One of the most abundant proteins in the yeast plasma membrane is the P-type H+-ATPase that pumps protons out of the cell, supplying the driving force for a wide array of H+-dependent cotransporters. The ATPase is a 100 kDa polypeptide, anchored in the lipid bilayer by 10 transmembrane α-helices. It is structurally and functionally related to the P-type Na+,K+-, H+,K+- and Ca2+-ATPases of animal cells and the H+-ATPases of plant cells, and it shares with them a characteristic reaction mechanism in which ATP is split to ADP and inorganic phosphate (Pi) via a covalent β-aspartyl phosphate intermediate. Cryoelectron microscopic images of the H+-ATPase of Neurospora crassa and the sarcoplasmic reticulum Ca2+-ATPase of animal cells have recently been obtained at 8 nm resolution. The membrane-embedded portion of the molecule, which presumably houses the cation translocation pathway, is seen to be connected via a narrow stalk to a large, multidomained cytoplasmic portion, known to contain the ATP-binding and phosphorylation sites. In parallel with the structural studies, efforts are being made to dissect structure/function relationships in several P-type ATPases by means of site-directed mutagenesis. This paper reviews three phenotypically distinct classes of mutant that have resulted from work on the yeast PMA1 H+-ATPase: (1) mutant ATPases that are poorly folded and retained in the endoplasmic reticulum; (2) mutants in which the conformational equilibrium has been shifted from the E2 state, characterized by high affinity for vanadate, to the E1 state, characterized by high affinity for ATP; and (3) mutants with altered coupling between ATP hydrolysis and proton pumping. Although much remains to be learned before the transport mechanism can be fully understood, these mutants serve to identify critical parts of the polypeptide that are required for protein folding, conformational change and H+:ATP coupling.

Fungal plasma-membrane H+-ATPases were first detected biochemically in the late 1970s, although their existence had been predicted earlier by electrophysiological measurements of a large, ATP-and H+-dependent membrane potential in Neurospora crassa (for a review, see Rao and Slayman, 1996). Over the next several years, methods were developed to purify the H+-ATPase from three fungal species: Schizosaccharomyces pombe (Dufour and Goffeau, 1978), Saccharomyces cerevisiae (Malpartida and Serrano, 1980) and Neurospora crassa (Bowman et al., 1981; Addison and Scarborough, 1981). In all cases, the ATPase was shown to be composed of a single, 100 kDa polypeptide, firmly embedded in the lipid bilayer and requiring detergents for solubilization. Upon reconstitution into liposomes, the 100 kDa polypeptide carried out ATP-dependent proton pumping (Villalobo et al., 1981; Malpartida and Serrano, 1981; Perlin et al., 1984; Goormaghtigh et al., 1986) with a measured stoichiometry of 1 H+ translocated per ATP split (Perlin et al., 1986). Parallel studies with 32P-labelled ATP pointed to the formation of a covalent β-aspartyl phosphate intermediate as an essential part of the reaction cycle (Amory et al., 1980; Dame and Scarborough, 1980, 1981; Amory and Goffeau, 1982), accompanied by sensitivity to micromolar concentrations of inorganic orthovanadate (Bowman and Slayman, 1979). Taken together, these findings placed the fungal H+-ATPases in the group now known as the P-type cation pumps.

The next major step came in 1986, when PMA1 genes encoding the yeast and Neurospora crassa H+-ATPases were cloned and the deduced amino acid sequences were compared with those of the sarcoplasmic reticulum Ca2+-ATPase and renal plasma-membrane Na+,K+-ATPase (Serrano et al., 1986; Hager et al., 1986; Addison, 1986). Hydropathy analysis led to a common topological model for all three enzymes (Fig. 1; for a review, see Lutsenko and Kaplan, 1995), with the 100 kDa polypeptide anchored in the membrane by four hydrophobic segments at the N-terminal end of the molecule and six at the C-terminal end. The central region protrudes into the cytoplasm and contains strongly conserved motifs for ATP binding and phosphorylation.

Fig. 1.

