H+-ATP synthase (F1Fo ATPase) catalyzes the synthesis and/or hydrolysis of ATP, and the reactions are strongly affected by all the substrates (products) in a way clearly distinct from that expected of a simple reversibly operating enzyme. Recent studies have revealed the structure of F1, which is ideally suited for the alternating binding change mechanism, with a rotating γ-subunit as the energy-driven coupling device. According to this mechanism ATP, ADP, inorganic phosphate (Pi) and Mg2+ participate in the forward and reverse overall reactions exclusively as the substrates and products. However, both F1 and F1Fo demonstrate non-trivial steady-state and pre-steady-state kinetics as a function of variable substrate (product) concentrations. Several effectors cause unidirectional inhibition or activation of the enzyme. When considered separately, the unidirectional effects of ADP, Pi, Mg2+ and energy supply on ATP synthesis or hydrolysis may possibly be explained by very complex kinetic schemes; taken together, the results suggest that different conformational states of the enzyme operate in the ATP hydrolase and ATP synthase reactions. A possible mechanism for an energy-dependent switch between the two states of F1Fo ATPase is proposed.

Mitochondrial H+-ATP synthase (FoF1-H+ ATPase) catalyzes the reaction:
where n is the quotient for stoichiometric vectorial proton translocation from the cytoplasm (H+ ) to the matrix space (H+ ) and m is the stoichiometric quotient for scalar proton consumption, which is determined by the difference in total acidity of the substrates (ADP and inorganic phosphate) and product (ATP). The value of n (2–4) is not definitely established; m is close to 1 at pH>7.5 and depends predominantly on pH and Mg2+ concentration, forming more (ATP) or less (ADP and Pi) stable complexes with the reaction product and substrates. The enzyme is composed of two distinct parts. The F1 sector (α3β3γδε minimal subunit structure) bears catalytic sites where nucleotide-substrate/product and presumably Pi bind. The membrane-spanning Fo sector forms a proton-conducting path. It is generally accepted that work done by the flow of H+ through Fo is transformed into the chemical work needed to form ATP as described by equation 1. Energy-dependent ATP synthesis is phenomenologically reversible: ATP hydrolysis in tightly coupled membranous preparations results in a formation of the difference between the electrochemical potentials of hydrogen ions on different sides of the coupling inner mitochondrial membrane . Whether the ATPase activity of these preparations is a simple reversal of ATP synthesis is not known because the precise mechanisms of ATP hydrolysis and its synthesis are not known. Spectacular achievements have been made in elucidating the ternary structure of F1 at an atomic resolution of 2.8×10−10 m (Abrahams et al., 1994; Bianchet et al., 1998). F1 appears as a trimer of tightly packed αβ-subunit pairs and a long rod-like γ-subunit, which is positioned in the central cavity of the spherical globular protein. The ternary structure of smaller δ- and ε-subunits and their positions relative to the α3β3γ heptamer have not been revealed by crystallography, although structural features of the Escherichia coli ε-subunit and possible sites of its interaction with the γ-subunit of F1 and Fo have been determined by nuclear magnetic resonance spectroscopy (Wilkens et al., 1995). The Fo sector has the structure ab2c12 (the minimal composition); only the structure of the c subunit, which seems to be a key component of the proton-conducting path, is well established (Fillingame, 1997). The catalytic sites residing on F1 are located far away from the plane of the coupling membrane. This spatial separation suggests that -dependent ATP synthesis must be a complex electro-mechano-chemical process. The rotational alternating binding change mechanism, originally based on quantitative measurements of isotope exchange reactions (Gresser et al., 1982) and supported by numerous studies including elegant demonstrations of strong negative cooperativity in ATP binding and positive cooperativity in ATP hydrolysis (Grubmeyer and Penefsky, 1981; Grubmeyer et al., 1982), is now widely accepted (Boyer, 1998). Essential features of this mechanism are as follows: (i) the free energy produced by H+ flow is used to release tightly bound ATP spontaneously (isoenergetically) formed at the very high affinity ATP-binding site of the enzyme; (ii) each catalytic site sequentially (in the turnover time scale) appears as: empty → tight → loose in terms of occupation and tightness of nucleotide binding; and (iii) the occupation and mode of nucleotide binding at one catalytic site within the trimer strictly determine the properties of two other sites. The most plausible way to realize such a mechanism is to rotate one element of the F1 structure relative to the trimerically arranged catalytic sites (Boyer, 1993). The spatial structure of F1 seems to fit ideally the proposed mechanism. Perhaps the most impressive achievement of bioenergetic enzymology during the last decade was the original prediction and further indirect demonstration (Duncan et al., 1995; Junge et al., 1996; Sabbert et al., 1996) and final direct visualization (Noji et al., 1997) of ATP-dependent rotation of the γ-subunit (the rotor) within three αβ-subunit pairs (the stator).

The rotational alternating binding change mechanism apparently leaves no alternative but for ADP, Pi, ATP and ΔμH+ to participate in the overall reaction in a way different from that required for the substrate/product binding/release and driving force as depicted in equation 1. However, the kinetic properties of either soluble F1 or membrane-bound F1Fo are far from being trivial. The purpose of this review is twofold. First, I will try to summarise the standard (steady-state) and non-standard (slow pre-steady-state) kinetic properties of the enzyme. They will be described mainly for bovine heart mitochondrial ATPase, and data for other ATPases of the F1 family (Escherichia coli, chloroplasts, thermophilic bacteria) will be referred to only when absolutely necessary. Countless papers relevant to the subject have been published during last 20–30 years. A newcomer to the field is often faced by a number of contradictory statements circulating in various reviews. This paper is by no means a comprehensive review, but rather a summary of some studies performed by my own group.

The second problem to be discussed briefly concerns the reversibility of F1Fo ATP synthase. Analysis and further development of the rotational mechanism of F1 ATPase, which is now often considered to be the smallest electro-mechanical machine, has resulted in the introduction of new terminology to describe the enzyme: a number of terms previously unknown to enzymology, such as rotor, stator, torque, shaft, spring, etc., are now widely used in the current literature. If, as according to several language dictionaries, a general definition of a machine is an ‘assemblage of parts that transit forces, motions and energy from one to another in a predetermined manner’ is to be accepted, the question arises: do reversible molecular machines exist? In my opinion, the molecular machines created by Nature and those created by mankind are made exclusively for unidirectional energy transduction. Thus, a special mechanism that can be switched on or off by the applied driving force, thus permitting unidirectional energy transmission, must be a component of any machine.

