Skip to main content
Elsevier Sponsored Documents logoLink to Elsevier Sponsored Documents
. 2015 Jan;78:100–106. doi: 10.1016/j.yjmcc.2014.09.023

The mitochondrial permeability transition pore: Molecular nature and role as a target in cardioprotection

Paolo Bernardi a,b,, Fabio Di Lisa a,b,⁎⁎,1
PMCID: PMC4294587  PMID: 25268651

Abstract

The mitochondrial permeability transition (PT) – an abrupt increase permeability of the inner membrane to solutes – is a causative event in ischemia–reperfusion injury of the heart, and the focus of intense research in cardioprotection. The PT is due to opening of the PT pore (PTP), a high conductance channel that is critically regulated by a variety of pathophysiological effectors. Very recent work indicates that the PTP forms from the F-ATP synthase, which would switch from an energy-conserving to an energy-dissipating device. This review provides an update on the current debate on how this transition is achieved, and on the PTP as a target for therapeutic intervention. This article is part of a Special Issue entitled "Mitochondria: from basic mitochondrial biology to cardiovascular disease".

Abbreviations: ANT, adenine nucleotide translocase; CsA, cyclosporin A; CyP, cyclophilin; Drp1, dynamin-related protein 1; Δψm, mitochondrial membrane potential; ERK, extracellular signal regulated kinase; GSK, glycogen synthase kinase; IMM, inner mitochondrial membrane; I/R, ischemia–reperfusion; MMC, mitochondrial megachannel; OMM, outer mitochondrial membrane; PKA, cyclic AMP-dependent protein kinase; PKG, cyclic GMP-dependent protein kinase; PPIase, peptidylprolyl cis-trans isomerase; PT, permeability transition; PTP, permeability transition pore; ROS, reactive oxygen species; TSPO, transport protein of 18 kDa; VDAC, voltage-dependent anion channel

Keywords: Mitochondria, Permeability transition pore, Ischemia–reperfusion injury

Highlights

  • The mitochondrial permeability transition pore plays a key role in heart disease.

  • Existing models for the permeability transition pore are critically reviewed.

  • F-ATP synthase is the best molecular candidate for pore formation.

  • Cyclophilin D is an important pore regulator but not a structural component.

  • Cyclophilin inhibitors are promising but promiscuous drugs.

1. Introduction

The permeability transition (PT) is an abrupt increase of the inner mitochondrial membrane (IMM) permeability to solutes, which in mammalian mitochondria has a cutoff of about 1500 Da. Occurrence of the PT and its inhibition by adenine nucleotides is known since the 1950s [1,2], and the phenomenon has been investigated in a number of laboratories (e.g. [3–13]). The term “permeability transition” was introduced in 1979 by Haworth and Hunter, who carried out a thorough characterization of its basic features in heart mitochondria, and provided the important insight – which is today generally accepted – that the PT could be due to opening of an IMM channel, the PTP [14–17]. This hypothesis was confirmed by patch-clamp studies on mammalian mitoplasts, which revealed the presence of a high-conductance (≈ 1 nS) channel, the mitochondrial megachannel (MMC) [18,19]. The MMC possesses all the basic features of the PTP [20,21] including sensitivity to cyclosporin A (CsA) [22], and represents the electrophysiological equivalent of the pore [23]. The study of mitochondrial channels has greatly contributed to our understanding of mitochondrial physiology, and to the acceptance of the pore theory of the PT (see [24] for a recent review).

PTP opening is traditionally linked to mitochondrial dysfunction because its occurrence leads to mitochondrial depolarization, cessation of ATP synthesis, Ca2 + release, pyridine nucleotide depletion, inhibition of respiration and, in vitro at least, matrix swelling; in turn, swelling causes mobilization of cytochrome c, outer mitochondrial membrane (OMM) rupture and eventually release of proapoptotic proteins such as cytochrome c itself, endonuclease G and AIF [25,26]. It should be mentioned that these detrimental effects on energy conservation and cell viability are only seen for long-lasting openings of the PTP [27], while short-term openings – which have been documented both in isolated mitochondria and in situ [27–30] – may be involved in physiological regulation of Ca2 + and reactive oxygen species (ROS) homeostasis [31], and provide mitochondria with a fast mechanism for Ca2 + release [32–35]. The potential role of the PTP in heart injury has been recognized very early [36,37], well before the role of mitochondria in apoptosis was discovered [38–40]. PTP desensitization with CsA proved beneficial in heart ischemia–reperfusion injury, as well as in pre- and post-conditioning through mechanisms that await clarification [41–49].

Matrix Ca2 + is an essential permissive factor for PTP opening, but the role of mitochondrial “Ca2 + overload” as a causative event in I/R injury of the heart has recently been challenged. In MCU null mitochondria – where Ca2 + overload does not occur during reperfusion – the extent of necrosis was the same as that observed in the hearts from wild type littermates, and the cardioprotective effect of CyPD ablation was abrogated [50]. These surprising observations raise many issues that still await an answer, such as the cause of cell death, the mechanism of activation of mitochondrial metabolism and the mechanism of PTP opening in MCU null mice. Yet these experiments do show that cardiomyocyte cell death can occur without mitochondrial “Ca2 + overload”; and suggest that there is enough Ca2 + in the matrix of MCU null mitochondria to allow pore PTP opening, possibly a consequence of the burst of ROS that follows reperfusion [31].

2. Molecular nature of the permeability transition pore: the early days

The molecular nature of the PTP has been the matter of debate for the last 30 years. In the early 1990s Snyder and coworkers found that the peripheral benzodiazepine receptor, an OMM protein today called TSPO [51], copurified with the adenine nucleotide translocator (ANT) and the voltage-dependent anion channel (VDAC) in protocols based on detergent extraction followed by hydroxylapatite chromatography; radiolabeled high-affinity ligands of TSPO were recovered in fractions where TSPO could be detected together with VDAC and ANT [52]. This finding was of great interest because nanomolar concentrations of the same TSPO ligands affected the channel properties of MMC in electrophysiological experiments, suggesting that all of these proteins could be involved in formation of the PTP [53].

This suggestion was strengthened a few years later by work from the Brdiczka laboratory during the characterization of OMM and IMM “contact sites”, i.e. specialized structures where the two membranes form close contacts mediated by protein–protein interactions [54]. These sites would include hexokinase on the cytosolic surface of, and VDAC within, the OMM, creatine kinase and nucleoside diphosphate kinase in the intermembrane space, and ANT in the IMM; they were proposed to mediate channeling of adenine nucleotides to and from mitochondria [54–56]. The link with the PTP was made when the same laboratory showed that hexokinase-enriched fractions from low detergent extracts of mitochondria formed channels with the conductance expected of the PTP, and conferred permeability properties to liposomes that could be inhibited by N-methylVal-4-cyclosporin [57]. It must be stressed that the preparation contained a very large number of proteins, which makes assignment of the channel activity to a specific species quite problematic. Furthermore – and unlike the case of PTP – currents were inhibited rather than induced by atractylate, and the active fractions were not enriched in VDAC and/or ANT [57]. The same preparations were shown to also contain proteins of the Bcl-2 family [58], and this set of observations led to a model where the PTP would be a multiprotein complex spanning both mitochondrial membranes and comprising ANT, VDAC, TSPO, cyclophilin (CyP) D as well as hexokinase and Bcl-2 proteins [59]. This model did not stand the test of genetics, as a CsA-sensitive PT could be easily detected in the absence of ANT [60], VDAC [61,62] as well as of TSPO [63]. An alternative model is the formation of the PTP by the Pi carrier following its interaction with CyP-D and ANT [64]. However, results obtained by patch-clamp analysis of the reconstituted Pi carrier do not match the electrophysiological PTP features [65] and genetic deletion of the Pi carrier does not support the idea that this protein is essential for PTP formation [66].

Studies on Ppif−/− mice (Ppif is the unique gene encoding CyPD in the mouse) have demonstrated that this protein is an important modulator which sensitizes the PTP to Ca2 + and confers sensitivity to CsA, but not an essential pore component [67–70]. By following the interactions of the matrix CyPD with other mitochondrial proteins it has recently been possible to identify a novel structure for the PTP, which will be described in the following paragraph.

