Driven by transmembrane electrochemical ion gradients, F-type ATP synthases are the primary source of the universal energy currency, adenosine triphosphate (ATP), throughout all domains of life. The ATP synthase found in the thylakoid membranes of photosynthetic organisms has some unique features not present in other bacterial or mitochondrial systems. Among these is a larger-than-average transmembrane rotor ring and a redox-regulated switch capable of inhibiting ATP hydrolysis activity in the dark by uniquely adapted rotor subunit modifications. Here, we review recent insights into the structure and mechanism of ATP synthases specifically involved in photosynthesis and explore the cellular physiological consequences of these adaptations at short and long time scales.

Cellular life depends on the energy that is ultimately powered by sun light. Its direct users are plants, algae and some clades of bacteria, e.g. cyanobacteria and purple bacteria. These photoautotrophic organisms convert solar energy into chemical forms of energy by photosynthesis, a process in which 7 × 1016 g of carbon is fixed by autotrophs globally each year [1]. The light-dependent reactions of photosynthesis involve a series of large, macromolecular complexes, which are capable of harvesting light energy to split water molecules releasing electrons for the reduction in NADP+ into NADPH and releasing protons (H+) into the thylakoid lumen generating the proton-motive force, pmf, required for ATP synthesis. The resultant pools of NADPH and ATP are essential substrates for the Calvin–Benson–Bassham cycle (CBB) where carbon fixation occurs [2,3]. The four large protein complexes involved in the production of NADPH and ATP are photosystem II, cytochrome b6f, photosystem I and ATP synthase. The first three complexes form the electron transport chain for NADP+ reduction and the latter complex synthesising ATP, the cells’ universal energy currency [4–6]. This review focuses on ATP synthase from photosynthetic organisms, which possesses some unique features not observed in other forms of life.

The pmf is the direct energy source of the rotary F1Fo-ATP synthase; a large, membrane-embedded protein complex that converts the substrates adenosine diphosphate (ADP) and inorganic phosphate (Pi) into ATP by a unique rotational mechanism. The process of ATP synthesis by ATP synthases is entirely reversible. Under particular cellular conditions, the ATP synthase can operate in reverse, converting energy from ATP hydrolysis and pumping ions across the membrane to generate a pmf [7]. The pmf itself is composed of two components, a transmembrane electrical potential (Δψ) and a transmembrane chemical proton gradient (ΔpH). The ATP synthase from bacterial, chloroplast and mitochondrial sources plays a key role in cellular bioenergetics and has been studied by biochemical and structural methods already for many decades [8–11]. More recently, electron cryo-microscopy (cryo-EM) has revealed the structure of fully intact F-type ATP synthases from chloroplasts at high-resolution, resolving the intermediary conformational states and other key mechanistic and regulatory aspects (Figure 1) [12,13].

Structure of the spinach chloroplast F1Fo-ATP synthase.

Figure 1.
Structure of the spinach chloroplast F1Fo-ATP synthase.

(A) The F1Fo-ATP synthase from spinach chloroplast is found embedded in the thylakoid membranes and couples the proton-motive force, pmf, with the synthesis of ATP. The water-soluble, catalytic F1 subcomplex consists of subunits α3β3γδε. The membrane-embedded Fo subcomplex consists of subunits ab2c14. During ATP synthesis, protons enter via an aqueous half-channel on the a-subunit exposed to the thylakoid lumen and bind successively to the binding sites on the c14-ring. As protons bind, the c14-ring rotates almost a full turn before the protons exit through a second half-channel formed by an aqueous wedge between the c-ring and the a-subunit into the stroma. The torque generated by the rotating c14-ring is transferred to the catalytic F1 head group via the central stalk (γε). In the catalytic β-subunits ADP and Pi condenses to ATP. Each of the three β-subunits is in different conformational state at any given time point, either empty (βE), bound to ADP (βDP) or bound to ATP (βTP) [15,76]. Arrows indicate the direction of proton flow through Fo during ATP synthesis. (B) Open view with βDP and one adjacent α-subunit removed; bound ADP and ATP nucleotides are indicated (sphere models). (C) Bottom view, from lumen side. (D) Top view from stroma, including cutaway showing a-subunit (cartoon model) in association with c14-ring at the level of the ion binding sites of the c-ring (see Figure 2). (E) Illustrative representation of ATP synthases embedded in a thylakoid membrane. The structure models are from [12], in surface representation.

Figure 1.
Structure of the spinach chloroplast F1Fo-ATP synthase.

(A) The F1Fo-ATP synthase from spinach chloroplast is found embedded in the thylakoid membranes and couples the proton-motive force, pmf, with the synthesis of ATP. The water-soluble, catalytic F1 subcomplex consists of subunits α3β3γδε. The membrane-embedded Fo subcomplex consists of subunits ab2c14. During ATP synthesis, protons enter via an aqueous half-channel on the a-subunit exposed to the thylakoid lumen and bind successively to the binding sites on the c14-ring. As protons bind, the c14-ring rotates almost a full turn before the protons exit through a second half-channel formed by an aqueous wedge between the c-ring and the a-subunit into the stroma. The torque generated by the rotating c14-ring is transferred to the catalytic F1 head group via the central stalk (γε). In the catalytic β-subunits ADP and Pi condenses to ATP. Each of the three β-subunits is in different conformational state at any given time point, either empty (βE), bound to ADP (βDP) or bound to ATP (βTP) [15,76]. Arrows indicate the direction of proton flow through Fo during ATP synthesis. (B) Open view with βDP and one adjacent α-subunit removed; bound ADP and ATP nucleotides are indicated (sphere models). (C) Bottom view, from lumen side. (D) Top view from stroma, including cutaway showing a-subunit (cartoon model) in association with c14-ring at the level of the ion binding sites of the c-ring (see Figure 2). (E) Illustrative representation of ATP synthases embedded in a thylakoid membrane. The structure models are from [12], in surface representation.

Close modal

The F-type ATP synthase found in bacteria and chloroplasts is ∼550 kDa in size with a protein subunit stoichiometry of α3β3γδεab2c8–17 (Figure 1). The complex can be divided into two biochemically and functionally distinct subcomplexes: The water-soluble, catalytic head subcomplex, F1, composed of α3β3γδε subunits and the membrane-embedded rotor–stator subcomplex, Fo, composed of ab2c8–17 subunits. F1 and Fo are connected by a central and peripheral stalk. From an operational point of view, terms such as rotor and stator are applicable, with the rotor (c8–17γε) and the stator (α3β3δab2) subcomplexes.