When soluble F1, oligomycin-sensitive purified F1Fo or submitochondrial particles are added to a reaction mixture containing ATP and Mg2+, rapid hydrolysis of ATP is observed. The enzyme activity under ‘optimal’ conditions has a turnover number of several hundreds per second. It might be expected that the standard kinetic parameters of this remarkably stable and highly active enzyme could easily be determined. However, this is not the case. Meaningful interpretation of enzyme kinetics is only possible when initial steady-state rates are measured. When the hydrolytic activity of F1-type ATPase is considered, the initial rate, as such, is not an unambiguous term. There are several reasons for this. First, many preparations contain a small natural inhibitory peptide, an ATPase inhibitor protein (Pullman and Monroy, 1963), which is an intrinsic part of the mitochondrial F1Fo ATP synthase. The inhibitory effect of this protein is time-, pH-, ionic-strength- and ATP-dependent (Harris and Das, 1991; Panchenko and Vinogradov, 1985, 1989) and may also be dependent on and on the presence of divalent cations (Vasilyeva et al., 1989; Yamada and Huzel, 1988). It appears that uncontrolled contamination of the protein inhibitor may be a source of serious errors in measurements of the actual initial velocities of ATPase in many early studies, including our own (Akimenko et al., 1972). Second, the product of ATP hydrolysis, ADP, is an inhibitor of all ATPase preparations. One type of inhibition (rapidly equilibrating competitive inhibition) can be easily prevented if ATP-regenerating systems, such as phosphoenolpyruvate and pyruvate kinase or phosphocreatine and creatine kinase, are used in the assays. However, another unusual time-dependent inhibition by ADP (Fitin et al., 1979) is a potential source of serious errors in the ATPase assays even in the presence of ATP-regenerating systems. When ATP hydrolysis is followed using continuous methods, the rate of product release is almost always biphasic or triphasic, depending on the ‘history’ of a particular preparation. It shows a ‘burst’ or lag, and the time-dependence of the active → inactive or inactive → active transition is a very complex function of the concentrations of nucleotides and [Mg2+] (Bulygin and Vinogradov, 1988; Vasilyeva et al., 1980, 1982a; Vinogradov, 1984). Thus, the possibility that erroneous conclusions may be reached from the measurements of apparent ‘initial’ rates is considerable. Although the details of the unusual hysteresis behaviour of ATPase have been discussed (Vinogradov, 1984), only a few research groups have paid attention to this phenomenon, and the results of ATPase kinetics studies should be interpreted with caution.

Some standard kinetic parameters of ATP hydrolysis by the mitochondrial ATPase are summarized in Table 1 and briefly discussed below.

Table 1.

Steady-state kinetic parameters for mitochondrial FoF1ATPase (submitochondrial particles)

Steady-state kinetic parameters for mitochondrial FoF1ATPase (submitochondrial particles)
Steady-state kinetic parameters for mitochondrial FoF1ATPase (submitochondrial particles)

Adenosine triphosphate

ATP and other nucleotides (ITP, GTP) are hydrolyzed at significant rates only in the presence of Mg2+ or some other divalent cations (Mn2+, Co2+, Fe2+). At high Mg2+ concentrations (2–10 mmol l−1 at pH 7.0–8.0), the dependence of the hydrolytic rate on ATP concentration is a simple hyperbolic function within the range 2–5000 μmol l−1 with an apparent Km of approximately 10−4 mol l−1. This finding contrasts with several reports in the literature in which two or three Km values for ATP hydrolysis have been found. There are strong reasons to believe that the non-hyperbolic dependence of ATPase activity on ATP concentration is due to an erroneous perception of the observed rates as the initial ones (see above). It is worth noting that the experimentally well-documented strong negative cooperativity in ATP binding and extremely positive cooperativity in hydrolytic performance (Grubmeyer and Penefsky, 1981; Grubmeyer et al., 1982) do not conflict with the simple Michaelis–Menten kinetics of steady-state ATP hydrolysis. So-called unisite catalysis is seen under conditions when the enzyme concentration is several times higher than that of ATP. The affinity for ATP binding to the ‘first’ catalytic site is extremely high (approximately 10−12 mol l−1; Grubmeyer et al., 1982; Penefsky, 1985; Souid and Penefsky, 1995), and this concentration range is never covered in the steady-state kinetic measurements. Also, simple Michaelis–Menten kinetics do not conflict with the alternating binding change mechanism (see Introduction): if cooperativity of binding is very strong or absolute (as in the ‘flip-flop’ mechanism; Lazdunski, 1972), a single ‘entrance’ for the substrate and a single ‘exit’ for the products would exist. An interesting possibility is that, during the steady-state performance of F1-type ATPase/synthase, a ‘gate’ and special ‘path’ operate to determine the penetration of ATP (or ADP) from the medium into a ‘catalytic cavity’ where three exchangeable sites (out of a total of six) are located. In the light of the structural arrangement of F1, it seems unlikely that what are assigned as ‘catalytic’ and ‘noncatalytic’ sites on the β and α subunits are readily accessible to nucleotides in solution. The Km value for ATP depends strongly on Mg2+ concentration. A high ATP concentration (a significant molar excess over [Mg2+]) inhibits ATPase activity. This inhibition is most likely due to a decrease in free [Mg2+], which is needed (in addition to that bound as an ATP.Mg complex) for the catalysis (Syroeshkin et al., 1999).

Adenosine diphosphate

ADP, the product of the ATPase reaction, is an inhibitor of the hydrolytic activity. Quantitative and qualitative parameters for inhibition by ADP in terms of steady-state kinetics are difficult to determine because ADP slowly deactivates the enzyme (the term ‘deactivation’ will be used throughout this article to discriminate between a decrease of only the initial velocity (i.e. appearance of the lag phase in the enzyme performance) and the usual inhibition of the activity (see below). From the initial rate measurements, it has been concluded that ADP is a purely competitive inhibitor (with ATP) of hydrolysis (Hammes and Hilborn, 1970). We found that free ADP, not the ADP.Mg complex, competes with the ATP.Mg complex for the ‘substrate-binding’ site (Syroeshkin et al., 1999). This finding is of particular interest because the ADP.Mg complex has been shown to be a true substrate for photophosphorylation in chloroplasts (Zhou and Boyer, 1992) and for oxidative phosphorylation in submitochondrial particles (Galkin and Syroeshkin, 1999). Certainly, as pointed out above, one needs to define clearly what is assumed to be the ‘substrate-binding site’: one of these that are seen in the resolved F1 structure, or the ‘gate-specific path’ step which may contribute significantly to the overall steady-state reaction.

Inorganic phosphate

Although Pi is equally as important as ADP as a substrate/product of ATP synthase, only a few studies, compared with those on nucleotide binding, are available on its specific effects on the mitochondrial ATPase. Soluble F1 has been shown to bind Pi with a relatively high affinity (Kasahara and Penefsky, 1978; Penefsky, 1977). It would be logical to expect that Pi must inhibit ATP hydrolysis with an apparent Ki similar to its Km value for oxidative phosphorylation (in the millimolar range). However, this is not the case, and Pi either has no effect on ATP hydrolysis or it slightly stimulates ATPase activity, as do several other anions. It is extremely difficult, if not impossible, to construct any kinetic scheme that includes reversible Pi release and that does not include a Pi inhibition term in an equation for the steady-state reaction rate. This problem has somehow been overlooked in detailed discussions and experimental studies of the enzyme. The only possible explanation for the absence of an inhibitory effect is that Pi irreversibly dissociates from the active site during ATP hydrolysis: this proposal immediately creates an unavoidable problem in viewing ATP synthase as a reversible ATPase.

