Skip to main content
Open Biology logoLink to Open Biology
. 2019 Apr 10;9(4):180221. doi: 10.1098/rsob.180221

An update of the chemiosmotic theory as suggested by possible proton currents inside the coupling membrane

Alessandro Maria Morelli 1,, Silvia Ravera 2, Daniela Calzia 1, Isabella Panfoli 2
PMCID: PMC6501646  PMID: 30966998

Abstract

Understanding how biological systems convert and store energy is a primary purpose of basic research. However, despite Mitchell's chemiosmotic theory, we are far from the complete description of basic processes such as oxidative phosphorylation (OXPHOS) and photosynthesis. After more than half a century, the chemiosmotic theory may need updating, thanks to the latest structural data on respiratory chain complexes. In particular, up-to date technologies, such as those using fluorescence indicators following proton displacements, have shown that proton translocation is lateral rather than transversal with respect to the coupling membrane. Furthermore, the definition of the physical species involved in the transfer (proton, hydroxonium ion or proton currents) is still an unresolved issue, even though the latest acquisitions support the idea that protonic currents, difficult to measure, are involved. Moreover, FoF1-ATP synthase ubiquitous motor enzyme has the peculiarity (unlike most enzymes) of affecting the thermodynamic equilibrium of ATP synthesis. It seems that the concept of diffusion of the proton charge expressed more than two centuries ago by Theodor von Grotthuss is to be taken into consideration to resolve these issues. All these uncertainties remind us that also in biology it is necessary to consider the Heisenberg indeterminacy principle, which sets limits to analytical questions.

Keywords: ATP synthesis, mitochondria, oxydative phosphorylation, chemiosmotic theory, FoF1-ATP synthase

1. Introduction

The ‘chemiosmotic theory’ formulated by Mitchell [1], a researcher with an Anglo-Saxon training in chemistry, dates back more than 50 years. The theory has universally been accepted, although it immediately raised several controversies, which lasted until today. An upgrading of the chemiosmotic theory appears necessary in light of the enormous progress of bioanalytic techniques defining the fine structure of the macromolecular complexes involved in oxidative phosphorylation (OXPHOS) [27], notably studies on complex I (NADH: ubiquinone oxidoreductase) and complex IV (cytochrome c oxidase) [2,3,8]. These data allow further insight into the proton pathway, a key issue of the theory. The actual proton path across the membrane, the putative proton concentrations on either sides of the membrane and the consequent membrane potential have been the subjects of countless studies. In all evidence, it appears that a free proton osmosis would be impossible, it being a quantum particle that binds to water forming hydronium ions (H3O+). In fact, any free proton in the membrane would quickly be drained by the aqueous phase, releasing the energy associated with the solvation process, to the detriment of the membrane. Also, free protons display a huge destructive force on any biological membrane they pass through. Accordingly, some membrane transporters (such as the potential-dependent proton pump Hv1) are designed specifically to prevent the proton destructive force [9]. Also, a number of reports argue that protons accumulating onto the respiring membrane never reside in the aqueous phase [1013].

Therefore, in this work, we hypothesize that the coupling modality between the electron transport chain and the protonic movement could happen inside the membrane to prevent a proton release, opening new scenarios to explain the basic mechanisms of aerobic metabolism. In other words, in this review, we debate the possibility to update the chemiosmotic theory and unravel the role of local processes in the coupling. This may help in developing new strategies for innovative research centred on cellular bioenergetics.

2. The chemiosmotic theory and FoF1-ATP synthase

The basic formulation of Mitchell's theory is schematically depicted in figure 1, where ATP synthase was also indicated, differently from the original release of 1961, where necessarily it was not depicted. At the time ATP synthesis was attributed to the membrane as a whole, in the form of a generic subtraction of H+ and OH to ADP and orthophosphate to form ATP.

Figure 1.

Figure 1.

Schematic of the 1961 Mitchell chemiosmotic theory. A delocalized coupling is depicted among protons extruded by the electron transport chain (ETC) and ATP synthesis. The overall process is arbitrarily divided in the two phases: the ‘RedOx coupling’, in which the proton movement is operated by the ETC, and the ‘proton coupling’, in which proton movement is coupled with ATP synthesis, by FoF1-ATP synthase.

The experimental data in support of the theory came successively and are reported in literature as a huge amount of contributions. Reviews have been published and we refer to them for a complete documentation [14,15], only the most significant issues being mentioned here. Particularly influential were the data produced in 1966 by A. T. Jagendorf & E. Uribe in the famous ‘acid bath experiment’ [16]. They obtained an ATP synthesis inducing a transmembrane leap of pH in chloroplasts in vitro. In the same year, Y. Kagawa & E. Racker ascertained that the synthesis of ATP occurred on the so-called ‘spheres’ referred to as F1 subunits of the ATP synthase [17]. Since then the basic contribution of FoF1-ATP synthase (ATP synthase) to the OXPHOS became clear.

Developments in the molecular knowledge regarding ATP synthase have been comprehensively addressed in many reviews [1820]. In summary, we can say that the theory is based on three basic postulates:

  • (1)

    an electron transport chain, providing the energy for H+ transfer from one side to other side of the membrane;

  • (2)

    ATP synthase, synthesizing/hydrolysing ATP through H+ translocation; and

  • (3)

    impermeability of the inner mitochondrial membrane to ionic species thereby including protons.

The basic requirement for the OXPHOS is a coupling between redox processes, proton translocation and ATP synthesis. Such global coupling can arbitrarily be divided into two distinct phases: a coupling between the oxidation-reductive process and the protonic translocation, referred as ‘RedOx coupling’ or ‘first coupling’ (see figure 1, left; see also recent review in [21]) and the coupling between protons accumulated on the p-side of the membrane moving to the n-side through the ATP synthase, which determines the synthesis of ATP, here referred as ‘proton coupling’ or ‘second coupling’ (see right side of figure 1). Considerable attention was devoted in the 1980s and 1990s to clarifying the structural–functional details of the respiratory complexes (I, II, III and IV) and ATP synthase (complex V). Important was the study of respiratory complexes organized in supercomplexes [2224] with the demonstration that the loss of their aggregation leads to an increase in the production of reactive oxygen species [24,25]. The possible participation of complex V to supercomplexes has never been demonstrated [23]. The study of supercomplexes has also benefited from extraordinary surveys carried out on X-rays [26].

By contrast, the ‘proton coupling’ or ‘second coupling’ (see figure 1, right) appears to be the most critical passage of the whole OXPHOS process. Literature reports several experiments performed with reconstructed systems [27,28] (i.e. ATP synthase incorporated into phospholipid vesicles) carried out about fifteen years after Mitchell's hypothesis. Vesicles obtained from the membranes from the purple Halobacterium salinarum synthesized ATP as a result of illumination, and it was thus demonstrated that proton movements through the membrane support the synthesis of ATP. In 1977, N. Sone et al. observed that ATP synthase purified by Thermophilic bacterium, inserted in artificial membranes, was able to synthesize ATP thanks to a transient shift in membrane potential (ΔΨ) induced by valinomycin, allowing rapid passage of K+ ions across a membrane on the sides of which different salt concentrations were set [28]. These experiments demonstrated that proton translocation is the crucial step for the ‘proton coupling’ between protonic movement and ATP synthesis. On this general topic, pivotal is the minireview by Junge [29, p. 197], which on one hand highlights the versatility of the ATP synthase, a nano-machine ‘unique in converting electrochemical, mechanical and chemical forms of energy’ and on the other hand points out that there is still much to be understood about the chemical–physical basis of such process.

3. Controversies about the chemiosmotic theory

A long struggle was necessary for the chemiosmotic theory formulated by Mitchell [1] to be widely accepted. The controversy, central to the history of bioenergetics for more than half a century, appears tackled by more than 200 articles and to have lasted until the most recent years. G. F. Azzone, many years ago (1972), published the manuscript ‘Oxidative phosphorylation, a history of unsuccessful attempts: is it only an experimental problem?’ [30], which already highlighted the non-convincing parts of the theory, wishing for answers from the fine analysis of the macromolecular structures involved in chemiosmosis.

The harshest criticisms came from J. Prebble, who emphasized the lack of experimental data in support of the theory [31]. W. Junge effectively described, in his recent review ‘Half a century of molecular bioenergetics’, the chronicle of the dispute, which even took harsh tones [32]. S. Brown & D. C. Simcook [33] considered the motivations that have convinced the scientific community to accept the chemiosmotic theory. The authors note that: ‘science shows tremendous resistance to change and it takes extraordinary perseverance to persuade the community’ [33, p. 178].

