Eukaryotic cells contain organelles bounded by a single membrane in the cytoplasm. These organelles have differentiated to carry out various functions in the pathways of endocytosis and exocytosis. Their lumina are acidic, with pH ranging from 4.5 to 6.5. This article describes recent studies on these animal cell organelles focusing on (1) the primary proton pump (vacuolar-type H+-ATPase) and (2) the functions of the organelle luminal acidity. We also discuss similarities and differences between vacuolar-type H+-ATPase and F-type ATPase. Our own studies and interests are emphasized.

Highly differentiated endomembrane (single-membrane-bounded) organelles are involved in the dynamic membrane trafficking processes essential for protein sorting and targeting, receptor-mediated endocytosis, neural or hormonal signal transduction and other processes (Mellman et al., 1986; Goldstein et al., 1985). These organelles include the Golgi apparatus, secretory vesicles, coated vesicles, endosomes, lysosomes and synaptic vesicles. Their lumina are acidic, ranging in pH from 4.5 to 6.5, thus generating a transmembrane electrochemical proton gradient with respect to the near-neutral cytoplasm. The acidic luminal pH is established by a proton-pumping vacuolar-type ATPase (V-ATPase) in combination with ion channels and transporters, whose varying distribution contributes to the organelle-specific luminal contents. The acidic pH is required for organelle functions such as hydrolysis of macromolecules, release of ligands from receptors and processing of preproproteins. The proton electrochemical gradient, consisting of the pH gradient (Δ pH) and/or membrane potential (Δ Ψ), provides a driving force for the accumulation of hormones or transmitters into secretory vesicles, synaptic vesicles or synaptic-like microvesicles (microvesicles). A diversity of organelles is ensured by the accurate sorting and transport of the specific components into the lumina or membranes of the corresponding organelles. In this regard, it has become apparent recently that certain organelle-specific membrane protein(s) function in docking/fusion with transport vesicles (Ferro-Novick and Jahn, 1994).

In this article, we briefly summarize recent studies on animal acidic organelles that possess V-ATPase. We include results from our own laboratories and emphasize our interests. Similarities between V-ATPase and F-ATPase (ATP synthase) are also discussed. Information about related fields not mentioned here can be found in reviews (Forgac, 1989; Anraku, 1996; Stevens and Forgac, 1997; Futai et al., 1998).

Visualization of the acidic luminal pH of endomembrane organelles

The acidic luminal pH of the endomembrane organelles can be visualized using pH-sensitive fluorescent dyes, such as Acridine Orange and fluorescein isothiocyanate (FITC)-dextran, or by immuno-gold electron microscopy using 3(2,4-dinitroanilino)-3’-amino-N-methyl dipropylamine (DAMP) (Anderson and Orci, 1988; Yoshimori et al., 1991). Acridine Orange and DAMP are lipophilic amines, and their non-protonated forms permeate membranes. They become non-permeant once protonated in the acidic lumina of the organelles, yielding pH-dependent distributions. The fluorescence of accumulated Acridine Orange is qualitatively proportional to the pH of the organelle (dark yellow to orange) in mouse or human cell lines (Yoshimori et al., 1991). The organelle pH can be estimated by counting gold particles in immuno-gold electron micrographs using DAMP (Yoshimori et al., 1991). Lysosomal pH was estimated to be approximately 5 using this method (Yoshimori et al., 1991).

The fluorescence of FITC-dextran incorporated into endosomes or lysosomes can be correlated with the intra-organellar pH by monitoring the excitation wavelength. That the acidic pH is generated by V-ATPase can be confirmed using inhibitors such as bafilomycin and concanamycin (Bowman et al., 1988; Dröse et al., 1993; Ito et al., 1995). The lysosomal pH of a cultured human carcinoma cell line was approximately 4.8, as estimated by FITC-dextran incorporation, and increased to approximately 6 upon treatment with bafilomycin (Yoshimori et al., 1991). The advantage of inhibitors is that they are effective at low concentration and do not inhibit F-ATPase or P-type ATPase (Bowman et al., 1988). Inhibition was prevented by adding the Vo sector (Hanada et al., 1990) or the a subunit (Zhang et al., 1994), indicating that bafilomycin inhibits the enzyme by binding to the Vo sector a subunit. In contrast to ionophores, bafilomycin does not alter organelle morphology even after prolonged administration (Yoshimori et al., 1991; Umata et al., 1990). However, ionophores such as nigericin can be used for a short-term assay to disrupt Δ pH in cultured cells. Analysis using inhibitors supported the hypothesis that an acidic luminal pH is required at the later stage of receptor-mediated endocytosis of epidermal growth factor (Yoshimori et al., 1991).

Organelles with an acidic luminal pH

At what developmental stage do mammals establish acidic compartments? We studied this question using preimplanation mouse embryos cultured in vitro (G.-H. Sun-Wada, Y. Wada and M. Futai, in preparation). Fertilized eggs to eight-cell embryos showed a weak diffuse granular staining with Acridine Orange throughout the cytoplasm. However, in compacted eight-cell embryos and morulae stages, staining became apparent. The staining patterns disappeared immediately when embryos were exposed to a weak base such as ammonium chloride. In the blastocyst, acidic compartments were concentrated exclusively in the perinuclear region of trophectoderm cells, while staining in the inner cell mass was still diffuse. These results suggest that acidic compartments are formed and that their morphologies change during an early developmental stage. The polarized distribution of the compartments may be essential for early mammalian development. Immunofluorescence microscopic analysis using an antibody to subunit B confirmed that the V-ATPase is localized to these acidic compartments.

Small regions of the plasma membrane, called coated pits, form clathrin-coated vesicles during receptor-mediated endocytosis (Mellman et al., 1986). The regions containing receptors with bound ligand are incorporated into the vesicles. After shedding the clathrin coat, these vesicles form early endosomes and fuse with lysosomes. During exocytosis, vesicles derived from endosomes or from the Golgi network fuse with plasma membranes and secrete neurotransmitters or hormones. We have been interested in the accumulation of transmitters into neural synaptic vesicles and endocrine cell microvesicles. The electrochemical proton gradient established by the V-ATPase in combination with ion channels or transporters provides the driving force for transporting transmitters into vesicles. Glutamate is accumulated into synaptic vesicles in a manner dependent on membrane potential (Moriyama and Futai, 1990; Moriyama et al., 1990), whereas the transport of monoamines and γ-aminobutyric acid (GABA) is dependent on Δ pH (Schuldiner et al., 1995). The acidic interior of these vesicles also resulted in the accumulation of neuron blockers (lipophilic cations), which penetrate the vesicles in a non-protonated lipophilic form and accumulate inside as protonated forms (Moriyama et al., 1993b). This accumulation dissipates the Δ pH and inhibits Δ pH-coupled transport. Thus, the lipophilic cations inhibit the accumulation of monoamines or GABA, but not glutamate transport driven by membrane potential. Compounds such as 1-methyl-4-phenylpyridinium (MPP+), which is known to cause selective degradation of dopaminergic neurons, are transported specifically by a Δ pH-coupled monoamine transporter (Moriyama et al., 1993a). These findings indicate that the acidic compartments are important pharmacologically for considering the mechanisms of action of hydrophobic drugs.