Topological model of the fungal plasma-membrane H+-ATPase. Residues highlighted in blue are present in the PMA1 and PMA2 ATPases from Saccharomyces cerevisiae (accession numbers P05030 and P19657), PMA1 and PMA2 ATPases from Schizosaccharomyces pombe (P09627 and P28876), PMA1 ATPases from Neurospora crassa (P07038), Zygosaccharomyces rouxii (P24545), Kluyveromyces lactis (P49380), Candida albicans (P28877), Pichia augusta (AF109913) and Histoplasma capsulatum (Q07421), PCA1 ATPase from Pneumocystis carinii (AAB06958) and PMAA ATPase from Aspergillus (Emericella) nidulans (AAC27991). M1–M10, membrane-spanning segments; P, phosphorylation site; ATP, ATP-binding site.

Fig. 1.

Topological model of the fungal plasma-membrane H+-ATPase. Residues highlighted in blue are present in the PMA1 and PMA2 ATPases from Saccharomyces cerevisiae (accession numbers P05030 and P19657), PMA1 and PMA2 ATPases from Schizosaccharomyces pombe (P09627 and P28876), PMA1 ATPases from Neurospora crassa (P07038), Zygosaccharomyces rouxii (P24545), Kluyveromyces lactis (P49380), Candida albicans (P28877), Pichia augusta (AF109913) and Histoplasma capsulatum (Q07421), PCA1 ATPase from Pneumocystis carinii (AAB06958) and PMAA ATPase from Aspergillus (Emericella) nidulans (AAC27991). M1–M10, membrane-spanning segments; P, phosphorylation site; ATP, ATP-binding site.

Until recently, the model of Fig. 1 was based only on limited experimental evidence. In 1989, Mandala and Slayman raised antibodies against the N and C termini of the Neurospora crassa H+-ATPase and showed that they could bind to inside-out plasma-membrane vesicles but not to intact protoplasts. Likewise, sites close to the N and C termini could be cleaved by trypsin in vesicles but not in protoplasts (Mandala and Slayman, 1988, 1989). These two sets of findings established that the ATPase must have an even number of membrane-spanning segments, with both termini exposed at the cytoplasmic surface of the membrane. Further information came from studies by Lin and Addison (1994), who examined the ability of fusion constructs from the Neurospora crassa H+-ATPase to undergo in vitro membrane insertion and glycosylation. As predicted by the model shown in Fig. 1, membrane-spanning segments M5 and M7 behaved as initiate-transfer sequences and M6, M8 and M10 as stop-transfer sequences. However, M9 failed to bring about transfer in vitro even though it became associated with the microsomal membranes, and M10 was able to initiate transfer as well as to stop transfer. Thus, although the data were generally consistent with a model for the H+-ATPase containing 10 membrane-spanning segments, it was obvious that another, more direct, approach was needed.

Very recently, an important step was taken by Scarborough, Kuhlbrandt and co-workers, who grew two-dimensional crystals of purified Neurospora crassa H+-ATPase, analyzed the crystals using cryoelectron microscopy and successfully computed a three-dimensional map at 8 nm resolution (Auer et al., 1998). The results confirmed the presence of 10 membrane-spanning α-helices, tilted at varying angles from the plane of the bilayer; four of the 10 helices twisted around one another in a distinctive righthanded bundle. The cytoplasmic portion of the ATPase had four distinct domains and was linked to the membrane by a complex stalk-like structure. Similar results were reported for the sarcoplasmic reticulum Ca2+-ATPase by Zhang et al. (1998).

Vigorous efforts are now under way to produce higher-resolution structures of P-ATPases. In the meantime, this paper will summarize the progress that has been made towards mapping structure/function relationships within the yeast H+-ATPase by means of site-directed mutagenesis.

Immediately after the yeast PMA1 gene had been cloned, work began in several laboratories to construct and characterize site-directed mutants in order to examine the functional role of individual amino acid residues and motifs throughout the protein. It quickly became clear that a special expression strategy was needed because of the central role played by the H+-ATPase in cell growth. In particular, the ATPase has been estimated to consume as much as one-quarter of cellular ATP, pumping H+ outwards and generating the electrochemical gradient that drives H+-dependent nutrient uptake (for a review, see Rao and Slayman, 1996). Thus, knockout of the PMA1 gene is lethal (Serrano et al., 1986), and a variety of studies have established that the ATPase must have at least 30 % of the usual wild-type activity to support cell growth (D. S. Perlin, personal communication; C. W. Slayman’s laboratory, unpublished results).