Mg2+

It is well established that Mg2+ (or some other divalent cation) is needed for ATPase activity (Selwin, 1967; Akimenko et al., 1972; Fleury et al., 1980). Since the early publications on ATPase kinetics (Selwin, 1967; Fleury et al., 1980), statements on the inhibition of ATPase by excess free Mg2+ have been widely circulated in the literature. Recently, we have reinvestigated the problem and found no inhibition of enzyme activity by Mg2+ (up to 10–15 mmol l−1) provided that true initial rates are measured. It appears that the previously reported inhibitory effects of Mg2+ were due to underestimations of another effect: Mg2+-induced, ADP-dependent (or vice versa) slow deactivation of the enzyme (Bulygin and Vinogradov, 1991; Guererro et al., 1990). Moreover, we found that free Mg2+, in addition to that bound in the true substrate (ATP.Mg+), is needed for ATPase activity. The kinetic analysis suggests that activating Mg2+ is involved in the overall reaction according to the so-called ‘ping-pong’ mechanism, i.e. it participates after irreversible dissociation of one of the products from the active site. When only a limiting concentration of free Mg2+ is available, ATPase activity becomes strongly pH-dependent, as to be expected if Mg2+ binds to a deprotonated group with a pKa of 7.4 (Syroeshkin et al., 1999).

Proton-motive force

The ATPase activity of intact mitochondria is greatly stimulated by uncoupling agents. Unfortunately, the ATPase activity of coupled submitochondrial particles is only slightly (not more than twofold) accelerated when is collapsed.

Our knowledge on the effects of on steady-state ATP hydrolysis is very limited.

To summarise, the simplest kinetic scheme for the steady-state ATPase reaction is depicted in Fig. 1. The initial steady-state rate reaction mechanism (under conditions where slow deactivation of the enzyme is prevented) is shown as the sequence of steps 1–5. ATP.Mg is the species taken up as the substrate. After the hydrolytic step (step 2), Pi leaves the active site before ADP. The absence of any inhibitory effect of Pi on F1Fo ATPase suggests that it leaves the enzyme during uncoupled ATP hydrolysis from a site with very low or no affinity and is taken as the thermodynamic basis for the kinetic sequence of product release. The simple competitive inhibition of ATPase by free ADP is in accord with the reaction sequence depicted in Fig. 1, in which free ADP leaves the catalytic site via a Mg2+-dependent mechanism. The data on tight binding of ADP and some nucleotide derivatives to F1 in the absence of Mg2+ and on rearrangements of the nucleotide-binding sites after subsequent addition of Mg2+ (Burgard et al., 1994; Yoshida and Allison, 1983) are in accord with the proposed sequence. Since the structure of α3β3γ at 2.8×10−10 m resolution is now available (Abrahams et al., 1994; Bianchet et al., 1998), an obvious question arises: where is the Mg2+-specific ‘activating’ site located? The structures of the nucleotide-binding sites in asymmetric F1 containing Mg.nucleotide complexes bound to the α and β subunits have been visualised (Abrahams et al., 1994). Surprisingly, no Mg2+ other than that originating from the Mg.nucleotide complexes was seen, despite the presence of adenosine-5′-phosphoimidophosphate, ADP, Mg2+ and sodium azide and the absence of Pi in the crystallization medium (Lutter et al., 1993). There are several ways of explaining the absence of visualized Mg2+. Significant rearrangement of F1 subunits induced by an interaction between F1 and Fo is one possibility (Böttcher et al., 1995). Another possibility is that the Mg2+-specific site is located on the disordered ‘invisible’ part of the γ-subunit. This proposal may be substantiated by the finding that ATP hydrolysis catalyzed by the reconstituted α3β3γ complex of Thermus thermophilus F1 is not sensitive to azide, which is known to inhibit complete α3β3 ATPase by a Mg2+-dependent mechanism (Bulygin and Vinogradov, 1991; Bulygin et al., 1993; Vasilyeva et al., 1982a). If the proposal for γ-subunit Mg2+-binding is correct, the structural arrangements of the events shown as the reaction scheme in Fig. 1 would be as follows: free ADP that is left at the catalytic site after the departure of Pi.Mg is then released to the solution via an interaction with the Mg2+-liganded arm of the γ-subunit. The key role of Mg2+ in energy-transducing enzymes was suggested many years ago (Jenkins et al., 1984; Racker, 1977, 1979) and merits further study.

Fig. 1.

Reaction pathways for ATP hydrolysis catalyzed by F1Fo ATPase. E, T, D, Pi, M and Ma stand for the enzyme, ATP, ADP, inorganic phosphate, Mg2+ originating from the ATP.Mg complex (T.M) and free activating Mg2+, respectively. The scheme does not attempt to portray the interactions between the enzyme subunits. The hydrolytic steps, steps 2–5, are almost certainly dependent on cooperative interactions within the multisubunit enzyme [the alternating binding change mechanism (Boyer, 1993), which may include γ-subunit rotation (Noji et al., 1997)]. The scheme is intended to describe the fate of the substrate and enzyme–substrate complexes during catalytic turnover steps 1–5. Steps 4 and 5 are shown as binding and release of Mg2+ for the sake of clarity; no differences in the kinetic behaviour are expected if activating Mg2+, participating in the reaction cycle after Pi release, remains permanently bound to the enzyme during catalysis. Ks is the dissociation constant of the first enzyme–substrate complex E.(TM); the relative contributions of the first-order rate constants of the dissociation of Pi (kp) and ADP (kcat) to the apparent Km for the ATP.Mg complex, compared with that of Ks, depends on the concentration of activating Mg2+. Slow deactivation of the enzyme is described by step 6 (isomerization of the enzyme–product complex, which is prevented by some anion activators such as sulphite) or by a combination of steps 5 and 6 (deactivation by added ADP and Mg2+). The irreversibility of step 5 holds only for the initial rates, when the concentration of ADP is small enough to be neglected. Step 8 describes ATP-dependent activation of the deactivated enzyme; the ionic species of ATP and the mechanism of activation (Fitin et al., 1979; Jault and Allison, 1994) remain to be established. Deactivated forms of the enzyme (E*) (whether or not stabilized by azide; step 7) are proposed to be transformed into ATP synthase via an energy-dependent mechanism (see Fig. 2). What particular component of the total proton-motive force

Δμ¯H+
[electrical potential difference (Δψ) or pH, or both] is involved in the energy-dependent transformation of the enzyme is not known.

Fig. 1.

