Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2013 Sep 23;110(41):16468–16473. doi: 10.1073/pnas.1305497110

Phosphate release coupled to rotary motion of F1-ATPase

Kei-ichi Okazaki a,b,c, Gerhard Hummer a,b,1
PMCID: PMC3799341  PMID: 24062450

Significance

F1-ATPase is the catalytic domain of FoF1-ATP synthase, the rotary molecular motor at the core of the energy transduction machinery in all of life. We use atomistic molecular dynamics simulations to study a key event in its catalytic cycle, the release of inorganic phosphate (Pi) produced by the hydrolysis of ATP. We determine the timing, kinetics, and molecular mechanism of Pi release and clarify its role in torque generation. We also obtain an atomically detailed structure of a crystallographically unresolved intermediate formed after the 40° substep. Our results help reconcile conflicting interpretations of earlier biochemical, crystallographic, and single-molecule studies; shed light on the functional requirements of efficient ATP synthesis; and establish connections to other motors such as myosin.

Abstract

F1-ATPase, the catalytic domain of ATP synthase, synthesizes most of the ATP in living organisms. Running in reverse powered by ATP hydrolysis, this hexameric ring-shaped molecular motor formed by three αβ-dimers creates torque on its central γ-subunit. This reverse operation enables detailed explorations of the mechanochemical coupling mechanisms in experiment and simulation. Here, we use molecular dynamics simulations to construct a first atomistic conformation of the intermediate state following the 40° substep of rotary motion, and to study the timing and molecular mechanism of inorganic phosphate (Pi) release coupled to the rotation. In response to torque-driven rotation of the γ-subunit in the hydrolysis direction, the nucleotide-free αβE interface forming the “empty” E site loosens and singly charged Pi readily escapes to the P loop. By contrast, the interface stays closed with doubly charged Pi. The γ-rotation tightens the ATP-bound αβTP interface, as required for hydrolysis. The calculated rate for the outward release of doubly charged Pi from the αβE interface 120° after ATP hydrolysis closely matches the ∼1-ms functional timescale. Conversely, Pi release from the ADP-bound αβDP interface postulated in earlier models would occur through a kinetically infeasible inward-directed pathway. Our simulations help reconcile conflicting interpretations of single-molecule experiments and crystallographic studies by clarifying the timing of Pi exit, its pathway and kinetics, associated changes in Pi protonation, and changes of the F1-ATPase structure in the 40° substep. Important elements of the molecular mechanism of Pi release emerging from our simulations appear to be conserved in myosin despite the different functional motions.


F1-ATPase (F1), the catalytic domain of FoF1-ATP synthase, is a rotary molecular motor that reversibly interconverts ATP hydrolysis free energy and mechanical work associated with the rotation of the central stalk (1). The minimal functional F1 consists of a hexameric ring formed by three αβ-subunit dimers, with the rod-like γ-subunit located at its center (2). The rotation of the γ-subunit is tightly coupled to the reactions in the three catalytic sites located at the αβ interfaces and hosted mainly by the β-subunits. As a result, F1 is a unique reversible motor that rotates γ by converting ATP hydrolysis energy at high efficiency (35) and, conversely, synthesizes ATP from ADP and inorganic phosphate (Pi) by forced rotation of γ in the reverse direction (6, 7). The three nucleotide-binding sites are in different phases of catalysis, reflecting the asymmetric structure of the γ-subunit. Correspondingly, the αβ-subunits hosting the catalytic interfaces are in different conformational states, empty (E), ATP-bound (TP), and ADP+Pi-bound (DP), as seen in crystal structures (2). They communicate through the γ-subunit or directly within the α3β3 ring (8, 9) and cooperatively drive the rotation of the γ-subunit.

Single-molecule experiments have shown that the γ-subunit rotates in 120° steps (4). Distinct substeps of 80° and 40° (10) are driven by ATP binding plus ADP release, and ATP hydrolysis plus Pi release, respectively (1013). We thus expect two metastable conformations of F1, one before the 80° substep (binding dwell) and the other before the 40° substep (catalytic dwell) (14, 15) (Fig. 1A). Most crystal structures correspond to the catalytic dwell state (1, 1419), with the F1 conformation of the binding dwell state still elusive (16).

Fig. 1.

Fig. 1.

Mechanochemical coupling scheme and Pi release of F1-ATPase. (A) Mechanochemical coupling scheme deduced from recent single-molecule experiments (13). ATP hydrolysis reaction steps in one of the three catalytic sites (yellow circles) as a function of the γ-rotation (red arrow). T, D−P, D+P, and P stand for ATP bound, hydrolyzing ATP, ADP+Pi bound, and Pi bound, respectively. At an angle of 0°, ATP is bound. States differing by integer multiples of 120° are functionally equivalent. The timing and pathway of Pi release (purple) has been controversial. (B) Crystal structure (PDB ID code 2jdi) of F1-ATPase that represents the ∼80° catalytic dwell state, with α-, β-, γ-subunits colored in blue, orange, and red, respectively. (C) Tightly bound Pi in the E site modeled on the basis of yeast structures (SI Text). The blue and orange ribbon backbones correspond to the αE- and βE-subunit, respectively. The five residues that participate in Pi binding are shown with thick bonds. VMD was used to prepare structural figures (64).

The transition from the catalytic dwell to the binding dwell involves Pi release, but the release site and the exact timing in the full cycle remain controversial. In most kinetic models of F1 (1), but not all (20), Pi is assumed to exit first, followed by ADP release. This order appears to be consistent with kinetic (21) and structural studies (22) showing all three sites occupied by nucleotide (22); by contrast, the single-molecule experiments of Watanabe et al. (13) and structural studies of yeast F1 (23) suggest that ADP is released before Pi. With γ-rotation stalled in the catalytic dwell state by magnetic tweezers, the hydrolysis reaction was found to be reversible without excess Pi in solution (13), suggesting that Pi is released at 320°, i.e., from the E site (13, 24) (where ATP binding defines 0°; Fig. 1A).