3. The permeability transition pore forms from F-ATP synthase

By monitoring the presence of CyPD in blue native gels of mitochondrial proteins Giorgio et al. discovered that CyPD interacts with the F-ATP synthase, and that it can be crosslinked to the stalk proteins b, d and OSCP [71]. Binding of CyPD to the F-ATP synthase required Pi, and caused a decrease of the enzyme's catalytic activity; while it was counteracted by CsA, which displaced CyPD and increased the catalytic activity [71]. It was then found that CyPD interacts with the OSCP subunit of F-ATP synthase [72]. Gel-purified dimers of F-ATP synthase incorporated into lipid bilayers displayed currents activated by Ca2 +, Bz-243 and phenylarsine oxide (but not atractylate) with a unit conductance of about 500 pS, which is identical to that of the bona fide mammalian MMC-PTP [72]. The channel-forming property is shared by purified F-ATP synthase dimers of yeast mitochondria, which also displayed Ca2 +-dependent currents of slightly lower conductance (about 300 pS) [73]. Furthermore, yeast strains lacking the e and/or g subunits, which are necessary for dimer formation, showed a remarkable resistance to PTP opening [73]. Although strains lacking subunits e [74] or g [75] display abnormal morphology, with balloon-shaped cristae and F-ATP synthase monomers distributed randomly in the membrane, they did develop a normal membrane potential [73], suggesting that the increased resistance to PTP opening may not depend on these structural differences. Based on these findings, it has been proposed that the PTP forms from F-ATP synthase dimers, possibly in the lipid region between two adjacent stalks [76].

The idea that the pore forms from the F-ATP synthase is also supported by two independent studies. Bonora et al. used targeted inactivation of the c subunit of F-ATP synthase – which forms the H+-transporting c ring of F-ATP synthases – to show that HeLa cells become resistant to PTP opening and cell death [77]; while Alavian et al. reconstituted the c subunit or the purified F-ATP synthase in liposomes, and measured Ca2 +-activated channels [78] with properties similar to those described by Giorgio et al. with purified dimers [72]. It is not possible to derive mechanistic insights about the nature of the PTP-forming channel from the study of Bonora et al. because the consequences of knockdown of the c subunit on other components of the F-ATP synthase and on other mitochondrial proteins were not addressed, and it is unclear whether and how many functional F-ATP synthases were left after the knockdown of the c subunit [77]. Alavian et al., on the other hand, suggested that the channel of the PTP forms within the c ring itself after Ca2 +-dependent extrusion of F1, i.e. of the γ subunit [78]. We think that this hypothesis is extremely unlikely for the following reasons:

  • Displacement of F1 from FO requires very drastic conditions, such as treatment with 2 M urea [79] yet a functional FOF1 complex can be easily reconstituted after treatment with urea, indicating that the γ/δ/ε subunit reinserts into FO. It is hard to envision a plausible mechanism through which matrix Ca2 + could cause release of F1, and then create within FO a channel that cannot be closed by subunit γ/δ/ε [78].

  • Alavian et al. reported that the “FO channel” can instead be closed by the β subunit, and suggested that this is the mechanism through which pore closure occurs in situ [78]. There are major problems with this proposal, because structural studies have established that subunit β does not interact with the c ring [80]; and it is not obvious where the free β subunit would come from, given the extreme resistance of the F1 subcomplex to denaturation. This hypothesis is also difficult to reconcile with the well established fact that PTP–MMC opening is readily and fully reversible upon chelation of Ca2 + in mitoplasts [21], intact mitochondria [81] as well as in reconstituted dimers of F-ATP synthase [72].

  • Channel openings were also seen with preparations of the whole F-ATP synthase, and these could be inhibited by CsA after the addition of Ca2 + [78]. If the mechanism of pore opening is “expulsion” of F1 by Ca2 +, it is not easy to explain how the current could be inhibited by CsA. Indeed, it is firmly established that CsA inhibits the pore by removing CyPD, which in turn interacts with F1, not FO [71,72].

  • F1 has binding sites that can accommodate directly the effects of Ca2 +, Mg2 +, adenine nucleotides and Pi; and through CyPD (un)binding those of H+, CsA and possibly of oxidants [82]. Any model of the pore must account for all inducers and inhibitors, which appear difficult to fit in the c ring, a multimer of identical c subunits.

  • Silencing of the ATP5E gene, which encodes the ε subunit, resulted in downregulation of the F-ATP synthase complex with accumulation of subunit c, yet mitochondria were more coupled [83], which is the opposite of what would be expected if the isolated c ring can form membrane channels.

  • McGeoch and coworkers have performed patch-clamp studies of highly purified c subunit (which gave a single silver-stained band and was validated by sequencing) with very different results, as currents were inhibited rather than activated by Ca2 + [84–86]. It is legitimate to wonder whether other F-ATP synthase components were present in the preparation of Alavian et al. that could explain this discrepancy.

Given that in our hands the PTP-MMC readily forms from F-ATP synthase dimers but not monomers [72]; and that inactivation of the “dimerization” subunits e and g in Saccharomyces cerevisiae increases resistance of the PTP to Ca2 + [73], we favor the idea that the pore forms at the interface between two monomers in the dimeric enzyme, as we will discuss further after covering the regulatory role of CyPD.

4. Modulation by cyclophilin D and cyclosporin A

CyPs are ubiquitous, conserved proteins possessing peptidyl prolyl cis-trans isomerase activity [87–89]. Sixteen isoforms have been found in man, and the most abundant (and the first to be discovered) is cytosolic CyPA [90]. The enzymatic activity of all CyPs is inhibited by CsA [91] and the CsA/CyPA complex inhibits the cytosolic phosphatase calcineurin [92]; as a result, NFAT is no longer dephosphorylated, an event that prevents its nuclear translocation causing immunosuppression [93,94].

Mammals possess a unique mitochondrial species called CyPD, which in the mouse is encoded by the Ppif gene (see [95] for a review). CyPD is the mitochondrial receptor for CsA and modulates the PTP, but it is not a structural pore component. Indeed, the PTP can still open in mitochondria from Ppif−/− mice, although higher Ca2 + loads are required [67–70]. Regulation by CyPD may be a relatively recent evolutionary event [96], since the PTPs of S. cerevisiae and D. melanogaster are insensitive to CsA [97,98]. As also discussed elsewhere, the effect of CsA on the mammalian PTP is best described as “desensitization” in the sense that the PTP can still occur but becomes more resistant to Ca2 +, Pi and other inducers [25,76]. This consideration is important because (i) CsA can desensitize but not block the PTP, and therefore lack of sensitivity to CsA does not necessarily imply that the PTP is not involved in the event being studied; and (ii) different cells express different levels of CyPD, and obviously only CyPD-expressing mitochondria can respond to CsA [99].

CyPD binds the lateral stalk of F-ATP synthase (OSCP, b and d subunits) [71]. Like PTP induction, binding requires Pi (and results in partial inhibition of ATP synthase activity); while CsA displaces CyPD resulting in enzyme reactivation [71]. We have recently identified the binding site of CyPD to the OSCP subunit, possibly in a region comprising helices 3 and 4 [72], which is also the binding site of Bz-423 [100,101]. In keeping with PTP formation by F-ATP synthase, decreased levels of OSCP halved the threshold Ca2 + load required for PTP opening [72].

As discussed in detail elsewhere [72,76] we assume that the PTP forms within the IMM at the interface between two adjacent FO sectors. Matrix Ca2 + would have an essential permissive role in PTP formation after binding to the catalytic Me2 +-binding site, which is usually occupied by Mg2 +, and could be influenced by OSCP. Our working hypothesis is that OSCP as such is a “negative” modulator, whose effect can be counteracted by binding of the “positive” effector CyPD (which indeed decreases the threshold Ca2 + required for PTP opening). Removal of OSCP, or CyPD binding to OSCP, would induce similar conformational effects on the rigid stalk proteins, leading to increased probability of PTP opening at the IMM — a working hypothesis that awaits experimental testing.

We note that PTP formation at the membrane interface between two stalks could also accommodate the PTP-modulating effects of fatty acids [5–7,102]. The role of cardiolipin should also be explored, as it stabilizes respiratory supercomplexes [103] and, due to its partitioning into high-curvature membrane regions, plays a role in cristae formation and morphology [104]. Indeed, cardiolipin increases the degree of oligomerization of F-ATP synthase by promoting the formation of extended dimer rows, which is compromised in D. melanogaster mutants defective for cardiolipin synthesis [104]. The high susceptibility of cardiolipin to oxidation might alter F-ATP synthase conformation, affecting in turn the PTP open probability.

