Skip to main content
Biophysical Journal logoLink to Biophysical Journal
. 2015 Mar 10;108(5):1144–1152. doi: 10.1016/j.bpj.2015.01.017

Torque Transmission Mechanism via DELSEED Loop of F1-ATPase

Rikiya Watanabe 1,2,3, Kazuma Koyasu 1, Huijuan You 1, Mizue Tanigawara 1, Hiroyuki Noji 1,3,
PMCID: PMC4375457  PMID: 25762326

Abstract

F1-ATPase (F1) is an ATP-driven rotary motor in which the three catalytic β subunits in the stator ring sequentially induce the unidirectional rotation of the rotary γ subunit. Many lines of evidence have revealed open-to-closed conformational transitions in the β subunit that swing the C-terminal domain inward. This conformational transition causes a C-terminal protruding loop with conserved sequence DELSEED to push the γ subunit. Previous work, where all residues of DELSEED were substituted with glycine to disrupt the specific interaction with γ and introduce conformational flexibility, showed that F1 still rotated, but that the torque was halved, indicating a remarkable impact on torque transmission. In this study, we conducted a stall-and-release experiment on F1 with a glycine-substituted DELSEED loop to investigate the impact of the glycine substitution on torque transmission upon ATP binding and ATP hydrolysis. The mutant F1 showed a significantly reduced angle-dependent change in ATP affinity, whereas there was no change in the equilibrium for ATP hydrolysis. These findings indicate that the DELSEED loop is predominantly responsible for torque transmission upon ATP binding but not for that upon ATP hydrolysis.

Introduction

FOF1 ATP synthase catalyzes the terminal reaction of oxidative phosphorylation, ATP synthesis from ADP and inorganic phosphate (Pi), using the electrochemical potential of protons (or sodium in some organisms) across the mitochondrial inner membrane, the chloroplast thylakoid membrane, and the bacterial plasma membrane (1–5). FOF1 ATP synthase is structurally and functionally separated into two distinct parts, F1 and FO. F1 is the water-soluble portion of ATP synthase responsible for catalysis. When isolated from the membrane domain, F1 exhibits strong ATP hydrolysis activity. F1 is an ATP-driven rotary motor in which the rotary shaft rotates against the surrounding stator upon ATP hydrolysis. FO is also a rotary motor driven by proton flow down the electrochemical potential of the proton gradient (6,7). FO and F1 are connected by a shared rotary shaft and a peripheral stalk that transmit their competing rotary torques. Under physiological conditions with sufficient electrochemical potential, FO generates greater torque than F1 and enforces rotation of F1 in the reverse direction to drive the reverse chemical reaction of ATP hydrolysis, ATP synthesis. When the proton electrochemical potential diminishes, F1 reverses direction and rotates FO to reestablish the proton electrochemical potential.

The subunit composition of bacterial F1 is α3β3γδε. The minimum complex of F1 as a rotary motor is the α3β3γ subcomplex (referred to here as F1 for simplicity). Three α and three β subunits form the α3β3 stator ring with a central cavity that accommodates the central rotary shaft, γ. The catalytic reaction centers for ATP hydrolysis/synthesis reside at the αβ interfaces, mainly on the β subunit (8–10).

The rotary dynamics of F1 have been well studied in single-molecule rotation assays (11–13). The basic rotational properties of F1 from thermophilic Bacillus PS3 (TF1), the best-characterized F1 in terms of rotational mechanism, are summarized below. F1 exerts a constant torque of 40 pN·nm irrespective of rotary angle, as estimated from the hydrodynamic friction of a rotation marker attached to the γ subunit (14,15). Higher torque was reported for F1 from Escherichia coli (EF1) (16,17), although another group reported rather lower torque for EF1 (18). Each rotational step is 120° (14), as expected from the pseudo-threefold symmetry of F1. Each 120° step is coupled with hydrolysis of a single molecule of ATP. The work per ATP hydrolysis is thus estimated as 80 pN·nm (torque times step size). This value corresponds to the free-energy change for hydrolysis of one ATP molecule, suggesting an extremely high energy conversion efficiency of F1. This point was later verified in an experiment in which constant external torque was applied to a rotating F1 (19). The 120° step can be further resolved into 80° and 40° substeps (20,21) that differ significantly in other F1 complexes (18,22). The 80° substep is triggered by ATP binding and ADP release, which occur on different β subunits (23–25). The 40° substep is initiated after hydrolysis of bound ATP and Pi release, which are also thought to occur on different β subunits (26–28). Hereafter, the angular positions of F1 before the 80° and 40° substeps are referred to as the binding and catalytic angles, respectively. When the rotary angle for ATP binding to a certain β subunit is defined as 0°, ATP hydrolysis, ADP release, and Pi release occur at 200°, 240°, and 320°, respectively (Fig. 1 a) (29). Thus, each β subunit catalyzes a single turnover of ATP hydrolysis for one turn of the γ subunit, but the phase of the catalytic state differs by 120° among the three β subunits. Although most of the reaction scheme has been established based on the single-molecule rotation assay, some elements are controversial in biochemical and x-ray structural studies, especially regarding the order of product release (30,31).

Figure 1.