A controversial issue was the correlation between the membrane potential (ΔΨ) and the protonmotive force, often considered equivalent entities [34]. It was postulated that a ΔΨ with positive charge on the external p-side of the internal mitochondrial membrane and negative on the n-side in contact with mitochondrial matrix (figure 1) would let protons enter the Fo rotor that synthesizes ATP in the matrix thanks to its mechanical connection with the F1 moiety. Protons would gather across the coupling membrane like chemical ions, creating a driving force for FoF1-ATP synthase to synthesize ATP, realizing the ‘proton coupling’. However, the yield in ATP poorly correlates with bulk-to-bulk membrane potential [35,36], questioning the basics of chemiosmotic theory [37].

The award of the 1978 Nobel Prize for chemistry to Peter Mitchell cooled the dispute, but not definitively. In fact, in 1979, there was a heated confrontation published in Trends in Biochemical Sciences between H. Tedeschi, who disproved the idea that the metabolic activity of mitochondria could contribute to membrane potential, and H. Rottenberg, which instead defended Mitchell's theory [38]. The original theory provides a protonated ‘delocalized coupling’ (as depicted in figure 1), as opposed to a ‘localized coupling’ supported by Williams [39]. Lee [40] reported a rigorous chemical/physical experiment in favour of localized coupling, demonstrating furthermore that the thylakoid membrane can be a ‘proton capacitor’. The putative existence of a proton capacitor is a matter of great importance, and later, Saeed & Lee [10] showed that protons can actually accumulate on the membrane surface even though they never reside in the aqueous phase. Moreover, concerning the experimental verification of the ‘proton coupling’, a recent elegant investigation in HeLa cells, bioengineered with green fluorescent protein as pH indicator inserted in respiratory complex III and in Fo moiety of ATP synthase, points to a localized coupling [41].

A report entitled ‘Proton migration along the membrane surface and retarded surface to bulk transfer’ by Heberle et al. [11] interestingly reconciles the two visions, providing proof that proton transfer from a proton generator (bacteriorhodopsin) to an acceptor (water-soluble pH indicators) is faster if occurring on the membrane rather than when protons are released in the aqueous bulk. Ferguson [12] emphasized Heberle et al.'s experiments, concluding that the delocalized coupling and lateral proton transfer (localized coupling) between the proton generator and user occurs very rapidly on the membrane, as compared with the slower and transversal passage through the aqueous bulk. In this context, the recent paper from C. von Ballmoos's group observed that ΔΨ and ΔpH are equivalent for the coupling with ATP synthase [34]. A primary role for membrane buffering on proton mobility in general can be hypothesized [42]. The experimental data [43,44] showing a close thermodynamic correlation between valinomycin-induced ΔΨ and ATP synthesis in reconstituted systems are very important, but it seems plausible that they induce a transmembrane protonic flow that probably differs from the path in a native environment.

Moreover, the eminent English chemist R. J. Williams clearly rejected the hypothesis of the accumulation of protons from p-side: ‘the p-phase corresponds to the infinitely extended external space. If protons are extruded into this “Pacific Ocean”, they would be diluted and the entropic component of the pmf would be lost’ [13, p. 18]. R. Williams observed that ‘protons in the membrane rather than an osmotic trans-membrane gradient of protons were required to drive ATP formation’, based on a series of considerations that excluded the presence of free protons from p-side [39, p. 123]. An elegant demonstration of Williams's localized coupling hypothesis came in 1976 [45] by an experiment in which purified ATP synthase was added to the octane–water interface. It was observed that protons accumulate in octane, a Brønsted acid, leading to ATP synthesis by ATP synthase. These data have also been recently confirmed [46]. Eighteen years later, it was reported that ‘our results suggest that protons can efficiently diffuse along the membrane surface between a source and a sink (for example H+-ATP synthase) without dissipation losses into the aqueous bulk’ [11, p. 379]. From all the cited data it can be concluded that protons (or more likely protonic currents) are confined into the membrane, while proton exit from the membrane is to be considered only as a fallback way of escape, mostly via in vitro reconstituted conditions.

A direct measurement of membrane potential of the mitochondrial inner membrane with microelectrodes was only be accomplished by H. Tedeschi, who showed the existence of a positive inside and negative outside mitochondrial ΔΨ [47,48]. Such potential (contrary to the canonical one) interestingly coincides with that calculated on the basis of the ionic species present on the membrane sides [49]. Clearly, knowledge of the entity and especially the sign of this potential is fundamental for understanding the basic functioning of chemiosmosis, as emerges from the already mentioned historical dispute between Tedeschi and Rottenberg [38].

To measure ΔΨ, laboratory tests currently use lipophilic fluorescent compounds whose response is considered to be related to ΔΨ. However, tests conducted with rhodamine have cast doubts on such correlation [50] since these indicators inhibited the mitochondrial respiration, so they disturb the system. As such compounds dissolve into the membranes, they may reflect the membrane behaviour; in fact they inhibit a membrane intrinsic process (i.e. the OXPHOS), but do not interfere with ΔΨ. Surely, in vitro, proton passage across membranes can be forced with rapid movements of potassium ions by addition of valinomycin [28,51]. A laboratory procedure using valinomycin and also nigericin has been widely used to create a transient ΔΨ operating a delocalized coupling linking ΔΨ and ATP synthesis, but this does not exclude that in the native membranes a localized coupling would operate independently of ΔΨ.

4. A crucial issue: the membrane permeability to protons

In a previous study (1986), M. Zoratti et al. highlighted the uncertainties of the proton cycle [52]. In the same year Grzesiek & Dencher [53] showed that the phospholipid membranes are intrinsically permeable to protons. Data show that phospholipid membranes, normally impermeant to ionic solutes (transversal or permeability coefficients varying between 10−12 to 10−14 cm s−1), exhibit a significant proton permeability, varying from 10−3 to 10−9 cm s−1. Such variability may be justified by a buffering capacity of the membranes for protons: proton diffusion value could depend on the higher or lower degree of pre-existing protonation. Recently, the proton leak through lipid bilayers was modelled as a concerted mechanism [5355]. Tepper & Voth [54] provided a theoretical interpretation of proton permeability, based on the formation of transient membrane spanning aqueous solvent structure. High proton permeability has also been confirmed in liposomes, independently form their phospholipid composition [56]. It is clear that a high degree of permeability to protons is per se in contrast to the third of the aforementioned basic postulates of the chemiosmotic theory. With regard to the relationship between the membrane and aqueous phases, many observations confirm the existence of a layer of water molecules on the two sides of the membrane, which to some extent isolates it from the aqueous phases present on its two sides [5759]. In particular, E. Deplazes et al. observed that at the membrane level, H3O+ forms strong and long-lived hydrogen bonds with the phosphate and carbonyl oxygens in phospholipids [60].

5. Proton solvation

A central issue is the actual chemical species of the proton: free, or in the form of H3O+? This depends on the phase in which the proton is located. Protons possess peculiar chemical properties, being essentially an atomic nucleus. Free protons do not exist in the aqueous phase, being solvated to H3O+, from which the extraction of a proton would be virtually impossible. In fact, in the transition from H3O+ to free proton a strong energy barrier higher than 500 meV must be overcome. The desolvation barrier [61] corresponds to the enormous amount of 262 400 Cal mol−1 [62]. An immense literature exists on the subject [40,61,63,64]. An interesting report [65], not sufficiently taken into account, calculated the number of free protons (actually in the form of H3O+) in the volume of a mitochondrion, whose the order of magnitude is femtolitres. Starting from basic physical chemical data (Avogadro number, ionic water product, mathematical pH expression and mitochondrial volume), this study demonstrated that free protons in a mitochondrial periplasmic space are too few (fewer than 10) to support any process dependent on proton translocation in the aqueous bulk across the membrane and absolutely inadequate to support the thousands of ATP synthase molecules present in a mitochondrion. Moreover, the pH value inside the mitochondrion was shown to differ by 0.5 units from what was previously believed [66]. Indeed, the huge energy associated with proton solvation would have a negative consequence: a free membrane proton would quickly be ‘sucked’ by the near aqueous phase, releasing the huge energy associated with the solvation process, to the detriment of the membrane.

