Skip to main content
Protein Science : A Publication of the Protein Society logoLink to Protein Science : A Publication of the Protein Society
. 2024 Mar 19;33(4):e4942. doi: 10.1002/pro.4942

Engineering of IF1‐susceptive bacterial F1‐ATPase

Yuichiro C Hatasaki 1, Ryohei Kobayashi 1,2, Ryo R Watanabe 1, Mayu Hara 1, Hiroshi Ueno 1,3, Hiroyuki Noji 1,3,
PMCID: PMC10949317  PMID: 38501464

Abstract

IF1, an inhibitor protein of mitochondrial ATP synthase, suppresses ATP hydrolytic activity of F1. One of the unique features of IF1 is the selective inhibition in mitochondrial F1 (MF1); it inhibits catalysis of MF1 but does not affect F1 with bacterial origin despite high sequence homology between MF1 and bacterial F1. Here, we aimed to engineer thermophilic Bacillus F1 (TF1) to confer the susceptibility to IF1 for elucidating the molecular mechanism of selective inhibition of IF1. We first examined the IF1‐susceptibility of hybrid F1s, composed of each subunit originating from bovine MF1 (bMF1) or TF1. It was clearly shown that only the hybrid with the β subunit of mitochondrial origin has the IF1‐susceptibility. Based on structural analysis and sequence alignment of bMF1 and TF1, the five non‐conserved residues on the C‐terminus of the β subunit were identified as the candidate responsible for the IF1‐susceptibility. These residues in TF1 were substituted with the bMF1 residues. The resultant mutant TF1 showed evident IF1‐susceptibility. Reversely, we examined the bMF1 mutant with TF1 residues at the corresponding sites, which showed significant suppression of IF1‐susceptibility, confirming the critical role of these residues. We also tested additional three substitutions with bMF1 residues in α and γ subunits that further enhanced the IF1‐susceptibility, suggesting the additive role of these residues. We discuss the molecular mechanism by which IF1 specifically recognizes F1 with mitochondrial origin, based on the present result and the structure of F1‐IF1 complex. These findings would help the development of the inhibitors targeting bacterial F1.

Keywords: ATP synthase, ATPase inhibitory factor 1, enzyme inhibitor, F1‐ATPase, IF1 , protein engineering, species specificity

1. INTRODUCTION

FoF1‐ATP synthase (FoF1) is one of the most ubiquitous enzymes; it is found in the inner membrane of mitochondria, the thylakoid membrane of chloroplast, and the plasma membrane of bacteria. This enzyme catalyzes ATP synthesis reaction from ADP and inorganic phosphate (Pi), using the proton motive force (pmf) across the membranes (Stewart et al., 2014; Walker, 2013). FoF1 consists of two structurally and functionally distinct rotary motor proteins, Fo and F1. Fo, the membrane‐embedded domain, mediates proton translocation, driving the rotation of the oligomeric c‐ring against the stator complex (ab 2). F1 is the catalytic portion of FoF1 and an ATP‐driven motor in which the rotary shaft rotates against the surrounding stator upon catalysis. In cells, Fo and F1 are connected via the complex of the rotor shafts and the peripheral stalk, which enables inter‐transmission of rotary torque between Fo and F1. The rotary direction of the rotor complex is determined by the balance of total free energy of ATP hydrolysis and pmf across the membranes. When pmf is sufficiently large, Fo rotates the rotor complex, reversing the rotation of F1 to induce ATP synthesis. Conversely, when pmf decreases, F1 hydrolyzes ATP to rotate the rotor complex in reverse, enforcing Fo to pump protons (Boyer, 1997; Junge & Nelson, 2015).

The minimal complex of bacterial F1 as a rotary motor is the α3β3γ subcomplex in which the rotor γ subunit is accommodated in the central cavity of the stator α3β3 ring (Abrahams et al., 1994). The mammalian mitochondrial F1 has the δ and the ε subunit that form the rotor shaft with the γ subunit. F1 has three catalytic sites for ATP hydrolysis/synthesis, each of which resides on the αβ interface of the α3β3‐ring. Most of the catalytical residues are found in the β subunit. F1 from bovine mitochondria (bMF1) in the ground state (Bowler et al., 2007) showed that three β subunits exhibit distinct nucleotide occupancies and conformational states; two β subunits bind to nucleotides, one bound to ATP‐analog, (denoted as βTP) and the other bound to ADP (βDP). Both βTP and βDP adopt a closed conformation, inwardly swinging the C‐terminal domain. The third β subunit, termed βE, has no bound nucleotide and takes an open conformation. The three β subunits operate catalysis in a well‐coordinated manner; the β subunits sequentially proceed the catalytic reactions, coupled with conformational transitions between open‐ and closed‐state, that induces the unidirectional rotation of the rotor (Kobayashi et al., 2020; Okazaki & Takada, 2011; Shimabukuro et al., 2003). The recent study found the intermediate conformational states (Sobti et al., 2021), revising the reaction scheme of F1 (Noji & Ueno, 2022). Due to such a highly cooperative catalytic mechanism, F1 is known to stop the overall catalysis when one of the β subunits is inhibited by an inhibitory compound or mutation (Amano et al., 1996; Gledhill & Walker, 2006).

Because FoF1 is principally obligated to function as ATP synthase, FoF1 possesses various mechanisms to suppress or inhibit the hydrolytic activity of F1 (Feniouk & Yoshida, 2006; Mendoza‐Hoffmann et al., 2022; Walker, 1994). The most universal suppression mechanism across species is ADP inhibited form that is observed as a transient pause during rotation in single‐molecule rotation assay (Hirono‐Hara et al., 2001; Kobayashi et al., 2020), although the physiological role of ADP inhibition is not well investigated. Mitochondrial FoF1 has an endogenous inhibitor protein that blocks ATP hydrolysis by F1, termed ATPase inhibitory factor 1 (IF1), which was discovered in 1963 by Pullman and Monroy (Pullman & Monroy, 1963). IF1 shows highly conserved sequence across species, which allows for an interspecies inhibition in other mitochondrial ATP synthases; bovine IF1 well suppresses ATPase activity of mitochondrial F1 from yeast (Klein et al., 1977). The bovine IF1 is 84 amino acid long. The N‐terminal region is responsible for the inhibition, binding to the αβ interface of F1 to block conformational transition of the β subunit (Kobayashi et al., 2023), while the C‐terminal region of IF1 is for a homodimer formation of IF1 (Domínguez‐Ramírez et al., 2001). When the C‐terminal region (61–84) is deleted, IF1 is always in the monomeric form retaining its full inhibitory capacity (van Raaij et al., 1996). Due to the simplicity and the ease to handle, the monomeric IF1 is often used for structural (Bason et al., 2014; Gledhill et al., 2007) and biochemical analyses (Bason et al., 2011; Kobayashi et al., 2021), including the present study.