In F1, the synthesis and hydrolysis of ATP occurs at a hexamer of alternating α- and β-subunits (α3β3) where one ATP molecule is able to bind at a Walker A motif (GxxxxGK[TS]), also known as the P-loop, found in each of the three β-subunits [14]. Additionally, each of the three α-subunits contains a non-catalytic nucleotide-binding site that usually bind unreactive (‘structural’) bound ATP molecules (Figure 1B) [15]. The central stalk, composed of the γ-subunit and ε-subunit, induces the necessary conformational changes in the α3β3 hexamer head for ATP synthesis by transferring torque generated in the membrane-embedded Fo subcomplex into the F1 catalytic head.

The peripheral/outer stalk, consisting of δ-, b- and b′-subunits, acts as a stabilising stator element between F1 and Fo preventing futile co-rotation. The δ-subunit has an N-terminal domain that connects to the three α-subunits in F1, and a C-terminal domain that binds to the b-subunits of the peripheral stalk. A flexible hinge connects the two domains of the δ-subunit and gives rise to a highly dynamic coupling of F1 and Fo [16]. The two b-subunits are copies of each other with high sequence identity; structurally they split at their membrane sections and sit on the a-subunit, similar to a horse saddle. In the soluble, outer stalk region they connect to form a right-handed coiled-coil motif (Figure 1B) with the purpose to counteract torque that is transmitted into F1 during ATP synthesis [17]. At the other end, the two b-subunits bind at the F1 hexamer head and form several contacts to different domains at the outer surface of α-subunit as well as contacts to the δ-subunit on top of the F1 headpiece at several locations.

F-type ATP synthases have only one outer stalk, while other rotary ATPases (A-type, V-type) have two or three [11]. One can, therefore, speculate safely that the determination of the peripheral stalk stoichiometry in F-type ATP synthases must lie in the structural attachment of the δ-subunit, which is asymmetrically bound on top of the F1 headpiece, which once bound, irreversibly determines the position of two b-subunits composing the peripheral stalk (Figure 1D).

The Fo subcomplex accommodates the membranous ion motor of the ATP synthase. The motor comprises the rotor ring, also known as the c-ring due to its oligomeric assembly of a species-specific number of c-subunit copies [18], and the stator a-subunit that lies outside, adjacent to the c-ring and forms two water accessible, separated aqueous half channels (Figure 1) [19]. The two half channels are laterally adjacent but spatially and electrostatically separated by a highly conserved and positively charged arginine side chain, and each channel provides an ion access pathway to and from the c-ring ion binding sites to opposite sides of the membrane [20–22].

The c-rings are structurally highly conserved and resemble hourglass-shaped cylinders with a central pore (Figure 2). The ion binding sites on the c-ring lie at their outside surface, each in between two adjacent c-subunits; all binding sites are accessible from the outside. Ion coordination is mediated by a particular set of conserved amino acid residues, which also determine ion specificity either for protons or Na+ [23–25]. Common to all binding sites in all rotary ATPases is a highly conserved carboxylate group, which plays the key role in H+ (and sometimes Na+) coordination. While in the ion bound state the site remains in a locked conformational state [18,25], it has been shown that the binding and unbinding events themselves are dependent on hydration events at the ion binding site [26]. These events can also be viewed from the perspective of shifts in pKa values of the binding site carboxylate groups, with a pKa of a hydrated carboxylate being ∼4.5 and a much-increased pKa (7.5 or higher, [24,27] for the same carboxylate in a hydrophobic environment, e.g. within a membrane or a detergent micelle. Hence, while the ion remains locked at its c-ring ion binding site during its travel along the hydrophobic membrane interface, upon reaching the hydrophilic amino acid residues of the exit channel, the site encounters a new chemical environment that lowers the pKa and leads the carboxylate to open its conformation and ultimately release the ion into the channel towards the opposite side of the membrane [26,28]. Recent structural studies of ATPases have elucidated a range of these ion translocation pathways [12,16,29–32]. While the overall mechanism of ion access and release involving the a-subunit and the ion binding mode on the c-ring is highly conserved in all rotary ATPases, including the V-type and A-type rotary ATPases/synthases, the details of their individual channels and the exact ion paths along conserved hydrophilic or charged amino acid residues varies between different species [10].

Examples of F-type ATP synthase cn-ring stoichiometries from c8 to c17.

Figure 2.
Examples of F-type ATP synthase cn-ring stoichiometries from c8 to c17.

The c-rings are shown in cartoon representation. Upper row: View from the cytoplasm (matrix/stroma). Lower row: view from the membrane plane in the same orientation as in Figure 1A. The membrane borders are indicated by grey lines. From left to right: c8-ring from B. taurus mitochondria [35], c11-ring from M. phlei [50], c11-ring from Ilyobacter tartaricus [18], c13-ring from Bacillus pseudofirmus OF4 [43], c15-ring from Spirulina platensis [25], c17-ring from Burkholderia pseudomallei [36]. The location of the conserved carboxylate amino acid residue required for ion binding is coloured red. All the examples here are proton-binding cn-rings, except for the c11-ring from I. tartaricus which is Na+-binding. Plant chloroplast ATP synthase harbours a c14-ring (Figure 1D), while cyanobacterial ATP synthases have c13c15 rings [49]. The resulting overall range of ion-to-ATP ratios in fully coupled ATP synthases is 2.7–5.7. For a functional description, see text.

Figure 2.
Examples of F-type ATP synthase cn-ring stoichiometries from c8 to c17.

The c-rings are shown in cartoon representation. Upper row: View from the cytoplasm (matrix/stroma). Lower row: view from the membrane plane in the same orientation as in Figure 1A. The membrane borders are indicated by grey lines. From left to right: c8-ring from B. taurus mitochondria [35], c11-ring from M. phlei [50], c11-ring from Ilyobacter tartaricus [18], c13-ring from Bacillus pseudofirmus OF4 [43], c15-ring from Spirulina platensis [25], c17-ring from Burkholderia pseudomallei [36]. The location of the conserved carboxylate amino acid residue required for ion binding is coloured red. All the examples here are proton-binding cn-rings, except for the c11-ring from I. tartaricus which is Na+-binding. Plant chloroplast ATP synthase harbours a c14-ring (Figure 1D), while cyanobacterial ATP synthases have c13c15 rings [49]. The resulting overall range of ion-to-ATP ratios in fully coupled ATP synthases is 2.7–5.7. For a functional description, see text.