To make it possible to study a full range of mutants, including those with low activity, our laboratory undertook to develop a transient expression system for the PMA1 ATPase and other homologous and heterologous plasma membrane proteins (Nakamoto et al., 1991). As illustrated in Fig. 2, this approach relies upon a temperature-sensitive mutation in the SEC6 gene to arrest newly synthesized mutant ATPase in intracellular secretory vesicles, just prior to fusion with the plasma membrane. The vesicles are abundant and readily isolated by gel filtration (Walworth and Novick, 1987) or sucrose gradient centrifugation (Ambesi et al., 1997). Equally important, they are uniformly oriented inside-out relative to the plasma membrane, allowing ATP-dependent H+ pumping to be assayed by Acridine Orange fluorescence quenching (Nakamoto et al., 1991; Ambesi et al., 1997). In the work to be described, PMA1 mutants were constructed, expressed in secretory vesicles and examined in detail for their ability to split ATP and pump protons.

Fig. 2.

Expression of H+-ATPase mutants in secretory vesicles. (A) Yeast cell with nucleus (N), endoplasmic reticulum (ER), Golgi (G), secretory vesicles (SV) and plasma membrane (PM). As described by Nakamoto et al. (1991), the wild-type copy of the PMA1 gene is placed under the control of the GAL1 promoter, and the mutant copy is introduced on a centromeric plasmid under heat-shock control. Fusion of the secretory vesicles to the plasma membrane is controlled by a temperature-sensitive allele of the SEC6 gene. (B) Secretory vesicle illustrating its inside-out orientation relative to the plasma membrane.

Fig. 2.

Expression of H+-ATPase mutants in secretory vesicles. (A) Yeast cell with nucleus (N), endoplasmic reticulum (ER), Golgi (G), secretory vesicles (SV) and plasma membrane (PM). As described by Nakamoto et al. (1991), the wild-type copy of the PMA1 gene is placed under the control of the GAL1 promoter, and the mutant copy is introduced on a centromeric plasmid under heat-shock control. Fusion of the secretory vesicles to the plasma membrane is controlled by a temperature-sensitive allele of the SEC6 gene. (B) Secretory vesicle illustrating its inside-out orientation relative to the plasma membrane.

As the list of known fungal H+-ATPase genes has grown, it has become possible to obtain a provisional view of functionally important regions by aligning the deduced amino acid sequences and mapping the residues that have remained unchanged over 1.1 billion years of fungal evolution. Fig. 1 shows that only 40 % of residues are conserved throughout a group of 10 fungal PMA1 and PMA2 ATPases. The identical residues are scattered along the small cytoplasmic loop between M2 and M3 and the large loop between M4 and M5, sometimes in regions with recognizable function (e.g. the ATP-binding and phosphorylation sites) but often with no direct evidence as to functional role. Most of the membrane-spanning segments have also been relatively well conserved, while three (M3, M7 and M10) have diverged considerably.

On the basis of a variety of considerations, it seems likely that the 60-amino-acid stretch from the start of M4 to the end of the phosphorylation region and the 60-amino-acid stretch from stalk segment 5 (S5) to M6 are central to the reaction mechanism of the H+-ATPase. Aspartyl and glutamyl residues in M4, M5 and M6 of the mammalian Na+,K+-, H+,K+-and Ca2+-ATPases are now known to be required for high-affinity cation binding (Clarke et al., 1989, 1990; Nielsen et al., 1998), making it reasonable to think that these helices line the transport pathway of the H+-ATPase as well. Furthermore, the stalk segments linking M4 and M5 to the central catalytic domain of the ATPase are strong candidates for transmitting the E1→E2 conformational change into the membrane. We have performed alanine-scanning (or, in the case of S5, cysteine-scanning) mutagenesis along the entire length of both regions to maximize the chance of identifying functionally important residues. Further details about many of the mutants can be found in recent papers by Ambesi et al. (1996), Nakamoto et al. (1998), Dutra et al. (1998), DeWitt et al. (1998) and Sen Gupta et al. (1998); results on the S4 and S5 mutants are now being prepared for publication.