Reaction pathways for ATP hydrolysis catalyzed by F1Fo ATPase. E, T, D, Pi, M and Ma stand for the enzyme, ATP, ADP, inorganic phosphate, Mg2+ originating from the ATP.Mg complex (T.M) and free activating Mg2+, respectively. The scheme does not attempt to portray the interactions between the enzyme subunits. The hydrolytic steps, steps 2–5, are almost certainly dependent on cooperative interactions within the multisubunit enzyme [the alternating binding change mechanism (Boyer, 1993), which may include γ-subunit rotation (Noji et al., 1997)]. The scheme is intended to describe the fate of the substrate and enzyme–substrate complexes during catalytic turnover steps 1–5. Steps 4 and 5 are shown as binding and release of Mg2+ for the sake of clarity; no differences in the kinetic behaviour are expected if activating Mg2+, participating in the reaction cycle after Pi release, remains permanently bound to the enzyme during catalysis. Ks is the dissociation constant of the first enzyme–substrate complex E.(TM); the relative contributions of the first-order rate constants of the dissociation of Pi (kp) and ADP (kcat) to the apparent Km for the ATP.Mg complex, compared with that of Ks, depends on the concentration of activating Mg2+. Slow deactivation of the enzyme is described by step 6 (isomerization of the enzyme–product complex, which is prevented by some anion activators such as sulphite) or by a combination of steps 5 and 6 (deactivation by added ADP and Mg2+). The irreversibility of step 5 holds only for the initial rates, when the concentration of ADP is small enough to be neglected. Step 8 describes ATP-dependent activation of the deactivated enzyme; the ionic species of ATP and the mechanism of activation (Fitin et al., 1979; Jault and Allison, 1994) remain to be established. Deactivated forms of the enzyme (E*) (whether or not stabilized by azide; step 7) are proposed to be transformed into ATP synthase via an energy-dependent mechanism (see Fig. 2). What particular component of the total proton-motive force

Δμ¯H+
[electrical potential difference (Δψ) or pH, or both] is involved in the energy-dependent transformation of the enzyme is not known.

Mitchell and Moyle (1971) were the first to observe a relatively slow time-dependence of some kinetic properties of the ATPase. This phenomenon, although demonstrated for ATPases from a variety of sources (Moyle and Mitchell, 1975; Rectenwald and Hess, 1977, 1979), was somehow ignored for a long time. In 1979, it was discovered that ADP is responsible for the slow transient kinetics of ATPase (Fitin et al., 1979). Since then, a number of detailed studies have shown that an ADP(Mg2+)-dependent slow active/inactive transition is a universal property of F1-type ATPases (Jault and Allison, 1994; Matsui et al., 1997; and references cited therein) and, as has been shown more recently, is also a characteristic of V-type ATPases (Yokoyama et al., 1998). Some essential features of this still enigmatic phenomenon are briefly summarized below.

When ATP at relatively low concentrations (::sKm) is hydrolyzed by ATPase free of protein inhibitor in the absence of an ATP-regenerating system, a rapid decline in the activity is observed which is not due to a decrease in the substrate concentration. The addition of further enzyme results in an increase in activity that would be expected if the enzyme added previously had been irreversibly inactivated (Fitin et al., 1979).

Brief preincubation of F1Fo (submitochondrial particles) or soluble F1 with very low, almost stoichiometric, concentrations of ADP in the presence of Mg2+ results in a considerable lag (on a time scale of minutes) in the onset of ATPase activity measured in the presence of an ATP-regenerating system. Qualitatively, the same phenomenon is seen for standard F1 preparations after preincubation with Mg2+ alone, but is absent when nucleotide-depleted F1 is assayed with an ATP-regenerating system after preincubation with Mg2+ (Minkov et al., 1979).

The initial rates of ATP hydrolysis by submitochondrial particles or soluble F1 measured under unified conditions (with or without regenerating enzymes) are poorly reproducible. The activities are markedly increased and become quite reproducible after prolonged preincubation of the preparations in the presence of phosphoenolpyruvate plus pyruvate kinase or phosphocreatine and creatine kinase. After such activation, the ATPase activity measured with an ATP-regenerating system becomes biphasic: a rapid initial rate slowly decreases to 50–70 % of the original level, and the reaction then proceeds at a constant rate (Vasilyeva et al., 1980, 1982a).

Neither azide (a specific inhibitor of F1 ATPase) nor sulphite (an activator of the enzyme) affects the initial rapid phase of ATP hydrolysis, whereas the delayed steady-state rate is inhibited by azide and stimulated by sulphite (Vasilyeva et al., 1982b). Hydrolysis of other nucleotides (ITP, GTP) by activated ATPase occurs at a constant rate (no lags or ‘bursts’) and is insensitive to azide or sulphite. Preincubation of the enzyme with IDP or GDP has no effect on its ATPase activity, whereas preincubation with ADP completely blocks the initial rates of hydrolysis of other nucleotides.

The duration of the lag phase in the ATP regeneration assay depends on the ATP concentration approximately in the same way as the steady-state ATPase activity does itself. Thus, activation of the enzyme during preincubation with pyruvate kinase plus phosphoenolpyruvate (see above) occurs much more slowly than in the assay system in the presence of added ATP.

When ATPase is first deactivated by ADP+Mg2+ and then passed through a Sephadex column in the presence of EDTA to remove Mg2+, the enzyme becomes ‘active’ and can again be deactivated by added Mg2+ (Bulygin and Vinogradov, 1991). In other words, F1Fo bears a specific site for Mg2+ binding with an apparent affinity close to that measured in the steady-state kinetic experiments. Mg2+ affinity, as measured using these two independent approaches, is strongly pH-dependent. It increases at alkaline pH and decreases upon acidification (Bulygin et al., 1993; Syroeshkin et al., 1999).

Pi, which does not affect the rate of ATP hydrolysis, strongly modulates the ADP(Mg2+)-induced deactivation. It decreases the affinity of the ADP-specific deactivating site, so that the enzyme can be reactivated equally well by either pyruvate kinase or Pi (Yalamova et al., 1982). Pi also significantly increases the rate of EDTA-induced activation (see above) (Bulygin and Vinogradov, 1991).

The results listed above show that, in addition to the rapid ATP hydrolytic steady-state turnovers (steps 1–5, Fig. 1), an additional pathway of enzyme transformation exists (as described in Fig. 1 by steps 6, 7 and 8) which explains the pre-steady-state kinetics of ATP hydrolysis and also the inhibitory and activating effects of azide and sulphite. This scheme as presented is an extension of that suggested previously (Vasilyeva et al., 1982b).

ADP dissociates from its ATPase inhibiting site very slowly (the first-order rate constant varies from 0.1 to 1 min−1). This site cannot therefore be considered as participating in the ATPase reaction. Many years ago, we showed that ADP(Mg2+)-deactivated ATPase is able to catalyze oxidative phosphorylation without any lag in ATP formation (Minkov et al., 1980). This observation led us to conclude that, when F1 catalyzes ATP hydrolysis or -dependent ATP synthesis, it is in an alternative, slowly interconvertible, conformation. In other words, the microreversibility principle, which is usually quantitatively expressed in enzymology as Haldane relationships, does not hold for F1 ATPase. The main difficulty in obtaining experimental evidence for or against such a proposal is that the conditions for measurement of the reaction in the two directions are quite different. Some recent findings relevant to this problem are briefly discussed below.

