPase and a cation-motive plasma membrane V-ATPase in its epithelia. Against this, however, it should be noted that the proteolipid subunits of both Manduca sexta and Drosophila melanogaster show no divergence from those of other species in which the identity of the transported ion is not in doubt (Dow et al. 1992; Meagher et al. 1990), and so it is unlikely that these subunits could support a significant flux of ions other than protons.

The big questions

From the preceding arguments, it can be seen that there are several interesting questions which need to be addressed, both in insects and more generally: are the plasma membrane V-ATPases different from those of endomembranes, in terms of expression levels, transcript choice or functional properties, and are such differences functionally important?

How can the roles of plasma membrane be dissociated from those of endomembranes? To study such problems, we require a system in which we can selectively manipulate gene expression in the intact organism. Although yeast is the preferred model for straightforward structure/function analysis of V-ATPases (Nelson and Nelson, 1990), it is not a suitable model for analysis of a plasma membrane function. Clearly, a transgenic animal model is required. There are three species for which genetics is sufficiently advanced; the mouse, the zebrafish and D. melanogaster. The mouse model allows elegant and precise disruption of a gene by homologous recombination, but the ability to intervene in a particular tissue, or at a particular phase of the life-cycle, remains poorly developed. It is thus a poor model for reverse genetics of most of the genes of interest in transport physiology, precisely because of their ubiquity: we can predict that many of these genes will be essential for organismal function, and so the phenotype of disruptant homozygotes will be lethal at a stage too early to permit detailed physiological analysis. Although the zebrafish holds considerable promise for the future, the genetics of the system is still underdeveloped and there is very little transport physiology to permit a phenotypic assay. By contrast, there is a wealth of transport physiological data for insects, although D. melanogaster ranks as one of the smallest yet studied.

Drosophila melanogaster as a model system

D. melanogaster is the higher organism with perhaps the best balance of structural complexity and genetic tractability available at present (Rubin, 1988). Indeed, the power of this model in producing information of general significance is illustrated by two famous examples; the developmental (Lawrence, 1993) and behavioural (Davis and Dauwalder, 1991; Tully, 1991) phenotypes. The advantages of D. melanogaster are as follows: (i) a short generation time and high fecundity, allowing crosses to be made and analysed within 2 weeks; (ii) the ease and relatively low cost of rearing, which permit large-scale genetic experimentation; (iii) the relatively small genome, such that a given recombination interval corresponds to a far smaller (and thus more tractable) piece of DNA than would be the case in mouse or human; (iv) the existence of a huge body of classical genetics, which gives an excellent density of mapped, visible markers (such as eye colour and bristle length) for use in experimentation; (v) the reconciliation of these data into a genome databank, which allows new data to be reconciled easily with previously existing data; (vi) the development of synthetic transposons (particularly those based on the P-element), which allow targetted mutagenesis of desired genes, the incorporation of transgenes into the genome at high efficiency and the identification of potentially interesting loci by enhancer-trapping (Kaiser, 1993; Kaiser and Goodwin, 1990; Rubin and Spradling, 1982).

In the face of such overwhelming advantages, it should be noted that there are deficiencies in D. melanogaster as a physiological model for the study of transport genes such as V-ATPases. These are (i) the lack of a pre-existing phenotype for studies of mutants, (ii) the extremely small size of the organism, which makes it difficult to perform such physiological analysis, and (iii) the different body architecture of invertebrates compared with vertebrates, which requires experiments to be carefully designed if they are to yield results of general significance. Nonetheless, we have been able to develop in the last few years a model which allows the detailed analysis of transport or cell signalling genes in a physiological context that is strongly reminiscent of several vertebrate epithelia.

The insect Malpighian tubule performs a function loosely analogous to that of the vertebrate kidney tubule. Despite its small size (a linear array of just 121 principal cells), the D. melanogaster tubule is remarkably robust and provides a valuable physiological phenotype (Dow et al. 1994a). Although originally developed as a screen for V-ATPase mutations because of the known sensitivity of insect tubules to bafilomycin (Dow et al. 1994a), it has also proved a valuable prototype for an epithelium controlled by nitric oxide and cyclic GMP (Dow et al. 1994b) and for the study of neuropeptides such as the leucokinins and CAP2b (Davies et al. 1995; O’Donnell et al. 1996). Potentially, then, the D. melanogaster Malpighian tubule may prove a useful tool for the study of plasma membrane V-ATPase function.

We have embarked on a characterisation of V-ATPase genes in D. melanogaster. In addition to those cloned explicitly by our group, two further subunits have been cloned serendipitously, one from an enhancer-trap study (Harvie and Bryant, 1996) and one from a yeast two-hybrid study of cytoskeletal proteins (He and Kramer, 1996). This means that, in terms of molecular characterisation, D. melanogaster V-ATPase approaches that of human, yeast and Manduca sexta in the number of subunits characterised (Table 1).

Table 1.

