Abstract
Following structural determination by recent advances in electron cryomicroscopy, it is now well established that the respiratory Complexes I–IV in oxidative phosphorylation (OXPHOS) are organized into supercomplexes in the respirasome. Nonetheless, the reason for the existence of the OXPHOS supercomplexes and their functional role remains an enigma. Several hypotheses have been proposed for the existence of these supercomplex supercomplexes. A commonly-held view asserts that they enhance catalysis by substrate channeling. However, this – and other views – has been challenged based on structural and biophysical information. Hence, new ideas, concepts, and frameworks are needed. Here, a new model of energy transfer in OXPHOS is developed on the basis of biochemical data on the pure competitive inhibition of anionic substrates like succinate by the classical anionic uncouplers of OXPHOS (2,4-dinitrophenol, carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone, and dicoumarol), and pharmacological data on the unique site-selective, energy-linked inhibition of energy conservation pathways in mitochondria induced by the guanidine derivatives. It is further found that uncouplers themselves are site-specific and exhibit differential selectivity and efficacy in reversing the inhibition caused by the Site 1/Complex I or Site 2/Complexes II–III-selective guanidine derivatives. These results lead to new vistas and sufficient complexity in the network of energy conservation pathways in the mitochondrial respiratory chain that necessitate discrete points of interaction with two classes of guanidine derivatives and uncoupling agents and thereby separate and distinct energy transfer pathways between Site 1 and Site 2 and the intermediate that energizes adenosine triphosphate (ATP) synthesis by Complex V. Interpretation based on Mitchell’s single-ion chemiosmotic theory that postulates only a single energy pool is inadequate to rationalize the data and account for the required complexity. The above results and available information are shown to be explained by Nath’s two-ion theory of energy coupling and ATP synthesis, involving coupled movement of succinate anions and protons, along with the requirement postulated by the theory for maintenance of homeostasis and ion translocation across the energy-transducing membrane of both succinate monoanions and succinate dianions by Complexes I–V in the OXPHOS supercomplexes. The new model of energy transfer in mitochondria is mapped onto the solved structures of the supercomplexes and integrated into a consistent model with the three-dimensional electron microscope computer tomography visualization of the internal structure of the cristae membranes in mammalian mitochondria. The model also offers valuable insights into diseased states induced in type 2 diabetes and especially in Alzheimer’s and other neurodegenerative diseases that involve mitochondrial dysfunction.
Abbreviations
- AD
-
Alzheimer’s disease
- ADP
-
adenosine diphosphate
- ATP
-
adenosine triphosphate
- Aβ
-
amyloid-β-peptide
- AMPK
-
AMP-activated protein kinase
- ANT
-
adenine nucleotide translocase
- CsA
-
cyclosporin A
- CypD
-
cyclophylin D
- DNP
-
2,4-dinitrophenol
- FCCP
-
carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone
- FOF1
-
ATP synthase
- Metformin
-
1,1-dimethylbiguanide
- OXPHOS
-
oxidative phosphorylation
- Pi-OH–
-
phosphate exchanger
- PT
-
permeability transition
- PTP
-
permeability transition pore
- ROS
-
reactive oxygen species
- TOM
-
translocase of outer mitochondrial membrane
Introduction
The development of blue native polyacrylamide gel electrophoresis led to the separation, visualization, and identification of the ∼2 MDa oxidative phosphorylation (OXPHOS) supercomplexes by Schägger and Pfeiffer ∼25 years ago [1,2,3]. The existence of the OXPHOS supercomplexes was initially criticized as artifacts of detergent solubilization. However, it was shown by later research work that the active mammalian respiratory chain supercomplex contained a single unit of Complex I, two units of Complex III, and a single unit of Complex IV, and I1III2IV1 was considered the functional unit of the respirasome [4]. OXPHOS supercomplexes have also been imaged in situ [5]. Supercomplex structures were subsequently solved to atomic resolution by three groups [6,7,8]. Yang and coworkers solved the structure of the mammalian OXPHOS supercomplex to a 3.9 Å resolution [9] and proposed a model of a human OXPHOS supercomplex containing all the four complexes of the respiratory chain [10].
The functional role of the OXPHOS supercomplexes has proved to be an even greater puzzle. Why are they present, and what advantages do they confer? The proposed advantages – substrate channeling [11,12], prevention of ROS generation, structural stabilization and packing, and assembly and activation of Complex I [13] – have been challenged based on biochemical and biophysical evidence [14]. On the one hand, individual OXPHOS complexes have been shown to possess normal catalytic activity [15,16]; yet on the other hand, a number of reports also point to defects in network morphology [17,18], ATP synthesis [19,20,21,22], apoptosis [22,23,24], and integrated mitochondrial function [15,25]. In particular, aging [26,27] and various mitochondrial diseases [28,29,30,31,32,33,34] have been shown to be characterized by a decreased ability to form supercomplexes, or a greater propensity of the OXPHOS supercomplexes to disassociate, among other defects.
The above inconclusive debates during the past 25 years and the current impasse suggest the need for new ideas on the raison d’etre of the OXPHOS supercomplexes. A unique perspective for this exploration is offered by the development and refinement of Nath’s torsional mechanism of energy transduction and ATP synthesis in both the FO and F1 portions of the ATP synthase enzyme [35,36,37,38,39,40,41,42,43,44] over the past 25 years. The full naming of the mechanism is due to other authors [45,46,47,48,49,50]. Other workers have also interpreted the mechanism from the standpoint of Complex V, ATP synthase [20,50,51,52,53,54,55,56,57,58,59,60,61]. Following Professor William E. Moerner’s suggestion [62], for greater clarity in understanding of Nath’s work work by other scientists, the detailed theory/mechanism within the membrane-bound FO portion of the ATP synthase was termed Nath’s two-ion theory of energy coupling and ATP synthesis [21,63,64,65,66,67,68,69,70]. What insights does the theory/mechanism have to offer on Complexes I–IV and supramolecular organization in the respirasome?
Section 3 is organized as follows. Section 3.1 reports biochemical experiments on the competitive inhibition of succinate with several anionic uncouplers of OXPHOS, 2,4-dinitrophenol (DNP), carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone (FCCP), and dicoumarol. Since inhibition could occur either due to inhibition of succinate oxidation or due to inhibition of succinate entry into mitochondria, experiments to discriminate between the two alternatives are performed and discussed in Section 3.2. Pharmacological experiments on the site-specific interaction of guanidine and octylguanidine–reagents first introduced by Hollunger [71,72]–with the energy transfer pathways in mitochondria are the subject of Section 3.3. The section also reports relevant experiments on the site-specific nature of uncoupler action and especially on the abilities of uncouplers to relieve inhibition induced by the site-selective, energy-linked action of octylguanidine and phenethylguanide to different extents. These experiments suggest a new model of energy transfer between the respiratory enzyme complexes (Complexes I–IV) and ATP synthase (Complex V) that is formulated and discussed in considerable detail in Section 3.4. In particular, a new idea is developed in this section and in Figure 6, namely that a correct model of mitochondrial energy transfer needs to include sufficient complexity that explains discrete points of interaction with two classes of guanidine derivatives and uncoupling agents. A model postulating a single energy pool is shown to be inadequate for explaining the experimental observations.
The model is integrated into new information on the internal structure of mitochondria [73,74,75,76,77] in Section 3.5. The question of what respiration is is posed and answered in Section 3.6. The need for a local potential to be sensed and the coupling of electron-coupled proton transfer to the translocation of the dicarboxylic acid anions across the energy-transducing membrane is developed in Section 3.6. This suggests a novel, hitherto unexplored rationale and functional role for the existence of the OXPHOS supercomplexes. The overall model of energy coupling during respiration, its biological implications, and its relevance to disease are discussed briefly in Section 3.7.
Finally, in response to referee comments, the difficult grand challenge of the specific application of the results contained in refs. 22,40 and the biochemical and pharmacological insights gained from the present work to type 2 diabetes and Alzheimer’s disease (AD) is taken up in Section 3.8.
Experimental
For the inhibition experiments (Figures 1 and 2) between anionic substrate (succinate) and anionic uncouplers (DNP, FCCP, and dicoumarol), the following procedure was used. Mitochondria were isolated from rat liver as described previously [78,79,80] and incubated at 25°C at 1 mg/mL concentration in a medium containing 50 mM Tris chloride, 50 mM KCl, 50 mM sucrose, 5 mM MgCl2, 1 mM EDTA, 10 mM Tris phosphate, 2 mM adenosine diphosphate (ADP), and 1 μg/mL rotenone and varying substrate concentrations of succinate between 0.5 and 6.7 mM. The final pH measured 7.4 and the volume 2 mL. Oxygen consumption was measured using a Clark-type oxygen electrode. FCCP concentrations varied in the range from 0 to 10 μM while the inhibitory effects of dicoumarol concentrations were studied in the range between 0 and 50 μM [78].

Competitive inhibition of succinate permeation/oxidation by the classical anionic uncoupler of OXPHOS FCCP. Kinetically pure competitive inhibition by FCCP is clearly indicated. The succinate concentration was varied between 1.0 and 6.7 mM, while the concentration of FCCP was altered from 0 to 10 μM. FCCP concentrations: 0 μM (○), 0.5 μM (◊), 5.0 μM (□), and 10.0 μM (∆).

Competitive inhibition of succinate by the OXPHOS uncoupler dicoumarol. The succinate concentration was varied between 0.5 and 6.7 mM, while the concentration of dicoumarol was varied in the range from 0 to 50 μM. Dicoumarol concentrations: 0 μM (○), 15 μM (◊), 30 μM (□), and 50 μM (∆).
In the case of the experiments on the inhibition of succinate uptake in the presence of respiratory inhibitors and oligomycin (Figure 4), the following conditions were used. The incubation mixture of 1 mL final volume contained 0.25 M sucrose; 20 mM tris(hydroxymethyl)aminomethane hydrochloride (Tris–HCl), pH 7.2; 1 mM ethylenediamine tetraacetic acid; 1 mM [14C] succinate; 1 μg rotenone; 0.5 μg antimycin; 10 μg oligomycin; FCCP at concentrations between 0 and 1 μM; and 4 mg mitochondrial protein. The temperature was 23°C, and the reaction time measured 1 min.
For experiments studying the behavior of succinate uptake in the presence of respiration (Figure 5), the reaction mixture contained 0.25 M sucrose, 20 mM triethanolamine, pH 7.3, 2 mM [14C] succinate, 1 μg rotenone, 2 mM phosphate, and 2 mg mitochondrial protein. The final volume measured 1 mL, and the temperature 23°C. The reaction time was 1 min.
In pharmacological experiments with the guanidine derivatives (Tables 1 and 2), the medium consisted of 0.25 M sucrose; 25 mM KCl; 15 mM Tris–HCl, pH 7.4; 5 mM Tris phosphate, pH 7.4; and 5 mM MgCl2. ADP, when added, was present at 1.5 mM concentration. Substrates, for example, succinate, were added to yield a final concentration of 3 mM. In the case of glutamate + malate, each substrate was present at a 3 mM concentration. The final volume measured 3 mL and contained 0.5 mg/mL mitochondrial protein. The temperature was 25°C. The concentration of octylguanidine was varied in the range from 0 to 250 μM. In order to achieve similar degrees of inhibition as shown by octylguanidine, guanidine concentration had to be varied between 0 and 100 mM.