Close modal

The c-ring in Fo is composed of a variable number (n) of monomeric c-subunits, with each monomer forming an ∼8–9 kDa hairpin of two transmembrane α-helices, connected by a loop that non-covalently connects to the F1 subcomplex via γ- and ε-subunits [33,34]. The known range covers most integers between n = 8 (Bos taurus) [35] and n = 17 (Burkholderia pseudomallei) [36] (exception: n = 16). The available evidence suggests that the stoichiometry of cn-rings is determined solely by the primary structure of the c-subunit and thus remains constant within a given species but varies across different species [10,18,25,37–39]. This notion is not only supported by the formation of functional chimeric ATP synthases, where resulting chimeric ATP synthases have a cn-ring stoichiometry matching the source organism [40–43], but it is also strongly suggested by heterologous expression of a variety of c-rings in Escherichia coli cells, all showing native (i.e. non E. coli) c-ring stoichiometries [44–46].

In tightly coupled F1 and Fo complexes a key bioenergetic parameter known as the ion-to-ATP ratio can be derived, which equals the number of c-subunits in the cn-ring (n) divided by the number of ATP molecules synthesised per full turnover, three. The ratio, therefore, determines the bioenergetic cost of making a single ATP molecule for a given organism [35] and it also contributes strongly to the thermodynamic efficiency of membrane systems [47]. Remarkably, when experimentally varying the c-subunit stoichiometry of a given ATP synthase by site-directed mutagenesis of the c-subunit and thereby experimentally inducing variation of the ion-to-ATP ratio, it effects both on the enzymatic level of ATP synthase kinetics [42] as well as on the growth level of the given bacterium [43]. This fact holds high hopes for similar future engineering approaches of ATP synthases to be adapted to externally imposed bioenergetic challenges, such as low light conditions in cyanobacteria or plants.

In the plant ATP synthase, the c-ring stoichiometry is c14 [48], while in cyanobacteria the stoichiometry varies between 13 and 15 [49]. Compared with most bacterial or mitochondrial ATP synthases this is a rather large stoichiometry, leading to an ion-to-ATP ratio of 4.7 in plants and 4.3–5.0 in cyanobacteria. With the total range of known c-rings known (c8c17, Figure 2) the variation of this ratio becomes impressively wide, 2.7–5.7 ions per ATP. The photosynthetic active organisms take their place in the upper regions of this range.

The differently primed functions as well as the cellular physiological environments of ATP synthase in various organisms may explain the large variation in cn-ring stoichiometries and hence also help understand the variations found in the ion-to-ATP ratios in plant and cyanobacterial ATP synthases. For an ATP synthase operating under sufficiently high pmf (or smf), smaller cn-ring stoichiometries are advantageous as they ‘consume’ fewer, albeit higher energised, protons per ATP synthesised. The hypothesis is supported by the smallest known stoichiometries, which belong to F-type ATP synthases found in animal (and plant) mitochondria: Mammalian mitochondrial ATP synthase operates with a c8-ring [35] and is known to operate almost exclusively in ATP synthesis direction. Some bacteria, such as Mycobacterium phlei, E. coli or some Bacillus species have cn-ring stoichiometries of n = 9, 10, 11, 12 and 13 [39,50-53]. Some of these bacteria are living in physiologically challenging environments, e.g. high temperature and high pH and hence specifically adapted in stoichiometry and ion binding chemistry [39,50,54], while others are challenged by aerobic–anaerobic environmental changes [55-57]. Hence in these latter cases, occasionally acting as ion pumps (ATPases) to maintain the pmf is more efficient with larger cn-ring stoichiometries as more ions can be translocated per ATP hydrolysed. In its extreme with the largest c-ring found so far (c17-ring, B. pseudomallei) the ATP synthase is suggested to exclusively consume ATP to pump protons out, allowing the pathogen to survive in acidic environments of macrophages during the early stages of an infection [36].

In photosynthetic organisms with their larger cn-ring stoichiometries, this disparity in stoichiometries and their relatively high ion-to-ATP ratios reflects the requirement for ATP synthesis to occur even under conditions where only a low total pmf is available. In photosynthetic organisms or organelles (Table 1) light fluctuations can lead to rapid changes of the pmf. Under high-light conditions, photosynthetically active organisms can establish a remarkably high ΔpH across the thylakoid membrane, up to a pH gradient of 2.5 [58,59]. This corresponds to −180 mV of total pmf, mainly provided by the ΔpH component of −150 mV. While the electric component (ΔΨ) in this proton gradient is much lower (−30 mV), the high concentration of protons reflected by the high ΔpH component in the lumen of thylakoid membranes compensates for the overall phosphorylation potential (ΔG0′) required for the synthesis of ATP between 42 kJ/mol (low light) and 60 kJ/mol (high light) that has been determined under various experimental conditions [60]. The c-rings in the ATP synthases from these organisms are generally enlarged in their stoichiometry, transporting more protons per ATP (higher ion-to-ATP ratio). In contrast, the pmf in animal mitochondria is mostly stored as Δψ (∼150 mV) and a steady supply of protons favours a small cn-ring stoichiometry as this requires fewer protons per ATP molecule synthesised (Table 1). In the case of bovine mitochondrial F-type ATP synthase, which has a c8-ring, the ion-to-ATP ratio is 8/3 = 2.7 [35]. A similarly high contribution to pmf by Δψ is not favourable in photosynthetic systems as high membrane potentials lead to more recombination events and results in photodamage [61]. The differences in cn-ring stoichiometries are an example of how ATP synthases have adapted to different physiological and environmental conditions over longer, evolutionary time scales. In photosynthetic organisms, the unique bioenergetic demands and environmental pressures have favoured larger cn-ring stoichiometries, resulting in the higher ion-to-ATP ratios.