Microvesicles are small organelles (approximately 50 nm in diameter) with a morphology similar to the synaptic vesicles and membrane components that are necessary for docking/fusion with plasma membranes and transmitter accumulation (Moriyama et al., 1996). They have specific transporters coupled to V-ATPase. Examples include pancreatic β cells (GABA; Thomas-Reetz et al., 1993), PC12 cells (acetylcholine; Bauerfeind et al., 1993) posterior pituitary cells (norepinephrine; Moriyama et al., 1995) and pinealocytes (glutamate; Moriyama and Yamamoto, 1995). Microvesicles and synaptic vesicles are similar to each other in protein composition, having V-ATPase, transmitter transporters and factors required for docking with plasma membranes. However, they are not exactly the same. Pinealocyte microvesicles are devoid of synapsin and contain synaptophysin with a slightly different electrophoretic mobility from that of synaptic vesicles (Moriyama and Yamamoto, 1995). Synaptic vesicle protein 2 (SV2) has two known isoforms: SV2A has been shown to be present in neuronal and endocrine cells, whereas SV2B is neuron-specific (Bajjalieh et al., 1994). However, SV2B has been found in pinealocyte and clonal pancreatic α cells (Hayashi et al., 1998). Thus, the distribution of SV2B does not always define a vesicle as being of neuronal origin. The rates of release of transmitters from the two vesicles are different: the rates are of the order of seconds for microvesicles but only milliseconds for synaptic vesicles (Yamada et al., 1996). The mechanism by which transmitters accumulate into microvesicles and synaptic vesicles are similar: membrane-potential-dependent transport of glutamate into pinealocyte microvesicles (Moriyama and Yamamoto, 1995) or Δ pH-driven transport of GABA into pancreatic β cell vesicles (Thomas-Reetz et al., 1993) and of monoamine into posterior pituitary cells (Moriyama et al., 1995). Studies on the roles of microvesicles in pinealocytes have contributed to our understanding of the roles of glutamate in endocrine cells (Yamada et al., 1996; Moriyama et al., 1996).

V-ATPase has also been localized to the plasma membrane of epithelial cells, including those in the seminal duct (Breton et al., 1996), in osteoclasts (Chatterjee et al., 1992), in kidney proximal tubules (Gluck, 1992) and in bladder (Tomochika et al., 1997). These localizations are consistent with the role of acidic pH in these tissues or cells. Plasma membrane V-ATPase is also present in insect midgut, Malpighian tubules and sensory sensilla (Wieczorek et al., 1999) as well as frog skin (Harvey, 1992). It is of interest to study the biochemical differences between V-ATPase localized to plasma membranes and to endomembrane organelles.

V-ATPase catalysis

The V-ATPase is a primary proton pump with a structure and mechanism similar to those of F-ATPase (Nelson, 1992; Anraku, 1996; Stevens and Forgac, 1997; Futai et al., 1998; Nelson and Harvey, 1999). The V-ATPase functions only as an ATP hydrolase, pumping protons away from the cytoplasmic compartment, whereas the F-ATPase is an ATP synthase driven by an electrochemical proton gradient and functions as an ATPase only in exceptional cases (Futai et al., 1989). ATP hydrolysis by V-ATPase does not follow simple Michaelis–Menten kinetics: the chromaffin granule V-ATPase has three Km values (Hanada et al., 1990) and, as for F-ATPase, single-site/multisite catalysis has been suggested from studies of the yeast enzyme (Uchida et al., 1988; Hirata et al., 1989). It is of interest to know whether V-ATPase is a reversible enzyme. Synthesis of ATP by V-ATPase is dependent on an electrochemical proton gradient, as has been shown in isolated plant vacuoles (tonoplast-enriched vesicles) by measuring the exchange reaction (Façanha and de Meis, 1998) that leads to the incorporation of radioactive phosphate into ATP. ATP synthesis has also been assayed using a luciferin/luciferase assay (Dupaix et al., 1989; Schmidt and Briskin, 1993). However, detailed studies with inhibitors for related enzymes including V-ATPase, F-ATPase and adenylate kinase are still required. The plant pyrophosphatase has been expressed in yeast vacuoles (Kim et al., 1994). Both V-ATPase and pyrophosphatase are functional in the isolated vacuoles and form electrochemical proton gradients depending on the presence of ATP and pyrophosphate, respectively. We were interested in the pyrophosphatase system to test the reversibility and regulation of V-ATPase. The yeast vacuole could synthesize ATP, depending on the electrochemical proton gradient established by the pyrophosphatase (Hirata et al., 2000). These results indicate that the V-ATPase is a reversible enzyme. The vacuole was a good system for studying the regulation of V-ATPase by Δ pH or membrane potential: analysis of Δ pH formation with ATP and/or pyrophosphate suggests that V-ATPase is not strictly regulated by proton gradient.

Catalytic V1 sector

Like the F1 and Fo sectors of F-ATPase, V-ATPase has a ball- and-stalk structure (see, for example, Moriyama et al., 1991; for a review, see Forgac, 1989) and membrane extrinsic and intrinsic sectors, termed V1 and Vo, respectively, from the analogy (Fig. 1). The V1 sector contains catalytic domains, and the Vo sector constitutes a proton pathway. The V1 subunits for the catalytic domains are A (67–73 kDa), B (55–60 kDa), C (41 kDa), D (34 kDa), the stalk subunits are E (33 kDa), F (14 kDa), G (15 kDa) and H (50–57 kDa) and the Vo subunits are a (100–116 kDa), c’’(19–23 kDa), d (38–39 kDa) and c and c’ (14–17 kDa) (Stevens and Forgac, 1997; Arai et al., 1988). The 9.7 kDa Vo subunit was found recently in insect, human, murine and bovine sources (Merzendorfer et al., 1999). The molecular mass of the V1 sector is approximately 500 kDa with a stoichiometry of A3B3C1D1E1F1G3H1, whereas the molecular mass of the Vo sector is approximately 250 kDa with a stoichiometry of a1d1c’’1(c and c’)6. Phenotypes of yeast lacking V-ATPase have been useful for defining the subunits of the enzyme (Anraku, 1996; Stevens and Forgac, 1997):mutants show no growth at pH 7.5, failure to accumulate pigment in an ade2 cell and sensitivity to Ca2+. Mutants lacking a V-ATPase can grow at pH 5.5, indicating that the loss of V-ATPase causes conditional lethality (Nelson and Nelson, 1990). Utilizing these phenotypes, yeast subunits and assembly factors have been identified genetically (Stevens and Forgac, 1997; Anraku, 1996). Certain V-ATPase subunits from other eukaryotes, such as Caenorhabditis elegans subunit C, have been identified functionally by complementation with yeast mutants (Oka et al., 1998). Most of the V1 subunits were identified biochemically in mammals (Stevens and Forgac, 1997) and insects (Wieczorek et al., 1999).

Fig. 1.

Structural models of F-ATPase and V-ATPase. Subunit structures and ATP-hydrolysis-dependent H+ transport by V-ATPase and F-ATPase are shown schematically. Homologous subunits, such as subunits A and β of V-ATPase and F-ATPase, respectively, are indicated by the same colours.

Fig. 1.

Structural models of F-ATPase and V-ATPase. Subunit structures and ATP-hydrolysis-dependent H+ transport by V-ATPase and F-ATPase are shown schematically. Homologous subunits, such as subunits A and β of V-ATPase and F-ATPase, respectively, are indicated by the same colours.