The crystal structure of the bMF1‐IF1 1:1 complex provides insights into the atomic details of the interactions between bMF1 and IF1 (Gledhill et al., 2007; Figure 1a). The central long helix of residues 21–46 in IF1 was bound on the C‐terminus domain of the βDP subunit, seemingly inserted along the crevice of the αβDP interface. The central helix has also contact with the βTP subunit at its N‐terminus end of the helix. The N‐terminus unstructured region (8–13) and the short helix (14–18), which were followed by a distinct kink and the central helix, were deeply inserted in F1, having contacts with the γ and αE subunit. Thus, the bMF1‐IF1 complex showed the intertwined binding mode, interacting with αE, αDP, βDP, βTP, and γ subunit. The crystal structure of the bMF1‐IF1 1:3 complex was reported (Bason et al., 2014), in which βTP and βE also possessed bound IF1. IF1s bound on βTP and βE showed partially folded state in comparison with the βDP‐bound IF1, suggesting the progressive binding mechanism of IF1 coupled with folding (Bason et al., 2014); IF1 loosely associates with the most‐open αβE pair, and then it isomerizes to the final‐locked state in the αβDP site, coupled with γ rotation driven by ATP hydrolysis (Figure 1b). This progressive inhibition mechanism was in good agreement with biochemical experiments (Corvest et al., 2005; Corvest et al., 2007; Kobayashi et al., 2021; Milgrom, 1989).

FIGURE 1.

FIGURE 1

Schematic image of IF1 inhibition mechanism. (A) IF1 binding with each subunit in F1. αE, αDP, βTP, βDP, and γ interact with the full‐folded IF1 in the final locked state (PDB: 2v7q). The N‐ and C‐terminus of IF1 are shown as ‘N’ and ‘C’, respectively. Details are described in the Tables S3 and S4. (B) Two‐step reaction mechanism of IF1 inhibition: the initial association to β and the following isomerization process to the locked state (PDB: 4tt3). The β, γ, and IF1 are shown in blue, yellow, and green, respectively.

One of the unique features of IF1 is the apparent unidirectional inhibition. IF1 completely blocks the catalysis of F1 when the chemical equilibrium of ATP hydrolysis/synthesis is toward hydrolysis (De Gómez‐Puyou et al., 1980; Harris & Crofts, 1978; Lippe et al., 1988; Power et al., 1983). When the chemical equilibrium is on ATP synthesis side with sufficiently high pmf, IF1 does not interfere with the catalysis. Previously, we showed that this condition‐dependent manner of inhibition is due to the rotation‐direction‐dependent activation; bound IF1 is ejected from F1 when torque is applied to FoF1 in ATP synthesis direction (Kobayashi et al., 2023). Another remarkable feature of IF1 is distinctive species‐specific inhibition. IF1 inhibits catalysis of mitochondrial F1 but does not interfere with the catalysis of bacterial F1s (Gledhill & Walker, 2005; Suzuki et al., 2014; Wu et al., 2014). The structure of the bMF1‐IF1 1:3 complex showed that, out of 22 residues of bMF1 in contact with bound IF1, 17 residues are on β subunit, while two residues on α subunits and three residues on γ subunit (Table S1; Bason et al., 2014). Consistent with this contention, the F1–IF1 interaction has buried surface area of 2200–2400 Å2 in total, with the βDP subunit contributing to nearly half of this interface, around 1100–1200 Å2. Considering the apparent contribution of the β subunit for IF1 binding, it would be reasonable to assume that the species‐specificity of IF1 inhibition primarily arises from the β subunit. On the other hand, the amino acid conservation of the 22 residues in contact with IF1 between bMF1 and TF1 implies the contribution of the α and γ subunits for the species‐specificity of IF1; the amino acid identity of the contacting residues is the highest for the β subunit (71%), while those are modest for the α subunit (50%) and the lowest for the γ subunit (33%) (Tables S1 and S2). Further, the structural comparison between bMF1 and TF1 indicated that the β subunit exhibits the smallest RMSD value (1.27 Å), followed by the α subunit with a second lower value (1.47 Å), whereas the γ subunit displays a markedly higher value (5.02 Å) (Figure S1). The relatively lower sequence and structural identities of the α and the γ subunits among species are considered to imply that the α or the γ subunits are responsible for the species‐specific recognition of IF1. Thus, the mechanism of the species‐specific inhibition of IF1 remains elusive.

This study aims to address this question by engineering TF1 to confer the susceptibility to bovine mitochondrial IF1. To determine which subunit is responsible for the species specificity, we first tested the hybrid F1s, which are composed of each α, β, γ subunit from bMF1 or TF1 (Watanabe et al., 2023). Then, based on the sequence analysis, we constructed the quintuple and the octuple mutants of TF1, where the non‐conserved five or eight residues in contact with IF1 were substituted to the corresponding residues of bMF1. We also tested the quintuple mutant of bMF1, where the same five residues were replaced to the TF1's counterparts. Our results clearly showed that the five non‐conserved residues on the C‐terminal domain of the β subunit are crucial for species‐specific recognition by IF1. These findings would provide new insights on the recognition mechanism of IF1 as well as on the development of inhibitors targeting F1 with bacterial origin, discriminating from mitochondrial F1s.