Table 1.
Comparison of overall pmf, the pmf components (ΔΨ and ΔpH) and the c-ring stoichiometry found in organisms from different kingdoms of life. The photosynthetic active organisms are shown in bold.
Kingdom (Phylum)Species (organelle or adaptation to pH)Overall pmf (mV)ΔΨ (mV)ΔpH (mV)c-ring stoichiometry
Plantae Spinacia oleraceae(chloroplast) −180 mV [58−30 mV [59−150 mV 14 [48
Bacteria (Cyanobacteria) Synechococcus PCC 7942*(neutrophile) −190 mV [77−90 mV [77−100 mV [7713-15 [49
Animalia Bos taurus (mitochondria) −210 mV [78−150 mV −60 mV 8 [35
Bacteria (Proteobacteria) E. coli (neutrophile) −170 mV −135 mV −35 mV 10 [51
Bacteria (Firmicutes) Bacillus pseudofirmus OF4 (alkaliphile) −50 mV −180 mV +130 mV 13 [39
Kingdom (Phylum)Species (organelle or adaptation to pH)Overall pmf (mV)ΔΨ (mV)ΔpH (mV)c-ring stoichiometry
Plantae Spinacia oleraceae(chloroplast) −180 mV [58−30 mV [59−150 mV 14 [48
Bacteria (Cyanobacteria) Synechococcus PCC 7942*(neutrophile) −190 mV [77−90 mV [77−100 mV [7713-15 [49
Animalia Bos taurus (mitochondria) −210 mV [78−150 mV −60 mV 8 [35
Bacteria (Proteobacteria) E. coli (neutrophile) −170 mV −135 mV −35 mV 10 [51
Bacteria (Firmicutes) Bacillus pseudofirmus OF4 (alkaliphile) −50 mV −180 mV +130 mV 13 [39

*previously known as: Anacystis nidulans or Synechococcus R-2.

A second unique feature of ATP synthases from plant chloroplasts and other photoautotrophic organisms is how they adapt their ATP synthesis activity in response to light intensity through a redox-modulated thioredoxin system [62–64]. Under sufficient light intensity, linear electron flow reduces thioredoxin which in turn reduces a redox switch found in the γ-subunit of ATP synthase. In this reduced state, ATP synthase remains active —– this can be artificially induced with the addition of the reducing agent dithiothreitol (DTT) [65]. On the other hand, with insufficient light intensity, the redox switch in ATP synthase is oxidised and enzyme activity is inhibited. This mechanism helps to down-regulate ATP synthase operation in the reverse direction (ATPase) and avoids wasteful ATP hydrolysis in the dark [57,63,64,66,67]. In plants, ATP is readily available for anabolic reactions of the CBB cycle.

To understand this regulatory feature from a structural point of view, the plant chloroplast F1Fo-ATP synthase has been recently solved by cryo-EM at high-resolution in both an oxidised (inactive) state [12] and a reduced (active) state [13]. The latter was obtained by reducing the sample with DTT and stabilising this state with the uncompetitive inhibitor, tentoxin. The structures reveal that the rotor γ-subunit is the key structural element in the redox regulation in this enzyme (Figure 3).

Chloroplast ATP synthase redox regulation by the γ-subunit.

Figure 3.
Chloroplast ATP synthase redox regulation by the γ-subunit.

(A) Overlay of the reduced and oxidised conformations in cyan and blue, respectively. (B) The 40-residue insert forming an L-shape with two β-hairpins is shown in yellow and the interacting β-subunit is in green. (C) In the dark, the γCys240 and γCys246, on the shorter β-hairpin, are oxidised (green box) and form a disulfide bond which causes conformational changes in the longer β-hairpin of γ-subunit that lead to clashes with the β-DELSEED motif when rotating in the ATP hydrolysis direction (cyan square). While in the light, γCys240 and γCys246 are reduced (red square) and the γ-subunit does not clash with the β-DELSEED motif (blue square). Figure 3 is based on knowledge of both oxidised and reduced structures of the plant chloroplast ATP synthase [12,13].

Figure 3.
Chloroplast ATP synthase redox regulation by the γ-subunit.

(A) Overlay of the reduced and oxidised conformations in cyan and blue, respectively. (B) The 40-residue insert forming an L-shape with two β-hairpins is shown in yellow and the interacting β-subunit is in green. (C) In the dark, the γCys240 and γCys246, on the shorter β-hairpin, are oxidised (green box) and form a disulfide bond which causes conformational changes in the longer β-hairpin of γ-subunit that lead to clashes with the β-DELSEED motif when rotating in the ATP hydrolysis direction (cyan square). While in the light, γCys240 and γCys246 are reduced (red square) and the γ-subunit does not clash with the β-DELSEED motif (blue square). Figure 3 is based on knowledge of both oxidised and reduced structures of the plant chloroplast ATP synthase [12,13].

Close modal

The γ-subunit across all organisms typically consists of an antiparallel coiled-coil of the N- and C-terminal α-helices and a globular Rossmann-fold domain [68]. Relative to ATP synthases from non-photosynthetic bacteria and mitochondria, the γ-subunit from chloroplasts and cyanobacteria contains a unique, 40-residue insertion that forms an L-shape with two β-hairpins. The longer C-terminal β-hairpin extends towards the central cavity of the F1 hexamer where it can interact with a highly conserved DELSEED motif found near the bottom of the β-subunit (Figure 3C). Exclusively in plants, the shorter, N-terminal γ-subunit β-hairpin contains a motif (CDxNGxC) with cysteine residues on each of the two strands, capable of forming a disulfide bond in response to the chloroplast redox potential [64,69]. In the darkness, the disulfide bond forms under the oxidising conditions leading to a conformational change in the C-terminal β-hairpin that inhibits the rotation of the rotor in ATP hydrolysis direction by causing clashes between the coiled-coil of γ-subunit and the β-DELSEED. In light, the disulfide bond is once again reduced and there are no such clashes preventing rotary catalysis.

In contrast with chloroplast ATP synthase, the insertion found in cyanobacterial γ-subunits is only ∼25 amino acid residues and missing the shorter N-terminal β-hairpin and cysteine motif. However, experimental evidence suggests that the shorter insert in cyanobacterial γ-subunit provide regulatory capabilities through the induction of ADP-dependent inhibition [70]. Additionally, the rotor ε-subunit that is attached to the γ-subunit and the c-ring (Figure 1A/B) is observed in a retracted conformation that is not consistent with the ε-inhibition seen elsewhere in non-photosynthetic bacteria [29,67,71-73], where regulation involves the C-terminal part of the ε-subunit to reversibly interact with β-DELSEED in a so-called up-state (inhibited) or down-state (active) conformation [67,73,74]. In cyanobacteria though, the ε-subunit may be responsible for inducing the required conformational change in the γ-subunit to inhibit rotation in the ATP hydrolysis direction [72,75]. Structural studies of the cyanobacterial ATP synthase will help to identify which regulatory mechanisms play the most significant roles in cyanobacteria.