Here, we reconcile these conflicting interpretations by determining the kinetics of Pi release from the E site and the DP site with atomistic molecular dynamics simulations. These simulations also allow us to explore the coupling between the 40° substep and Pi release, and to examine the underlying molecular mechanisms. F1 has been studied by molecular simulations at various levels of resolution, including quantum chemical calculations of ATP hydrolysis (2527), all-atom simulations of conformational changes in the β-subunit (28, 29), of ATP release (30), of fluctuations in the complex (31, 32), and of γ-rotation (33, 34), as well as coarse-grained simulations of γ-rotation (3537). As in experiment and in earlier simulations (33), we apply external torque to modulate the rates of the functional processes, including Pi release.

We first characterize the molecular motions of F1 during the 40° substep in atomically detailed simulations. By rotating the γ-subunit with the help of a newly developed flexible rotor method, we show how the γ-angle and the protonation state of Pi (i.e., HPO42− or H2PO4) affect the conformation of αβ-dimers and the stability of Pi binding in the E site. We then estimate rates of Pi release from the E site and from the DP site with the help of metadynamics simulations (38). From the simulations and by relating the calculated rates to experiment, we determine the timing, pathway, and charge state dependence of Pi release. Finally, we show that key elements of Pi release appear to be conserved in myosin, despite the different functional motions.

Results

Rotation of γ and Conformational Response of αβ-Dimers.

In all-atom explicit-solvent simulations of the F1 motor, the γ-angle was rotated in the hydrolysis direction with an angular velocity of 1°/ns by the newly developed flexible rotor method (Methods, Fig. S1, and Movie S1). The angular velocity may seem higher than in experiment. However, the millisecond timescale of F1 rotation in experiment mainly reflects the catalytic dwell waiting for hydrolysis and Pi release. The rotary motion during the 40° substep transition should thus be much faster, yet it has not been measured without load to the best of our knowledge. For structurally more complex transitions like protein and RNA folding, transition path times have been measured in the low-microsecond range, many orders of magnitude faster than the corresponding mean first-passage times of (un)folding (39, 40). With the γ-subunit allowed to twist freely in our simulations, the angle of its core lagged behind the average, while the protruded part moved ahead (Fig. 2B). This twisting in response to torque reflects both on the torsional stiffness of γ and on the friction experienced by the core part due to its tight interactions with the α3β3 ring. The enhanced rotation of the protruded part (or “foot”) seen here is consistent with large variations in its rotational angle in crystal structures (22, 41) and the low torsional stiffness of the γ-rotor (42). Importantly, despite the twisting the γ-subunit remains structurally intact, as indicated by root-mean-square deviations (RMSDs) of <3.5 Å (Fig. S2A). The bottom part of the C-terminal helix of γ resists rotation, which seems to be a main source of the twisting (Fig. S2B). The elastic energy stored in this region may contribute to the reversibility and efficiency of the F1 motor (32).

Fig. 2.

Fig. 2.

Rotation of γ-subunit in the hydrolysis direction. (A) Rotation of γ against α3β3 ring in the hydrolysis direction. The N- and C-terminal helices (γ1–33, 224–273), colored in purple, form the core that penetrates the center of the α3β3 ring. The protruded part is colored in red. (B) Rotational angle for the whole (blue), the core (purple), and the protruded parts of γ (red) as a function of time for F1 with H2PO4 in the E site.

Rotation of the γ-subunit induces conformational changes in all three αβ dimers, αβE, αβTP, and αβDP. We monitor their conformational responses by projection onto the principal component axes PC1 and PC2 obtained previously from the analysis of X-ray structures (43). PC1 reports on the hinge closing motion of the β-subunit, and PC2 on the tightening motion of the αβ interface in response to ligand binding and γ-rotation. PC1 and PC2 are in units of Ångstroms. During the γ-rotation, the αβDP interface remained close to the X-ray structures (Fig. 3A). By contrast, the interface structure of αβTP tightened in both simulations, moving toward αβDP already after ∼20° γ-rotation (Fig. 3A, green points; and difference ΔDRMSD of the RMSD with respect to the initial αβTP and αβDP structures in Fig. S2C). αβTP thus indeed moved in the direction expected for hydrolysis. This motion is concentrated mainly in αTP (Fig. 3B, blue), with rotation of γ in the hydrolysis direction pushing a protruded β-hairpin of γ (γ165–173) away from the helix-turn-helix motif of the C-terminal domain of αTP, and αTP occupying the vacated space (Fig. 3B and Movie S1). We note that this conformational change was not observed in free simulations without torque (Fig. S3B).

Fig. 3.

Fig. 3.

Conformational response of αβ-dimers to γ-rotation in the hydrolysis direction. (A) Projections of αβ-dimer structures onto the principal component axes PC1 and PC2 obtained from the analysis of X-ray structures (43) during the (Upper) first and (Lower) second half of the simulations with H2PO4 (Left) and HPO42− (Right) in the E site. Red, green, and blue points correspond to αβ that starts from αβE, αβTP, and αβDP, respectively, and purple points correspond to X-ray crystal structures (43). The arrows indicate directions of αβ-motion in response to γ-subunit rotation in the hydrolysis direction. (B) Conformational change of αβTP in response to 30° rotation of the γ-subunit during the trajectory with HPO42−. The snapshot at 30 ns is colored as in Fig. 1B, and the reference structure at 0 ns is colored in cyan. Only the C-terminal domains of αβ are shown for the 0-ns conformation, and the β-hairpin and the C-terminal helix are shown for the γ-subunit. (C) Conformational change of αβE in response to 30° rotation of the γ-subunit during the trajectory with H2PO4. The snapshots at 30 and 0 ns are shown as in B.