6. Grotthuss mechanism and proton translocation through the membranes

A putative mechanism for proton diffusion was hypothesized more than two centuries ago by von Grotthuss [67] and is synthetically explained by S. Serowy and colleagues: ‘Proton diffusion according to the Grotthuss mechanism occurs much faster than molecular diffusion because it is uncoupled from the self-diffusion of its mass’ [68, p. 1031]. Protons would not diffuse as a mass, rather as a charge, the latter moving between water molecules or protonable groups of suitable macromolecules of the membranes. The Grotthuss mechanism allows better understanding of the possible ways in which protons or species derived from them move in biological systems. T. E. DeCoursey published a review [68] which exhaustively analyses proton transfer pathways in water and biological membranes [9], including Hv1 channels that specifically transfer protons in aqueous phase, therefore actual acidity from one side of the membrane to the other. The mechanism of the voltage gated proton channel Hv1 is not as yet resolved, as emblematically stated by the title of the paper: ‘The voltage-gated proton channel: a riddle, wrapped in a mystery, inside an enigma’ [69]. Two mechanisms have been proposed for Hv1, as schematically depicted in figure 2. On the left side of figure 2 is the so-called ‘frozen water’ mechanism, in which the channel traps one or more molecules of water allowing protons to pass through with Grotthuss-style proton hopping, as in the typical case of Gramicidin [70]. On the right side of figure 2 is a passage of protons with protonation/deprotonation of amino acid side-chains, which would realize the so-called ‘proton wire’ already proposed in the paper by Nagle & Morowitz [71]. The topic was consolidated successively by Nagle & Tristram-Nagle [72]. T. E. DeCoursey in a debate recently published in the Journal of Physiology supports the mechanism shown in the right part of figure 2 (i.e. proton wires through the membrane [73]), while Bennett & Ramsey [74] support a mechanism of passage through water molecules as schematized on figure 2, left. Interestingly, both mechanisms are based on the Grotthuss proton movement between (i) water molecules in the case of the water channel (on the left) and (ii) amino acid side chains in the case of protons wires (on the right).

Figure 2.

Figure 2.

Mechanism for H+ transfer through the membrane by HV1. On the left is the water channel model: the water molecules allow protons to pass through with Grotthuss-style H+ hopping. On the right is the proton wire model: a charge migration occurs (through with Grotthuss-style H+ hopping) on polar groups of side chains of amino acids of HV1.

The proton movement in membranes (both biological and artificial) has been analysed by many rigorous chemical/physical studies. The article ‘“Proton holes” in long-range proton transfer reactions in solution and enzymes: a theoretical analysis’ shows that other compounds in addition to water are involved in the ‘proton hopping’ [75], and is interesting that a quantum-mechanical approach is applied [76]. The article ‘Grotthuss mechanisms: from proton transport in proton wires to bioprotonic devices’ presents devices such as proton diodes, transistors, memories and transducers, semiconductor electronic devices that use the Grotthuss mechanism [77].

Notably, Hv1 only allows the passage of protons in the form of hydronium ions, balancing their concentration between two aqueous compartments separated by a membrane. This property does depend on the membrane potential ΔΨ, similarly to other ion membrane devices abundant in biological membranes (for example the Na+ and K+ voltage gated channels), acting on the conformation of Hv1. It appears therefore that to carry protons through the membrane there are ad hoc structures that deeply differ from the respiratory complexes, both for their finality and the molecular mechanism. From this and other evidence it can be concluded that the native proton movements in the respiring membranes, when there is no need for acidification of the milieu on one side of a membrane, must take place entirely inside the membrane [78]. As far as the respiratory complexes are concerned, on the other hand, it is theorized that the proton and membrane potential movements are mutually dependent, in that the pumping of protons would generate the membrane potential, which is impossible for thermodynamic considerations that we will elaborate later on.

Moreover, this comparison between proton movement in support of the OXPHOS and the actual protonic movement in nature sheds light on the fact that when protons are really transferred through a membrane (i) they are never in the form of free protons and (ii) they are subject to the Grotthuss-style proton hopping. Hence the need for chemiosmotic theory to be updated in the light of all this emerges.

7. Inhibition of ATP hydrolysis: an open topic

ATP synthase inherent catalytic properties differentiate it from the vast majority of other enzymes/catalysts. In fact, it transfers energy to the reaction it catalyses, therefore it influences the equilibrium of the reaction while the other enzymes are irrelevant with respect to it. This peculiar property also requires that the activity of ATP synthase be subjected to proper control, as once the coupling with the oxide/reduction systems is lost, the nanomotor itself rapidly hydrolyses ATP, resulting fatal for cell survival. The significant free energy release (ΔG° = −7300 cal mol−1) involved in ATP hydrolysis thus establishes a clear directionality towards the process of hydrolysis itself. It should be emphasized that for most enzymes, the direction of the reaction depends on the concentration of the reactants/substrates, which obviously does not occur for ATP synthase.

To inhibit the reversal of ATP synthase in uncoupled conditions, the action of the inhibitor protein IF1 (UniProtKB: P01096) is essential, particularly in the brain [79], which is relatively devoid of energy reserves (glycogen, triglycerides), where anoxia would lead to the tumultuous ATP hydrolysis. IF1 knock down in HeLa cells [80] equipped with a FRET sensor for measuring the intracellular ATP concentration, surprisingly, did not affect cell viability or mitochondrial morphology, even though the cellular ATP concentration decreased by about a third. Moreover, recently it was reported that deleting the IF1-like ζ subunit from Paracoccus denitrificans has little influence on ATP hydrolysis by ATP synthase [81], confirming that the inhibition of the hydrolysis of ATP by uncoupled ATP synthase does not depend only on IF1 and that the subject requires further investigations. Also, in bacteria, when the concentration of the ATP reaches the physiological value, the ɛ subunit inhibits ATP synthase by binding to the c ring, and it has been proposed as a target for the design of anti-tuberculosis drugs [82].

8. Local processes for the coupling inside the respiring membranes

Having established the clear divergence between the pathways of the solvated proton and of the proton alone, detailed molecular structural data on the respiratory complexes able to handle protons (i.e. complex I, III and IV) are now available thanks to the progress of X-ray analysis and of cryo-microscopy [27]. Excellent investigations are available on complex I; here for simplicity we only mention the studies from L. Sazanov and collaborators [2,3].

Numerous structural X-ray studies were conducted on complex I from Escherichia coli [4], Thermus thermophilus [2] and mammalian ovine (Ovis aries) mitochondria [5], and with cryo-microscopy on complex I from Bos taurus [6,7]. The complexity of the macromolecular aggregation of complex I is impressive: in mammals it is formed by as many as 45 polypeptides and its assembly needs an unknown number of chaperones, so indispensable that the impairment of just one of them (B17.2 L) causes a progressive encephalopathy [83]. Ogilvie et al. [83, p. 2784] state that ‘results demonstrate that B17.2 L is a bona fide molecular chaperone that is essential for the assembly of complex I and for the normal function of the nervous system.’

We may seek for the proton plausible pathways inside the respiring membrane, even if it is a plasma-membrane. As only five alpha-helices have been found in the complex I structure, in order to allow for proton translocation, it was necessary to postulate the existence of two hemi channels: one from the matrix (n) side to the centre of the respiratory complex, referred here to as the ‘proton entrance hemi channel’, and the other from the centre to the periplasmatic (p) side, here indicated as ‘proton exit hemi channel’. However, only the latter was well identifiable in complex I [4]. Instead, the entry pathway has not been identified with certainty, so we only can talk about putative pathways labelled with ‘?’ [4]. Furthermore, it clearly emerges from the X-ray studies that there is an obvious proton tunnelling at the centre of the complex I. Comprehensive studies on the protonic movement inside complex I have been carried out by the Helsinki Bioenergetic Group of M. Wikström, which highlighted uncertainty margins on the stoichiometry of protonic extrusion that appear closer to 3 H+/2e [84] instead to the classic 4 H+/2e. Also, in the review of Verkhovskaya & Bloch [85] (of Helsinki Bioenergetic Group), four mechanisms for proton translocation are proposed and the ‘proton entrance half channel’ is not identified with certainty, while the ‘proton exit half channel’ is clearly identifiable. Emblematically, on the website of the Helsinki Bioenergetic Group it is written (www.biocenter.helsinki.fi/bi/hbg/cl/complex_I_main.html) that ‘the mechanism of proton transfer in complex I remains completely enigmatic’.

As far as complex IV (cytochrome c oxidase) is concerned, the proton translocation of has been studied in depth [86]. In their recent review, M. Wikström & V. Sharma talk of an ‘anniversary’: ‘Proton pumping by cytochrome c oxidase—A 40 year anniversary’ [87]. Among the many works cited in this review the Chemical Review article of the M. Wikström group stands out [8]. It goes into the details of the possible molecular processes carried out by complex IV [8], thereby including proton translocation. The topic is complex, and more putative pathway of protons are well developed in the review, which we cannot here detail here. For the ‘proton entrance half channel’, for each molecule of oxygen reduced to water they need 4 H+, and it is hypothesized that another 4 H+ (for a total of 8) enter through this channel. It seems unlikely that there exists equivalence between protons that exist as particles and link to water molecules and protons that should move as a charge, according to the Grotthuss mechanism.