Because some mutations could prevent proper folding or insertion of newly synthesized ATPase into the endoplasmic reticulum membrane, the first step was to determine the relative amount of mutant protein that reached the secretory vesicles. This was achieved by quantitative immunoblotting with polyclonal ATPase antibody. Indeed, mutations at 12 of the 120 positions studied completely prevented normal biogenesis. In most cases, the mutant polypeptides have been epitope-tagged and shown by indirect immunofluorescence to accumulate in intracellular structures derived from the endoplasmic reticulum (e.g. Harris et al., 1994; Portillo, 1997; DeWitt et al., 1998; Sen Gupta et al., 1998). The same mutants are exceedingly sensitive to low concentrations of trypsin, consistent with the idea that they are poorly folded and arrested by the endoplasmic reticulum quality control machinery. Many of them behave genetically in a dominant lethal fashion, causing co-expressed wild-type ATPase to be trapped in the endoplasmic reticulum (Harris et al., 1994; Portillo, 1997; DeWitt et al., 1998); studies are under way to determine whether the wild-type and mutant forms interact directly by forming oligomers or whether they compete for some component of the endoplasmic reticulum biosynthetic machinery such as a translocation channel or chaperone.

Perhaps the most unexpected finding has been the nearly uninterrupted stretch of biogenesis-defective mutants in the phosphorylation region of the H+-ATPase, from Asp378 (the phosphorylation site) to Thr384. When the data described above were superimposed upon secondary structural predictions for this region, there was a clear correspondence between the cluster of biogenesis-defective mutants and a likely β-turn in the vicinity of Asp378. According to the algorithm of Garnier et al. (1978), substitution by Ala at each position should introduce a more helical nature to the β-turn and could therefore perturb the structure of the cytoplasmic loop, preventing the insertion of M5–M10 into the lipid bilayer. It will be interesting to see whether these predictions are borne out when a high-resolution structure becomes available for the H+-ATPase.

For two other residues that are essential for normal folding and biogenesis, further studies have uncovered a likely mechanism. When either Arg695 in M5 or Asp730 in M6 is replaced by a neutral amino acid, the ATPase becomes trypsin-sensitive and is retained in the endoplasmic reticulum. But when Arg695 and Asp730 are replaced simultaneously by Ala, or when they are swapped to produce an R695D, D730R double mutant, the ATPase undergoes proper maturation and displays normal rates of ATP hydrolysis and ATP-dependent H+ pumping (Sen Gupta et al., 1998). These results indicate that neither Arg695 nor Asp730 is an essential residue; instead, the two take part in a charge–charge interaction, presumably lending stability to the M5–M6 hairpin.

Also encountered in the present study have been mutants with a characteristic pattern of kinetic changes, including an elevated IC50 for orthovanadate, a decreased Km for MgATP and, in some cases, a higher-than-normal pH optimum for ATP hydrolysis (Ambesi et al., 1996; Dutra et al., 1998; A. Ambesi, M. Miranda and C. W. Slayman, in preparation). The most likely interpretation is that these mutants have undergone a shift in equilibrium from the E2 conformation, which is expected to bind vanadate with high affinity and ATP with low affinity, towards the E1 conformation, which should have a high affinity for ATP and a low affinity for vanadate. Whether the alkaline pH optimum seen in some but not all of the mutants reflects a greater affinity of the E1 form for the transported proton, or whether it is simply a spurious pKa change at some unrelated site, remains to be established.