Like the α and β subunits of F1, V-ATPase has three pairs of A and B subunits. The catalytic subunit A of V-ATPase shares approximately 25 % identity with the β subunit of F-ATPase (Bowman et al., 1992). Subunit A contains an insertion of approximately 100 amino acid residues not found in the β subunit. This insertion reduces the identity between the two subunits and also suggests that structural and functional differences may exist between the two ATPases. Catalytic residues identified in the P-loop (Gly-X-X-X-X-Gly-Lys-Thr) and the GERXXE sequence (Gly-Glu-Arg-X-X-Glu) of the β subunit are conserved in subunit A (Futai et al., 1989; Futai and Omote, 1996; Omote and Futai, 1998) (Fig. 2). A cysteine residue is conserved in the P-loop (Gly-X-X-X-Cys-Gly-Lys-Thr) of V-ATPase but not in that of F-ATPase. N-ethylmaleimide (NEM) modifies this cysteine and inhibits ATPase activity (Feng and Forgac, 1992). The regulatory role of the cysteine was suggested by Feng and Forgac (1994): the P-loop cysteine (Cys261) of the bovine coated-vesicle ATPase forms a disulphide bond with Cys539 of the same subunit and inactivates the enzyme. The two cysteines are predicted to be approximately 1.3 nm apart, suggesting that the distortion of the nucleotide binding site may occur upon disulphide bond formation. F-ATPase lacks the cysteine in the P-loop, consistent with its insensitivity to NEM. However, the introduction of a cysteine into the corresponding position makes the F-ATPase sensitive to NEM (Iwamoto et al., 1994). Yeast mutagenesis experiments confirmed that the lysine (Lys263) in the P-loop and the glutamate (Glu286) in the GERXXE sequence are catalytically essential (Liu et al., 1997). In the proposed mechanism of F-ATPase catalysis, the lysine binds the β/γ phosphate of ATP, and the glutamate activates a water molecule (Abrahams et al., 1994; Futai and Omote, 1996). The positive charge of the arginine in GERXXE is also required for catalysis, and the last glutamate of the GERXXE sequence is essential for catalytic cooperativity in F-ATPase (Futai and Omote, 1996). The conservation of these residues suggests that V-ATPase may have similar catalytic mechanism. The V1 subunit B is a counterpart of the F1 α subunit, and they share appoximately 25 % identity (Bowman et al., 1992; Inatomi et al., 1989): the α subunit residues required for the catalytic cooperativity of the F1 sector are conserved in the B subunit of the V1 sector. The V1 subunit corresponding to the γ subunit of F1 is not known, although V1 subunit D (yeast Vma8p) may be the counterpart, as suggested by their predicted structural similarity (Nelson et al., 1995; Graham et al., 1995).

Fig. 2.

Alignment of the sequence of the F-ATPase β subunit of Escherichia coli and the V-ATPase A subunits of Saccharomyces cereviciae and ox. Parts of the sequence of the F-ATPase β subunit and V-ATPase A subunits are aligned. The P-loop (Gly-X-X-X-X-Gly-Lys-Thr) and the GERXXE sequence (Gly-Glu-Arg-X-X-Glu) are boxed, and conserved residues in both V- and F-ATPase are emphasized with a red bar. The roles of the cysteine residue in the V-ATPase P-loop are discussed in the text; both motifs contain catalytic residues.

Fig. 2.

Alignment of the sequence of the F-ATPase β subunit of Escherichia coli and the V-ATPase A subunits of Saccharomyces cereviciae and ox. Parts of the sequence of the F-ATPase β subunit and V-ATPase A subunits are aligned. The P-loop (Gly-X-X-X-X-Gly-Lys-Thr) and the GERXXE sequence (Gly-Glu-Arg-X-X-Glu) are boxed, and conserved residues in both V- and F-ATPase are emphasized with a red bar. The roles of the cysteine residue in the V-ATPase P-loop are discussed in the text; both motifs contain catalytic residues.

Proton channel Vo sector

The Vo sector consists of at least four different subunits. The extremely hydrophobic Voc subunit (16 kDa), also called the proteolipid because of its lipophilicity, has been cloned from yeast Saccharomyces cerevisiae (Nelson and Nelson, 1989; Umemoto et al., 1991), Caenorhabditis elegans (Oka et al., 1997, 1998), Manduca sexta (Dow et al., 1992), Drosophila melanogaster (Finbow et al.,1994), ox (Mandel et al., 1988), mouse (Hanada et al., 1991) and human (Hasebe et al., 1992). It is the counterpart of the Foc subunit (8 kDa). The amino and carboxyl halves of the Voc subunit are homologous, suggesting that the subunit may have resulted from a duplication of an ancestral gene (Nelson, 1992; Mandel et al., 1988). The c subunit has four hydrophobic transmembrane α helices (I, II, III, IV), and the amino (I, II) and carboxyl (III, IV) domains are individually homologous to the Foc subunit. However, unlike the Foc subunit, the Vo amino-terminal domain does not have glutamate or aspartate in the middle of domain II. The glutamate is only present in domain IV, where it is critical for proton translocation, because it could not be replaced by other amino acid residues including glutamine (Noumi et al., 1991). Domain IV has the highest degree of sequence conservation of any domain in the c subunits so far sequenced, supporting its essential role in proton pumping. Six copies of the c subunit are present in the Vo complex (Stevens and Forgac, 1997); thus, 24 transmembrane α helices from the six c subunits are present in Vo, whereas the same number of α helices in Fo requires 12 copies of the c subunit, as clearly confirmed recently (Jones and Fillingame, 1998). The V-ATPase proteolipid (subunit c) has also been suggested to be a component of the insect gap junction (Finbow et al., 1994).

Vo has another proteolipid (subunit c’’, 23 kDa) found in yeast (Hirata et al., 1997), C. elegans (Oka et al., 1997), mouse (G. H. Sun-Wada, Y. Wada and M. Futai, in preparation) and human (Nishigori et al., 1998). The sequence of this proteolipid can be aligned with the Voc subunit except that an additional 50 amino-terminal residues are present. Like the c subunit, the amino- and carboxyl-terminal halves of the 23 kDa subunit share homology. This protein has five putative transmembrane domains. C. elegans subunit c’’ (vha-4) exhibited 52 % identity with yeast Vma16p and 64 % with a human homologue (ATP6F). A mouse homologue is 96 % identical with ATP6F. A glutamate residue in the middle of the third transmembrane segment is conserved in all proteolipids so far sequenced. Despite the homology with the c subunit, the VMA16 gene coding for the 23 kDa proteolipid is also essential for the activity of yeast V-ATPase (Hirata et al., 1997). No isoform for the c’’ subunit is found in yeast and C. elegans, in which the entire genomes have already been sequenced.

The Vo subunit a is a large protein (116 kDa in mammals and 100 kDa in yeast) (Stevens and Forgac, 1997). This subunit may be a counterpart of F1 subunit a, which is proposed to function as a part of the stator in F1Fo (Junge et al., 1997). This subunit can be divided into two domains; a hydrophilic amino-terminal domain and a carboxyl half forming 6–7 transmembrane domains. V-ATPase subunit G shows significant homology with the b subunit of Fo (Hunt and Bowman, 1997). Subunit b has a single membrane-spanning domain and an extra-membrane domain, both possibly α helices, and is suggested to serve as a stator that holds the α3β3 complex in place while the γec complex rotates. However, V-ATPase subunit G apparently lacks a membrane-spanning domain. Thus, if the G and b subunits play the same role, the G subunit should form a complex with a hydrophilic domain of the a subunits of Vo and also V1 subunit(s). From the sequence identities, it is not easy to identify the counterparts of F1 d and e subunits that may interact with G.

Overall structural similarities (Fig. 1) suggest that V- and F-ATPase utilize the same basic mechanisms for ATP hydrolysis, proton translocation and energy coupling. Thus, the mechanism of action of V-ATPase may include the rotation of the subunit complex during ATP-dependent proton translocation, as shown for the F-ATPase (Noji et al., 1997; Junge et al., 1997; Omote et al., 1999).

As described above, V-ATPase pumps protons into diverse endomembrane organelles to generate their unique acidic luminal pH, and the same enzyme in the plasma membranes of epithelial cells acidifies external compartments, such as those in the urinary bladder (Tomochika et al., 1997), and alkalizes other compartments, such as those in caterpillar midgut (Wieczorek et al., 1999). It is tempting to speculate that different V-ATPase subunit isoforms contribute to the organelle-specific luminal pH, and are localized to the plasma membrane to acidify or alkalize the external cytosol. Since V-ATPase is a complicated multisubunit enzyme, changing a single subunit may affect the properties of the entire enzyme. Thus, the roles of V-ATPase subunit isoforms are of interest and should be studied carefully.