Inhibition of mitochondrial respiration by the alkylguanidines. Differential effect of octylguanidine on systems respiring on succinate and glutamate–malate substrates
Concentration of octylguanidine (μM) | % Inhibition of respiration on 3 mM succinate | % Inhibition of respiration on 3 mM glutamate + 3 mM malate |
---|---|---|
0 | 0 | 0 |
50 | 19 | 72 |
100 | 37 | 86 |
150 | 50 | 90 |
200 | 57 | 93 |
250 | 57 | 95 |
Release of inhibition due to guanidine derivatives by OXPHOS uncouplers. Differential ability of the classical anionic uncouplers of OXPHOS dicoumarol and DNP to restore mitochondrial respiration inhibited by the alkylguanidines and phenethylbiguanides
Uncoupler | Uncoupler concentration (μM) | % Release of inhibition caused by 100 μM octylguanidine | % Release of inhibition caused by 3 mM phenethylbiguanide |
---|---|---|---|
DNP | 100 | 50 | 100 |
Dicoumarol | 10 | 2 | 100 |
Results and discussion
Competition between succinate anions and the acidic uncouplers of OXPHOS
The inhibition of uptake/oxidation of succinate in mitochondria by anions of various acidic uncouplers of OXPHOS is an important phenomenon in cellular metabolism. Figures 1 and 2 present the results of experiments that strongly support such competition between anions. Figure 1 shows the kinetically pure competitive inhibition observed between succinate (1–6.7 mM) and the uncoupler FCCP (0–10 μM). Figure 2 illustrates the competition process between substrate and uncoupler anions for succinate (0.5–6.7 mM) and dicoumarol uncoupler (0–50 μM). Similar competitive behavior between succinate (0.5–5 mM) and DNP anions in the range from 0 to 1.5 mM has been reported previously, along with a determination of the enzymological parameters K m , V max, and K i of the inhibition process [78].
Why should several anionic uncouplers compete with anionic substrates for entry/conversion in mitochondria? The competition phenomenon (Figures 1 and 2 and ref. 78) is not readily accounted for by Mitchell’s single-ion chemiosmotic theory, but is naturally explained by Nath’s two-ion theory of energy coupling and ATP synthesis [36,38,39,40,42,49,50,51,59,63,64,65,70,78]. The inhibition phenomenon (Figures 1 and 2) can occur either by inhibition of the entry of succinate or be due to the inhibition of the oxidation of succinate (Figure 3). If the inhibition is caused by the former process, then the intramitochondrial concentration of succinate would be expected to decrease, whereas if the inhibition is due to the latter, then the process would show an increase in the internal concentration of succinate in mitochondria (Figure 3).

Possible causes of succinate inhibition by uncoupler in mitochondria. The inhibition could occur either due to inhibition of the succinate entry process (1, green arrow) or by inhibition of the succinate oxidation process (2, red arrow).
Inhibition of succinate entry or succinate oxidation?
The competition between uncoupler and substrate anions shown by the rate experiments in Figures 1 and 2 can be demonstrated directly by following the internal concentration of labeled substrates. If the inhibition by uncouplers is accompanied by a decrease in the intramitochondrial succinate concentration, then the inhibition process can be attributed to the inhibition of the entry of substrate by uncoupler. If an increase in the internal succinate concentration is found, then inhibition of the oxidation process would be operative. The two possibilities are illustrated diagrammatically in Figure 3.
Figure 4 shows the experimental results that uncouplers strongly inhibit the uptake of succinate in the presence of respiratory inhibitors and oligomycin, that is, when both respiration and ATP energy are cut off. Similar behavior is found in the presence of respiration (Figure 5). The results of Figure 5 clearly show that the intramitochondrial concentration of succinate is progressively decreased by an increase in the concentration of DNP uncoupler. Since a decrease in the internal succinate concentration is found (Figures 4 and 5), it is concluded that inhibition of entry of succinate by uncoupler (Figure 3) has been demonstrated.

Effect of increasing FCCP uncoupler concentrations from 0 to 1 μM on succinate uptake by rat liver mitochondria. Inhibition of succinate uptake by FCCP in the presence of respiratory inhibitors rotenone, antimycin, and oligomycin.

Influence of increasing concentrations of DNP uncoupler on succinate uptake. Inhibition of the rate of entry of substrate by uncoupler, as shown by the progressive decrease in intramitochondrial succinate concentration upon increase in DNP concentration from 0 to 1 mM during respiration.
A detailed molecular mechanism of coupling of ion transport to ATP synthesis [81] and uncoupling of ion transport from ATP synthesis [82] has been developed based on these concepts by the two-ion theory [21,70]. Permeability of the energy-transducing membrane to both the uncoupler acid and the uncoupler anion is necessary [82], and according to the mechanism, the uncoupler is translocated through the half-access channels in the membrane in its anionic form exactly like a substrate anion. Once inside, the anion picks up a proton and leaves the mitochondrion in its neutral form by diffusion [82]. A detailed quantitative kinetic model of the process of coupling and uncoupling has been formulated [81,82]. The mathematical equations used to explain coupling and uncoupling in the framework are the same, except for the imposition of the respective conditions during coupling/uncoupling (e.g., translocation as separate H+ and succinate A– ions through the access channels during the process of coupling, as opposed to the formation of the neutral acid form UH due to the lipid solubility of the uncoupler anion U–, which results in uncoupling in OXPHOS) [81,82].
It has also been pointed out by us that Mitchell’s single-ion chemiosmotic theory [83,84] is forced to postulate active transport of the uncoupler U– anion in order to explain the dissipation of Δψ during uncoupling [82]. Where is the energy for this active anionic transport coming from? This is not addressed by the chemiosmotic theory [83,84]. Above all, there are major contradictions and inconsistencies in the chemiosmotic theory that, as one of its central tenets, postulates H+ as the sole ion involved in coupling oxidation and ATP synthesis in OXPHOS [21,45,50,51,56,59,61,83,84,85,86,87], because the theory is forced to fall back on anion transport to explain uncoupling [82], yet it envisages no role for the anion in the process of energy coupling [83,84,85]! These difficulties are overcome, and the longstanding inconsistencies are satisfactorily resolved by the two-ion theory of energy coupling/uncoupling involving succinate/uncoupler anions and protons [36,38−40,42,43,63–70,78–82].
The work has many other important biological implications. Uncoupler anions U− uncouple not by a general ability to ferry protons and act as proton conductors, as believed by some workers, but by binding to a specific site at the a–c interface of FO and forming a neutral UH species, thereby collapsing both driving forces Δ (pH) and Δ(ψ). This drive to form the undissociated species is not operative in the case of the physiological anion succinate and H+. It is also known that K+ (or Na+) can substitute for H+ in ATP synthesis by certain organisms. In such a situation, succinate and K+ or an “uncoupling” anion U− (for example, FCCP anion) and K+ shall translocate as individual (i.e. separate) ions through the access half-channels in the FO portion, and thereby enable energy coupling in ATP synthase. In other words, one can also have anion substitutes, depending on the context.
The classical chemical uncouplers of OXPHOS exert two effects on respiration. The results shown in Figures 1–5 have uncovered the first of these effects, for example, by competitive inhibition studies (Figures 1 and 2) and by the lowering of the internal substrate concentration by investigation of the direct action of uncouplers on the entry of substrate and succinate accumulation (Figures 4 and 5). However, uncouplers can also affect respiration directly by interacting with the Complexes of the electron transport chain in mitochondria (Section 3.3).
Network of energy transfer pathways in OXPHOS: the site-selective, energy-linked inhibition by guanidine derivatives, and the differential release of their inhibition by OXPHOS uncouplers
The pioneering pharmacological work of Gunnar Hollunger at the University of Lund in the mid-1950s was the first to describe the site-selective inhibition of mitochondria by guanidine and its derivatives [71,72]. Hollunger showed that these agents induced a preferential inhibition at Site 1 of the mitochondrial respiratory chain; the energy-linked character of this inhibition was clear from the reversal of the inhibition by uncouplers [71,72]. Hollunger’s visionary research has been largely forgotten, although it led to the development of the FDA-approved drug metformin, the gold standard for the management of type 2 diabetes since 1960 (Section 3.8.1). Metformin (or 1,1-dimethylbiguanide) inhibits oxidation of the reduced cofactor nicotinamide adenine dinucleotide (NADH) in preference to that of succinate and flavin adenine dinucleotide-linked carbon compounds.
The differential effect of octylguanidine on systems respiring on succinate and glutamate–malate substrates is listed in Table 1. Low (50 μM) concentrations of octylguanidine are effective in the inhibition of glutamate–malate system in state 3, compared to succinate as substrate (Table 1). At higher concentration levels (>200 μM), inhibition of glutamate-malate reaches ∼100% completion; however, the maximum inhibition of succinate oxidation only approaches ∼60%. Similar results were obtained with guanidine; however, the maximal concentrations of the inhibitor used were on the order of 100 mM.
The results of Table 1 confirm the original findings [71,72] that the alkylguanidines are more effective in inhibiting respiration on NADH-linked substrates than succinate; thus, they preferentially inhibit the first energy conservation site or Complex I. On the other hand, the related pharmacological agent, phenethylbiguanide, preferentially inhibits energy transfer with succinate as substrate; that is, it inhibits energy conservation at Site 2. Furthermore, we found Site 2 selectivity of biguanides only over a narrow structural range and side chain length of the biguanide derivatives. Thus, butylbiguanides and pentylbiguanides, like their phenethylbiguanide cousin, exhibited Site 2 selectivity; however, derivatives with shorter or longer side chains showed the more common Site 1 selectivity and inhibition of Complex I. Moreover, uncouplers themselves show site-selective effects. Thus, the results listed in Table 2 reveal that uncouplers such as dicoumarol are completely ineffective in releasing the respiratory inhibition induced by octylguanidine. However, dicoumarol has the ability to fully release inhibition due to phenethylbiguanide (Table 2). On the other hand, uncoupling agents such as DNP can release the inhibition induced by both types of guanidines, that is, octylguanidine and phenethylbiguanide (Table 2).
The results of inhibition by the guanidine derivatives (Tables 1 and 2) reveal the site-specificity of two types of uncoupling agents, one such as dicoumarol that relieves the inhibition by Site 2-selective phenethylbiguanide but not the inhibition induced by the Site 1-selective alkylguanidines, and another like DNP that release inhibition effected by both classes of guanidines. The results cannot be reconciled with the chemiosmotic theory that predicts only a single energy pool between respiration and ATP synthesis.
The observations lead to a more complex scheme involving separate and distinct energy transfer pathways between Site 1 and Site 2 in the mitochondrial respiratory chain and the intermediate that energizes ATP synthesis by Complex V (Section 3.4 and Figure 6).