An interesting aspect of the E1→E2 mutants is their distribution within the ATPase polypeptide (Fig. 3). Recent work (A. Ambesi, M. Miranda and C. W. Slayman, in preparation) has identified a stretch of eight consecutive residues in the stalk 4 region along which mutations elevate the IC50 for vanadate from 1 μmol l−1 to 10–200 μmol l−1 and decrease the Km for MgATP from 1.5 mmol l−1 to 0.1–0.5 mmol l−1. This stretch contains Ser368, where a mutation to Phe had previously been isolated on the basis of hygromycin B resistance and found to display similar IC50 and Km changes (Perlin et al., 1989). The properties of the kinetically altered mutants suggest that the middle of stalk segment 4 may play a critical role in the E1→E2 transition and thus in communication between the cytoplasmic and membrane portions of the H+-ATPase.

Fig. 3.

Region from membrane-spanning segments M4 to M6 of the yeast PMA1 H+-ATPase, indicating residues that have been studied using site-directed mutagenesis (see text). Mutations that disrupt protein folding and biogenesis are highlighted in black; mutations that alter the E1/E2 conformational equilibrium are highlighted in red; and mutations that affect H+:ATP coupling are highlighted in blue. S4, S5, stalk 4 and stalk 5 regions.

Fig. 3.

Region from membrane-spanning segments M4 to M6 of the yeast PMA1 H+-ATPase, indicating residues that have been studied using site-directed mutagenesis (see text). Mutations that disrupt protein folding and biogenesis are highlighted in black; mutations that alter the E1/E2 conformational equilibrium are highlighted in red; and mutations that affect H+:ATP coupling are highlighted in blue. S4, S5, stalk 4 and stalk 5 regions.

A third type of mutant, found during a systematic survey of charged residues within the transmembrane segments of the H+-ATPase (V. V. Petrov, K. P. Padmanabha, R. K. Nakamoto and C. W. Slayman, in preparation), is altered in H+:ATP coupling. When secretory vesicles expressing the wild-type ATPase are incubated with increasing concentrations of MgATP, a linear relationship is seen between the initial rate of ATP-dependent Acridine Orange fluorescence quenching (as a measure of H+ transport) and the initial rate of ATP hydrolysis, consistent with the idea that the pump operates at a constant stoichiometry over a considerable range of velocities (Ambesi et al., 1996). For most mutants studied in the same way, the data points have been found to fall along the wild-type line, suggesting little or no change in H+:ATP coupling. For E703Q, E703L and E803N, however, the slope of the line was significantly lowered, pointing to a partial uncoupling of the pump; and for E803Q, the slope was elevated, as if the pump had somehow increased its functional stoichiometry (V. V. Petrov, K. P. Padmanabha, R. K. Nakamoto and C. W. Slayman, in preparation).

Glu703 and Glu803 are of special interest because the corresponding residues have been implicated in high-affinity cation binding by mammalian Ca2+-and Na+,K+-ATPases (Clarke et al., 1989, 1990; Nielsen et al., 1998). Studies are under way to express the E703Q, E703L, E803N and E803Q H+-ATPases in the yeast plasma membrane, where it should be possible to analyze the charge-transfer step of the transport cycle by patch-clamping and therefore to test the effect of the mutations directly.

The mutants described in this brief review, together with mutants that have been constructed to study M1 and M2 (Seto-Young et al., 1996), the small cytoplasmic loop (Wang et al., 1996), the stalk segments extending into the cytoplasm from M2 and M3 (Soteropoulos and Perlin, 1998) and several other parts of the polypeptide (Maldonado and Portillo, 1995; Portillo, 1997), have produced a broad picture of structure/function relationships throughout the H+-ATPase. Important questions remain concerning the architecture of the ATP-binding and phosphorylation sites, the way in which ATP hydrolysis and phosphorylation lead to the E1→E2 conformational change, the coupling between that change and cation transport, and the detailed pathway along which cations move through the membrane. Progress towards answering these questions will speed up dramatically once high-resolution structures are available for the E1 and E2 conformations, which will provide a framework for refining existing data and for pursuing new biochemical, biophysical and genetic tests of the pump mechanism.

Work in our laboratory has been supported by NIH grant GM15761 and by postdoctoral fellowships to A.A. (NIH), M.M. (Fogarty) and V.P. (Yale School of Medicine).

