Main

Bacteria use multiple systems to maintain K+ homeostasis2. TrkH and KtrB are gated channels that belong to the superfamily of K+ transporters (SKT)3, which supply K+ to the cell under normal growth conditions. When K+ concentrations fall into the micromolar range, many bacteria use the inducible Kdp system, which produces the four-subunit KdpFABC membrane complex that actively drives K+ into the cell. This complex has high selectivity, binding affinity in the low micromolar range and can maintain cytoplasmic K+ concentrations against up to 104-fold gradients4. Mutagenesis has been used to establish that K+ is transported through KdpA5,6 and that the energy of ATP is harnessed by KdpB7. These subunits are joined by KdpC, which has been proposed to be a catalytic chaperone8, and KdpF9; both have single transmembrane helices and no known homologues outside Kdp.

As a P-type ATPase, KdpB operates according to the Post–Albers scheme, which involves two main conformational states: E1 and E2 (ref. 10). In the E1 state, ATP is bound by the cytoplasmic domains in order to autophosphorylate a conserved aspartate, thus stepping to E1P; this high-energy phosphoenzyme is typically formed in response to cytoplasmic ions binding at a canonical transmembrane site. The energy is used in converting E1P to E2P, where ion binding sites are exposed to the other side of the membrane with lowered affinity. After ions leave, the aspartyl phosphate is hydrolysed to produce E2, which then reverts back to E1 to complete the cycle. As K+ is bound by a different subunit in KdpFABC, it is unclear whether E1P formation in KdpB is associated with ion binding from the periplasm or with release to the cytoplasm by KdpA. Furthermore, it is unclear whether counterions, which generally facilitate E2 formation in other P-type ATPases, are involved in this process.

K+ is expected to move through KdpA, like all members of the SKT family, by way of a selectivity filter descended from that of the bacterial channel KcsA3. The selectivity filter has multiple, tandem binding sites for dehydrated K+ ions that are derived from four repeated M1PM2 motifs, in which two transmembrane helices (M) sandwich a reentrant pore helix (P). Whereas the KcsA channel is a homotetramer, TrkH, KtrB and KdpA are all single polypeptides with four pseudo repeats (D1–D4). Structures of TrkH and KtrB11,12 show a kinked helix in the third repeat (D3M2) with a loop that forms a regulatory gate on the cytoplasmic side of the selectivity filter13.

For this work, we used the KdpFABC complex from E. coli carrying the Gln116Arg mutation in KdpA. This mutant exhibits lowered apparent K+ affinity (Michaelis constant (Km) = 6 mM versus 10 μM for wild-type)5 and has been widely used in previous biochemical studies4,7,14. The structure was solved by X-ray crystallography to 2.9 Å resolution using experimental phases from tungsten and mercury with an Rfree of 27.5% (Extended Data Table 1). The large asymmetric unit contains three KdpFABC complexes that adopt identical conformations (Extended Data Fig. 1). KdpA has ten transmembrane helices with four M1PM2 repeats (D1–D4) and a K+ ion bound in the central selectivity filter (Fig. 1). KdpB has seven transmembrane helices (bM1–bM7) and three cytoplasmic domains found in all P-type ATPases: phosphorylation (P) domain, nucleotide-binding (N) domain, and dephosphorylating actuator (A) domain15. KdpC has a single transmembrane helix (cM1), but the topology is inverted relative to previous models (Extended Data Fig. 1d). This topology puts the soluble domain, which appears to have a novel fold, on the periplasmic side of the membrane. KdpF is a single transmembrane helix with a position that is distinct from those of transmembrane helices or accessory elements in other P-type ATPases. KdpF is not present in some species16 and the E. coli complex lacking KdpF is fully functional in the membrane-bound state or after addition of lipids9, suggesting that it plays a structural role in stabilizing the complex.

Figure 1: Overview of the KdpFABC complex.
figure 1

a, ATPase activity using purified, detergent-solubilized KdpFABC demonstrates robust stimulation by K+ with an apparent affinity of 24.8 mM and Vmax of 6.6 μmol mg−1 min−1. There is negligible activity with other ions. Error bars indicate s.e., n = 3. b, KdpA (green) has a K+ ion bound in the middle of the membrane (purple sphere). KdpB has a transmembrane domain (brown) and three cytoplasmic domains responsible for nucleotide binding (N, red), phosphoenzyme formation (P, blue) and dephosphorylation (A, yellow). KdpC (purple) and KdpF (cyan) have a single transmembrane helix each and the soluble domain of KdpC is periplasmic. c, Transmembrane topology diagram, coloured as in b.