New model of the network of energy transfer pathways between the respiratory chain (Complexes I–IV) and ATP synthase (Complex V) based on the results of this work. Guanidine and the alkylguanidines induce a preferential inhibition of energy conservation at Site 1 (Complex I), shown by the dashed red lines with downward arrows in the upper limb. On the other hand, phenethylbiguanide shows Site 2 selectivity and reveals a predilection for energy-linked inhibition of energy conservation and transfer at Site 2 (Complexes II–III), drawn in the diagram with downward pointing dashed red arrows in the lower limb. Furthermore, uncouplers are intrinsically site-specific too. Thus, DNP effectively reverses the respiratory inhibition caused by either Site 1- or Site 2-selective guanidine derivatives (shown by bold green upward pointing arrows in the two limbs). However, dicoumarol has the ability to release only the inhibition induced by phenethylbiguanide, but not by the alkylguanidines, delineated by the green arrow in the lower limb (Tables 1 and 2). Thus, the guanidine derivatives specifically interact at different sites in the respiratory energy transfer pathways in mitochondria. To sum up, the network of energy transfer pathways in the mitochondrial respiratory chain requires discrete points of interaction with two classes of guanidine derivatives and uncoupling agents.
New model of network of energy transfer pathways between the respiratory chain (Complexes I–IV) and ATP synthase (Complex V)
As discussed in Section 3.3, the experimental observation on the site-specific inhibition unleashed by the guanidine derivatives, and the differential restoration of respiration induced by two classes of uncoupling agents necessitates a new model of interaction in mitochondrial energy transfer. This model is illustrated by the scheme shown in Figure 6. The preferential inhibition of energy conservation at Site 1 (Complex I) induced by guanidine and octylguanidine is shown by the dashed red lines with downward arrows in the upper limb of Figure 6. The energy-linked inhibition of energy conservation and transfer at Site 2 (Complexes II–III) is illustrated by the downward pointing dashed red arrows in the lower limb of Figure 6. The relief by uncouplers of the respiratory inhibition caused by either Site 1- or Site 2-selective guanidine derivatives is shown by bold green upward-pointing arrows in the two limbs (Figure 6). Thus, two energy transfer pathways between electron transport and ATP synthesis are required for a more complete understanding of the OXPHOS system (Figure 6).
The internal structure of mitochondria
The pioneering work of Lea and Hollenberg had visualized the internal structure of liver mitochondria (Figure 7) using high-resolution scanning electron microscopy [73]. The tubular cristae and lumen are clearly seen in the structure (Figure 7). Higher-resolution visualization of the internal structure was made possible by Frey, Perkins, and Mannella by using powerful three-dimensional (3D) electron microscope (EM) tomography techniques [74,75,76] (Figure 8).
![Figure 7
Internal structure of liver mitochondria in rat hepatocytes based on the pioneer high-resolution scanning electron microscopy work of Lea and Hollenberg, 1989 [73] that clearly revealed tubular cristae. O, outer membrane; I, inner membrane; T, tubular cristae (arrows); L, lumens within tubular cristae. Modified with permission.](/document/doi/10.1515/bmc-2022-0021/asset/graphic/j_bmc-2022-0021_fig_007.jpg)
Internal structure of liver mitochondria in rat hepatocytes based on the pioneer high-resolution scanning electron microscopy work of Lea and Hollenberg, 1989 [73] that clearly revealed tubular cristae. O, outer membrane; I, inner membrane; T, tubular cristae (arrows); L, lumens within tubular cristae. Modified with permission.
![Figure 8
Internal structure of mitochondria in rat liver and chick cerebellum based on 3D EM tomography work of Frey, Mannella, and Perkins [74,75,76,77], reproduced with permission. (a) The model showing all cristae in yellow, the inner mitochondrial membrane in light blue, and the outer mitochondrial membrane in dark blue. (b) Outer membrane, inner membrane, and four representative cristae shown in different colors.](/document/doi/10.1515/bmc-2022-0021/asset/graphic/j_bmc-2022-0021_fig_008.jpg)
Internal structure of mitochondria in rat liver and chick cerebellum based on 3D EM tomography work of Frey, Mannella, and Perkins [74,75,76,77], reproduced with permission. (a) The model showing all cristae in yellow, the inner mitochondrial membrane in light blue, and the outer mitochondrial membrane in dark blue. (b) Outer membrane, inner membrane, and four representative cristae shown in different colors.
Further progress on the internal structure of mitochondria was made using immunolabeling and transmission electron microscopy [88,89]. It was shown that the OXPHOS Complexes I–V were preferentially localized in the cristae membranes, and hence, the cristae membranes were the principal site of OXPHOS [88,89]. The recent solution of the structure of the OXPHOS supercomplexes [6,7,8,9,10] by cryo-EM techniques led to further molecular insights. These developments enable us to link mitochondrial ultrastructure to function in OXPHOS and provide a novel raison d’etre for the organization of the OXPHOS complexes into supercomplexes (Section 3.6).
Furthermore, ion translocation into sealed vesicles formed by the cristae membranes in OXPHOS has different implications from postulating ion movement across the inner mitochondrial membrane into a permeable outer mitochondrial membrane that is open to the cytosol.
What is respiration? Linking mitochondrial ultrastructure to function in OXPHOS
The current view of mitochondrial respiration sees Complexes I–IV as being involved in electron-coupled proton translocation (Figure 9a) after which the process is finished. If respiration is defined in this way, then no likely explanation appears possible for multiple energy conservation and transfer pathways determined in this work (Tables 1 and 2 and Figure 6). Nor can the role of succinate anions (Figures 1–5) be rationalized, and a single energy pool is sufficient. However, the results of this work require at least two distinct pathways between the redox and ATP sides (Figure 6) to explain the site-specific inhibitory interaction of the guanidine derivatives with the respiratory chain in mitochondria and the release of the induced inhibition by uncouplers. How can this be realized?

(a) Traditional and (b) new model of respiration. Primary ion translocations are denoted by bold arrows, and secondary ion translocations by dashed arrows.
Obviously, at the level of protons, no distinction is possible, since all redox complexes are involved in the same function and carry out electron-coupled proton transport (Figure 9a). However, if there exists another step subsequently of anion translocation, that is, the proton transport step is coupled to translocation of metabolic anions against their concentration gradient on the redox side [21,36,39,42,50,70,78], then a distinction is possible. In particular, since succinate is a dicarboxylic acid anion, there is a need to translocate both succinate monoanions and succinate dianions back and forth across the cristae membrane by the redox Complexes I–IV and Complex V in order to maintain perfect homeostasis, as shown in the schematic of Figure 9b. The manifold central roles of succinate in cellular metabolism are being increasingly recognized by new research [57,59,78,90,91]. Thus, respiration should be viewed as the process of electron-coupled proton translocation–coupled dicarboxylic acid anion translocation (Figure 9b). Such coupling requires sensing of the local Δψ produced by primary H+ translocation by the redox enzyme complexes/subunits, for which the complexes need to be organized into supercomplexes. If the subunits are separated or are disassociated from each other or lie at great distance from each other, such local electrical potential sensing would be precluded. It should also be noted that the measurements [92,93,94,95] cannot distinguish between a transient, local Δψ [36,38,39,40,63,64,70,78,85,95], and a steady-state delocalized potential, Δφ [83,84]. Hence, care should be taken to consider both possibilities in the interpretation of biochemical data.
Finally, if OXPHOS Complexes I and IV are specialized to recognize, bind, and translocate succinate dianions coupled to protons, and Complexes II and III are involved in the coupled translocation of succinate monoanions and protons, respectively, then the differential inhibition induced by the guanidine derivatives, the site-selective, energy-linked release of inhibition by uncouplers, and the network of two energy transfer pathways (Figure 6) can be explained in a natural way. The new view of respiration would also allow maintenance of overall electrical neutrality within the intracristal space and the matrix space of mitochondria [63]. Furthermore, the consensus H+/2e– stoichiometry of 4 for Complexes I and IV, and H+/2e– = 2 for Complex III of the mitochondrial electron transport chain [67,96] would also find a ready explanation [43,78]. Such coupling of proton and anion/counteraction transport had been a longstanding prediction of the two-ion theory and torsional mechanism of energy transduction and ATP synthesis based on various considerations of theory, experiment, and computation. A new model of integrated mitochondrial function with a possible organization of the OXPHOS supercomplexes is shown in Figure 10.
![Figure 10
Model of supramolecular organization of Complexes I–V in the cristae membranes of mammalian mitochondria along with the overall molecular mechanism of their functioning based on Nath’s two-ion theory of energy coupling and ATP synthesis [21,63,64,65,66,67,68,69,70,117] involving succinate anions and protons. The structures of Complexes I–V are assembled from the structures of the respirasome supercomplex I1III2IV1 (PDB ID: 5XTH; CI in blue, CIII in gold, and CIV in magenta), the structure of Complex II (PDB ID: 1ZOY; CII in green), and the structure of the tetramer of ATP synthase, Complex V (PDB ID: 6J5K) [135]. Primary ion translocations by Complexes I–V are denoted by bold arrows, and secondary ion translocations by dashed arrows. The translocation of succinate dianions and succinate monoanions along with protons by the respective Complexes are clearly distinguished and provide a novel explanation of the differential interactions and effects of various pharmacological and uncoupling agents at discrete points of the respiratory chain found in this work (Figure 6).](/document/doi/10.1515/bmc-2022-0021/asset/graphic/j_bmc-2022-0021_fig_010.jpg)
Model of supramolecular organization of Complexes I–V in the cristae membranes of mammalian mitochondria along with the overall molecular mechanism of their functioning based on Nath’s two-ion theory of energy coupling and ATP synthesis [21,63,64,65,66,67,68,69,70,117] involving succinate anions and protons. The structures of Complexes I–V are assembled from the structures of the respirasome supercomplex I1III2IV1 (PDB ID: 5XTH; CI in blue, CIII in gold, and CIV in magenta), the structure of Complex II (PDB ID: 1ZOY; CII in green), and the structure of the tetramer of ATP synthase, Complex V (PDB ID: 6J5K) [135]. Primary ion translocations by Complexes I–V are denoted by bold arrows, and secondary ion translocations by dashed arrows. The translocation of succinate dianions and succinate monoanions along with protons by the respective Complexes are clearly distinguished and provide a novel explanation of the differential interactions and effects of various pharmacological and uncoupling agents at discrete points of the respiratory chain found in this work (Figure 6).
Biological implications in cell life, cell death and apoptosis, and mitochondrial diseases
The above has several biological implications in health and disease [17,18,19,20,22,23,24,26,27,28,29,33,34,97]. As discussed, a widely held view of the existence of supercomplexes is that they enhance catalysis by channeling mobile substrates [13,98]. However, it has also been asserted that no robust experimental information exists to justify this view [14]. Evidence from flux control analysis [11,99,100] and biochemical evidence supporting the existence of Coenzyme Q and cytochrome c pools [101] and functional segmentation of the pools [12] have also been challenged [16,102]. However, views questioning substrate channeling [16,102] have themselves been questioned, for instance, based on the different time scales of the processes being probed [103].
Other suggestions such as a decrease in ROS production in the context of Complex III [6] have also been disputed [14]. For instance, it has been asserted that, contrary to the suggestion of decreased ROS at the Q i site, it is the Q o site that dominates ROS production [104], not the Q i site.