9. Proposal for a localized complex I-ATP synthase coupling

Taking into account what is reported above, it is possible to trace a plausible proton pathway within the respiring membrane. In 2006 a direct proton transfer was proposed to couple the respiratory pathway of complexes I, III, IV with ATP synthase [88]. However, in a recent review [21] the concept of transmembrane proton motor force to move the ATP synthase is reinforced, although it is noted that many aspects of coupling are not yet clarified. Clear-cut consideration was proposed some years ago (1991) by Akeson & Deamer [89] about speed of proton translocation through putative proton channel as a limiting step for ATP synthesis by ATP synthase. For the sake of simplicity, we only examine the coupling between the respiratory complex I and the ATP synthase. The existence of a proton pathway at the centre of complex I is quite clear, and we can hypothesize that the protons are sent to the well-identified ‘exit half channel’, as shown in figure 3. The proton donor at the centre of complex I has already been tentatively identified in the phospholipid cardiolipin (CL) [78], essential for OXPHOS. However, for example, phosphatidylethanolamine (PE) can effectively sustain the functioning of OXPHOS [90]. Data from C. von Ballmoos's group [91,92] unconventionally clarify such a role of phospholipids. In fact, pure phosphatidylcholine (PC) is excellent for the coupling, increasing when the membrane is formed by PC + PE. By contrast, coupling is dramatically inhibited if the membrane is formed by PC + CL, incredibly diverging from the traditional role assigned to CL. This research is also important because experiments are performed with a reconstructed system, more adherent to the ‘H+/ATP coupling’ or ‘second coupling’ (figure 1). Here, the driving force that feeds ATP synthase was not the rapid transfer of K+ generated by valinomycin, a widely used method, but the bo3 oxidase of Escherichia coli, analogous to the respiratory complex I of vertebrates. It emerges that all membrane phospholipids can act as mobile proton transporters, hindering proton relocation in the near-aqueous medium, so that these are never free. The exergonic process of proton solvation would release an impressive amount of energy (262 400 Cal mol−1) [62], which, if not transferred on a generic acceptor, would generate heat devastating not only the membrane but also the cell integrity. Complex I would transfer protons to the p side of the membrane, but these would not be dispersed in the aqueous bulk. In fact it is well documented that there exists a barrier of water molecules attached to the membrane, determining the lateral displacement of protons on the phosphate heads of phospholipids [68,93,94] to meet a sink that is the ATP synthase subunit which, through a proton wire [95,96], would lead the proton to the rotor (figure 3), accommodating on the central glutamic or aspartic residue (depending on the species) of the c subunit, which in variable numbers from 8 to 15 make up the rotor Fo [97]. As Fo rotates it is conceivable that the proton returns to the centre of complex I, a highly hydrophobic environment, thus closing the protonic circuit. This last passage, from the rotor exit to the centre of the respiratory complex, appears plausible for the operativity of the Brownian motion (diffusion) of particles in a highly anisotropic environment that can occur efficiently [45,98] and for a theoretically possible direct passage of the proton exiting the subunit at the entrance to the complexes. Moreover the investigations of the C. von Ballmoos group have produced exhaustive studies with reconstructed systems [91,92,99], in which a direct transfer of proton on the p-side of the membrane from complex I to ATP synthase is evident. Their recent paper also demonstrates that proximity to the membrane between the respiratory complex and ATP synthase is required for the ‘proton coupling’ [92]. However, the fact remains that the last part of the protonic circuitry (i.e. from Fo moiety of ATP synthase to respiratory complexes) is only hypothetical.

Figure 3.

Figure 3.

A possible H+ circuit inside respiring membrane. The phosphate groups of phospholipids on both sides of the membrane are shown by brown ellipsoids. The image proposes that the H+ (red dotted line) are transferred to the Glu 58 (E58) at the centre of subunit c through subunit a of ATP synthase by proton tunnelling. H+ would flow from the periplasmic side, always bound to phospholipid heads. This can be arranged in each layer of the membrane.

In this controversial scenario, a decisive contribution is undoubtedly the sophisticated bioengineering experiment that labelled both complex IV and ATP synthase with proteins of the GFP family, to experimentally observe a local ΔpH triggered by the respiratory substrate galactose [41]. Observations were conducted in cultured HeLa cells. The authors concluded [41, p. 1]: ‘the observed lateral variation in the proton-motive force necessitates a modification to Peter Mitchell's chemiosmotic proposal’. In other words, the experimentally proven lateral proton motive force is in line with the hypothesis of localized coupling.

10. Extramitochondrial oxidative phosphorylation

The chemiosmotic theory [1] as it was formulated envisions a process that can only take place in organelles possessing double membrane systems forming closed compartments to entrap protons, such as mitochondrial cristae, bacteria and thylakoids (here, for the sake of simplicity, we have considered the mitochondrial inner membrane). However, in the last years, several authors have described a functional expression of OXPHOS machinery in cellular districts devoid of mitochondria [100,101], suggesting a possible build-up of a transversal proton gradient across the membrane, out of mitochondria. In particular, an extramitochondrial OXPHOS has been described in rod outer segment (OS) discs [102105], myelin sheath [79,106110], cell plasma membrane [111119], platelets [120], and extracellular vesicles shedding from cells such as exosomes and microvesicles [121,122], which seem to carry an unsuspected metabolic signature [121,123]. Notably, since the plasma membrane potential is positive on the outside and negative on the inside, it would favour ATP hydrolysis rather than its synthesis. The extramitochondrial ATP synthesis is in line with the postulated independence of ATP synthase activity from the membrane potential, and therefore it is plausible that this synthesis of ATP depends on the proton intramembranous coupling.

11. Conclusion

The impressive amount of experimental data cited appears globally in contrast with the above-cited three assumptions underlying the chemiosmotic theory. First of all, it excludes that the protons can accumulate on the coupling membrane surface, whose high permeability would dissipate, since it would correspond to an extreme acidity incompatible with any vital process. Second, since proton diffusion does occur, it is clear that the membrane potential is irrelevant to proton translocation. Third, it can be excluded that the respiratory complexes operate as transmembrane proton transfer from the aqueous bulk, where the proton would exist as hydroxonium ion. This appears to rule out the actual possibility that ATP synthase can overcome the energy barrier, higher than 500 meV [61], required to extract the proton from water, leaving space for the localized coupling hypothesis [63,64], as a dehydration–hydration reaction from hydroxonium ion to free protons would require 262 400 Cal mol−1 [62]. Which are the actual processes acting on the ATP synthase? There is no certainty, as highlighted by the emblematic title of an article by J. Walker: ‘The ATP synthase: the understood, the uncertain and the unknown’ [124].

Indeed, today a constellation of clues leads us to hypothesize the existence of protonic currents internal to the membrane, with the formation of possible circuits travelled by the positive elementary charge, thus realizing a localized coupling that excludes an osmotic nature of the process. To trace this circuit, at least for the possible coupling between respiratory complex I and ATP synthase, it appears realistic that complex I may transfer protons from its central part to the periplasmatic side, allowing them to travel on the membrane surface thanks to the heads of phospholipids [93,94] finally tunnelling inside the a subunit of ATP synthase [95,125] (figure 3). Moreover, several studies show that the membrane is isolated from the aqueous bulk thanks to a layer of water molecules on both sides of the membrane, which consolidates the idea that the membrane is radically distinct and isolated from the liquid phase [59]. Considering the two isolated phases, the only way evolution could pursue to link proton movement to ATP synthesis was a nanomachine connecting the proton movement inside the membrane to the deformation mechanics of the F1 sphere immersed in the aqueous phase. We can reasonably assume that the proton movement inside the membranes occurs as a charge, according to the proton-hopping Grotthuss mechanism, with the establishment of ‘protonic currents’ inside the membrane [126]. In the non-biological field we find a remarkable adherence to this theory for the development of protonic devices (as proton diodes, transistors, memories and transducers) [77]. It is surprising that the biological and the physical–chemical areas have ignored each other. Since in the history of biology the application of physical–chemical methodologies has led to dramatic advances in biology (we may remind the reader that the resolution of DNA structure [127] was obtained with the fundamental application of X-ray crystallography, developed in 1913 by Williams Henry Bragg and his son Lawrence for the study of inorganic crystals), it is desirable that this fusion of knowledge can be realized in the years to come.

Furthermore, the classical mechanical approach cannot be used to approach these currents. In fact, any time there is a movement of charge bound to a mass, the dualism that cannot be assessed by classical mechanics, instead quantum mechanics must be applied, and in this perspective lay the promising recent quantum-mechanical approaches by Riccardi et al. [76] and Ivontsin et al. [125]. There is a need to take into account the Heisenberg indeterminacy principle as Philip Hunter already highlighted in the title of a 2006 paper: ‘A quantum leap in biology: one inscrutable field helps another, as quantum physics unravels consciousness’ [128].

Supplementary Material

Reviewer comments

Acknowledgements

The authors are very grateful to Professor Giorgio Lenaz, Bologna University (Italy), for the critical reading of the manuscript and for the valuable comments and discussions, and Professor Matthew Stephen Hodgart, University of Surrey, for help in revising the manuscript language.

Data accessibility

This article has no additional data.

Competing interests

We declare we have no competing interests.

Funding

We received no funding for this study.