2. RESULTS

2.1. IF1 inhibition in wild‐type bMF1 and TF1

ATPase activity was quantified by NADH absorbance monitored at 340 nm. ATP hydrolysis reaction was initiated upon the addition of F1 to the assay mixture of an ATP‐regenerating system, leading to an immediate decay in NADH absorbance upon ATP hydrolysis. As previously reported (Hirono‐Hara et al., 2001; Kobayashi et al., 2020, 2023), F1 showed initial rapid ATP hydrolysis, followed by gradual deceleration to reach the steady‐state catalysis. This time‐dependent inactivation is due to the ADP inhibition, where F1 molecules transiently lapse into an inactive state during catalysis (Figure S2 and Table S3; Lapashina & Feniouk, 2018; Turina, 2022). After confirming the steady‐state catalysis of F1, that is, 180 s after F1 injection, IF1 inhibition measurement was initiated by adding IF1 into the reaction cuvette. The blue plot in Figure 2a shows the typical time‐course of IF1 inhibition in bMF1(WT) in the presence of 1 mM ATP and 1 μM IF1, where IF1 inhibition in bMF1 was almost maximum. After IF1 injection, the slope of NADH absorbance abruptly went to almost zero, suggesting that IF1 blocked the catalysis of bMF1 within the mixing time. This rapid and complete inhibition is a significant feature of IF1 inhibition in bMF1. By contrast, TF1 was not susceptible to IF1 at all (Figure 2b). One may concern that the residual ATPase activity of TF1 was decreased down to 80% after IF1 injection. Notably, the same decay of ATPase activity was also observed in the IF1‐free condition (Figure S2). Thus, this is due to slow inactivation by ADP inhibition that slowly proceeds even after 180 s.

FIGURE 2.

FIGURE 2

IF1 inhibition assay with bMF1(WT), TF1(WT) and four hybrid F1s. (A) Time‐courses of six F1s standardized by each initial activity. The absorbance value on the y‐axis is taken from the raw data of bMF1, and the others are standardized. The triplet characters represent the origins (‘b’ from bMF1 or ‘T' from TF1) of the three subunits, α, β, and γ, in this order; bbT represents the hybrid F1 with α and β from bMF1 and γ from TF1. Experiments were conducted at 1 mM ATP and 1 μM IF1. The raw data set of the time‐courses is presented in Figure S4. (B) Residual ATPase Activity after IF1 injection. Residual ATPase activity was measured at the end of the measurement (300 s after IF1 injection). Values represent the mean value of three measurements, and error bars represent SD (N = 3 for each measurement). (C) Time‐courses of the bbT hybrid at the indicated [IF1]s with 1 mM ATP. Dkinhibitionapp plotted against [IF1]. kinhibitionapp was determined by the fitting of the NADH absorbance change by (Equation 1). The solid lines represent the fitting curve of (Equation 2). Circles and error bars represent the mean value and SD, respectively (N = 3 for each measurement). (E) Residual ATPase activity versus [IF1] at the end of the measurement. The inset is the expansion of the plots. The residual ATPase activity was <5% in each datapoint, and they mostly overlap in this figure. Circles and error bars represent the mean value and SD, respectively (N = 3 for each measurement).

2.2. IF1 inhibition in hybrid F1s

We prepared four kinds of hybrid F1s composed of subunits from bMF1 or TF1 along our previous article, bbT, bTb, bTT, and TTb (Watanabe et al., 2023). The triplet characters represent the origins (‘b’ from bMF1 or ‘T' from TF1) of the three subunits, α, β, and γ, in this order; bbT represents the hybrid F1 with α and β from bMF1 and γ from TF1. Analysis of ATPase activity in hybrid F1s is summarized in Figure S3 and Table S3. The IF1 susceptibility of these hybrid F1s was investigated under the same condition for bMF1(WT) and TF1(WT). Figure 2a shows the time‐courses standardized with the initial activities (=slope before IF1 injection). The raw data set of the time‐courses is presented in Figure S4. The residual ATPase activities at 300 s after IF1 injection are summarized in Figure 2b. The hybrid F1s did not show the IF1 susceptibility. The exception is bbT with β from bMF1 (the green plot in Figure 2a). All of the other hybrids: bTb, bTT, and TTb carry β from TF1 while α and/or γ originate from bMF1. Thus, it is evident that the β subunit is the determinant subunit for the IF1 susceptibility.

We have investigated IF1 inhibition in bbT hybrid at various concentrations of IF1 in the presence of 1 mM ATP (Figure 2c–e). The time‐courses of IF1 inhibition in bbT hybrid show clear inhibition at 0.1–1 μM IF1 (Figure 2c). We also analyzed the residual ATPase activity at 300 s after IF1 injection, and found that IF1 completely suppressed ATPase activity of bbT hybrid (Figure 2e). These features are well similar to the inhibition in bMF1(WT). To estimate the rate constant of IF1 inhibition, kinhibitionapp, we have fitted the time‐courses with the exponential decay function (Kobayashi et al., 2021; see Section 5),

yty0=Vt+V0Vkinhibitionapp1expkinhibitionappt (1)

The resultant kinhibitionapp plotted against [IF1] shows the typical saturation curve, indicating that IF1 inhibition in bbT hybrid follows the two‐step mechanism as found for bMF1(WT) (Figure 2d). According to the equation, [IF1]‐dependent kinhibitionapp is expressed as follows:

kinhibitionapp=klockIF1KMIF1+IF1 (2)

where klock is the rate constant for isomerization to the fully inhibited state, and KMIF1 represents an affinity with IF1, respectively. These values characterize kinetic features of IF1 inhibition; higher klock explains rapid inhibition and lower KMIF1 explains tight inhibition. The comparison of these kinetic parameters between bbT hybrid and bMF1 provides important insights on the molecular mechanism of IF1 inhibition. At low [IF1], where the initial IF1 association is the rate‐limiting step, bMF1 and bbT hybrid showed the similar inhibition rate, that is seen in the productive binding rate of IF1 estimated from klock/KMIF1 (Table S4; Kobayashi et al., 2021). It suggests that the interspecies substitution of γ subunit does not affect the association step of IF1. By contrast, at high [IF1], where the locking process to the final state is dominant in the overall reaction, bMF1 reached the final inhibited state 1.6 times faster than bbT, as seen in the estimation of klock. This result indicates that γ subunit can enhance the rate of the final locking process to some extent.

2.3. IF1 ‐susceptive TF1 mutants

The experiments with hybrid F1s highlighted the dominant role of β subunit in IF1 susceptibility. According to the bMF1‐(IF1)3 structure, where three IF1s were bound to each αβ pair, 22 residues in αE, αDP, βDP, βTP, and γ subunit were identified as the IF1‐contacting residues (Table S1): 17 residues are on β subunit, two residues on α subunit and three residues on γ subunit (Figure 3a). The sequence comparison between bMF1 and TF1 shows that five out of 17 residues in β subunit are non‐conserved, one out of two residues in α subunit, and two out of three residues in γ subunit, respectively (Figure 3a).

FIGURE 3.