The view that individual OXPHOS complexes have similar catalytic activities as in supercomplexes is based on assaying electron-coupled H+ activity only, but not possible subsequent coupling to anion translocation (Figure 9b). A number of reports on aging and mitochondrial neurodegenerative diseases show that integrated mitochondrial function is indeed compromised in diseased states, even though no defect exists at the level of the individual OXPHOS Complexes I–IV [15,25]. It is very difficult to explain these observations if supramolecular organization of the OXPHOS Complexes did not confer any functional advantages. The need to sense local potentials, for example, between Complexes II and III, in order to couple proton and anion transport, maintain homeostasis of both monoanionic and dianionic forms of succinate, and distribute and compartmentalize the processing of these anionic forms among Complexes I/IV and Complexes II/III (Figures 9b and 10) helps us progress beyond current narratives.
Another constructive possibility to progress beyond the inconclusive debates over the past 25 years and break the current impasse in the field is to consider pools not only in terms of material, for example, Coenzyme Q pools or cytochrome c pools and their functional segmentation [13]. It is important to also think, interpret, and integrate the available biochemical information in terms of energy pools, attempted in the present work (Figure 6 and Sections 3.3, 3.4, and 3.6).
The concept of local potentials in channels and transporters [105,106,107,108,109,110,111,112,113,114,115,116] has been shown to be important for a better understanding of apoptosis and cell death [22]. The integration of the above concepts in biological energy coupling, transduction, and ATP synthesis in cell life and death has also been discussed recently [22,70,117]. These ideas provide new avenues and may even be crucial to catalyze the progress of OXPHOS research in cell life, cell death, and mitochondrial dysfunction and disease.
Applications of the results to specific diseases: type 2 diabetes and AD
In response to referee comments and suggestions, the grand challenge of application of the novel biochemical and pharmacological results contained in this work to the supercomplex problem afforded by specific diseases, for example, type 2 diabetes and AD, has been attempted. Several caveats need to be made before such a task is initiated.
The above multifactorial diseases have specialized journals dedicated to them, and it is quite impossible to even summarize the recent medical advances. For instance, let us take the role of metformin in type 2 diabetes. Metformin has been the first-line oral therapy for the treatment of type 2 diabetes for over 60 years as per consensus American and European guidelines [118]. Its safety and efficacy have been studied extensively, and its action leading to the specific inhibition of mitochondrial Complex I has been documented thoroughly [71,72,119,120,121]. Almost 20,000 research articles have been published on metformin to date; however, despite this large body of research, its exact mechanism of action is not completely understood [122]. The uncertainties have to do with a large number of factors. For example, studies of metformin performed on submitochondrial particles may not retain the supramolecular organization of intact mitochondria dealt with in this work. Furthermore, the concentrations of the drug used in cells and in mitochondrial studies vary substantially, and studies that correlate the metformin concentration in tissues/cells to the metformin-induced inhibition of mitochondrial Complex I are not readily available. In this sense, intact mitochondrial studies on supramolecular organization of the OXPHOS complexes, as in this work, stand on surer ground. Furthermore, unlike in our experiments, Complex I inhibitors such as rotenone cannot be used in a clinical medicine setting in view of their toxicity effects. Similarly, cyclosporine A (CsA), a routinely used inhibitor of the permeability transition (PT) in the laboratory, has little therapeutic potential owing to its immunosuppressive side effects and cannot be used in long-term diseases such as hyperglycemia-induced angiopathy.
In this section, it is assumed that the clinical effects of the anti-diabetic drug metformin [118,121,122] and the early events in the neurodegenerative diseases such as AD and Parkinson’s disease have a mitochondrial origin and involve mitochondrial network dysfunctions (for reviews, see refs. [17,123]). It is also taken as given that the concepts of the author’s work on ATP synthesis and OXPHOS in cell life, apoptosis, and cell death ([22,40,44]; this work) can be integrated to achieve a better understanding. Thus, in the spirit of refs. [22,40], chemical agents with pro-apoptotic/necrotic and anti-apoptotic/necrotic effects can be discussed based on their inducing or inhibitory activities and regulatory effects and their interactions that have the ability to open or close the permeability transition pore (PTP) in mitochondria. This is also in line with the pioneering work of Bernardi and colleagues on cell death and the PTP (for a review, see ref. [24]).
Type 2 diabetes
Inhibitory compounds in apoptosis and cell death were classified into agents that block/inhibit the various components of the ATP synthasome, that is, the adenine nucleotide transporter, the phosphate exchanger, and the FOF1, and exert an indirect shutting-down effect on the redox electron transfer processes coupled to ATP synthesis [40]. However, as discussed in Sections 3.3 and 3.4, pharmacological evidence shows that these site-specific compounds can also affect respiration directly by interaction with the Complexes of the electron transport chain in mitochondria. Such direct interactions of metformin with respiratory Complex I in mitochondria inhibit the opening of the PTP and lie at the heart of the protective effects of metformin in hyperglycemia.
The blood glucose-lowering effect of metformin is well established [124]. However, the antidiabetic properties of metformin are not related only to blood glucose normalization but have also to do with the weak and specific inhibition metformin and related guanides exert on mitochondrial Complex I (Sections 3.3, 3.4 and refs. [119,120]). Thus, elevated glucose concentration in hyperglycemia and oxidative stress lead to PTP opening and cell death in several cell types, and therapeutic concentrations of metformin inhibit PTP opening and offer cytoprotection; combination therapy with metformin, and CsA totally prevented hyperglycemia-induced β-cell death on exposure to 30 mM glucose in insulinoma cell lines INS-1 [125,126].
The literature on diabetes shows that when experiments are designed in the laboratory to mimic conditions of a diabetic state, cells undergo apoptotic cell death by the opening of the PTP. In other words, the intrinsic mitochondrial apoptosis pathway is linked to glucose metabolism. Inhibition of mitochondrial Complex I by rotenone, CsA, metformin, or combinations thereof led to PTP inhibition and conferred cell protection [24,125].
The molecular mechanisms by which hyperglycemia-induced cell death and ischemia-reperfusion injury open PTP channels require further study. It has been proposed that metformin affects cell death by first activating the AMP-activated protein kinase (AMPK), which in turn causes the observed inhibitory effects on Complex I and PTP regulation [127,128]. However, AMPK is a cytoplasmic and nucleus protein and is not localized in the mitochondrion, so it is difficult to visualize the proposed effects on PTP opening/closing. Moreover, the addition of succinate that activates mitochondrial site 2 respiration and acts as a shunt bypassing site 1 (Complex I) respiration has been shown to abolish metformin-induced AMPK activation [129]. Hence, it is more likely that primary Complex I inhibition by metformin is responsible for the observed effects.
So how does the action of metformin on mitochondrial Complex I help prevent hyperglycemia-induced cell death? An important clue comes from the finding that deletion of the Ppif gene that encodes for the mitochondrial matrix protein cyclophylin D (CypD, that associates with the cristae membrane during PT) restores cell mass and prevents diabetes in mice [125]. CypD, a peptidyl prolyl isomerase, is the universal regulatory component of the mitochondrial permeability transition that favors the onset of PT [24]. CypD is the molecular target of the PTP inhibitory drug CsA [22,24]. It was found that both metformin and CsA target the same site/s on the PTP in Complex I, and these sites are shielded by CypD. Moreover, Bernardi and colleagues have shown that the inhibitory effects of rotenone/metformin and CsA are additive and complementary, with the maximum inhibition being constant, and that genetic ablation of CypD restored the inhibition of PTP by rotenone in tissues that were normally resistant to its effects [125]. This suggests that if CypD bound to its target on Complex I is displaced or detached by any treatment protocol, or genetically ablated such that the CypD bound to Complex I is no longer available for PTP opening leading to cell death, then rotenone/metformin and CsA can readily bind to the Complex I sites unmasked by CypD, and thereby exert their synergistic inhibitory effect. This latter effect leads to an inhibition of PTP opening and thereby offers protection to cells from cell death arising from oxidative stress, hyperglycemia, and other insults [22,24,125]. Ultimately, this serves to desensitize the cell to Ca2+ and leads to a pro-survival state of cells, as explained by Bernardi and Nath [22,24,44,125]. Hence, CypD is a valid molecular target for therapy that offers an effective pharmacological treatment for the prevention of type 2 diabetes.
AD
The therapy outlined in Section 3.8.1 relies on the inhibition of the opening of the PTP by various pharmacological compounds that lead to a pro-survival state of the cell. However, the opposite type of therapeutic intervention, where the drug or compound needs to help induce a pro-apoptotic state of the cell may be useful in other diseases. This may be the case in AD triggered by early events of the uptake and accumulation of amyloid β-peptide (Aβ) in mitochondria [130,131].
Research on AD shows that Aβ cannot be produced locally inside mitochondria, but must be imported into mitochondria. Specific import mechanisms through the translocase of outer mitochondrial membrane complexes in mitochondria have been uncovered [130]. Confocal microscopy studies have also imaged the association and colocalization of Aβ with the mitochondrial respiratory chain complexes [131]. It has also been found that mitochondrial dysfunction and bioenergetics deficits occur early in pathogenesis of AD and precede the development of plaque formation in female mouse models of the disease [132].
In another development, Yan and coworkers provided an important piece of evidence that Aβ binds and forms a complex with CypD in the cortical mitochondria of AD patients [133]. As discussed in Section 3.8.1, CypD favors the onset of PT. Some workers interpreted the interaction of Aβ with CypD as causing the formation of the PTP. However, the work of Bernardi and colleagues shows that lack of CypD desensitizes the PTP to Ca2+ [24,125]. In other words, the non-availability of CypD will increase the concentration of Ca2+ required to induce PT. How exactly the binding of matrix Ca2+ to the F1-portion of ATP synthase – the key factor for PT induction in our view – triggers the formation of the PTP has been explained by us previously [22,44]. Thus, since the CypD bound to Aβ is no longer available for the opening of the PTP, it is difficult to understand how the Aβ–CypD interaction can provoke the formation of PTP events leading to neuronal degeneration. Rather, chemical agents that interfere with the formation of the Aβ–CypD interaction should decrease neurotoxicity due to Aβ.
The above implies that an increased expression of CypD or selection of experimental conditions that free CypD from its interactions with Aβ and allow CypD to bind, interact, and shield sites, for example, on Complex I may have a beneficial effect. Such interventions commit CypD to assist in the opening of the PTP, generate a pro-apoptotic state of programmed cell death, and reduce or prevent mitochondrial dysfunction due to uptake of Aβ into the mitochondria of cells. Hence, the PTP is an important therapeutic target for the treatment of AD [134].
Section 3.8 has also shown that the work of Nath on ATP and cell life and the work of Bernardi on ATP and cell death – both independent efforts spanning >30 years – agree with and also complement each other. Integration of these works is worth exploring. Taken together, the work has the power to offer novel insights and further a deeper understanding of both physiological and diseased states in mitochondrial systems.
Conclusion
The following conclusions were arrived at in this work based on a biochemical and pharmacological investigation and analysis of OXPHOS:
The classical anionic OXPHOS inhibitors DNP, FCCP, and dicoumarol have been shown to exhibit pure competitive inhibition with substrate succinate anions.
A decrease in the intramitochondrial concentration of succinate is observed with increasing concentrations of uncoupler during respiration.
Taking (a) and (b) together, it is concluded that inhibition of the entry of succinate by uncoupler takes place in mitochondria.
