Abstract
Mitochondrial permeability transition (MPT) is a phenomenon that the inner mitochondrial membrane (IMM) loses its selective permeability, leading to mitochondrial dysfunction and cell injury. Electrophysiological evidence indicates the presence of a mega-channel commonly called permeability transition pore (PTP) whose opening is responsible for MPT. However, the molecular identity of the PTP is still under intensive investigations and debates, although cyclophilin D that is inhibited by cyclosporine A (CsA) is the established regulatory component of the PTP. PTP can also open transiently and functions as a rapid mitochondrial Ca2+ releasing mechanism. Mitochondrial fission and fusion, the main components of mitochondrial dynamics, control the number and size of mitochondria, and have been shown to play a role in regulating MPT directly or indirectly. Studies by us and others have indicated the potential existence of a form of transient MPT that is insensitive to CsA. This “non-conventional” MPT is regulated by mitochondrial dynamics and may serve a protective role possibly by decreasing the susceptibility for a frequent or sustained PTP opening; hence, it may have a therapeutic value in many disease conditions involving MPT.
Keywords: mitochondrial permeability transition, permeability transition pore, non-conventional mitochondrial permeability transition, mitochondrial dynamics
Introduction
Mitochondria play a critical role in eukaryotic cells, most notably, providing cellular energy in the form of ATP via oxidative phosphorylation (OXPHOS). Complete oxidation of carbon-containing organic nutrients to CO2 occurs through the Krebs cycle in mitochondria, during which reducing equivalents (NADH and FADH2) containing high-energy electrons are generated. The electron transport chain (ETC) in the inner mitochondrial membrane (IMM) converts high-energy electrons into a proton motive force across the IMM that is used by the ATP synthase for ATP production. The essential requirement for the OXPHOS is the well “sealed” IMM to maintain the electrochemical gradient created by the ETC. Maintaining the proper mitochondrial membrane potential (MMP) is also necessary for in-and-out of ions and metabolites. Therefore, the IMM maintains relative impermeability, and permits selective passages of ions and metabolites only through the specified channels, carriers, and transporters in regulated manners.
Mitochondrial permeability transition (MPT) is a phenomenon that the IMM loses selective permeability. It is thought to occur via the opening of a high conductance mega-channel called permeability transition pore (PTP), which allows free passage of molecules of up to 1.5 kDa in size. The phenomenon of MPT was initially appreciated as mitochondrial swelling and a loss of membrane integrity in the presence of Ca2+ in isolated mitochondria [1–4]. However, it is relatively recent that MPT drew attention due to its involvement in cell death in pathology. MPT often results in necrotic cell death because of the cellular energy crisis caused by the inability of mitochondria to produce ATP [5]. MPT is also involved in apoptosis. Mitochondrial swelling from massive influx of water during MPT can cause rupture of the outer mitochondrial membrane (OMM), which releases apoptotic effectors residing in the intermembrane space to the cytosol [6].
Despite decades of studies, the molecular identity of PTP is still not fully understood. The original model involving voltage-dependent anion channel (VDAC) of the OMM, adenine nucleotide translocase (ANT) of the IMM, and cyclophilin D (CypD) of the matrix was widely accepted for a while until the studies using mouse gene knockout (KO) technology. Gene KO studies demonstrated that VDAC and ANT are dispensable for MPT, although it requires higher Ca2+ concentrations to induce MPT in ANT KO [7, 8]. Most notably, CypD deletion greatly decreases PTP opening [9–12], establishing that CypD, which is inhibited by cyclosporine A (CsA), is the major regulator of PTP, albeit not a core component of the pore. Hence, PTP is conventionally defined by CsA-inhibitable large-conductance non-selective channel.
More recently, additional models have been proposed for PTP [13–15]. However, debates still continue about which protein (or its subunit) is responsible for the formation of PTP or whether there are multiple PTPs. Considering the long history of PTP studies and a recent development in the renewed identity of the PTP, there are many reviews that the readers can refer to [16–19]. In this short article, we will briefly describe the current state of conventional PTP, and then focus on a less-appreciated CsA-insensitive form of MPT that is regulated by mitochondrial dynamics.
Mitochondrial Permeability Transition Pore
Original model
Electrophysiological studies demonstrated that PTP opens in multiple sub-conductance states [20, 21], suggesting that the PTP has a multi-subunit structure that assembles into different oligomerization states. The initial clue for the molecular identity of the PTP was the sensitivity of MPT to atractyloside and bongkrekic acid that bind to the ANT [22]. ANT was shown to form complex with VDAC at the contact site along with the peripheral benzodiazepine receptor (now known as translocator protein of 18 kDa (TSPO)) [23–25]. VDAC as a PTP component was also suggested from the patch-clamp observation that the maximal conductance of PTP was twice of the VDAC conductance and that the most abundant sub-conductance state of the PTP happened to be the half-maximal state [26, 27]. Inhibition of MPT by CsA was known early on [28–30], and the CsA target protein CypD in the matrix was considered a regulatory component of the PTP. This series of information formulated an idea that VDAC and ANT form a core pore complex that is regulated by TSPO in the OMM and CypD in the matrix and possibly further associated with additional proteins found at the contact site including mitochondrial creatine kinase, hexokinase, and Bcl-2-related proteins [31, 32]. As indicated above, studies of genetic deletions of core pore components ANT and VDAC, as well as TSPO disproved this hypothetical model [7, 8, 33], but CypD remains as a main regulator of the PTP [9–12]. Indirect evidence also showed that the VDAC is a dispensable part of the PTP [34]. Inorganic phosphate sensitizes Ca2+-induced PTP opening [29]. Phosphate carrier (PiC) in the IMM binds to CypD in a CsA-sensitive manner and was also proposed to be a part of the PTP complex [35]. However, genetic manipulation of PiC did not affect PTP [36, 37]. Another report showed that PiC KO does not block PTP opening, but increases resistance to MPT, indicating that it is not the pore-forming component [38]. Since then, renewed efforts to identify PTP components resulted in new models involving the ATP synthase and spastic paraplegia 7 (SPG7).
Newly proposed models of PTP
It was found that CypD binds to oligomeric forms of the ATP synthase in a CsA-sensitive manner [39], which resembles the regulation of PTP by CypD. Follow-up studies showed that purified dimers of ATP synthase reconstituted in the lipid bilayer exhibit the electrophysiological characteristics of the PTP, and proposed that ATP synthase dimers form the PTP [13]. Oligomycin sensitivity conferring protein (OSCP) of the peripheral stalk was shown to be the site of CypD binding. In this model, the channel is formed at the interface of the two ATP synthases that is formed by a collection of small subunits of the Fo complex linked to the peripheral stalk [40, 41]. However, genetic deletions of the OSCP, the b subunit of the peripheral stalk, or other components of the ATP synthase dimerizing domain did not prevent PTP opening, disputing this model [42–46].
The ATP synthase as a PTP was also proposed by another group. In this case, however, the c-ring of the Fo complex was proposed to be the PTP [14]. An earlier study demonstrated that depletion of ATPase c subunit blocks MPT to the similar extent to CsA, and decreases cell death and neuronal excitotoxicity, suggesting a critical role of the ATPase c subunit in MPT [47]. In this new study, reconstituted c-subunit ring was shown to form a voltage-sensitive channel that has a current similar to that of PTP. Similarly, silencing c subunit in cells prevented ionomycin-induced MPT, mitochondrial depolarization, and cell death in a CsA-sensitive manner [14]. It was proposed that the c-ring becomes the PTP when the F1 complex dissociates from the Fo complex, which occurs with sustained high matrix Ca2+ concentration [14]. However, this model has also been disputed by the study employing gene KO of the three genes encoding c-subunits. The cells completely lacking c-subunits were shown to maintain PTP [46].