Figure 1

DELSEED loop mutants. (a) Chemomechanical coupling scheme of F1 at low ATP concentration. The circles and red arrows represent the catalytic state of the β subunits and the angular positions of the γ subunit, respectively. One catalytic site (green) is shown undergoing the binding and catalytic events. The other two catalytic sites are undergoing the same events simultaneously, but offset by 120° and 240°. (b) Crystal structure of the γ subunit (red) and the β subunit in the empty state with (green) or without (gray) the glyloop mutation. (PDB ID 2JDI or 1BMF). The structure of the glyloop mutant was determined by molecular dynamics simulations (50). (c) Time courses of rotary motion in the presence of 1 mM ATP. Gray, red, and blue represent the time courses of F1WT, F1glyloop, and F1glyloop/E190D, respectively. The traces at left are a rescaled view of those at right. The data for F1glyloop and F1WT were measured byTanigawara et al. (50). (d) Rotational velocity (V) of F1glyloop (red), F1glyloop/E190D (blue), and F1WT (gray) at various ATP concentrations. The curves represent Michaelis-Menten fits with V = Vmax[ATP]/([ATP]+Km), where Vmaxglyloop = 119 s−1, Vmaxglyloop/E190D = 0.43 s−1, VmaxWT = 169 s−1, Kmglyloop = 23 μM, Kmglyloop/E190D = 1.0 μM, and KmWT = 22 μM. The corresponding rate constants for ATP binding, kon = 3 × Vmax/Km, are konglyloop = 1.5 × 107 M−1·s−1, konglyloop/E190D = 1.2 × 106 M−1·s−1, and konWT = 2.6 × 107 M−1·s−1. The data for F1glyloop and F1WT were measured by Tanigawara et al. (50). To see this figure in color, go online.

In early single-molecule studies, the contribution of each catalytic step to torque generation was estimated from the substep sizes induced by each reaction (20). More quantitative estimation was provided from a recent single-molecule manipulation experiment using stall and release, where paused F1 waiting for a certain reaction step to occur was stalled with and released from magnetic tweezers (32,33). Because F1 does not exert torque in the absence of its catalytic reaction, the probability of the reaction can be measured as the probability that F1 rotates after release from the magnetic tweezers. This stall-and-release experiment determined the rate and equilibrium constants of ATP binding and ATP hydrolysis as functions of the rotary angle and showed that the affinity of F1 for ATP significantly increases upon the counterclockwise rotation in the hydrolysis direction, whereas the equilibrium constant for ATP hydrolysis is relatively independent of angle. The estimated energy released for each step suggests that F1 generates three- to fourfold higher torque through the affinity change for ATP than in the hydrolysis step (32,34).

Crystal structures of bovine mitochondrial F1 (MF1) provided essential structural insights into torque generation by the catalytic β subunits and torque transmission to the γ subunit. In the original reference structure (8) or the ground-state structure (35) that represent the common structural features shared among most other crystal structures of MF1 or F1 from other species (9,10), the three β subunits are termed βTP, βDP, and βempty according to the bound nucleotide. βTP is bound to an analog of ATP, β,γ-imido-triphosphate (AMP-PNP), βDP is bound to ADP and azide, and βempty has no nucleotide bound. Both βTP and βDP assume a closed conformation that rotates the C-terminal helical domain inward toward the rotational axis and enclosing the bound nucleotide, whereas βempty has an open conformation (Fig. 1 b). The open-to-closed conformational transition of β in solution was measured by NMR (36) and single-molecule imaging (37). Although no significant conformational differences were identified between βTP and βDP, changes in the αβ interface were observed: the αβDP interface is more tightly closed than the αβTP interface. Because tight packing of the αβ interface accompanies a positional shift of the catalytically critical arginine residue on the α subunit, referred to as the arginine finger, toward bound ATP (38–40), the αβDP interface is thought to be very similar to the catalytically active state. This contention is supported by several lines of experimental (41–43) and theoretical evidence (40,44). Based on these structural features, nucleotide binding and release are thought to induce the open-to-closed or closed-to-open conformational transitions of the β subunit, respectively, whereas ATP hydrolysis accompanies changes in the αβ interface. Recent theoretical studies suggest that Pi release is also coupled with the conformational transition of the αβ interface (45,46).

The α3β3 and γ subunits interact in the upper and lower regions. α3β3 forms a hydrophobic sleeve in the lower region that does not show obvious differences among the three β subunits. Therefore, the lower interface is thought to act mainly as an axle holder for the C-terminal helix of the γ subunit, although some reports suggest that this interaction is also involved in torque transmission (47). The upper contact is mainly formed by the γ subunit and the C-terminal domains of the β subunits, and this region shows distinct differences among the β subunits. The C-terminal domain of the β subunits contacts the γ subunit with a helical loop, referred to as the DELSEED loop for its conserved amino acid sequence. The DELSEED loop of TF1 used in this study has the amino acid sequence DELSDED (amino acids 386–394). Because the DELSEED loop rotates remarkably upon the open-to-closed conformational transition, it is thought to be critical for torque transmission and has been extensively studied (48–50). Deletion or glycine substitution of DELSEED residues significantly reduced the torque production of F1.

Thus, the critical role of the DELSEED loop in torque transmission between the β-γ subunits has been determined. However, the specific reaction steps involved in torque transmission via the DELSEED loop remain unclear. In this work, we studied the effect of glycine substitution in the DELSEED loop on torque generation in each catalytic step. We conducted stall-and-release experiments of glycine-substituted DELSEED-loop mutant F1 (F1glyloop) to examine the angle dependence of ATP binding and hydrolysis. Our data indicate that although the angle dependence of the hydrolysis step was not affected by glycine substitution, the angle dependence of ATP binding (affinity change) was significantly decreased, suggesting that F1glyloop is deficient in torque transmission coupled with the affinity change for bound ATP.

Materials and Methods

Rotation assay

To visualize the rotation of F1, the stator subunits (α3β3), which have His-tags at the N-terminus, were fixed on a glass surface modified with Ni-NTA, and a streptavidin-coated magnetic bead (ϕ = 0.2–0.4 μm; Seradyn, Indianapolis, IN) was attached on the top of a biotinylated rotor subunit (γ) to probe rotation. The flow chamber was constructed from an uncoated top coverslip and a bottomcover slip modified with Ni-NTA for immobilization of His-tagged F1. F1 was diluted with buffer A (50 mM MOPS-KOH and 50 mM KCl, pH 7.0) to a final concentration of 200 pM and infused into the flow chamber. After 5 min, unbound F1 molecules were washed out with buffer A containing 10% bovine serum albumin, then streptavidin-coated beads in buffer A were infused. After 10 min, unbound beads were washed out with buffer A. Finally, buffer A containing the indicated amount of Mg-ATP and Pi was infused. The rotating beads were observed under a phase-contrast microscope (IX-70 or IX-71, Olympus, Center Valley, PA) with a 100× objective lens. The rotation assay was performed at 25°C ± 3°C.