References

  • 1.Mitchell P. 1961. Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature 191, 144–148. ( 10.1038/191144a0) [DOI] [PubMed] [Google Scholar]
  • 2.Baradaran R, Berrisford JM, Minhas GS, Sazanov LA. 2013. Crystal structure of the entire respiratory complex I. Nature 494, 443–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Sazanov LA. 2015. A giant molecular proton pump: structure and mechanism of respiratory complex I. Nat. Rev. Mol. Cell Biol. 16, 375–388. ( 10.1038/nrm3997) [DOI] [PubMed] [Google Scholar]
  • 4.Efremov RG, Sazanov LA. 2011. Structure of the membrane domain of respiratory complex I. Nature 476, 414–420. ( 10.1038/nature10330) [DOI] [PubMed] [Google Scholar]
  • 5.Fiedorczuk K, Letts JA, Degliesposti G, Kaszuba K, Skehel M, Sazanov LA. 2016. Atomic structure of the entire mammalian mitochondrial complex I. Nature 538, 406–410. ( 10.1038/nature19794) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Vinothkumar KR, Zhu J, Hirst J. 2014. Architecture of mammalian respiratory complex I. Nature 515, 80–84. ( 10.1038/nature13686) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Zhu J, Vinothkumar KR, Hirst J. 2016. Structure of mammalian respiratory complex I. Nature 536, 354–358. ( 10.1038/nature19095) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Wikström M, Krab K, Sharma V. 2018. Oxygen activation and energy conservation by cytochrome c oxidase. Chem. Rev. 118, 2469–2490. ( 10.1021/acs.chemrev.7b00664) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Kreuer K-D. 1996. Proton conductivity: materials and applications. Chem. Mater. 8, 610–641. ( 10.1021/cm950192a) [DOI] [Google Scholar]
  • 10.Saeed HA, Lee JW. 2015. Experimental demonstration of localized excess protons at a water-membrane interface. Bioenergetics 4, 1–7. ( 10.4172/2167-7662.1000127) [DOI] [Google Scholar]
  • 11.Heberle J, Riesle J, Thiedemann G, Oesterhelt D, Dencher NA. 1994. Proton migration along the membrane surface and retarded surface to bulk transfer. Nature 370, 379–382. ( 10.1038/370379a0) [DOI] [PubMed] [Google Scholar]
  • 12.Ferguson SJ. 1995. Chemiosmotic coupling. Protons fast and slow. Curr. Biol. 5, 25–27. ( 10.1016/S0960-9822(95)00008-X) [DOI] [PubMed] [Google Scholar]
  • 13.Williams RJ. 1978. The multifarious couplings of energy transduction. Biochim. Biophys. Acta 505, 1–44. ( 10.1016/0304-4173(78)90007-1) [DOI] [PubMed] [Google Scholar]
  • 14.Ernster L, Schatz G. 1981. Mitochondria: a historical review. J. Cell Biol. 91, 227s–255s. ( 10.1083/jcb.91.3.227s) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Hatefi Y. 1985. The mitochondrial electron transport and oxidative phosphorylation system. Annu. Rev. Biochem. 54, 1015–1069. ( 10.1146/annurev.bi.54.070185.005055) [DOI] [PubMed] [Google Scholar]
  • 16.Jagendorf AT, Uribe E. 1966. ATP formation caused by acid-base transition of spinach chloroplasts. Proc. Natl Acad. Sci. USA 55, 170–177. ( 10.1073/pnas.55.1.170) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Kagawa Y, Racker E. 1966. Partial resolution of the enzymes catalyzing oxidative phosphorylation. X. Correlation of morphology and function in submitochondrial particles. J. Biol. Chem. 241, 2475–2482. [PubMed] [Google Scholar]
  • 18.Walker JE, Fearnley IM, Lutter R, Todd RJ, Runswick MJ. 1990. Structural aspects of proton-pumping ATPases. Phil. Trans. R. Soc. Lond. B 326, 367–378. ( 10.1098/rstb.1990.0018) [DOI] [PubMed] [Google Scholar]
  • 19.Stock D, Gibbons C, Arechaga I, Leslie AG, Walker JE. 2000. The rotary mechanism of ATP synthase. Curr. Opin. Struct. Biol. 10, 672–679. ( 10.1016/S0959-440X(00)00147-0) [DOI] [PubMed] [Google Scholar]
  • 20.Junge W, Nelson N. 2015. ATP synthase. Annu. Rev. Biochem. 84, 631–657. ( 10.1146/annurev-biochem-060614-034124) [DOI] [PubMed] [Google Scholar]
  • 21.Papa S, Capitanio G, Papa F. 2018. The mechanism of coupling between oxido-reduction and proton translocation in respiratory chain enzymes. Biol. Rev. Camb. Philos. Soc. 93, 322–349. ( 10.1111/brv.12347) [DOI] [PubMed] [Google Scholar]
  • 22.Schägger H, Pfeiffer K. 2001. The ratio of oxidative phosphorylation complexes I-V in bovine heart mitochondria and the composition of respiratory chain supercomplexes. J. Biol. Chem. 276, 37 861–37 867. [DOI] [PubMed] [Google Scholar]
  • 23.Wittig I, Carrozzo R, Santorelli FM, Schagger H. 2006. Supercomplexes and subcomplexes of mitochondrial oxidative phosphorylation. Biochim. Biophys. Acta 1757, 1066–1072. ( 10.1016/j.bbabio.2006.05.006) [DOI] [PubMed] [Google Scholar]
  • 24.Bianchi C, Genova ML, Parenti Castelli G, Lenaz G. 2004. The mitochondrial respiratory chain is partially organized in a supercomplex assembly. J. Biol. Chem. 279, 36 562–36 569. ( 10.1074/jbc.M405135200) [DOI] [PubMed] [Google Scholar]
  • 25.Lenaz G, Tioli G, Falasca AI, Genova ML. 2016. Complex I function in mitochondrial supercomplexes. Biochim. Biophys. Acta 1857, 991–1000. ( 10.1016/j.bbabio.2016.01.013) [DOI] [PubMed] [Google Scholar]
  • 26.Letts JA, Fiedorczuk K, Sazanov LA. 2016. The architecture of respiratory supercomplexes. Nature 537, 644–648. ( 10.1038/nature19774) [DOI] [PubMed] [Google Scholar]
  • 27.Yoshida M, Sone N, Hirata H, Kagawa Y. 1975. ATP synthesis catalyzed by purified DCCD-sensitive ATPase incorporated into reconstituted purple membrane vesicles. Biochem. Biophys. Res. Commun. 67, 1295–1300. ( 10.1016/0006-291X(75)90167-9) [DOI] [PubMed] [Google Scholar]
  • 28.Sone N, Yoshida M, Hirata H, Kagawa Y. 1977. Adenosine triphosphate synthesis by electrochemical proton gradient in vesicles reconstituted from purified adenosine triphosphatase and phospholipids of thermophilic bacterium. J. Biol. Chem. 252, 2956–2960. [PubMed] [Google Scholar]
  • 29.Junge W. 2004. Protons, proteins and ATP. Photosynth. Res. 80, 197–221. ( 10.1023/B:PRES.0000030677.98474.74) [DOI] [PubMed] [Google Scholar]
  • 30.Azzone GF. 1972. Oxidative phosphorylation, a history of unsuccessful attempts: is it only an experimental problem? J. Bioenerg. 3, 95–103. ( 10.1007/BF01516000) [DOI] [PubMed] [Google Scholar]
  • 31.Prebble JN. 2001. The philosophical origins of Mitchell's chemiosmotic concepts: the personal factor in scientific theory formulation. J. Hist. Biol. 34, 433–460. ( 10.1023/A:1012946715673) [DOI] [PubMed] [Google Scholar]
  • 32.Junge W. 2013. Half a century of molecular bioenergetics. Biochem. Soc. Trans. 41, 1207–1218. ( 10.1042/BST20130199) [DOI] [PubMed] [Google Scholar]
  • 33.Brown S, Simcock DC. 2011. Half a century of the chemiosmotic hypothesis and the practice of science. Orbital Electron. J. Chem. 3, 174–179. [Google Scholar]
  • 34.Wiedenmann A, Dimroth P, von Ballmoos C.. 2008. Δψ and ΔpH are equivalent driving forces for proton transport through isolated F0 complexes of ATP synthases. Biochim. Biophys. Acta Bioenerg. 1777, 1301–1310. ( 10.1016/j.bbabio.2008.06.008) [DOI] [PubMed] [Google Scholar]
  • 35.Krulwich TA, Ito M, Gilmour R, Sturr MG, Guffanti AA, Hicks DB. 1996. Energetic problems of extremely alkaliphilic aerobes. Biochim. Biophys. Acta 1275, 21–26. ( 10.1016/0005-2728(96)00044-8) [DOI] [PubMed] [Google Scholar]
  • 36.Mulkidjanian AY, Heberle J, Cherepanov DA. 2006. Protons @ interfaces: implications for biological energy conversion. Biochim. Biophys. Acta 1757, 913–930. ( 10.1016/j.bbabio.2006.02.015) [DOI] [PubMed] [Google Scholar]