Isoforms of the c subunit

Two genes (VMA3, VMA11) code for the yeast c and c’ subunits, respectively, and their products are 56 % identical in amino acid sequence (Anraku, 1996). They are not redundant genes, and Vma3p and Vma11p are both essential for the activity of V-ATPase, indicating that they are subunits of the same enzyme (Umemoto et al., 1991). We have detected three C. elegans genes coding for the c subunit (Fig. 3). Two of them (vha-1, vha-2) form an operon on chromosome III (Oka et al., 1997), and the third gene (vha-3) and the subunit C gene (vha-11) form a cluster on chromosome IV (Oka et al., 1998). The vha-1 gene product is 66 % identical to that of vha-2 or vha-3 and 61 % identical to that of yeast Vma3p. Since the vha-1 and vha-2 gene products share equal similarities with the Vma3p and Vma11p proteins, we cannot conclude that the two yeast genes correspond to the C. elegans counterparts. vha-2 and vha-3 produce an identical protein, although the nucleotide sequences of their coding regions differ by 15 %. In this regard, two cDNAs for closely similar proteolipids have been isolated from cotton, and their coding region shared 95 % identity (Hasenfratz et al., 1995).

Fig. 3.

The Caenorhabditis elegans vha-1, vha-2 and vha-3 genes coding for the c subunits (16 kDa proteolipid). (A) Gene organization of vha-1, vha-2 and vha-3 genes. Filled and hatched boxes represent coding and untranslated regions, respectively. (B) Expression of the vha-1 gene in the H-shaped excretory cell. Expression of the vha-1::lacZ fusion gene (pCV02) in a young adult transgenic animal is shown. lacZ expression is seen in the H-shaped excretory cell. (C) Expression of the vha-3 gene in gastric and intestinal cells. The lacZ reporter gene (pCV3-03) is under the control of a 2.2 kilobase (kb) upstream regulatory region of vha-3. vha-11 is the gene for subunit C. See text and Oka et al. (1997, 1998) for more details.

Fig. 3.

The Caenorhabditis elegans vha-1, vha-2 and vha-3 genes coding for the c subunits (16 kDa proteolipid). (A) Gene organization of vha-1, vha-2 and vha-3 genes. Filled and hatched boxes represent coding and untranslated regions, respectively. (B) Expression of the vha-1 gene in the H-shaped excretory cell. Expression of the vha-1::lacZ fusion gene (pCV02) in a young adult transgenic animal is shown. lacZ expression is seen in the H-shaped excretory cell. (C) Expression of the vha-3 gene in gastric and intestinal cells. The lacZ reporter gene (pCV3-03) is under the control of a 2.2 kilobase (kb) upstream regulatory region of vha-3. vha-11 is the gene for subunit C. See text and Oka et al. (1997, 1998) for more details.

Northern blot analysis indicated that three proteolipid genes are expressed in a similar pattern during the life cycle of C. elegans (Oka et al., 1998). However, studies with transgenic worms indicated that the genes are expressed differently in a cell-specific manner. The vha-1 and vha-2 genes were expressed in most of the larval cells, but predominantly in an H-shaped excretory cell in the adult (Oka et al., 1997). The H-shaped cell is a large mononuclear cell with bilateral excretory canals extending along the length of the worm’s body. The cell body forming a bridge between the two canals is positioned on the ventral epidermal ridge slightly posterior to the nerve ring. The expression pattern of the c subunit was similar to that of the subunit c’’ proteolipid coded by the vha-4 gene and of the V1 sector B subunit, indicating that an entire V-ATPase is highly expressed in an H-shaped cell. Immunochemical localization of the subunit C protein confirmed the high level of expression of V-ATPase in the H-shaped cell (T. Oka and M. Futai, in preparation). A P-glycoprotein (PGP-3) is also predominantly expressed in the H-shaped cell, and the pgp-3 deletion mutant is sensitive to colchicine and chloroquine, suggesting that the P-glycoprotein exports toxic compounds or metabolic waste from the H-shaped cell to the exterior (Broeks et al., 1995). Thus, the V-ATPase that is highly expressed in the H-shaped cell may be supporting such a transport function.

The vha-3 gene was mainly expressed in gastrointestinal and hypodermal cells in addition to the H-shaped cell (Oka et al., 1998). In contrast, vha-2 was highly expressed in the H-shaped cell and was not detectable in gastrointestinal and hypodermal cells even after prolonged staining with the reporter gene product. These results strongly suggest that expression of the two vha genes is differentially regulated according cell type.

The presence of cell-specific isoforms in C. elegans prompted us to survey the c subunit isoforms in mammals. However, only one gene was found for the human c subunit, although three pseudogenes were identified (Hasebe et al., 1992). Similarly, two pseudogenes were identified for mouse, and Southern blot analysis suggested that no closely related homologue was present (T. Noumi, H. Inoue and H. Kanazawa, in preparation). It should be noted that negative results in the blotting analysis do not always indicate an absence of a gene. In this regard, the presence of the C. elegans vha-3 isoform could be deduced from Southern blotting with the reading frame of vha-2 as a probe, but not with that of vha-1 (Oka et al., 1998).

Kidney-type and ubiquitous isoforms of the B subunit

Two cDNAs encoding homologous but distinct B subunits (B1, B2) have been cloned from ox, human and mouse (Nelson et al., 1992). Their amino acid sequences share approximately 80 % identity: the B1 subunit is known to be a kidney-type subunit, whereas B2 is ubiquitous. The V-ATPase in kidney is responsible for acid secretion by distal nephrons, and the B1 subunit is found at the apical surface of acid-secreting cells. Karet et al. (1999) identified 15 different mutations (premature termination, frameshift, missense substitutions and splice site mutations) in the human ATP6B1 gene coding for the B1 subunit. All the altered residues in the missense mutations are conserved in the B subunit homologues from human to Neurospora crassa (Bowman et al., 1992). Patients with ATP6B1 mutations had distal renal tubular acidosis, a condition characterized by impaired renal acid secretion and resulting in metabolic acidosis (Karet et al., 1999). Patients also had sensorineural hearing loss: consistent with this finding, expression of ATP6B1 in the cochlea and endolymphatic sac has been demonstrated in mouse. These results suggest that expression of ATP6B1 is required for endolymph pH homeostasis and normal auditory function.

Isoforms of subunit a

Yeast has two genes, VPH1 and STV1, coding for isoforms of Vo subunit a (Manolson et al., 1992, 1994). Their gene products (Vph1p and Stv1p) are found in distinct compartments, probably vacuoles and endosomes/Golgi network, respectively. On the basis of the sequence similarity, four candidate genes for C. elegans subunit a could be found in its genome (T. Oka, unpublished results). These subunit a isoforms may be responsible for the organelle-or cell-specific localization.

Three subunit a isoforms seem to be present in mammals: the subunit for coated-vesicles from bovine brain (Peng et al., 1994), the protein for which cDNA was originally cloned as coding for a putative immune regulator from murine T cells (Lee et al., 1990) and a putative subunit from human osteoclast (Li et al., 1996). Sequence divergence between the three subunits is more than that expected from species differences.

We have cloned mouse cDNAs coding for three subunit a isoforms, termed a1, a2 and a3, coding for polypeptides of 832, 856 and 834 amino acid residues, respectively (T. Toyomura, T. Oka, C. Yamaguchi, Y. Wada and M. Futai, in preparation). Their amino acid sequences are appoximately 50 % identical: the homology is higher in the hydrophobic region of the carboxyl-terminal half and weak in the amino-terminal half between residues 70 and 300. It is also noteworthy that the regions between residues 660 and 720 have essentially no homology. These regions may contribute to the tissue-or organelle-specific distributions. All three proteins seem to have similar topologies, giving 6–9 transmembrane domains (Stevens and Forgac, 1997; Leng et al., 1999). Immunoblots have indicated different tissue distributions of the three isoforms. It should be noted that the subunit a isoforms have been found in organisms ranging from yeast to mammals.

The results of immunoreactivity and northern blotting suggest that the distribution of mammalian V-ATPase subunit isoforms is different among cells and among organelles. We may further speculate that the subunit isoform determines the specific location of V-ATPase. Tissue-specific isoforms of the bovine subunit a1, formed by alternative RNA splicing, have also been identified (Peng et al., 1994).