Manipulation with magnetic tweezers

The microscope stage was equipped with magnetic tweezers (51) controlled by custom software (Celery, Library). Using progressive-scan cameras, the rotary motion of the bead was imaged at 30 frames/s for the angle-dependence experiments in Figs. 3, 4, and 5 (FC300M, Takex, Kyoto, Japan) or at 1000 frames/s for the torque and rotary potential experiments in Fig. 2 (FASTCAM 1024PCI-SE, Photoron, Tokyo, Japan). Images were stored in the HDD of a computer as AVI files and analyzed using custom software.

Figure 3.

Figure 3

Single-molecule manipulation of F1. (a) Schematic of manipulation procedures. When F1 paused at the ATP binding or hydrolysis dwell, the magnetic tweezers were turned on to stall F1 at the target angle then off to release the motor after the set time. Released F1 either steps forward (ON) or returns to the original pause angle (OFF). These behaviors indicate that the reaction under investigation has occurred or not, respectively. (b) Examples of stall-and-release traces for ATP binding at 200 nM ATP. During a pause, F1glyloop was stalled for 1 s and then released. (Left) After release, F1glyloop stepped to the next binding angle without moving back, indicating that ATP had already bound to F1glyloop before release. (Right) When stalled for 1 s, F1glyloop rotated back to the original binding angle, indicating that no ATP binding had occurred. To see this figure in color, go online.

Figure 4.

Figure 4

Angle dependence of ATP binding to F1glyloop. (a) Time courses of PON of F1glyloop at 200 nM ATP after stalling at θb = −30° (cyan), 0° (black), +30° (green), or +50° (yellow) from the original ATP-binding angle. The gray line represents the time course for free rotation. konATP and koffATP were determined by fits to a single exponential function: PON = (konATP·[ATP]/(konATP[ATP] + koffATP)) × (1 − exp(−(konATP[ATP] + koffATP) × t)), according to the reversible reaction scheme F1 + ATP ⇄ F1 × ATP. Each data point was obtained from 20–151 trials using five molecules. The error in PON is given as PON(100PON)/N, where N is the number of trials for each stall measurement. (b) Histograms of ATP binding dwell times at 200 nM ATP in free rotation (yellow) and for F1 after an OFF (green) or ON (red) event. Each analysis was completed for experiments with stall times long enough for PON to reach a plateau level. The solid lines represent curves fit to a single-order reaction scheme, y = Cexp(−kt). The first bin for F1 after an OFF event (left gray dot) was unusually small, probably because the dwell time is sometimes too short to be recognized as a pause at 200 nM ATP. Such events would be counted as ON events, decreasing the number of short dwells observed. Therefore, the first bin in the histogram for OFF events was omitted from the fit. The rate constants were determined to be 2.9 s–1 (yellow), 1.4 s–1 (green), and 2.0 s–1 (red). (ce) Angle dependence of konATP, koffATP, and KdATP plotted against θγ. Zero degrees corresponds to the ATP binding angle in Fig. 1a. Red and gray symbols represent the values for F1glyloop (red), determined from Fig. 4a, and F1WT (gray), taken from Watanabe et al. (32). Open symbols in (c) represent the kon of free rotation. To see this figure in color, go online.

Figure 5.

Figure 5

Angle dependence of ATP hydrolysis by F1glyloop/E190D. (a) Time courses for PON of F1glyloop/E190D at 1 mM ATP after stalling at θb = 0° (black), +15° (green), +45° (blue), and +75° (red) from the original catalytic angle. Time courses were fit with a function shown in Materials and Methods. The gray line represents the time course of free rotation. Each data point was obtained from 14–69 trials using 5–13 molecules. The error in PON is given by PON(100PON)/N, where N is the number of trials for each stall measurement. Cyan represents PON in the presence of 10 mM Pi after stalling at θb = +45°. (b) Zoom-up of the plotted traces in (a). (c) Histograms for the ATP hydrolysis dwell of F1glyloop/E190D in free rotation (yellow) and for F1 after an OFF (green) or ON (red) event. Each analysis was conducted for experiments with stall times long enough for PON to reach a plateau. The solid lines represent curves fit to a single-order reaction scheme, y = Cexp(−kt). The rate constants were determined to be 2.9 s–1 (yellow), 1.4 s–1 (green), and 2.0 s–1 (red). (df) Angle dependence of khydATP, ksynATP, and KEATP plotted against θγ. Here, we defined 0° as the angle for ATP hydrolysis in Fig. 1a. Blue and gray symbols represent the values for F1glyloop/E190D (blue), determined from (a), and F1E190D (gray), taken from Watanabe et al. (32). To see this figure in color, go online.

Figure 2.

Figure 2

Torque and rotary potential. (a) The fluctuation theorem was employed for torque measurement of F1WT, F1glyloop, and F1glyloop/E190D. The plot shows ln[Pθ)/P(−Δθ)] versus Δθ/kBT. The slope represents the rotary torque generated by F1. The average torque was determined from a linear approximation of all data points (solid lines). The data for F1glyloop and F1WT are from Tanigawara et al. (50). (b) The rotary torques (N) generated by F1glyloop, F1glyloop/E190D, and F1WT are 24, 18, and 41 pN·nm, respectively. (c) Rotary potential of the F1. Probability densities of angular positions during pauses from the five molecules, i.e., ATP binding pause for F1glyloop and hydrolysis pause for F1glyloop/E190D, were transformed into rotary potentials according to Boltzmann’s law: F1glyloop (red), and F1glyloop/βE190D (blue). The determined potentials were fit to the harmonic function ΔG = 1/2 × κtotal × θb2, where κtotal is the torsion stiffness. Determined stiffness values were 40 and 44 pN·nm for F1glyloop and F1glyloop/E190D, respectively. To see this figure in color, go online.