Rotary F-type ATP synthases from photosynthetic active organisms such as plants (chloroplasts) and cyanobacteria show a set of unique rotor subunit adaptations for the regulation of their activities and functionality. On a short and immediate time scale, they control wasteful ATP hydrolysis during the anabolic biochemical reactions at day/night times by the γ-subunit. On the longer, evolutionary time scale, they adapted their c-ring stoichiometry to the ATP synthesis’ driving force, the pmf. The effects of these adaptations affect the operation modes of these ATP synthases and manifest in measurable cellular effects. Most conserved biochemical reactions in living organisms are achieved by enzymes with remarkable differences in their individual subunit compositions and structural features. In the case of ATP synthases from photosynthetic organisms, the adaptations to individual physiological requirements of the species are restricted to specific subunit modifications in their rotor subunits c and γ. These adaptations reflect the power of life to cope with environmental challenges at short and long time scales, but also bear interesting potential for future biotechnological advances.

  • F-type ATP synthases play a major role in the light-dependent reactions of oxygenic photosynthesis and have unique features compared with those found in mitochondria and non-photosynthetic bacteria.

  • A wide range of ATP synthase structures are available, with mitochondrial and bacterial ATP synthases being particularly interesting with regards to health and medicine. Whilst, studying photosynthetic ATP synthases may help to reveal the complex interplay of the many cellular reactions of photosynthesis and their evolutionary past.

  • Photosynthetic ATP synthases themselves may be a suitable target for engineering photosynthetic efficiency in plants or as a way to address an altered demand in ATP/NADPH, e.g. in C4 photosynthesis.

The authors declare that there are no competing interests associated with the manuscript.

This work was funded by the Wellcome Trust [WT110068/Z/15/Z] and the BBSRC Doctoral Training Program [BB/M011178/1].

Open access for this article was enabled by the participation of Imperial College London in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.

Both authors wrote the manuscript. AC made the figures and TM conceived the concept and directed the project.

We thank Maike Bublitz for reading the manuscript and providing helpful suggestions.

ADP

adenosine diphosphate

AMP

adenosine monophosphate

ATP

adenosine triphosphate

NADP+/NADPH

nicotinamide adenine dinucleotide phosphate (oxidised/reduced)