The ADP(Mg2+)-deactivated enzyme can be trapped by azide (see Fig. 1). Surprisingly, tightly coupled submitochondrial particles in which ADP-deactivated F1 is trapped by azide catalyze succinate-supported ATP synthesis at the same rate as control samples, whereas both uncoupled ATPase activity and ATP-dependent generation cease (Syroeshkin et al., 1995). Further studies of this unidirectional inhibition have revealed that (a ‘substrate’ or ‘product’ of the reactions) is a strong effector of the F1Fo complex (Galkin and Vinogradov, 1999). If ATP synthesis catalyzed by the azide-trapped deactivated enzyme is stopped because of the collapse of no further hydrolysis of added ATP occurs. However, if ATP is added first, rapid hydrolysis occurs after subsequent addition of an uncoupler to the submitochondrial particles, and the reaction rate decreases slowly over time because of inhibition by azide. It has been shown that this striking effect of is not due to back pressure (inhibition by a ‘product’), and the only ligand that prevents dependent enzyme transformation is free ADP. A similar (if not identical) phenomenon of dependent transformation of ATPase in chloroplasts is well documented (Du and Boyer, 1989; Groth and Junge, 1995; Junge, 1970; Zhou and Boyer, 1993; and references cited therein) and has been interpreted using a quite complex kinetic scheme (Gräber, 1994). More recently, the unidirectional inhibition of ATPase activity by azide has been confirmed in elegant studies on a reconstituted system composed of Thermus thermophilus F1Fo ATP synthase and bacterial rhodopsin (Bald et al., 1998).

How the different reaction pathways in ATP synthesis and hydrolysis are catalyzed by F1Fo has been discussed in our previous publications (Vinogradov, 1984; Vinogradov et al., 1985) using a language for enzyme kinetics that is not agreeably accepted by most scholars in the field. The remarkable progress in structural studies of the enzyme enable us to present our working hypothesis as the scheme shown in Fig. 2. The key point in this scheme is that the ‘clutch’ between a rotating γ-subunit and the α3β3 complex is changed when the enzyme operates in different directions. The simplest driving mechanism to shift the ‘clutch’ from one mode of operation to the other could be the vertical (perpendicular to the plane of the membrane) dependent movement of one or several subunits of Fo that are firmly attached to the F1 sector. The minimal requirement for such a movement would be the positioning of a charged (or potentially charged) group within the membrane-embedded area to function as a sensitive sensor. This model is different from those in which an elastic spring (γ-subunit) is proposed for the transmission of a torque from Fo to the catalytic sites of F1 (Cherepanov et al., 1999; Yasuda et al., 1998). I believe that a rigid ratchet mechanism operates in the electro-mechanical energy transduction. If Nature has created an electro-mechanical motor, it has also provided it with a special mechanism to allow it to operate effectively as a mechano-electrical generator, just like similar devices created by mankind are differently constructed and operated although based on the same physical principles. It would be of great interest to apply several modern techniques, such as cryoelectron microscopy (Gogol et al., 1990) and scanning force microscopy (Singh et al., 1996), to visualise the structure of F1Fo under the conditions of generation.

Fig. 2.

Schematic representation of the unidirectional operation of the mitochondrial ATPase/ATP synthase. Only one (out of three functioning αβ-subunit pair) is shown as that which has the γ–αβ catch. The relative positions and molecular masses of the ab2δε complex are not specified. Readers should consult the series of excellent mini-reviews describing this aspect and the dynamics of the different subunits of F1Fo (Pedersen, 1996). The key proposal is that there is an energy-driven switch between three types of transmission: (B) no enzyme turnover, no leak for H+ through the Fo component; (A) the ATP synthase state; and (C) the ATP hydrolase state. For the sake of simplicity, the proton-motive force

(Δμ¯H+)
‘sensor’ is shown as an acidic group AH, although the same type of mechanism can easily be constructed by placing a positive charge in the appropriate position. It is proposed that the ‘switch’ mechanism is operated by both
Δμ¯H+
and nucleotides (ATP or ADP in italic type), which are bound at non-catalytic sites. The model predicts at least two different
Δμ¯H+
-dependent conformations for the enzyme. It is worth noting that different forms of F1 knob in submitochondrial particles (collapsed and extended) were observed many years ago (Hatase et al., 1972).

Fig. 2.

Schematic representation of the unidirectional operation of the mitochondrial ATPase/ATP synthase. Only one (out of three functioning αβ-subunit pair) is shown as that which has the γ–αβ catch. The relative positions and molecular masses of the ab2δε complex are not specified. Readers should consult the series of excellent mini-reviews describing this aspect and the dynamics of the different subunits of F1Fo (Pedersen, 1996). The key proposal is that there is an energy-driven switch between three types of transmission: (B) no enzyme turnover, no leak for H+ through the Fo component; (A) the ATP synthase state; and (C) the ATP hydrolase state. For the sake of simplicity, the proton-motive force

(Δμ¯H+)
‘sensor’ is shown as an acidic group AH, although the same type of mechanism can easily be constructed by placing a positive charge in the appropriate position. It is proposed that the ‘switch’ mechanism is operated by both
Δμ¯H+
and nucleotides (ATP or ADP in italic type), which are bound at non-catalytic sites. The model predicts at least two different
Δμ¯H+
-dependent conformations for the enzyme. It is worth noting that different forms of F1 knob in submitochondrial particles (collapsed and extended) were observed many years ago (Hatase et al., 1972).

What are the biological advantages of an enzyme operating differently in the forward and reverse directions? At least two possible answers can be offered. First, this arrangement may increase the efficiency of energy transmission. Second, perhaps more importantly, it makes independent metabolic control of the enzyme activities possible, since different conformations of the protein are likely to be susceptible in different ways to the same (or different) ligands. An analogy is anabolic and catabolic pathways, which are widely exploited for metabolic regulation in cells. Different enzymes, e.g. phosphofructokinase and fructoso-bis-phosphatase, operate in glycolysis and gluconeogenesis; again, isoenzymes, such as different lactate dehydrogenases, participate in the metabolism of the same substrate in different organs or cell compartments. The rapidly expanding concept that considers an enzyme as a molecular machine leads me to believe that ‘the one-way traffic’ principal is also realized at the level of a single enzyme. Several recent findings indicate that such a principal applies to another mitochondrial molecular machine, i.e. complex I in the mitochondrial respiratory chain (Vinogradov, 1998).

Tremendous experimental effort has been made for more than half a century to determine the molecular mechanism of ATP synthesis during oxidative phosphorylation and photophosphorylation. One recent review on the subject contains 284 references, citing almost exclusively papers published since 1990 (Weber and Senior, 1997)! In spite of spectacular progress in the field, it cannot be denied that this problem remains to be solved. The great majority of mechanistic studies in the past have focused on the ATP hydrolysis reaction catalyzed by F1 which, by definition, is a hydrolase. These studies have been performed with the explicit or implicit assumption that the enzyme operates reversibly within the F1Fo-dependent ATP synthase complex. I believe that, in the future, more emphasis will be focused on F1Fo operating in the ‘physiological direction’ (at least for the mammalian enzyme) and that many current dogmas concerning the mechanism of the enzyme will be re-evaluated. Another direction for future studies, which at present is at a very primitive stage, is the physiological regulation of ATP synthesis. It is hard to believe that ‘The ATP synthase – a splendid molecular machine’ (Boyer, 1997) is not the subject of fine metabolic control.