Addison
,
R.
(
1986
).
Primary structure of the Neurospora plasma membrane H+-ATPase deduced from gene sequence
.
J. Biol. Chem
.
261
,
14896
14901
.
Addison
,
R.
and
Scarborough
,
G. A.
(
1981
).
Solubilization and purification of the Neurospora plasma-membrane H+-ATPase
.
J. Biol. Chem
.
257
,
10421
10426
.
Ambesi
,
A.
,
Allen
,
K. E.
and
Slayman
,
C. W.
(
1997
).
Isolation of transport-competent secretory vesicles from Saccharomyces cerevisiae
.
Analyt. Biochem
.
251
,
127
129
.
Ambesi
,
A.
,
Pan
,
R. L.
and
Slayman
,
C. W.
(
1996
).
Alanine-scanning mutagenesis along membrane segment 4 of the yeast plasma membrane H+-ATPase
.
J. Biol. Chem
.
271
,
22999
23005
.
Amory
,
A.
,
Foury
,
F.
and
Goffeau
,
A.
(
1980
).
The purified plasma membrane ATPase of the yeast Schizosaccharomyces pombe forms a phosphorylated intermediate
.
J. Biol. Chem
.
255
,
9353
9357
.
Amory
,
A.
and
Goffeau
,
A.
(
1982
).
Characterization of the β-aspartyl phosphate intermediate formed by the H+-translocating ATPase from the yeast Schizosaccharomyces pombe
.
J. Biol. Chem
.
257
,
4723
4730
.
Auer
,
M.
,
Scarborough
,
G. A.
and
Kuhlbrandt
,
W.
(
1998
).
Three-dimensional map of the plasma membrane H+-ATPase in the open conformation
.
Nature
392
,
840
843
.
Bowman
,
B. J.
,
Blasco
,
F.
and
Slayman
,
C. W.
(
1981
).
Purification and characterization of the plasma membrane ATPase of Neurospora crassa
.
J. Biol. Chem
.
256
,
12343
12349
.
Bowman
,
B. J.
and
Slayman
,
C. W.
(
1979
).
The effects of vanadate on the plasma membrane ATPase of Neurospora crassa
.
J. Biol. Chem
.
254
,
2928
2934
.
Clarke
,
D. M.
,
Loo
,
T. W.
,
Inesi
,
G.
and
MacLennan
,
D. H.
(
1989
).
Location of high affinity Ca2+-binding sites within the predicted transmembrane domain of the sarcoplasmic reticulum Ca2+-ATPase
.
Nature
339
,
476
478
.
Clarke
,
D. M.
,
Loo
,
T. W.
and
MacLennan
,
D. H.
(
1990
).
Functional consequences of alterations to polar amino acids located in the transmembrane domain of the Ca2+-ATPase of sarcoplasmic reticulum
.
J. Biol. Chem
.
265
,
6262
6267
.
Dame
,
J. B.
and
Scarborough
,
G. A.
(
1980
).
Identification of the hydrolytic moiety of the Neurospora plasma membrane H+-ATPase and demonstration of a phosphoryl-enzyme intermediate in its catalytic mechanism
.
Biochemistry
19
,
2931
2937
.
Dame
,
J. B.
and
Scarborough
,
G. A.
(
1981
).
Identification of the phosphorylated intermediate of the Neurospora plasma membrane H+-ATPase as β-aspartyl phosphate
.
J. Biol. Chem
.
256
,
10724
10730
.
DeWitt
,
N. D.
,
Tourinho dos Santos
,
C. F.
,
Allen
,
K. E.
and
Slayman
,
C. W.
(
1998
).
Phosphorylation region of the yeast plasma-membrane ATPase. Role in protein folding and biogenesis
.
J. Biol. Chem
.
273
,
21744
21751
.
Dufour
,
J. P.
and
Goffeau
,
A.
(
1978
).
Solubilization by lysolecithin and purification of the plasma membrane ATPase of the yeast Schizosaccharomyces pombe
.