5. Regulation of the permeability transition by the outer mitochondrial membrane

The PT is an inner membrane event, since it also occurs in mitoplasts, i.e. mitochondria stripped of the OMM [105], yet the OMM does play a role in pore modulation. Lê-Quôc and Lê-Quôc were the first to show that induction of the PTP by substituted maleimides requires the OMM [106]. We have confirmed that PTP opening by N-ethylmaleimide is no longer present in mitoplasts [63] and extended the “sensitizing” role of the OMM to protocols where the PTP is modulated by dicarboxylic porphyrins plus visible light, a treatment that leads to the efficient production of singlet oxygen [107]. Low light doses inactivate the PTP through degradation of hystidyl residues, which in turn prevents matrix cysteine oxidation [108], possibly through a conformational change; while higher light doses activate the PTP through OMM cysteine oxidation [109]. PTP activation strictly required an intact OMM, since the inducing effect of hematoporphyrin plus light was completely lost in mitoplasts [105]. Largely based on the effect of its ligands on the PTP, it was proposed that the OMM protein responsible for PTP sensitization was the peripheral benzodiazepine receptor (today called TSPO) [52,53,105,110–112], but this turned out to be incorrect.

It had been known for some time that many cellular effects of “TSPO ligands” were not due to an interaction with TSPO, suggesting the existence of a different site of action [113,114]. The relevant target for PTP modulation, however, was identified only recently by 2 independent sets of observations. The Glick laboratory discovered that “TSPO ligands” (including the benzodiazepine Bz-423 and PK11195) interact with the mitochondrial F-ATP synthase and affect its activity [100,101,115,116]; while Giorgio et al. demonstrated that Bz-423 mimics the PTP-inducing effects of CyPD, and is able to favor the Ca2 +- and oxidant-induced transition of F-ATP synthase dimers to unselective high-conductance channels [72]. Thus, TSPO ligands directly modulate the propensity of F-ATP synthase to form channels; and it may therefore come as no surprise that mitochondria from mice with genetic ablation of TSPO (which completely lack high-affinity binding sites for PK11195) respond with PTP induction to PK11195, N-ethylmaleimide and photooxidative stress [63]. Thus, TSPO is not responsible for the PTP-modulating effects of “TSPO ligands”, and the protein(s) responsible for pore induction by N-ethylmaleimide and photooxidative stress await clarification. A potential candidate is Abcb6, an ATP binding cassette transporter of the OMM involved in heme and porphyrin homeostasis [117].

It is worth mentioning that the vast majority of F-ATP synthase dimers is located in rows of oligomers inside cristae [75,118,119] where a direct interaction with the OMM is not possible. Thus, either the PTP forms in the small population of dimers facing the intermembrane space, where direct contact with the OMM can occur; or the effect of the OMM is exerted by controlling the diffusion of PTP-regulating metabolites and ions, including Ca2 + itself [120,121] in a process that would be greatly favored by, and contribute to, cristae remodeling [122].

The OMM may also affect the outcome of PTP opening without directly affecting its open-closed transitions. Indeed, it has been shown that the PT causes matrix swelling and OMM rupture only in the presence of Bax and Bak, whose genetic ablation confers OMM resistance to swelling, thus preventing organelle rupture and cell death [123].

6. Cyclosporin A and cardioprotection

Treatment with CsA confers remarkable protection against acute myocardial injury induced by post-ischemic reperfusion [37], and this is matched by genetic ablation of CyPD [67,69], which suggests that the protective effects of CsA depend on CyPD inhibition and, by inference, on PTP desensitization.

Cardioprotection by CsA has been described in a wide variety of experimental models and animal species including humans [124]. All in all, we think that CsA provided a terrific proof of concept about the causative role of the PTP in I/R injury of the heart, and as a target for cardioprotection. There may be exceptions, however, as lack of protection by CsA has been reported in in vivo I/R in rat [125] and pig hearts [126], while a recent report has instead confirmed the cardioprotective efficacy in pig hearts [127]. The basis for these discrepancies is not immediately obvious, but key elements could be the duration of ischemia, which determines severity of damage [128], and CsA dosage [37]. In an apparent paradox, in perfused hearts of Ppif−/− (CyPD null) mice the extent of necrosis increased when reperfusion was established after a relatively short ischemic episode (30 min), while the expected protection was observed when ischemia was prolonged to 60 min [128]. Additional studies should address the factors involved in mild I/R damage that are not affected by treatments targeting CyPD.

Regarding dosage, it should be recalled that the protective effect of CsA is observed in a very narrow dosage window; this was first shown in perfused rat hearts where protection was seen at 0.2 but not 1 μM [37], a finding that has been confirmed by several laboratories, including ours, and never disputed. It seems possible that the effective CsA concentration may differ in different species, and therefore that conditions defined as optimal in one model may prove ineffective in others. The limitations discussed above, however, must be seriously taken into account when devising possible clinical applications for CsA, also considering the potentially adverse effects of prolonged PTP inhibition in failing hearts [129].

7. Additional effects of cyclophilin D and cyclosporin A

In the mouse CyPD is encoded by a unique nuclear gene (Ppif), and the transcript includes a matrix targeting sequence that is removed from the protein after mitochondrial import [130]. Consistently, the protein localizes to the mitochondrial matrix as detected both by electron microscopy and by trypsin titrations of mitochondrial fractions [131,132]. Protection afforded by deletion of CyPD or its inhibition is generally referred to its localization in the mitochondrial matrix, which allows regulatory interactions with F-ATP synthase and the PTP. CyPD has been proposed as a key interactor for the Hsp90–Trap-1 complex [133] and for p53 [134]. By sequestering a relevant fraction of CyPD these proteins would reduce the probability of PTP opening thus favoring survival of cancer cells, a provocative hypothesis that is the subject of controversy [135].

Genetic ablation of CyPD also caused mild ER stress, as judged on the basis of increased phosphorylation of eIF2a and expression of GRP78 without changes of the potentially detrimental CHOP [136]. This mild ER stress appears necessary to protect the heart against I/R injury by preventing the severe ER stress associated with ischemia and reperfusion. This concept is supported by the observation that inhibition of ER stress by tauroursodeoxycholic acid abrogated cardioprotection resulting from CyPD ablation or inhibition, while a moderate ER stress induced by tunicamycin reduced post-ischemic reperfusion injury [136]. More recent work from the Ovize laboratory suggests that CyPD contributes to Ca2 + trafficking between ER/SR and mitochondria by binding to a protein complex which also includes VDAC1, Grp75 and IP3R1 [137]. This complex appears to be located at the interface between the OMM and the ER, and may facilitate the transfer of Ca2 + released from the ER into the mitochondrial matrix [138]. Interestingly, the CyPD–VDCA1–Grp75–IP3R1 interaction was facilitated by mitofusin 2, and increased under conditions of hypoxia and reoxygenation [137]. This complex may be extremely relevant because the genetic ablation or the pharmacological inhibition of any of its members decreased the occurrence of cell death induced by protocols of anoxia/reoxygenation or ischemia/reperfusion [137].

Whether this additional role for CyPD is relevant to PTP regulation in relation to Ca2 + is an open question. PTP opening does not depend on VDAC [61,62], and mitochondrial Ca2 + uptake occurs through MCU in a Δψm-dependent process [139,140]. PTP opening curtails mitochondrial Ca2 + uptake both by disrupting Δψm and by providing a possible pathway for Ca2 + efflux [32–35], consistent with the finding that CyPD ablation or inhibition leads to an increase of matrix [Ca2 +] [34,129,141].

An obvious question is whether can CyPD also be found outside the matrix and in the ER/SR. This question applies both to CyPD interactions with the VDAC1–Grp75–IP3R1 complex, and to those with other proteins shown to bind CyPD mostly by immunoprecipitation such as Bcl-2 [142], p53 [134,135], GSK-3 [131] and the Hsp90–Trap-1 complex [133]. These findings would imply that a fraction of CyPD is localized at sites other than the mitochondrial matrix, or that under specific pathophysiological conditions CyPD translocates to other cellular sites. Until such mechanism is identified, chances remain that results obtained with detergent extraction and immunoprecipitation may not reflect the protein–protein interactions involving CyPD in situ.