It has been concluded that the uncoupler is translocated through access channels in the membrane in its anionic form exactly like a substrate anion. Once inside, the anion picks up a proton and leaves the mitochondrion in its neutral form by diffusion. A detailed quantitative kinetic model of the process of coupling and uncoupling has been worked out [81,82]. It ought to be stressed that the same mathematical equations are used to explain coupling and uncoupling in the unified framework, except for the imposition of the respective conditions during coupling/uncoupling.
Guanidine and alkylguanidine have been shown to be more effective in inhibiting mitochondrial respiration on NADH-linked substrates than succinate and preferentially inhibit the first energy conservation Site 1, or OXPHOS Complex I. The related pharmacological compound phenethylbiguanide preferentially inhibits energy transfer with succinate as substrate; that is, it inhibits energy conservation at Site 2 or Complexes II–III.
The site-selective guanidine derivatives have been concluded to unmask the site-specific action of uncouplers and to allow the grouping of uncoupling agents into two classes. Thus, uncouplers such as dicoumarol have been concluded to be ineffective in releasing the respiratory inhibition induced by octylguanidine; however, dicoumarol has the ability to fully release inhibition due to phenethylbiguanide. On the other hand, uncoupling agents such as DNP have the ability to release the inhibition induced by both types of guanidines, that is, octylguanidine and phenethylbiguanide.
The observations suggest a more complex scheme involving two separate and distinct networks of energy transfer pathways between Site 1 and Site 2 in the mitochondrial respiratory chain and the intermediate that energizes ATP synthesis by Complex V.
The results cannot be reconciled with the single-ion chemiosmotic theory that predicts only a single energy pool between respiration and ATP synthesis, but are shown to be explained by the two-ion theory of energy coupling and ATP synthesis.
The OXPHOS Complexes I and IV are proton-succinate dianion cotransporters, while Complexes (II–III) and Complex V (FOF1–ATP synthase) function as proton–succinate monoanion cotransporters.
The process of respiration should be viewed as one involving electron-coupled proton translocation–coupled dicarboxylic acid anion translocation.
A new raison d’etre for the supramolecular organization of the respiratory complexes in OXPHOS has been suggested. The coupling of proton and anion transport as in (j) requires sensing of the local potential, Δψ produced by primary H+ translocation by the redox enzyme Complexes/subunits, which necessitates that the complexes be organized into supercomplexes.
The local Δψ existing between primary and secondary ion translocations in the access channels is immediately changed to a transmembrane potential and recorded (or rather calculated) as a measure of the bulk delocalized Δφ, created by the conditions for its measurement/calculation (for a thorough discussion of this subtle deception, see refs. [42,85]).
With succinate as a dicarboxylic acid coupling anion, the requirement to translocate both succinate monoanions and succinate dianions (along with protons) back and forth across the cristae membrane by the redox Complexes I–IV and Complex V in order to maintain cellular homeostasis, regulation, and overall electroneutrality has been shown to be satisfied.
A new model of energy transfer in mitochondria linking mitochondrial ultrastructure to function in OXPHOS has been formulated. An integrated model of energy transfer and supramolecular organization of Complexes I–V in the cristae membranes of mammalian mitochondria along with the overall molecular mechanism of their functioning based on the two-ion theory of energy coupling and ATP synthesis involving succinate anions and protons has been proposed. The model better explains the differential interactions uncovered by biochemical and pharmacological investigation in this work (along with their differential release by uncouplers) and the new structural information obtained by recent advances in cryo-EM.
Some biological implications arising for cell life, cell death and apoptosis, and mitochondrial dysfunction and disease have been discussed.
The difficult challenge of applying the novel results of the present experimental study in conjunction with Nath’s unified theory of ATP synthesis/hydrolysis in cell life and cell death [22,40] to type 2 diabetes and AD has been taken up. A study exploring the integration of the work of Nath on ATP and cell life with the pioneering work of Bernardi and colleagues on cell death and ATP [24,125] is seen as a promising prospect for future research.
Acknowledgments
The author is grateful to the Villum Foundation, Copenhagen for generous financial support through a VELUX Visiting Professorship tenable at the Technical University of Denmark. He thanks J. Woodley for arranging facilities to conduct the research and J. Villadsen for his keen interest and for several helpful discussions.
-
Conflict of interest: Author states no conflict of interest.
-
Data availability statement: The data sets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.
References
[1] Schägger H, Pfeiffer K. Supercomplexes in the respiratory chains of yeast and mammalian mitochondria. EMBO J. 2000;19:1777–83.10.1093/emboj/19.8.1777Search in Google Scholar PubMed PubMed Central
[2] Schägger H. Respiratory chain supercomplexes. IUBMB Life. 2001;52:119–28.10.1080/15216540152845911Search in Google Scholar PubMed
[3] Schägger H, Pfeiffer K. The ratio of oxidative phosphorylation complexes I–V in bovine heart mitochondria and the composition of respiratory chain supercomplexes. J Biol Chem. 2001;276:37861–7.10.1074/jbc.M106474200Search in Google Scholar PubMed
[4] Schäfer E, Seelert H, Reifschneider NH, Krause F, Dencher NA, Vonck J. Architecture of active mammalian respiratory chain supercomplexes. J Biol Chem. 2006;281:15370–5.10.1074/jbc.M513525200Search in Google Scholar PubMed
[5] Rieger B, Shalaeva DN, Söhnel A-C, Kohl W, Duwe P, Mulkidjanian AY, et al. Lifetime imaging of GFP at CoxVIIIa reports respiratory supercomplex assembly in live cells. Sci Rep. 2017;7:46055.10.1038/srep46055Search in Google Scholar PubMed PubMed Central
[6] Letts JA, Fiedorczuk K, Sazanov LA. The architecture of respiratory supercomplexes. Nature. 2016;537:644–8.10.1038/nature19774Search in Google Scholar PubMed
[7] Gu J, Wu M, Guo R, Yan K, Lei J, Gao N, et al. The architecture of mammalian respirasome. Nature. 2016;537:639–43.10.1038/nature19359Search in Google Scholar PubMed
[8] Sousa JS, Mills DJ, Vonck J, Kühlbrandt W. Functional asymmetry and electron flow in the bovine respirasome. eLife. 2016;5:e21290.10.7554/eLife.21290Search in Google Scholar PubMed PubMed Central
[9] Wu M, Gu J, Guo R, Huang Y, Yang M. Structure of mammalian respiratory supercomplex I1III2IV1. Cell. 2016;167:1598–1609.10.1016/j.cell.2016.11.012Search in Google Scholar PubMed
[10] Guo R, Zong S, Wu M, Gu J, Yang M. Architecture of human mitochondrial respiratory megacomplex I2III2IV2. Cell. 2017;170:1247–57.10.1016/j.cell.2017.07.050Search in Google Scholar PubMed
[11] Lenaz G, Genova ML. Supramolecular organisation of the mitochondrial respiratory chain: a new challenge for the mechanism and control of oxidative phosphorylation. Adv Exp Med Biol. 2012;748:107–44.10.1007/978-1-4614-3573-0_5Search in Google Scholar PubMed
[12] Benard G, Faustin B, Galinier A, Rocher C, Bellance N, et al. Functional dynamic compartmentalization of respiratory chain intermediate substrates: implications for the control of energy production and mitochondrial diseases. Int J Biochem Cell Biol. 2008;40:1543–54.10.1016/j.biocel.2007.11.023Search in Google Scholar PubMed
[13] Enríquez JA. Supramolecular organization of respiratory complexes. Annu Rev Physiol. 2016;78:533–61.10.1146/annurev-physiol-021115-105031Search in Google Scholar PubMed
[14] Hirst J. Open questions: respiratory chain supercomplexes–why are they there and what do they do? BMC Biol. 2018;16:111.10.1186/s12915-018-0577-5Search in Google Scholar PubMed PubMed Central
[15] Kayser E-B, Sedensky MM, Morgan PG, Hoppel CL. Mitochondrial oxidative phosphorylation is defective in the long-lived mutant clk-1. J Biol Chem. 2004;279:54479–86.10.1074/jbc.M403066200Search in Google Scholar PubMed
[16] Blaza JN, Serreli R, Jones AJY, Mohammed K, Hirst J. Kinetic evidence against partitioning of the ubiquinone pool and the catalytic relevance of the respiratory-chain supercomplexes. Proc Natl Acad Sci USA. 2014;111:15735–40.10.1073/pnas.1413855111Search in Google Scholar PubMed PubMed Central
[17] Zanin M, Santos BFR, Balling R, Ollert M, He FQ, Diederich NJ, et al. Mitochondria interaction networks show altered topological patterns in Parkinson’s disease. Npj Systems Biol Appl. 2020;6:38.10.1101/2020.03.09.984195Search in Google Scholar
[18] Witkov C. Bond graphs: graphical cross-disciplinary systems modeling based on power and energy relations. Smithfield, RI, USA: GRC Physics Research and Education, Novel Research in Energy Topics; 2018.Search in Google Scholar
[19] Nath S. Interpretation of the mechanism of action of antituberculosis drug bedaquiline based on a novel two-ion theory of energy coupling in ATP synthesis. Bioeng Transl Med. 2019;4:164–70.10.1002/btm2.10106Search in Google Scholar PubMed PubMed Central
[20] Nesci S, Trombetti F, Algieri C, Pagliarani A. A therapeutic role for the F1FO-ATP synthase. SLAS Discov. 2019;24:893–903.10.1177/2472555219860448Search in Google Scholar PubMed
[21] Nath S. Electrophysiological experiments revalidate the two-ion theory of energy coupling and ATP synthesis. Function. 2022;3(2):zqac004.10.1093/function/zqac004Search in Google Scholar PubMed PubMed Central
[22] Nath S. A novel conceptual model for the dual role of FOF1-ATP synthase in cell life and cell death. Biomol Concepts. 2020;11:143–52.10.1515/bmc-2020-0014Search in Google Scholar PubMed
[23] Bruns K, Studtrucker N, Sharma A, Fossen T, Mitzner D, Eissman A, et al. Structural characterization and oligomerization of PB1-F2, a proapoptotic influenza A virus protein. J Biol Chem. 2007;282:353–63.10.1074/jbc.M606494200Search in Google Scholar PubMed
[24] Bernardi P, Carraro M, Lippe G. The mitochondrial permeability transition: recent progress and open questions. FEBS J. 2021. 10.1111/febs.16254.Search in Google Scholar PubMed