An shRNA-based loss-of function screening for PTP yielded several candidates for PTP components [15]. Among those, shRNAs for SPG7, VDAC1, and CypD were shown to decrease ROS-sensitized, Ca2+-induced PTP opening. This study showed that SPG7 interacts with VDAC1 and CypD and proposed that this tripartite complex forms the PTP at the contact sites [15]. However, SPG7 depletion did not completely block PTP opening, but decreased it, similar to CypD KO, indicating that SPG7 may be a regulatory component instead of a pore component of the PTP. SPG7 is a subunit of m-AAA protease of the IMM. In another study, however, it was shown that the loss of m-AAA induces an accumulation of the constitutively active mitochondrial calcium uptake channel MCU complex, resulting in Ca2+ overload and PTP opening [48]. This study further showed that depletion of SPG7 has no effect on Ca2+-induced PTP opening [48]. Therefore, SPG7 as a putative PTP is still far from conclusive.
Low conductance transient PTP (tPTP)
MPT by PTP opening is considered a terminal process that ends up with cell death as mentioned at the beginning. However, PTP was shown to open in multiple sub-conductance states by electrophysiological studies [20, 21]. In addition to the high conductance maximal state, PTP was observed to operate reversibly, opened and closed at a half-maximal state in a very high frequency [27, 49]. The bursting “flickering” of the half-conductance activity of the PTP was thought to be due to a cooperative nature of the mega-channel opening and/or a potential binary structure of the PTP comprised of two pore units [26, 27]. This transient low-conductance mode of PTP opening (tPTP) has been shown to function as a Ca2+-releasing channel subsequent to mitochondrial Ca2+ uptake [50–53]. In isolated mitochondria, tPTP induces transient depolarization of mitochondria, causing Ca2+ release [50]. The known inhibitors of PTP opening, CsA and ADP, completely blocked transient depolarization and Ca2+ release, indicating that the PTP opening is responsible for this event [50, 53]. In cells, tPTP was responsible for Ca2+ efflux following mitochondrial uptake of Ca2+ released from the ER/SR, suggesting that tPTP plays a role in the amplification of intracellular Ca2+ signals [50, 53].
Mitochondrial dynamics and MPT
Mitochondrial dynamics is the cellular process describing dynamic changes in mitochondrial shape and location. Mitochondrial fission and fusion are the main processes that regulate mitochondrial size and number, and are mediated by mitochondria-associated dynamin-related large GTPases. Dynamin-like/related protein 1 (DLP1/Drp1) is a cytosolic protein that is recruited to the mitochondrial outer surface where it constricts and divides mitochondria for fission. Mitochondrial fusion requires two separate machineries for fusion of OMM and IMM. Two isoforms of mitofusin, Mfn1 and Mfn2, in the OMM and the optic atrophy 1 (OPA1) protein in the IMM mediate fusion of respective membranes. The mechanisms and additional factors necessary for mitochondrial fission and fusion can be found in recent review articles [54–59].
Mitochondrial fission and fusion are implicated in pathological conditions, as they are involved in regulating mitochondrial function and cell death. As mentioned earlier, MPT can lead to necrosis and apoptosis. It was first observed that mitochondria in cells undergoing apoptosis were fragmented and swollen, which was prevented by inhibition of mitochondrial fission, potentially implicating mitochondrial fission in MPT [60]. Additional studies showed that ischemia-reperfusion (I-R) injury and heart failure (HF) as well as neuronal excitotoxicity where the MPT is a major cell death mechanism were reduced by inhibiting fission (Drp1 siRNA / Drp1-K38A dominant negative / Drp1 chemical inhibitor mdivi-1) or increasing fusion (Mfn / OPA1 overexpression) [61–65]. Therefore, general notion is that mitochondrial fragmentation by fission sensitizes to cell death [66]. Although preventing mitochondrial fragmentation by fission/fusion manipulations is protective in pathological conditions, only limited studies directly assessed MPT in altered mitochondrial fission or fusion. It was found that expressing the dominant-negative fission mutant Drp1-K38A or the fusion protein Mfn1 or Mfn2 in cardiac cells decreased MPT and cell death upon simulated I-R injury [61]. A similar observation was made in HF model [65]. However, this study used a loss of MMP as a surrogate marker of MPT, which is not a direct measure of the MPT. Quenching of mitochondrial calcein fluorescence by cobalt upon MPT is a well-established method to test the MPT [67, 68]. Using this assay, it was shown that inhibition of mitochondrial fission by Drp1-K38A decreases MPT and cell death in hyperglycemia [69].
These studies demonstrated that mitochondrial dynamics plays a role in MPT, apparently acting upstream of PTP opening. Reactive oxygen species (ROS) is a major permissive factor for PTP opening [70–73]. Our studies showed that increased mitochondrial ROS levels in hyperglycemic conditions increase MPT and cell death, and that inhibiting mitochondrial fission blocks the ROS increase and MPT [69, 74]. Similar experiments in prediabetic heart demonstrate an increase in ROS production, which trigger mitochondrial swelling together with changes in fission/fusion protein expression, leading to cell death [75, 76]. As I-R injury and excitotoxicity also involve mitochondrial ROS increase, the reduced MPT by fission inhibition may be due to indirectly decreasing ROS production, not directly acting on PTP components. In another study, it is shown that overexpression of Drp1 decreases mitochondrial calcium retention capacity (mCRC) in permeabilized cells, indicating that increasing Drp1 levels promotes MPT [77]. This finding raises the possibility that Drp1 may directly regulate PTP opening; however, it is also plausible that the effect is secondary to Drp1 influencing on the mitochondrial Ca2+ influx mechanism. A recent study suggested that Drp1 promotes hypoxia-induced PTP opening by binding to potential PTP constituents, Bax and PiC, and releasing hexokinase II from the PTP [78]. Bax, PiC, and hexokinase have been suggested to be regulatory proteins for PTP; therefore, it is possible that Drp1 directly interacts with components in the PTP supercomplexes, and thereby changes pore opening probability. Further studies in identifying the interacting proteins in the PTP complexes (interactome) may help to elucidate and distinguish the direct and indirect effects of Drp1 in MPT (Fig. 1).
Figure 1. Mitochondrial dynamics and MPT.

Mitochondria undergo fission-fusion cycle. Various insults including hyperglycemia (HG), ischemia-reperfusion (I-R), and neuronal excitotoxicity drive fission and cause mitochondrial ROS overproduction and Ca2+ overload, leading to PTP opening and MPT for cell injury. During hypoxia, the fission protein Drp1 may interact with Bax and phosphate carrier (PiC), and lead to a release of hexokinase II (HKII) from mitochondria, which induces PTP opening. Bax may prevent fusion, allowing mitochondria to remain fragmented.
Transient loss and recovery of MMP has been observed in cells and isolated mitochondria [73, 79–85]. This ‘flickering’ of trans-membrane potential is reminiscent of aforementioned tPTP. However, there are mixed conclusions regarding its dependence on ROS, Ca2+ or MPT. We reported earlier that mitochondrial fission plays a profound role in regulating mitochondrial function by controlling this transient mitochondrial depolarization [86, 87]. We will describe a possibility of a new form of tPTP regulated by mitochondrial dynamics, and discuss potential mechanisms of how it controls mitochondrial function and thus its application for therapeutic strategy.