Simulation of PON for F1glyloop/E190D

The kinetic scheme of stalled F1glyloop/E190D is given as

βE190D×ATPksynATPkhydATPβE190D×(ADP+Pi)koffPiβE190D×ADP, (1)

where koffPi is the rate constant for Pi release from β immediately after hydrolysis. Because F1 generates torque after ATP hydrolysis, PON is represented as the sum of probabilities of [βE190D × (ADP+Pi)] and [βE190D × ADP]. Thus, PON is represented as a function of time, t, by

PON=[βE190D×(ADP+Pi)]+[βE190D×ADP][βE190D]total=100+ε50{(ηε)eεtεη}e12(δ+ε)t (2)
δ=khydATP+ksynATP+koffPi
ε=(khydATP+ksynATP+koffPi)24khydATPkoffPi
η=khydATPksynATPkoffPi.

By fitting to the curves of PON, the fitting parameters khydATP and ksynATP were determined (see Fig. 5, d and e).

Angular position of γ

Fig. 2 c shows that the total stiffness of the system (κtotal) was determined to be 55, 40, and 44 pN·nm for wild-type F1 (F1WT), F1glyloop, and F1glyloop/E190D, respectively. By modeling the system as a composition of two elastic parts, internal and external, connected in series, the total stiffness (κtotal) can be expressed as

1/κtotal=1/κinternal+1/κexternal, (3)

where κinternal and κexternal represent the stiffness of the internal and external parts, respectively. Because glycine substitutions in the DELSEED loop do not affect the external part, κexternal is assumed to be the same as for the wild-type, 72.6 pN·nm (52). From Eq. 3, κinternal is thus determined to be 223, 90, and 113 pN·nm for F1WT, F1glyloop, and F1glyloop/E190D, respectively. Because the internal and external parts are connected in series, the angular position of γ (θγ) can be calculated by

θγ=κexternal/(κinternal+κexternal)×θb, (4)

where θb is the angular position of the bead. Based on Eq. 4, we determined the angular position of γ from the bead position.

Results

Rotation of DELSEED-loop mutants

In this work, we used three F1 complexes derived from thermophilic Bacillus PS3 (TF1), wild-type F1 (F1WT), F1glyloop, and F1glyloop/E190D, for stall-and-release experiments. F1glyloop and F1glyloop/E190D contain glycine substitutions at all residues constituting the DELSEED loop of the β subunit (β384–394 in TF1 numbering) to disrupt the specific interaction with the γ subunit and the conformational rigidity of the loop (50) (Fig. 1 b). In F1glyloop/E190D, the βE190D mutation was introduced to slow ATP hydrolysis and extend the hydrolysis waiting dwell to facilitate the stall-and-release experiment (32). Because F1glyloop/E190D was not previously characterized, we conducted a rotation assay for F1glyloop/E190D (Fig. 1 c). Magnetic beads were attached as rotation markers to the γ subunit of F1 immobilized on a glass surface coated with Ni-NTA. Similar to F1E190D, F1glyloop/E190D exhibited 120° steps separated by long pauses for slowed hydrolysis at saturating ATP (21). Figs. 1 d and S1 show the Michaelis-Menten curves and the Lineweaver-Burk plots for the rotation of F1glyloop/E190D (obtained herein), F1WT, and F1glyloop (from our previous reports) (50) for comparison of kinetic parameters. The maximum rotation rate (Vmax = 0.43 s−1) and Michaelis constant (Km = 1.0 μM) for F1glyloop/E190D were comparable to those of F1E190D (21). Thus, the rate constant for ATP binding of F1glyloop/E190D was determined to be 1.2 × 106 M−1·s−1 (3 × Vmax/Km).

The rotary torque of F1glyloop/E190D was measured using the fluctuation theorem (15), which estimates torque as entropy generation from the time series of the centroid of a rotation probe (Fig. 2, a and b). From the time trajectories of the 120° rotation steps, the time domains of the steps were extracted. We calculated the ratio between the forward and backward movement probabilities, [Pθ)/P(−Δθ)], over a set period and determined the torque from the slope of this ratio versus Δθ/kBT (Fig. 2 a). The results showed that F1glyloop/E190D generates 18 pN·nm torque (Fig. 2 b), which is less than that generated by F1WT (41 pN·nm) and comparable to that generated by F1glyloop (24 pN·nm). This indicates that glycine substitutions in the DELSEED loop dominantly affect torque generation for F1glyloop/E190D, whereas the kinetics of hydrolysis is mainly determined by the βE190D mutation.

We also examined the mutational effects on the rotary potential during the reaction waiting pauses. The probability densities of bead position (θb) during pauses were measured, and the rotary potentials were determined according to Boltzmann’s law (Fig. 2 c). The determined potentials fit well to a harmonic function, ΔG =1/2 × κtotal × θb2, where κtotal is the torsion stiffness. The determined stiffness values were 55, 40, and 44 pN·nm for F1WT, F1glyloop, and F1glyloop/E190D, respectively. The lower stiffness of the mutants suggested that their rotary potentials became more gradual than that of F1WT.

Manipulation of single F1 rotation

To investigate the impact of glycine substitutions in the DELSEED loop on the angle-dependent affinity change of ATP binding and equilibrium change of ATP hydrolysis, we conducted stall-and-release experiments using F1glyloop to measure ATP binding and F1glyloop/E190D to measure ATP hydrolysis. Experimental procedures were essentially the same as for previous stall-and-release experiments (32) (Fig. 3 a). Briefly, we turned on the magnetic tweezers to arrest F1 at the target angle when F1 paused for ATP binding or hydrolysis (Fig. 3 a). At the designated time, we turned off the magnetic tweezers to release F1 from arrest. After release, F1 showed one of two behaviors, taking a forward 120° step (on event) or resuming pause at the original pause angle (off event) (Fig. 3, a and b). The on event indicates that F1 completed the waiting reaction and exerted a torque on the magnetic beads (Fig. 3 b, left); conversely, the off event shows that the waiting reaction has not occurred because F1 cannot generate torque without completing the waiting reaction (Fig. 3 b, right). We conducted the stalling experiment primarily in the angle range of ±50°. The subsequent sections discuss analysis of the probability of on events against total trials, PON.