The αβE-subunit motions induced by γ-subunit rotation depend sensitively on the protonation state of Pi in the E site. To account for the near-neutral pKa2 of 7.2, we studied HPO42− and H2PO4 in separate simulations. The interface of αβE loosened after a few degrees of rotation with H2PO4, as reflected in the large change of PC2 to negative values (Fig. 3A, Left). The loosening of the αβE interface with γ-rotation in the hydrolysis direction is consistent with a yeast structure at a 12° γ-angle (23), as indicated by their PC2 values of −10 ± 5 Å (molecular dynamics at γ= 12°) and −7.9 Å (X-ray) (43). By contrast, the αβE interface remained relatively tight with HPO42− (Fig. 3A, Right). The doubly charged Pi locks the interface, forcing α- and β-subunits to move together in response to γ-rotation (Fig. S2D). By comparison, with the singly charged Pi, the motion is concentrated in the α-subunit, reflecting the weaker interactions of Pi with the αβE dimer. Thus, the motion of αE, pushed by the N-terminal helix of γ, is primarily responsible for the loosening of the E-site interface (Fig. 3C), which in turn weakens Pi binding to the E site, as studied below.

Protonation State of Pi, Interface Conformation, and Direction of γ-Rotation Affect Binding Stability in the E Site.

Pi moves in the E site as the γ-subunit is rotated. To cancel out the overall rotational motion, we superimposed the binding site on the initial structure, using residues within 10 Å from Pi in the initial structure. Fig. 4A shows the resulting Pi distance from the initial position as a function of time, with γ rotated in the hydrolysis direction by 1°/ns. H2PO4 and HPO42− in the E site responded quite differently to γ-rotation. The singly charged Pi readily moved toward the P-loop region. This motion of Pi was tightly coupled to the loosening of the interface conformation of αβE (Fig. 5A) and was not observed in the free simulation without torque (Fig. S3C). The observed coupling supports the earlier suggestion that Pi binding is regulated by the αβE interface conformation, as inferred from a comparison of different F1 structures (43). The transient structure at 3.5 ns is shown in Fig. 4B, with αARG373 and βLYS162 coordinating Pi. The so-called Arg-finger, αARG373, seems to play a role of guiding Pi to the P loop. In fact, αARG373 remained coordinated to Pi even as Pi moved toward the P loop (Fig. 4A, green line). Pi stayed at the P loop for the rest of the trajectory, frequently interacting with αARG373. The rapid escape of Pi from the E site to the P loop during the 40° substep, as probed in the torque simulations, is consistent with the dramatic increase in the rate of Pi release between 320° and 360° reported from experiment (24). By contrast, the doubly charged Pi remained quite stably bound in the initial position and maintained interactions with its coordinating residues throughout the trajectory (Fig. 4C). With two negative charges, the doubly charged Pi was, in effect, trapped by the four positively charged residues in its immediate surrounding, holding the interface tight during the torque simulations.

Fig. 4.

Fig. 4.

Pi motion in the E site during rotation of γ in the hydrolysis direction. (A) Pi distance from the initial position during the torque simulations. The trajectories with H2PO4 and HPO42− are shown in red and blue, respectively. The distance between H2PO4 and the Cζ atom of αARG373 is plotted in green. (B) Pi (H2PO4) migration toward the P loop at 3.5 ns (Upper) and 39 ns (Lower). (C) Tightly bound Pi (HPO42−) at 20 ns (details as in Fig. 1C).

Fig. 5.

Fig. 5.

Interface conformation and γ-rotational–direction dependence of Pi binding stability. (A) PC2 of the E site (left scale; blue) reporting on αβ interface tightening, and Pi distance relative to starting conformation (right scale; red) during the first 10 ns of the hydrolytic torque simulation with H2PO4. (B) PC2 of the E site in the hydrolytic (blue) and synthetic (purple) torque simulations with H2PO4. (C) (Left) Distance of the singly charged Pi from its starting point for the hydrolytic (red) and synthetic (green) torque simulations. (Right) Distance of the doubly charged Pi for the hydrolytic (red) and synthetic (green) torque simulations.

We also studied the dependence of the Pi binding stability in the E site on the direction of γ-rotation. The singly charged Pi, with the γ-subunit rotated in the synthesis direction, stayed at the initial position for the first ∼30 ns (Fig. 5C, Left). After moving transiently to the P loop, Pi returned back to the initial position at around 70 ns and remained there. These motions suggest that Pi binding in the E site is more stable when γ is rotated in the synthesis direction than in the hydrolysis direction. This increased stability would be important for ATP synthesis, where ADP should bind into the E site already occupied by Pi (see below). The loosening of the αβ interface in response to γ-rotation in the hydrolysis direction, and the tightening in the synthesis direction (Fig. 5B), are likely the main factors changing the Pi binding stability. The doubly charged Pi (Fig. 5C, Right) stayed at the initial position, independent of the direction of the γ-rotation, and maintained tight interactions with the αβE residues.

As a direct demonstration of the coupling between Pi release and γ-rotation, three independent 40-ns free simulations without torque were conducted: (i) without Pi, and with (ii) singly and (iii) doubly charged Pi in the E site (Fig. S3D). The resulting γ-angles are distributed narrowly around 85° with doubly charged Pi, with a wider distribution for singly charged Pi indicating a looser γ-shaft. Importantly, after doubly charged Pi was released, γ indeed rotated by ∼10° in the hydrolysis direction (Fig. S3E). These free simulations thus further support the hypothesis that release of doubly charged Pi from the E site drives γ-rotation in the hydrolysis direction.

Metadynamics Simulation of Pi Release from the E Site and DP Site.