[25] Boumans H, Grivell LA, Berden JA. The respiratory chain in yeast behaves as a single functional unit. J Biol Chem. 1998;273:4872–7.10.1074/jbc.273.9.4872Search in Google Scholar PubMed
[26] Gómez LA, Monette JS, Chavez JD, Maier CS, Hagen TM. Supercomplexes of the mitochondrial electron transport chain decline in the aging rat heart. Arch Biochem Biophys. 2009;490:30–5.10.1016/j.abb.2009.08.002Search in Google Scholar PubMed PubMed Central
[27] Frenzel M, Rommelspacher H, Sugawa MD, Dencher NA. Ageing alters the supramolecular architecture of OxPhos complexes in rat brain cortex. Exp Gerontol. 2010;45:563–72.10.1016/j.exger.2010.02.003Search in Google Scholar PubMed
[28] Rosca MG, Vazquez EJ, Kerner J, Parland W, Chandler MP, Hoppel CL, et al. Cardiac mitochondria in heart failure: decrease in respirasomes and oxidative phosphorylation. Cardiovasc Res. 2008;80:30–9.10.1093/cvr/cvn184Search in Google Scholar PubMed PubMed Central
[29] McKenzie M, Lazarou M, Thorburn DR, Ryan MT. Mitochondrial respiratory chain supercomplexes are destabilized in Barth syndrome patients. J Mol Biol. 2006;361:462–9.10.1016/j.jmb.2006.06.057Search in Google Scholar PubMed
[30] Adhikari A, Mondal S, Darbar S, Pal SK. Role of nanomedicine in redox mediated healing at molecular level. Biomol Concepts. 2019;10:160–74.10.1515/bmc-2019-0019Search in Google Scholar PubMed
[31] Brzezinski P, Moe A, Ådelroth P. Structure and mechanism of respiratory III–IV supercomplexes in bioenergetic membranes. Chem Rev. 2021;121:9644–73.10.1021/acs.chemrev.1c00140Search in Google Scholar PubMed PubMed Central
[32] González-Rodriguez P, Zampese E, Stout KA, Guzman JN, Ilijic E, Yang B, et al. Disruption of mitochondrial complex I induces progressive Parkinsonism. Nature. 2021;599:650–6.10.1038/s41586-021-04059-0Search in Google Scholar
[33] Nesci S, Trombetti F, Pagliarani A, Ventrella V, Algieri C, Tioli G, et al. Molecular and supramolecular structure of the mitochondrial oxidative phosphorylation system: implications for pathology. Life. 2021;11:242.10.3390/life11030242Search in Google Scholar
[34] Pizzo P, Pozzan T. Mitochondrialand: what will come next? Function. 2021;3(1):zqab073.10.1093/function/zqab073Search in Google Scholar
[35] Nath S, Rohatgi H, Saha A. The torsional mechanism of energy transfer in ATP synthase. Curr Sci. 1999;77:167–9.Search in Google Scholar
[36] Nath S. The molecular mechanism of ATP synthesis by F1F0-ATP synthase: a scrutiny of the major possibilities. Adv Biochem Eng Biotechnol. 2002;74:65–98.10.1007/3-540-45736-4_4Search in Google Scholar
[37] Nath S, Jain S. The detailed molecular mechanism of ATP synthesis in the F0 portion of ATP synthase reveals a non-chemiosmotic mode of energy coupling. Thermochim Acta. 2002;394:89–98.10.1016/S0040-6031(02)00242-3Search in Google Scholar
[38] Nath S. Molecular mechanisms of energy transduction in cells: engineering applications and biological implications. Adv Biochem Eng Biotechnol. 2003;85:125–80.10.1007/3-540-36466-8_5Search in Google Scholar PubMed
[39] Nath S. The torsional mechanism of energy transduction and ATP synthesis as a breakthrough in our understanding of the mechanistic, kinetic and thermodynamic details. Thermochim Acta. 2004;422:5–17.10.1016/j.tca.2004.08.004Search in Google Scholar
[40] Nath S. The new unified theory of ATP synthesis/hydrolysis and muscle contraction, its manifold fundamental consequences and mechanistic implications and its applications in health and disease. Int J Mol Sci. 2008;9:1784–840.10.3390/ijms9091784Search in Google Scholar PubMed PubMed Central
[41] Nath SS, Nath S. Energy transfer from adenosine triphosphate: quantitative analysis and mechanistic insights. J Phys Chem B. 2009;113:1533–7.10.1021/jp809678nSearch in Google Scholar PubMed
[42] Nath S. Beyond the chemiosmotic theory: analysis of key fundamental aspects of energy coupling in oxidative phosphorylation in the light of a torsional mechanism of energy transduction and ATP synthesis − Invited review part 2. J Bioenerg Biomembr. 2010;42:301–9.10.1007/s10863-010-9295-6Search in Google Scholar PubMed
[43] Nath S. The thermodynamic efficiency of ATP synthesis in oxidative phosphorylation. Biophys Chem. 2016;219:69–74.10.1016/j.bpc.2016.10.002Search in Google Scholar PubMed
[44] Mehta R, Singh J, Nath S. Time-resolved oxygen exchange measurements offer novel mechanistic insights into enzyme-catalyzed ATP synthesis during photophosphorylation. J Phys Chem B. 2020;124:5139–48.10.1021/acs.jpcb.0c03505Search in Google Scholar PubMed
[45] Villadsen J, Nielsen J, Lidén G. Bioreaction engineering principles. 3rd edn. New York: Springer; 2011. Chapter 4.10.1007/978-1-4419-9688-6Search in Google Scholar
[46] Channakeshava C. New paradigm for ATP synthesis and consumption. J Biosci. 2011;36:3–4.10.1007/s12038-011-9015-3Search in Google Scholar PubMed
[47] Agarwal B. A role for anions in ATP synthesis and its molecular mechanistic interpretation. J Bioenerg Biomembr. 2011;43:299–310.10.1007/s10863-011-9358-3Search in Google Scholar PubMed
[48] Xu L, Liu F. The chemo-mechanical coupled model for F1F0-motor. Prog Biophys Mol Biol. 2012;108:139–48.10.1016/j.pbiomolbio.2012.01.002Search in Google Scholar PubMed
[49] Agarwal B. Revisiting the ‘chemiosmotic theory’: coupled transport of anion and proton for ATP synthesis. Bioenergetics. 2013;2:e116.10.4172/2167-7662.1000e116Search in Google Scholar
[50] Wray V. Commentary on “Oxidative phosphorylation revisited”. Biotechnol Bioeng. 2015;112:1984–5.10.1002/bit.25627Search in Google Scholar PubMed
[51] Jain S, Murugavel R, Hansen LD. ATP synthase and the torsional mechanism: resolving a 50-year-old mystery. Curr Sci. 2004;87:16–9.Search in Google Scholar
[52] Rao NM. Medical biochemistry. New Delhi: New Age International; 2014. p. 280–1.Search in Google Scholar
[53] Matta CF, Massa L. Information theory and the thermodynamic efficiency of biological sorting systems: case studies of the kidney and of mitochondrial ATP-synthase. In: Bagchi B, editor. Sustained energy for enhanced human function and activity. Amsterdam: Elsevier; 2017. p. 25–8.10.1016/B978-0-12-805413-0.00001-6Search in Google Scholar
[54] Ji S. The cell language theory: connecting mind and matter. Singapore: World Scientific; 2018. Chapter 3.10.1142/p758Search in Google Scholar
[55] Ivanischev VV. Problems of the mathematical description of the chemiosmotic theory. News Tula State Univ Natl Sci. 2018;3:129–35.Search in Google Scholar
[56] Hua X, Du G, Han J, Xu Y. Bioprocess intensification for whole-cell catalysis of catabolized chemicals with 2,4-dinitrophenol uncoupling. ACS Sustainable Chem Eng. 2020;8:15782–90.10.1021/acssuschemeng.0c06466Search in Google Scholar
[57] Bose HS, Marshall B, Debnath DK, Perry EW, Whittal RM. Electron transport chain Complex II regulates steroid metabolism. iScience. 2020;23:101295.10.1016/j.isci.2020.101295Search in Google Scholar PubMed PubMed Central
[58] Karapetyan L, Mikoyan G, Vassilian A, Valle A, Bolivar J, Trchounian A, et al. Escherichia coli Dcu C4-dicarboxylate transporters dependent proton and potassium fluxes and FOF1-ATPase activity during glucose fermentation at pH 7.5. Bioelectrochemistry. 2021;141:107867.10.1016/j.bioelechem.2021.107867Search in Google Scholar PubMed
[59] Levy WB, Calvert VG. Communication consumes 35 times more energy than computation in the human cortex, but both costs are needed to predict synapse number. Proc Natl Acad Sci USA. 2021;118:e2008173118.10.1073/pnas.2008173118Search in Google Scholar PubMed PubMed Central
[60] Apps S. Electric field of ATP synthase suggests enzyme has functions beyond catalysis. Chem World. 2022 Feb 28Search in Google Scholar
[61] Bertero E, Maack C. Rethinking Mitchell’s chemiosmotic theory: potassium dominates over proton flux to drive mitochondrial F1Fo-ATP synthase. Function. 2022;3(2):zqac012.10.1093/function/zqac012Search in Google Scholar PubMed PubMed Central
[62] Moerner WE. Personal communication to Sunil Nath at Stanford University, Stanford, CA, USA; 2017 June 13.Search in Google Scholar
[63] Nath S. Analysis of molecular mechanisms of ATP synthesis from the standpoint of the principle of electrical neutrality. Biophys Chem. 2017;224:49–58.10.1016/j.bpc.2017.03.002Search in Google Scholar PubMed
[64] Nath S. Two-ion theory of energy coupling in ATP synthesis rectifies a fundamental flaw in the governing equations of the chemiosmotic theory. Biophys Chem. 2017;230:45–52.10.1016/j.bpc.2017.08.005Search in Google Scholar PubMed
[65] Nath S. Optimality principle for the coupled chemical reactions of ATP synthesis and its molecular interpretation. Chem Phys Lett. 2018;699:212–7.10.1016/j.cplett.2018.03.068Search in Google Scholar
[66] Nath S. Coupling in ATP synthesis: test of thermodynamic consistency and formulation in terms of the principle of least action. Chem Phys Lett. 2019;723:118–22.10.1016/j.cplett.2019.03.029Search in Google Scholar
[67] Nath S. Integration of demand and supply sides in the ATP energy economics of cells. Biophys Chem. 2019;252:106208.10.1016/j.bpc.2019.106208Search in Google Scholar PubMed
[68] Nath S. Modern theory of energy coupling and ATP synthesis. Violation of Gauss’s law by the chemiosmotic theory and validation of the two-ion theory. Biophys Chem. 2019;255:106271.10.1016/j.bpc.2019.106271Search in Google Scholar
[69] Nath S. Consolidation of Nath’s torsional mechanism of ATP synthesis and two-ion theory of energy coupling in oxidative phosphorylation and photophosphorylation. Biophys Chem. 2020;257:106279.10.1016/j.bpc.2019.106279Search in Google Scholar
[70] Nath S. Energy landscapes and dynamics of ion translocation through membrane transporters: a meeting ground for physics, chemistry, and biology. J Biol Phys. 2021;47:401–33.10.1007/s10867-021-09591-8Search in Google Scholar
[71] Hollunger G. Guanidines and oxidative phosphorylations. Acta Pharmacol Toxicol (Copenhagen). 1955;11(Suppl 1):1–84.10.1111/j.1600-0773.1955.tb02972.xSearch in Google Scholar
[72] Chance B, Hollunger G. Inhibition of electron and energy transfer in mitochondria. J Biol Chem. 1963;238:432–8.10.1016/S0021-9258(19)84015-2Search in Google Scholar
[73] Lea PJ, Hollenberg MJ. Mitochondrial structure revealed by high-resolution scanning electron microscopy. Am J Anatomy. 1989;184:245–57.10.1002/aja.1001840308Search in Google Scholar