Is V-ATPase essential for the growth of higher eukaryotes? Yeast cells lacking V-ATPase cannot grow at neutral pH but can grow at acidic pH. Growth at acidic pH may be explained because an acidic luminal pH in the organelles can be generated by fusion with pinocytotic vesicles containing acidic fluid (Nelson and Nelson, 1990). These results suggest that V-ATPase is not essential for yeast growth as long as the medium is acidic.

Higher animals, including mammals, are not likely to survive without V-ATPase because it is ubiquitously expressed in tissues and physiologically important for various functions. Thus, deletions of only a tissue-specific protein such as the B1 subunit are likely to result in a viable phenotype (Karet et al., 1999). However, it is not known at what stage of animal development the defective V-ATPase has its lethal effect.

The obvious test to answer this question is to create an animal lacking an essential V-ATPase subunit. We were interested in the vha-11 gene coding for V1 subunit C of C. elegans. The VHA-11 protein exihibits 37, 56 and 56 % sequence identity with the C subunit of yeast (Vma5p), ox and human, respectively (Oka et al., 1998). The yeast Vma5p protein is essential for growth at neutral pH, and only one corresponding gene is found in the C. elegans genome. Expression of vha-11 completely restored the growth of the yeast vma5 mutant, confirming that vha-11 codes for a functional subunit C. The transgene with the vha-11 promoter region was expressed in an adult H-shaped cell, as for other V-ATPase subunits. Thus, it should be obvious that the disruption of this gene would affect the animal seriously. Gene function in C. elegans is conveniently disrupted using RNA interference (RNAi) by double-stranded RNA (Montgomery et al., 1998). The double-stranded RNA has been shown to target and disrupt the corresponding gene activity. When we introduced double-stranded RNA corresponding to the vha-11 transcript into the adult gonad, no live offspring were obtained from the animal (T. Oka and M. Futai, in preparation). These results suggest that V-ATPase, and therefore an acidic organelle lumen, is essential for the hatching of the worm.

The first knock-out of V-ATPase in an animal was obtained from Drosophila melanogaster. A fly mutant with lethal P-element insertions in the vha55 gene for the B subunit was identified by Davies et al. (1996). Deletion of the B subunit locus was shown to be lethal, whereas point mutations gave varying phenotypes that ranged from lethal to surviving flies. These results suggest that V-ATPase in organelles or plasma membranes is essential for the development from larva to fly. The same question can be asked regarding mammalian V-ATPase. The mouse subunit c gene (PL16) is an appropriate target to knock out because no isoforms have been found. Chromosomal PL16 was replaced by the neomycin-resistance gene using target-directed mutagenesis (Inoue et al., 2000). No PL16-/- mouse has been identified at birth, while PL16+/- embryos were viable and showed no noticeable abnormality. These results indicate that the homozygous loss of the c subunit gene caused embryonic lethality. However, the PL16-/- embryo could develop into the blastocyst stage in vitro (G. H. Sun-Wada, Y. Wada and M. Futai, in preparation). Furthermore, the mutant blastocysts formed acidic compartments similar to those of the wild type, suggesting that the maternal mRNA is sufficient for establishing an acidic compartment up to the blastocyst stage. Embryos of the same stage were also detected in vivo, but none was attached to the uterine epithelium after 7 days post coitus. These results suggest that V-ATPase or acidic organelles are essential immediately before gastrulation in mammals.

It is also of great interest to disrupt V-ATPase in specific cells or tissues using a conditional knock-out system. An initial attempt has been made to rescue a PL16-/-mouse by introducing a transgene for subunit c, and to knock out the transgene in defined cells at specific developmental stage. These experiments are currently in progress (T. Yoshimizu, G. H. Sun-Wada, Y. Wada and M. Futai, in preparation).

Studies of our laboratory cited in this article are supported in part by the Japanese Ministry of Education, Science and Culture and Japan Science and Technology Corporation. We are grateful to the coworkers whose names appear in the text and references.