Characterisation of Drosophila melanogaster genes encoding V-ATPase subunits

Characterisation of Drosophila melanogaster genes encoding V-ATPase subunits
Characterisation of Drosophila melanogaster genes encoding V-ATPase subunits

In principle, the identification of a chromosomal localisation for a gene in D. melanogaster unlocks a wealth of published data, as effectively all known mutations have been mapped by recombination and documented as part of a visionary genome project that has run from the early decades of this century. It is thus possible that, when a novel genetic locus is characterised, there may be existing mutations in the region that merit further analysis as candidate alleles of the gene under study. Over the last few years, the probability of such findings has been increased greatly by the systematic physical mapping of the genome, the production of comprehensive panels of thousands of lines carrying lethal P-element insertions, which must presumably have inactivated a large number of essential genes (Török et al. 1993), and the development of economical screening strategies that allow the easy identification of candidate lines for a particular genes (Guo et al. 1996a).

Our V-ATPase work has been greatly assisted by this integration of work from the D. melanogaster community, and we now have a panel of flies in which V-ATPase subunits have been inactivated (Table 1). Although the discovery of a V-ATPase C-subunit mutant was serendipitous (Harvie and Bryant, 1996), we have identified mutant alleles of the A, B, c and E subunit by screening lethal P-element insertions which had been mapped previously to the same genetic interval by in situ hybridisation (Davies et al. 1996; Guo et al. 1996a). Although such mutations might lie anywhere within approximately 200 kb of the target gene, we have been fortunate in identifying at least one line for each subunit in which a P-element is situated either within an intron, or directly 5´ to the start of the gene.

In one case, the B-subunit, this lethal insertion has proved invaluable in demonstrating that vha55, the gene encoding the subunit, is synonymous with l(3)87Ca (Davies et al. 1996), a lethal locus whose detailed characterisation (Gausz et al. 1979) predates the discovery of V-ATPases by a decade. It is thus the first knockout of a V-ATPase in an animal.

What phenotype could be expected for V-ATPase mutations?

Our working model is that there are two distinct roles for V-ATPases in animals: the classical endomembrane role first characterised in plants and fungi, and a plasma membrane role in specialised cell types and epithelia. There may be differences in subunit composition or splice site choices in at least some subunits of the holoenzyme, and there is evidence in vertebrates that B-subunit transcript choice influences the choice of other subunits. There is also a more general difference; while only a few V-ATPase holoenzymes may suffice to acidify a given endomembrane compartment to the desired level, V-ATPase activity is the primary driving force, and thus the rate-limiting factor, in epithelial function. We thus expect to find a far higher holoenzyme level on plasma membranes of epithelia energised by V-ATPases than on any endomembrane; and indeed this is the case. V-ATPases form effectively close-packed arrays of ‘portasomes’, as seen by electron microscopy of most insect epithelia (Harvey et al. 1983b). This property is also shared by the plasma membrane of kidney intercalated cells (Brown et al. 1992).

Given that V-ATPase levels are critical to epithelial function, it is reasonable to suppose that an epithelial phenotype might be one of the earlier manifestations of a V-ATPase null phenotype.

The SzA tubule phenotype

Although there is no detectable heterozygous phenotype of any V-ATPase mutation we have so far characterised in D. melanogaster, there is a homozygous phenotype manifest in tubules. This was first described by Gausz et al. (1979) in some alleles of the SzA, or l(3)87Ca, lethal complementation group at 87C, which we have subsequently shown to be synonymous with vha55 (Davies et al. 1996). Classically, those mutant homozygotes that survived to late embryonic or early larval stages showed transparent Malpighian tubules, without the luminal white material observed in healthy larvae. The phenotype was variable, and not all lethal alleles displayed it. If this phenotype is indeed associated with a defective V-ATPase B-subunit, then it is reasonable to suppose that defects in other V-ATPase subunits would be equally effective in eliciting the phenotype. Indeed, we have been able to replicate this phenotype with other alleles of vha68-2, one of the genes encoding the A-subunit (Fig. 1). Taken together, these results suggest that genetic intervention, not just in V-ATPase subunit genes but in any gene essential for plasma membrane V-ATPase function, is likely to show this characteristic phenotype, which may thus constitute a general screen for such genes.

Fig. 1.

The SzA tubule phenotype is replicated by a lethal allele of vha68-2, a gene encoding an A-subunit of V-ATPase. The vha68-2 homozygote is shown on the left; a heterozygous sib is shown on the right. In transmitted light, the relatively poor development of luminal insoluble material in the mutant tubules is clear.

Fig. 1.

The SzA tubule phenotype is replicated by a lethal allele of vha68-2, a gene encoding an A-subunit of V-ATPase. The vha68-2 homozygote is shown on the left; a heterozygous sib is shown on the right. In transmitted light, the relatively poor development of luminal insoluble material in the mutant tubules is clear.

We interpret this phenotype as a failure to acidify the tubule luminal contents, thus impeding the precipitation either of uric acid or of the mineral concretions that accumulate in the tubule lumen.