[74] Frey TG, Mannella CA. The internal structure of mitochondria. Trends Biochem Sci. 2000;25:319–24.10.1016/S0968-0004(00)01609-1Search in Google Scholar
[75] Mannella CA, Pfeiffer DR, Bradshaw PC, Moraru II, Slepchenko B, Loew LM. et al. Topology of the mitochondrial inner membrane: dynamics and bioenergetics implications. IUBMB Life. 2001;52:93–100.10.1080/15216540152845885Search in Google Scholar PubMed
[76] Perkins GA, Tjong J, Brown JM, Poquiz PH, Scott RT, Kolson DR, et al. The micro-architecture of mitochondria at active zones: electron tomography reveals novel anchoring scaffolds and cristae structured for high-rate metabolism. J Neurosci. 2010;30:1015–26.10.1523/JNEUROSCI.1517-09.2010Search in Google Scholar PubMed PubMed Central
[77] Ghochani M, Nulton JD, Salamon P, Frey TG, Rabinovitch A, Baljon ARC. Tensile forces and shape entropy explain observed crista structure in mitochondria. Biophys J. 2010;99:3244–54.10.1016/j.bpj.2010.09.038Search in Google Scholar PubMed PubMed Central
[78] Nath S, Villadsen J. Oxidative phosphorylation revisited. Biotechnol Bioeng. 2015;112:429–37.10.1002/bit.25492Search in Google Scholar
[79] Nath S. Entropy production and its application to the coupled nonequilibrium processes of ATP synthesis. Entropy. 2019;21(746):1–22.10.3390/e21080746Search in Google Scholar
[80] Nath S. Novel molecular insights into ATP synthesis in oxidative phosphorylation based on the principle of least action. Chem Phys Lett. 2022;796:139561.10.1016/j.cplett.2022.139561Search in Google Scholar
[81] Nath S. Molecular mechanistic insights into coupling of ion transport to ATP synthesis. Biophys Chem. 2018;241:20–6.10.1016/j.bpc.2018.07.006Search in Google Scholar
[82] Nath S. Molecular mechanistic insights into uncoupling of ion transport from ATP synthesis. Biophys Chem. 2018;242:15–21.10.1016/j.bpc.2018.08.006Search in Google Scholar
[83] Mitchell P. Chemiosmotic coupling in oxidative and photosynthetic phosphorylation. Biol Rev. 1966;41:445–502.10.1111/j.1469-185X.1966.tb01501.xSearch in Google Scholar
[84] Mitchell P. Bioenergetic aspects of unity in biochemistry: evolution of the concept of ligand conduction in chemical, osmotic and chemiosmotic reaction mechanisms. In: Semenza G, editor. Of oxygen, fuels and living matter, part 1. New York: John Wiley; 1981. p. 30–56.Search in Google Scholar
[85] Nath S. Beyond the chemiosmotic theory: analysis of key fundamental aspects of energy coupling in oxidative phosphorylation in the light of a torsional mechanism of energy transduction and ATP synthesis − invited review part 1. J Bioenerg Biomembr. 2010;42:293–300.10.1007/s10863-010-9296-5Search in Google Scholar
[86] Bal W, Kurowska E, Maret W. The final frontier of pH and the undiscovered country beyond. PLoS ONE. 2012;7:e45832.10.1371/journal.pone.0045832Search in Google Scholar
[87] Juhaszova M, Kobrinsky E, Zorov DB, Nuss HB, Yaniv Y, Fishbein KW, et al. ATP synthase K+ - and H+ -fluxes drive ATP synthesis and enable mitochondrial K+ -‘uniporter’ function: I. Characterization of ion fluxes. Function. 2022;3(2):zqab065.10.1093/function/zqab065Search in Google Scholar
[88] Gilkerson RW, Selker JML, Capaldi RA. The cristal membrane of mitochondria is the principal site of oxidative phosphorylation. FEBS Lett. 2003;546:355–8.10.1016/S0014-5793(03)00633-1Search in Google Scholar
[89] Vogel G, Bornhövd C, Neupert W, Reichert AS. Dynamic subcompartmentalization of the mitochondrial inner membrane. J Cell Biol. 2006;175:237–47.10.1083/jcb.200605138Search in Google Scholar
[90] Tretter L, Patocs A, Chinopoulos C. Succinate, an intermediate in metabolism, signal transduction, ROS, hypoxia, and tumorigenesis. Biochim Biophys Acta. 2016;1857:1086–101.10.1016/j.bbabio.2016.03.012Search in Google Scholar
[91] Zhou T, Mu X, You Q. Succinate: an initiator in tumorigenesis and progression. Oncotarget. 2017;8(32):53819–28.10.18632/oncotarget.17734Search in Google Scholar
[92] Kell DB. A protet-based, protonic charge transfer model of energy coupling in oxidative and photosynthetic phosphorylation. Adv Microb Physiol. 2021;78:1–177.10.1016/bs.ampbs.2021.01.001Search in Google Scholar
[93] Williams RJP. Some unrealistic assumptions in the theory of chemi-osmosis and their consequences. FEBS Lett. 1979;102:126–32.10.1016/0014-5793(79)80943-6Search in Google Scholar
[94] Williams RJP. Chemical advances in evolution by and changes in use of space during time. J Theor Biol. 2011;268:146–59.10.1016/j.jtbi.2010.09.021Search in Google Scholar PubMed
[95] Nath S. Coupling mechanisms in ATP synthesis: rejoinder to “Response to molecular-level understanding of biological energy coupling and transduction”. Biophys Chem. 2021;272:106579.10.1016/j.bpc.2021.106579Search in Google Scholar PubMed
[96] Hinkle PC, Arun Kumar M, Resetar A, Harris DL. Mechanistic stoichiometry of mitochondrial oxidative phosphorylation. Biochemistry. 1991;30:3576–82.10.1021/bi00228a031Search in Google Scholar PubMed
[97] Seyfried TN, Chinopoulos C. Can the mitochondrial metabolic theory explain better the origin and management of cancer than can the somatic mutation theory? Metabolites. 2021;11(9):572.10.3390/metabo11090572Search in Google Scholar PubMed PubMed Central
[98] Acin-Pérez R, Fernández-Silva P, Peleato ML, Pérez-Martos A, Enriquez JA. Respiratory active mitochondrial supercomplexes. Mol cell. 2008;32:529–39.10.1016/j.molcel.2008.10.021Search in Google Scholar PubMed
[99] Bianchi C, Genova ML, Parenti Castelli G, Lenaz G. The mitochondrial respiratory chain is partially organized in a supercomplex assembly: kinetic evidence using flux control analysis. J Biol Chem. 2004;279:36562–9.10.1074/jbc.M405135200Search in Google Scholar
[100] Lenaz G, Genova ML. Kinetics of integrated electron transfer in the mitochondrial respiratory chain: random collisions vs. solid state electron channeling. Am J Physiol Cell Physiol. 2007;292:C1221–39.10.1152/ajpcell.00263.2006Search in Google Scholar
[101] Lapuente-Brun E, Moreno-Loshuertos R, Acin-Pérez R, Latorre-Pellicer A, Colás C, Balsa E, et al. Supercomplex assembly determines electron flux in the mitochondrial electron transport chain. Science. 2013;340:1567–70.10.1126/science.1230381Search in Google Scholar
[102] Trouillard M, Meunier B, Rappaport F. Questioning the functional relevance of mitochondrial supercomplexes by time-resolved aanalysis of the respiratory chain. Proc Natl Acad Sci USA. 2011;108:E1027–34.10.1073/pnas.1109510108Search in Google Scholar
[103] Guo R, Gu J, Zong S, Wu M, Yang M. Structure and mechanism of mitochondrial electron transport chain. Biomed J. 2018;41:9–20.10.1016/j.bj.2017.12.001Search in Google Scholar
[104] Sarewicz M, Osyczka A. Electronic connection between the quinone and cytochrome C redox pools and its role in regulation of mitochondrial electron transport and redox signaling. Physiol Rev. 2015;95:219–43.10.1152/physrev.00006.2014Search in Google Scholar
[105] Zander K, Sherman MP, Tessmer U, Bruns K, Wray V, Prechtel AT, et al. Cyclophilin A interacts with HIV-1 Vpr and is required for its functional expression. J Biol Chem. 2003;278:43202–13.10.1074/jbc.M305414200Search in Google Scholar
[106] Dai G. Symmetry breaking in photoreceptor cyclic nucleotide-gated channels. Nature Struct Mol Biol. 2022;29:7–9.10.1038/s41594-021-00711-5Search in Google Scholar
[107] Jou D, Ferrer F. A simple nonequilibrium thermodynamic description of some inhibitors of oxidative phosphorylation. J Theor Biol. 1985;117:471–88.10.1016/S0022-5193(85)80155-7Search in Google Scholar
[108] Juretić D. Bioenergetics: a bridge across life and universe. Boca Raton, FL, USA: CRC Press; 2022.10.1201/9781351172769Search in Google Scholar
[109] Nath S, Schütte H, Hustedt H, Deckwer W-D. Application of continuous zone electrophoresis to preparative separation of proteins. Biotechnol Bioeng. 1993;42:829–35.10.1002/bit.260420707Search in Google Scholar
[110] Nath S, Schütte H, Hustedt H, Deckwer W-D, Weber G. Separation of enzymes from microorganism crude extracts by free-flow zone electrophoresis. Biotechnol Bioeng. 1996;51:15–22.10.1002/(SICI)1097-0290(19960705)51:1<15::AID-BIT2>3.0.CO;2-3Search in Google Scholar
[111] Nath S. Complexation behavior of proteins with polyelectrolytes and random acrylic polyampholytes using turbidimetric titration. J Chem Tech. Biotechnol. 1995;62:295–300.10.1002/jctb.280620313Search in Google Scholar
[112] Nath S, Schütte H, Hustedt H, Deckwer W-D. Separation of enzymes from Candida boidinii crude extract in continuous-flow zone electrophoresis. Electrophoresis. 1990;11:937–41.10.1002/elps.1150111110Search in Google Scholar
[113] Nath S, Schütte H, Hustedt H, Deckwer W-D. Correlation of migration behavior in free-flow zone electrophoresis and electrophoretic titration curve. Electrophoresis. 1990;11:612–6.10.1002/elps.1150110805Search in Google Scholar
[114] Dierks T, Salentin A, Heberger C, Krämer R. The mitochondrial aspartate/glutamate and ADP/ATP carrier switch from obligate counterexchange to unidirectional transport after modification by SH-reagents. Biochim Biophys Acta. 1990;1028:268–80.10.1016/0005-2736(90)90176-OSearch in Google Scholar
[115] Kasai M, Kawasaki T, Yamamoto K. Permeation of neutral molecules through calcium channel in sarcoplastic reticulum vesicles. J Biochem. 1992;112:197–203.10.1093/oxfordjournals.jbchem.a123877Search in Google Scholar
[116] Kolbe M, Besir H, Essen L-O, Oesterhelt D. Structure of the light-driven chloride pump halorhodopsin at 1.8 Å resolution. Science. 2000;288:1390–6.10.1126/science.288.5470.1390Search in Google Scholar
[117] Nath S. Network representation and analysis of energy coupling mechanisms in cellular metabolism by a graph-theoretical approach. Theory Biosci. 2022;141:1–12. 10.1007/s12064-022-00370-0.Search in Google Scholar
[118] Nathan DM, Buse JB, Davidson MB, Ferrannini E, Holman RR, Sherwin R, et al. Medical management of hyperglycemia in type 2 diabetes: a consensus algorithm for the initiation and adjustment of therapy: a consensus statement of the American Diabetes Association and the European Association for the Study of Diabetes. Diabetes Care. 2009;32:193–203.10.2337/dc08-9025Search in Google Scholar
[119] El-Mir MY, Nogueira V, Fontaine E, Averet N, Rigoulet M, Leverve X. Dimethylbiguanide inhibits cell respiration via an indirect effect targeted on the respiratory chain complex I. J Biol Chem. 2000;275:223–8.10.1074/jbc.275.1.223Search in Google Scholar PubMed
[120] Owen MR, Doran E, Halestrap AD. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem J. 2000;348:607–14.10.1042/bj3480607Search in Google Scholar