PowerPoint slide

KdpA displays pseudo-four-fold symmetry with M1PM2 motifs surrounding a central selectivity filter (Fig. 2a). A large number of mutations that affect apparent K+ affinity5,6 map directly onto the selectivity filter or the associated pore helices (Extended Data Fig. 2). The presence of a K+ ion within the selectivity filter is evidenced by a strong density peak that colocalizes with an anomalous density peak (Extended Data Fig. 1e). In K+ channels, four distinct sites (designated S1–S4) are characterized by ‘rings’ of oxygen ligands above and below each site17 (Extended Data Fig. 3). In KdpA, most of these oxygens contribute to the selectivity filter and the observed K+ ion occupies site S3 (Fig. 2b). Sites S2 and S4 lack ions and are distorted, whereas the S1 site is occupied by the charged side chain of mutated Arg116, which is hydrogen bonded to selectivity filter carbonyls (Gly232, Gly345, Gly468) and to Asn239, all of which produce lowered K+ affinity when mutated5,6. The position of the Arg116 side chain suggests that it is a surrogate for a K+ ion in the S1 site. Thus, this side chain is likely to interfere with K+ ions entering the selectivity filter in the Gln116Arg mutant, whereas this site would assist in recruiting K+ to the selectivity filter of the native protein and could thus contribute to its high apparent affinity.

Figure 2: Potassium binding by KdpA.
figure 2

a, KdpA, as seen from the periplasmic side, has four M1PM2 units (D1–D4) that form the selectivity filter (yellow backbone) and bind K+ (purple sphere). KdpB (brown) includes a water molecule (red sphere) at its canonical cation binding site. b, Side view of the KdpA selectivity filter shows Arg116 in the S1 site and main chain carbonyls coordinating K+ in the S3 site. c, Side view shows the gating loop below the selectivity filter with the coupling helix (pink). On the periplasmic side, the soluble domain of KdpC (purple) is firmly held by loops from repeats D2 (dark green) and D3 (light green) of KdpA (Extended Data Fig. 2).

PowerPoint slide

Unlike permeation through channels, the reaction cycle of an ATP-driven transporter involves transient occlusion of ions in order to prevent the leakage of ions and futile ATP hydrolysis10. Such occlusion requires gating elements on both sides of the binding site. Towards the cytoplasm, the selectivity filter of KdpA is blocked by a loop directly below the K+ ion (Fig. 2c). As in TrkH and KtrB, this loop is derived from D3M2, which forms a kinked helix (Fig. 1c, Extended Data Fig. 3). On the periplasmic side, access to the filter is unimpeded, except by Arg116, which is not present in the native sequence. Nevertheless, the soluble domain of KdpC is held nearby by two loops from the D2 and D3 repeats of KdpA (Fig. 2c). The robustness of the hydrogen bond network that mediates this interaction may explain why KdpA and KdpC can be co-purified in the absence of KdpB18. Furthermore, this interaction suggests that relative movements between the repeats of KdpA could move the soluble domain of KdpC into an occluding position in response to the relevant conformational change in KdpB.

KdpB is characterized by seven transmembrane helices, the first six of which are consistent with the ‘core’ of other P-type ATPases19. The middle of bM4 is unwound at the conserved proline motif (IP264TTI) and is very similar to M4 of Ca2+-ATPase (SERCA1a) in the Ca2+-bound E1 state10 (Extended Data Fig. 4). We observe a strong density peak at the unwound part of bM4 (Extended Data Fig. 1f). Neither K+ nor Na+ fit the binding geometry and both introduce positive charge adjacent to Lys586. There is no anomalous signal, and refinement with K+ led to an anomalously high temperature factor for this atom. Therefore, this peak was assigned as water on the basis of the presence of four binding ligands within 2.4–2.5 Å, although the planar geometry of these bonds indicates a strained environment. Given the canonical nature of this cation site in most P-type ATPases, this water could serve as a substrate mimic to assist communication between the cytosolic and transmembrane domains of KdpB. Although the cytosolic domains of KdpB adopt a unique configuration due to the unexpected presence of a phosphoserine in the A domain (Fig. 1b, Extended Data Fig. 5), the general decoupling of the A domain, the juxtaposition of the N and P domains, the lack of phosphorylation at Asp307, and the unwound configuration of bM4 are all consistent with an E1 enzymatic state.