An important line of evidence for differences between plasma membrane and endomembrane V-ATPases is the study of cell-specific expression patterns. As described above, it has already been established by immunocytochemistry, in both insects and vertebrates, that plasma membranes of epithelia energised by V-ATPases contain far higher levels of protein than endosomal membranes of cells in general (Klein, 1992; Nelson et al. 1992). Is this upregulation in epithelia through higher levels of expression of the gene, enhanced stability of the mRNA or enhanced stability of the protein product? There is now evidence to support the first of these models. Synthetic P-elements contain the lacZ reporter gene downstream of a very weak,‘permissive’ promoter. A P-element inserted near a given gene (they have a tendency to insert non-randomly either upstream or in early introns) will thus fall under the influence of the control elements (promoters and enhancers) that control expression of that gene. lacZ expression can thus be taken as at least an approximate estimate of the rate of transcription of the gene; and as it is an exogenous gene, its expression will not be affected by post-transcriptional tissue-specific effects that might modulate the longevity of the neighbouring mRNA. Such reporter expression patterns can thus distinguish between the first two possibilities outlined above.

P-element insertions in either vha68-2, one of at least two genes encoding A-subunits (Y. Guo, in preparation), or in vha55, the single gene encoding a B-subunit (Davies et al. 1996), both show that transcription is selectively upregulated in epithelial tissues known to be energised by V-ATPases (Fig. 2). Enhanced expression is first detected in embryos, in line with the embryonic lethal phenotype elicited by severe disruptants (Davies et al. 1996).

Fig. 2.

lacZ reporter gene expression in a P-element insertional mutant of vha55. (A) Early embryo; (B) later embryo; (C) Malpighian tubules; (D) rectum; (E) antennal palps; (F) oviduct. (G) Higher-power view of Malpighian tubule main segment reporter expression, showing that expression is confined to the larger principal cell nuclei. (H) Enhancer trap line c724 stained with fluorescent anti-β-galactosidase to reveal stellate cells and counterstained with ethidium bromide to reveal nuclei. A–F are reproduced from Davies et al. (1996) and G–H are reproduced from Sozen, M. A., Armstrong, J. D., Yang, M.-Y., Kaiser, K. and Dow, J. A. T., in preparation.

Fig. 2.

lacZ reporter gene expression in a P-element insertional mutant of vha55. (A) Early embryo; (B) later embryo; (C) Malpighian tubules; (D) rectum; (E) antennal palps; (F) oviduct. (G) Higher-power view of Malpighian tubule main segment reporter expression, showing that expression is confined to the larger principal cell nuclei. (H) Enhancer trap line c724 stained with fluorescent anti-β-galactosidase to reveal stellate cells and counterstained with ethidium bromide to reveal nuclei. A–F are reproduced from Davies et al. (1996) and G–H are reproduced from Sozen, M. A., Armstrong, J. D., Yang, M.-Y., Kaiser, K. and Dow, J. A. T., in preparation.

Expression is also detected in epithelia known to be energised by V-ATPases, the Malpighian tubules, the antennal palps (rich in tormogen cells) and the rectum (Fig. 2). Expression is also elevated in the female oviduct and accessory glands (Fig. 2F), not a tissue in which V-ATPase action had previously been implicated: although plasma membrane V-ATPases have been shown to play an important part in regulating the pH of the male reproductive tract, and thus of sperm motility, in vertebrates (Breton et al. 1996).

Recently, physiological analysis of the Malpighian tubules of D. melanogaster has shown that they are stimulated by each of the second messengers cyclic AMP, cyclic GMP and Ca2+ (Davies et al. 1995; Dow et al. 1994a,b). These act independently, with cyclic nucleotides stimulating an apical V-ATPase, whereas Ca2+ acts to elevate Cl permeability. This separate control of cation and anion transport led to the speculation that these transport processes might also be separately located in the different cell subtypes of the tubule (O’Donnell et al. 1996). The enhancer detector mapping of V-ATPase expression levels described above (Davies et al. 1996) has recently provided the first demonstration of such segregation; elevated V-ATPase expression is unique to the principal cells of the main segment, and not to the stellate, secondary cells (Sozen et al. 1996). Cation transport is thus confined to a genetically defined population of principal cells in this epithelium, which would allow the stellate cells, in turn, to be specialised for transport of anions and perhaps water (Dow et al. 1995).