OXPHOS

oxidative phosphorylation

1
Phillips
,
R.
and
Milo
,
R.
(
2009
)
A feeling for the numbers in biology
.
Proc. Natl Acad. Sci. U.S.A.
106
,
21465
21471
2
Lin
,
M.T.
,
Occhialini
,
A.
,
Andralojc
,
P.J.
,
Parry
,
M.A.J.
and
Hanson
,
M.R.
(
2014
)
A faster Rubisco with potential to increase photosynthesis in crops
.
Nature
513
,
547
550
3
Blankenship
,
R.E.
,
Tiede
,
D.M.
,
Barber
,
J.
,
Brudvig
,
G.W.
,
Fleming
,
G.
,
Ghirardi
,
M.
et al (
2011
)
Comparing photosynthetic and photovoltaic efficiencies and recognizing the potential for improvement
.
Science
332
,
805
4
Yamamoto
,
H.
and
Shikanai
,
T.
(
2019
)
PGR5-dependent cyclic electron flow protects photosystem I under fluctuating light at donor and acceptor sides
.
Plant. Physiol.
179
,
588
600
5
Nelson
,
N.
and
Junge
,
W.
(
2015
)
Structure and energy transfer in photosystems of oxygenic photosynthesis
.
Annu. Rev. Biochem.
84
,
659
683
6
Laughlin
,
T.G.
,
Savage
,
D.F.
and
Davies
,
K.M.
(
2020
)
Recent advances on the structure and function of NDH-1: The complex I of oxygenic photosynthesis
.
Biochim. Biophys. Acta
1861
,
148254
7
Blbulyan
,
S.
,
Avagyan
,
A.
,
Poladyan
,
A.
and
Trchounian
,
A.
(
2011
)
Role of different Escherichia coli hydrogenases in H+ efflux and F1Fo-ATPase activity during glycerol fermentation at different pH values
.
Biosci Rep.
31
,
179
184
8
Walker
,
J.E.
(
2013
)
The ATP synthase: the understood, the uncertain and the unknown
.
Biochem. Soc. Trans.
41
,
1
16
9
Junge
,
W.
and
Nelson
,
N.
(
2015
)
ATP synthase
.
Annu. Rev. Biochem.
84
,
631
657
10
Meier
,
T.
,
Faraldo-Gómez
,
J.
and
Börsch
,
M.
(
2011
) ATP synthase – a paradigmatic molecular machine. In
Molecular Machines in Biology
(
Frank
,
J.
, ed.), pp.
208
238
,
Cambridge University Press
,
Cambridge, U.K.
11
Stewart
,
A.G.
,
Sobti
,
M.
,
Harvey
,
R.P.
and
Stock
,
D.
(
2013
)
Rotary ATPases: models, machine elements and technical specifications
.
Bioarchitecture
3
,
2
12
12
Hahn
,
A.
,
Vonck
,
J.
,
Mills
,
D.J.
,
Meier
,
T.
and
Kühlbrandt
,
W.
(
2018
)
Structure, mechanism, and regulation of the chloroplast ATP synthase
.
Science
360
,
eaat4318
13
Yang
,
J.H.
,
Williams
,
D.
,
Kandiah
,
E.
,
Fromme
,
P.
and
Chiu
,
P.L.
(
2020
)
Structural basis of redox modulation on chloroplast ATP synthase
.
Commun. Biol.
3
,
482
14
Walker
,
J.E.
,
Saraste
,
M.
,
Runswick
,
M.J.
and
Gay
,
N.J.
(
1982
)
Distantly related sequences in the (- and (-subunits of ATP synthase, myosin, kinases and other ATP-requiring enzymes and a common nucleotide binding fold
.
EMBO J.
1
,
945
951
15
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
16
Murphy
,
B.J.
,
Klusch
,
N.
,
Langer
,
J.
,
Mills
,
D.J.
,
Yildiz
,
Ö.
and
Kühlbrandt
,
W.
(
2019
)
Rotary substates of mitochondrial ATP synthase reveal the basis of flexible F1-Fo coupling
.
Science
364
,
eaaw9128
17
Del Rizzo
,
P.A.
,
Bi
,
Y.
and
Dunn
,
S.D.
(
2006
)
ATP synthase b subunit dimerization domain: a right-handed coiled coil with offset helices
.
J. Mol. Biol.
364
,
735
746
18
Meier
,
T.
,
Polzer
,
P.
,
Diederichs
,
K.
,
Welte
,
W.
and
Dimroth
,
P.
(
2005
)
Structure of the rotor ring of F-type Na+-ATPase from Ilyobacter tartaricus
.
Science
308
,
659
662
19
Allegretti
,
M.
,
Klusch
,
N.
,
Mills
,
D.J.
,
Vonck
,
J.
,
Kühlbrandt
,
W.
and
Davies
,
K.M.
(
2015
)
Horizontal membrane-intrinsic α-helices in the stator a-subunit of an F-type ATP synthase
.
Nature
521
,
237
240
20
Vik
,
S.B.
and
Antonio
,
B.J.
(
1994
)
A mechanism of proton translocation by F1F0 ATP synthases suggested by double mutants of the a subunit
.
J. Biol. Chem.
269
,
30364
30369
21
Junge
,
W.
,
Lill
,
H.
and
Engelbrecht
,
S.
(
1997
)
ATP synthase: an electrochemical transducer with rotatory mechanics
.
Trends Biochem. Sci.
22
,
420
423
22
Mitome
,
N.
,
Ono
,
S.
,
Sato
,
H.
,
Suzuki
,
T.
,
Sone
,
N.
and
Yoshida
,
M.
(
2010
)
Essential arginine residue of the Fo-a subunit in FoF1-ATP synthase has a role to prevent the proton shortcut without c-ring rotation in the Fo proton channel
.
Biochem. J.
430
,
171
177
23
Meier
,
T.
,
Krah
,
A.
,
Bond
,
P.J.
,
Pogoryelov
,
D.
,
Diederichs
,
K.
and
Faraldo-Gómez
,
J.D.
(
2009
)
Complete ion-coordination structure in the rotor ring of Na+-dependent F-ATP synthases
.
J. Mol. Biol.
391
,
498
507
24
Krah
,
A.
,
Pogoryelov
,
D.
,
Meier
,
T.
and
Faraldo-Gómez
,
J.D.
(
2010
)
On the structure of the proton-binding site in the Fo rotor of chloroplast ATP synthases
.
J. Mol. Biol.
395
,
20
27
25
Pogoryelov
,
D.
,
Yildiz
,
Ö.
,
Faraldo-Gómez
,
J.D.
and
Meier
,
T.
(
2009
)
High-resolution structure of the rotor ring of a proton-dependent ATP synthase
.
Nat. Struct. Mol. Biol.
16
,
1068
1073
26
Pogoryelov
,
D.
,
Krah
,
A.
,
Langer
,
J.D.
,
Yildiz
,
Ö.
,
Faraldo-Gómez
,
J.D.
and
Meier
,
T.
(
2010
)
Microscopic rotary mechanism of ion translocation in the Fo complex of ATP synthases
.
Nat. Chem. Biol.
6
,
891
899
27
Meier
,
T.
,
Matthey
,
U.
,
von Ballmoos
,
C.
,
Vonck
,
J.
,
von Nidda
,
K.
,
Kühlbrandt
,
T.
et al (
2003
)
Evidence for structural integrity in the undecameric c-rings isolated from sodium ATP synthases
.
J. Mol. Biol.
325
,
389
397
28
Symersky
,
J.
,
Osowski
,
D.
,
Walters
,
D.E.
and
Mueller
,
D.M.
(
2012
)
Oligomycin frames a common drug-binding site in the ATP synthase
.
Proc. Natl Acad. Sci. U.S.A.
109
,
13961
13965
29
Guo
,
H.
,
Suzuki
,
T.
and
Rubinstein
,
J.L.
(
2019
)
Structure of a bacterial ATP synthase
.
Elife
8
,