Transient depolarization of mitochondria: MMP flickering
Transient depolarization of mitochondria has been observed since late 1990s using the mitochondrial potentiometric probe tetramethylrhodamine ethyl ester (TMRE) or its methyl version (TMRM). Some of these observations attributed the transient depolarization to an artifact of ROS generated from excited dye molecules interacting with oxygen under high energy light [73, 82, 83, 88, 89]. In these cases, tPTP facilitated by ROS is responsible for transient depolarization and thus sensitive to CsA. In other studies, however, transient mitochondrial depolarization was shown to be spontaneous, and independent of ROS, Ca2+, or MPT [79, 84, 85]. In addition, spontaneous mitochondrial pH flashes occur concurrently with transient mitochondrial depolarization [90]. This matrix pH flash reflects a compensatory increase in proton pumping immediately following the spontaneous loss of MMP, so that the proton motive force can be maintained. The transient mitochondrial depolarization coupled to pH flash was shown to be independent of ROS, Ca2+, and MPT [90]. These observations indicate that there likely exist different forms of transient mitochondrial depolarization, and that, aside from the light-induced artifactual phenomenon, spontaneous flickering of MMP is a normal cellular process. Then, why do mitochondria do this? What is the role of mitochondrial dynamics in this?
MMP flickering: why and how?
Spontaneous transient depolarization (MMP flickering) is an intrinsic property of mitochondria [79, 84–87, 90, 91]. The MMP generated by electrogenic proton efflux by the ETC is the major component of the proton motive force that drives ATP synthesis [92]. Therefore, maintaining proper MMP is critical for mitochondrial and cellular health, as it supports not only ATP synthesis, but also in-and-out of ions and metabolites across the IMM, and mitochondrial protein import. Barring inhibitory conditions, the ETC is a self-regulating system, which strives to maintain the MMP within an adequate range, by accelerating the electron transport and proton extrusion when MMP is low whereas decelerating when it is high [93, 94]. An example of this property, in their extremities, is the experimental evaluation of the ETC activity by oxygen consumption measurement, in which the maximum capacity of electron transport is measured in the presence of an uncoupler (e.g. with FCCP), and the proton leak by blocking the proton channel activity of the ATP synthase (e.g. with oligomycin). Mitochondria do not function properly when their MMPs remain outside of the adequate range for a prolonged period. While sustained low MMP is unable to support mitochondrial function, excessively high MMP (hyperpolarization) is also harmful, because hyperpolarization increases the effective concentration of free radical semiquinones (QH•) at CoQ, FMN, and FAD of the ETC that can directly react with oxygen for ROS production [95, 96]. Mitochondrial hyperpolarization also increases the risk of Ca2+ overload for MPT. Therefore, one can envision that spontaneous transient depolarization of mitochondria is the cellular protective mechanism to prevent harmful ROS overproduction and mitochondrial Ca2+ overload caused by mitochondrial hyperpolarization.
MMP flickering is stochastic, occurring in random frequency and intervals [84, 86, 87]. Furthermore, there is no correlation between MMP flickering and the extent of MMP, as we observed that some mitochondria with low MMP flickers more robustly than those with high MMP and that some mitochondria with high MMP were quiescent [86, 87]. These observations suggest that individual mitochondria in a cell are in different energetic states. Indeed, in cardiac and skeletal muscle cells, MMP differs among the different subpopulations of mitochondria as well as their energetic characteristics and its Ca2+ and ROS regulation [97–99]. How this spatial difference in MMP is achieved is unclear. However, it is likely that the intracellular compartmentalization of mitochondria and thus MMP is dependent on spatial relationships with other organelles and localized metabolic processes under different energy demands, as well as on the cytoskeletal system. For example, Ca2+ is an important activator of mitochondrial metabolism [100, 101]; therefore, the energetic state of mitochondria in the vicinity of the ER and plasma membrane where locally high concentration of Ca2+ is available would be different from that in mitochondria away from them. The same would be true for mitochondria in the metabolic carrier-enriched areas. Furthermore, cytosolic Ca2+ regulates mitochondrial motility, and high Ca2+ concentrations stop mitochondrial movement [102], indicating that mitochondria may move to the high Ca2+ areas where they remain and provide energy locally [103]. Thus, dynamic changes in localized mitochondrial energetic states would contribute to the intracellular heterogeneity of the MMP.
As to how the MMP flickering occurs, we can speculate that the MMP above a ‘set ceiling’ for the maximum tolerable MMP evokes spontaneous transient depolarization. Indeed, we showed that transient mitochondrial depolarization occurs with a buildup of MMP through active electron transport [87]. Furthermore, oligomycin treatment that increases the MMP activates MMP flickering, whereas the complex III inhibitor antimycin A decreases it [87, 104], indicating that MMP flickering is likely caused by elevation of MMP. As just described, mitochondrial energetic states differ locally inside a cell. Therefore, it is likely that individual mitochondria have their own ‘set ceilings’ for the MMP, and a buildup of MMP above the set MMP ceiling would evoke transient depolarization (Fig. 2A). Individual mitochondria in a cell experience fluctuations of metabolic fluxes because of spatial and temporal differences of metabolic inputs [99]. Accordingly, MMPs of individual mitochondria also fluctuate. The randomness of the MMP flickering is likely from the differences in MMP ceilings as well as metabolic states and their fluctuations in individual mitochondria.
Figure 2. MMP flickering and the impact of mitochondrial connectivity on mitochondrial energetics.

(A) Fluctuation of metabolic input into mitochondria results in temporal changes in MMP. When MMP goes above a hypothetical MMP upper limit (green dashed line), mitochondria become depolarized to prevent Ca2+ overload and ROS production. (B) TMRE images of primary hepatocytes with and without Drp1-K38A expression. Ball-shape mitochondria that are mostly discrete are prevalent in control cells, and they undergo small, isolated flickering (arrowhead). Inhibiting fission by expressing Drp1-K38A induces mitochondrial interconnection and enlargement. Large-scale MMP flickering occurs in fission-deficient cells, as all connected mitochondria undergo simultaneous loss and recovery of MMP (bottom panels). Partly reproduced from reference #86. (C) Drawing illustration of (B). Spontaneous MMP flickering is a normal cellular process. Isolated MMP flickering in discrete mitochondria (grey mitochondrion) represents normal level of proton leak. When mitochondria become interconnected by fusion, all connected mitochondria are simultaneously depolarized during MMP flickering, which is reflected as increased proton leak in respiration measurement of a cell population.
The MMP flickering may possibly be the manifestation of the aforementioned tPTP that induces mitochondrial depolarization and functions as a mitochondrial Ca2+ releasing mechanism. Unlike tPTP, however, the spontaneous MMP flickering is insensitive to CsA [84, 86, 87, 90, 99], suggesting a distinct cellular process and possibly a novel form of tPTP. Our studies demonstrated that mitochondrial connectivity regulated by mitochondrial dynamics has a profound impact on MMP flickering and thus mitochondrial function.
The role of mitochondrial dynamics in MMP flickering and cellular energetics
While we observed rare but consistent occurrence of MMP flickering in normal conditions, the inhibition of mitochondrial fission, thereby increasing mitochondrial interconnection drastically augments it [86, 87, 104]. It is known that a single mitochondrion is electrically connected regardless of its length and size [105, 106]. In normal cells, spontaneous depolarization occurs in discrete mitochondria as localized events (Fig. 2B control cell). In contrast, with fission inhibition, many of mitochondria become interconnected, and thus, local depolarization spreads to all connected mitochondria (Fig. 2B Drp1-K38A cell), causing large-scale MMP flickering [86, 87, 104]. Therefore, mitochondrial interconnection amplifies small local flickering into large-scale MMP flickering (Fig. 2B and C).