FIGURE 3

Design concept of TF15) and TF11β5γ2). (A) Sequence comparison between bMF1 and TF1. The colored characters indicate the amino acid residues involved in IF1 binding, as found in the bMF1‐(IF1)3 structure (Table S1). The red ones represent the different residues between bMF1 and TF1, whereas the blue ones represent the identical residues. (B) Side (left) and top (right) views of the positions of mutations in the TF15) and TF11β5γ2). The structural data of PDB: 2v7q for the bMF1‐IF1 complex was used for these illustrations. αDP, βDP, γ, and IF1 are shown in pink, cyan, yellow, and green, respectively. The following mutations are included in the TF15) or TF11β5γ2): H401S, K467D, M469L, G470A, V473H in the β subunit, Q345E in the α subunit, K16N and T17I in the γ subunit (TF1 numbering).

To assess the role of these non‐conserved residues in IF1 susceptibility, we prepared two kinds of TF1 mutants, by substituting these non‐conserved amino acid residues of TF1 with the corresponding bMF1 residues. The resultant mutants were named TF15) and TF11β5γ2) (Figure 3b); TF15) is a quintuple mutant that possesses H401S, K467D, M469L, G470A, and V473H on the β subunit, and TF11β5γ2) is an octuple mutant that possesses Q345E on the α subunit, and K16N and T17I on the γ subunit, in addition to the five mutations of TF15). ATPase activity measurement of these mutants showed the maximum ATPase activity at 3 mM for TF15) and 1 mM for TF11β5γ2), although the ATPase activity of TF11β5γ2) was a few times smaller than TF15) and TF1(WT) (Figure S5). Under the above conditions, the mean time for ADP inhibition was 20 s for TF15) and 90 s for TF11β5γ2), respectively (Table S3). To minimize the possible effect of ADP inhibition on IF1 inhibition measurements, IF1 was injected at 240 s after F1 injection.

Biochemical assay of IF1 inhibition with TF15) and TF11β5γ2) was conducted under the saturating ATP conditions (Figure 4). The time‐courses of IF1 inhibition with the mutants (Figure 4a) showed gradual deceleration of ATPase activity after IF1 injection, while TF1 showed continuous ATPase activity (left panel in Figure 4a). These results confirmed that the engineered TF1s acquired IF1 sensitivity as expected. The kinetic analysis of the time‐courses revealed that the mutant TF1s also obey the two‐step inhibition mechanism, giving kinhibitionapp, (Figure 4b and Table S4). The resultant klock was determined as 0.015 s−1 of TF15) and 0.022 s−1 of TF11β5γ2), which were evidently lower than that of bMF1, 0.032 s−1. KMIF1 was estimated as 3.6 μM of TF15) and 2.2 μM of TF11β5γ2), which were higher than that of bMF1, 0.3 μM. Thus, although TF15) and TF11β5γ2) acquired IF1‐susceptibility, the affinity with IF1 and the rate of the locking process are both lower than those of bMF1. The comparison between TF15) and TF11β5γ2) shows that TF11β5γ2) is more susceptive to IF1 than TF15), suggesting the additional role of the mutations at α and γ subunit. It should be noted that both of the mutants did not show complete inhibition that was seen in the IF1 inhibition in bMF1(WT) (Figure 4c). The remarkable example is shown in the measurement of TF11β5γ2) at 10 μM IF1, where 20% of the residual ATPase was still observed. These observations indicate that the complex of the mutants and IF1 is not sufficiently stable, so that the mutants undergo reversible association and dissociation with IF1.

FIGURE 4.

FIGURE 4

IF1 inhibition assay with TF15) and TF11β5γ2). (A) Time‐courses of TF1(WT) (left), TF15) (center), and TF11β5γ2) (right) at the indicated concentrations of IF1. IF1 inhibition measurement was performed at 1 mM for TF1(WT) and TF11β5γ2) and 3 mM for TF15), respectively. Bkinhibitionapp plotted against [IF1] in TF15) (yellow), TF11β5γ2) (red), and bMF1(WT) (blue). kinhibitionapp was determined by the fitting of the NADH absorbance change by (Equation 1). The solid lines represent the fitting curve of (Equation 2). Circles and error bars represent the mean value and SD, respectively (N = 3 for each measurement). (C) Residual ATPase activity of TF15) (yellow), TF11β5γ2) (red), and TF1(WT) (gray) versus [IF1] at the end of the measurement (300 s after IF1 injection). Circles and error bars represent the mean value and SD, respectively (N = 3 for each measurement).

2.4. IF1 ‐less‐susceptive bMF15)

For further confirmation about the important role of the five non‐conserved residues in the IF1 inhibition, we prepared bMF15), in which these residues were substituted with the corresponding TF1 residues (Figure 5). The resultant mutant, termed bMF15), includes S405H, D471K, L473M, A474G, and H477V in the β subunit (in bMF1 numbering). Figure 5a shows the time‐courses of IF1 inhibition in bMF15) at 1 mM ATP. The mutant required evidently higher concentration of IF1 than bMF1(WT); the mutant required 10 μM IF1 to reach the maximum inhibition. Figure 5b shows [IF1]‐dependent kinhibitionapp plot in bMF15), which follows a typical hyperbolic curve. The resultant klock was determined as 0.021 s−1, which is also lower than that in bMF1(WT). Furthermore, the affinity parameter, KMIF1 was determined as 2.3 μM, which was seven times larger than bMF1(WT). Thus, these analyses again confirmed the crucial role of these five non‐conserved residues in IF1‐susceptibility. Interestingly, the residual ATPase activity of bMF15) can reach almost zero (Figure 5c), which is in contrast to the TF15) and the TF11β5γ2) mutants. This observation suggests that non‐mutated residues in the β‐subunit, or residues in the α and/or γ subunit, support complete inhibition.

FIGURE 5.

FIGURE 5

IF1 inhibition assay with bMF15). (A) Time‐courses of bMF15) at the indicated concentrations of IF1 with 1 mM ATP. Bkinhibitionapp plotted against [IF1] in bMF15) (purple) and bMF1(WT) (blue). kinhibitionapp was determined by the fitting of the NADH absorbance change by (Equation 1). The solid lines represent the fitting curve of (Equation 2). Circles and error bars represent the mean value and SD, respectively (N = 3 for each measurement). (C) Residual ATPase activity of bMF15) (purple) and bMF1(WT) (blue) versus [IF1] at the end of the measurement (300 s after IF1 injection). Circles and error bars represent the mean value and SD, respectively (N = 3 for each measurement).