The work carried out in my laboratory was partially supported by the Russian Foundation for Fundamental Research (Grants 96-04-48185, 99-04-48082) and by the Programme for Advanced Schools in Science (Grant 96-15-97822). I thank all my colleagues past and present for many stimulating discussions. The great help of Mr F. Kasparinsky and Dr V. Grivennikova in the preparation of the manuscript is gratefully acknowledged. Thanks are due to Ms Julia Grechneva for her helpful linguistic assistance.

Abrahams
,
J. P.
,
Leslie
,
A. G. W.
,
Lutter
,
R.
and
Walker
,
J. E.
(
1994
).
Structure at 2.8 Å resolution of F1-ATPase from bovine heart mitochondria
.
Nature
370
,
621
628
.
Akimenko
,
V. K.
,
Minkov
,
I. B.
,
Bakeeva
,
L. E.
and
Vinogradov
,
A. D.
(
1972
).
Preparation and properties of soluble ATPase from beef heart mitochondria
.
Biokhimia (USSR)
37
,
348
358
.
Bald
,
D.
,
Amano
,
T.
,
Muneyuki
,
E.
,
Pitard
,
B.
,
Rigaud
,
J.-L.
,
Kruip
,
J.
,
Hisabori
,
T.
,
Yoshida
,
M.
and
Shibata
,
M.
(
1998
).
ATP synthesis by FoF1-ATP synthase independent of noncatalytic nucleotide binding sites and insensitive to azide inhibition
.
J. Biol. Chem.
273
,
865
870
.
Bianchet
,
M. A.
,
Hullihen
,
J.
,
Pedersen
,
P. L.
and
Amzel
,
L. M.
(
1998
).
The 2.8-Å structure of the rat liver F1-ATPase: configuration of a critical intermediate in ATP synthesis/hydrolysis
.
Proc. Natl. Acad. Sci. USA
95
,
11065
11070
.
Böttcher
,
B.
,
Gräber
,
P.
,
Boekema
,
E. J.
and
Lücken
,
U.
(
1995
).
Electron cryomicroscopy of two-dimensional crystals of the H+-ATPase from chloroplasts
.
FEBS Lett.
373
,
262
264
.
Boyer
,
P.
(
1993
).
The binding change mechanism for ATP synthase – some probabilities and possibilities
.
Biochim. Biophys. Acta
1140
,
215
250
.
Boyer
,
P.
(
1997
).
The ATP synthase – a splendid molecular machine
.
Annu. Rev. Biochem.
66
,
717
749
.
Boyer
,
P.
(
1998
).
Energy, life and ATP
.
Nobel Lecture. Biosci. Reports
18
,
97
117
.
Bulygin
,
V. V.
,
Syroeshkin
,
A. V.
and
Vinogradov
,
A. D.
(
1993
).
Nucleotide/H+-dependent change in Mg2+affinity at the ATPase inhibitory site of the mitochondrial F1-FoATP synthase
.
FEBS Lett.
328
,
193
196
.
Bulygin
,
V. V.
and
Vinogradov
,
A. D.
(
1988
).
Three adenine nucleotide binding sites in F1-Fomitochondrial ATPase as revealed by presteady-state and steady-state kinetics of ATP hydrolysis
.
FEBS Lett.
236
,
497
500
.
Bulygin
,
V. V.
and
Vinogradov
,
A. D.
(
1991
).
Interaction of Mg2+with Fo.F1mitochondrial ATPase as related to its slow active/inactive transition
.
Biochem. J.
276
,
149
156
.
Burgard
,
S.
,
Nett
,
J.
,
Sauer
,
H. E.
,
Kagawa
,
Y.
,
Schäfer
,
H.-J.
,
Wise
,
J. G.
,
Vogel
,
P. D.
and
Trommer
,
W. E.
(
1994
).
Effect of magnesium ions on the relative conformation of nucleotide binding sites of F1-ATPases as studied by electron spin resonance spectroscopy
.
J. Biol. Chem.
269
,
17815
17819
.
Cherepanov
,
D. A.
,
Mulkidjanian
,
A. Y.
and
Junge
,
W.
(
1999
).
Transient accumulation of elastic energy in proton translocating ATP synthase
.
FEBS Lett.
449
,
1
6
.
Du
,
Z.
and
Boyer
,
P. D.
(
1989
).
Control of ATP hydrolysis by ADP bound at the catalytic site of chloroplast ATP synthase as related to protonmotive force and Mg2+
.
Biochemistry
28
,
873
879
.
Duncan
,
T. M.
,
Bulygin
,
V. V.
,
Zhou
,
Y.
,
Hutcheon
,
M. L.
and
Cross
,
R.
(
1995
).
Rotation of subunits during catalysis by Escherichia coli F1-ATPase
.
Proc. Natl. Acad. Sci. USA
92
,
10964
10968
.
Fillingame
,
R. H.
(
1997
).
Coupling H+transport and ATP synthesis in FoF1-ATP synthases: glimpses of interacting parts in a dynamic molecular machine
.
J. Exp. Biol.
200
,
217
224
.
Fitin
,
A. F.
,
Vasilyeva
,
E. A.
and
Vinogradov
,
A. D.
(
1979
).
An inhibitory high affinity binding site for ADP in the oligomycin-sensitive ATPase of beef heart submitochondrial particles
.
Biochem. Biophys. Res. Commun.
86
,
434
439
.
Fleury
,
B.
,
di Pietro
,
A.
,
Godinot
,
C.
and
Gautheron
,
D. C.
(
1980
).
Role of magnesium on kinetic parameters of soluble F1-ATPase from pig heart mitochondria
.
Biochimie
62
,
733
737
.
Galkin
,
M. A.
and
Syroeshkin
,
A. V.
(
1999
).
Kinetic mechanism of ATP synthesis catalyzed by the mitochondrial Fo.F1-ATPase
.
Biochemistry (Moscow)
64
,
1176
1185
.
Galkin
,
M. A.
and
Vinogradov
,
A. D.
(
1999
).
Energy-dependent transformation of the catalytic activities of the mitochondrial Fo.F1-ATP synthase
.
FEBS Lett
.
448
,
123
126
.
Gogol
,
E. P.
,
Johnston
,
E.
,
Aggeler
,
R.
and
Capaldi
,
R. A.
(
1990
).
Ligand-dependent structural variations in Escherichia coli F1ATPase revealed by cryoelectron microscopy
.
Proc. Natl. Acad. Sci. USA
87
,
9585
9589
.
Gräber
,
P.
(
1994
).
The H+-ATPase from chloroplasts: energetic of the catalytic cycle
.
Biochim. Biophys. Acta
1187
,
171
176
.
Gresser
,
M. J.
,
Myers
,
J. A.
and
Boyer
,
P. D.
(
1982
).
Catalytic site cooperativity of beef heart mitochondrial F1adenosine triphosphatase. Correlations of initial velocity, bound intermediates and oxygen exchange measurement with an alternating three-site model