e43128
30
Mühleip
,
A.
,
McComas
,
S.E.
and
Amunts
,
A.
(
2019
)
Structure of a mitochondrial ATP synthase with bound native cardiolipin
.
Elife
8
,
e51179
31
Sobti
,
M.
,
Walshe
,
J.L.
,
Wu
,
D.
,
Ishmukhametov
,
R.
,
Zeng
,
Y.C.
,
Robinson
,
C.V.
et al (
2020
)
Cryo-EM structures provide insight into how E. coli F1Fo ATP synthase accommodates symmetry mismatch
.
Nat. Commun.
11
,
2615
32
Spikes
,
T.E.
,
Montgomery
,
M.G.
and
Walker
,
J.E.
(
2020
)
Structure of the dimeric ATP synthase from bovine mitochondria
.
Proc. Natl Acad. Sci. U.S.A.
117
,
23519
23526
33
Fillingame
,
R.H.
(
1992
)
H+ transport and coupling by the F0 sector of the ATP synthase: insights into the molecular mechanism of function
.
J. Bioenerg. Biomembr.
24
,
485
491
34
Girvin
,
M.E.
,
Rastogi
,
V.K.
,
Abildgaard
,
F.
,
Markley
,
J.L.
and
Fillingame
,
R.H.
(
1998
)
Solution structure of the transmembrane H+-transporting subunit c of the F1Fo ATP synthase
.
Biochemistry
37
,
8817
8824
35
Watt
,
I.N.
,
Montgomery
,
M.G.
,
Runswick
,
M.J.
,
Leslie
,
A.G.W.
and
Walker
,
J.E.
(
2010
)
Bioenergetic cost of making an adenosine triphosphate molecule in animal mitochondria
.
Proc. Natl Acad. Sci. U.S.A.
107
,
16823
16827
36
Schulz
,
S.
,
Wilkes
,
M.
,
Mills
,
D.J.
,
Kühlbrandt
,
W.
and
Meier
,
T.
(
2017
)
Molecular architecture of the N-type ATPase rotor ring from Burkholderia pseudomallei
.
EMBO Rep.
18
,
526
535
37
Müller
,
D.J.
,
Dencher
,
N.A.
,
Meier
,
T.
,
Dimroth
,
P.
,
Suda
,
K.
,
Stahlberg
,
H.
et al (
2001
)
ATP synthase: constrained stoichiometry of the transmembrane rotor
.
FEBS Lett.
504
,
219
222
38
Müller
,
P.
,
Li
,
X.P.
and
Niyogi
,
K.K.
(
2001
)
Non-photochemical quenching. A response to excess light energy
.
Plant Physiol.
125
,
1558
1566
39
Preiss
,
L.
,
Yildiz
,
O.
,
Hicks
,
D.B.
,
Krulwich
,
T.A.
and
Meier
,
T.
(
2010
)
A new type of proton coordination in an F1Fo-ATP synthase rotor ring
.
PLoS Biol.
8
,
e1000443
40
Laubinger
,
W.
,
Dimroth
,
P.
,
Deckers-Hebestreit
,
G.
and
Altendorf
,
K.
(
1990
)
A hybrid adenosinetriphosphatase composed of F1 of Escherichia coli and F0 of Propionigenium modestum is a functional sodium ion pump
.
Biochemistry
29
,
5458
5463
41
Suzuki
,
T.
,
Ozaki
,
Y.
,
Sone
,
N.
,
Feniouk
,
B. A.
and
Yoshida
,
M.
(
2007
)
The product of uncI gene in F1Fo-ATP synthase operon plays a chaperone-like role to assist c-ring assembly
.
Proc. Natl Acad. Sci. U.S.A.
104
,
20776
20781
42
Pogoryelov
,
D.
,
Klyszejko
,
A.L.
,
Krasnoselska
,
G.O.
,
Heller
,
E.-M.
,
Leone
,
V.
,
Langer
,
J.D.
et al (
2012
)
Engineering rotor ring stoichiometries in the ATP synthase
.
Proc. Natl Acad. Sci. U.S.A.
109
,
E1599
E1608
43
Preiss
,
L.
,
Klyszejko
,
A.L.
,
Hicks
,
D.B.
,
Liu
,
J.
,
Fackelmayer
,
O.J.
,
Yildiz
,
Ö.
et al (
2013
)
The c-ring stoichiometry of ATP synthase is adapted to cell physiological requirements of alkaliphilic Bacillus pseudofirmus OF4
.
Proc. Natl Acad. Sci. U.S.A.
110
,
7874
7879
44
Meier
,
T.
,
Yu
,
J.
,
Raschle
,
T.
,
Henzen
,
F.
,
Dimroth
,
P.
and
Müller
,
D.J.
(
2005
)
Structural evidence for a constant c11 ring stoichiometry in the sodium F-ATP synthase
.
FEBS J.
272
,
5474
5483
45
Meier
,
T.
and
Dimroth
,
P.
(
2002
)
Intersubunit bridging by Na+ ions as a rationale for the unusual stability of the c-rings of Na+-translocating F1F0 ATP synthases
.
EMBO Rep.
3
,
1094
1098
46
Meier
,
T.
,
Morgner
,
N.
,
Matthies
,
D.
,
Pogoryelov
,
D.
,
Keis
,
S.
,
Cook
,
G.M.
et al (
2007
)
A tridecameric c ring of the adenosine triphosphate (ATP) synthase from the thermoalkaliphilic Bacillus sp. strain TA2.A1 facilitates ATP synthesis at low electrochemical proton potential
.
Mol. Microbiol.
65
,
1181
1192
47
Silverstein
,
T.P.
(
2014
)
An exploration of how the thermodynamic efficiency of bioenergetic membrane systems varies with c-subunit stoichiometry of F1-Fo ATP synthases
.
J. Bioenerg. Biomembr.
46
,
229
241
48
Seelert
,
H.
,
Poetsch
,
A.
,
Dencher
,
N.A.
,
Engel
,
A.
,
Stahlberg
,
H.
and
Müller
,
D.J.
(
2000
)
Structural biology. proton-powered turbine of a plant motor
.
Nature
405
,
418
419
49
Pogoryelov
,
D.
,
Reichen
,
C.
,
Klyszejko
,
A.L.
,
Brunisholz
,
R.
,
Müller
,
D.J.
,
Dimroth
,
P.
et al (
2007
)
The oligomeric state of c rings from cyanobacterial F-ATP synthases varies from 13 to 15
.
J. Bacteriol.
189
,
5895
5902
50
Preiss
,
L.
,
Langer
,
J.D.
,
Yildiz
,
Ö.
,
Eckhardt-Strelau
,
L.
,
Guillemont
,
J.E.G.
Koul
,
A.
et al (
2015
)
Structure of the mycobacterial ATP synthase Fo rotor ring in complex with the anti-TB drug bedaquiline
.
Sci. Adv.
1
,
e1500106
51
Jiang
,
W.
,
Hermolin
,
J.
and
Fillingame
,
R.H.
(
2001
)
The preferred stoichiometry of c subunits in the rotary motor sector of Escherichia coli ATP synthase is 10
.
Proc. Natl Acad. Sci. U.S.A.
98
,
4966
4971
52
Toei
,
M.
,
Gerle
,
C.
,
Nakano
,
M.
,
Tani
,
K.
,
Gyobu
,
N.
,
Tamakoshi
,
M.
et al (
2007
)
Dodecamer rotor ring defines H+/ATP ratio for ATP synthesis of prokaryotic V-ATPase from Thermus thermophilus
.
Proc. Natl Acad. Sci. U.S.A.
104
,
20256
20261
53
Stahlberg
,
H.
,
Müller
,
D.J.
,
Suda
,
K.
,
Fotiadis
,
D.
,
Engel
,
A.
,
Meier
,
T.
et al (
2001
)
Bacterial Na+-ATP synthase has an undecameric rotor
.
EMBO Rep.
2
,
229
233
54
Matthies
,
D.
,
Preiss
,
L.
,
Klyszejko
,
A.L.
,
Muller
,
D.J.
,
Cook
,
G.M.
,
Vonck
,
J.
et al (
2009
)
The c13 ring from a thermoalkaliphilic ATP synthase reveals an extended diameter due to a special structural region