It should also be mentioned that the effects of CyPD on proteins located outside mitochondria need not be only due to direct protein–protein interactions, but might also be related to functional modifications induced by CyPD deletion or inhibition, as well as by covalent changes in response to the activation of signaling pathways [143]. These include (de)phosphorylation [131], (de)acetylation [144], nitrosylation [82,145] and oxidation of Cys203, which inhibits the enzymatic activity of CyPD and plays a crucial role in favoring PTP opening. Importantly, expression of a CyPDC203S mutant desensitized the PTP to both Ca2 +- and H2O2 providing a mechanistic link between oxidative stress and PTP opening [82]. This wide array of CyPD modifications, which has largely been characterized in E. Murphy's laboratory, is likely to affect mitochondrial metabolism, eventually resulting in variations of responses to physiological and pathological stimuli. In keeping with this possibility ablation of CyPD causes significant changes in the level of mitochondrial proteins involved in intermediary metabolism [146] as well as in the mitochondrial acetylome [147]. Interestingly, increased acetylation of β-hydroxyacyl-CoA dehydrogenase resulted in a 50% decrease of its activity [147], which may explain the decreased fatty acid oxidation observed in Ppif−/− hearts [129].

CsA inhibits all CyP isoforms, and this can be an important confounder in evaluating its mitochondrial effects. Indeed, inhibition of calcineurin, which is mediated by the complex of CsA with cytosolic CyPA [92,93], prevents dephosphorylation of the pro-fission protein Drp-1 at Ser637, thus hampering its mitochondrial translocation and therefore fragmentation, a cytoprotective mitochondrial effect of CsA that is not related to PTP inhibition [148,149]. Of note, genetic ablation or pharmacological inhibition of Drp1 has been shown to afford significant protection against cardiomyocyte injury caused by ceramide, hyperglycemia and post-ischemic reperfusion [150–152].

This effect of CsA may be counteracted by other processes in the context of myocardial ischemia. Thus, lack of Ser637 dephosphorylation due to calcineurin inhibition can be contrasted by the activating effect of Ser616 phosphorylation [153], a process catalyzed by various protein kinases including PKCδ that is generally considered detrimental for the survival of ischemic cardiomyocytes [153]. The protective effect of CsA could be further limited by direct stimulation of HIF-1α Pro564 hydroxylation, which leads to increased removal of HIF-1α and may thus prevent adaptation to hypoxia, an effect originally described in glioma cells [154] that should now be investigated in cardiomyocytes.

8. Summary and perspectives

The discovery that the PTP forms from the F-ATP synthase is redirecting research toward the mechanisms that switch this key enzyme from an energy-conserving to an energy-dissipating device. These hold great promise to improve our understanding of the pathophysiological events that trigger the transition in heart diseases, and to set a logical frame for therapeutic strategies. CyPD remains a viable target, and we are confident that the current hurdles between preclinical studies and clinical application of PTP-inhibitory strategies will be overcome by more specific inhibitors of CyP isoforms [155], by drugs targeting the PTP at sites other than CyPD [156] and by their combinatorial use.

Conflict of interest

The authors declare no conflict of interest.

Acknowledgments

Research in our laboratories is supported by the AIRC (IG13392), Telethon (GGP11082 and GPP10005A), the Ministero dell'Istruzione, dell'Università e della Ricerca (FIRB RBAP11S8C3 and PRIN 20107Z8XBW), NIH-PHS (R01GM069883 and R03DA033978-01), the University of Padova and Fondazione Cassa di Risparmio di Padova e Rovigo.

Contributor Information

Paolo Bernardi, Email: bernardi@bio.unipd.it.

Fabio Di Lisa, Email: dilisa@bio.unipd.it.