Angle dependence of ATP binding

Stall-and-release experiments were conducted for F1glyloop at 200 nM ATP, where the ATP waiting dwell is 0.34 s. Fig. 4 a shows a plot of PON versus stall time. PON increased with both the stall angle and stall time, which is similar to our previous finding for F1WT. Furthermore, PON was not always 100% saturated; it converged to a certain value, e.g., 60% for a +30° stall (Fig. 4 a, green line). These observations imply that ATP binding is reversible and that ATP release occurs during stalling. Accordingly, the plateau level indicates the equilibrium between ATP binding and release. To confirm the reversibility, we analyzed the dwell time for F1 to spontaneously conduct a 120° step after an off event (Fig. 3 b, blue). To avoid including data from before the equilibrium, only experiments with longer stalling times in which PON achieved a plateau were analyzed. The dwell-time histogram exhibited single-exponential decay, allowing determination of the rate constant (Fig. 4 b). As expected, the determined rate constant corresponded to that of free rotation (Fig. 4 b). This correspondence discounted the possibility that unexpected inactivation during stalling competes with ATP binding. We also plotted a histogram of the dwell time for the second 120° step after an on event (Fig. 3 b, red). The determined rate constant was also in good agreement with that of free rotation (Fig. 4 b), confirming that manipulation did not alter the kinetic properties of F1glyloop.

By fitting time courses of PON based on the reversible reaction scheme F1glyloop + ATP ⇄ F1glyloop·ATP, the rate constants for ATP binding and release, konATP and koffATP, respectively, were determined for each stall angle of magnetic beads (θb). The actual γ position (θγ) deviates from the bead position (θb) because the system has elastic components: the outwardly protruding domain of γ, streptavidin, and the α3β3 stator ring, which reduces the torsional stress by twisting itself. We corrected the data with the apparent rotary stiffness of the mutant F1 complexes (see Materials and Methods). The corrected data for the angle dependence of ATP binding are shown in Fig. 4, c and d. konATP for F1glyloop increased exponentially with the stall angle by ∼2.8-fold per 20°, similar to previously reported corresponding increases for the wild-type (32). In contrast, koffATP for F1glyloop was almost constant, which is markedly different from the ∼2.6-fold decrease per 20° for the wild-type. Therefore, the dissociation constant for ATP, KdATP, decreased 2.8-fold per 20° for F1glyloop, which is significantly smaller than the decrease for wild-type, 6.3-fold per 20° (Fig. 4 e), suggesting that glycine substitution of the DELSEED loop impairs torque transmission during the change in affinity between F1 and bound ATP.

Angle dependence of hydrolysis

Stall-and-release experiments to measure hydrolysis were conducted for F1glyloop/E190D, because the hydrolysis waiting dwell for F1glyloop is only 1 ms, which is too short for the stalling experiment. Under saturating ATP conditions, F1glyloop/E190D showed a 120° stepping rotation with hydrolysis pauses of 1.1 s (Fig. 5 c, yellow). During the 1.1 s hydrolysis waiting dwells, we stalled F1glyloop/E190D with magnetic tweezers to determine PON. As shown previously (26), PON increased in two phases (Fig. 5, a and b). After a rapid increase, PON gradually approached 100%. The first increase was nearly complete within 1 s, which is consistent with the time constant of ATP hydrolysis (Fig. 5 c, yellow). The subsequent increase was extremely slow for an effective catalytic reaction. We previously (26) showed that this slow increase results from the release of inorganic phosphate (Pi) by the β subunit that hydrolyzed the bound ATP (Fig. 1 a, 200° state) during stalling. This slow Pi release is suppressed by 10 mM Pi, which is also consistent with the previous study (26). We fit the time courses of PON using a consecutive reaction model with a reversible hydrolysis step followed by irreversible Pi release at 200°: F1glyloop/E190D × ATP ⇄ F1glyloop/E190D × ADP × Pi → F1glyloop/E190D × ADP + Pi (see Materials and Methods). The rate constants for hydrolysis and synthesis, khydATP and ksynATP, respectively, were determined from the fitted data, and the data points were corrected based on the rotary stiffness, as for the angle dependence of ATP binding. The khydATP for F1glyloop/E190D increased exponentially with the stall angle (Fig. 5 d), whereas ksynATP remained constant (Fig. 5 e). These features follow essentially the same trend observed in our previous stall-and-release experiment with F1E190D (32). Hence, the equilibrium constant for hydrolysis (khydATP/ksynATP = KEHyd-ATP) showed angle dependence similar to F1E190D (Fig. 5 f). The koffPi for F1glyloop/E190D at 200° was constant, ∼0.07 s−1 (data not shown), and almost the same as that of F1E190D (0.04 s−1). The similarity of these values suggests that the DELSEED loop is not involved in torque transmission during the hydrolysis step.