To evaluate the accessibility of the bound water molecule in KdpB, we searched for exit pores and found that the site is blocked from both the cytosol and the periplasm. However, this analysis revealed a tunnel that connects the water site in KdpB with the K+ ion binding site in KdpA (Fig. 3). This tunnel runs for 40–45 Å parallel to the membrane surface and to the coupling helix of D3M2 and is completely encased within the membrane domain of the complex. The KdpA part of the tunnel is reminiscent of fenestrations found in other channels20,21. Unlike these fenestrations, however, the tunnel stays within the membrane and connects functional sites in the two subunits. Whereas the middle is hydrophobic or neutral, the two ends are distinctly electronegative owing to Glu370 and carbonyl oxygens in the gating loop of KdpA, and to Asp583 and side chain oxygens of Ser579 and Thr266 in KdpB. Although no density is visible within the tunnel, its diameter would accommodate water molecules and, like cavities identified in other membrane proteins, it is likely to be filled with water22.

Figure 3: Coupling KdpA and KdpB.
figure 3

a, A tunnel (>1.4 Å radius, blue surface) connects the water molecule in KdpB to the K+ ion in KdpA. The coupling helix (pink) runs adjacent to the tunnel on the cytosolic side of the membrane. b, Salt bridges attach the coupling helix to the P domain of KdpB (blue). c, Close-up from the periplasmic side shows charged residues at either end of the tunnel: Glu370 and Arg493 in KdpA and Asp583 and Lys586 in KdpB. The tunnel could facilitate charge transfer between these two sites. d, In KdpB, an unwound portion of transmembrane helix bM4 (shown in stick representation) forms a binding site similar to those of other P-type ATPases. The modelled water is coordinated by main chain carbonyls from bM4 and side chain oxygens from Thr265 and Asn624.

PowerPoint slide

Coupling between KdpA and KdpB requires two-way communication. On the one hand, the presence or absence of K+ in KdpA should initiate autophosphorylation of Asp307 in KdpB and, on the other hand, conformational change in KdpB should control gating of ion sites in KdpA. The latter step is likely to involve the kinked helix from the distal part of the D3M2 element of KdpA, which forms a strong interaction with the P domain of KdpB involving a salt bridge between Arg400 from KdpA and Asp300 and Asp302 from KdpB (Fig. 3b). In TrkH and KtrB, this same ‘coupling helix’ interacts with regulatory cytoplasmic subunits that undergo conformational changes in response to ATP binding12,23. In P-type ATPases, the E1P-to-E2P transition is associated with a marked inclination of the P domain24, which in KdpFABC would pull on this coupling helix and open the cytosolic gate. Indeed, this helix is relatively free to move, in part owing to the tunnel that is located directly above it, which ensures minimal interactions with other parts of KdpA. This idea is supported by mutation of Asp300 in KdpB, which has been shown to increase ATPase activity and decrease K+ affinity, as would be expected from an uncoupled complex25. To initiate autophosphorylation (E1-to-E1P transition), P-type ATPases rely on binding of cations to their canonical transmembrane site. The water-filled intramembrane tunnel provides a possible proton wire for introducing charge into that site in response to K+ binding to the selectivity filter of KdpA. The tunnel begins near Asp583 and Lys586 in KdpB, and passes Glu370 before ending next to Arg493 in KdpA (Fig. 3c, d). The functional importance of the tunnel is supported by mutation of Asp583, which uncouples K+ transport and ATP hydrolysis26,27, and mutation of Arg493, which disrupts KdpFABC function27. All of these residues are conserved, except for Glu370, which is a Gln or an Asn in about half of bacterial species (Extended Data Figs 6, 7, 8, 9).