An intriguing byproduct of the Drosophila genetic approach as that a gene may map to the same position as some, apparently quite unrelated, known locus. This proves to be the case for vha55, the gene encoding the B-subunit, which maps to the same interval (87C on chromosome 3) as the far more fashionable gene Polycomb. Products of the Polycomb (Pc) group of genes are thought to act cooperatively to elicit transcriptional silencing of genomic DNA in response to the initial transcription levels of homeotic genes in the embryo by binding untranscribed DNA and packing it into heterochromatin (Moehrle and Paro, 1994). They thus help to preserve the ‘memory’ of the homeotic selector genes by shielding inactivated genes from later access by transcription factors that might otherwise tend to activate them ectopically. The mutant phenotypes of any of the Pc group (ectopic expression of sex combs on T2 and T3 legs of males) thus resemble a homeotic misassignment of T2 and T3 legs to a T1 fate. This model for transcriptional silencing may have general significance for the long-term inactivation of genes during development. A previous screen for dosage-dependent suppressors or enhancers of Pc had identified a number of homeotic genes, together with several that were mapped but unidentified (Kennison and Tamkun, 1988). One of the powerful suppressing loci was shown to be allelic to l(3)87Ca (Kennison and Tamkun, 1988), a locus that we have now demonstrated to be vha55 (Davies et al. 1996). This implies that V-ATPases might play a role in transcriptional silencing. This finding could reflect some specific but uninformative interaction between these two mutant proteins, it could reflect a general sensitivity of Pc to the V-ATPase status of the insect, or it could reflect some unusual genetic event involving 87C, unrelated to the role of l(3)87Ca as a gene encoding a V-ATPase subunit. For example, most of the loci previously identified as modulating Pc were known homeotic genes (Kennison and Tamkun, 1988), and at 87C, vha55 is placed between the two major homeotic clusters on chromosome 3. It is thus conceivable that some alleles that inactivate the V-ATPase might also destroy some exceptionally long-range control element of a homeotic gene in one of the neighbouring clusters, although the distance makes this unlikely.

The finding that the effects of mutations in Polycomb can be modulated by hemizygosity for a gene encoding a proton pump subunit is quite unexpected and raises the general question of whether the V-ATPase holoenzyme or particular subunits may serve multiple roles, in addition to straightforward endosomal acidification. This proves to be the case, and an increasing number of cellular processes are being shown to be dependent on V-ATPase function.

V-ATPase subunits may affect holoenzyme assembly either directly or by interaction with the cytoskeleton. The B-subunit can be found preferentially attached to early endosomal vesicles of kidney collecting duct, in the absence of other major V-ATPase subunits, suggesting an independent role in vesicle anchoring or trafficking (Sabolic et al. 1992). Similarly, preliminary data from a yeast two-hybrid screen implicates the D. melanogaster D-subunit in binding to the cytoskeletal protein hook (He and Kramer, 1996). If the V-ATPase interacts directly with the cytoskeleton, it could play a major active role in vesicle trafficking and in the sorting processes that govern the relative acidity of different endosomal compartments.

Recently, several emerging roles of V-ATPases in processes that may have developmental implications could offer additional explanations of this interaction. One is the high-affinity binding of oncoproteins, such as bovine papillomavirus E5 (Goldstein et al. 1991) or human lymphotrophic leukaemia virus type I P12 proteins (Koralnik et al. 1995), to the transmembrane proteolipid subunit of the V-ATPase (Koralnik et al. 1995). Binding may reduce the efficiency of dissociation of a growth factor (probably platelet-derived growth factor or epidermal growth factor) from its receptor in endosomes by inhibiting V-ATPases and thus raising endosomal pH. The growth factor receptor would then signal occupancy for longer than normal and so predispose a cell to inappropriate growth. The sensing of the gradients of D. melanogaster segmentation gene products may involve analogous ligand–receptor interactions. A slightly underactive endosomal compartment would then cause the position of the cell in such a gradient to be misread, resulting in subsequent mis-setting of homeotic selector genes, and thus explain an interaction with Pc.

V-ATPases have also been implicated in the regulation of cytoplasmic pH, a putative signalling system. Pharmacological blockade of V-ATPases in culture increases susceptibility to apoptosis (Nishihara et al. 1995), whereas granulocyte colony-stimulating factor delays apoptotic cell death in neutrophils by upregulating V-ATPase (Gottlieb et al. 1995), suggesting that downregulation of V-ATPases and concomitant cytoplasmic adicification may be an integral part of the apoptotic machinery. Inhibition of apoptotic processes in development could lead to survival of cells expressing inappropriate gene products.

V-ATPases have also been found to co-localise with calcineurin, an important Ca2+-sensitive phosphatase; and, in yeast, mutations in calcineurin or other Ca2+ regulating genes and in V-ATPase have been shown to have synthetic lethality, suggesting an important role for V-ATPases in regulating intracellular Ca2+ (Garrettengele et al. 1995; Tanida et al. 1995).

Another possibility is that a slightly higher than normal endosomal pH would produce aberrant post-translational processing of genes not normally associated with ion transport; differential endocytic processing of wild-type and mutant β-amyloid is observed when V-ATPase activity is reduced by the toxin bafilomycin-A1 (Haass et al. 1995).

It is also possible that a human genetic disease may be associated with haploabnormality for a V-ATPase gene. Recently, cat eye syndrome, a disease characterised by multifaceted developmental abnormalities, has been shown to be caused by a supernumerary minute ring chromosome, corresponding to the centromeric region of chromosome 22 which contains the gene encoding the V-ATPase subunit-E (Baud et al. 1994). The E-subunit probe is the most distal marker duplicated in all patients screened to date (Mears et al. 1995), suggesting that it might be involved in the pathogenesis of the syndrome.