In mitochondria with well-sealed IMM, translocation of protons through the ATP synthase to the matrix powers the phosphorylation of ADP, representing a “coupled” respiration. Protons can “leak” to the matrix away from the ATP synthase, decreasing the OXPHOS coupling (“uncoupling”). Proton leak is measured by leak respiration, which is the respiration (oxygen consumption) in the presence of oligomycin that blocks the proton translocation through the ATP synthase. As protons leak to the matrix, proton gradient and MMP become lower, and mitochondria are depolarized. Oligomycin does not block MMP flickering, indicating that the transient depolarization in MMP flickering occurs by proton leak. In normal cells, at any given instance, the population of mitochondria undergoing transient depolarization is small, having little impact on overall cellular energetics, which represents the normal cellular activity of the ETC. Upon increasing mitochondrial interconnection, however, the spatial amplification of MMP flickering in individual cells is collectively manifested as an increase of proton leak (i.e. respiration uncoupling) in respiration measurement of a cell population (Fig. 2C) [86, 87]. An increase of proton leak accelerates electron transport, and thus, decreases the chance of ectopic electron-oxygen reaction, preventing superoxide production in the ETC. This is likely the reason why inhibition of mitochondrial fission (or enhancing fusion) decreases ROS and is protective in hyperglycemic conditions, cardiac I-R, neuronal excitotoxicity and other pathologies [61–65]. Therefore, mitochondrial connectivity that is determined by mitochondrial dynamics is an important factor that regulates the OXPHOS activity.
Does MMP flickering represent a novel form of tPTP?
Genetic studies have thus far disputed all models of PTP including the recently proposed ones – ATP synthase dimer, c-ring of Fo-ATP synthase, and SPG7 [42, 43, 46, 48]. In addition, the presence of an “unregulated” PTP (versus conventional regulated PTP) that is CypD-independent and is not regulated by Ca2+ has been reported [107]. It was also shown that CsA-insensitive PTP or channels can be formed by experimental addition of palmitic acid, prooxidants, or mitochondrial targeting sequences to isolated mitochondria [108–110]. Possibly, one of the reasons for the ongoing unresolved identity of PTP might be the existence of multiple forms of pores [13–15, 27, 35, 107–111]. As mentioned above, MMP flickering is not inhibited by CsA [84, 86, 87, 90], indicating that it is not mediated by conventional tPTP opening. Then, the question is whether MMP flickering represents a novel form of tPTP.
The cobalt-quenching calcein assay has been used for testing MPT [67, 68]. This assay uses the property of cobalt ion quenching calcein fluorescence. IMM is normally impermeable to cobalt ions, but once PTP opens, cobalt ions as well as calcein molecules (MW 622 Da) freely go in and out of mitochondria and thus calcein fluorescence is quenched. Our experiments with adult cardiomyocytes indicate that depolarizing mitochondria concomitantly lose calcein fluorescence (Fig. 3). These observations support the idea that transient depolarization occurs through a non-specific pore, which we like to refer to as non-conventional tPTP (nc-tPTP). Isolated mitochondria were also used to test the presence of nc-tPTP. We loaded calcein to isolated heart mitochondria by using the membrane-permeable calcein-AM. Mitochondrial calcein fluorescence was stable without cobalt. The addition of cobalt caused a minimal loss of calcein fluorescence. However, upon addition of the complex I-linked substrates glutamate/malate to establish the electrochemical gradient along with cobalt in calcium-free buffer, we found that a significant number of mitochondria lost or decreased calcein fluorescence, indicating that those mitochondria underwent MPT (Fig. 4A). This loss of calcein fluorescence in energized mitochondria was mostly insensitive to CsA, indicating that the observed MPT was likely nc-tPTP (Fig. 4A). Fluorometric analyses also detected Ca2+- and CsA-independent nc-tPTP. Whereas Ca2+ caused a rapid loss of calcein fluorescence by CsA sensitive PTP opening, simply energizing mitochondria with glutamate/malate also induced a significant decrease of calcein fluorescence in the absence of Ca2+ (Fig. 4B), indicating spontaneous pore opening independent of Ca2+. CsA has no effect on this energy-dependent spontaneous pore opening. Of note, cobalt alone in non-energized mitochondria causes a small steady decrease of calcein fluorescence, likely reflecting background cobalt permeability and/or MPT of deteriorating mitochondria. Additionally, we observed that oligomycin does not inhibit nc-tPTP, but FCCP completely blocks it (Fig. 4C). The requirement of ETC substrates and MMP for nc-tPTP opening, and its insensitivity to CsA and independence of Ca2+ are the properties observed for MMP flickering. It can be speculated that MMP flickering occurs by nc-tPTP.
Figure 3. Mitochondrial depolarization during MMP flickering occurs by MPT.

Mitochondria of adult cardiomyocytes were co-labeled with TMRE and calcein. Time sequence imaging shows a loss of mitochondrial calcein fluorescence coinciding with MMP loss, demonstrating mitochondrial depolarization by PTP opening.
Figure 4. The nc-tPTP in isolated mitochondria.

(A) Isolated mouse heart mitochondria loaded with calcein (5μM calcein-AM) were imaged in the presence of 5mM glutamate/2.5mM malate plus CoCl2 (1mM) with and without CsA (1μM). A significant number of mitochondria show decreases or losses of calcein fluorescence regardless of CsA addition. Scale bar: 5μm. (B) Calcein fluorometry of isolated mouse heart mitochondria. The addition of CoCl2 along with glutamate/malate (arrow) in Ca2+-free buffer significantly decreases calcein fluorescence by nc-tPTP. Including CsA does not affect the nc-tPTP. Arrowheads indicate the addition of Triton X-100 to solubilize mitochondria. (C) FCCP addition completely inhibits nc-tPTP opening whereas oligomycin has little impact on it. Co: cobalt, GM: glutamate/malate, Ca: calcium, Olm: oligomycin
The identity and mechanism of nc-tPTP are currently unknown. It was found that MMP flickering requires the mitochondrial dynamics protein OPA1 [87, 90, 104]. OPA1 is a membrane remodeling dynamin-related protein that mediates IMM fusion and cristae maintenance [112–114]. Initially, the MMP flickering was proposed to be the process of MMP equilibration between fusing mitochondria through the fusion pore formed by OPA1 [90]. However, MMP flickering occurs in already fused interconnected mitochondrial networks of fission-deficient cells [87, 104]. Furthermore, fusion between IMMs would not cause MPT, and thus, it is unlikely that OPA1-mediated IMM fusion forms nc-tPTP. As mentioned earlier, mitochondrial fission and fusion play a critical role for the extent of MMP flickering by determining mitochondrial connectivity. While how OPA1 participates in MMP flickering and nc-tPTP is undefined, OPA1 may play an additional role possibly in membrane leak for nc-tPTP through its novel IMM remodeling activity [87]. Indeed, it has been shown that OPA1 along with components of the mitochondrial contact site and cristae organizing system (MICOS complex) can establish tight cristae junctions that cause heterogeneity in localized MMP and pH [115], which may enhance the susceptibility for nc-tPTP opening.