3. DISCUSSION

In this study, we have successfully conferred IF1‐susceptibility to TF1(WT), by replacing the five non‐conserved amino acids at the C‐terminus of the β subunit to the corresponding residues of bMF1 (S405, D471, L473, A474, and H477 in bMF1 numbering). In the previous study with yeast mitochondrial F1 (yMF1) and IF1 (Wu et al., 2014), E471 and A474 that correspond to D471 and A474 of bMF1 were substituted to Lys and Glu, respectively. The yMF1 with double mutations decreased the rate constant of IF1 association by 7‐fold. This result was consistent with the present study where we observed the 7‐fold decrement in KMIF1 of bMF15) in comparison with that of bMF1(WT) (Figure 5), highlighting the important role of D471 and A474 of bMF1 for the IF1 recognition.

Here, we discuss the possible molecular mechanism on the IF1 recognition based on the present results, with the sequence comparison of the β subunits derived from mitochondrial and non‐mitochondrial enzymes (Figure S6). One straightforward idea is the contribution of the electrostatic interactions. Assuming that electrostatic interactions work as a long‐range attraction force, it is reasonable to consider that D471 of bMF1 β has an important role in the recognition process. The sequence alignment shows that Asp or Glu at this position are highly conserved in mitochondrial F1s, whereas these are not conserved among bacterial F1s (Figure S6). The negative charge of Asp/Glu found in mitochondrial F1s is likely to facilitate the initial association via electrostatic interactions with the positive charges of IF1. K46 and K47 of IF1 are the possible counterpart residues to form electrostatic interaction with Asp/Glu of the β subunit in MF1s (Figure S7).

Another possible mechanism for the IF1 recognition is that these mutations increase hydrophobic interaction with IF1, considering that D471 and A474 of bMF1 β make direct contacts with IF1 residues such as L42 and L45 in IF1 (Figure 6a and Table S1). Interestingly, the sequence alignment of β subunit showed that the A474 of bMF1 is strictly conserved in other mitochondrial F1s, whereas the corresponding residue in bacterial ones are non‐conserved (Figure S6). It should be noted that some bacterial enzymes lack a few residues after the residue 474 in bMF1 numbering. In addition, the corresponding residue in TF1 is Gly, which is likely to prevent helix formation after this residue. These would suggest the important role of the helix formation at the C‐terminus of the β subunit; the hydrophobic interaction of the C‐terminal helix with IF1 has the crucial role for the specific‐recognition of mitochondrial F1 by IF1. For further investigation of this possibility, we performed the structural prediction of a single β subunit in the IF1‐free condition using AlphaFold2 (Jumper et al., 2021; Mirdita et al., 2022). As expected, the mutant β subunit in TF15) took the α‐helix at the C‐terminus, although the wild‐type β subunit failed (Figure 6b). These would implicate that a helical structure at the C‐terminus of β subunit may facilitate the formation of the F1‐IF1 complex. One may consider why bMF15) still kept IF1‐susceptibility despite the introduction of helix‐breaking glycine. The structural prediction suggests that bMF15) can retain the C‐terminal helix even with the Gly mutation. These might be a clue for the molecular mechanism of the species‐specific recognition by IF1. The precise role of these residues in IF1 inhibition would be tested in future experiments.

FIGURE 6.

FIGURE 6

The C‐terminus of the β subunit. (A) Interaction of IF1 with the C‐terminus of the βDP subunit in the final inhibited form (PDB: 2v7q). The side chains of D471 and A474 in βDP and L42 and L45 in IF1 are represented by stick form. (B) Structural prediction of a single β subunit by AlphaFold2. The last 16 residues are illustrated in these figures.

Another interesting finding in the present study is that the relatively small number of mutations has a remarkable impact on the IF1‐susceptibility. To discuss the impact of the introduced five mutations from structural point of view, we estimated the buried surface area of hydrophobic residues of IF1 in the bMF1‐IF1 complex using PyMOL and PDBePISA (Krissinel & Henrick, 2007; Table S5 and Section 5). The difference between bMF1(WT) and bMF15) was 16 Å2, which was so small compared to the total buried surface area of the βDP subunit with IF1, 1200 Å2. Furthermore, the resultant energy difference by hydrophobic interaction was estimated to be <0.6 kcal/mol (1 k BT), implying that the stabilization/destabilization of the complex by these mutations should be, at most, a 3‐fold change in the equilibrium constant. Thus, it is difficult to account for the observed 7‐fold difference in KMIF1 between bMF1(WT) and bMF15). The more evident impact of the five mutations on the association was found in the difference in the productive binding rate of IF1 that is estimated from klock/KMIF1; bMF1(WT) has 11 times higher rate than bMF15). Taking this into account, it is likely that the principal role of the five residues is to stabilize the initial association complex with IF1 rather than to stabilize the final locked state. The positions of these residues in the three‐dimensional structure are along this contention; these residues are located at the entrance of the crevice of the αβ interface, from which IF1 would be inserted.

Probably, relevant to the above argument, the mutant TF15) did not achieve complete inhibition. In the assay with TF15), we observed some residual ATPase activity remaining under the saturating IF1 conditions, even after sufficient time has elapsed after IF1 injection (Figure 4c). This means that the five residues are not sufficient for the stable formation of the final locked state. The additional mutations into α and γ does not evidently enhance the stability of the locked state. Considering that the bbT hybrid achieved complete inhibition even with γ subunit from TF1 (Figure 2d), the incomplete inhibition in TF15) and TF11β5γ2) is attributable to more detailed difference in the structure of the β subunit between bMF1 and TF1; more bMF1‐like structure of the β subunit is required for the stable formation of the locked state. The crystal structure of bMF1‐(IF1)3 revealed that Glu30 of IF1 and the Arg408 in β subunit make a salt bridge that augments the binding of IF1 to the β subunit (Gledhill et al., 2007). The importance of this salt bridge for irreversible inhibition was demonstrated by several biochemical assays (Bason et al., 2011; Kobayashi et al., 2021). Thus, more optimized interaction between IF1 and the β subunit is a key to achieve complete IF1 inhibition.

4. CONCLUSION

Overall, we investigated the molecular origins of species‐specific IF1 inhibition by engineering bacterial F1 to confer the susceptibility to IF1. We found that the non‐conserved residues on the C‐terminus helix of the β subunit are responsible for the IF1‐susceptibility. We discuss the molecular mechanism by which IF1 specifically recognizes F1 with mitochondrial origin, based on the present result and the crystal structure of F1‐IF1 complex. These findings would provide new insights into the development of inhibitors targeting F1 with bacterial origin. The findings of the present study also prove important implications for the de novo design of the inhibitory peptides against F1s with bacterial origin.