The proteolipid c-subunit of the V-ATPase has also been argued to play an essential role in, if not form the pore of, gap junctions, at least insects (Finbow et al. 1995), a process with wide-ranging significance both in development and in adult function of all organisms, and in the ‘mediatophore’ (vesicle fusion apparatus) of Torpedo marmorata (Leroy and Meunier, 1995). Evidence for this proposal includes the purification of a 17 kDa integral membrane protein, with peptide sequence indistinguishable from V-ATPase proteolipid, from several invertebrate gap junctional preparations (Dow et al. 1992; Finbow et al. 1992; Finbow and Pitts, 1993). This view is heretical, but in its support is the patent failure to clone invertebrate homologues of connexins, despite their close conservation across other phyla. In rat embryos, proteolipid gene expression is transiently elevated at sites of mesenchymal/epithelial interactions, in a pattern resembling that of connexin 43 (Numata et al. 1995). Unusually, this model invokes membrane insertion in both possible orientations (reviewed elsewhere, Dow, 1994); however, it is prudent to allow the possibility that V-ATPases may not be passengers in membrane fusion, cycling and communication events, but essential participants.

The results presented above lead to the slightly suprising conclusion that, despite its small size, D. melanogaster may be the ideal model for the multidisciplinary investigation of V-ATPases and other transport-or signalling-related genes in an organismal context. The advanced genetics unique to D. melanogaster can be combined with a novel physiological phenotype to allow interesting experiments in integrative physiology to be designed. Taken together with the evidence presented here for a great richness in roles for the V-ATPase, both in D. melanogaster and in other organisms, there is great potential in this synthesis of experimental technologies.

This work was supported by MRC Grant G9120579CB, a Nuffield Foundation Research Fellowship and a Royal Society Grant to J.A.T.D., and by general Funds of the University of Glasgow.