Pi exits the E site (at 320°) and DP site (at 200°) on entirely different pathways according to the metadynamics (38) simulations (Fig. 6A). From the E site, singly and doubly charged Pi both escaped outward via a transient intermediate at the P loop, consistent with the escape route of H2PO4 in the torque simulations. In the DP site, this outward path is blocked by ADP, forcing Pi instead to escape inward and exit at the central hole of the ring subunits (Table S1). This alternate pathway was confirmed by separate simulations with a different metadynamics hill height (0.2 kcal/mol) and for a different F1 structure with a slightly open DP site [Protein Data Bank (PDB) ID code 4asu; Fig. S4].

Fig. 6.

Fig. 6.

Pi release from the E site and DP site. (A) Pathways and (B) free-energy profiles of Pi release determined from metadynamics simulations of singly charged Pi (brown) and doubly charged Pi (yellow) from the E site, and of singly charged Pi from the DP site (green). In A, βE (cyan cartoon) is superimposed on βDP (orange cartoon), and the P loops are colored in purple. ADP, Mg2+, and Pi in the DP site are shown as sticks. (C) Cut through the binding site in the DP state, with arrows indicating exit pathways.

These two exit pathways have drastically different free-energy barriers (Fig. 6B). Escaping from the E site in the z direction, singly and doubly charged Pi face free-energy barriers of ∼5 and ∼10 kcal/mol, respectively, to transition from the tightly bound state to the P loop. The free-energy barrier for Pi to be released from the P loop into solution is ∼5 kcal/mol in both cases. By contrast, the free-energy barrier for Pi to be released from the DP site, moving along the x direction, is almost ∼30 kcal/mol in total, and thus insurmountable on the functional timescale. This high barrier is a result of the tight interface of the DP site, and of ADP blocking the pathway to the P loop, which effectively prevent Pi escape to the outside (Fig. 6C). Our results are consistent with the interpretation of recent single-molecule experiments (13) of Pi being released from the E site, i.e., at 320° in Fig. 1A during the catalytic dwell. The free-energy barriers were confirmed by additional long simulations with a different protocol (SI Text and Fig. S5).

As listed in Table 1, the calculated timescale of ∼2 ms for the release of doubly charged Pi agrees well with the ∼1-ms estimate from experiment (10, 12, 13). The mean first-passage times were calculated from the free-energy profiles and diffusion coefficients (SI Text, Fig. S6, and Table S2). As described above, the Pi movement from the tightly bound state to the P loop is coupled to the global interface motion that occurred within a few nanoseconds (Fig. 5A). This rapid conformational change prevents backtransfer of Pi, which is thus ignored in our timescale estimates. Overall, our simulations suggest that doubly charged Pi is released from the E site at 320° before the 40° substep.

Table 1.

Estimated mean first-passage times of Pi release from the E site

Pi Tightly bound to P loop P loop to tightly bound P loop to release
H2PO4 2.4 μs 0.11 μs 0.34 μs
HPO42− 2.0 ms 0.014 μs 41 μs

Discussion

We found that Pi release from the E site is kinetically realistic and that release from the DP site is not. The calculated time of ∼2 ms for the release of doubly charged Pi from the E site matches the rate of Pi release during the catalytic dwell reported from experiment (13). With this timescale matching also the catalytic dwell time, Pi release could thus be rate-limiting. Our results suggest the need to correct earlier models in which Pi was assumed to exit first from the DP site, just after the hydrolysis reaction, followed by ADP release. Instead, Pi is likely released from the E site at 320° (Fig. 1A), 120° after the hydrolysis reaction, and after ADP release.

Our simulation results help reconcile conflicting interpretations of earlier experiments. Shimo-Kon et al. (21) showed that Pi binding from solution inhibits ATP binding, concluding that the E site must be devoid of Pi. However, at 320°, Pi could have escaped with a ∼1/ms rate and ATP rebound with an estimated on rate of ∼1/ms at 1 mM concentration (24), making the empty E site short-lived in the cycle. Nucleotide binding to all three sites at ∼1 mM concentration (21) could also explain the presence of three ADPs in a recent F1 crystal structure (22). With these ADPs not produced during turnover, but likely taken up from solution (1), Pi would not be kinetically trapped in the DP and E sites. By contrast, tightly bound Pi observed in the E site of some structures (23, 44) likely mimics a product Pi of hydrolysis before release in the catalytic dwell. Gao et al. (20), on the basis of kinetic data on the inhibition of ATP hydrolysis by ADP and Pi, concluded that Pi is released after ADP, consistent with our calculations. We also note that, in the synthesis direction, preferential binding of Pi in the empty E site is advantageous by favoring the subsequent binding of substrate ADP (45) over the abundant but inhibitory product ATP. The ADP/ATP preference with Pi bound could be probed in experiment (21).

The combination of experiment (13) and simulation suggests that Pi is released as HPO42−. According to quantum chemical calculations, Pi is doubly protonated (singly charged) just after hydrolysis (25). However, after ADP and Mg2+ (net charge is −1) leave the site, four positively charged residues surround Pi, shifting its pKa. This electrostatic environment would indeed favor the loss of one proton, resulting in a singly protonated (doubly charged) Pi in the E site. Probing the pH dependence of the dwell time of the Pi release (or the 40° substep) could test this hypothesis, assuming protonic equilibrium with the solvent.