5. MATERIALS AND METHODS

5.1. Construct and purifications of proteins

The proteins of bMF1(WT), TF1(WT), four hybrid F1s, and IF1 were expressed and purified as described in the reference article (Watanabe et al., 2023). For TF15), TF11β5γ2), and bMF15), the artificially synthesized construct with mutations was introduced into the wild‐type TF1 and bMF1 plasmids, respectively. Purifications were performed according to the method of TF1 (for TF15) and TF11β5γ2)) and bMF1 (for bMF15)). The sequence of the mutant proteins was confirmed using the Fasmac sequencing service (Fasmac, Japan). The purified mutant F1s were confirmed by sodium dodecyl sulfate poly‐acrylamide gel electrophoresis (SDS‐PAGE) analysis (Figure S8).

5.2. Biochemical assay and analysis

The ATPase activity of F1 was quantified from the oxidation rate of NADH, monitored at 340 nm (Kobayashi et al., 2021). Experiments were performed at 25°C, using a UV/VIS spectrophotometer equipped with a temperature controller. The buffer contained 50 mM HEPES‐KOH (pH 7.5; for bMF1(WT), hybrid F1s, bMF15)) or 50 mM MOPS‐KOH (pH 7.5; for TF1(WT), TF15), TF11β5γ2)), 50 mM KCl, an excess amount of MgCl2 over [ATP], 0.2 mg/mL pyruvate kinase, 2.5 mM phosphoenolpyruvate, 0.2 mM NADH and 50 μg/ml lactate dehydrogenase. NADH was added during the measurement if necessary.

The ATPase reaction was initiated by adding purified F1 into the buffer including ATP. Time‐course of ATPase activity is exemplified in Figures S2 and S3. IF1 inhibition measurement was started after the activity of F1 reached a steady state, that is, 180 s for bMF1(WT), TF1(WT), four hybrids, and bMF15), 240 s for TF15) and TF11β5γ2). As described previously (Kobayashi et al., 2021), the rate constant of IF1 inhibition (kinhibitionapp) was quantified by fitting the time‐course using a following single‐exponential decay function:

yty0=Vt+V0Vkinhibitionapp1expkinhibitionappt (3)

The residual ATPase activity was determined from the activity 300 s after IF1 injection relative to that just before IF1 injection.The progressive inhibition equation of IF1 with the initial association process followed by the isomerization process is as follows:

F1+IF1konkoffF1IF1klockF1IF1lock (4)

where F1IF1lock represents the final inhibited form, and F1IF1 is the intermediate state before isomerization. The kinetic parameters, kon and koff represent the rate constants of association and dissociation with F1, respectively. klock is the rate constant of the isomerization to the fully folded state of IF1 in the 2nd process. According to this equation, kinhibitionapp and KMIF1 are defined as follows:

kinhibitionapp=klockIF1KMIF1+IF1 (5)
KMIF1koff+klockkon (6)

The resultant kinhibitionapp plot against [IF1] follows a hyperbolic curve.

5.3. Structural comparison between bMF1 and TF1

The catalysis‐waiting structures of bMF1 (PDB: 2jdi; Bowler et al., 2007) and TF1 (PDB:7l1r; Sobti et al., 2021) were used for analysis (Figure S1). The RMSD (root mean square deviation) for each subunit was estimated from the Cα atoms of the backbone using the Python packages, MDAnalysis (Liu et al., 2010; Michaud‐Agrawal et al., 2011; Theobald, 2005). The amino acid residues analyzed here are as follows. For the α subunit, residues 28–187, 196–401, 413–483, 494, and 497–509 in bMF1, and residues 28–393, 405–475, 486–499 in TF1 were analyzed. For the β subunit, residues 10–126, 129–206, 213–310, 312–387, 396–474 in bMF1, and residues 3–21, 23–35, 43–115, 117–128, 131–306, 308–383, 392–470 were analyzed. For the γ subunit, residues 1–47, 67–69, 72–86, 105–116, 130–148, 159–173, 206, 211–271 in bMF1, and residues 6–49, 61–63, 83–100, 119–130, 140–145, 148–157, 161–163, 173–187, 226–287 in TF1 were analyzed.

5.4. Estimation of buried surface area and binding energy estimation

Buried surface areas of hydrophobic amino acids were calculated with PDBePISA (Krissinel & Henrick, 2007; Table S5). For this purpose, modeling of bMF15) was performed using the PyMOL mutagenesis tool, with the least clashed rotamer candidate selected for each mutation.

5.5. Structural prediction by AlphaFold2

Structures of a single β subunit of bMF1(WT), bMF15), TF1(WT), and TF15) (Figure 6b) were predicted using AlphaFold2 in the Google Collaboratory with default parameters (ColabFold v1.5.2‐patch: AlphaFold2 using MMseqs2; Jumper et al., 2021; Mirdita et al., 2022).

AUTHOR CONTRIBUTIONS

Ryohei Kobayashi: Investigation; visualization; conceptualization; methodology; writing – original draft; writing – review and editing. Yuichiro Hatasaki C: Investigation; methodology; writing – review and editing. Ryo Watanabe R: Investigation; writing – review and editing. Mayu Hara: Investigation. Hiroshi Ueno: Investigation; conceptualization; writing – review and editing. Hiroyuki Noji: Writing – original draft; supervision; conceptualization; writing – review and editing.

CONFLICT OF INTEREST STATEMENT

The authors declare no conflict of interest.

Supporting information

DATA S1. Supporting Information.

PRO-33-e4942-s001.pdf (1.3MB, pdf)

ACKNOWLEDGMENTS

We thank Dr. Kei‐ichi Okazaki (Institute for Molecular Science) for the fruitful discussion, and all members of the Noji Laboratory for their valuable comments. This study was supported in part by Grant‐in‐aid for JSPS Fellows (JP22KJ3188 to Ryohei Kobayashi), Grant‐in‐Aid for Scientific Research on Innovation Areas (JP21H00388 to Hirosh Ueno), and Grant‐in‐Aid for Challenging Research (Exploratory; JP23K18092 to Hiroshi Ueno), and Grant‐in‐Aid for Scientific Research (S) (JP19H05624 to Hiroyuki Noji), and a Research Grant from Human Science Frontier Program (Ref. No: RGP0054/2020 to Hiroyuki Noji).