Our results suggest the following model for transport by KdpFABC (Fig. 4). With KdpB in the E1 conformation, as observed in the structure, K+ ions enter the selectivity filter of KdpA from the periplasmic side of the membrane. Selectivity filters have innate binding affinity for K+ ions at the micromolar level28 and the relatively slow rate of transport by KdpFABC would allow equilibration of these sites and explain the selectivity for K+ over Na+. Permeation of ions into the S4 site would induce a proton charge transfer through the tunnel to the structural water molecule bound to the canonical ion binding site in KdpB. This mimic of cation binding to a P-type ATPase would induce a conformational change in KdpB that would lead to phosphorylation of Asp307. The consequent conformational changes associated with E1P formation in KdpB would lead to outward occlusion of the selectivity filter in KdpA, perhaps by movement of the periplasmic KdpC domain. As in SERCA1a, the transition of KdpB to E2P would induce an inclination of the P domain away from KdpA24,29, which would pull the coupling helix towards KdpB and open the gating loop to allow K+ to escape to the cytoplasm. This could be associated with a reverse charge movement from KdpB to KdpA through the tunnel due to expected changes in the conformation of bM429. A removal of the transmembrane charge in KdpB would then trigger hydrolysis of E2P, returning the system to the E1 state24,30.

Figure 4: Mechanism for KdpFABC.
figure 4

According to the Post–Albers scheme, transitions between open and closed states (black boxes) are driven by phosphorylation and dephosphorylation events in the cytosolic domains of KdpB. The cycle is initiated by K+ binding to the E1 state from the periplasm (1). The presence of K+ in the S4 site of the selectivity filter of KdpA leads to charge transfer through the tunnel to the transmembrane domain of KdpB (2). The presence of charge at the canonical site in KdpB triggers phosphorylation through a conserved P-type ATPase mechanism (3). K+ occlusion, which may involve the periplasmic domain of KdpC, leads to the occluded E1P state (4). The transition to the E2P state in P-type ATPases involves inclination of the P domain away from KdpA, which will pull the D3 coupling helix (pink) of KdpA (5). This movement opens the cytoplasmic gate, thereby allowing K+ release to the cytosol (6). The models are derived from our structure of an inhibited E1 state and SERCA1a structures of E1P (1T5T) and E2P (3B9B) states.

PowerPoint slide

Ion transport by channels and by transporters such as P-type ATPases are fundamentally different processes and KdpFABC is, so far, unique in pairing these two disparate functionalities to enable active transport. Although the structure provides the first hints as to how this is accomplished, there are many questions to be answered. This structure provides a template with which we can design experiments to address these questions for a better understanding of this system and of transport mechanisms in general.

Methods

Sample expression

The kdp operon encoding the four subunit KdpFABC complex with a single mutation (Q116R) in the KdpA subunit and an 8 × histidine tag at the C terminus of the KdpC subunit was cloned into plasmid pSD107, which was originally derived by inserting the KdpFABC operon with a C-terminal histidine tag into the pBR322 plasmid, thus providing ampicillin resistance6. This plasmid was transformed into the E. coli KdpFABC knockout strain TK2498, with genotype: F thi lacZ nagA rha trkA405 trkD1 Δ(KdpFAB)5 Δ(ompT). Both the plasmid and the E. coli strain were obtained from W. Epstein (University of Chicago). Expression was controlled by the native kdp promoter9,31, with KdpD and KdpE being present in the chromosome. Expression was induced by a potassium-free K0-medium (46 mM Na2PO4, 23 mM NaH2PO4, 25 mM (NH4)2SO4, 0.4 mM MgSO4, 6 μM FeSO4, 1 mM sodium citrate, 0.2% glucose, 1 μg ml−1 thiamine, 50 μg ml−1 carbenicillin) supplemented with different amounts of KCl32,33. To start, cells from a glycerol stock were incubated overnight at 37 °C in 10 ml K5-medium (K0-medium supplemented with 5 mM KCl). This culture was transferred to 500 ml K1-medium (K0-medium supplemented with 1 mM KCl) and incubated at 37 °C for 8 h. The 500 ml cell culture was transferred again to 18 l K0.2-medium (K0-medium supplemented with 0.2 mM KCl) and incubated at 31 °C to induce expression of KdpFABC. Cells were harvested when the culture density reached OD600 ~1. For production of protein with selenomethionine (SeMet) substitution, we used high concentrations of isoleucine, leucine, phenylalanine, lysine, and threonine to inhibit the methionine biosynthesis pathway in E. coli34. Specifically, 100 mg/l each of lysine, phenylalanine, and threonine, 50 mg/l each of isoleucine, leucine, and valine, and 60 mg ml−1 SeMet were added to the K0.2 culture medium as the protein expression was induced.