J. Biol. Chem
.
253
,
7026
7032
.
Dutra
,
M. B.
,
Ambesi
,
A.
and
Slayman
,
C. W.
(
1998
).
Structure–function relationships in membrane segment 5 of the yeast Pma1 H+-ATPase
.
J. Biol. Chem
.
273
,
17411
17417
.
Garnier
,
J.
,
Osguthorpe
,
D. J.
and
Robson
,
B.
(
1978
).
Analysis of the accuracy and implications of simple methods for predicting the secondary structure of globular proteins
.
J. Mol. Biol
.
120
,
97
120
.
Goormaghtigh
,
E.
,
Chadwick
,
C.
and
Scarborough
,
G. A.
(
1986
).
Monomers of the Neurospora plasma membrane H+-ATPase catalyze efficient proton translocation
.
J. Biol. Chem
.
261
,
7466
7471
.
Hager
,
K. M.
,
Mandala
,
S. M.
,
Davenport
,
J. W.
,
Speicher
,
D. E.
,
Benz
,
E. J.
and
Slayman
,
C. W.
(
1986
).
Amino acid sequence of the plasma membrane ATPase of Neurospora crassa: deduction from genomic and cDNA sequences
.
Proc. Natl. Acad. Sci. USA
83
,
7693
7697
.
Harris
,
S. L.
,
Na
,
S.
,
Zhu
,
X.
,
Seto-Young
,
D.
,
Perlin
,
D. S.
and
Haber
,
J. E.
(
1994
).
Dominant lethal mutations in the plasma membrane H+-ATPase gene of Saccharomyces cerevisiae
.
Proc. Natl. Acad. Sci. USA
91
,
10531
10535
.
Lin
,
J.
and
Addison
,
R.
(
1994
).
Topology of the Neurospora plasma membrane H+-ATPase. Localization of a transmembrane segment
.
J. Biol. Chem
.
269
,
3887
3890
.
Lutsenko
,
S.
and
Kaplan
,
J. H.
(
1995
).
Organization of P-type ATPases: significance of structural diversity
.
Biochemistry
34
,
15607
15613
.
Maldonado
,
A.
and
Portillo
,
F.
(
1995
).
Genetic analysis of the fluorescein isothiocyanate binding site of the yeast plasma membrane H+-ATPase
.
J. Biol. Chem
.
270
,
8655
8659
.
Malpartida
,
F.
and
Serrano
,
R.
(
1980
).
Purification of the yeast plasma membrane ATPase solubilized with a novel zwitterionic detergent
.
FEBS Lett
.
111
,
69
72
.
Malpartida
,
F.
and
Serrano
,
R.
(
1981
).
Proton translocation catalyzed by the purified yeast plasma membrane ATPase reconstituted in liposomes
.
FEBS Lett
.
131
,
351
354
.
Mandala
,
S. M.
and
Slayman
,
C. W.
(
1988
).
Identification of tryptic cleavage sites for two conformational states of the Neurospora plasma membrane H+-ATPase
.
J. Biol. Chem
.
263
,
15122
15128
.
Mandala
,
S. M.
and
Slayman
,
C. W.
(
1989
).
The amino and carboxy termini of the Neurospora plasma membrane H+-ATPase are cytoplasmically located
.
J. Biol. Chem
.
264
,
16276
16281
.
Nakamoto
,
R. K.
,
Rao
,
R.
and
Slayman
,
C. W.
(
1991
).
Expression of the yeast plasma membrane [H+] ATPase in secretory vesicles. A new strategy for directed mutagenesis
.
J. Biol. Chem
.
266
,
6425
6428
.
Nakamoto
,
R. K.
,
Verjovski-Almeida
,
S.
,
Allen
,
K. E.
,
Ambesi
,
A.
,
Rao
,
R.
and
Slayman
,
C. W.
(
1998
).
Substitutions of aspartate-378 in the phosphorylation domain of the yeast PMA1 H+-ATPase disrupt protein folding and biogenesis
.
J. Biol. Chem
.
273
,
7338
7344
.
Nielsen
,
J. M.
,
Pedersen
,
P. A.
,
Karlish
,
S. J. D.