Therapeutic Perspectives of nc-tPTP
Sustained opening of the PTP is inherently pathologic as it causes mitochondrial dysfunction. I-R injury is caused by PTP opening, and therefore, it is predicted that targeting the PTP is beneficial in these pathological conditions by decreasing mitochondrial dysfunction and cell death [116–119]. However, except for CypD, the molecular identity of the PTP is still unresolved, which makes molecular targeting difficult. Surprisingly, in a clinical trial involving about 1,000 patients with acute myocardial infarction, cyclosporine A did not result in better clinical outcomes than those with placebo [120]. Although the molecular identity nc-tPTP is also unknown, it can be controlled by manipulating known mitochondrial dynamics proteins. In early reperfusion during I-R injury, increased metabolic influx along with limited ATP synthase activity causes mitochondrial hyperpolarization, Ca2+ overload, and ROS increase for PTP opening [121–123]. Under this condition, the nc-tPTP can serve as a relief valve for excess proton gradient and matrix Ca2+ during reperfusion, which would prevent PTP opening. Indeed, inhibiting mitochondrial fission, thereby increasing nc-tPTP, decreases PTP opening, ameliorating heart I-R injury in both cell and animal models [61, 64, 124]. In this context, the conventional tPTP functioning as Ca2+ releasing channel [50, 51, 125] would be a relief valve as well. However, during I-R injury, the opening of PTP is sustained not transient. Finally, PTP and tPTP are molecularly the same channel, which makes separate manipulations difficult.
The functional significance of mitochondrial dynamics has mainly been appreciated as to maintain healthy population of mitochondria, which involves the segregation of dysfunctional mitochondria by fission and mitochondrial content mixing by fusion [126–128]. However, as discussed in this review, mitochondrial dynamics can play a more direct and fundamental role in regulating mitochondrial energetics by modulating mitochondrial connectivity and mediating nc-tPTP. Unlike the conventional PTP, nc-tPTP is likely a cellular protective mechanism. As nc-tPTP affects mitochondrial proton leak, pathological conditions involving mitochondrial ROS overproduction benefit from manipulating nc-tPTP, including I-R and excitotoxic injuries and diabetes. In addition, many other pathological conditions are also accompanied by PTP opening and mitochondrial dysfunction at some stage of disease progression. This pathologic PTP opening causes mitochondrial dysfunction-induced mitochondrial fragmentation, decreasing the nc-tPTP effect, which predicts the exacerbation of injury by forming a vicious cycle. Therefore, we can postulate that shifts in balance between the extents of PTP (“push” to injury) and nc-tPTP (“pull” from injury) determine pathological outcomes (Fig. 5). An increase of nc-tPTP would put a brake on this vicious cycle and slow pathological progression. Manipulation of mitochondrial dynamics has been shown to benefit many pathological models. Possibly, those are the result of unknowingly targeting nc-tPTP. Identifying the true nature of nc-tPTP will further increase therapeutic opportunities for many diseases.
Figure 5. Opposite roles of PTP and nc-tPTP in pathology.

Sustained PTP opening results in mitochondrial dysfunction and tissue injury whereas nc-tPTP and tPTP may function as a relief valve for mitochondria. In PTP-prone conditions, shifting the balance to nc-tPTP may improve pathological outcomes.
Highlights.
This review summarizes the studies on a form of mitochondrial permeability transition (MPT) that occurs stochastically and transiently and is insensitive to cyclosporine A and Ca2+
We discuss the regulation of this non-conventional transient opening of permeability transition pore (nc-tPTP) by mitochondrial dynamics
We propose the “push and pull” concept of PTP (“push” to injury) and nc-tPTP (“pull” from injury) in controlling life and death of cells and therapeutic perspectives of nc-tPTP
Acknowledgments
This work was supported by NIH R01HL093671 to SSS and YY, and R21EY031483 to YY. We thank Dr. Joannes Hoek for the comments on the manuscript.
References
- 1.Hunter FE Jr. and Ford L, Inactivation of oxidative and phosphorylative systems in mitochondria by preincubation with phosphate and other ions. J Biol Chem, 1955. 216(1): p. 357–69. [PubMed] [Google Scholar]
- 2.Tapley DF, The effect of thyroxine and other substances on the swelling of isolated rat liver mitochondria. J Biol Chem, 1956. 222(1): p. 325–39. [PubMed] [Google Scholar]
- 3.Chappell JB and Crofts AR, Calcium Ion Accumulation and Volume Changes of Isolated Liver Mitochondria. Calcium Ion-Induced Swelling. Biochem J, 1965. 95: p. 378–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Crofts AR and Chappell JB, Calcium Ion Accumulation and Volume Changes of Isolated Liver Mitochondria. Reversal of Calcium Ion-Induced Swelling. Biochem J, 1965. 95: p. 387–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Lemasters JJ, V. Necrapoptosis and the mitochondrial permeability transition: shared pathways to necrosis and apoptosis. Am. J. Physiol, 1999. 276: p. G1–G6. [DOI] [PubMed] [Google Scholar]
- 6.Yang JC and Cortopassi GA, Induction of the mitochondrial permeability transition causes release of the apoptogenic factor cytochrome c. Free Radic Biol Med, 1998. 24(4): p. 624–31. [DOI] [PubMed] [Google Scholar]
- 7.Kokoszka JE, et al. , The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore. Nature, 2004. 427(6973): p. 461–465. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Baines CP, et al. , Voltage-dependent anion channels are dispensable for mitochondrial-dependent cell death. Nat Cell Biol, 2007. 9(5): p. 550–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Baines CP, et al. , Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature, 2005. 434(7033): p. 658–62. [DOI] [PubMed] [Google Scholar]
- 10.Schinzel AC, et al. , Cyclophilin D is a component of mitochondrial permeability transition and mediates neuronal cell death after focal cerebral ischemia. Proc Natl Acad Sci U S A, 2005. 102(34): p. 12005–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Basso E, et al. , Properties of the permeability transition pore in mitochondria devoid of Cyclophilin D. J Biol Chem, 2005. 280(19): p. 18558–61. [DOI] [PubMed] [Google Scholar]
- 12.Nakagawa T, et al. , Cyclophilin D-dependent mitochondrial permeability transition regulates some necrotic but not apoptotic cell death. Nature, 2005. 434(7033): p. 652–8. [DOI] [PubMed] [Google Scholar]
- 13.Giorgio V, et al. , Dimers of mitochondrial ATP synthase form the permeability transition pore. Proc Natl Acad Sci U S A, 2013. 110(15): p. 5887–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Alavian KN, et al. , An uncoupling channel within the c-subunit ring of the F1FO ATP synthase is the mitochondrial permeability transition pore. Proc Natl Acad Sci U S A, 2014. 111(29): p. 10580–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Shanmughapriya S, et al. , SPG7 Is an Essential and Conserved Component of the Mitochondrial Permeability Transition Pore. Mol Cell, 2015. 60(1): p. 47–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Morciano G, et al. , The mitochondrial permeability transition pore: an evolving concept critical for cell life and death. Biol Rev Camb Philos Soc, 2021. 96(6): p. 2489–2521. [DOI] [PubMed] [Google Scholar]
- 17.Hurst S, Hoek J, and Sheu SS, Mitochondrial Ca(2+) and regulation of the permeability transition pore. J Bioenerg Biomembr, 2017. 49(1): p. 27–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Bonora M, Giorgi C, and Pinton P, Molecular mechanisms and consequences of mitochondrial permeability transition. Nature Reviews Molecular Cell Biology, 2021. [DOI] [PubMed] [Google Scholar]