.
J. Mol. Biol.
388
,
611
618
55
Bagramyan
,
K.
,
Mnatsakanyan
,
N.
and
Trchounian
,
A.
(
2003
)
Formate increases the F0F1-ATPase activity in Escherichia coli growing on glucose under anaerobic conditions at slightly alkaline pH
.
Biochem. Biophys. Res. Commun.
306
,
361
365
56
Kato-Yamada
,
Y.
,
Bald
,
D.
,
Koike
,
M.
,
Motohashi
,
K.
,
Hisabori
,
T.
and
Yoshida
,
M.
(
1999
)
Epsilon subunit, an endogenous inhibitor of bacterial F1-ATPase, also inhibits F0F1-ATPase
.
J. Biol. Chem.
274
,
33991
33994
57
Feniouk
,
B.A.
, and
Yoshida
,
M.
(
2008
) Regulatory mechanisms of proton-translocating F0F1-ATPSynthase. In
Bioenergetics: Energy Conservation and Conversion
(
Schäfer
,
G.
and
Penefsky
,
H. S.
, eds.), pp.
279
308
,
Springer Berlin Heidelberg
,
Berlin, Heidelberg
58
Junesch
,
U.
and
Gräber
,
P.
(
1991
)
The rate of ATP-synthesis as a function of ΔpH and Δψ catalyzed by the active, reduced H+-ATPase from chloroplasts
.
FEBS Lett.
294
,
275
278
59
Cruz
,
J.A.
,
Sacksteder
,
C.A.
,
Kanazawa
,
A.
and
Kramer
,
D.M.
(
2001
)
Contribution of electric field Δψ to steady-state transthylakoid proton motive force (pmf) in vitro and in vivo. Control of pmf parsing into Δψ and ΔpH by ionic strength
.
Biochemistry
40
,
1226
1237
60
Giersch
,
C.
,
Heber
,
U.
,
Kobayashi
,
Y.
,
Inoue
,
Y.
,
Shibata
,
K.
and
Heldt
,
H.W.
(
1980
)
Energy charge, phosphorylation potential and proton motive force in chloroplasts
.
Biochim. Biophys. Acta
590
,
59
73
61
Davis
,
G.A.
and
Kramer
,
D.M.
(
2020
)
Optimization of ATP synthase c-rings for oxygenic photosynthesis
.
Front. Plant Sci.
10
,
1778
62
Wolosiuk
,
R.A.
and
Buchanan
,
B.B.
(
1977
)
Thioredoxin and glutathione regulate photosynthesis in chloroplasts
.
Nature
266
,
565
567
63
Hisabori
,
T.
,
Sunamura
,
E.I.
,
Kim
,
Y.
and
Konno
,
H.
(
2013
)
The chloroplast ATP synthase features the characteristic redox regulation machinery
.
Antioxid. Redox Signal.
19
,
1846
1854
64
Sekiguchi
,
T.
,
Yoshida
,
K.
,
Okegawa
,
Y.
,
Motohashi
,
K.
,
Wakabayashi
,
K.I.
and
Hisabori
,
T.
(
2020
)
Chloroplast ATP synthase is reduced by both f-type and m-type thioredoxins
.
Biochim. Biophys. Acta
1861
,
148261
65
Ketcham
,
S.R.
,
Davenport
,
J.W.
,
Warncke
,
K.
and
McCarty
,
R.E.
(
1984
)
Role of the gamma subunit of chloroplast coupling factor 1 in the light-dependent activation of photophosphorylation and ATPase activity by dithiothreitol
.
J. Biol. Chem.
259
,
7286
7293
66
Junesch
,
U.
and
Gräber
,
P.
(
1987
)
Influence of the redox state and the activation of the chloroplast ATP synthase on proton-transport-coupled ATP synthesis/hydrolysis
.
Biochim. Biophys. Acta
893
,
275
288
67
Sobti
,
M.
,
Ishmukhametov
,
R.
,
Bouwer
,
J.C.
,
Ayer
,
A.
,
Suarna
,
C.
,
Smith
,
N.J.
et al (
2019
)
Cryo-EM reveals distinct conformations of E. coli ATP synthase on exposure to ATP
.
Elife
8
,
e43864
68
Gibbons
,
C.
,
Montgomery
,
M.G.
,
Leslie
,
A.G.
and
Walker
,
J.E.
(
2000
)
The structure of the central stalk in bovine F1-ATPase at 2.4 Å resolution
.
Nat. Struct. Biol.
7
,
1055
1061
69
Hisabori
,
T.
,
Ueoka-Nakanishi
,
H.
,
Konno
,
H.
and
Koyama
,
F.
(
2003
)
Molecular evolution of the modulator of chloroplast ATP synthase: origin of the conformational change dependent regulation
.
FEBS Lett.
545
,
71
75
70
Sunamura
,
E.-I.
,
Konno
,
H.
,
Imashimizu-Kobayashi
,
M.
,
Sugano
,
Y.
and
Hisabori
,
T.
(
2010
)
Physiological impact of intrinsic ADP inhibition of cyanobacterial FoF1 conferred by the inherent sequence inserted into the γ subunit
.
Plant Cell Physiol.
51
,
855
865
71
Masaike
,
T.
,
Suzuki
,
T.
,
Tsunoda
,
S.P.
,
Konno
,
H.
and
Yoshida
,
M.
(
2006
)
Probing conformations of the beta subunit of F0F1-ATP synthase in catalysis
.
Biochem. Biophys. Res. Commun.
342
,
800
807
72
Murakami
,
S.
,
Kondo
,
K.
,
Katayama
,
S.
,
Hara
,
S.
,
Sunamura
,
E.I.
,
Yamashita
,
E.
et al (
2018
)
Structure of the γ–ε complex of cyanobacterial F1-ATPase reveals a suppression mechanism of the γ subunit on ATP hydrolysis in phototrophs
.
Biochem. J.
475
,
2925
2939
73
Tsunoda
,
S.P.
,
Rodgers
,
A.J.
,
Aggeler
,
R.
,
Wilce
,
M.C.
,
Yoshida
,
M.
and
Capaldi
,
R.A.
(
2001
)
Large conformational changes of the epsilon subunit in the bacterial F1F0 ATP synthase provide a ratchet action to regulate this rotary motor enzyme
.
Proc. Natl Acad. Sci. U.S.A.
98
,
6560
6564
74
Nakanishi-Matsui
,
M.
,
Sekiya
,
M.
,
Yano
,
S.
and
Futai
,
M.
(
2014
)
Inhibition of F1-ATPase rotational catalysis by the carboxyl-terminal domain of the ɛ subunit
.
J. Biol. Chem.
289
,
30822
30831
75
Akiyama
,
K.
,
Kondo
,
K.
,
Inabe
,
K.
,
Murakami
,
S.
,
Wakabayashi
,
K.-I.
and
Hisabori
,
T.
(
2019
)
The β-hairpin region of the cyanobacterial F1-ATPase γ-subunit plays a regulatory role in the enzyme activity
.
Biochem. J.
476
,
1771
1780
76
Boyer
,
P.D.
(
1997
)
The ATP synthase—a splendid molecular machine
.
Annu. Rev. Biochem.
66
,
717
749
77
Ritchie
,
R.J
. (
1991
)
Membrane potential and pH control in the cyanobacterium Synechococcus R-2 (Anacystis nidulans) PCC 7942
.
J. Plant Physiol.
137
,
409
418
78
Mitchell
,
P
. (
2011
)
Chemiosmotic coupling in oxidative and photosynthetic phosphorylation
. 1966.
Biochim. Biophys. Acta.
1807
,
1507
1538
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY). Open access for this article was enabled by the participation of Imperial College London in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.