Azuma
,
M.
,
Harvey
,
W. R.
and
Wieczorek
,
H.
(
1995
).
Stoichiometry of K+/H+ antiport helps to explain extracellular pH 11 in a model epithelium
.
FEBS Lett
.
361
,
153
156
.
Baud
,
V.
,
Mears
,
A. J.
,
Lamour
,
V.
,
Scamps
,
C.
,
Duncan
,
A.
,
Mcdermid
,
H. E.
and
Lipinski
,
M.
(
1994
).
The E subunit of vacuolar H+-ATPase localizes close to the centromere on human chromosome 22
.
Human molec. Genetics
3
,
335
339
.
Bertram
,
G.
,
Shleithoff
,
L.
,
Zimmermann
,
P.
and
Wessing
,
A.
(
1991
).
Bafilomycin-A1 is a potent inhibitor of urine formation by Malpighian tubules of Drosophila hydei – is a vacuolar-type ATPase involved in ion and fluid secretion?
J. Insect Physiol
.
37
,
201
209
.
Breton
,
S.
,
Smith
,
P. J. S.
,
Lui
,
B.
and
Brown
,
D.
(
1996
).
Acidification of the male reproductive-tract by a proton-pumping H+-ATPase
.
Nature Medicine
2
,
470
472
.
Brown
,
D.
,
Sabolic
,
I.
and
Gluck
,
S.
(
1992
).
Polarized targeting of V-ATPase in kidney epithelial cells
.
J. exp. Biol
.
172
,
231
243
.
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. (in press)
.
Davis
,
R. L.
and
Dauwalder
,
B.
(
1991
).
The Drosophila dunce locus: learning and memory genes in the fly
.
Trends Genetics
7
,
224
229
.
Dow
,
J. A. T.
(
1984
).
Extremely high pH in biological systems: a model for carbonate transport
.
Am. J. Physiol
.
246
,
R633
R635
.
Dow
,
J. A. T.
(
1992
).
pH gradients in lepidopteran midgut
.
J. exp. Biol
.
172
,
355
375
.
Dow
,
J. A. T.
(
1994
).
V-ATPases in insects
.
In Organellar Proton-ATPases
(ed.
N.
Nelson
), pp.
75
102
.
Austin, TX
:
R. G. Landes Company
.
Dow
,
J.
,
Goodwin
,
S. F.
and
Kaiser
,
K.
(
1992
).
Analysis of the gene encoding a 16-kDa proteolipid subunit of the vacuolar H+-ATPase from Manduca sexta midgut and tubules
.
Gene
122
,
355
360
.
Dow
,
J. A. T.
,
Gupta
,
B. L.
,
Hall
,
T. A.
and
Harvey
,
W. R.
(
1984
).
X-ray microanalysis of elements in frozen-hydrated sections of an electrogenic K+ transport system: the posterior midgut of tobacco hornworm (Manduca sexta) in vivo and in vitro
.
J. Membr. Biol
.
77
,
223
241
.
Dow
,
J. A. T.
,
Kelly
,
D. C.
,
Davies
,
S. A.
,
Maddrell
,
S. H. P.
and
Brown
,
D.
(
1995
).
A member of the Major Intrinsic Protein family in Drosophila tubules
.
J. Physiol., Lond
.
489
,
110P
.
Dow
,
J. A. T.
,
Maddrell
,
S. H. P.
,
Görtz
,
A.
,
Skaer
,
N. V.
,
Brogan
,
S.
and
Kaiser
,
K.
(
1994a
).
The Malpighian tubules of Drosophila melanogaster: a novel phenotype for studies of fluid secretion and its control
.
J. exp. Biol
.
197
,
421
428
.
Dow
,
J. A. T.
,
Maddrell
,
S. H. P.
,
Davies
,
S.-A.
,
Skaer
,
N. J. V.
and
Kaiser
,
K.
(
1994b
).
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.
and
Peacock
,
J. M.
(
1989
).
Microelectrode evidence for the electrical isolation of goblet cavities of the middle midgut of Manduca sexta
.
J. exp. Biol
.
143
,
101
114
.
Finbow
,
M. E.
,
Eliopoulos
,
E. E.
,
Jackson
,
P. J.
,
Keen
,
J. N.
,
Meagher
,
L.
,
Thompson
,
P.
,
Jones
,
P.
and
Findlay
,
J. B. C.
(
1992
).
Structure of a 16 kDa integral membrane protein that has identity to the putative proton channel of the vacuolar H+-ATPase
.
Protein Engineering
5
,
7
15
.
Finbow
,
M. E.
,
Harrison
,
M.
and
Jones
,
P.
(
1995
).
Ductin – a proton pump component, a gap junction channel and a neurotransmitter release channel
.
BioEssays
17
,
247
255
.
Finbow
,
M. E.
and
Pitts
,
J. D.
(
1993
).
Is the gap junction channel – the connexon – made of connexin or ductin?
J. Cell Sci
.
106
,
463
471
.
Garrettengele
,
P.
,
Moilanen
,
B.
and
Cyert
,
M. S.
(
1995
).
Calcineurin, the Ca2+/calmodulin-dependent protein phosphatase, is essential in yeast mutants with cell integrity defects and in mutants that lack a functional vacuolar H+-ATPase
.
Molec. cell. Biol
.
15
,
4103
4114
.
Gausz
,
J.
,
Bencze
,
G.
,
Gyurkovics
,
H.
,
Ashburner
,
M.
,
Ish-Horowitz
,
D.
and
Holden
,
J. J.
(
1979
).
Genetic characterization of the 87C region of the third chromosome of Drosophila melanogaster
.
Genetics
93
,
917
934
.
Gluck
,
S.
(
1992
).
V-ATPases of the plasma-membrane
.
J. exp. Biol
.
172
,
29
37
.
Gluck
,
S.
and
Nelson
,
R.
(
1992
).
The role of the V-ATPase in renal epithelial H+ transport
.