- 19.Bernardi P, Carraro M, and Lippe G, The mitochondrial permeability transition: Recent progress and open questions. FEBS J, 2021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Kinnally KW, Campo ML, and Tedeschi H, Mitochondrial channel activity studied by patch-clamping mitoplasts. J Bioenerg Biomembr, 1989. 21(4): p. 497–506. [DOI] [PubMed] [Google Scholar]
- 21.Zorov DB, et al. , Multiple conductance levels in rat heart inner mitochondrial membranes studied by patch clamping. Biochim Biophys Acta, 1992. 1105(2): p. 263–70. [DOI] [PubMed] [Google Scholar]
- 22.Hunter DR and Haworth RA, The Ca2+-induced membrane transition in mitochondria. I. The protective mechanisms. Arch Biochem Biophys, 1979. 195(2): p. 453–9. [DOI] [PubMed] [Google Scholar]
- 23.Bucheler K, Adams V, and Brdiczka D, Localization of the ATP/ADP translocator in the inner membrane and regulation of contact sites between mitochondrial envelope membranes by ADP. A study on freeze-fractured isolated liver mitochondria. Biochim. Biophys. Acta, 1991. 1056: p. 233–242. [DOI] [PubMed] [Google Scholar]
- 24.Brdiczka D, Contact sites between mitochondrial envelope membranes. Structure and function in energy- and protein-transfer. Biochim Biophys Acta, 1991. 1071(3): p. 291–312. [DOI] [PubMed] [Google Scholar]
- 25.McEnery MW, et al. , Isolation of the mitochondrial benzodiazepine receptor: association with the voltage-dependent anion channel and the adenine nucleotide carrier. Proc Natl Acad Sci U S A, 1992. 89(8): p. 3170–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Szabo I, Bernardi P, and Zoratti M, Modulation of the mitochondrial megachannel by divalent cations and protons. J Biol Chem, 1992. 267(5): p. 2940–6. [PubMed] [Google Scholar]
- 27.Szabo I and Zoratti M, The mitochondrial permeability transition pore may comprise VDAC molecules. I. Binary structure and voltage dependence of the pore. FEBS Lett, 1993. 330(2): p. 201–5. [DOI] [PubMed] [Google Scholar]
- 28.Fournier N, Ducet G, and Crevat A, Action of cyclosporine on mitochondrial calcium fluxes. J Bioenerg Biomembr, 1987. 19(3): p. 297–303. [DOI] [PubMed] [Google Scholar]
- 29.Crompton M, Ellinger H, and Costi A, Inhibition by cyclosporin A of a Ca2+−dependent pore in heart mitochondria activated by inorganic phosphate and oxidative stress. Biochem J, 1988. 255(1): p. 357–60. [PMC free article] [PubMed] [Google Scholar]
- 30.Broekemeier KM, Dempsey ME, and Pfeiffer DR, Cyclosporin A is a potent inhibitor of the inner membrane permeability transition in liver mitochondria. J Biol Chem, 1989. 264(14): p. 7826–30. [PubMed] [Google Scholar]
- 31.Marzo I, et al. , Bax and adenine nucleotide translocator cooperate in the mitochondrial control of apoptosis. Science, 1998. 281(5385): p. 2027–31. [DOI] [PubMed] [Google Scholar]
- 32.Zamzami N and Kroemer G, The mitochondrion in apoptosis: how Pandora’s box opens. Nat. Rev. Mol. Cell Biol, 2001. 2: p. 67–71. [DOI] [PubMed] [Google Scholar]
- 33.Sileikyte J, et al. , Regulation of the mitochondrial permeability transition pore by the outer membrane does not involve the peripheral benzodiazepine receptor (Translocator Protein of 18 kDa (TSPO)). J Biol Chem, 2014. 289(20): p. 13769–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Chiara F, et al. , Hexokinase II detachment from mitochondria triggers apoptosis through the permeability transition pore independent of voltage-dependent anion channels. PLoS One, 2008. 3(3): p. e1852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Leung AW, Varanyuwatana P, and Halestrap AP, The mitochondrial phosphate carrier interacts with cyclophilin D and may play a key role in the permeability transition. J Biol Chem, 2008. 283(39): p. 26312–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Gutierrez-Aguilar M, et al. , Genetic manipulation of the cardiac mitochondrial phosphate carrier does not affect permeability transition. J Mol Cell Cardiol, 2014. 72: p. 316–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Varanyuwatana P and Halestrap AP, The roles of phosphate and the phosphate carrier in the mitochondrial permeability transition pore. Mitochondrion, 2012. 12(1): p. 120–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Kwong JQ, et al. , Genetic deletion of the mitochondrial phosphate carrier desensitizes the mitochondrial permeability transition pore and causes cardiomyopathy. Cell Death Differ, 2014. 21(8): p. 1209–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Giorgio V, et al. , Cyclophilin D modulates mitochondrial F0F1-ATP synthase by interacting with the lateral stalk of the complex. J Biol Chem, 2009. 284(49): p. 33982–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Bernardi P, The mitochondrial permeability transition pore: a mystery solved? Front Physiol, 2013. 4: p. 95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Bernardi P, et al. , The Mitochondrial Permeability Transition Pore: Channel Formation by F-ATP Synthase, Integration in Signal Transduction, and Role in Pathophysiology. Physiol Rev, 2015. 95(4): p. 1111–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.He J, et al. , Permeability transition in human mitochondria persists in the absence of peripheral stalk subunits of ATP synthase. Proc Natl Acad Sci U S A, 2017. 114(34): p. 9086–9091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Carroll J, et al. , Persistence of the permeability transition pore in human mitochondria devoid of an assembled ATP synthase. Proc Natl Acad Sci U S A, 2019. 116(26): p. 12816–12821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Marchetti P, et al. , Apoptosis-associated derangement of mitochondrial function in cells lacking mitochondrial DNA. Cancer Res, 1996. 56(9): p. 2033–8. [PubMed] [Google Scholar]
- 45.Masgras I, Rasola A, and Bernardi P, Induction of the permeability transition pore in cells depleted of mitochondrial DNA. Biochim Biophys Acta, 2012. 1817(10): p. 1860–6. [DOI] [PubMed] [Google Scholar]
- 46.He J, et al. , Persistence of the mitochondrial permeability transition in the absence of subunit c of human ATP synthase. Proc Natl Acad Sci U S A, 2017. 114(13): p. 3409–3414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Bonora M, et al. , Role of the c subunit of the FO ATP synthase in mitochondrial permeability transition. Cell Cycle, 2013. 12(4): p. 674–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Konig T, et al. , The m-AAA Protease Associated with Neurodegeneration Limits MCU Activity in Mitochondria. Mol Cell, 2016. 64(1): p. 148–162. [DOI] [PubMed] [Google Scholar]
- 49.Zoratti M and Szabo I, The mitochondrial permeability transition. Biochim Biophys Acta, 1995. 1241(2): p. 139–76. [DOI] [PubMed] [Google Scholar]
- 50.Ichas F, Jouaville LS, and Mazat JP, Mitochondria are excitable organelles capable of generating and conveying electrical and calcium signals. Cell, 1997. 89(7): p. 1145–53. [DOI] [PubMed] [Google Scholar]
- 51.Ichas F, et al. , Mitochondrial calcium spiking: a transduction mechanism based on calcium-induced permeability transition involved in cell calcium signalling. FEBS Lett, 1994. 348(2): p. 211–5. [DOI] [PubMed] [Google Scholar]
- 52.Korge P, et al. , Protective role of transient pore openings in calcium handling by cardiac mitochondria. J Biol Chem, 2011. 286(40): p. 34851–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Lu X, et al. , Individual Cardiac Mitochondria Undergo Rare Transient Permeability Transition Pore Openings. Circ Res, 2016. 118(5): p. 834–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Tilokani L, et al. , Mitochondrial dynamics: overview of molecular mechanisms. Essays Biochem, 2018. 62(3): p. 341–360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Yapa NMB, et al. , Mitochondrial dynamics in health and disease. FEBS Lett, 2021. 595(8): p. 1184–1204. [DOI] [PubMed] [Google Scholar]