We also investigated the conformational changes of F1 in response to the 40° rotation step of the γ-subunit, and their coupling to Pi release. γ was rotated both in the hydrolysis and synthesis directions, for two possible protonation states of Pi in the E site. By rotating for 40° in the hydrolysis direction relative to the available crystal structures, we obtained a first fully atomistic conformation of the intermediate state following the 40° substep. We found that the structure of the ATP-bound αβTP interface became as tight as that of ADP+Pi-bound αβDP interface, thus facilitating hydrolysis. By comparison, the αβE interface became looser, and Pi escaped to the P loop. We expect these structures to be representative of the conformations of the two sites in the binding dwell state. The αβDP site remains relatively tight and shows only a slight opening, as indicated by the displacement of purple and blue points in Fig. 3A. We conclude that αβDP motions in response to γ-rotation are considerably slower than those of αβE and αβTP (16). The salt bridges between γ and α3β3 formed in the intermediate state are listed in SI Text. To test whether the intermediate state is stable without torque applied to γ, two 40-ns free simulations were conducted with the torque switched off at 40° and 50° from the catalytic dwell, respectively. In both simulations, the γ-angle settled around 110° (30° from the catalytic dwell), indicating the formation of a metastable state (Fig. S7A). Importantly, the torque simulations drive the system out-of-equilibrium (as would the Fo motor in intact ATP synthase). In such a nonequilibrium process, the γ-rotor will be slightly over-twisted. Releasing the torque that drives the rotation is thus expected to result in a fast relaxation of the angle. We indeed observed a relaxation on a ∼5-ns timescale after releasing the torque at 120°, close to the presumed metastable minimum (Fig. S7A, green curve). When we released the torque at 130°, beyond the minimum, we again saw this fast recoil, and in addition a more gradual relaxation to a state at ∼110°, consistent with an overdamped, weakly driven motion in a metastable minimum at ∼110°. The interface conformations of αβE and αβTP relaxed partially, but importantly in a reversible way (Fig. S7B).

These interfacial motions are coupled to torque generation, which should make it possible to probe our results by mutation. Torque generation as a result of nucleotide binding/dissociation has been probed extensively by mutations of β–γ interactions (46). By contrast, relatively little is known about the specific mechanisms of torque generation as a result of Pi release. The interfacial motions seen in our simulations suggest that torque generation by Pi release should be sensitive to the hydrophobic interactions between αPHE403/αPHE406 and γILE16/γILE19/γLEU77, and the electrostatic interactions between αASP409 and γARG133, αASP411 and γLYS30, αGLU355 and γLYS11, and αGLU399 and γLYS18. Perturbations of these interactions by mutation should affect the coupling between the αβE interface and the γ-subunit, and should thus report directly on torque generation as a result of Pi release.

Remarkably, the two Pi exit pathways observed in this study appear to be conserved in myosin, and possibly other members of the large family of P-loop ATPases. In particular, a so-called “back door” exit pathway for Pi release has been proposed for myosin, to circumvent the “front door” pathway blocked by ADP in the binding pocket (47). Molecular dynamics simulations showed that the Pi release pathway is tightly linked to the cleft-closing motion of myosin upon binding to actin (48, 49). The back door route dominated for an open cleft and a closed binding site, whereas the front door and a variant of the back door route dominated for a closed cleft and an open binding site (49). We have a similar situation in F1, although the global motions are quite different. With the binding site closed (DP site), Pi escaped via an inward route that corresponds to the back door in myosin (toward switch II). By contrast, with the binding site open (E site), we observed the outward route corresponding to the front door in myosin (Fig. S8). Despite large differences in the global functional motions of different P-loop ATPases, local motions during the hydrolysis cycle thus appear to be conserved. Functional requirements then dictate the dominant route of Pi exit in different ATPases, and possibly also GTPases. Myosin uses both back and front doors, whereas F1 takes the front door exclusively, as shown here.

Methods

Initial Structures.

The initial structures of F1 were taken from the azide-free crystal structure (PDB ID code 2jdi) (50). Residues 24–510 were used for the α-subunits, 9–478 were used for the β-subunits, and all residues were used for the γ-subunit. MODELER (51) was used to model structures of missing residues. Missing residues in the γ-subunit were modeled according to the dicyclohexylcarbodiimide-inhibited structure (PDB ID code 1e79) (52). In the azide-free structure, the TP site and DP site bind an ATP analog (AMPPNP) and Mg2+, and the E site is empty. For the simulation system, the ATP analog in the TP site was replaced with ATP, and that in the DP site was replaced with ADP and Pi (H2PO4). Pi (HPO42− or H2PO4) was modeled in the E site, representing the catalytic dwell state just after ATP hydrolysis (SI Text). A variety of analyses (1, 1419) consistently indicate that the initial crystal structure represents the catalytic dwell, which corresponds to a γ-angle of 80° by definition (Fig. 1A). The half-closed structure by Menz et al. (44) was not used here because it possibly corresponds to an E site occupied by ADP rebound from solution at high ADP concentration, and thus would not be part of the actual sequence of product release. Pi in the DP site was modeled as doubly protonated following the quantum chemical studies of ATP hydrolysis in this site (25). Crystal water molecules were retained unless they clearly overlap with replaced ligands.

Simulation Setups.

The modeled F1 with crystal waters was solvated with TIP3P water (53) in a rectangular box such that the minimum distance to the edge of the box is 11 Å, and neutralized with potassium ions. Then, ∼100 mM KCl was added to the system. The total number of atoms is ∼313,000 (Fig. S1). The Amber ff99SB force field was used for the protein (54), with modified parameters for KCl (55), and for ATP and ADP (56). The point charges for singly and doubly protonated Pi were determined using Gaussian 03 (57) and the Antechamber module of Amber (58) (SI Text). The system was energy minimized and equilibrated at isothermal–isobaric (NPT) conditions with Ewald electrostatics and restraints on all heavy atoms in the protein for 500 ps, and subsequently, with restraints on only Cα atoms for 1 ns. After the equilibration, production runs were conducted with restraints on the Cα atoms of the 10 N-terminal residues of each of the three β-subunits, mimicking the role of His-tag anchoring in single-molecule experiments. NAMD 2.8 was used for the molecular dynamics simulations with periodic boundary condition (59). Langevin dynamics with 1 ps−1 damping coefficient was used for temperature control at 310 K, and the Nose–Hoover Langevin piston was used for pressure control at 1 atm (60).

Flexible Rotor Simulation.