J. exp. Biol
.
172
,
205
218
.
Goldstein
,
D. J.
,
Finbow
,
M. E.
,
Andresson
,
T.
,
Mclean
,
P.
,
Smith
,
K.
,
Bubb
,
V.
and
Schlegel
,
R.
(
1991
).
Bovine papillomavirus-E5 oncoprotein binds to the 16k component of vacuolar H+-ATPases
.
Nature
352
,
347
349
.
Gottlieb
,
R. A.
,
Giesing
,
H. A.
,
Zhu
,
J. Y.
,
Engler
,
R. L.
and
Babior
,
B. M.
(
1995
).
Cell acidification in apoptosis – granulocyte-colony-stimulating factor delays programmed cell-death in neutrophils by up-regulating the vacuolar H+-ATPase
.
Proc. natn. Acad. Sci. U.S.A
.
92
,
5965
5968
.
Guo
,
Y.
,
Gillan
,
A.
,
Török
,
T.
,
Kiss
,
I.
,
Dow
,
J. A. T.
and
Kaiser
,
K.
(
1996a
).
Site-selected mutagenesis of the Drosophila second chromosome via plasmid rescue of lethal P-element insertions
.
Genome Res. (in press)
.
Guo
,
Y.
,
Kaiser
,
K.
,
Wieczorek
,
H.
and
Dow
,
J. A. T.
(
1996b
).
The Drosophila melanogaster gene vha14 encoding a 14-kDa F-subunit of the vacuolar ATPase
.
Gene
172
,
239
243
.
Guo
,
Y.
,
Wang
,
Z.
,
Carter
,
A.
,
Kaiser
,
K.
and
Dow
,
J. A. T.
(
1996c
).
Characterisation of vha26, the Drosophila gene for a 26kDa E-subunit of the vacuolar ATPase
.
Biochim. biophys. Acta
1283
,
4
9
.
Haass
,
C.
,
Capell
,
A.
,
Citron
,
M.
,
Teplow
,
D. B.
and
Selkoe
,
D. J.
(
1995
).
The vacuolar H+-ATPase inhibitor bafilomycin A1 differentially affects proteolytic processing of mutant and wild-type β-amyloid precursor protein
.
J. biol. Chem
.
270
,
6186
6192
.
Harvey
,
B. J.
(
1992a
).
Energization of sodium-absorption by the H+-ATPase pump in mitochondria-rich cells of frog skin
.
J. exp. Biol
.
172
,
289
309
.
Harvey
,
W. R.
(
1992b
).
Physiology of V-ATPases
.
J. exp. Biol
.
172
,
1
17
.
Harvey
,
W. R.
,
Cioffi
,
M.
,
Dow
,
J. A. T.
and
Wolfersberger
,
M. G.
(
1983a
).
Potassium ion transport ATPase in insect epithelia
.
J. exp. Biol
.
106
,
91
117
.
Harvey
,
W. R.
,
Cioffi
,
M.
and
Wolfersberger
,
M. G.
(
1983b
).
Chemiosmotic ion pump of insect epithelia
.
Am. J. Physiol
.
244
,
R163
R175
.
Harvie
,
P. D.
and
Bryant
,
P. J.
(
1996
).
Cloning of the C subunit of V-ATPase from Drosophila melanogaster
.
A. Conf. Drosophila Res
.
37
,
92
.
He
,
J.
and
Kramer
,
H.
(
1996
).
Search for proteins that bind to Hook, a novel component of the endocytic compartment
.
A. Conf. Drosophila Res
.
37
,
93
.
Kaiser
,
K.
(
1993
).
Second-generation enhancer traps
.
Current Biol
.
3
,
560
562
.
Kaiser
,
K.
and
Goodwin
,
S. F.
(
1990
).
“Site-selected” transposon mutagenesis of Drosophila
.
Proc. natn. Acad. Sci. U.S.A
.
87
,
1686
1690
.
Kennison
,
J. A.
and
Tamkun
,
J. W.
(
1988
).
Dosage-dependent modifiers of Polycomb and Antennapedia mutations in Drosophila
.
Proc. natn. Acad. Sci. U.S.A
.
85
,
8136
8140
.
Klein
,
U.
(
1992
).
The insect V-ATPase, a plasma-membrane proton pump energizing secondary active transport – immunological evidence for the occurrence of a V-ATPase in insect ion-transporting epithelia
.
J. exp. Biol
.
172
,
345
354
.
Klein
,
U.
,
Löffelmann
,
G.
and
Wieczorek
,
H.
(
1991
).
The midgut as a model system for insect K+-transporting epithelia: immunocytochemical localization of a vacuolar-type H+ pump
.
J. exp. Biol
.
161
,
61
75
.
Koralnik
,
I. J.
,
Mulloy
,
J. C.
,
Andresson
,
T.
,
Fullen
,
J.
and
Franchini
,
G.
(
1995
).
Mapping of the intermolecular association of the human T-cell leukaemia/lymphotropic virus type-i p12(i) and the vacuolar H+-ATPase 16 kda subunit protein
.
J. gen. Virol
.
76
,
1909
1916
.
Küppers
,
J.
and
Bunse
,
I.
(
1996
).
A primary cation transport by a V-type ATPase of low specificity
.
J. exp. Biol
.
199
,
1327
1334
.
Lawrence
,
P. A.
(
1993
).
The Making of a Fly
.
Oxford
:
Blackwell Scientific
.
Leroy
,
C.
and
Meunier
,
F. M.
(
1995
).
Differential targeting to the plasma membrane of the Torpedo 15-kDa proteolipid expressed in oocytes
.
J. Neurochem
.
65
,
1789
1797
.
Maddrell
,
S. H. P.
and
O’donnell
,
M. J.
(
1992
).
Insect Malpighian tubules: V-ATPase action in ion and fluid transport
.
J. exp. Biol
.
172
,
417
429
.
Meagher
,
L.
,
Mclean
,
P.
and
Finbow
,
M. E.
(
1990
).
Sequence of a cDNA from Drosophila coding for the 16 kD proteolipid component of the vacuolar H+-ATPase
.
Nucleic Acids Res
.
18
,
6712