Hatasaki YC, Kobayashi R, Watanabe RR, Hara M, Ueno H, Noji H. Engineering of IF1‐susceptive bacterial F1‐ATPase. Protein Science. 2024;33(4):e4942. 10.1002/pro.4942

Yuichiro C. Hatasaki and Ryohei Kobayashi have contributed equally to this study.

Review Editor: John Kuriyan.

REFERENCES

  1. Abrahams JP, Leslie AGW, Lutter R, Walker JE. Structure at 2.8 Å resolution of F1‐ATPase from bovine heart mitochondria. Nature. 1994;370(6491):621–628. 10.1038/370621a0 [DOI] [PubMed] [Google Scholar]
  2. Amano T, Hisabori T, Muneyuki E, Yoshida M. Catalytic activities of α3β3γ complexes of F1‐ATPase with 1, 2, or 3 incompetent catalytic sites*. J Biol Chem. 1996;271(30):18128–18133. 10.1074/jbc.271.30.18128 [DOI] [PubMed] [Google Scholar]
  3. Bason JV, Montgomery MG, Leslie AGW, Walker JE. Pathway of binding of the intrinsically disordered mitochondrial inhibitor protein to F1‐ATPase. Proc Natl Acad Sci. 2014;111(31):11305–11310. 10.1073/pnas.1411560111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Bason JV, Runswick MJ, Fearnley IM, Walker JE. Binding of the inhibitor protein IF1 to bovine F1‐ATPase. J Mol Biol. 2011;406(3):443–453. 10.1016/J.JMB.2010.12.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bowler MW, Montgomery MG, Leslie AGW, Walker JE. Ground state structure of F1‐ATPase from bovine heart mitochondria at 1.9 Å resolution. J Biol Chem. 2007;282(19):14238–14242. 10.1074/JBC.M700203200 [DOI] [PubMed] [Google Scholar]
  6. Boyer PD. The ATP synthase: a splendid molecular machine. Annu Rev Biochem. 1997;66(1):717–749. 10.1146/annurev.biochem.66.1.717 [DOI] [PubMed] [Google Scholar]
  7. Corvest V, Sigalat C, Haraux F. Insight into the bind−lock mechanism of the yeast mitochondrial ATP synthase inhibitory peptide. Biochemistry. 2007;46(29):8680–8688. 10.1021/bi700522v [DOI] [PubMed] [Google Scholar]
  8. Corvest V, Sigalat C, Venard R, Falson P, Mueller DM, Haraux F. The binding mechanism of the yeast F1‐ATPase inhibitory peptide: Role of catalytic intermediates and enzyme turnover. J Biol Chem. 2005;280(11):9927–9936. 10.1074/jbc.M414098200 [DOI] [PubMed] [Google Scholar]
  9. de Gómez‐Puyou MT, Nordenbrand K, Muller U, Gómez‐Puyou A, Ernster L. The interaction of mitochondrial F1‐ATPase with the natural ATPase inhibitor protein. Biochim Biophys Acta. 1980;592(3):385–395. 10.1016/0005-2728(80)90086-9 [DOI] [PubMed] [Google Scholar]
  10. Domínguez‐Ramírez L, Mendoza‐Hernandez G, Carabez‐Trejo A, Gómez‐Puyou A, Tuena de Gómez‐Puyou M. Equilibrium between monomeric and dimeric mitochondrial F1‐inhibitor protein complexes. FEBS Lett. 2001;507(2):191–194. 10.1016/S0014-5793(01)02979-9 [DOI] [PubMed] [Google Scholar]
  11. Feniouk BA, Suzuki T, & Yoshida M. The role of subunit epsilon in the catalysis and regulation of FOF1‐ATP synthase. Biochimica et Biophysica Acta (BBA) ‐ Bioenergetics. 2006;1757(5–6):326–338. 10.1016/J.BBABIO.2006.03.022 [DOI] [PubMed] [Google Scholar]
  12. Gledhill JR, Walker JE. Inhibitors of the catalytic domain of mitochondrial ATP synthase. Biochem Soc Trans. 2006;34(5):989–992. 10.1042/BST0340989 [DOI] [PubMed] [Google Scholar]
  13. Gledhill JR, Montgomery MG, Leslie AGW, Walker JE. How the regulatory protein, IF1, inhibits F1‐ATPase from bovine mitochondria. Proc Natl Acad Sci. 2007;104(40):15671–15676. 10.1073/pnas.0707326104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Gledhill JR, Walker JE. Inhibition sites in F1‐ATPase from bovine heart mitochondria. Biochem J. 2005;386(3):591–598. 10.1042/BJ20041513 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Harris DA, Crofts AR. The initial stages of photophosphorylation. Studies using excitation by saturating, short flashes of light. Biochim Biophys Acta. 1978;502(1):87–102. 10.1016/0005-2728(78)90134-2 [DOI] [PubMed] [Google Scholar]
  16. Hirono‐Hara Y, Noji H, Nishiura M, Muneyuki E, Hara KY, Yasuda R, et al. Pause and rotation of F1‐ATPase during catalysis. Proc Natl Acad Sci. 2001;98(24):13649–13654. 10.1073/pnas.241365698 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596(7873):583–589. 10.1038/s41586-021-03819-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Junge W, Nelson N. ATP synthase. Annu Rev Biochem. 2015;84(1):631–657. 10.1146/annurev-biochem-060614-034124 [DOI] [PubMed] [Google Scholar]
  19. Klein G, Satre M, Vignais P. Natural protein ATPase inhibitor from Candida utilis mitochondria binding properties of the radiolabeled inhibitor. FEBS Lett. 1977;84(1):129–134. 10.1016/0014-5793(77)81072-7 [DOI] [PubMed] [Google Scholar]
  20. Kobayashi R, Mori S, Ueno H, Noji H. Kinetic analysis of the inhibition mechanism of bovine mitochondrial F1‐ATPase inhibitory protein using biochemical assay. J Biochem. 2021;170(1):79–87. 10.1093/jb/mvab022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Kobayashi R, Ueno H, Li C‐B, Noji H. Rotary catalysis of bovine mitochondrial F1‐ATPase studied by single‐molecule experiments. Proc Natl Acad Sci. 2020;117(3):1447–1456. 10.1073/pnas.1909407117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Kobayashi R, Ueno H, Okazaki K, Noji H. Molecular mechanism on forcible ejection of ATPase inhibitory factor 1 from mitochondrial ATP synthase. Nat Commun. 2023;14(1):1682. 10.1038/s41467-023-37182-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Krissinel E, Henrick K. Inference of macromolecular assemblies from crystalline state. J Mol Biol. 2007;372(3):774–797. 10.1016/j.jmb.2007.05.022 [DOI] [PubMed] [Google Scholar]