In the torque simulations, only the average rotational angle of the γ-subunit was restrained. As in an earlier method (61), the γ-subunit is allowed to twist freely, but here without arbitrary subdivision into slabs. We also allow free translation of its center of mass. Details of the method and explicit expressions for the restraint forces are in SI Text.

Metadynamics Simulation of Pi Release.

To enhance the sampling and estimate free-energy profiles for Pi release, metadynamics (38) simulations were conducted using the PLUMED 1.3 plugin with NAMD (62). The relative Cartesian coordinates x, y, and z of Pi from the binding site were used as collective variables (63), where the binding site was defined by the Cα atoms of β-subunit residues 188, 256, 257, 259, and 309–312, which are within 8 Å from Pi in the DP site. Flexible α and P-loop residues are excluded. Repulsive Gaussian potentials with heights of 0.1 kcal/mol and widths of 0.6 Å were deposited every 1 ps during the metadynamics simulations. The free-energy profiles were calculated by a PLUMED utility, sum_hills.

Supplementary Material

Supporting Information

Acknowledgments

We thank Drs. Kazuhiko Kinosita and Hiroyuki Noji for insightful discussions, and Drs. Edina Rosta, Fangqiang Zhu, Jürgen Köfinger, Pilar Cossio, and Ikuo Kurisaki for technical help. This work was supported by the Intramural Research Program of the National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health and by the Max Planck Society. K.-i.O. is supported by research fellowship for young scientists and postdoctoral fellowship for research abroad of the Japan Society for the Promotion of Science. The research used the high-performance computational capabilities of the Biowulf Linux cluster at the National Institutes of Health (http://biowulf.nih.gov).

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1305497110/-/DCSupplemental.