Discussion

Our results indicate that glycine substitution of the DELSEED loop impairs torque transmission during the power stroke of β by affecting the affinity change to bound ATP. This is consistent with the structural features of bovine mitochondrial F1 (bMF1) suggesting that β undergoes a conformational transition from an open to a closed conformation upon ATP binding to push the γ rotor with its C-terminal DELSEED loop (8). Thus, the crystal structures suggest that the DELSEED loop is predominantly involved in torque transmission upon ATP binding. From this perspective, it is surprising that F1glyloop has any capacity for angle-dependent modulation of konATP. The energy released by the affinity change, estimated from /θ{kBT×ln[KE(θ)]}kBT, was 65% of that for F1WT. This implies that angle-dependent affinity change is mediated not only by the DELSEED-γ interaction but also by other contact regions, such as the lower interface between the β and γ subunits. Remote communication among β subunits via the neighboring α subunits could also support the robust mechanism of the angle-dependent affinity change. The recent finding of allostery between β subunits in the isolated α3β3 ring supports this model (53). On the other hand, the glycine substitution did not affect the angle-dependent equilibrium change between hydrolysis and synthesis of ATP (the change is slightly enhanced compared to the wild-type), suggesting that the DELSEED loop is not responsible for torque transmission upon hydrolysis. This agrees with observations from the crystal structure indicating that hydrolysis accompanies closure of the αβ interface, which does not directly involve the DELSEED loop (41). Thus, this study shows that the principal role of the DELSEED loop is torque transmission upon the affinity change of ATP coupled with the transition between the open and closed conformations of β, and not torque transmission upon hydrolysis. For further confirmation, we conducted a rotation assay for a hybrid F1 carrying a single copy of βglyloop/E190D. Fluctuation theorem analysis showed that torque generation upon affinity change of ATP was reduced by 42% compared with the wild-type, whereas torque generation for ATP hydrolysis was unchanged (You et al., unpublished data), supporting the results of this study.

The remarkable change in angle dependence upon glycine substitution of the DELSEED loop appears only for koffATP, and the change of konATP is very moderate (Fig. 4 e). One possible interpretation of the asymmetric effect is that the torque transmission upon ATP binding is composed of two processes; DELSEED loop is involved in only one of them, which presumably modulates koffATP, and the lower β-γ interface compensates for the other. The stall-and-release experiment using γ-truncated mutant F1, which eliminates the interaction via the lower β-γ interface (47), will provide detailed insights into the torque transmission mechanism upon ATP binding.

In this study, the impact of glycine substitution on torque transmission upon ATP binding and hydrolysis was studied using the stall-and-release experiment, which provides insights into the torque transmission mechanism of F1 at a resolution of elementary reaction steps, e.g., ATP binding and hydrolysis. However, that for product release (ADP release and Pi release) remained elusive due to technical difficulties. Considering the conformational change in the β subunit depending on the presence of nucleotide (53,54), it is highly probable that the DELSEED loop is also responsible for torque transmission upon ADP release. In contrast, some theoretical studies propose that Pi release is coupled with the opening of the αβ interface (45,55), which suggests that torque transmission for this step would not be affected by the glycine substitution of the DELSEED loop. To comprehensively understand the torque transmission mechanism of F1, we hope to evaluate the impact of glycine substitution on torque transmission upon product release using the stall-and-release experiment.

Author Contributions

R.W. and K.K. designed and performed experiments and analyzed data; H.Y. and M.T. provided technical support; H.N. designed experiments and wrote the paper with R.W.

Acknowledgments

The authors thank all members of the Noji Laboratory for technical support. R.W. and K.K. designed and performed experiments and analyzed data; H.Y. and M.T. provided technical support; H.N. designed experiments and wrote the article with R.W.

This work was supported by Grants-in-Aid for Scientific Research to H.N. (No. 18074005) and R.W. (No. 30540108) from the Japan Ministry of Education, Culture, Sports, Science, and Technology.

Footnotes

Rikiya Watanabe and Kazuma Koyasu contributed equally to this work.

Supporting Material

Document S1. One figure
mmc1.pdf (191.4KB, pdf)
Document S2. Article plus Supporting Material
mmc2.pdf (1.2MB, pdf)