Abrahams
,
J. P.
,
Leslie
,
A. G. W.
,
Lutter
,
R.
and
Walker
,
J. E.
(
1994
).
Structure at 2.8 Å resolution of F1 ATPase from bovine heart mitochondria
.
Nature
370
,
621
628
.
Anderson
,
R. G.
and
Orci
,
L.
(
1988
).
A view of acidic intracellular compartments
.
J. Cell Biol
.
106
,
539
543
.
Anraku
,
Y.
(
1996
).
Structure and function of the yeast vacuolar membrane H+ATPase
.
Handbook of Biological Physics
, vol.
2
(ed.
W. N.
Konings
,
H. R.
Kaback
and
J. S.
Lolkema
), pp.
93
109
.
Amsterdam
:
Elsevier Science B.V
.
Arai
,
H.
,
Terres
,
G.
,
Pink
,
S.
and
Forgac
,
M.
(
1988
).
Topography and subunit stoichiometry of the coated vesicle proton pump
.
J. Biol. Chem
.
263
,
8796
8802
.
Bajjalieh
,
S. M.
,
Frantz
,
G. D.
,
Weimann
,
J. M.
,
McConnel
,
S. K.
and
Scheller
,
R. H.
(
1994
).
Differential expression of synaptic protein isoforms
.
J. Neurosci
.
14
,
5223
5235
.
Bauerfeind
,
R.
,
Regnier-Vigouronx
,
A.
,
Flatmark
,
T.
and
Huttner
,
W. B.
(
1993
).
Selective storage of acetylcholine, but not catecholamines, in neuroendocrine synaptic-like microvesicles of early endosomal origin
.
Neuron
11
,
105
121
.
Bowman
,
B. J.
,
Vázquez-Laslop
,
N.
and
Bowman
,
E. J.
(
1992
).
The vacuolar ATPase of Neurospora crassa
.
J. Bioenerg. Biomembr
.
24
,
361
370
.
Bowman
,
E. J.
,
Siebers
,
A.
and
Altendorf
,
K.
(
1988
).
Bafilomycins: a class of inhibitors of membrane ATPases from microorganisms, animal cells and plant cells
.
Proc. Natl. Acad. Sci. USA
85
,
7972
7976
.
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
.
Broeks
,
A.
,
Janssen
,
H. W. R. M.
,
Calafat
,
J.
and
Plasterk
,
B. H. A.
(
1995
).
A P-glycoprotein protects Caenorhabditis elegans against natural toxins
.
EMBO J
.
14
,
1858
1866
.
Chatterjee
,
D.
,
Chakraborty
,
M.
,
Leit
,
M.
,
Neff
,
L.
,
Jamsakello-Kumpu
,
S.
,
Fuchs
,
R.
,
Bartkiewicz
,
M.
,
Hernando
,
N.
and
Baron
,
R.
(
1992
).
The osteoclast proton pump differs in its pharmacology and catalytic subunit from other vacuolar H+-ATPases
.
J. Exp. Biol
.
172
,
193
204
.
Davies
,
S. A.
,
Goodwin
,
S. F.
,
Kelly
,
D. C.
,
Wang
,
Z.
,
Sözen
,
M. A.
,
Kaiser
,
K.
and
Dow
,
J. A. T.
(
1996
).
Analysis and inactivation of vha55, the gene encoding the vacuolar ATPase B-subunit in Drosophila melanogaster reveals a larval lethal phenotype
.
J. Biol. Chem
.
271
,
30677
30684
.
Dow
,
J. A.
,
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
.
Dröse
,
S.
,
Bindseil
,
K. U.
,
Bowman
,
E. J.
,
Siebers
,
A.
Zeeck
,
A.
and
Altendorf
,
K.
(
1993
).
Inhibitory effect of modified bafilomycins and concanamycins on P- and V-type adenosinetriphosphatases
.
Biochemistry
32
,
3902
3906
.
Dupaix
,
A.
,
Jahannin
,
G.
and
Arrio
,
B.
(
1989
).
ATP synthesis and pyrophosphate-driven proton transport in tonoplast-enriched vesicles isolated from Catharanthres roseus
.
FEBS Lett
.
249
,
13
16
.
Façanha
,
A. R.
and
de Meis
,
L.
(
1998
).
Reversibility of H+-ATPase and H+-pyrophosphatase in tonoplast vesicles from maize coleoptiles and seeds
.
Plant Physiol
.
116
,
1487
1495
.
Feng
,
Y.
and
Forgac
,
M.
(
1992
).
Cysteine 254 of the 73 kDa A subunit is responsible for inhibition of the coated vesicle (H+)-ATPase upon modification by sulfhydryl reagents
.
J. Biol. Chem
.
267
,
5817
5822
.
Feng
,
Y.
and
Forgac
,
M.
(
1994
).
Inhibition of vacuolar H+-ATPase by disulfide bond formation between cysteine 254 and cysteine 532 in subunit A
.
J. Biol. Chem
.
269
,
13224
13230
.
Ferro-Novick
,
R. S.
and
Jahn
,
R.
(
1994
).
Vesicle fusion from yeast to man
.
Nature
370
,
191
199
.
Finbow
,
M. E.
,
Goodwin
,
S. F.
,
Meagher
,
L.
,
Lane
,
N. J.
,
Keen
,
J.
,
Findlay
,
J. B.
and
Kaiser
,
K.
(
1994
).
Evidence that the 16 kDa proteolipid (subunit c) of the vacuolar H+-ATPase and ductin from gap junctions are the same polypeptide in Drosophila and Manduca: molecular cloning of the Vha16k gene from Drosophila
.
J. Cell Sci
.
107
,
1817
1824
.
Forgac
,
M.
(
1989
).
Structure and function of vacuolar class of ATP driven proton pumps
.
Physiol. Rev
.
69
,
765
796
.
Futai
,
M.
,
Noumi
,
T.
and
Maeda
,
M.
(
1989
).
ATP synthase (H+-ATPase): results by combined biochemical and molecular biological approaches
.
Annu. Rev. Biochem
.
58
,
111
136
.
Futai
,
M.
,
Oka
,
T.
,
Moriyama
,
Y.
and
Wada
,
Y.
(
1998
).
Diverse roles of single membrane organelles: factors establishing the acid lumenal pH
.
J. Biochem
.
124
,
259
267
.
Futai
,
M.
and
Omote
,
H.
(
1996
).
F-type H+ ATPase (ATP synthase): catalytic site and energy coupling
.
Handbook of Biological Physics
, vol.
2
(ed.
W. N.
Konings
,
H. R.
Kaback
and
J. S.
Lolkema
), pp.
47
74
.
Amsterdam
:
Elsevier Science B.V
.
Gluck
,
S. L.
(
1992
).
The structure and biochemistry of the vacuolar H+ATPase in proximal and distal urinary acidification
.
J. Bioenerg. Biomembr
.
24
,
351
359
.
Goldstein
,
J. L.
,
Brown
,
M. S.
,
Anderson
,
R. G. W.
,
Russel
,
D. W.
and
Schneiden
,
W. J.
(
1985
).
Receptor-mediated endocytosis: concepts emerging from the LDL receptor system
.
Annu. Rev. Cell Biol
.
1
,
1
39
.
Graham
,
L. A.
,
Hill
,
K. J.
and
Stevens
,
T. H.
(
1995
).
VMA8 encodes a 32 kDa V1 subunit of Saccharomyces cerevisiae vacuolar H+-ATPase required for functions and assembly of the enzyme complex
.
J. Biol. Chem
.
270
,
15037
15044
.
Hanada
,
H.
,
Hasebe
,
M.
,
Moriyma
,
Y.
,
Maeda
,
M.
and
Futai
,
M.
(
1991
).
Molecular cloning of cDNA encoding the 16 kDa subunit of vacuolar H+-ATPases from mouse cerebellum
.
Biochem. Biophys. Res. Commun
.
176
,
1062
1067
.
Hanada
,
H.
,
Moriyama
,
Y.
,
Maeda
,
M.
and
Futai
,
M.
(
1990
).
Kinetic studies of chromaffin H+-ATPase and effects of bafilomycin A1
.
Biochem. Biophys. Res. Commun
.
170
,
873
818
.
Harvey
,
B. J.
(
1992
).
Energization of sodium absorption by the H+-ATPase pump in mitochondria-rich cells of frog skin
.
J. Exp. Biol
.
172
,
289
309
.
Hasebe
,
M.
,
Hanada
,
H.
,
Moriyama
,
Y.
,
Maeda
,
M.
and
Futai
,
M.
(
1992
).
Vacuolar type H+-ATPase genes: presence of four genes including pseudogenes for the 16 kDa proteolipid subunit in human genome
.
Biochem. Biophys. Res. Commun
.
183
,
856
863
.
Hasenfratz
,
M. P.
,
Tsou
,
C. L.
and
Wilkins
,
T. A.
(
1995
).
Expression of two related vacuolar H+-ATPase 16-kilodalton proteolipid genes is differentially regulated in a tissue-specific manner