References

  • 1.Junge W, Sielaff H, Engelbrecht S. Torque generation and elastic power transmission in the rotary FOF1-ATPase. Nature. 2009;459(7245):364–370. doi: 10.1038/nature08145. [DOI] [PubMed] [Google Scholar]
  • 2.Abrahams JP, Leslie AG, Lutter R, Walker JE. Structure at 2.8 A resolution of F1-ATPase from bovine heart mitochondria. Nature. 1994;370(6491):621–628. doi: 10.1038/370621a0. [DOI] [PubMed] [Google Scholar]
  • 3.Noji H, Yasuda R, Yoshida M, Kinosita K., Jr Direct observation of the rotation of F1-ATPase. Nature. 1997;386(6622):299–302. doi: 10.1038/386299a0. [DOI] [PubMed] [Google Scholar]
  • 4.Yasuda R, Noji H, Kinosita K, Jr, Yoshida M. F1-ATPase is a highly efficient molecular motor that rotates with discrete 120 degree steps. Cell. 1998;93(7):1117–1124. doi: 10.1016/s0092-8674(00)81456-7. [DOI] [PubMed] [Google Scholar]
  • 5.Toyabe S, Watanabe-Nakayama T, Okamoto T, Kudo S, Muneyuki E. Thermodynamic efficiency and mechanochemical coupling of F1-ATPase. Proc Natl Acad Sci USA. 2011;108(44):17951–17956. doi: 10.1073/pnas.1106787108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Itoh H, et al. Mechanically driven ATP synthesis by F1-ATPase. Nature. 2004;427(6973):465–468. doi: 10.1038/nature02212. [DOI] [PubMed] [Google Scholar]
  • 7.Rondelez Y, et al. Highly coupled ATP synthesis by F1-ATPase single molecules. Nature. 2005;433(7027):773–777. doi: 10.1038/nature03277. [DOI] [PubMed] [Google Scholar]
  • 8.Furuike S, et al. Axle-less F1-ATPase rotates in the correct direction. Science. 2008;319(5865):955–958. doi: 10.1126/science.1151343. [DOI] [PubMed] [Google Scholar]
  • 9.Uchihashi T, Iino R, Ando T, Noji H. High-speed atomic force microscopy reveals rotary catalysis of rotorless F1-ATPase. Science. 2011;333(6043):755–758. doi: 10.1126/science.1205510. [DOI] [PubMed] [Google Scholar]
  • 10.Yasuda R, Noji H, Yoshida M, Kinosita K, Jr, Itoh H. Resolution of distinct rotational substeps by submillisecond kinetic analysis of F1-ATPase. Nature. 2001;410(6831):898–904. doi: 10.1038/35073513. [DOI] [PubMed] [Google Scholar]
  • 11.Nishizaka T, et al. Chemomechanical coupling in F1-ATPase revealed by simultaneous observation of nucleotide kinetics and rotation. Nat Struct Mol Biol. 2004;11(2):142–148. doi: 10.1038/nsmb721. [DOI] [PubMed] [Google Scholar]
  • 12.Adachi K, et al. Coupling of rotation and catalysis in F1-ATPase revealed by single-molecule imaging and manipulation. Cell. 2007;130(2):309–321. doi: 10.1016/j.cell.2007.05.020. [DOI] [PubMed] [Google Scholar]
  • 13.Watanabe R, Iino R, Noji H. Phosphate release in F1-ATPase catalytic cycle follows ADP release. Nat Chem Biol. 2010;6(11):814–820. doi: 10.1038/nchembio.443. [DOI] [PubMed] [Google Scholar]
  • 14.Yasuda R, et al. The ATP-waiting conformation of rotating F1-ATPase revealed by single-pair fluorescence resonance energy transfer. Proc Natl Acad Sci USA. 2003;100(16):9314–9318. doi: 10.1073/pnas.1637860100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Okuno D, et al. Correlation between the conformational states of F1-ATPase as determined from its crystal structure and single-molecule rotation. Proc Natl Acad Sci USA. 2008;105(52):20722–20727. doi: 10.1073/pnas.0805828106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Masaike T, Koyama-Horibe F, Oiwa K, Yoshida M, Nishizaka T. Cooperative three-step motions in catalytic subunits of F(1)-ATPase correlate with 80 degrees and 40 degrees substep rotations. Nat Struct Mol Biol. 2008;15(12):1326–1333. doi: 10.1038/nsmb.1510. [DOI] [PubMed] [Google Scholar]
  • 17.Sielaff H, Rennekamp H, Engelbrecht S, Junge W. Functional halt positions of rotary FOF1-ATPase correlated with crystal structures. Biophys J. 2008;95(10):4979–4987. doi: 10.1529/biophysj.108.139782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Sielaff H, Börsch M. Twisting and subunit rotation in single FOF1-ATP synthase. Philos Trans R Soc Lond B Biol Sci. 2013;368(1611):20120024. doi: 10.1098/rstb.2012.0024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Sun SX, Wang HY, Oster G. Asymmetry in the F1-ATPase and its implications for the rotational cycle. Biophys J. 2004;86(3):1373–1384. doi: 10.1016/S0006-3495(04)74208-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Gao YQ, Yang W, Karplus M. A structure-based model for the synthesis and hydrolysis of ATP by F1-ATPase. Cell. 2005;123(2):195–205. doi: 10.1016/j.cell.2005.10.001. [DOI] [PubMed] [Google Scholar]
  • 21.Shimo-Kon R, et al. Chemo-mechanical coupling in F(1)-ATPase revealed by catalytic site occupancy during catalysis. Biophys J. 2010;98(7):1227–1236. doi: 10.1016/j.bpj.2009.11.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Rees DM, Montgomery MG, Leslie AGW, Walker JE. Structural evidence of a new catalytic intermediate in the pathway of ATP hydrolysis by F1-ATPase from bovine heart mitochondria. Proc Natl Acad Sci USA. 2012;109(28):11139–11143. doi: 10.1073/pnas.1207587109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kabaleeswaran V, Puri N, Walker JE, Leslie AG, Mueller DM. Novel features of the rotary catalytic mechanism revealed in the structure of yeast F1 ATPase. EMBO J. 2006;25(22):5433–5442. doi: 10.1038/sj.emboj.7601410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Watanabe R, et al. Mechanical modulation of catalytic power on F1-ATPase. Nat Chem Biol. 2012;8(1):86–92. doi: 10.1038/nchembio.715. [DOI] [PubMed] [Google Scholar]
  • 25.Hayashi S, et al. Molecular mechanism of ATP hydrolysis in F1-ATPase revealed by molecular simulations and single-molecule observations. J Am Chem Soc. 2012;134(20):8447–8454. doi: 10.1021/ja211027m. [DOI] [PubMed] [Google Scholar]
  • 26.Beke-Somfai T, Lincoln P, Nordén B. Rate of hydrolysis in ATP synthase is fine-tuned by α-subunit motif controlling active site conformation. Proc Natl Acad Sci USA. 2013;110(6):2117–2122. doi: 10.1073/pnas.1214741110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Dittrich M, Hayashi S, Schulten K. On the mechanism of ATP hydrolysis in F1-ATPase. Biophys J. 2003;85(4):2253–2266. doi: 10.1016/s0006-3495(03)74650-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Böckmann RA, Grubmüller H. Conformational dynamics of the F1-ATPase beta-subunit: A molecular dynamics study. Biophys J. 2003;85(3):1482–1491. doi: 10.1016/S0006-3495(03)74581-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Ito Y, Oroguchi T, Ikeguchi M. Mechanism of the conformational change of the F1-ATPase β subunit revealed by free energy simulations. J Am Chem Soc. 2011;133(10):3372–3380. doi: 10.1021/ja1070152. [DOI] [PubMed] [Google Scholar]
  • 30.Antes I, Chandler D, Wang HY, Oster G. The unbinding of ATP from F1-ATPase. Biophys J. 2003;85(2):695–706. doi: 10.1016/S0006-3495(03)74513-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Ito Y, Ikeguchi M. Structural fluctuation and concerted motions in F1-ATPase: A molecular dynamics study. J Comput Chem. 2010;31(11):2175–2185. doi: 10.1002/jcc.21508. [DOI] [PubMed] [Google Scholar]