References

  • 1.Noji H., Yasuda R., Kinosita K., Jr. Direct observation of the rotation of F1-ATPase. Nature. 1997;386:299–302. doi: 10.1038/386299a0. [DOI] [PubMed] [Google Scholar]
  • 2.Boyer P.D. A research journey with ATP synthase. J. Biol. Chem. 2002;277:39045–39061. doi: 10.1074/jbc.X200001200. [DOI] [PubMed] [Google Scholar]
  • 3.Kinosita K., Jr., Adachi K., Itoh H. Rotation of F1-ATPase: how an ATP-driven molecular machine may work. Annu. Rev. Biophys. Biomol. Struct. 2004;33:245–268. doi: 10.1146/annurev.biophys.33.110502.132716. [DOI] [PubMed] [Google Scholar]
  • 4.Nakamoto R.K., Baylis Scanlon J.A., Al-Shawi M.K. The rotary mechanism of the ATP synthase. Arch. Biochem. Biophys. 2008;476:43–50. doi: 10.1016/j.abb.2008.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Junge W., Sielaff H., Engelbrecht S. Torque generation and elastic power transmission in the rotary FOF1-ATPase. Nature. 2009;459:364–370. doi: 10.1038/nature08145. [DOI] [PubMed] [Google Scholar]
  • 6.Watanabe R., Tabata K.V., Noji H. Biased Brownian stepping rotation of FOF1-ATP synthase driven by proton motive force. Nat. Commun. 2013;4:1631. doi: 10.1038/ncomms2631. [DOI] [PubMed] [Google Scholar]
  • 7.Diez M., Zimmermann B., Gräber P. Proton-powered subunit rotation in single membrane-bound FOF1-ATP synthase. Nat. Struct. Mol. Biol. 2004;11:135–141. doi: 10.1038/nsmb718. [DOI] [PubMed] [Google Scholar]
  • 8.Abrahams J.P., Leslie A.G., Walker J.E. Structure at 2.8 Å resolution of F1-ATPase from bovine heart mitochondria. Nature. 1994;370:621–628. doi: 10.1038/370621a0. [DOI] [PubMed] [Google Scholar]
  • 9.Cingolani G., Duncan T.M. Structure of the ATP synthase catalytic complex (F1) from Escherichia coli in an autoinhibited conformation. Nat. Struct. Mol. Biol. 2011;18:701–707. doi: 10.1038/nsmb.2058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Kabaleeswaran V., Puri N., Mueller D.M. Novel features of the rotary catalytic mechanism revealed in the structure of yeast F1 ATPase. EMBO J. 2006;25:5433–5442. doi: 10.1038/sj.emboj.7601410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Okuno D., Iino R., Noji H. Rotation and structure of FOF1-ATP synthase. J. Biochem. 2011;149:655–664. doi: 10.1093/jb/mvr049. [DOI] [PubMed] [Google Scholar]
  • 12.Futai M., Nakanishi-Matsui M., Nakamoto R.K. Rotational catalysis in proton pumping ATPases: from E. coli F-ATPase to mammalian V-ATPase. Biochim. Biophys. Acta. 2012;1817:1711–1721. doi: 10.1016/j.bbabio.2012.03.015. [DOI] [PubMed] [Google Scholar]
  • 13.Hornung T., Martin J., Frasch W.D. Microsecond resolution of single-molecule rotation catalyzed by molecular motors. Methods Mol. Biol. 2011;778:273–289. doi: 10.1007/978-1-61779-261-8_18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Yasuda R., Noji H., Yoshida M. F1-ATPase is a highly efficient molecular motor that rotates with discrete 120° steps. Cell. 1998;93:1117–1124. doi: 10.1016/s0092-8674(00)81456-7. [DOI] [PubMed] [Google Scholar]
  • 15.Hayashi K., Ueno H., Noji H. Fluctuation theorem applied to F1-ATPase. Phys. Rev. Lett. 2010;104:218103. doi: 10.1103/PhysRevLett.104.218103. [DOI] [PubMed] [Google Scholar]
  • 16.Cherepanov D.A., Junge W. Viscoelastic dynamics of actin filaments coupled to rotary F-ATPase: curvature as an indicator of the torque. Biophys. J. 2001;81:1234–1244. doi: 10.1016/S0006-3495(01)75781-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Spetzler D., Ishmukhametov R., Frasch W.D. Single molecule measurements of F1-ATPase reveal an interdependence between the power stroke and the dwell duration. Biochemistry. 2009;48:7979–7985. doi: 10.1021/bi9008215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Bilyard T., Nakanishi-Matsui M., Berry R.M. High-resolution single-molecule characterization of the enzymatic states in Escherichia coli F1-ATPase. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2013;368:20120023. doi: 10.1098/rstb.2012.0023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Toyabe S., Watanabe-Nakayama T., Muneyuki E. Thermodynamic efficiency and mechanochemical coupling of F1-ATPase. Proc. Natl. Acad. Sci. USA. 2011;108:17951–17956. doi: 10.1073/pnas.1106787108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Yasuda R., Noji H., Itoh H. Resolution of distinct rotational substeps by submillisecond kinetic analysis of F1-ATPase. Nature. 2001;410:898–904. doi: 10.1038/35073513. [DOI] [PubMed] [Google Scholar]
  • 21.Shimabukuro K., Yasuda R., Yoshida M. Catalysis and rotation of F1 motor: cleavage of ATP at the catalytic site occurs in 1 ms before 40° substep rotation. Proc. Natl. Acad. Sci. USA. 2003;100:14731–14736. doi: 10.1073/pnas.2434983100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Suzuki T., Tanaka K., Yoshida M. Chemomechanical coupling of human mitochondrial F1-ATPase motor. Nat. Chem. Biol. 2014;10:930–936. doi: 10.1038/nchembio.1635. [DOI] [PubMed] [Google Scholar]
  • 23.Adachi K., Oiwa K., Kinosita K., Jr. Coupling of rotation and catalysis in F1-ATPase revealed by single-molecule imaging and manipulation. Cell. 2007;130:309–321. doi: 10.1016/j.cell.2007.05.020. [DOI] [PubMed] [Google Scholar]
  • 24.Adachi K., Oiwa K., Kinosita K., Jr. Controlled rotation of the F₁-ATPase reveals differential and continuous binding changes for ATP synthesis. Nat. Commun. 2012;3:1022. doi: 10.1038/ncomms2026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Martin J.L., Ishmukhametov R., Frasch W.D. Anatomy of F1-ATPase powered rotation. Proc. Natl. Acad. Sci. USA. 2014;111:3715–3720. doi: 10.1073/pnas.1317784111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Watanabe R., Iino R., Noji H. Phosphate release in F1-ATPase catalytic cycle follows ADP release. Nat. Chem. Biol. 2010;6:814–820. doi: 10.1038/nchembio.443. [DOI] [PubMed] [Google Scholar]
  • 27.Watanabe R., Noji H. Timing of inorganic phosphate release modulates the catalytic activity of ATP-driven rotary motor protein. Nat. Commun. 2014;5:3486. doi: 10.1038/ncomms4486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Watanabe R., Iino R., Noji H. Temperature-sensitive reaction intermediate of F1-ATPase. EMBO Rep. 2008;9:84–90. doi: 10.1038/sj.embor.7401135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Ariga T., Muneyuki E., Yoshida M. F1-ATPase rotates by an asymmetric, sequential mechanism using all three catalytic subunits. Nat. Struct. Mol. Biol. 2007;14:841–846. doi: 10.1038/nsmb1296. [DOI] [PubMed] [Google Scholar]
  • 30.Shimo-Kon R., Muneyuki E., Kinosita K., Jr. Chemo-mechanical coupling in F1-ATPase revealed by catalytic site occupancy during catalysis. Biophys. J. 2010;98:1227–1236. doi: 10.1016/j.bpj.2009.11.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Rees D.M., Montgomery M.G., Walker J.E. 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;109:11139–11143. doi: 10.1073/pnas.1207587109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Watanabe R., Okuno D., Noji H. Mechanical modulation of catalytic power on F1-ATPase. Nat. Chem. Biol. 2012;8:86–92. doi: 10.1038/nchembio.715. [DOI] [PubMed] [Google Scholar]
  • 33.Watanabe R., Noji H. Characterization of the temperature-sensitive reaction of F1-ATPase by using single-molecule manipulation. Sci. Rep. 2014;4:4962. doi: 10.1038/srep04962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Watanabe R., Noji H. Chemomechanical coupling mechanism of F1-ATPase: catalysis and torque generation. FEBS Lett. 2013;587:1030–1035. doi: 10.1016/j.febslet.2013.01.063. [DOI] [PubMed] [Google Scholar]
  • 35.Bowler M.W., Montgomery M.G., Walker J.E. Ground state structure of F1-ATPase from bovine heart mitochondria at 1.9 Å resolution. J. Biol. Chem. 2007;282:14238–14242. doi: 10.1074/jbc.M700203200. [DOI] [PubMed] [Google Scholar]
  • 36.Kobayashi M., Akutsu H., Yagi H. Analysis of the open and closed conformations of the β subunits in thermophilic F1-ATPase by solution NMR. J. Mol. Biol. 2010;398:189–199. doi: 10.1016/j.jmb.2010.03.013. [DOI] [PubMed] [Google Scholar]
  • 37.Masaike T., Koyama-Horibe F., Nishizaka T. Cooperative three-step motions in catalytic subunits of F1-ATPase correlate with 80° and 40° substep rotations. Nat. Struct. Mol. Biol. 2008;15:1326–1333. doi: 10.1038/nsmb.1510. [DOI] [PubMed] [Google Scholar]
  • 38.Turina P., Aggeler R., Capaldi R.A. The cysteine introduced into the α subunit of the Escherichia coli F1-ATPase by the mutation α R376C is near the α-β subunit interface and close to a noncatalytic nucleotide binding site. J. Biol. Chem. 1993;268:6978–6984. [PubMed] [Google Scholar]
  • 39.Komoriya Y., Ariga T., Noji H. Principal role of the arginine finger in rotary catalysis of F1-ATPase. J. Biol. Chem. 2012;287:15134–15142. doi: 10.1074/jbc.M111.328153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Dittrich M., Hayashi S., Schulten K. ATP hydrolysis in the βTP and βDP catalytic sites of F1-ATPase. Biophys. J. 2004;87:2954–2967. doi: 10.1529/biophysj.104.046128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Kagawa R., Montgomery M.G., Walker J.E. The structure of bovine F1-ATPase inhibited by ADP and beryllium fluoride. EMBO J. 2004;23:2734–2744. doi: 10.1038/sj.emboj.7600293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Okuno D., Fujisawa R., Noji H. Correlation between the conformational states of F1-ATPase as determined from its crystal structure and single-molecule rotation. Proc. Natl. Acad. Sci. USA. 2008;105:20722–20727. doi: 10.1073/pnas.0805828106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Sielaff H., Rennekamp H., Junge W. Functional halt positions of rotary FOF1-ATPase correlated with crystal structures. Biophys. J. 2008;95:4979–4987. doi: 10.1529/biophysj.108.139782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Hayashi S., Ueno H., Noji H. Molecular mechanism of ATP hydrolysis in F1-ATPase revealed by molecular simulations and single-molecule observations. J. Am. Chem. Soc. 2012;134:8447–8454. doi: 10.1021/ja211027m. [DOI] [PubMed] [Google Scholar]
  • 45.Okazaki K., Hummer G. Phosphate release coupled to rotary motion of F1-ATPase. Proc. Natl. Acad. Sci. USA. 2013;110:16468–16473. doi: 10.1073/pnas.1305497110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.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;133:3372–3380. doi: 10.1021/ja1070152. [DOI] [PubMed] [Google Scholar]
  • 47.Kohori A., Chiwata R., Kinosita K., Jr. Torque generation in F1-ATPase devoid of the entire amino-terminal helix of the rotor that fills half of the stator orifice. Biophys. J. 2011;101:188–195. doi: 10.1016/j.bpj.2011.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Furuike S., Hossain M.D., Kinosita K., Jr. Axle-less F1-ATPase rotates in the correct direction. Science. 2008;319:955–958. doi: 10.1126/science.1151343. [DOI] [PubMed] [Google Scholar]
  • 49.Usukura E., Suzuki T., Yoshida M. Torque generation and utilization in motor enzyme FOF1-ATP synthase: half-torque F1 with short-sized pushrod helix and reduced ATP synthesis by half-torque F0F1. J. Biol. Chem. 2012;287:1884–1891. doi: 10.1074/jbc.M111.305938. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Tanigawara M., Tabata K.V., Noji H. Role of the DELSEED loop in torque transmission of F1-ATPase. Biophys. J. 2012;103:970–978. doi: 10.1016/j.bpj.2012.06.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hirono-Hara Y., Ishizuka K., Noji H. Activation of pausing F1 motor by external force. Proc. Natl. Acad. Sci. USA. 2005;102:4288–4293. doi: 10.1073/pnas.0406486102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Okuno D., Iino R., Noji H. Stiffness of γ subunit of F1-ATPase. Eur. Biophys. J. 2010;39:1589–1596. doi: 10.1007/s00249-010-0616-9. [DOI] [PubMed] [Google Scholar]
  • 53.Uchihashi T., Iino R., Noji H. High-speed atomic force microscopy reveals rotary catalysis of rotorless F₁-ATPase. Science. 2011;333:755–758. doi: 10.1126/science.1205510. [DOI] [PubMed] [Google Scholar]
  • 54.Tsunoda S.P., Muneyuki E., Noji H. Cross-linking of two β subunits in the closed conformation in F1-ATPase. J. Biol. Chem. 1999;274:5701–5706. doi: 10.1074/jbc.274.9.5701. [DOI] [PubMed] [Google Scholar]
  • 55.Ito Y., Yoshidome T., Ikeguchi M. Molecular dynamics simulations of yeast F1-ATPase before and after 16° rotation of the γ subunit. J. Phys. Chem. B. 2013;117:3298–3307. doi: 10.1021/jp312499u. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Document S1. One figure
mmc1.pdf (191.4KB, pdf)
Document S2. Article plus Supporting Material
mmc2.pdf (1.2MB, pdf)

Articles from Biophysical Journal are provided here courtesy of The Biophysical Society

RESOURCES