.
Plant Physiol
.
108
,
1395
1404
.
Hayashi
,
M.
,
Yamamoto
,
A.
,
Yatsushiro
,
S.
,
Yamada
,
H.
,
Futai
,
M.
,
Yamaguchi
,
A.
and
Moriyama
,
Y.
(
1998
).
Synaptic vesicle protein SV2B, but not SV2A, is predominantly expressed and associated with microvesicles in rat pinealocytes
.
J. Neurochem
.
71
,
356
365
.
Hirata
,
R.
,
Graham
,
L. A.
,
Takatuki
,
A.
,
Stevens
,
T. M.
and
Anraku
,
Y.
(
1997
).
VMA11 and VMA16 encode second and third proteolipid subunits of the Saccharomyces cerevisiae vacuolar membrane H+-ATPase
.
J. Biol. Chem
.
272
,
4795
4803
.
Hirata
,
R.
,
Ohsumi
,
Y.
and
Anraku
,
Y.
(
1989
).
Functional molecular masses of vacuolar membrane H+-ATPase from Saccharomyces cerevisiae as studied by radiation inactivation analysis
.
FEBS Lett
.
244
,
397
401
.
Hirata
,
T.
,
Nakamura
,
N.
,
Wada
,
Y.
and
Futai
,
M.
(
2000
).
Regulation and reversibility of vacuolar H+-ATPase
.
J. Biol. Chem. (in press)
.
Hunt
,
I. F.
and
and Bowman
,
B. J.
(
1997
).
The intriguing evolution of the ‘b’ and ‘G’ subunits in F-type and V-type ATPases: Isolation of the vma-10 gene from Neurospora crassa
.
J. Bioenerg. Biomembr
.
29
,
533
540
.
Inoue
,
H.
,
Noumi
,
T.
,
Nagata
,
M.
,
Murakami
,
H.
and
Kanazawa
,
H.
(
1999
).
Targeted disruption of the gene encoding the proteolipid subunit of mouse vacuolar H+-ATPase leads to early embryonic lethality
.
Biochem. Biophys. Acta (in press)
.
Inatomi
,
K.
,
Eya
,
S.
,
Maeda
,
M.
and
Futai
,
M.
(
1989
).
Amino acid sequence of the α and β subunits of Methanosarcina barkeri ATPase deduced from cloned genes: similarity to subunits of eukaryotic vacuolar and FoF1-ATPases
.
J. Biol. Chem
.
264
,
10954
10959
.
Ito
,
K.
,
Kobayashi
,
T.
,
Moriyama
,
Y.
,
Toshima
,
K.
,
Tatsuta
,
K.
,
Kakiuchi
,
T.
,
Futai
,
M.
,
Ploegh
,
H. L.
and
Miwa
,
K.
(
1995
).
Concanamycin B inhibits the expression of newly-synthesized MHC class II molecules on the cell surface
.
J. Antibiotics
48
,
488
494
.
Iwamoto
,
A.
,
Orita
,
Y.
,
Maeda
,
M.
and
Futai
,
M.
(
1994
).
N-ethylmaleimide-sensitive mutant (βVal-153®Cys) Escherichia coli F1 ATPase: cross-linking of the mutant β subunit with the α subunit
.
FEBS Lett
.
352
,
243
246
.
Jones
,
P. C.
and
Fillingame
,
R. H.
(
1998
).
Genetic fusions of subunit c in the Fo sector of H+ transporting ATP synthase. Functional dimers and trimers and determination of stoichiometry by cross-linking analysis
.
J. Biol. Chem
.
273
,
29701
29705
.
Junge
,
W.
,
Lill
,
H.
and
Engelbrecht
,
S.
(
1997
).
ATP synthase: an electrochemical transducer with rotatory mechanics
.
Trends Biol. Sci
.
22
,
420
423
.
Karet
,
F. E.
,
Finberg
,
K. E.
,
Nelson
,
R. D.
,
Nayir
,
A.
,
Mocan
,
H.
,
Sanjad
,
S. A.
,
Rodriguez-Soriano
,
J.
,
Santos
,
F.
,
Cremers
,
C. W.
,
Di Pietro
,
A.
,
Hoffbrand
,
B. I.
,
Winiarski
,
J.
,
Bakkaloglu
,
A.
,
Ozen
,
S.
,
Dusunse
,
L. R.
,
Goodyer
,
P.
,
Hulton
,
S. A.
,
Wu
,
D. K.
,
Skvorak
,
A. B.
,
Morton
,
C. C.
,
Cunningham
,
M.
,
Jha
,
V.
and
Lifton
,
R. P.
(
1999
).
Mutations in the gene encoding B1 subunit of H+-ATPase cause renal tubular acidosis with sensorineural deafness
.
Nature Genetics
21
,
84
90
.
Kim
,
E. J.
,
Zhen
,
R.-G.
and
Rea
,
P. A.
(
1994
).
Heterologous expression of plant vacuolar pyrophosphatase in yeast demonstrates sufficiency of the substrate-binding subunit for proton transport
.
Proc. Natl. Acad. Sci. USA
91
,
6128
6132
.
Lee
,
C.-K.
,
Ghoshal
,
K.
and
Beaman
,
K. D.
(
1990
).
Cloning of a cDNA for a T cell produced molecule with a putative immune regulatory role
.
Mol. Immunol
.
27
,
1137
1144
.
Leng
,
X.-H.
,
Nishi
,
T.
and
Forgac
,
M.
(
1999
).
Transmembrane topology of the 100-kDa a subunit (Vph1p) of the yeast vacuolar proton-translocating ATPase
.
J. Biol. Chem
.
274
,
14655
14664
.
Li
,
Y. P.
,
Chen
,
W.
and
Stashenko
,
P.
(
1996
).
Molecular cloning and characterization of a putative novel human osteoclast-specific 116-kDa vacuolar proton pump subunit
.
Biochem. Biophys. Res. Commun
.
218
,
813
821
.
Liu
,
Q.
,
Leng
,
X.-H.
,
Neuman
,
P. R.
,
Vasilyeva
,
E.
,
Kane
,
P. M.
and
Forgac
,
M.
(
1997
).
Site-directed mutagenesis of the yeast V-ATPase A subunit
.
J. Biol. Chem
.
272
,
11750
11756
.
Mandel
,
M.
,
Moriyama
,
Y.
,
Hulmes
,
J. D.
,
Pan
,
Y.-C. E.
,
Nelson
,
H.
and
Nelson
,
N.
(
1988
).
cDNA sequence encoding the 16-kDa proteolipid of chromaffin granule implies gene duplication in the evolution of H+-ATPases
.
Proc. Natl. Acad. Sci. USA
85
,
5521
5524
.
Manolson
,
M. F.
,
Proteau
,
D.
,
Preston
,
R. A.
,
Slenbit
,
A.
and
Roberts
,
B. T.
(
1992
).
The VPH1 gene encodes a 95 kDa integral membrane polypeptide required for in vivo assembly and activity of the yeast vacuolar H+-ATPase
.
J. Biol. Chem
.
267
,
14294
14303
.
Manolson
,
M. F.
,
Wo
,
B.
,
Proteau
,
D.
,
Tailon
,
B. E.
,
Roberts
,
B. T.
,
Hoyt
,
M. A.
and
Jones
,
E. W.
(
1994
).
STV1 gene encodes functional homologue of 95-kDa yeast vacuolar H+-ATPase subunit Vph1p
.
J. Biol. Chem
.
269
,
14064
14074
.
Mellman
,
I.
,
Fuchs
,
R.
and
Helenius
,
A.
(
1986
).
Acidification of the endocytic and exocytic pathway
.
Annu. Rev. Biochem
.
55
,
663
700
.
Merzendorfer
,
H.
,
Huss
,
M.
,
Schmid
,
R.
,
Harvey
,
W. R.
and
Wieczorek
,
H.
(
1999
).
A novel insect V-ATPase subunit M9.7 is glycosylated extensively
.
J. Biol. Chem
.
274
,
17372
17378
.
Montgomery
,
M. K.
,
Xu
,
S.
and
Fire
,
A.
(
1998
).
RNA as a target of double-stranded RNA-mediated genetic interference in Caenorhabditis elegans
.
Proc. Natl. Acad. Sci. USA
95
,
15502
15507
.
Moriyama
,
Y.
,
Amakatsu
,
K.
and
Futai
,
M.
(
1993a
).
Uptake of the neurotoxin, 1-methyl-4-phenylpyridinium, a neurotoxin causing Parkinsonism, into chromafin granules and synaptic vesicles through monoamine transporter
.
Arch. Biochem. Biophys
.
305
,
271
277
.
Moriyama
,
Y.
and
Futai
,
M.
(
1990
).
H+-ATPase, a primary pump for accumulation of neurotransmitters, is a major constituent of brain synaptic vesicles
.
Biochem. Biophys. Res. Commun
.
173
,
443
448
.
Moriyama
,
Y.
,
Maeda
,
M.
and
Futai
,
M.
(
1990
).
Energy coupling of L-glutamate transport and vacuolar H+-ATPase in brain synaptic vesicles
.
J. Biochem
.
108
,
689
693
.
Moriyama
,
Y.
,
Tsai
,
H. L.
and
Futai
,
M.
(
1993b
).