References

  • 1.Raaflaub J. Die schwellung isolierter leberzell mitochondrien und ihre physikalisch beeinfluβarkeit. Helv Physiol Pharmacol Acta. 1953;11:142–156. [PubMed] [Google Scholar]
  • 2.Raaflaub J. Über den wirkungsmechanismus von adenosintriphosphat (ATP) als cofaktor isolierter mitochondrien. Helv Physiol Pharmacol Acta. 1953;11:157–165. [PubMed] [Google Scholar]
  • 3.Hunter F.E., Jr., Ford L. Inactivation of oxidative and phosphorylative systems in mitochondria by preincubation with phosphate and other ions. J Biol Chem. 1955;216:357–369. [PubMed] [Google Scholar]
  • 4.Tapley D.F. The effect of thyroxine and other substances on the swelling of isolated rat liver mitochondria. J Biol Chem. 1956;222:325–339. [PubMed] [Google Scholar]
  • 5.Lehninger A.L., Remmert L.F. An endogenous uncoupling and swelling agent in liver mitochondria and its enzymic function. J Biol Chem. 1959;234:2459–2464. [PubMed] [Google Scholar]
  • 6.Lehninger A.L. Reversal of various types of mitochondrial swelling by adenosine triphosphate. J Biol Chem. 1959;234:2465–2471. [PubMed] [Google Scholar]
  • 7.Wojtczak L., Lehninger A.L. Formation and disappearance of an endogenous uncoupling factor during swelling and contraction of mitochondria. Biochim Biophys Acta. 1961;51:442–456. doi: 10.1016/0006-3002(61)90600-x. [DOI] [PubMed] [Google Scholar]
  • 8.Zborowski J., Wojtczak L. Induction of swelling of liver mitochondria by fatty acids of various chain length. Biochim Biophys Acta. 1963;70:596–598. doi: 10.1016/0006-3002(63)90799-6. [DOI] [PubMed] [Google Scholar]
  • 9.Azzi A., Azzone G.F. Swelling and shrinkage phenomena in liver mitochondria. I. Large amplitude swelling induced by inorganic phosphate and by ATP. Biochim Biophys Acta. 1965;105:253–264. doi: 10.1016/s0926-6593(65)80150-3. [DOI] [PubMed] [Google Scholar]
  • 10.Azzone G.F., Azzi A. Volume changes in liver mitochondria. Proc Natl Acad Sci U S A. 1965;53:1084–1089. doi: 10.1073/pnas.53.5.1084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Pfeiffer D.R., Tchen T.T. The role of Ca2 + in control of malic enzyme activity in bovine adrenal cortex mitochondria. Biochem Biophys Res Commun. 1973;50:807–813. doi: 10.1016/0006-291x(73)91316-8. [DOI] [PubMed] [Google Scholar]
  • 12.Pfeiffer D.R., Tchen T.T. The activation of adrenal cortex mitochondrial malic enzyme by Ca2 + and Mg2 + Biochemistry. 1975;14:89–96. doi: 10.1021/bi00672a015. [DOI] [PubMed] [Google Scholar]
  • 13.Pfeiffer D.R., Kuo T.H., Tchen T.T. Some effects of Ca2 +, Mg2 +, and Mn2 + on the ultrastructure, light-scattering properties, and malic enzyme activity of adrenal cortex mitochondria. Arch Biochem Biophys. 1976;176:556–563. doi: 10.1016/0003-9861(76)90199-5. [DOI] [PubMed] [Google Scholar]
  • 14.Hunter D.R., Haworth R.A., Southard J.H. Relationship between configuration, function, and permeability in calcium-treated mitochondria. J Biol Chem. 1976;251:5069–5077. [PubMed] [Google Scholar]
  • 15.Hunter D.R., Haworth R.A. The Ca2 +-induced membrane transition in mitochondria. I. The protective mechanisms. Arch Biochem Biophys. 1979;195:453–459. doi: 10.1016/0003-9861(79)90371-0. [DOI] [PubMed] [Google Scholar]
  • 16.Haworth R.A., Hunter D.R. The Ca2 +-induced membrane transition of rat liver mitochondria. II. Nature of the Ca2 + trigger site. Arch Biochem Biophys. 1979;195:460–467. doi: 10.1016/0003-9861(79)90372-2. [DOI] [PubMed] [Google Scholar]
  • 17.Hunter D.R., Haworth R.A. The Ca2 +-induced membrane transition in mitochondria. III. Transitional Ca2 + release. Arch Biochem Biophys. 1979;195:468–477. doi: 10.1016/0003-9861(79)90373-4. [DOI] [PubMed] [Google Scholar]
  • 18.Kinnally K.W., Campo M.L., Tedeschi H. Mitochondrial channel activity studied by patch-clamping mitoplasts. J Bioenerg Biomembr. 1989;21:497–506. doi: 10.1007/BF00762521. [DOI] [PubMed] [Google Scholar]
  • 19.Petronilli V., Szabó I., Zoratti M. The inner mitochondrial membrane contains ion-conducting channels similar to those found in bacteria. FEBS Lett. 1989;259:137–143. doi: 10.1016/0014-5793(89)81513-3. [DOI] [PubMed] [Google Scholar]
  • 20.Bernardi P., Vassanelli S., Veronese P., Colonna R., Szabo I., Zoratti M. Modulation of the mitochondrial permeability transition pore. Effect of protons and divalent cations. J Biol Chem. 1992;267:2934–2939. [PubMed] [Google Scholar]
  • 21.Szabó I., Bernardi P., Zoratti M. Modulation of the mitochondrial megachannel by divalent cations and protons. J Biol Chem. 1992;267:2940–2946. [PubMed] [Google Scholar]
  • 22.Szabó I., Zoratti M. The giant channel of the inner mitochondrial membrane is inhibited by cyclosporin A. J Biol Chem. 1991;266:3376–3379. [PubMed] [Google Scholar]
  • 23.Szabó I., Zoratti M. The mitochondrial megachannel is the permeability transition pore. J Bioenerg Biomembr. 1992;24:111–117. doi: 10.1007/BF00769537. [DOI] [PubMed] [Google Scholar]
  • 24.Szabò I., Zoratti M. Mitochondrial channels: ion fluxes and more. Physiol Rev. 2014;94:519–608. doi: 10.1152/physrev.00021.2013. [DOI] [PubMed] [Google Scholar]
  • 25.Bernardi P., Krauskopf A., Basso E., Petronilli V., Blachly-Dyson E., Di Lisa F. The mitochondrial permeability transition from in vitro artifact to disease target. FEBS J. 2006;273:2077–2099. doi: 10.1111/j.1742-4658.2006.05213.x. [DOI] [PubMed] [Google Scholar]
  • 26.Rasola A., Bernardi P. Mitochondrial permeability transition in Ca2 +-dependent apoptosis and necrosis. Cell Calcium. 2011;50:222–233. doi: 10.1016/j.ceca.2011.04.007. [DOI] [PubMed] [Google Scholar]
  • 27.Petronilli V., Penzo D., Scorrano L., Bernardi P., Di Lisa F. The mitochondrial permeability transition, release of cytochrome c and cell death. Correlation with the duration of pore openings in situ. J Biol Chem. 2001;276:12030–12034. doi: 10.1074/jbc.M010604200. [DOI] [PubMed] [Google Scholar]
  • 28.Hüser J., Rechenmacher C.E., Blatter L.A. Imaging the permeability pore transition in single mitochondria. Biophys J. 1998;74:2129–2137. doi: 10.1016/S0006-3495(98)77920-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Hüser J., Blatter L.A. Fluctuations in mitochondrial membrane potential caused by repetitive gating of the permeability transition pore. Biochem J. 1999;343(Pt 2):311–317. [PMC free article] [PubMed] [Google Scholar]
  • 30.Petronilli V., Miotto G., Canton M., Brini M., Colonna R., Bernardi P. 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;76:725–734. doi: 10.1016/S0006-3495(99)77239-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Zorov D.B., Juhaszova M., Sollott S.J. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol Rev. 2014;94:909–950. doi: 10.1152/physrev.00026.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Bernardi P., Petronilli V. The permeability transition pore as a mitochondrial calcium release channel: a critical appraisal. J Bioenerg Biomembr. 1996;28:131–138. doi: 10.1007/BF02110643. [DOI] [PubMed] [Google Scholar]
  • 33.Ichas F., Jouaville L.S., Mazat J.P. Mitochondria are excitable organelles capable of generating and conveying electrical and calcium signals. Cell. 1997;89:1145–1153. doi: 10.1016/s0092-8674(00)80301-3. [DOI] [PubMed] [Google Scholar]
  • 34.Barsukova A., Komarov A., Hajnoczky G., Bernardi P., Bourdette D., Forte M. Activation of the mitochondrial permeability transition pore modulates Ca2 + responses to physiological stimuli in adult neurons. Eur J Neurosci. 2011;33:831–842. doi: 10.1111/j.1460-9568.2010.07576.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Bernardi P., von Stockum S. The permeability transition pore as a Ca2 + release channel: new answers to an old question. Cell Calcium. 2012;52:22–27. doi: 10.1016/j.ceca.2012.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Duchen M.R., McGuinness O., Brown L.A., Crompton M. On the involvement of a cyclosporin A sensitive mitochondrial pore in myocardial reperfusion injury. Cardiovasc Res. 1993;27:1790–1794. doi: 10.1093/cvr/27.10.1790. [DOI] [PubMed] [Google Scholar]