.
Mears
,
A. J.
,
Elshanti
,
H.
,
Murray
,
J. C.
,
Mcdermid
,
H. E.
and
Patil
,
S. R.
(
1995
).
Minute supernumerary ring chromosome 22 associated with cat eye syndrome: Further delineation of the critical region
.
Am. J. Human Genet
.
57
,
667
673
.
Moehrle
,
A.
and
Paro
,
R.
(
1994
).
Spreading the silence: Epigenetic transcriptional regulation during Drosophila development
.
Devl Genet
.
15
,
478
484
.
Nelson
,
H.
and
Nelson
,
N.
(
1990
).
Disruption of genes encoding subunits of yeast vacuolar H+-ATPase causes conditional lethality
.
Proc. natn. Acad. Sci. U.S.A
.
87
,
3503
3507
.
Nelson
,
R. D.
,
Guo
,
X. L.
,
Masood
,
K.
,
Brown
,
D.
,
Kalkbrenner
,
M.
and
Gluck
,
S.
(
1992
).
Selectively amplified expression of an isoform of the vacuolar H+-ATPase 56-kilodalton subunit in renal intercalated cells
.
Proc. natn. Acad. Sci. U.S.A
.
89
,
3541
3545
.
Nishihara
,
T.
,
Akifusa
,
S.
,
Koseki
,
T.
,
Kato
,
S.
,
Muro
,
M.
and
Hanada
,
N.
(
1995
).
Specific inhibitors of vacuolar-type H+-ATPases induce apoptotic cell death
.
Biochem. biophys. Res. Commun
.
212
,
255
262
.
Numata
,
M.
,
Ohkuma
,
S.
and
Iseki
,
S.
(
1995
).
Expression and localization of messenger RNA for the 16-kD subunit of V-ATPase in the rat embryo
.
J. Histochem. Cytochem
.
43
,
761
769
.
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
.
Pedersen
,
P. L.
and
Carafoli
,
E.
(
1987a
).
Ion motive ATPases. I. Ubiquity, properties and significance to cell function
.
Trends biochem. Sci
.
12
,
146
150
.
Pedersen
,
P. L.
and
Carafoli
,
E.
(
1987b
).
Ion motive ATPases. II. Energy coupling and work output
.
Trends biochem. Sci
.
12
,
186
189
.
Rubin
,
G. M.
(
1988
).
Drosophila melanogaster as an experimental organism
.
Science
240
,
1453
1459
.
Rubin
,
G. M.
and
Spradling
,
A. C.
(
1982
).
Genetic transformation of Drosophila with transposable element vectors
.
Science
218
,
348
353
.
Sabolic
,
I.
,
Wuarin
,
F.
,
Shi
,
L. B.
,
Verkman
,
A. S.
,
Ausiello
,
D. A.
,
Gluck
,
S.
and
Brown
,
D.
(
1992
).
Apical endosomes isolated from kidney collecting duct principal cells lack subunits of the proton pumping ATPase
.
J. Cell Biol
.
119
,
111
122
.
Takase
,
K.
,
Kakinuma
,
S.
,
Yamato
,
I.
,
Konishi
,
K.
,
Igarashi
,
K.
and
Kakinuma
,
Y.
(
1994
).
Sequencing and characterization of the ntp gene cluster for vacuolar-type Na+-translocating ATPase of Enterococcus hirae
.
J. biol. Chem
.
269
,
11037
11044
.
Tanida
,
I.
,
Hasegawa
,
A.
,
Iida
,
H.
,
Ohya
,
Y.
and
Anraku
,
Y.
(
1995
).
Cooperation of calcineurin and vacuolar H+-ATPase in intracellular Ca2+ homeostasis of yeast cells
.
J. biol. Chem
.
270
,
10113
10119
.
Torok
,
T.
,
Tick
,
G.
,
Alvarado
,
M.
and
Kiss
,
I.
(
1993
).
P-lacW insertional mutagenesis on the second chromosome of Drosophila melanogaster: Isolation of lethals with different overgrowth phenotypes
.
Genetics
135
,
71
80
.
Tully
,
T.
(
1991
).
Physiology of mutations affecting learning and memory in Drosophila – the missing link between gene product and behaviour
.
Trends Neurosci
.
14
,
163
164
.
Van Kerkhove
,
E.
(
1994
).
Cellular mechanisms of salt secretion by the Malpighian tubules of insects
.
Belg. J. Zool
.
124
,
73
90
.
Weltens
,
R.
,
Leyssens
,
A.
,
Zhang
,
A. L.
,
Lohhrmann
,
E.
,
Steels
,
P.
and
Van Kerkhove
,
E.
(
1992
).
Unmasking of the apical electrogenic H pump in isolated Malpighian tubules (Formica polyctena) by the use of barium
.
Cell. Physiol. Biochem
.
2
,
101
116
.
Wieczorek
,
H.
(
1992
).
The insect V-ATPase, a plasma-membrane proton pump energizing secondary active transport – molecular analysis of electrogenic potassium transport in the tobacco hornworm midgut
.
J. exp. Biol
.
172
,
335
343
.
Wieczorek
,
H.
and
Harvey
,
W. R.
(
1995
).
Energization of animal plasma membranes by the proton-motive force
.
Physiol. Zool
.
68
,
15
23
.
Wieczorek
,
H.
,
Putzenlechner
,
M.
,
Zeiske
,
W.
and
Klein
,
U.
(
1991
).
A vacuolar-type proton pump energizes K+/H+ antiport in an animal plasma membrane
.
J. biol. Chem
.
266
,
15340
15347
.
Zhang
,
Y.
and
Fillingame
,
R. F.
(
1995
).
Changing the ion-binding specificity of the Escherichia coli H+-ATP synthase by directed mutagenesis of subunit c
.
J. biol. Chem
.
270
,
87
93
.