  24. Lapashina AS, Feniouk BA. ADP‐inhibition of H+‐FOF1‐ATP synthase. Biochemistry (Moscow). 2018;83(10):1141–1160. 10.1134/S0006297918100012 [DOI] [PubMed] [Google Scholar]
  25. Lippe G, Sorgato MC, Harris DA. The binding and release of the inhibitor protein are governed independently by ATP and membrane potential in ox‐heart submitochondrial vesicles. Biochim Biophys Acta. 1988;933(1):12–21. 10.1016/0005-2728(88)90051-5 [DOI] [PubMed] [Google Scholar]
  26. Liu P, Agrafiotis DK, Theobald DL. Fast determination of the optimal rotational matrix for macromolecular superpositions. J Comput Chem. 2010;31(7):1561–1563. 10.1002/jcc.21439 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Mendoza‐Hoffmann F, Zarco‐Zavala M, Ortega R, Celis‐Sandoval H, Torres‐Larios A, García‐Trejo J. Evolution of the inhibitory and non‐inhibitory ε, ζ, and IF1 subunits of the F1FO‐ATPase as related to the endosymbiotic origin of mitochondria. Microorganisms. 2022;10(7):1372. 10.3390/microorganisms10071372 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Michaud‐Agrawal N, Denning EJ, Woolf TB, Beckstein O. MDAnalysis: a toolkit for the analysis of molecular dynamics simulations. J Comput Chem. 2011;32(10):2319–2327. 10.1002/jcc.21787 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Milgrom YM. An ATP dependence of mitochondrial F1‐ATPase inactivation by the natural inhibitor protein agrees with the alternating‐site binding‐change mechanism. FEBS Lett. 1989;246(1–2):202–206. 10.1016/0014-5793(89)80283-2 [DOI] [PubMed] [Google Scholar]
  30. Mirdita M, Schütze K, Moriwaki Y, Heo L, Ovchinnikov S, Steinegger M. ColabFold: making protein folding accessible to all. Nat Methods. 2022;19(6):679–682. 10.1038/s41592-022-01488-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Noji H, Ueno H. How does F1‐ATPase generate torque?: analysis from Cryo‐electron microscopy and rotational catalysis of thermophilic F1 . Front Microbiol. 2022;13:1–7. 10.3389/fmicb.2022.904084 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Okazaki KI, Takada S. Structural comparison of F1‐ATPase: interplay among enzyme structures, catalysis, and rotations. Structure. 2011;19(4):588–598. 10.1016/j.str.2011.01.013 [DOI] [PubMed] [Google Scholar]
  33. Power J, Cross RL, Harris DA. Interaction of F1‐ATPase, from ox heart mitochondria with its naturally occurring inhibitor protein. Studies using radio‐iodinated inhibitor protein. Biochim Biophys Acta. 1983;724(1):128–141. 10.1016/0005-2728(83)90034-8 [DOI] [PubMed] [Google Scholar]
  34. Pullman ME, Monroy GC. A naturally occurring inhibitor of mitochondrial adenosine triphosphatase. J Biol Chem. 1963;238(11):3762–3769. 10.1016/S0021-9258(19)75338-1 [DOI] [PubMed] [Google Scholar]
  35. Shimabukuro K, Yasuda R, Muneyuki E, Hara KY, Kinosita K, 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. 2003;100(25):14731–14736. 10.1073/pnas.2434983100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Sobti M, Ueno H, Noji H, Stewart AG. The six steps of the complete F1‐ATPase rotary catalytic cycle. Nat Commun. 2021;12(1):1–10. 10.1038/s41467-021-25029-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Stewart AG, Laming EM, Sobti M, Stock D. Rotary ATPases: dynamic molecular machines. Curr Opin Struct Biol. 2014;25:40–48. 10.1016/J.SBI.2013.11.013 [DOI] [PubMed] [Google Scholar]
  38. Suzuki T, Tanaka K, Wakabayashi C, Saita E, Yoshida M. Chemomechanical coupling of human mitochondrial F1‐ATPase motor. Nat Chem Biol. 2014;10(11):930–936. 10.1038/nchembio.1635 [DOI] [PubMed] [Google Scholar]
  39. Theobald DL. Rapid calculation of RMSDs using a quaternion‐based characteristic polynomial. Acta Crystallogr A. 2005;61(4):478–480. 10.1107/S0108767305015266 [DOI] [PubMed] [Google Scholar]
  40. Turina P. Modulation of the H+/ATP coupling ratio by ADP and ATP as a possible regulatory feature in the F‐type ATP synthases. Front Mol Biosci. 2022;9:1–8. 10.3389/fmolb.2022.1023031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. van Raaij MJ, Orriss GL, Montgomery MG, Runswick MJ, Fearnley IM, Skehel JM, et al. The ATPase inhibitor protein from bovine heart mitochondria: the minimal inhibitory sequence. Biochemistry. 1996;35(49):15618–15625. 10.1021/bi960628f [DOI] [PubMed] [Google Scholar]
  42. Walker JE. The regulation of catalysis in ATP synthase. Curr Opin Struct Biol. 1994;4(6):912–918. 10.1016/0959-440X(94)90274-7 [DOI] [PubMed] [Google Scholar]
  43. Walker JE. The ATP synthase: the understood, the uncertain and the unknown. Biochem Soc Trans. 2013;41(1):1–16. 10.1042/BST20110773 [DOI] [PubMed] [Google Scholar]
  44. Watanabe RR, Kiper BT, Zarco‐Zavala M, Hara M, Kobayashi R, Ueno H, et al. Rotary properties of hybrid F1‐ATPases consisting of subunits from different species. IScience. 2023;26(5):106626. 10.1016/j.isci.2023.106626 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Wu Q, Andrianaivomananjaona T, Tetaud E, Corvest V, Haraux F. Interactions involved in grasping and locking of the inhibitory peptide IF1 by mitochondrial ATP synthase. Biochim Biophys Acta. 2014;1837(6):761–772. 10.1016/J.BBABIO.2014.01.023 [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

DATA S1. Supporting Information.

PRO-33-e4942-s001.pdf (1.3MB, pdf)

Articles from Protein Science : A Publication of the Protein Society are provided here courtesy of The Protein Society

RESOURCES