and
Jorgensen
,
P. L.
(
1998
).
Importance of intramembrane carboxylic acids for occlusion of K+ ions at equilibrium in renal Na+,K+-ATPase
.
Biochemistry
37
,
1961
1968
.
Perlin
,
D. S.
,
Harris
,
S. L.
,
Seto-Young
,
D.
and
Haber
,
J. E.
(
1989
).
Defective H+-ATPase of hygromycin B-resistant pma1 mutants from Saccharomyces cerevisiae
.
J. Biol. Chem
.
264
,
21857
21864
.
Perlin
,
D. S.
,
Kasamo
,
K.
,
Brooker
,
R. J.
and
Slayman
,
C. W.
(
1984
).
Electrogenic H+ translocation by plasma membrane ATPase of Neurospora. Studies on plasma membrane vesicles and reconstituted enzyme
.
J. Biol. Chem
.
259
,
7884
7892
.
Perlin
,
D. S.
,
San Francisco
,
M. J. D.
,
Slayman
,
C. W.
and
Rosen
,
B. P.
(
1986
).
H+/ATP stoichiometry of proton pumps from Neurospora crassa and Escherichia coli
.
Arch. Biochem. Biophys
.
248
,
53
61
.
Portillo
,
F.
(
1997
).
Characterization of dominant lethal mutations in the yeast plasma membrane H+-ATPase gene
.
FEBS Lett
.
402
,
136
140
.
Rao
,
R.
and
Slayman
,
C. W.
(
1996
).
Plasma membrane and related ATPases
.
In The Mycota
, vol.
III
(ed.
R.
Brambl
and
G.
Marzluf
), pp.
29
56
.
Berlin
:
Springer-Verlag
.
Sen Gupta
,
S.
,
DeWitt
,
N. D.
,
Allen
,
K. E.
and
Slayman
,
C. W.
(
1998
).
Evidence for a salt bridge between transmembrane segments 5 and 6 of the yeast plasma-membrane H+-ATPase
.
J. Biol. Chem
.
273
,
34328
34334
.
Serrano
,
R.
,
Kielland-Brandt
,
M. C.
and
Fink
,
G. R.
(
1986
).
Yeast plasma membrane ATPase is essential for growth and has homology with (Na++K+)-, K+-and Ca2+-ATPases
.
Nature
319
,
689
693
.
Seto-Young
,
D.
,
Hall
,
M. J.
,
Na
,
S.
,
Haber
,
J. E.
and
Perlin
,
D. S.
(
1996
).
Genetic probing of the first and second transmembrane helices of the plasma membrane H+-ATPase from Saccharomyces cerevisiae
.
J. Biol. Chem
.
271
,
581
587
.
Soteropoulos
,
P.
and
Perlin
,
D. S.
(
1998
).
Genetic probing of stalk segments associated with M2 and M3 of the plasma membrane H+-ATPase from Saccharomyces cerevisiae
.
J. Biol. Chem
.
273
,
26426
26431
.
Villalobo
,
A.
,
Boutry
,
M.
and
Goffeau
,
A.
(
1981
).
Electrogenic proton translocation coupled to ATP hydrolysis by the plasma membrane Mg2+-dependent ATPase of yeast in reconstituted proteoliposomes
.
J. Biol. Chem
.
256
,
12081
12087
.
Walworth
,
N. C.
and
Novick
,
P. J.
(
1987
).
Purification and characterization of constitutive secretory vesicles from yeast
.
J. Cell Biol
.
105
,
163
174
.
Wang
,
G.
,
Tamas
,
M. J.
,
Hall
,
M. J.
,
Pascual-Ahuir
,
A.
and
Perlin
,
D. S.
(
1996
).
Probing conserved regions of the cytoplasmic LOOP1 segment linking transmembrane segments 2 and 3 of the Saccharomyces cerevisiae plasma membrane H+-ATPase
.
J. Biol. Chem
.
271
,
25438
25445
.
Zhang
,
P.
,
Toyoshima
,
C.
,
Yonekura
,
K.
,
Green
,
N. M.
and
Stokes
,
D. L.
(
1998
).
Structure of the calcium pump from sarcoplasmic reticulum at 8 Å resolution
.
Nature
392
,
835
839
.