  • 37.Griffiths E.J., Halestrap A.P. Protection by cyclosporin A of ischemia/reperfusion-induced damage in isolated rat hearts. J Mol Cell Cardiol. 1993;25:1461–1469. doi: 10.1006/jmcc.1993.1162. [DOI] [PubMed] [Google Scholar]
  • 38.Liu X., Kim C.N., Yang J., Jemmerson R., Wang X. Induction of apoptotic program in cell-free extracts: requirement for dATP and cytochrome c. Cell. 1996;86:147–157. doi: 10.1016/s0092-8674(00)80085-9. [DOI] [PubMed] [Google Scholar]
  • 39.Susin S.A., Zamzami N., Castedo M., Hirsch T., Marchetti P., Macho A. Bcl-2 inhibits the mitochondrial release of an apoptogenic protease. J Exp Med. 1996;184:1331–1341. doi: 10.1084/jem.184.4.1331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Zamzami N., Marchetti P., Castedo M., Hirsch T., Susin S.A., Masse B. Inhibitors of permeability transition interfere with the disruption of the mitochondrial transmembrane potential during apoptosis. FEBS Lett. 1996;384:53–57. doi: 10.1016/0014-5793(96)00280-3. [DOI] [PubMed] [Google Scholar]
  • 41.Javadov S.A., Clarke S., Das M., Griffiths E.J., Lim K.H., Halestrap A.P. Ischaemic preconditioning inhibits opening of mitochondrial permeability transition pores in the reperfused rat heart. J Physiol. 2003;549:513–524. doi: 10.1113/jphysiol.2003.034231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Argaud L., Gateau-Roesch O., Raisky O., Loufouat J., Robert D., Ovize M. Postconditioning inhibits mitochondrial permeability transition. Circulation. 2005;111:194–197. doi: 10.1161/01.CIR.0000151290.04952.3B. [DOI] [PubMed] [Google Scholar]
  • 43.Hausenloy D.J., Yellon D.M., Mani-Babu S., Duchen M.R. Preconditioning protects by inhibiting the mitochondrial permeability transition. Am J Physiol Heart Circ Physiol. 2004;287:H841–H849. doi: 10.1152/ajpheart.00678.2003. [DOI] [PubMed] [Google Scholar]
  • 44.Staat P., Rioufol G., Piot C., Cottin Y., Cung T.T., L'Huillier I. Postconditioning the human heart. Circulation. 2005;112:2143–2148. doi: 10.1161/CIRCULATIONAHA.105.558122. [DOI] [PubMed] [Google Scholar]
  • 45.Halestrap A.P., Clarke S.J., Khaliulin I. The role of mitochondria in protection of the heart by preconditioning. Biochim Biophys Acta. 2007;1767:1007–1031. doi: 10.1016/j.bbabio.2007.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Hausenloy D.J., Yellon D.M. Preconditioning and postconditioning: united at reperfusion. Pharmacol Ther. 2007;116:173–191. doi: 10.1016/j.pharmthera.2007.06.005. [DOI] [PubMed] [Google Scholar]
  • 47.Murphy E., Steenbergen C. Mechanisms underlying acute protection from cardiac ischemia–reperfusion injury. Physiol Rev. 2008;88:581–609. doi: 10.1152/physrev.00024.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Zorov D.B., Juhaszova M., Yaniv Y., Nuss H.B., Wang S., Sollott S.J. Regulation and pharmacology of the mitochondrial permeability transition pore. Cardiovasc Res. 2009;83:213–225. doi: 10.1093/cvr/cvp151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Ovize M., Baxter G.F., Di Lisa F., Ferdinandy P., Garcia-Dorado D., Hausenloy D.J. Postconditioning and protection from reperfusion injury: where do we stand? Position paper from the Working Group of Cellular Biology of the Heart of the European Society of Cardiology. Cardiovasc Res. 2010;87:406–423. doi: 10.1093/cvr/cvq129. [DOI] [PubMed] [Google Scholar]
  • 50.Pan X., Liu J., Nguyen T., Liu C., Sun J., Teng Y. The physiological role of mitochondrial calcium revealed by mice lacking the mitochondrial calcium uniporter. Nat Cell Biol. 2013;15:1464–1472. doi: 10.1038/ncb2868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Fan J., Lindemann P., Feuilloley M.G., Papadopoulos V. Structural and functional evolution of the translocator protein (18 kDa) Curr Mol Med. 2012;12:369–386. doi: 10.2174/1566524011207040369. [DOI] [PubMed] [Google Scholar]
  • 52.McEnery M.W., Snowman A.M., Trifiletti R.R., Snyder S.H. 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;89:3170–3174. doi: 10.1073/pnas.89.8.3170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Kinnally K.W., Zorov D.B., Antonenko Y.N., Snyder S.H., McEnery M.W., Tedeschi H. Mitochondrial benzodiazepine receptor linked to inner membrane ion channels by nanomolar actions of ligands. Proc Natl Acad Sci U S A. 1993;90:1374–1378. doi: 10.1073/pnas.90.4.1374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kottke M., Adam V., Riesinger I., Bremm G., Bosch W., Brdiczka D. Mitochondrial boundary membrane contact sites in brain: points of hexokinase and creatine kinase location, and control of Ca2 + transport. Biochim Biophys Acta. 1988;935:87–102. doi: 10.1016/0005-2728(88)90111-9. [DOI] [PubMed] [Google Scholar]
  • 55.Adams V., Bosch W., Schlegel J., Wallimann T., Brdiczka D. Further characterization of contact sites from mitochondria of different tissues: topology of peripheral kinases. Biochim Biophys Acta. 1989;981:213–225. doi: 10.1016/0005-2736(89)90031-x. [DOI] [PubMed] [Google Scholar]
  • 56.Bucheler K., Adams V., 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;1056:233–242. doi: 10.1016/s0005-2728(05)80054-4. [DOI] [PubMed] [Google Scholar]
  • 57.Beutner G., Rück A., Riede B., Welte W., Brdiczka D. Complexes between kinases, mitochondrial porin and adenylate translocator in rat brain resemble the permeability transition pore. FEBS Lett. 1996;396:189–195. doi: 10.1016/0014-5793(96)01092-7. [DOI] [PubMed] [Google Scholar]
  • 58.Marzo I., Brenner C., Zamzami N., Susin S.A., Beutner G., Brdiczka D. The permeability transition pore complex: a target for apoptosis regulation by caspases and bcl-2-related proteins. J Exp Med. 1998;187:1261–1271. doi: 10.1084/jem.187.8.1261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Zamzami N., Kroemer G. The mitochondrion in apoptosis: how Pandora's box opens. Nat Rev Mol Cell Biol. 2001;2:67–71. doi: 10.1038/35048073. [DOI] [PubMed] [Google Scholar]
  • 60.Kokoszka J.E., Waymire K.G., Levy S.E., Sligh J.E., Cai J., Jones D.P. The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore. Nature. 2004;427:461–465. doi: 10.1038/nature02229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Krauskopf A., Eriksson O., Craigen W.J., Forte M.A., Bernardi P. Properties of the permeability transition in VDAC1−/− mitochondria. Biochim Biophys Acta. 2006;1757:590–595. doi: 10.1016/j.bbabio.2006.02.007. [DOI] [PubMed] [Google Scholar]
  • 62.Baines C.P., Kaiser R.A., Sheiko T., Craigen W.J., Molkentin J.D. Voltage-dependent anion channels are dispensable for mitochondrial-dependent cell death. Nat Cell Biol. 2007;9:550–555. doi: 10.1038/ncb1575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Šileikyte J., Blachly-Dyson E., Sewell R., Carpi A., Menabò R., Di Lisa F. 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;289:13769–13781. doi: 10.1074/jbc.M114.549634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Leung A.W., Halestrap A.P. Recent progress in elucidating the molecular mechanism of the mitochondrial permeability transition pore. Biochim Biophys Acta. 2008;1777:946–952. doi: 10.1016/j.bbabio.2008.03.009. [DOI] [PubMed] [Google Scholar]
  • 65.Herick K., Krämer R., Lühring H. Patch clamp investigation into the phosphate carrier from Saccharomyces cerevisiae mitochondria. Biochim Biophys Acta. 1997;1321:207–220. doi: 10.1016/s0005-2728(97)00050-9. [DOI] [PubMed] [Google Scholar]
  • 66.Kwong J.Q., Davis J., Baines C.P., Sargent M.A., Karch J., Wang X. Genetic deletion of the mitochondrial phosphate carrier desensitizes the mitochondrial permeability transition pore and causes cardiomyopathy. Cell Death Differ. 2014;21:1209–1217. doi: 10.1038/cdd.2014.36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Baines C.P., Kaiser R.A., Purcell N.H., Blair N.S., Osinska H., Hambleton M.A. Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature. 2005;434:658–662. doi: 10.1038/nature03434. [DOI] [PubMed] [Google Scholar]
  • 68.Basso E., Fante L., Fowlkes J., Petronilli V., Forte M.A., Bernardi P. Properties of the permeability transition pore in mitochondria devoid of Cyclophilin D. J Biol Chem. 2005;280:18558–18561. doi: 10.1074/jbc.C500089200. [DOI] [PubMed] [Google Scholar]