- 56.Fenton AR, Jongens TA, and Holzbaur ELF, Mitochondrial dynamics: Shaping and remodeling an organelle network. Curr Opin Cell Biol, 2021. 68: p. 28–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Chan DC, Mitochondrial Dynamics and Its Involvement in Disease. Annu Rev Pathol, 2020. 15: p. 235–259. [DOI] [PubMed] [Google Scholar]
- 58.Lee H and Yoon Y, Mitochondrial fission and fusion. Biochem Soc Trans, 2016. 44(6): p. 1725–1735. [DOI] [PubMed] [Google Scholar]
- 59.Pernas L and Scorrano L, Mito-Morphosis: Mitochondrial Fusion, Fission, and Cristae Remodeling as Key Mediators of Cellular Function. Annu Rev Physiol, 2016. 78: p. 505–31. [DOI] [PubMed] [Google Scholar]
- 60.Frank S, et al. , The role of dynamin-related protein 1, a mediator of mitochondrial fission, in apoptosis. Dev. Cell, 2001. 1(4): p. 515–525. [DOI] [PubMed] [Google Scholar]
- 61.Ong SB, et al. , Inhibiting mitochondrial fission protects the heart against ischemia/reperfusion injury. Circulation, 2010. 121(18): p. 2012–22. [DOI] [PubMed] [Google Scholar]
- 62.Jahani-Asl A, et al. , Mitofusin 2 protects cerebellar granule neurons against injury-induced cell death. J Biol Chem, 2007. 282(33): p. 23788–98. [DOI] [PubMed] [Google Scholar]
- 63.Jahani-Asl A, et al. , The mitochondrial inner membrane GTPase, optic atrophy 1 (Opa1), restores mitochondrial morphology and promotes neuronal survival following excitotoxicity. J Biol Chem, 2011. 286(6): p. 4772–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Zepeda R, et al. , Drp1 loss-of-function reduces cardiomyocyte oxygen dependence protecting the heart from ischemia-reperfusion injury. J Cardiovasc Pharmacol, 2014. 63(6): p. 477–87. [DOI] [PubMed] [Google Scholar]
- 65.Xu S, et al. , CaMKII induces permeability transition through Drp1 phosphorylation during chronic beta-AR stimulation. Nat Commun, 2016. 7: p. 13189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Picard M, et al. , Mitochondrial morphology transitions and functions: implications for retrograde signaling? Am J Physiol Regul Integr Comp Physiol, 2013. 304(6): p. R393–406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Petronilli V, et al. , Transient and long-lasting openings of the mitochondrial permeability transition pore can be monitored directly in intact cells by changes in mitochondrial calcein fluorescence. Biophys J, 1999. 76(2): p. 725–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Scorrano L, et al. , Commitment to apoptosis by GD3 ganglioside depends on opening of the mitochondrial permeability transition pore. J. Biol. Chem, 1999. 274: p. 22581–22585. [DOI] [PubMed] [Google Scholar]
- 69.Yu T, et al. , Mitochondrial fission mediates high glucose-induced cell death through elevated production of reactive oxygen species. Cardiovasc Res, 2008. 79(2): p. 341–351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Kowaltowski AJ, Castilho RF, and Vercesi AE, Opening of the mitochondrial permeability transition pore by uncoupling or inorganic phosphate in the presence of Ca2+ is dependent on mitochondrial-generated reactive oxygen species. FEBS Lett, 1996. 378(2): p. 150–2. [DOI] [PubMed] [Google Scholar]
- 71.Halestrap AP, Woodfield KY, and Connern CP, Oxidative stress, thiol reagents, and membrane potential modulate the mitochondrial permeability transition by affecting nucleotide binding to the adenine nucleotide translocase. J Biol Chem, 1997. 272(6): p. 3346–54. [DOI] [PubMed] [Google Scholar]
- 72.Scarlett JL, et al. , Alterations to glutathione and nicotinamide nucleotides during the mitochondrial permeability transition induced by peroxynitrite. Biochem Pharmacol, 1996. 52(7): p. 1047–55. [DOI] [PubMed] [Google Scholar]
- 73.Zorov DB, et al. , Reactive oxygen species (ROS)-induced ROS release: a new phenomenon accompanying induction of the mitochondrial permeability transition in cardiac myocytes. J. Exp. Med, 2000. 192: p. 1001–1014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Yu T, Robotham JL, and Yoon Y, Increased production of reactive oxygen species in hyperglycemic conditions requires dynamic change of mitochondrial morphology. Proc. Natl. Acad. Sci. U S A, 2006. 103: p. 2653–2658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Federico M, et al. , Calcium-calmodulin-dependent protein kinase mediates the intracellular signalling pathways of cardiac apoptosis in mice with impaired glucose tolerance. J Physiol, 2017. 595(12): p. 4089–4108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Federico M, et al. , CaMKII activation in early diabetic hearts induces altered sarcoplasmic reticulum-mitochondria signaling. Sci Rep, 2021. 11(1): p. 20025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Kong D, et al. , Regulation of Ca2+-induced permeability transition by Bcl-2 is antagonized by Drp1 and hFis1. Mol. Cell. Biochem, 2005. 272: p. 187–199. [DOI] [PubMed] [Google Scholar]
- 78.Duan C, et al. , Mitochondrial Drp1 recognizes and induces excessive mPTP opening after hypoxia through BAX-PiC and LRRK2-HK2. Cell Death Dis, 2021. 12(11): p. 1050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Buckman JF and Reynolds IJ, Spontaneous changes in mitochondrial membrane potential in cultured neurons. J. Neurosci, 2001. 21(14): p. 5054–5065. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Jacobson J and Duchen MR, Mitochondrial oxidative stress and cell death in astrocytes--requirement for stored Ca2+ and sustained opening of the permeability transition pore. J Cell Sci, 2002. 115(Pt 6): p. 1175–88. [DOI] [PubMed] [Google Scholar]
- 81.Diaz G, et al. , Homogeneous longitudinal profiles and synchronous fluctuations of mitochondrial transmembrane potential. FEBS Lett, 2000. 475(3): p. 218–24. [DOI] [PubMed] [Google Scholar]
- 82.De Giorgi F, Lartigue L, and Ichas F, Electrical coupling and plasticity of the mitochondrial network. Cell Calcium, 2000. 28(5–6): p. 365–70. [DOI] [PubMed] [Google Scholar]
- 83.Huser J, Rechenmacher CE, and Blatter LA, Imaging the permeability pore transition in single mitochondria. Biophys J, 1998. 74(4): p. 2129–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Hattori T, et al. , Repetitive transient depolarizations of the inner mitochondrial membrane induced by proton pumping. Biophys J, 2005. 88(3): p. 2340–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Vergun O, Votyakova TV, and Reynolds IJ, Spontaneous changes in mitochondrial membrane potential in single isolated brain mitochondria. Biophys J, 2003. 85(5): p. 3358–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Galloway CA, et al. , Transgenic control of mitochondrial fission induces mitochondrial uncoupling and relieves diabetic oxidative stress. Diabetes, 2012. 61(8): p. 2093–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Lee H and Yoon Y, Transient contraction of mitochondria induces depolarization through the inner membrane dynamin OPA1 protein. J Biol Chem, 2014. 289(17): p. 11862–11872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Huser J and Blatter LA, Fluctuations in mitochondrial membrane potential caused by repetitive gating of the permeability transition pore. Biochem J, 1999. 343 Pt 2: p. 311–7. [PMC free article] [PubMed] [Google Scholar]
- 89.Aon MA, et al. , Synchronized whole cell oscillations in mitochondrial metabolism triggered by a local release of reactive oxygen species in cardiac myocytes. J Biol Chem, 2003. 278(45): p. 44735–44. [DOI] [PubMed] [Google Scholar]