Energy-dependent accumulation of neuron blockers causes selective inhibition of neurotransmitter uptakes by brain synaptic vesicles
.
Arch. Biochem. Biophys
.
305
,
278
281
.
Moriyama
,
Y.
and
Yamamoto
,
A.
(
1995
).
Microvesicles isolated from bovine pineal gland specifically accumulate L-glutamate
.
FEBS Lett
.
367
,
233
236
.
Moriyama
,
Y.
,
Yamamoto
,
A.
,
Tashiro
,
Y.
and
Futai
,
M.
(
1991
).
Chromaffin granule H+-ATPase has F1-like structure
.
FEBS Lett
.
291
,
92
96
.
Moriyama
,
Y.
,
Yamamoto
,
A.
,
Yamada
,
H.
,
Tashiro
,
Y.
and
Futai
,
M.
(
1996
).
Role of endocrine cell microvesicles in inter-cellular chemical transduction
.
Biol. Chem. Hoppe-Seyler
377
,
155
165
.
Moriyama
,
Y.
,
Yamamoto
,
A.
,
Yamada
,
H.
,
Tashiro
,
Y.
,
Tomochika
,
K.
,
Takahashi
,
M.
,
Maeda
,
M.
and
Futai
,
M.
(
1995
).
Microvesicles isolated from bovine posterior pituitary accumulate norepinephrine
.
J. Biol. Chem
.
279
,
11424
11429
.
Nelson
,
H.
,
Mandian
,
S.
and
Nelson
,
N.
(
1995
).
A bovine cDNA and yeast gene (VMA8) encoding the subunit D of vacuolar H+ ATPase
.
Proc. Natl. Acad. Sci. USA
92
,
497
501
.
Nelson
,
H.
and
Nelson
,
N.
(
1989
).
The progenitor of ATP synthases was closely related to the current vacuolar H+-ATPase
.
FEBS Lett
.
247
,
147
153
.
Nelson
,
H.
and
Nelson
,
N.
(
1990
).
Disruption of genes encoding subunits of yeast vacuolar H+ ATPase causes conditional lethality
.
Proc. Natl. Acad. Sci. USA
87
,
3503
3507
.
Nelson
,
N.
(
1992
).
Structural conservation and functional diversity of V-ATPases
.
J. Bioenerg. Biomembr
.
24
,
407
414
.
Nelson
,
N.
and
Harvey
,
W. R.
(
1999
).
Vacuolar and plasma membrane proton-adenosinetriphosphatases
.
Physiol. Rev
.
79
,
361
385
.
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. Natl. Acad. Sci. USA
89
,
3541
3545
.
Nishigori
,
H.
,
Yamada
,
S.
,
Tomura
,
H.
,
Fernald
,
A. A.
,
Le Beau
,
M. M.
,
Takeuchi
,
J.
and
Takeda
,
T.
(
1998
).
Identification and characterization of the gene encoding a second proteolipid subunit of human vacuolar H+-ATPase (ATP6F)
.
Genomics
50
,
222
228
.
Noji
,
H.
,
Yasuda
,
R.
,
Yoshida
,
M.
and
Kinoshita
,
K.
, Jr
(
1997
).
Direct observation of the rotation of F1-ATPase
.
Nature
386
,
299
302
.
Noumi
,
T.
,
Beltrán
,
C.
,
Nelson
,
H.
and
Nelson
,
N.
(
1991
).
Mutational analysis of yeast vacuolar H+-ATPase
.
Proc. Natl. Acad. Sci. USA
88
,
1938
1942
.
Oka
,
T.
,
Yamamoto
,
R.
and
Futai
,
M.
(
1997
).
Three vha genes encode proteolipids of Caenorhabditis elegans vacuolar-type ATPase. Gene structure and preferential expression in an H-shaped excretory cell and rectal cells
.
J. Biol. Chem
.
272
,
24387
24392
.
Oka
,
T.
,
Yamamoto
,
R.
and
Futai
,
M.
(
1998
).
Multiple genes for vacuolar-type ATPase proteolipids in Caenorhabditis elegans. A new gene, vha-3, has a distinct cell-specific distribution
.
J. Biol. Chem
.
273
,
22570
22576
.
Omote
,
H.
and
Futai
,
M.
(
1998
).
Mutational analysis of F1Fo ATPase: Catalysis and energy coupling
.
Acta Physiol. Scand
.
163
,
177
183
.
Omote
,
H.
,
Sambonmatsu
,
N.
,
Saito
,
K.
,
Sambongi
,
Y.
,
Iwamoto-Kihara
,
A.
,
Yanagida
,
T.
,
Wada
,
Y.
and
Futai
,
M.
(
1999
).
The γ subunit rotation and torque generation in F1-ATPase from wild-type or uncoupled mutant Escherichia coli
.
Proc. Natl. Acad. Sci. USA
96
,
7780
7784
.
Peng
,
S. B.
,
Crider
,
B. P.
,
Xie
,
X. S.
and
Stone
,
D. K.
(
1994
).
Alternative mRNA splicing generates tissue-specific isoforms of 116-kDa polypeptide of vacuolar proton pump
.
J. Biol. Chem
.
269
,
17262
17266
.
Schmidt
,
A. L.
and
Briskin
,
D. P.
(
1993
).
Reversal of the red beet tonoplast H+-ATPase by a pyrophosphate-generated proton electrochemical gradient
.
Arch. Biochem. Biophys
.
306
,
407
414
.
Schuldiner
,
S.
,
Shirvan
,
A.
and
Linial
,
M.
(
1995
).
Vesicular neurotransmitter transporters: from bacteria to human
.
Physiol. Rev
.
75
,
369
392
.
Stevens
,
T. H.
and
Forgac
,
M.
(
1997
).
Structure, function and regulation of the vacuolar (H+)-ATPases
.
Annu. Rev. Cell Dev. Biol
.
13
,
779
808
.
Thomas-Reetz
,
A.
,
Hell
,
J. W.
,
During
,
M. J.
,
Walch-Solimena
,
C.
,
Jahn
,
R.
and
De Camilli
,
P.
(
1993
).
A γ-aminobutyric acid transporter driven by a proton pump is present in synaptic-like microvesicles of pancreatic β-cells
.
Proc. Natl. Acad. Sci. USA
90
,
5317
5321
.
Tomochika
,
K.-I.
,
Shinoda
,
S.
,
Kumon
,
H.
,
Mori
,
M.
,
Moriyama
,
Y.
and
Futai
,
M.
(
1997
).
Vacuolar-type H+-ATPase in mouse bladder epithelium is responsible for urinary acidification
.
FEBS Lett
.
404
,
61
64
.
Uchida
,
E.
,
Ohsumi
,
Y.
and
Anraku
,
Y.
(
1988
).
Characterization and function of catalytic subunit A of H+-translocating adenosine triphosphatase from vacuolar membrane of Saccharomyces cerevisiae. A study with 7-chloro-4-nitrobenzo-2-oxa,1,3-diazole
.
J. Biol. Chem
.
263
,
45
51
.
Umata
,
T.
,
Moriyama
,
Y.
,
Futai
,
M.
and
Mekada
,
E.
(
1990
).
The cytotoxic action of diphtheria toxin and its degradation in intact Vero cells are inhibited by bafilomycin A1, a specific inhibitor of vacuolar-type H+-ATPase
.
J. Biol. Chem
.
265
,
21940
21945
.
Umemoto
,
N.
,
Ohya
,
Y.
and
Anraku
,
Y.
(
1991
).
VMA11, a novel gene that encodes a putative proteolipid, is indispensable for expression of yeast vacuolar membrane H+ ATPase activity
.
J. Biol. Chem
.
266
,
24526
24532
.
Wieczorek
,
H.
,
Gruber
,
G.
,
Harvey
,
W. R.
,
Huss
,
M.
and
Merzendorfer
,
H.
(
1999
).
The plasma membrane H+-V-ATPase from tobacco hornworm midgut
.
J. Bioenerg. Biomembr
.
31
,
67
74
.
Yamada
,
H.
,
Yamamoto
,
A.
,
Yodozawa
,
S.
,
Kozaki
,
S.
,
Takahashi
,
M.
,
Michibata
,
H.
,
Morita
,
M.
,
Furuichi
,
T.
,
Mikoshiba
,
K.
and
Moriyama
,
Y.
(
1996
).
Microvesicle-mediated exocytosis of glutamate is a novel paracrine-like chemical transduction mechanism and inhibits melatonin secretion in rat pinealocytes
.
J. Pineal Res
.
21
,
175
191
.
Yoshimori
,
T.
,
Yamamoto
,
A.
,
Moriyama
,
Y.
,
Futai
,
M.
and
Tashiro
,
Y.
(
1991
).
Bafilomycin A1, a specific inhibitor of vacuolar-type H+-ATPase, inhibits acidification and protein degradation in lysosomes of cultured cells
.
J. Biol. Chem
.
266
,
17707
17712
.
Zhang
,
J.
,
Feng
,
Y.
and
Forgac
,
M.
(
1994
).
Proton conduction and bafilomycin binding by the Vo domain of the coated vesicle V-ATPase
.
J. Biol. Chem
.
269
,
23518
23523
.