  • 69.Nakagawa T., Shimizu S., Watanabe T., Yamaguchi O., Otsu K., Yamagata H. Cyclophilin D-dependent mitochondrial permeability transition regulates some necrotic but not apoptotic cell death. Nature. 2005;434:652–658. doi: 10.1038/nature03317. [DOI] [PubMed] [Google Scholar]
  • 70.Schinzel A.C., Takeuchi O., Huang Z., Fisher J.K., Zhou Z., Rubens J. 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;102:12005–12010. doi: 10.1073/pnas.0505294102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Giorgio V., Bisetto E., Soriano M.E., Dabbeni-Sala F., Basso E., Petronilli V. Cyclophilin D modulates mitochondrial F0F1-ATP synthase by interacting with the lateral stalk of the complex. J Biol Chem. 2009;284:33982–33988. doi: 10.1074/jbc.M109.020115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Giorgio V., von Stockum S., Antoniel M., Fabbro A., Fogolari F., Forte M. Dimers of mitochondrial ATP synthase form the permeability transition pore. Proc Natl Acad Sci U S A. 2013;110:5887–5892. doi: 10.1073/pnas.1217823110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Carraro M., Giorgio V., Šileikyte J., Sartori G., Forte M., Lippe G. Channel formation by yeast F-ATP synthase and the role of dimerization in the mitochondrial permeability transition. J Biol Chem. 2014;289:15980–15985. doi: 10.1074/jbc.C114.559633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Giraud M.F., Paumard P., Soubannier V., Vaillier J., Arselin G., Salin B. Is there a relationship between the supramolecular organization of the mitochondrial ATP synthase and the formation of cristae? Biochim Biophys Acta. 2002;1555:174–180. doi: 10.1016/s0005-2728(02)00274-8. [DOI] [PubMed] [Google Scholar]
  • 75.Davies K.M., Anselmi C., Wittig I., Faraldo-Gomez J.D., Kühlbrandt W. Structure of the yeast F1FO-ATP synthase dimer and its role in shaping the mitochondrial cristae. Proc Natl Acad Sci U S A. 2012;109:13602–13607. doi: 10.1073/pnas.1204593109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Bernardi P. The mitochondrial permeability transition pore: a mystery solved? Front Physiol. 2013;4:95. doi: 10.3389/fphys.2013.00095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Bonora M., Bononi A., De Marchi E., Giorgi C., Lebiedzinska M., Marchi S. Role of the c subunit of the FO ATP synthase in mitochondrial permeability transition. Cell Cycle. 2013;12:674–683. doi: 10.4161/cc.23599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Alavian K.N., Beutner G., Lazrove E., Sacchetti S., Park H.A., Licznerski P. 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;111:10580–10585. doi: 10.1073/pnas.1401591111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Kagawa Y., Racker E. Partial resolution of the enzymes catalyzing oxidative phosphorylation. 8. Properties of a factor conferring oligomycin sensitivity on mitochondrial adenosine triphosphatase. J Biol Chem. 1966;241:2461–2466. [PubMed] [Google Scholar]
  • 80.Abrahams J.P., Leslie A.G., Lutter R., Walker J.E. Structure at 2.8 A resolution of F1-ATPase from bovine heart mitochondria. Nature. 1994;370:621–628. doi: 10.1038/370621a0. [DOI] [PubMed] [Google Scholar]
  • 81.Petronilli V., Nicolli A., Costantini P., Colonna R., Bernardi P. Regulation of the permeability transition pore, a voltage-dependent mitochondrial channel inhibited by cyclosporin A. Biochim Biophys Acta. 1994;1187:255–259. doi: 10.1016/0005-2728(94)90122-8. [DOI] [PubMed] [Google Scholar]
  • 82.Nguyen T.T., Stevens M.V., Kohr M., Steenbergen C., Sack M.N., Murphy E. Cysteine 203 of cyclophilin D is critical for cyclophilin D activation of the mitochondrial permeability transition pore. J Biol Chem. 2011;286:40184–40192. doi: 10.1074/jbc.M111.243469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Havlickova V., Kaplanova V., Nuskova H., Drahota Z., Houstek J. Knockdown of F1 epsilon subunit decreases mitochondrial content of ATP synthase and leads to accumulation of subunit c. Biochim Biophys Acta. 2010;1797:1124–1129. doi: 10.1016/j.bbabio.2009.12.009. [DOI] [PubMed] [Google Scholar]
  • 84.McGeoch J.E., Guidotti G. A 0.1–700 Hz current through a voltage-clamped pore: candidate protein for initiator of neural oscillations. Brain Res. 1997;766:188–194. doi: 10.1016/s0006-8993(97)00618-5. [DOI] [PubMed] [Google Scholar]
  • 85.McGeoch J.E., McGeoch M.W., Mao R., Guidotti G. Opposing actions of cGMP and calcium on the conductance of the F0 subunit c pore. Biochem Biophys Res Commun. 2000;274:835–840. doi: 10.1006/bbrc.2000.3231. [DOI] [PubMed] [Google Scholar]
  • 86.McGeoch J.E., Guidotti G. Batten disease and the control of the Fo subunit c pore by cGMP and calcium. Eur J Paediatr Neurol. 2001;5(Suppl. A):147–150. doi: 10.1053/ejpn.2000.0452. [DOI] [PubMed] [Google Scholar]
  • 87.Takahashi N., Hayano T., Suzuki M. Peptidyl-prolyl cis-trans isomerase is the cyclosporin A-binding protein cyclophilin. Nature. 1989;337:473–475. doi: 10.1038/337473a0. [DOI] [PubMed] [Google Scholar]
  • 88.Fischer G., Wittmann-Liebold B., Lang K., Kiefhaber T., Schmid F.X. Cyclophilin and peptidyl-prolyl cis-trans isomerase are probably identical proteins. Nature. 1989;337:476–478. doi: 10.1038/337476a0. [DOI] [PubMed] [Google Scholar]
  • 89.Pemberton T.J., Kay J.E. Identification and comparative analysis of the peptidyl-prolyl cis/trans isomerase repertoires of H. sapiens, D. melanogaster, C. elegans, S. cerevisiae and Sz. pombe. Comp Funct Genomics. 2005;6:277–300. doi: 10.1002/cfg.482. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Wang P., Heitman J. The cyclophilins. Genome Biol. 2005;6:226.1–226.6. doi: 10.1186/gb-2005-6-7-226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Borel J.F., Feurer C., Magnee C., Stahelin H. Effects of the new anti-lymphocytic peptide cyclosporin A in animals. Immunology. 1977;32:1017–1025. [PMC free article] [PubMed] [Google Scholar]
  • 92.Liu J., Farmer J.D.J., Lane W.S., Friedman J., Weissman I., Schreiber S.L. Calcineurin is a common target of cyclophilin–cyclosporin A and FKBP–FK506 complexes. Cell. 1991;66:807–815. doi: 10.1016/0092-8674(91)90124-h. [DOI] [PubMed] [Google Scholar]
  • 93.Clipstone N.A., Crabtree G.R. Identification of calcineurin as a key signalling enzyme in T-lymphocyte activation. Nature. 1992;357:695–697. doi: 10.1038/357695a0. [DOI] [PubMed] [Google Scholar]
  • 94.Walsh C.T., Zydowsky L.D., McKeon F.D. Cyclosporin A, the cyclophilin class of peptidylprolyl isomerases, and blockade of T cell signal transduction. J Biol Chem. 1992;267:13115–13118. [PubMed] [Google Scholar]
  • 95.Giorgio V., Soriano M.E., Basso E., Bisetto E., Lippe G., Forte M.A. Cyclophilin D in mitochondrial pathophysiology. Biochim Biophys Acta. 2010;1797:1113–1118. doi: 10.1016/j.bbabio.2009.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Azzolin L., von Stockum S., Basso E., Petronilli V., Forte M.A., Bernardi P. The mitochondrial permeability transition from yeast to mammals. FEBS Lett. 2010;584:2504–2509. doi: 10.1016/j.febslet.2010.04.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Jung D.W., Bradshaw P.C., Pfeiffer D.R. Properties of a cyclosporin-insensitive permeability transition pore in yeast mitochondria. J Biol Chem. 1997;272:21104–21112. doi: 10.1074/jbc.272.34.21104. [DOI] [PubMed] [Google Scholar]
  • 98.von Stockum S., Basso E., Petronilli V., Sabatelli P., Forte M.A., Bernardi P. Properties of Ca2 + transport in mitochondria of Drosophila melanogaster. J Biol Chem. 2011;286:41163–41170. doi: 10.1074/jbc.M111.268375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Li B., Chauvin C., De Paulis D., De Oliveira F., Gharib A., Vial G. Inhibition of complex I regulates the mitochondrial permeability transition through a phosphate-sensitive inhibitory site masked by cyclophilin D. Biochim Biophys Acta. 1817;2012:1628–1634. doi: 10.1016/j.bbabio.2012.05.011. [DOI] [PubMed] [Google Scholar]