Sample purification

The harvested cells were resuspended in 50 mM Tris pH 7.5, 1.2 M NaCl, 10 mM MgCl2, 10% glycerol, protease inhibitor tablets (Roche), 25 mg ml−1 DNase, and 1 mM DTT, and lysed using an Emulsiflex C3 high-pressure homogenizer (Avestin). After centrifugation at 10,600g for 15 min to remove unbroken cells and debris, the supernatant was centrifuged at 90,140g for 2 h to pellet cell membranes. This membrane fraction was solubilized in 50 mM Tris pH 7.5, 600 mM NaCl, 10 mM MgCl2, 10% glycerol, 1 mM DTT and 1.2% n-decyl-β-maltoside (DM) at 4 °C for at least 2 h, and then centrifuged at 90,140g for 30 min to remove insoluble components. The supernatant containing solubilized protein was loaded onto a 5 ml Ni+-charged HiTrap chelating column (GE Healthcare), which had been pre-equilibrated with buffer A (50 mM Tris pH 7.5, 600 mM NaCl, 10 mM MgCl2, 10% glycerol, 0.15% DM, 20 mM imidazole). A linear gradient of imidazole (20–500 mM based on buffer A) was used to elute the KdpFABC complex using a Biologic LP chromatography system (Bio-Rad). The eluted fractions were pooled together and concentrated using 100 kDa cut-off concentrators to ~10 mg ml−1. This sample was further purified by applying 0.5 ml onto a preparative size exclusion column Superdex 200 (GE Healthcare) equilibrated with OG-DMPC buffer composed of 25 mM Tris pH 7.5, 100 mM KCl, 10% glycerol, 1.1% n-octyl-β-d-glucoside (OG), and 0.5 mg ml−1 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC). The eluted fractions containing KdpFABC complex were combined, concentrated to 15–20 mg ml−1 and stored at −80 °C. An identical protocol was used for purification of SeMet-substituted KdpFABC complex.

ATPase assay and dephosphorylation

A coupled enzyme assay was used to measure ATPase activity35. The assay solution contained 50 mM Tris pH 7.5, 5 mM MgCl2, 0.15% DM, 2.4 mM ATP, 0.18 mM NADH, 0.5 mM phosphoenol pyruvate, 4.8 units of pyruvate kinase, and 4.8 units of lactate dehydrogenase in 10 ml. This assay couples the hydrolysis of ATP with the oxidation of NADH, which was followed in real time using the absorbance at 340 nm. For each reaction, 10 mg KdpFABC complex and various concentrations of cations were mixed with 0.5 ml assay buffer at 25 °C and specific activity (μmol Pi min−1 mg−1) was calculated as (OD340 per min) × (1/6.22) × (1/mg protein) × 0.5 ml. To study the effect of Ser162 phosphorylation, a given preparation was incubated in the presence and absence of lambda phosphatase at 30 °C for either 4 or 16 h and then assayed for ATPase activity in the presence of 50 mM KCl and for the level of phosphorylation using mass spectrometry.

Mass spectrometry

The presence of phosphoserine in the sample was confirmed using electrospray ionization with liquid chromatography and tandem mass spectrometry (ESI–LC–MS/MS). KdpB was isolated by SDS–PAGE from various different samples and digested with trypsin within the gel. After elution, the peptides were fractionated using high-performance liquid chromatography integrated with tandem mass spectrometers. The peptide containing the non-phosphorylated or phosphorylated Ser162 was identified by MS1, the sequence of which was further confirmed by MS2. The relative abundance of the non-phosphorylated and phosphorylated Ser162 of KdpB was calculated from the peak heights of the two from MS1.

Crystallization

Purified KdpFABC complex was adjusted to 8 mg ml−1 using OG-DMPC buffer and supplemented with 5 mM Mg-AMPPCP for crystallization. Mg-AMPPCP was required for crystallization, despite not being visible in the crystal structure. We speculate that the Mg-AMPPCP may trap the KdpFABC complexes that did not carry the Ser162 phosphorylation in an unfavourable conformation, thus excluding them from the crystal lattice. Native and SeMet-derived crystals were grown at 18 °C using the hanging-drop vapour diffusion method: 2 μl protein solution was mixed with 2 μl reservoir solution containing 20% (w/v) PEG3350, 0.5–1.0 M NaCl, 0.05 M sodium citrate pH 5.5. Crystals appeared within a week and grew to full size (100 × 300 × 40 μm3) after three weeks. Drops containing suitable crystals were dehydrated against a reservoir containing 36% glycerol for 1 day36 and flash-frozen in liquid nitrogen. To obtain heavy atom derivatives, crystals were soaked with osmium, platinum and tantalum salts; despite strong anomalous signal, however, these derivatives could not be used to solve the structure. Tungsten derivatives were obtained by soaking native crystals in 1 mM Na6(H2W12O40) for 1 day. The mercury derivative was obtained by soaking native crystals in 5 mM Hg(OOCCH3)2 for 2–3 days. Data used for structure solution were collected at the Advanced Photon Source beamline 23-ID-B and 23-ID-D. Additional screening and preliminary data collection was done at the National Synchrotron Light Source X4A and X4C and at Stanford Synchrotron Radiation Lightsource 14-1.