- 90.Santo-Domingo J, et al. , OPA1 promotes pH flashes that spread between contiguous mitochondria without matrix protein exchange. Embo J, 2013. 32(13): p. 1927–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Wang W, et al. , Superoxide flashes in single mitochondria. Cell, 2008. 134(2): p. 279–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Mitchell P, Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature, 1961. 191: p. 144–148. [DOI] [PubMed] [Google Scholar]
- 93.Nicholls DG, Mitochondrial membrane potential and aging. Aging Cell, 2004. 3: p. 35–40. [DOI] [PubMed] [Google Scholar]
- 94.Bagkos G, Koufopoulos K, and Piperi C, A new model for mitochondrial membrane potential production and storage. Med Hypotheses, 2014. 83(2): p. 175–81. [DOI] [PubMed] [Google Scholar]
- 95.Skulachev VP, Role of uncoupled and non-coupled oxidations in maintenance of safely low levels of oxygen and its one-electron reductants. Q Rev Biophys, 1996. 29(2): p. 169–202. [DOI] [PubMed] [Google Scholar]
- 96.Korshunov SS, Skulachev VP, and Starkov AA, High protonic potential actuates a mechanism of production of reactive oxygen species in mitochondria. FEBS Lett, 1997. 416(1): p. 15–8. [DOI] [PubMed] [Google Scholar]
- 97.Kuznetsov AV, et al. , Mitochondrial subpopulations and heterogeneity revealed by confocal imaging: possible physiological role? Biochim Biophys Acta, 2006. 1757(5–6): p. 686–91. [DOI] [PubMed] [Google Scholar]
- 98.Lombardi A, et al. , Characterisation of oxidative phosphorylation in skeletal muscle mitochondria subpopulations in pig: a study using top-down elasticity analysis. FEBS Lett, 2000. 475(2): p. 84–8. [DOI] [PubMed] [Google Scholar]
- 99.Romashko DN, Marban E, and O’Rourke B, Subcellular metabolic transients and mitochondrial redox waves in heart cells. Proc Natl Acad Sci U S A, 1998. 95(4): p. 1618–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.McCormack JG and Denton RM, The role of Ca2+ ions in the regulation of intramitochondrial metabolism and energy production in rat heart. Mol. Cell. Biochem, 1989. 89: p. 121–125. [PubMed] [Google Scholar]
- 101.Das AM, Regulation of the mitochondrial ATP-synthase in health and disease. Mol. Genet. Metab, 2003. 79: p. 71–82. [DOI] [PubMed] [Google Scholar]
- 102.Yi M, Weaver D, and Hajnoczky G, Control of mitochondrial motility and distribution by the calcium signal: a homeostatic circuit. J Cell Biol, 2004. 167(4): p. 661–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Yoon Y, Regulation of mitochondrial dynamics: Another process modulated by Ca2+ signal? Sci. STKE, 2005. 2005: p. pe18. [DOI] [PubMed] [Google Scholar]
- 104.Murata D, et al. , Mitochondrial Safeguard: a stress response that offsets extreme fusion and protects respiratory function via flickering-induced Oma1 activation. EMBO J, 2020. 39(24): p. e105074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Amchenkova AA, et al. , Coupling membranes as energy-transmitting cables. I. Filamentous mitochondria in fibroblasts and mitochondrial clusters in cardiomyocytes. J. Cell Biol, 1988. 107(2): p. 481–495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Skulachev VP, Mitochondrial filaments and clusters as intracellular power-transmitting cables. Trends Biochem. Sci, 2001. 26(1): p. 23–29. [DOI] [PubMed] [Google Scholar]
- 107.He L and Lemasters JJ, Regulated and unregulated mitochondrial permeability transition pores: a new paradigm of pore structure and function? FEBS Lett, 2002. 512(1–3): p. 1–7. [DOI] [PubMed] [Google Scholar]
- 108.Sultan A and Sokolove PM, Palmitic acid opens a novel cyclosporin A-insensitive pore in the inner mitochondrial membrane. Arch Biochem Biophys, 2001. 386(1): p. 37–51. [DOI] [PubMed] [Google Scholar]
- 109.Kushnareva YE and Sokolove PM, Prooxidants open both the mitochondrial permeability transition pore and a low-conductance channel in the inner mitochondrial membrane. Arch Biochem Biophys, 2000. 376(2): p. 377–88. [DOI] [PubMed] [Google Scholar]
- 110.Kushnareva YE, et al. , Signal presequences increase mitochondrial permeability and open the multiple conductance channel. Arch Biochem Biophys, 1999. 366(1): p. 107–15. [DOI] [PubMed] [Google Scholar]
- 111.Brustovetsky N and Klingenberg M, Mitochondrial ADP/ATP carrier can be reversibly converted into a large channel by Ca2+. Biochemistry, 1996. 35(26): p. 8483–8. [DOI] [PubMed] [Google Scholar]
- 112.Frezza C, et al. , OPA1 controls apoptotic cristae remodeling independently from mitochondrial fusion. Cell, 2006. 126: p. 177–189. [DOI] [PubMed] [Google Scholar]
- 113.Patten DA, et al. , OPA1-dependent cristae modulation is essential for cellular adaptation to metabolic demand. Embo J, 2014. 33(22): p. 2676–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Lee H, Smith SB, and Yoon Y, The Short Variant of the Mitochondrial Dynamin OPA1 Maintains Mitochondrial Energetics and Cristae Structure. J Biol Chem, 2017. 292(17): p. 7115–7130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Wolf DM, et al. , Individual cristae within the same mitochondrion display different membrane potentials and are functionally independent. EMBO J, 2019. 38(22): p. e101056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Baines CP, The mitochondrial permeability transition pore as a target of cardioprotective signaling. Am J Physiol Heart Circ Physiol, 2007. 293(2): p. H903–4. [DOI] [PubMed] [Google Scholar]
- 117.Townsend PA, et al. , Urocortin prevents mitochondrial permeability transition in response to reperfusion injury indirectly by reducing oxidative stress. Am J Physiol Heart Circ Physiol, 2007. 293(2): p. H928–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Walters AM, Porter GA Jr., and Brookes PS, Mitochondria as a drug target in ischemic heart disease and cardiomyopathy. Circ Res, 2012. 111(9): p. 1222–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Bonora M, et al. , Physiopathology of the Permeability Transition Pore: Molecular Mechanisms in Human Pathology. Biomolecules, 2020. 10(7). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Cung TT, et al. , Cyclosporine before PCI in Patients with Acute Myocardial Infarction. N Engl J Med, 2015. 373(11): p. 1021–31. [DOI] [PubMed] [Google Scholar]
- 121.Burwell LS, Nadtochiy SM, and Brookes PS, Cardioprotection by metabolic shut-down and gradual wake-up. J Mol Cell Cardiol, 2009. 46(6): p. 804–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Chouchani ET, et al. , Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature, 2014. 515(7527): p. 431–435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Pell VR, et al. , Moving Forwards by Blocking Back-Flow: The Yin and Yang of MI Therapy. Circ Res, 2016. 118(5): p. 898–906. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Disatnik MH, et al. , Acute inhibition of excessive mitochondrial fission after myocardial infarction prevents long-term cardiac dysfunction. J Am Heart Assoc, 2013. 2(5): p. e000461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Bernardi P and von Stockum S, The permeability transition pore as a Ca(2+) release channel: new answers to an old question. Cell Calcium, 2012. 52(1): p. 22–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Twig G, et al. , Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. Embo J, 2008. 27(2): p. 433–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Youle RJ and van der Bliek AM, Mitochondrial fission, fusion, and stress. Science, 2012. 337(6098): p. 1062–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Ni HM, Williams JA, and Ding WX, Mitochondrial dynamics and mitochondrial quality control. Redox Biol, 2015. 4: p. 6–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