[121] Vial G, Detaille D, Guigas B. Role of mitochondria in the mechanism(s) of action of metformin. Front Endocrinol. 2019;10:294.10.3389/fendo.2019.00294Search in Google Scholar PubMed PubMed Central
[122] Fontaine E. Metformin-induced mitochondrial complex I inhibition: facts, uncertainties, and consequences. Front Endocrinol. 2018;9:753.10.3389/fendo.2018.00753Search in Google Scholar PubMed PubMed Central
[123] Demetrius LA, Magistretti PJ, Pellerin L. Alzheimer’s disease: the amyloid hypothesis and the inverse Warburg effect. Front Physiol. 2015;5:522.10.3389/fphys.2014.00522Search in Google Scholar PubMed PubMed Central
[124] Bailey CJ. Biguanides and NIDDM. Diabetes Care. 1992;15:755–75.10.2337/diacare.15.6.755Search in Google Scholar PubMed
[125] Li B, Chauvin C, De Paulis D, Bernardi P, Ovize M, Fontaine E, et al. Inhibition of complex I regulates the mitochondrial permeability transition through a phosphate-sensitive inhibitory site masked by cyclophilin D. Biochim Biophys Acta. 2012;1817:1628–34.10.1016/j.bbabio.2012.05.011Search in Google Scholar PubMed
[126] Fujimoto K, Chen Y, Polonsky KS, Dorn GW. Targeting cyclophilin D and the mitochondrial permeability transition enhances beta-cell survival and prevents diabetes in Pdx1 deficiency. Proc Natl Acad Sci USA. 2010;107:10214–9.10.1073/pnas.0914209107Search in Google Scholar PubMed PubMed Central
[127] Zhou G, Myers R, Li Y, Goodyear LJ, Moller DE, Moller DE, et al. Role of AMP-activated protein kinase in mechanism of metformin action. J Clin Invest. 2001;108:1167–74.10.1172/JCI13505Search in Google Scholar PubMed PubMed Central
[128] Viollet B, Guigas B, Sanz Garcia N, Leclerc J, Foretz M, Andreeilli F. Cellular and molecular mechanisms of metformin: an overview. Clin Sci. 2012;122:253–70.10.1042/CS20110386Search in Google Scholar PubMed PubMed Central
[129] Hinke SA, Martens GA, Cai Y, Finsi J, Heimberg H, Pipeleers D, et al. Methyl succinate antagonises biguanide-induced AMPK-activation and death of pancreatic beta-cells through restoration of mitochondrial electron transfer. Br J Pharmacol. 2007;150:1031–43.10.1038/sj.bjp.0707189Search in Google Scholar PubMed PubMed Central
[130] Petersen CA, Alikhani N, Behbahani H, Wiehager B, Pavlov PF, Alafuzoff I, et al. The amyloid β-peptide is imported into mitochondria via the TOM import machinery and localized to mitochondrial cristae. Proc Natl Acad Sci USA. 2008;105:13145–50.10.1073/pnas.0806192105Search in Google Scholar PubMed PubMed Central
[131] Tillement L, Lecanu L, Yao W, Gleeson J, Papadopoulos V. Spirostenol blocks mitochondrial uptake of Aβ in neuronal cells and prevents Aβ-induced impairment of mitochondrial function. Steroids. 2006;71:725–35.10.1016/j.steroids.2006.05.003Search in Google Scholar PubMed
[132] Yao J, Irwin RW, Zhao L, Nilsen J, Hamilton RC, Brinton RD. Mitochondrial bioenergetics deficit precedes Alzheimer’s pathology in female mouse model of Alzheimer’s disease. Proc Natl Acad Sci USA. 2009;106:14670–5.10.1073/pnas.0903563106Search in Google Scholar PubMed PubMed Central
[133] Du H, Yan SS. Mitochondrial permeability transition pore in Alzheimer’s disease: cyclophilin D and amyloid beta. Biochim Biophys Acta. 2010;1802:198–204.10.1016/j.bbadis.2009.07.005Search in Google Scholar PubMed PubMed Central
[134] Rao VK, Carlson EA, Yan SS. Mitochondrial permeability transition pore is a potential drug target for neurodegeneration. Biochim Biophys Acta. 2014;1842:1267–72.10.1016/j.bbadis.2013.09.003Search in Google Scholar PubMed PubMed Central
[135] Gu J, Zhang L, Zhong S, Guo R, Liu T, Yi J, et al. Cryo-EM structure of the mammalian ATP synthase tetramer bound with inhibitory protein IF1. Science. 2019;364:1068–75.10.1126/science.aaw4852Search in Google Scholar PubMed
© 2022 Sunil Nath, published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Research Articles
- Diagnostic accuracy of genetic markers for identification of the Lr46/Yr29 “slow rusting” locus in wheat (Triticum aestivum L.)
- NADPH-derived ROS generation drives fibrosis and endothelial-to-mesenchymal transition in systemic sclerosis: Potential cross talk with circulating miRNAs
- Effect of omega-3 fatty acids on the telomere length: A mini meta-analysis of clinical trials
- Analysis of differentially expressed genes and signaling pathways involved in atherosclerosis and chronic obstructive pulmonary disease
- The epigenetic dimension of protein structure
- Gamma-induced mutants of Bacillus and Streptomyces display enhanced antagonistic activities and suppression of the root rot and wilt diseases in pulses
- Corticosterone potentiates ochratoxin A-induced microglial activation
- Supercomplex supercomplexes: Raison d’etre and functional significance of supramolecular organization in oxidative phosphorylation
- Insights into functional connectivity in mammalian signal transduction pathways by pairwise comparison of protein interaction partners of critical signaling hubs
- The effects of supplementation of Nannochloropsis oculata microalgae on biochemical, inflammatory and antioxidant responses in diabetic rats
- Molecular epidemiology of human papillomavirus in pregnant women in Burkina Faso
- Review Articles
- Interaction of cervical microbiome with epigenome of epithelial cells: Significance of inflammation to primary healthcare
- Seaweeds’ pigments and phenolic compounds with antimicrobial potential
- The capture of host cell’s resources: The role of heat shock proteins and polyamines in SARS-COV-2 (COVID-19) pathway to viral infection
- Erratum
- Erratum to “Plant growth-promoting properties of bacterial endophytes isolated from roots of Thymus vulgaris L. and investigate their role as biofertilizers to enhance the essential oil contents”
- Special Issue on XXV Congress of the Italian Society for Pure and Applied Biophysics
- Low-temperature librations and dynamical transition in proteins at differing hydration levels
- The phosphoinositide PI(3,5)P2 inhibits the activity of plant NHX proton/potassium antiporters: Advantages of a novel electrophysiological approach
- Targeted photoimmunotherapy for cancer
- Calorimetry of extracellular vesicles fusion to single phospholipid membrane
- Calcium signaling in prostate cancer cells of increasing malignancy
- Oxygen diffusion pathways in mutated forms of a LOV photoreceptor from Methylobacterium radiotolerans: A molecular dynamics study
- A photosensitizing fusion protein with targeting capabilities
- Ion channels and neuronal excitability in polyglutamine neurodegenerative diseases
- Is styrene competitive for dopamine receptor binding?
- Diffusion of molecules through nanopores under confinement: Time-scale bridging and crowding effects via Markov state model
- Quantitative active super-resolution thermal imaging: The melanoma case study
- Innovative light sources for phototherapy
- Electrophysiological study of the effects of side products of RuBi-GABA uncaging on GABAA receptors in cerebellar granule cells
- Subcellular elements responsive to the biomechanical activity of triple-negative breast cancer-derived small extracellular vesicles
Articles in the same Issue
- Research Articles
- Diagnostic accuracy of genetic markers for identification of the Lr46/Yr29 “slow rusting” locus in wheat (Triticum aestivum L.)
- NADPH-derived ROS generation drives fibrosis and endothelial-to-mesenchymal transition in systemic sclerosis: Potential cross talk with circulating miRNAs
- Effect of omega-3 fatty acids on the telomere length: A mini meta-analysis of clinical trials
- Analysis of differentially expressed genes and signaling pathways involved in atherosclerosis and chronic obstructive pulmonary disease
- The epigenetic dimension of protein structure
- Gamma-induced mutants of Bacillus and Streptomyces display enhanced antagonistic activities and suppression of the root rot and wilt diseases in pulses
- Corticosterone potentiates ochratoxin A-induced microglial activation
- Supercomplex supercomplexes: Raison d’etre and functional significance of supramolecular organization in oxidative phosphorylation
- Insights into functional connectivity in mammalian signal transduction pathways by pairwise comparison of protein interaction partners of critical signaling hubs
- The effects of supplementation of Nannochloropsis oculata microalgae on biochemical, inflammatory and antioxidant responses in diabetic rats
- Molecular epidemiology of human papillomavirus in pregnant women in Burkina Faso
- Review Articles
- Interaction of cervical microbiome with epigenome of epithelial cells: Significance of inflammation to primary healthcare
- Seaweeds’ pigments and phenolic compounds with antimicrobial potential
- The capture of host cell’s resources: The role of heat shock proteins and polyamines in SARS-COV-2 (COVID-19) pathway to viral infection
- Erratum
- Erratum to “Plant growth-promoting properties of bacterial endophytes isolated from roots of Thymus vulgaris L. and investigate their role as biofertilizers to enhance the essential oil contents”
- Special Issue on XXV Congress of the Italian Society for Pure and Applied Biophysics
- Low-temperature librations and dynamical transition in proteins at differing hydration levels
- The phosphoinositide PI(3,5)P2 inhibits the activity of plant NHX proton/potassium antiporters: Advantages of a novel electrophysiological approach
- Targeted photoimmunotherapy for cancer
- Calorimetry of extracellular vesicles fusion to single phospholipid membrane
- Calcium signaling in prostate cancer cells of increasing malignancy
- Oxygen diffusion pathways in mutated forms of a LOV photoreceptor from Methylobacterium radiotolerans: A molecular dynamics study
- A photosensitizing fusion protein with targeting capabilities
- Ion channels and neuronal excitability in polyglutamine neurodegenerative diseases
- Is styrene competitive for dopamine receptor binding?
- Diffusion of molecules through nanopores under confinement: Time-scale bridging and crowding effects via Markov state model
- Quantitative active super-resolution thermal imaging: The melanoma case study
- Innovative light sources for phototherapy
- Electrophysiological study of the effects of side products of RuBi-GABA uncaging on GABAA receptors in cerebellar granule cells
- Subcellular elements responsive to the biomechanical activity of triple-negative breast cancer-derived small extracellular vesicles