.
J. Biol. Chem.
257
,
12030
12038
.
Groth
,
G.
and
Junge
,
W.
(
1995
).
ATP synthase: activating versus catalytic proton transfer
.
FEBS Lett.
358
,
142
144
.
Grubmeyer
,
C.
,
Cross
,
R. L.
and
Penefsky
,
H. S.
(
1982
).
Mechanism of ATP hydrolysis by beef heart mitochondrial ATPase. Rate constants for elementary step in catalysis at a single site
.
J. Biol. Chem.
257
,
12092
12100
.
Grubmeyer
,
C.
and
Penefsky
,
H. S.
(
1981
).
Cooperativity between catalytic sites in the mechanism of action of beef heart mitochondrial adenosine triphosphatase
.
J. Biol. Chem.
256
,
3728
3734
.
Guerrero
,
K. J.
,
Xue
,
Z.
and
Boyer
,
P. D.
(
1990
).
Active/inactive state transition of the chloroplast F1ATPase are induced by a slow binding and release of Mg2+
.
J. Biol. Chem.
265
,
16280
16287
.
Hammes
,
G. G.
and
Hilborn
,
D. A.
(
1971
).
Steady state kinetics of soluble and membrane-bound mitochondrial ATPase
.
Biochim. Biophys. Acta
233
,
580
590
.
Harris
,
D. A.
and
Das
,
A. M.
(
1991
).
Control of mitochondrial ATP synthesis in the heart
.
Biochem. J.
280
,
561
573
.
Hatase
,
O.
,
Wakabayashi
,
T.
,
Hayashi
,
H.
and
Green
,
D. E.
(
1972
).
Collapse and extension of the headpiece-stalk projections in mitochondrial electron transport particles
.
Bioenergetics
3
,
509
514
.
Jault
,
J.-M.
and
Allison
,
W. S.
(
1994
).
Hysteretic inhibition of the bovine heart mitochondrial F1-ATPase is due to saturation of noncatalytic sites with ADP which blocks activation of the enzyme by ATP
.
J. Biol. Chem.
269
,
319
325
.
Jenkins
,
W. T.
,
Marshall
,
M. M.
and
Lewin
,
A.
(
1984
).
Hypothesis – A chemical mechanism for the biosynthesis of ATP involving ion-exchange reactions
.
Arch. Biochem. Biophys.
232
,
496
504
.
Junge
,
W.
(
1970
).
The critical electric potential difference for photophosphorylation
.
Eur. J. Biochem.
14
,
582
592
.
Junge
,
W.
,
Sabbert
,
D.
and
Engelbrecht
,
S.
(
1996
).
Rotatory catalysis by F-ATPase: real-time recording of intersubunit rotation
.
Ber. Bunsenges. Phys. Chem.
100
,
2014
2019
.
Kasahara
,
M.
and
Penefsky
,
H. S.
(
1978
).
High affinity binding of monovalent Piby beef heart mitochondrial adenosine triphosphatase
.
J. Biol. Chem.
253
,
4180
4187
.
Lazdunski
,
M.
(
1972
).
Flip-flop mechanisms and half-site enzymes
. In
Current Topics in Cellular Regulation, vol. 6 (ed. B. L. Horecker and E. R. Stadman), pp. 267–278. New York, London: Academic Press
.
Lutter
,
R.
,
Abrahams
,
J. P.
,
Van raaij
,
M. J.
,
Todd
,
R. J.
,
Lundqvist
,
T.
,
Buchanan
,
S. K.
,
Leslie
,
A. G. W.
and
Walker
,
J. E.
(
1993
).
Crystallization of F1-ATPase from bovine heart mitochondria
.
J. Mol. Biol.
229
,
787
790
.
Matsui
,
T.
,
Muneyuki
,
E.
,
Honda
,
M.
,
Allison
,
W. S.
,
Dou
,
C.
and
Yoshida
,
M.
(
1997
).
Catalytic activity of the α3β3complex of F1-ATPase without noncatalytic nucleotide binding site
.
J. Biol. Chem.
272
,
8215
8221
.
Minkov
,
I. B.
,
Fitin
,
A. F.
,
Vasilyeva
,
E. A.
and
Vinogradov
,
A. D.
(
1979
).
Mg2+-induced ADP-dependent inhibition of the ATPase activity of beef heart mitochondrial coupling factor F1
.
Biochem. Biophys. Res. Commun.
89
,
1300
1306
.
Minkov
,
I. B.
,
Vasilyeva
,
E. A.
,
Fitin
,
A. F.
and
Vinogradov
,
A. D.
(
1980
).
Differential effects of ADP on ATPase and oxidative phosphorylation in submitochondrial particles
.
Biochem. Int.
1
,
478
485
.
Mitchell
,
P.
and
Moyle
,
J.
(
1971
).
Activation and inhibition of mitochondrial adenosine triphosphatase by various anions and other agents
.
Bioenergetics
2
,
1
11
.
Moyle
,
J.
and
Mitchell
,
P.
(
1975
).
Active/inactive state transitions of mitochondrial ATPase molecules influenced by Mg2+, anions and aurovertin
.
FEBS Lett.
56
,
55
61
.
Noji
,
H.
,
Yasuda
,
R.
,
Yoshida
,
M.
and
Kinosita
,
K.
(
1997
).
Direct observation of the rotation F1-ATPase
.
Nature
386
,
299
302
.
Panchenko
,
M. V.
and
Vinogradov
,
A. D.
(
1985
).
Interaction between the mitochondrial ATP synthase and ATPase inhibitor protein
.
FEBS Lett.
184
,
226
230
.
Panchenko
,
M. V.
and
Vinogradov
,
A. D.
(
1989
).
Kinetics of interaction between ATPase and ATPase inhibitor protein in submitochondrial particles
.
Biokhimia (USSR)
54
,
569
579
.
Pedersen
,
P. L.
(
1996
).
(ed.) Minireview series. ATP synthase: Current views about subunit movement during catalysis
.
J. Bioenerg. Biomembr.
28
,
389
457
.
Penefsky
,
H. S.
(
1977
).
Reversible binding of Piby beef heart mitochondrial adenosine triphosphatase
.
J. Biol. Chem.
252
,
2891
2899
.
Penefsky
,
H. S.
(
1985
).
Energy-dependent dissociation of ATP from high affinity catalytic sites of beef heart mitochondrial adenosine triphosphatase
.
J. Biol. Chem.
260
,
13735
13741
.
Pullman
,
M. E.
and
Monroy
,
G. C.
(
1963
).
A naturally occurring inhibitor of mitochondrial adenosine triphosphatase
.
J. Biol. Chem.
238
,
3762
3769
.
Racker
,
E.
(
1977
).
Mechanisms of energy transformations
.
Annu. Rev. Biochem.
46
,
1006
1014
.
Racker
,
E.
(
1979
).
Reconstitution: past, present and future
. In
Membrane Bioenergetics
(ed.
C. P.
Lee
,
G.
Schatz
and
L.
Ernster
), pp.
569
591
.
London, Amsterdam, Don Mills, Ontario, Sydney, Tokyo
:
Addison-Wesley Publishing Company
.
Recktenwald
,
D.
and
Hess
,
B.
(
1977
).
A slow conformation change in the transient state kinetics of soluble ATPase of yeast mitochondria
.
FEBS Lett.
80
,
187
189
.
Recktenwald
,
D.
and
Hess
,
B.
(
1979
).
Pre-steady-state kinetics of nucleotide-triphosphate hydrolysis by mitochondrial F1-ATPase from yeast
.
FEBS Lett.
108
,
257
260
.
Sabbert
,
D.
,
Engelbrecht
,
S.
and
Junge
,
W.
(
1996
).
Intersubunit rotation in active F-ATPase
.
Nature
381
,
623
625
.
Selwyn
,
M. J.
(
1967
).
Preparation and general properties of a soluble adenosine triphosphatase from mitochondria
.
Biochem. J.
105
,
279
288
.
Singh
,
S.
,
Turina
,
P.
,
Bustamante
,
C. J.
,
Keller
,
D. J.
and
Capaldi
,
R.
(
1996
).
Topographical structure of membrane-bound Escherichia coli F1FoATP synthase in aqueous buffer
.
FEBS Lett
.
397
,
30
34
.
Souid
,
A.-K.
and
Penefsky
,
H. S.
(
1995
).
Energetics of ATP dissociation from the mitochondrial ATPase during oxidative phosphorylation
.
J. Biol. Chem.
270
,
9074
9082
.
Syroeshkin
,
A. V.
,
Galkin
,
M. A.
,
Sedlov
,
A. V.
and
Vinogradov
,
A. D.
(
1999
).
Kinetic mechanism of ATP hydrolysis by the mitochondrial F1.Fo-ATPase: Mg2+ requirement for Mg.ATP hydrolysis
.
Biochemistry (Moscow)
64
,
1128
1137
.
Syroeshkin
,
A. V.
,
Vasilyeva
,
E. A.
and
Vinogradov
,
A. D.
(
1995
).
ATP synthesis catalyzed by the mitochondrial F1-FoATP synthase is not a reversal of its ATPase activity
.
FEBS Lett.
366
,
29
32
.
Vasilyeva
,
E. A.
,
Fitin
,
A. F.
,
Minkov
,
I. B.
and
Vinogradov
,
A. D.
(
1980
).
Kinetics of interaction of adenosine diphosphate and adenosine triphosphate with ATPase of bovine heart submitochondrial particles
.
Biochem. J.
188
,
807
815
.
Vasilyeva
,
E. A.
,
Minkov
,
I. B.
,
Fitin
,
A. F.
and
Vinogradov
,
A. D.
(
1982a
).
Kinetic mechanism of mitochondrial adenosine triphosphatase. ADP-specific inhibition as revealed by the steady-state kinetics
.
Biochem. J.
202
,
9
14
.
Vasilyeva
,
E. A.
,
Minkov
,
I. B.
,
Fitin
,
A. F.
and
Vinogradov
,
A. D.
(
1982b
).
Kinetic mechanism of mitochondrial adenosine triphosphatase. Inhibition by azide and activation by sulphite
.
Biochem. J.
202
,
15
23
.
Vasilyeva
,
E. A.
,
Panchenko
,
M. V.
and
Vinogradov
,
A. D.
(
1989
).
Interaction of ATPase and natural protein inhibitor in submitochondrial particles under the condition of ΔμH+ generation
.
Biokhimia (USSR)
54
,
1490
1498
.
Vinogradov
,
A. D.
(
1984
).
Catalytic properties of the mitochondrial ATP-synthase
.
Biokhimia (USSR)
49
,
1220
1238
.
Vinogradov
,
A. D.
(
1998
).
Catalytic properties of the mitochondrial NADH–ubiquinone oxidoreductase (Complex I) and pseudo-reversible active/inactive enzyme transition
.
Biochim. Biophys. Acta
1364
,
169
185
.
Vinogradov
,
A. D.
,
Vasilyeva
,
E. A.
and
Evtushenko
,
O. A.
(
1985
).
Kinetics of ATP synthesis/hydrolysis by the mitochondrial ATP-synthase complex
. In
Proceeding of the 16th FEBS Congress, VNU Science Press Part B
, pp.
291
299
.
Weber
,
J.
and
Senior
,
A. E.
(
1997
).
Catalytic mechanism of F1-ATPase
.
Biochim. Biophys. Acta
1319
,
19
58
.
Wilkens
,
S.
,
Dahlquist
,
F. M.
,
Mcintosh
,
L. P.
,
Donaldson
,
L. W.
and
Capaldi
,
R. A.
(
1995
).
Structural features of the subunit of Escherichia coli ATP synthase determined by nmr spectroscopy
.
Nature Struct. Biol.
2
,
961
967
.
Yalamova
,
M. V.
,
Vasilyeva
,
E. A.
and
Vinogradov
,
A. D.
(
1982
).
Mutually dependent influence of ADP and Pion the activity of mitochondrial adenosine triphosphatase
.
Biochem. Int.
4
,
337
344
.
Yamada
,
E. W.
and
Huzel
,
N. J.
(
1988
).
The calcium-binding ATPase inhibitor protein from bovine heart mitochondria
.
J. Biol. Chem.
263
,
11498
11503
.
Yasuda
,
R.
,
Noji
,
H.
,
Kinosita
,
K.
and
Yoshida
,
M.
(
1998
).
F1-ATPase is a highly efficient molecular motor that rotates with discrete 120 ° steps
.
Cell
93
,
1117
1124
.
Yokoyama
,
K.
,
Muneyuki
,
E.
,
Amano
,
T.
,
Mizutani
,
S.
,
Yoshida
,
M.
,
Ishida
,
M.
and
Ohkuma
,
S.
(
1998
).
V-ATPase of Thermus thermophilus is inactivated during ATP hydrolysis but can synthesize ATP
.
J. Biol. Chem.
273
,
20504
20510
.
Yoshida
,
M.
and
Allison
,
W. S.
(
1983
).
Modulation by ADP and Mg2+ of the inactivation of the F1-ATPase from the thermophilic bacterium, PS3, with dicyclohexylcarbodiimide
.
J. Biol. Chem.
258
,
14407
14412
.
Zhou
,
J.-M.
and
Boyer
,
P. D.
(
1992
).
MgADP and free Pias the substrates and Mg2+requirement for photophosphorylation
.
Biochemistry
31
,
3166
3171
.
Zhou
,
J.-M.
and
Boyer
,
P. D.
(
1993
).
Evidence that energization of the chloroplast ATP synthase favors ATP formation at the tight binding catalytic site and increases the affinity for ADP at another catalytic site
.
J. Biol. Chem.
268
,
1531
1538
.