  • 37.Wolf MG, Grubmüller H, Groenhof G. 2014. Anomalous surface diffusion of protons on lipid membranes. Biophys. J. 107, 76–87. ( 10.1016/j.bpj.2014.04.062) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Tedeschi H, Rottenberg H. 1979. Is there a significant metabolically dependent membrane potential in mitochondria? Trends Biochem. Sci. 4, N182–N185. ( 10.1016/0968-0004(79)90407-9) [DOI] [Google Scholar]
  • 39.Williams RJ. 1975. Proton-driven phosphorylation reactions in mitochondrial and chloroplast membranes. FEBS Lett. 53, 123–125. ( 10.1016/0014-5793(75)80001-9) [DOI] [PubMed] [Google Scholar]
  • 40.Weifu Lee J. 2012. Proton-electrostatics hypothesis for localized proton coupling bioenergetics. Bioenergetics 1, 1–8. ( 10.4172/2167-7662.1000104) [DOI] [Google Scholar]
  • 41.Rieger B, Junge W, Busch KB. 2014. Lateral pH gradient between OXPHOS complex IV and F0F1 ATP-synthase in folded mitochondrial membranes. Nat. Commun. 5, 3103 ( 10.1038/ncomms4103) [DOI] [PubMed] [Google Scholar]
  • 42.Junge W, McLaughlin S. 1987. The role of fixed and mobile buffers in the kinetics of proton movement. Biochim. Biophys. Acta 890, 1–5. ( 10.1016/0005-2728(87)90061-2) [DOI] [PubMed] [Google Scholar]
  • 43.Capaldi RA, Aggeler R, Turina P, Wilkens S. 1994. Coupling between catalytic sites and the proton channel in F1F0-type ATPases. Trends Biochem. Sci. 19, 284–289. ( 10.1016/0968-0004(94)90006-X) [DOI] [PubMed] [Google Scholar]
  • 44.Turina P, Samoray D, Graber P. 2003. H+/ATP ratio of proton transport-coupled ATP synthesis and hydrolysis catalysed by CF0F1-liposomes. Embo J. 22, 418–426. ( 10.1093/emboj/cdg073) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Yaguzhinsky LS, Boguslavsky LI, Volkov AG, Rakhmaninova AB. 1976. Synthesis of ATP coupled with action of membrane protonic pumps at the octane-water interface. Nature 259, 494–496. ( 10.1038/259494a0) [DOI] [PubMed] [Google Scholar]
  • 46.Eroshenko LV, Marakhovskaya AS, Vangeli IM, Semenyuk PI, Orlov VN, Yaguzhinsky LS. 2012. Brønsted acids bounded to the mitochondrial membranes as a substrate for ATP synthase. Dokl. Biochem. Biophys. 444, 158–161. ( 10.1134/S160767291203009X) [DOI] [PubMed] [Google Scholar]
  • 47.Tupper JT, Tedeschi H. 1969. Mitochondrial membrane potentials measured with microelectrodes: probable ionic basis. Science 166, 1539–1540. ( 10.1126/science.166.3912.1539) [DOI] [PubMed] [Google Scholar]
  • 48.Tupper JT, Tedeschi H. 1969. Microelectrode studies on the membrane properties of isolated mitochondria. Proc. Natl Acad. Sci. USA 63, 370–377. ( 10.1073/pnas.63.2.370) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Harris EJ, Pressman BC. 1969. The direction of polarity of the mitochondrial trans-membrane potential. Biochim. Biophys. Acta 172, 66–70. ( 10.1016/0005-2728(69)90092-9) [DOI] [PubMed] [Google Scholar]
  • 50.Scaduto RC Jr, Grotyohann LW. 1999. Measurement of mitochondrial membrane potential using fluorescent rhodamine derivatives. Biophys. J. 76, 469–477. ( 10.1016/S0006-3495(99)77214-0) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Schmidt G, Gräber P. 1987. The rate of ATP synthesis catalyzed by reconstituted CF0F1-liposomes: dependence on ΔpH and Δψ. Biochim. Biophys. Acta Bioenerg. 890, 392–394. ( 10.1016/0005-2728(87)90168-X) [DOI] [Google Scholar]
  • 52.Zoratti M, Favaron M, Pietrobon D, Azzone GF. 1986. Intrinsic uncoupling of mitochondrial proton pumps. 1. Non-ohmic conductance cannot account for the nonlinear dependence of static head respiration on delta microH. Biochemistry 25, 760–767. ( 10.1021/bi00352a004) [DOI] [PubMed] [Google Scholar]
  • 53.Grzesiek S, Dencher NA. 1986. Dependency of delta pH-relaxation across vesicular membranes on the buffering power of bulk solutions and lipids. Biophys. J. 50, 265–276. ( 10.1016/S0006-3495(86)83460-9) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Tepper HL, Voth GA. 2006. Mechanisms of passive ion permeation through lipid bilayers: insights from simulations. J. Phys. Chem. B 110, 21 327–21 337. ( 10.1021/jp064192h) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Shinoda W. 2016. Permeability across lipid membranes. Biochim. Biophys. Acta 1858, 2254–2265. ( 10.1016/j.bbamem.2016.03.032) [DOI] [PubMed] [Google Scholar]
  • 56.Brookes PS, Hulbert AJ, Brand MD. 1997. The proton permeability of liposomes made from mitochondrial inner membrane phospholipids: no effect of fatty acid composition. Biochim. Biophys. Acta 1330, 157–164. ( 10.1016/S0005-2736(97)00160-0) [DOI] [PubMed] [Google Scholar]
  • 57.Mulkidjanian AY, Cherepanov DA. 2006. Probing biological interfaces by tracing proton passage across them. Photochem. Photobiol. Sci. 5, 577–587. ( 10.1039/b516443e) [DOI] [PubMed] [Google Scholar]
  • 58.von Hansen Y, Gekle S, Netz RR.. 2013. Anomalous anisotropic diffusion dynamics of hydration water at lipid membranes. Phys. Rev. Lett. 111, 118103 ( 10.1103/PhysRevLett.111.118103) [DOI] [PubMed] [Google Scholar]
  • 59.Kundacina N, Shi M, Pollack GH. 2016. Effect of local and general anesthetics on interfacial water. PLoS ONE 11, e0152127 ( 10.1371/journal.pone.0152127) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Deplazes E, Poger D, Cornell B, Cranfield CG. 2018. The effect of H3O+ on the membrane morphology and hydrogen bonding of a phospholipid bilayer. Biophys. Rev. 10, 1371–1376. ( 10.1007/s12551-018-0454-z) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Mulkidjanian AY. 2006. Proton in the well and through the desolvation barrier. Biochim. Biophys. Acta 1757, 415–427. ( 10.1016/j.bbabio.2006.04.023) [DOI] [PubMed] [Google Scholar]
  • 62.Zhan CG, Dixon DA. 2001. Absolute hydration free energy of the proton from first-principles electronic structure calculations. J. Phys. Chem. A 105, 11 534–11 540. ( 10.1021/jp012536s) [DOI] [Google Scholar]
  • 63.Cherepanov DA, Junge W, Mulkidjanian AY. 2004. Proton transfer dynamics at the membrane/water interface: dependence on the fixed and mobile pH buffers, on the size and form of membrane particles, and on the interfacial potential barrier. Biophys. J. 86, 665–680. ( 10.1016/S0006-3495(04)74146-6) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Mulkidjanian AY, Cherepanov DA, Heberle J, Junge W. 2005. Proton transfer dynamics at membrane/water interface and mechanism of biological energy conversion. Biochemistry (Mosc) 70, 251–256. ( 10.1007/s10541-005-0108-1) [DOI] [PubMed] [Google Scholar]
  • 65.Bal W, Kurowska E, Maret W. 2012. The final frontier of pH and the undiscovered country beyond. PLoS ONE 7, e45832 ( 10.1371/journal.pone.0045832) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Żurawik TM, Pomorski A, Belczyk-Ciesielska A, Goch G, Niedźwiedzka K, Kucharczyk R, Krężel A, Bal W. 2016. Revisiting mitochondrial pH with an improved algorithm for calibration of the ratiometric 5(6)-carboxy-SNARF-1 probe reveals anticooperative reaction with H+ ions and warrants further studies of organellar pH. PLoS ONE 11, e0161353 ( 10.1371/journal.pone.0161353) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Grotthuss T.1805. Memoire sur la décomposition de l'eau et des corps qu'elle tient en dissolution à l'aide de l’électricité galvanique. Rome.