  • 32.Czub J, Grubmüller H. Torsional elasticity and energetics of F1-ATPase. Proc Natl Acad Sci USA. 2011;108(18):7408–7413. doi: 10.1073/pnas.1018686108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Böckmann RA, Grubmüller H. Nanoseconds molecular dynamics simulation of primary mechanical energy transfer steps in F1-ATP synthase. Nat Struct Biol. 2002;9(3):198–202. doi: 10.1038/nsb760. [DOI] [PubMed] [Google Scholar]
  • 34.Ma J, et al. A dynamic analysis of the rotation mechanism for conformational change in F(1)-ATPase. Structure. 2002;10(7):921–931. doi: 10.1016/s0969-2126(02)00789-x. [DOI] [PubMed] [Google Scholar]
  • 35.Koga N, Takada S. Folding-based molecular simulations reveal mechanisms of the rotary motor F1-ATPase. Proc Natl Acad Sci USA. 2006;103(14):5367–5372. doi: 10.1073/pnas.0509642103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Pu J, Karplus M. How subunit coupling produces the gamma-subunit rotary motion in F1-ATPase. Proc Natl Acad Sci USA. 2008;105(4):1192–1197. doi: 10.1073/pnas.0708746105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Mukherjee S, Warshel A. Electrostatic origin of the mechanochemical rotary mechanism and the catalytic dwell of F1-ATPase. Proc Natl Acad Sci USA. 2011;108(51):20550–20555. doi: 10.1073/pnas.1117024108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Laio A, Parrinello M. Escaping free-energy minima. Proc Natl Acad Sci USA. 2002;99(20):12562–12566. doi: 10.1073/pnas.202427399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Chung HS, McHale K, Louis JM, Eaton WA. Single-molecule fluorescence experiments determine protein folding transition path times. Science. 2012;335(6071):981–984. doi: 10.1126/science.1215768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Hummer G, Eaton WA. Transition path times for DNA and RNA folding from force spectroscopy. Physics. 2012;5:87. [Google Scholar]
  • 41.Walker JE. The ATP synthase: The understood, the uncertain and the unknown. Biochem Soc Trans. 2013;41(1):1–16. doi: 10.1042/BST20110773. [DOI] [PubMed] [Google Scholar]
  • 42.Wächter A, et al. Two rotary motors in F-ATP synthase are elastically coupled by a flexible rotor and a stiff stator stalk. Proc Natl Acad Sci USA. 2011;108(10):3924–3929. doi: 10.1073/pnas.1011581108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Okazaki K, Takada S. Structural comparison of F1-ATPase: Interplay among enzyme structures, catalysis, and rotations. Structure. 2011;19(4):588–598. doi: 10.1016/j.str.2011.01.013. [DOI] [PubMed] [Google Scholar]
  • 44.Menz RI, Walker JE, Leslie AG. Structure of bovine mitochondrial F(1)-ATPase with nucleotide bound to all three catalytic sites: Implications for the mechanism of rotary catalysis. Cell. 2001;106(3):331–341. doi: 10.1016/s0092-8674(01)00452-4. [DOI] [PubMed] [Google Scholar]
  • 45.Ahmad Z, Senior AE. Identification of phosphate binding residues of Escherichia coli ATP synthase. J Bioenerg Biomembr. 2005;37(6):437–440. doi: 10.1007/s10863-005-9486-8. [DOI] [PubMed] [Google Scholar]
  • 46.Tanigawara M, et al. Role of the DELSEED loop in torque transmission of F1-ATPase. Biophys J. 2012;103(5):970–978. doi: 10.1016/j.bpj.2012.06.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Yount RG, Lawson D, Rayment I. (1995) Is myosin a “back door” enzyme? Biophys J 68(4 Suppl):44S–47S; discussion 47S–49S. [PMC free article] [PubMed]
  • 48.Lawson JD, Pate E, Rayment I, Yount RG. Molecular dynamics analysis of structural factors influencing back door Pi release in myosin. Biophys J. 2004;86(6):3794–3803. doi: 10.1529/biophysj.103.037390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Cecchini M, Alexeev Y, Karplus M. Pi release from myosin: A simulation analysis of possible pathways. Structure. 2010;18(4):458–470. doi: 10.1016/j.str.2010.01.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Bowler MW, Montgomery MG, Leslie AG, Walker JE. Ground state structure of F1-ATPase from bovine heart mitochondria at 1.9 A resolution. J Biol Chem. 2007;282(19):14238–14242. doi: 10.1074/jbc.M700203200. [DOI] [PubMed] [Google Scholar]
  • 51.Eswar N, et al. (2006) Comparative protein structure modeling using Modeller. Curr Protoc Bioinformatics Chap 5:Unit 5.6. [DOI] [PMC free article] [PubMed]
  • 52.Gibbons C, Montgomery MG, Leslie AG, Walker JE. The structure of the central stalk in bovine F(1)-ATPase at 2.4 A resolution. Nat Struct Biol. 2000;7(11):1055–1061. doi: 10.1038/80981. [DOI] [PubMed] [Google Scholar]
  • 53.Jorgensen WL, Chandrasekhar J, Madura J, Klein ML. Comparison of simple potential functions for simulating liquid water. J Chem Phys. 1983;79(2):926–935. [Google Scholar]
  • 54.Hornak V, et al. Comparison of multiple Amber force fields and development of improved protein backbone parameters. Proteins. 2006;65(3):712–725. doi: 10.1002/prot.21123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Joung IS, Cheatham TE., 3rd Determination of alkali and halide monovalent ion parameters for use in explicitly solvated biomolecular simulations. J Phys Chem B. 2008;112(30):9020–9041. doi: 10.1021/jp8001614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Meagher KL, Redman LT, Carlson HA. Development of polyphosphate parameters for use with the AMBER force field. J Comput Chem. 2003;24(9):1016–1025. doi: 10.1002/jcc.10262. [DOI] [PubMed] [Google Scholar]
  • 57.Frisch MJ, et al. (2004) Gaussian 03, Revision C.02 (Gaussian, Inc., Wallingford, CT)
  • 58.Wang J, Wang W, Kollman PA, Case DA. Automatic atom type and bond type perception in molecular mechanical calculations. J Mol Graph Model. 2006;25(2):247–260. doi: 10.1016/j.jmgm.2005.12.005. [DOI] [PubMed] [Google Scholar]
  • 59.Phillips JC, et al. Scalable molecular dynamics with NAMD. J Comput Chem. 2005;26(16):1781–1802. doi: 10.1002/jcc.20289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Feller SE, Zhang Y, Pastor RW, Brooks BR. Constant pressure molecular dynamics simulation: The Langevin piston method. J Chem Phys. 1995;103(11):4613–4621. [Google Scholar]
  • 61.Kutzner C, Czub J, Grubmüller H. Keep it flexible: Driving macromolecular rotary motions in atomistic simulations with GROMACS. J Chem Theory Comput. 2011;7(5):1381–1393. doi: 10.1021/ct100666v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Bonomi M, et al. PLUMED: A portable plugin for free-energy calculations with molecular dynamics. Comput Phys Commun. 2009;180(10):1961–1972. [Google Scholar]
  • 63.Nishihara Y, Hayashi S, Kato S. A search for ligand diffusion pathway in myoglobin using a metadynamics simulation. Chem Phys Lett. 2008;464(4):220–225. [Google Scholar]
  • 64. Humphrey W, Dalke A, Schulten K (1996) VMD: Visual molecular dynamics. J Mol Graph 14(1):33–38. [DOI] [PubMed]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information
Download video file (3.5MB, mpg)

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES