Abstract
The synthesis of ATP, the key reaction of biological energy metabolism, is accomplished by the rotary motor protein; FoF1-ATP synthase (FoF1). In vivo, FoF1, located on the cell membrane, carries out ATP synthesis by using the proton motive force. This heterologous energy conversion is supposed to be mediated by the mechanical rotation of FoF1; however, it still remained unclear. Recently, we developed the novel experimental setup to reproduce the proton motive force in vitro and succeeded in directly observing the proton-driven rotation of FoF1. In this review, we describe the interesting working principles determined so far for FoF1 and then introduce results from our recent study.
Keywords: ATP synthase, molecular motor, heterogeneous energy conversion
FoF1 ATP synthase (FoF1) is a molecular energy-converter which catalyzes physiologically important synthesis of ATP from ADP and inorganic phosphate (Pi) by using the electrochemical energy in proton gradient; the proton motive force (pmf) across bio-membranes1–3. The prominent feature of FoF1 is that mechanical rotation of the inner rotor complex mediates the aforementioned heterologous energy conversion with high efficiency and reversibility4,5, which is not found in other biological systems, and therefore, FoF1 attracts great interest from many researchers across a wide range of research fields. FoF1 comprises 2 rotary motors, F1 and Fo (Fig. 1a). F1 (α3β3γδɛ), a water-soluble part of FoF1, is an ATP-driven rotary motor, which couples ATP hydrolysis or synthesis at α3β3δ stator ring to the mechanical rotation of γɛ complex in counterclockwise direction (Fig. 1b)6. Each αβ interface possesses a catalytic site for ATP hydrolysis7; F1 hydrolyzes 3 ATP molecules per turn8. Extensive studies have been done to understand the chemo-mechanical coupling mechanism of F19–18, and therefore, currently, F1 is one of the best characterized molecular motor proteins.
Figure 1.
Structural model of bacterial FoF1-ATP synthase
(a) The space-filling model of FoF1 (side view) is assembled from several partial structures obtained by X-ray crystallography and NMR (PDB codes 1C17, 1B9U, 2KHK, 1L2P, 3UDO, 3OAA, and 1ABV). FoF1 has subunit stoichiometry of α3β3 γδεab2c8–15. FoF1 is composed of 2 rotary motors; a membrane-embedded Fo (ab2c8–15) and a water-soluble F1 (α3β3γδε). Proton flows through the channels located at the a–c interface of Fo, while ATP synthesis is carried out at the catalytic sites located at the α–β interface of F1. (b) Top view of Fo (left) and F1 (right).
Fo, a membrane-embedded part of FoF1, is the proton-driven rotary motor, which couples proton translocation to mechanical rotation of the oligomer ring of the c-subunit against the ab2 complex (Fig. 1b)5,19,20. Each c-subunit has a proton-binding site and mediates proton translocation: 1 proton per c-subunit per rotation. Thus, the number of c-subunits in the c-ring is thought to determine the total number of protons translocated per rotation. While the number of c-subunits varies among species from 8 to 1521,22, the bacterial Fo, such as E-coli and thermophilic Bacillus PS3, has 10 c-subunits that form a c10-ring 23,24. It is difficult to handle Fo since it is embedded in the membrane, and therefore, the rotary mechanism of Fo remains elusive.
In a cell, F1 and Fo are connected via central and peripheral stalks (Fig. 1a), which allow the torque transmission between the 2 motors. Under physiological conditions, Fo generates a larger torque than F1 and reverses the rotary direction of F1, thereby inducing the reverse reaction of ATP hydrolysis, i.e. ATP synthesis (Fig. 2a). In contrast, when pmf diminishes, F1 hydrolyzes ATP and reverses the rotary direction of Fo, thereby enforcing pumping of protons by Fo in order to generate pmf (Fig. 2b). Thus, FoF1 manifests reversibility with regard to the process of energy conversion, and the rotation of inner rotor complex plays an important role in this reversibility.
Figure 2.
Heterologous energy conversion by FoF1
Energy conversion mechanism of FoF1 in ATP synthesis (a) or hydrolysis (b) conditions. Orange and light green represent the rotor and stator subcomplex, respectively.
To understand the precise role of the rotation on the energy conversion of FoF1, some single-molecule studies have been carried out5,19,20,25–27. Due to ease of handling, ATP-driven rotation of detergent-solubilized FoF1, which is no longer part of a bio-membrane, has been studied; however, pmf-driven rotation has not been studied sufficiently. In this year, we developed a novel experimental setup that for the first time allowed direct observation of the proton-driven rotation of FoF1 20. In this review, we focus on the role of rotary motion in the energy conversion of FoF1, and introduce the fundamental working principles determined so far, including the results from our recent study20.
Energy conversion mechanism
As mentioned above, ATP synthesis/hydrolysis reaction is reversibly coupled with proton translocation across the membrane. To understand the reversible energy conversion mechanism of FoF1, extensive biochemical studies have been performed. Since it is difficult to measure the time course of pmf generated by FoF1, most studies pertained to the measurement of ATP synthesis activity under a given pmf. In general, pmf mainly comprises 2 components: the trans-membrane proton gradient (ΔpH) and the potential difference (Δψ). At first, biochemical studies examined ATP synthesis activity by changing the amplitude of each of these components of pmf, and examined the component-dependence of ATP synthesis activity of FoF1 28–34. In recent studies, FoF1 was purified and reconstituted into liposomes, and the ATP synthesis rate was quantitatively measured by the acid–base transition or valinomycin-mediated K+ diffusion potential for producing ΔpH or Δψ. It was found that ΔpH and Δψ contribute equally to ATP synthesis rate of FoF1, and, moreover, either ΔpH or Δψ alone can drive synthesis of ATP29. Kinetic equivalence implies that FoF1 can use ΔpH or Δψ to drive rotation at an equal efficiency.
Next, to determine the coupling efficiency in FoF1, biochemical studies examined the number of protons translocated per synthesis of 1 molecule of ATP; the H+/ATP ratio, by enzymes from various organisms, e.g., Escherichia coli, yeast, and chloroplasts4,35,36. In view of the rotary catalysis model of FoF1, the H+/ATP ratio should coincide with the ratio of the number of proton-binding c-subunits to the 3 catalytic nucleotide-binding β-subunits when proton-translocation and ATP synthesis are highly coupled. Analyses of the equilibrium point, where the free energy of ATP synthesis is balanced with that of proton translocation, allowed determination of the H+/ATP ratio. The determined H+/ATP ratios were dependent on the stoichiometry of the c and β subunits, although they were not identical to the c/β ratios. In particular, H+/ATP ratios in a recent study were smaller than the c/β ratios36, which implies that proton translocation is stochastically coupled to the synthesis of ATP; however, the coupling efficiency is extremely high.
ATP-driven rotation
To understand the precise role of rotation in the energy conversion, the rotation of FoF1 was observed at single-molecule level. At first, due to the ease of handling, ATP-driven rotation of solubilized FoF1 from E. coli and thermophilic Bacillus PS3 was observed (Fig. 3)25,26,37. In this condition where FoF1 was not a part of membrane, pmf was not imposed on FoF1, and Fo did not generate the rotary torque. Therefore, as same as isolated F1, FoF1 rotated in a counter-clockwise direction, and showed a 120°-stepping rotation at low ATP concentrations, in which ATP binding was the rate-limiting step of rotation. This 120°-stepping rotation reflects the structural symmetry of F1; 3 catalytic sites for ATP hydrolysis/synthesis are located on a single molecule of F17.
Figure 3.
ATP-driven rotation of FoF1
Time course of the ATP-driven rotation of solubilized FoF1 in the presence of 1 mM ATP (left) and 50 nM ATP (right). The insets show the centroid trajectories of the rotating particles. At low ATP concentration, FoF1 showed a 120° stepping rotation.
Next, the rotation of FoF1 reconstituted in a membrane was observed. Ishmukhametov et al. developed an experimental setup with a gold nanorod and phospholipid bilayer nanodisc, which has been shown to provide a good model for a lipid bilayer membrane, and attempted to visualize the rotary motion of membrane-constituted E. coli FoF1 driven by ATP hydrolysis27. In their setup, the gold nanorod was attached to the c-subunit as a rotation probe. The intensity of scattered red light from a nanorod changes in a sinusoidal manner as a function of the rotary position, and therefore, the rotary motion of FoF1 could be visualized from the analysis of red light scattered from the nanorod. On the other hand, the Fo module was buried within the phospholipid bilayer nanodisc, which is large enough to allow incorporation of Fo, but which is on the same scale as the FoF1 complex, and thus, it was difficult to generate a pmf across the nanodisc. Using this experimental setup, the 36° stepping rotation of FoF1 in the presence of a high concentration of polyethylene glycol was observed for the first time. This reflects the structural symmetry of the Fo module; 10 proton binding sites are located on a single molecule of Fo24. In this setup, as mentioned above, pmf was not imposed on FoF1, and Fo did not generate rotary torque; and therefore, it could not be confirmed whether the 36° step was coupled to the translocation of protons.
pmf-driven rotation
Diez et al. developed a method to indirectly visualize the rotation of E. coli FoF1 that had been reconstituted in liposomes by using single molecule Förster resonance energy transfer (sm-FRET)5. In their method, they introduced a pair of FRET probes at a stator and rotor subunit of FoF1 for visualization of the rotary motion of FoF1, and generated the pmf by using the acid–base transition or valinomycin-mediated K+ diffusion potential method, as mentioned above. By using this method, they observed the 120° stepping rotation driven by ATP hydrolysis5, and moreover, for the first time observed the obscure 36º stepping rotation driven by pmf 19. However, due to the low signal-to-noise ratio and fast photobleaching of the fluorescent dyes used in sm-FRET, high resolution tracking and long-term recording of the rotational dynamics of FoF1 has not yet been achieved (recording time < approximately 300 ms). Therefore, the fundamental features of the pmf-driven rotation of FoF1, such as the exact step size, unidirectionality of the rotation, and stochasticity of the steps, has remained elusive to date.
To solve this problem, we recently developed a novel experimental setup that allows long-term direct observation of the pmf-driven rotation of E. coli FoF1 with a high spa-tiotemporal resolution (Fig. 4a)20. In this setup, the FoF1-reconstituted, supported membrane was expanded on a cover-slip covered with Ni-NTA-modified agarose, where FoF1 molecules were anchored via His-tags that had been introduced to the periplasmic side of the c-subunits. The 80-nm gold colloid was attached as a rotation probe onto the β subunits of F1 to allow visualization using a total internal reflection dark-field illumination system, which facilitated the long-term recording of rotation (approximately 10 s) with a high spatiotemporal resolution of about 5 nm and <0.5 ms38. In addition, pmf across the supported lipid bilayer was generated by photolysis of caged protons [1-(2-Nitrophenyl) ethyl sulfate] with a total internal reflection illumination of UV light (λ=404 nm) that selectively acidified the space between the coverslip and the lipid bilayer (the interspace). This novel setup can stably generate ΔpH of 1.8–3.7 for several tens of seconds, while the conventional method, i.e. acid–base transition, can generate ΔpH for only a few seconds. The magnitude of ΔpH upon photolysis of the caged protons was measured using a pH-sensitive fluorescent dye, pHrodo-Red (pHrodo)39, which increased the fluorescent signal upon acidification.
Figure 4.
Proton-driven rotation of FoF1
(a) Schematic model of the novel experimental setup used to directly visualize the proton-driven rotation of FoF1. Evanescent illumination by 532-nm and 404-nm lasers was used. (b) Time course of the proton-driven rotation of FoF1 in the presence of 50 μM ADP, 1 μM ATP, and 200 μM Pi before (blue) and after (red) UV irradiation. The left inset shows an enlarged view of the time course. Orange and light blue points represent the pauses before clockwise and counterclockwise steps, respectively. The right inset shows the centroid trajectory of the rotating particle during UV irradiation. Under these conditions, FoF1 showed the stochastic 120° stepping rotation. (c) Rotational speed plotted against the trans-membrane proton gradient (ΔpH).
By using this experimental setup, we for the first time observed the clockwise rotary motion of FoF1 upon UV irradiation (averaged velocity was 2.7 rps.), while no rotating particles were observed prior to UV irradiation (Fig. 4b). To investigate the correlation between the rotational rate and ΔpH, we measured ΔpH by using the pHrodo located at the region where FoF1 showed rotations. Although FoF1 showed a large variation in velocity, it was still evident that faster rotation occurred at higher ΔpH (Fig. 4c). In addition, the data points of higher velocity at a given ΔpH qualitatively agreed with the aforementioned biochemical measurement of ATP synthesis activity28, showing that the rotation observed in this study was coupled to the proton translocation and ATP synthesis, and vice versa, the rotation of FoF1 can mediate the energy conversion with high efficiency.
We also observed a 120° stepping rotation of FoF1 driven by pmf (Fig. 4b). The step size of rotation, viz., 120°, implies that the rotary potential of F1 with 3-fold symmetry dominated the overall rotary potential of FoF1. In other words, the kinetic bottleneck of the pmf-driven rotation was not proton-translocation in Fo, but a catalytic event(s) on F1, such as ATP release or Pi binding14,40. This result is consistent with a previous sm-FRET measurement of rotation of FoF1 in ATP synthesis condition where a pair of FRET probes were introduced at a stator of Fo and a rotor subunit of F1 5, while small steps that was estimated to be 36° were recorded when FRET probes were introduced into Fo19. In the present study, we immobilized the rotor part of Fo on a coverslip and attached the rotation probe at the stator part of F1. Therefore, the observed rotation reflects the stepping behavior both of F1 and Fo as shown in the other works, in which 36°-steps of ATP-driven rotation of FoF1 were observed in a similar experimental setup27. However, we do not exclude the possibility that the difference in the probe position caused the apparently different step size of the rotation. Another possible reason for the inconsistency in step size is the difference in the components of the pmf; while pmf in the sm-FRET measurements was composed of both ΔpH and Δψ, pmf was essentially composed only of ΔpH in our study. To confirm this, a method for direct observation of the proton-driven rotation of FoF1 by applying Δψ is crucial.
On the other hand, noted that the stepping rotation of FoF1 was highly stochastic; FoF1 showed forward-and-backward (clockwise-and-counterclockwise) steps during rotation (Fig. 4b). This is a prominent feature of the pmf-driven rotation of FoF1 that is not seen in the ATP-driven rotation of F1 or FoF1. Surprisingly, the stepping was also observed in the absence of pmf, suggesting that pmf biased the rotary diffusion of FoF1 to the clockwise direction. To confirm this, we analyzed the pause durations between the 120°-steps. In the absence of pmf, the histograms of the pause duration before clockwise or counterclockwise steps showed single exponential decay. The rate constants of the clockwise and counterclockwise steps were determined to be 65 and 61 s−1, respectively. The equilibrium constant of clockwise rotation was, thus, almost 1. In the presence of pmf, the rate constant of clockwise stepping markedly increased about two-fold, while that of counterclockwise step decreased slightly. Thus, the equilibrium constant in the presence of pmf increased to 2, showing that pmf actually biased step direction. This result also suggests that chemical equilibrium was slightly biased toward ATP synthesis by pmf. The stochastic rotation of FoF1 would represent rotation under physiological conditions, where free energy of ATP synthesis almost balances pmf.
Future prospects
Owing to the progress of single molecule observation techniques, we can directly observe the rotary motion of FoF1 both in ATP hydrolysis and synthesis conditions. In particular, the introduced novel experimental setup that allows us to stably apply, under an optical microscope, pmf to the membrane will push forward to understand the rotary catalysis mechanism of FoF1 in ATP synthesis condition, which had been unclear for long time. This experimental strategy is fundamentally applicable to the study on the dynamics of other membrane proteins driven by electrochemical potential. The most promising experiment is the application of this protocol to transporters and ion channels since there are several caged compounds that release specific ions or chemicals. Such studies would reveal the generality and uniqueness of the finding in single-molecule studies on FoF1.
Acknowledgments
We thank Y. Moriizumi for the experiment on solubilized FoF1, and all members of Noji Laboratory for critical discussion. This work was supported by Grant-in-Aid for Scientific Research (No. 30540108) to R. W. from the Ministry of Education, Culture, Sports, Science and Technology, Japan.
References
- 1.Yoshida M, Muneyuki E, Hisabori T. ATP synthase—a marvellous rotary engine of the cell. Nat Rev Mol Cell Biol. 2001;2:669–677. doi: 10.1038/35089509. [DOI] [PubMed] [Google Scholar]
- 2.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]
- 3.Weber J. Structural biology: Toward the ATP synthase mechanism. Nat Chem Biol. 2010;6:794–795. doi: 10.1038/nchembio.458. [DOI] [PubMed] [Google Scholar]
- 4.Turina P, Samoray D, Gräber P. H+/ATP ratio of proton transport-coupled ATP synthesis and hydrolysis catalysed by CFoF1-liposomes. EMBO J. 2003;22:418–426. doi: 10.1093/emboj/cdg073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Diez M, Zimmermann B, Börsch M, König M, Schweinberger E, Steigmiller S, Reuter R, Felekyan S, Kudryavtsev V, Seidel CA, 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]
- 6.Noji H, Yasuda R, Yoshida M, Kinosita K., Jr Direct observation of the rotation of F1-ATPase. Nature. 1997;386:299–302. doi: 10.1038/386299a0. [DOI] [PubMed] [Google Scholar]
- 7.Abrahams JP, Leslie AG, Lutter R, Walker JE. 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]
- 8.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;93:1117–1124. doi: 10.1016/s0092-8674(00)81456-7. [DOI] [PubMed] [Google Scholar]
- 9.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]
- 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;410:898–904. doi: 10.1038/35073513. [DOI] [PubMed] [Google Scholar]
- 11.Shimabukuro K, Yasuda R, Muneyuki E, Hara KY, Kinosita K, Jr, Yoshida M. Catalysis and rotation of F1 motor: cleavage of ATP at the catalytic site occurs in 1 ms before 40 degree substep rotation. Proc Natl Acad Sci USA. 2003;100:14731–14736. doi: 10.1073/pnas.2434983100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Nishizaka T, Oiwa K, Noji H, Kimura S, Muneyuki E, Yoshida M, Kinosita K., Jr Chemomechanical coupling in F1-ATPase revealed by simultaneous observation of nucleotide kinetics and rotation. Nat Struct Mol Biol. 2004;11:142–148. doi: 10.1038/nsmb721. [DOI] [PubMed] [Google Scholar]
- 13.Adachi K, Oiwa K, Nishizaka T, Furuike S, Noji H, Itoh H, Yoshida M, 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]
- 14.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]
- 15.Watanabe R, Okuno D, Sakakihara S, Shimabukuro K, Iino R, Yoshida M, Noji H. Mechanical modulation of catalytic power on F1-ATPase. Nat Chem Biol. 2011;8:86–92. doi: 10.1038/nchembio.715. [DOI] [PubMed] [Google Scholar]
- 16.Adachi K, Oiwa K, Yoshida M, Nishizaka T, 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]
- 17.Rondelez Y, Tresset G, Nakashima T, Kato-Yamada Y, Fujita H, Takeuchi S, Noji H. Highly coupled ATP synthesis by F1-ATPase single molecules. Nature. 2005;433:773–777. doi: 10.1038/nature03277. [DOI] [PubMed] [Google Scholar]
- 18.Itoh H, Takahashi A, Adachi K, Noji H, Yasuda R, Yoshida M, Kinosita K. Mechanically driven ATP synthesis by F1-ATPase. Nature. 2004;427:465–468. doi: 10.1038/nature02212. [DOI] [PubMed] [Google Scholar]
- 19.Düser MG, Zarrabi N, Cipriano DJ, Ernst S, Glick GD, Dunn SD, Börsch M. 36 degrees step size of proton-driven c-ring rotation in FoF1-ATP synthase. EMBO J. 2009;28:2689–2696. doi: 10.1038/emboj.2009.213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Watanabe R, Tabata KV, Iino R, Ueno H, Iwamoto M, Oiki S, 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]
- 21.Dimroth P, von Ballmoos C, Meier T. Catalytic and mechanical cycles in F-ATP synthases. Fourth in the cycles review series. EMBO Rep. 2006;7:276–282. doi: 10.1038/sj.embor.7400646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.von Ballmoos C, Cook GM, Dimroth P. Unique rotary ATP synthase and its biological diversity. Annu Rev Biophys. 2008;37:43–64. doi: 10.1146/annurev.biophys.37.032807.130018. [DOI] [PubMed] [Google Scholar]
- 23.Mitome N, Suzuki T, Hayashi S, Yoshida M. Thermophilic ATP synthase has a decamer c-ring: indication of non-integer 10:3 H+/ATP ratio and permissive elastic coupling. Proc Natl Acad Sci USA. 2004;101:12159–12164. doi: 10.1073/pnas.0403545101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Jiang W, Hermolin J, Fillingame RH. The preferred stoichiometry of c subunits in the rotary motor sector of Escherichia coli ATP synthase is 10. Proc Natl Acad Sci USA. 2001;98:4966–4971. doi: 10.1073/pnas.081424898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Sambongi Y, Iko Y, Tanabe M, Omote H, Iwamoto-Kihara A, Ueda I, Yanagida T, Wada Y, Futai M. Mechanical rotation of the c subunit oligomer in ATP synthase (FoF1): direct observation. Science. 1999;286:1722–1724. doi: 10.1126/science.286.5445.1722. [DOI] [PubMed] [Google Scholar]
- 26.Ueno H, Suzuki T, Kinosita K, Jr, Yoshida M. ATP-driven stepwise rotation of FoF1-ATP synthase. Proc Natl Acad Sci USA. 2005;102:1333–1338. doi: 10.1073/pnas.0407857102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Ishmukhametov R, Hornung T, Spetzler D, Frasch WD. Direct observation of stepped proteolipid ring rotation in E. coli FoF1-ATP synthase. EMBO J. 2010;29:3911–3923. doi: 10.1038/emboj.2010.259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Iino R, Hasegawa R, Tabata KV, Noji H. Mechanism of inhibition by C-terminal alpha-helices of the epsilon subunit of Escherichia coli FoF1-ATP synthase. J Biol Chem. 2009;284:17457–17464. doi: 10.1074/jbc.M109.003798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Soga N, Kinosita K, Yoshida M, Suzuki T. Kinetic Equivalence of Transmembrane pH and Electrical Potential Differences in ATP Synthesis. J Biol Chem. 2012;287:9633–9639. doi: 10.1074/jbc.M111.335356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Bokranz M, Morschel E, Kroger A. Phosphorylation and phosphate-ATP exchange catalyzed by the ATP synthase isolated from Wolinella succinogenes. Biochim Biophys Acta. 1985;810:332–339. doi: 10.1016/0005-2728(85)90218-x. [DOI] [PubMed] [Google Scholar]
- 31.Slooten L, Vandenbranden S. ATP-synthesis by proteoliposomes incorporating Rhodospirillum rubrum FoF1 as measured with firefly luciferase: dependence on delta psi and delta pH. Biochim Biophys Acta. 1989;976:150–160. doi: 10.1016/s0005-2728(89)80224-5. [DOI] [PubMed] [Google Scholar]
- 32.Junesch U, Gräber P. The rate of ATP-synthesis as a function of delta pH and delta psi catalyzed by the active, reduced H+-ATPase from chloroplasts. FEBS Lett. 1991;294:275–278. doi: 10.1016/0014-5793(91)81447-g. [DOI] [PubMed] [Google Scholar]
- 33.Wiedenmann A, Dimroth P, von Ballmoos C. Functional asymmetry of the Fo motor in bacterial ATP synthases. Mol Microbiol. 2009;72:479–490. doi: 10.1111/j.1365-2958.2009.06658.x. [DOI] [PubMed] [Google Scholar]
- 34.Fischer S, Graber P. Comparison of ΔpH- and Δψ-driven ATP synthesis catalyzed by the H+-ATPases from Escherichia coli or chloroplasts reconstituted into liposomes. FEBS Lett. 1999;457:327–332. doi: 10.1016/s0014-5793(99)01060-1. [DOI] [PubMed] [Google Scholar]
- 35.Steigmiller S, Turina P, Graber P. The thermodynamic H+/ATP ratios of the H+-ATPsynthases from chloroplasts and Escherichia coli. Proc Natl Acad Sci USA. 2008;105:3745–3750. doi: 10.1073/pnas.0708356105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Petersen J, Forster K, Turina P, Gräber P. Comparison of the H+/ATP ratios of the H+-ATP synthases from yeast and from chloroplast. Proc Natl Acad Sci USA. 2012;109:11150–11155. doi: 10.1073/pnas.1202799109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Tsunoda SP, Aggeler R, Yoshida M, Capaldi RA. Rotation of the c subunit oligomer in fully functional F1Fo ATP synthase. Proc Natl Acad Sci USA. 2001;98:898–902. doi: 10.1073/pnas.031564198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Ueno H, Nishikawa S, Iino R, Tabata KV, Sakakihara S, Yanagida T, Noji H. Simple dark-field microscopy with nanometer spatial precision and microsecond temporal resolution. Biophys J. 2010;98:2014–2023. doi: 10.1016/j.bpj.2010.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Han J, Burgess K. Fluorescent indicators for intracellular pH. Chem Rev. 2010;110:2709–2728. doi: 10.1021/cr900249z. [DOI] [PubMed] [Google Scholar]
- 40.Rosing J, Kayalar C, Boyer PD. Evidence for energy-dependent change in phosphate binding for mitochondrial oxidative-phosphorylation based on measurements of medium and Intermediate phosphate-water exchanges. J Biol Chem. 1977;252:2478–2485. [PubMed] [Google Scholar]