  • 68.Serowy S, Saparov SM, Antonenko YN, Kozlovsky W, Hagen V, Pohl P. 2003. Structural proton diffusion along lipid bilayers. Biophys. J. 84, 1031–1037. ( 10.1016/S0006-3495(03)74919-4) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.DeCoursey TE. 2015. The voltage-gated proton channel: a riddle, wrapped in a mystery, inside an enigma. Biochemistry 54, 3250–3268. ( 10.1021/acs.biochem.5b00353) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Pomès R, Roux B. 2002. Molecular mechanism of H+ conduction in the single-file water chain of the gramicidin channel. Biophys. J. 82, 2304–2316. ( 10.1016/S0006-3495(02)75576-8) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Nagle JF, Morowitz HJ. 1978. Molecular mechanisms for proton transport in membranes. Proc. Natl Acad. Sci. USA 75, 298–302. ( 10.1073/pnas.75.1.298) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Nagle JF, Tristram-Nagle S. 1983. Hydrogen bonded chain mechanisms for proton conduction and proton pumping. J. Membr. Biol. 74, 1–14. ( 10.1007/BF01870590) [DOI] [PubMed] [Google Scholar]
  • 73.DeCoursey TE. 2017. CrossTalk proposal: proton permeation through HV1 requires transient protonation of a conserved aspartate in the S1 transmembrane helix. J. Physiol. 595, 6793–6795. ( 10.1113/JP274495) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Bennett AL, Ramsey IS. 2017. CrossTalk opposing view: proton transfer in Hv1 utilizes a water wire, and does not require transient protonation of a conserved aspartate in the S1 transmembrane helix. J. Physiol. 595, 6797–6799. ( 10.1113/JP274553) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Riccardi D, König P, Prat-Resina X, Yu H, Elstner M, Frauenheim T, Cui Q. 2006. ‘Proton holes’ in long-range proton transfer reactions in solution and enzymes: a theoretical analysis. J. Am. Chem. Soc. 128, 16 302–16 311. ( 10.1021/ja065451j) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Riccardi D, et al. 2006. Development of effective quantum mechanical/molecular mechanical (QM/MM) methods for complex biological processes. J. Phys. Chem. B 110, 6458–6469. ( 10.1021/jp056361o) [DOI] [PubMed] [Google Scholar]
  • 77.Miyake T, Rolandi M. 2016. Grotthuss mechanisms: from proton transport in proton wires to bioprotonic devices. J. Phys. Condens. Matter 28, 023001 ( 10.1088/0953-8984/28/2/023001) [DOI] [PubMed] [Google Scholar]
  • 78.Morelli AM, Ravera S, Calzia D, Panfoli I. 2013. Hypothesis of lipid-phase-continuity proton transfer for aerobic ATP synthesis. J. Cereb. Blood Flow Metab. 33, 1838–1842. ( 10.1038/jcbfm.2013.175) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Ravera S, Panfoli I, Aluigi MG, Calzia D, Morelli A. 2011. Characterization of myelin sheath FoF1-ATP synthase and its regulation by IF1. Cell Biochem. Biophys. 59, 63–70. ( 10.1007/s12013-010-9112-1) [DOI] [PubMed] [Google Scholar]
  • 80.Fujikawa M, Imamura H, Nakamura J, Yoshida M. 2012. Assessing actual contribution of IF1, inhibitor of mitochondrial FoF1, to ATP homeostasis, cell growth, mitochondrial morphology, and cell viability. J. Biol. Chem. 287, 18 781–18 787. ( 10.1074/jbc.M112.345793) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Varghese F, Blaza JN, Jones AJY, Jarman OD, Hirst J. 2018. Deleting the IF1-like ζ subunit from Paracoccus denitrificans ATP synthase is not sufficient to activate ATP hydrolysis. Open Biol. 8, 170206 ( 10.1098/rsob.170206) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Biukovic G, Basak S, Manimekalai MSS, Rishikesan S, Roessle M, Dick T, Rao SPS, Hunke C, Grüber G. 2013. Variations of subunit ɛ of the Mycobacterium tuberculosis F1Fo ATP synthase and a novel model for mechanism of action of the tuberculosis drug TMC207. Antimicrob. Agents Chemother. 57, 168–176. ( 10.1128/AAC.01039-12) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Ogilvie I, Kennaway NG, Shoubridge EA. 2005. A molecular chaperone for mitochondrial complex I assembly is mutated in a progressive encephalopathy. J. Clin. Invest. 115, 2784–2792. ( 10.1172/JCI26020) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Wikström M, Hummer G. 2012. Stoichiometry of proton translocation by respiratory complex I and its mechanistic implications. Proc. Natl Acad. Sci. USA 109, 4431–4436. ( 10.1073/pnas.1120949109) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Verkhovskaya M, Bloch DA. 2013. Energy-converting respiratory Complex I: on the way to the molecular mechanism of the proton pump. Int. J. Biochem. Cell Biol. 45, 491–511. ( 10.1016/j.biocel.2012.08.024) [DOI] [PubMed] [Google Scholar]
  • 86.Brzezinski P. 2004. Redox-driven membrane-bound proton pumps. Trends Biochem. Sci. 29, 380–387. ( 10.1016/j.tibs.2004.05.008) [DOI] [PubMed] [Google Scholar]
  • 87.Wikström M, Sharma V. 2018. Proton pumping by cytochrome c oxidase—a 40 year anniversary. Biochim. Biophys. Acta Bioenerg. 1859, 692–698. ( 10.1016/j.bbabio.2018.03.009) [DOI] [PubMed] [Google Scholar]
  • 88.Papa S, Lorusso M, Di Paola M.. 2006. Cooperativity and flexibility of the protonmotive activity of mitochondrial respiratory chain. Biochim. Biophys. Acta 1757, 428–436. ( 10.1016/j.bbabio.2006.03.015) [DOI] [PubMed] [Google Scholar]
  • 89.Akeson M, Deamer DW. 1991. Proton conductance by the gramicidin water wire. Model for proton conductance in the F1F0 ATPases? Biophys. J. 60, 101–109. ( 10.1016/S0006-3495(91)82034-3) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Tasseva G, Bai HD, Davidescu M, Haromy A, Michelakis E, Vance JE. 2013. Phosphatidylethanolamine deficiency in mammalian mitochondria impairs oxidative phosphorylation and alters mitochondrial morphology. J. Biol. Chem. 288, 4158–4173. ( 10.1074/jbc.M112.434183) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.von Ballmoos C, Biner O, Nilsson T, Brzezinski P.. 2016. Mimicking respiratory phosphorylation using purified enzymes. Biochim. Biophys. Acta Bioenerg. 1857, 321–331. ( 10.1016/j.bbabio.2015.12.007) [DOI] [PubMed] [Google Scholar]
  • 92.Nilsson T, Lundin CR, Nordlund G, Ädelroth P, von Ballmoos C, Brzezinski P.. 2016. Lipid-mediated protein-protein interactions modulate respiration-driven ATP synthesis. Sci. Rep. 6, 24113 ( 10.1038/srep24113) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Prats M, Tocanne JF, Teissié J. 1985. Lateral proton conduction at a lipid/water interface: its modulation by physical parameters. Experimental and mathematical approaches. Eur. J. Biochem. 149, 663–668. ( 10.1111/j.1432-1033.1985.tb08975.x) [DOI] [PubMed] [Google Scholar]
  • 94.Teissié J, Prats M, Soucaille P, Tocanne JF. 1985. Evidence for conduction of protons along the interface between water and a polar lipid monolayer. Proc. Natl Acad. Sci. USA 82, 3217–3221. ( 10.1073/pnas.82.10.3217) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Panfoli I, Ponassi M, Ravera S, Calzia D, Beitia M, Morelli A, Rosano C. 2017. Tracking protons from respiratory chain complexes to ATP synthase c-subunit: the critical role of serine and threonine residues. Biochem. Biophys. Res. Commun. 482, 922–927. ( 10.1016/j.bbrc.2016.11.134) [DOI] [PubMed] [Google Scholar]
  • 96.Brändén M, Sandén T, Brzezinski P, Widengren J. 2006. Localized proton microcircuits at the biological membrane-water interface. Proc. Natl Acad. Sci. USA 103, 19 766–19 770. ( 10.1073/pnas.0605909103) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Watt IN, Montgomery MG, Runswick MJ, Leslie AG, Walker JE. 2010. Bioenergetic cost of making an adenosine triphosphate molecule in animal mitochondria. Proc. Natl Acad. Sci. USA 107, 16 823–16 827. ( 10.1073/pnas.1011099107) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Saffman PG, Delbrück M. 1975. Brownian motion in biological membranes. Proc. Natl Acad. Sci. USA 72, 3111–3113. ( 10.1073/pnas.72.8.3111) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.von Ballmoos C, Wiedenmann A, Dimroth P.. 2009. Essentials for ATP synthesis by F1F0 ATP synthases. Annu. Rev. Biochem. 78, 649–672. ( 10.1146/annurev.biochem.78.081307.104803) [DOI] [PubMed] [Google Scholar]
  • 100.Morelli A, Ravera S, Calzia D, Panfoli I. 2011. Exportability of the mitochondrial oxidative phosphorylation machinery into myelin sheath. Theor. Biol. Forum 104, 67–74. [PubMed] [Google Scholar]
  • 101.Panfoli I, Ravera S, Bruschi M, Candiano G, Morelli A. 2011. Proteomics unravels the exportability of mitochondrial respiratory chains. Expert Rev. Proteomics 8, 231–239. ( 10.1586/epr.11.1) [DOI] [PubMed] [Google Scholar]
  • 102.Calzia D, Garbarino G, Caicci F, Manni L, Candiani S, Ravera S, Morelli A, Traverso CE, Panfoli I. 2014. Functional expression of electron transport chain complexes in mouse rod outer segments. Biochimie 102, 78–82. ( 10.1016/j.biochi.2014.02.007) [DOI] [PubMed] [Google Scholar]
  • 103.Calzia D, et al. 2013. Are rod outer segment ATP-ase and ATP-synthase activity expression of the same protein? Cell. Mol. Neurobiol. 33, 637–649. ( 10.1007/s10571-013-9926-7) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Bianchini P, et al. 2008. Live imaging of mammalian retina: rod outer segments are stained by conventional mitochondrial dyes. J. Biomed. Opt. 13, 054017 ( 10.1117/1.2982528) [DOI] [PubMed] [Google Scholar]
  • 105.Panfoli I, et al. 2009. Evidence for aerobic metabolism in retinal rod outer segment disks. Int. J. Biochem. Cell Biol. 41, 2555–2565. ( 10.1016/j.biocel.2009.08.013) [DOI] [PubMed] [Google Scholar]
  • 106.Ravera S, Panfoli I, Calzia D, Aluigi MG, Bianchini P, Diaspro A, Mancardi G, Morelli A. 2009. Evidence for aerobic ATP synthesis in isolated myelin vesicles. Int. J. Biochem. Cell Biol. 41, 1581–1591. ( 10.1016/j.biocel.2009.01.009) [DOI] [PubMed] [Google Scholar]
  • 107.Ravera S, Bartolucci M, Calzia D, Aluigi MG, Ramoino P, Morelli A, Panfoli I. 2013. Tricarboxylic acid cycle-sustained oxidative phosphorylation in isolated myelin vesicles. Biochimie 95, 1991–1998. ( 10.1016/j.biochi.2013.07.003) [DOI] [PubMed] [Google Scholar]
  • 108.Ravera S, et al. 2016. Support of nerve conduction by respiring myelin sheath: role of connexons. Mol. Neurobiol. 53, 2468–2479. ( 10.1007/s12035-015-9216-0) [DOI] [PubMed] [Google Scholar]
  • 109.Bartolucci M, Ravera S, Garbarino G, Ramoino P, Ferrando S, Calzia D, Candiani S, Morelli A, Panfoli I. 2015. Functional expression of electron transport chain and FoF1-ATP synthase in optic nerve myelin sheath. Neurochem. Res. 40, 2230–2241. ( 10.1007/s11064-015-1712-0) [DOI] [PubMed] [Google Scholar]
  • 110.Ravera S, Bartolucci M, Garbati P, Ferrando S, Calzia D, Ramoino P, Balestrino M, Morelli A, Panfoli I. 2016. Evaluation of the acquisition of the aerobic metabolic capacity by myelin, during its development. Mol. Neurobiol. 53, 7048–7056. ( 10.1007/s12035-015-9575-6) [DOI] [PubMed] [Google Scholar]
  • 111.Mangiullo R, Gnoni A, Leone A, Gnoni GV, Papa S, Zanotti F. 2008. Structural and functional characterization of FoF1-ATP synthase on the extracellular surface of rat hepatocytes. Biochim. Biophys. Acta 1777, 1326–1335. ( 10.1016/j.bbabio.2008.08.003) [DOI] [PubMed] [Google Scholar]
  • 112.Arakaki N, et al. 2003. Possible role of cell surface H+-ATP synthase in the extracellular ATP synthesis and proliferation of human umbilical vein endothelial cells. Mol. Cancer Res. 1, 931–939. [PubMed] [Google Scholar]
  • 113.Arakaki N, Kita T, Shibata H, Higuti T. 2007. Cell-surface H+-ATP synthase as a potential molecular target for anti-obesity drugs. FEBS Lett. 581, 3405–3409. ( 10.1016/j.febslet.2007.06.041) [DOI] [PubMed] [Google Scholar]
  • 114.Kita T, Arakaki N. 2015. Contribution of extracellular ATP on the cell-surface F1F0-ATP synthase-mediated intracellular triacylglycerol accumulation. Biomed. Res. 36, 115–120. ( 10.2220/biomedres.36.115) [DOI] [PubMed] [Google Scholar]
  • 115.Chang HY, Huang HC, Huang TC, Yang PC, Wang YC, Juan HF. 2012. Ectopic ATP synthase blockade suppresses lung adenocarcinoma growth by activating the unfolded protein response. Cancer Res. 72, 4696–4706. ( 10.1158/0008-5472.CAN-12-0567) [DOI] [PubMed] [Google Scholar]
  • 116.Lee H, Kim S-H, Lee J-S, Yang Y-H, Nam J-M, Kim B-W, Ko Y-G. 2016. Mitochondrial oxidative phosphorylation complexes exist in the sarcolemma of skeletal muscle. BMB Rep. 49, 116–121. ( 10.5483/BMBRep.2016.49.2.232) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Taurino F, Gnoni A. 2018. Systematic review of plasma-membrane ecto-ATP synthase: a new player in health and disease. Exp. Mol. Pathol. 104, 59–70. ( 10.1016/j.yexmp.2017.12.006) [DOI] [PubMed] [Google Scholar]
  • 118.Taurino F, Giannoccaro C, Sardanelli AM, Cavallo A, De Luca E, Santacroce S, Papa S, Zanotti F, Gnoni A.. 2016. Function and expression study uncovered hepatocyte plasma membrane ecto-ATP synthase as a novel player in liver regeneration. Biochem. J. 473, 2519–2530. ( 10.1042/BCJ20160065) [DOI] [PubMed] [Google Scholar]
  • 119.Gao J, Zhang T, Kang Z, Ting W, Xu L, Yin D. 2017. The F0F1 ATP synthase regulates human neutrophil migration through cytoplasmic proton extrusion coupled with ATP generation. Mol. Immunol. 90, 219–226. ( 10.1016/j.molimm.2017.08.004) [DOI] [PubMed] [Google Scholar]
  • 120.Ravera S, et al. 2018. Extramitochondrial energy production in platelets. Biol. Cell 110, 97–108. ( 10.1111/boc.201700025) [DOI] [PubMed] [Google Scholar]
  • 121.Panfoli I, et al. 2016. Exosomes from human mesenchymal stem cells conduct aerobic metabolism in term and preterm newborn infants. FASEB J. 30, 1416–1424. ( 10.1096/fj.15-279679) [DOI] [PubMed] [Google Scholar]
  • 122.Bruschi M, et al. 2015. Human urinary exosome as a metabolic effector cargo. Expert Rev. Proteomics 12, 425–432. ( 10.1586/14789450.2015.1055324) [DOI] [PubMed] [Google Scholar]
  • 123.Bruschi M, et al. 2017. Metabolic signature of microvesicles from umbilical cord mesenchymal stem cells of preterm and term infants. Proteomics Clin. Appl. 12, 1700082 ( 10.1002/prca.201700082) [DOI] [PubMed] [Google Scholar]
  • 124.Walker JE. 2013. The ATP synthase: the understood, the uncertain and the unknown. Biochem. Soc. Trans. 41, 1–16. ( 10.1042/BST20110773) [DOI] [PubMed] [Google Scholar]
  • 125.Ivontsin LA, Mashkovtseva EV, Nartsissov YR. 2017. Simulation of proton movement in FoF1-ATP synthase by quantum-mechanical approach. J. Phys. Conf. Ser. 784, 012021 ( 10.1088/1742-6596/784/1/012021) [DOI] [Google Scholar]
  • 126.Tepper HL, Voth GA. 2005. Protons may leak through pure lipid bilayers via a concerted mechanism. Biophys. J. 88, 3095–3108. ( 10.1529/biophysj.104.056184) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Watson JD, Crick FH. 1953. Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid. Nature 171, 737–738. ( 10.1038/171737a0) [DOI] [PubMed] [Google Scholar]
  • 128.Hunter P. 2006. A quantum leap in biology. One inscrutable field helps another, as quantum physics unravels consciousness. EMBO Rep. 7, 971–974. ( 10.1038/sj.embor.7400802) [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Reviewer comments

Data Availability Statement

This article has no additional data.


Articles from Open Biology are provided here courtesy of The Royal Society

RESOURCES