Data processing

Datasets were processed and scaled using XDS37 in space group P21, which suggested the presence of three KdpFABC complexes in the asymmetric unit (~70% solvent content). The positions of tungsten clusters were determined by single-wavelength anomalous dispersion (SAD) in SHELXC/D38 and initial phases were calculated in SHARP39. After solvent flattening with Solomon40 or DM40, the electron density map based on data from the tungsten cluster derivative at 4.0 Å resolution enabled us to identify KdpA and KdpB subunits and the expected 19 transmembrane helices, but the quality of this map was not sufficient to establish the orientation of the pseudo-four-fold symmetric KdpA or to place the soluble domains of KdpC unambiguously. Independent phase information to higher resolution (3.3 Å) was then obtained from the mercury derivatives. Owing to excessive non-isomorphism, no native or derivative datasets could be combined for phase calculation. However, mercury sites could be identified by anomalous difference Fourier maps using the phases from the tungsten cluster derivative. Mercury derivative SAD phases were calculated in SHARP, and were refined and extended using DM to 3.3 Å resolution, exploiting histogram matching, solvent flattening and three-fold non-crystallographic symmetry averaging. The resulting electron density map was of high quality, allowing for a continuous trace of the main chain. The model was built using COOT41 with KtrB12 and Cu+-ATPase42 as guides for building initial models of KdpA and KdpB, respectively. Objective aides in building this model came from 44 selenomethiones and 7 mercury atoms bound to cysteines in each KdpFABC complex, which were identified in anomalous difference Fourier maps from the SeMet and mercury derivatives, respectively (Extended Data Fig. 1b). The anomalous peak from K+ was identified from the native 2.9 Å dataset. It was not possible to use Rb+ as a K+ congener as the mutation Q116K not only lowers K+ affinity but also inhibits Rb+ coupling43 (Fig. 1a). Iterative model building in COOT and refinement using phenix.refine44 gradually improved the model and the fit to the experimental map. At later stages, the model was of sufficient quality to be used for molecular replacement into the native dataset (2.9 Å) using the program PHASER45; afterwards, model building was guided by 2mFo−DFc maps using model phases. Final refinement in phenix.refine exploited three-fold non-crystallographic symmetry with a refinement strategy of individual sites, individual ADP, and group TLS (18 groups), against a maximum likelihood (ML) target with reflections in the 20–2.9 Å range. The final model yielded an Rwork of 24.3% and an Rfree of 27.5% (Extended Data Table 1). MolProbity46 evaluation of the Ramachandran plot gave 95.9% in favoured regions and 0.1% outliers. The KdpA–KdpB tunnel was identified with CAVER47 using default settings and a probe radius of 1.4 Å, which is equivalent to the radius of water. All structural figures were prepared using PyMOL (The PyMOL Molecular Graphics System, Version 1.5.0.4 (Schrödinger LLC, 2012)).

Sequence alignment

Sequence conservation of KdpA and KdpB were initially evaluated by applying Clustal Omega alignment48 to a list of genes obtained from the Divblast server49. The Divblast server ensures that the list represents the diversity across all KdpFABC homologues, which are otherwise dominated by a large number of E. coli strains. Alignment of KdpA and KdpB with distantly related members of the SKT and P-Type ATPase families was done with the Promals3d server50. The ability of Promals3d to consider structural features as well as sequence produced a reasonably accurate alignment that nevertheless required minor manual adjustments based on comparison of tertiary structures of the respective atomic models.

Data availability

Atomic coordinates and structure factors have been deposited in the Protein Data Bank (PDB) with accession code 5MRW. All other data are available from the corresponding authors upon reasonable request.