  • 100.Johnson K.M., Chen X., Boitano A., Swenson L., Opipari A.W., Jr., Glick G.D. Identification and validation of the mitochondrial F1F0-ATPase as the molecular target of the immunomodulatory benzodiazepine Bz-423. Chem Biol. 2005;12:485–496. doi: 10.1016/j.chembiol.2005.02.012. [DOI] [PubMed] [Google Scholar]
  • 101.Stelzer A.C., Frazee R.W., Van Huis C., Cleary J., Opipari A.W., Jr., Glick G.D. NMR studies of an immunomodulatory benzodiazepine binding to its molecular target on the mitochondrial F1F0-ATPase. Biopolymers. 2010;93:85–92. doi: 10.1002/bip.21306. [DOI] [PubMed] [Google Scholar]
  • 102.Bernardi P., Penzo D., Wojtczak L. Mitochondrial energy dissipation by fatty acids. Mechanisms and implications for cell death. Vitam Horm. 2002;65:97–126. doi: 10.1016/s0083-6729(02)65061-7. [DOI] [PubMed] [Google Scholar]
  • 103.Pfeiffer K., Gohil V., Stuart R.A., Hunte C., Brandt U., Greenberg M.L. Cardiolipin stabilizes respiratory chain supercomplexes. J Biol Chem. 2003;278:52873–52880. doi: 10.1074/jbc.M308366200. [DOI] [PubMed] [Google Scholar]
  • 104.Acehan D., Malhotra A., Xu Y., Ren M., Stokes D.L., Schlame M. Cardiolipin affects the supramolecular organization of ATP synthase in mitochondria. Biophys J. 2011;100:2184–2192. doi: 10.1016/j.bpj.2011.03.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Šileikyte J., Petronilli V., Zulian A., Dabbeni-Sala F., Tognon G., Nikolov P. Regulation of the inner membrane mitochondrial permeability transition by the outer membrane translocator protein (peripheral benzodiazepine receptor) J Biol Chem. 2011;286:1046–1053. doi: 10.1074/jbc.M110.172486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Lê-Quôc K., Lê-Quôc D. Crucial role of sulfhydryl groups in the mitochondrial inner membrane structure. J Biol Chem. 1985;260:7422–7428. [PubMed] [Google Scholar]
  • 107.Ricchelli F., Šileikyte J., Bernardi P. Shedding light on the mitochondrial permeability transition. Biochim Biophys Acta. 1807;2011:482–490. doi: 10.1016/j.bbabio.2011.02.012. [DOI] [PubMed] [Google Scholar]
  • 108.Salet C., Moreno G., Ricchelli F., Bernardi P. Singlet oxygen produced by photodynamic action causes inactivation of the mitochondrial permeability transition pore. J Biol Chem. 1997;272:21938–21943. doi: 10.1074/jbc.272.35.21938. [DOI] [PubMed] [Google Scholar]
  • 109.Petronilli V., Šileikyte J., Zulian A., Dabbeni-Sala F., Jori G., Gobbo S. Switch from inhibition to activation of the mitochondrial permeability transition during hematoporphyrin-mediated photooxidative stress. Unmasking pore-regulating external thiols. Biochim Biophys Acta. 2009;1787:897–904. doi: 10.1016/j.bbabio.2009.03.014. [DOI] [PubMed] [Google Scholar]
  • 110.Pastorino J.G., Simbula G., Gilfor E., Hoek J.B., Farber J.L. Protoporphyrin IX, an endogenous ligand of the peripheral benzodiazepine receptor, potentiates induction of the mitochondrial permeability transition and the killing of cultured hepatocytes by rotenone. J Biol Chem. 1994;269:31041–31046. [PubMed] [Google Scholar]
  • 111.Chelli B., Falleni A., Salvetti F., Gremigni V., Lucacchini A., Martini C. Peripheral-type benzodiazepine receptor ligands: mitochondrial permeability transition induction in rat cardiac tissue. Biochem Pharmacol. 2001;61:695–705. doi: 10.1016/s0006-2952(00)00588-8. [DOI] [PubMed] [Google Scholar]
  • 112.Krestinina O.V., Grachev D.E., Odinokova I.V., Reiser G., Evtodienko Y.V., Azarashvili T.S. Effect of peripheral benzodiazepine receptor (PBR/TSPO) ligands on opening of Ca2 +-induced pore and phosphorylation of 3.5-kDa polypeptide in rat brain mitochondria. Biochemistry (Mosc) 2009;74:421–429. doi: 10.1134/s0006297909040105. [DOI] [PubMed] [Google Scholar]
  • 113.Gonzalez-Polo R.A., Carvalho G., Braun T., Decaudin D., Fabre C., Larochette N. PK11195 potently sensitizes to apoptosis induction independently from the peripheral benzodiazepin receptor. Oncogene. 2005;24:7503–7513. doi: 10.1038/sj.onc.1208907. [DOI] [PubMed] [Google Scholar]
  • 114.Hans G., Wislet-Gendebien S., Lallemend F., Robe P., Rogister B., Belachew S. Peripheral benzodiazepine receptor (PBR) ligand cytotoxicity unrelated to PBR expression. Biochem Pharmacol. 2005;69:819–830. doi: 10.1016/j.bcp.2004.11.029. [DOI] [PubMed] [Google Scholar]
  • 115.Cleary J., Johnson K.M., Opipari A.W., Jr., Glick G.D. Inhibition of the mitochondrial F1F0-ATPase by ligands of the peripheral benzodiazepine receptor. Bioorg Med Chem Lett. 2007;17:1667–1670. doi: 10.1016/j.bmcl.2006.12.102. [DOI] [PubMed] [Google Scholar]
  • 116.Sundberg T.B., Swenson L., Wahl D.R., Opipari A.W., Jr., Glick G.D. Apoptotic signaling activated by modulation of the F0F1-ATPase: implications for selective killing of autoimmune lymphocytes. J Pharmacol Exp Ther. 2009;331:437–444. doi: 10.1124/jpet.109.156422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Krishnamurthy P., Xie T., Schuetz J.D. The role of transporters in cellular heme and porphyrin homeostasis. Pharmacol Ther. 2007;114:345–358. doi: 10.1016/j.pharmthera.2007.02.001. [DOI] [PubMed] [Google Scholar]
  • 118.Strauss M., Hofhaus G., Schröder R.R., Kühlbrandt W. Dimer ribbons of ATP synthase shape the inner mitochondrial membrane. EMBO J. 2008;27:1154–1160. doi: 10.1038/emboj.2008.35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Davies K.M., Strauss M., Daum B., Kief J.H., Osiewacz H.D., Rycovska A. Macromolecular organization of ATP synthase and complex I in whole mitochondria. Proc Natl Acad Sci U S A. 2011;108:14121–14126. doi: 10.1073/pnas.1103621108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Madesh M., Hajnoczky G. VDAC-dependent permeabilization of the outer mitochondrial membrane by superoxide induces rapid and massive cytochrome c release. J Cell Biol. 2001;155:1003–1015. doi: 10.1083/jcb.200105057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Rapizzi E., Pinton P., Szabadkai G., Wieckowski M.R., Vandecasteele G., Baird G. Recombinant expression of the voltage-dependent anion channel enhances the transfer of Ca2 + microdomains to mitochondria. J Cell Biol. 2002;159:613–624. doi: 10.1083/jcb.200205091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Scorrano L., Ashiya M., Buttle K., Weiler S., Oakes S.A., Mannella C.A. A distinct pathway remodels mitochondrial cristae and mobilizes cytochrome c during apoptosis. Dev Cell. 2002;2:55–67. doi: 10.1016/s1534-5807(01)00116-2. [DOI] [PubMed] [Google Scholar]
  • 123.Karch J., Kwong J.Q., Burr A.R., Sargent M.A., Elrod J.W., Peixoto P.M. Bax and Bak function as the outer membrane component of the mitochondrial permeability pore in regulating necrotic cell death in mice. Elife. 2013;2:e00772. doi: 10.7554/eLife.00772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Piot C., Croisille P., Staat P., Thibault H., Rioufol G., Mewton N. Effect of cyclosporine on reperfusion injury in acute myocardial infarction. New Engl J Med. 2008;359:473–481. doi: 10.1056/NEJMoa071142. [DOI] [PubMed] [Google Scholar]
  • 125.Dow J., Kloner R.A. Postconditioning does not reduce myocardial infarct size in an in vivo regional ischemia rodent model. J Cardiovasc Pharmacol Ther. 2007;12:153–163. doi: 10.1177/1074248407300897. [DOI] [PubMed] [Google Scholar]
  • 126.Karlsson L.O., Zhou A.X., Larsson E., Astrom-Olsson K., Mansson C., Akyurek L.M. Cyclosporine does not reduce myocardial infarct size in a porcine ischemia–reperfusion model. J Cardiovasc Pharmacol Ther. 2010;15:182–189. doi: 10.1177/1074248410362074. [DOI] [PubMed] [Google Scholar]
  • 127.Skyschally A., Schulz R., Heusch G. Cyclosporine A at reperfusion reduces infarct size in pigs. Cardiovasc Drugs Ther. 2010;24:85–87. doi: 10.1007/s10557-010-6219-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Ruiz-Meana M., Inserte J., Fernandez-Sanz C., Hernando V., Miro-Casas E., Barba I. The role of mitochondrial permeability transition in reperfusion-induced cardiomyocyte death depends on the duration of ischemia. Basic Res Cardiol. 2011;106:1259–1268. doi: 10.1007/s00395-011-0225-5. [DOI] [PubMed] [Google Scholar]
  • 129.Elrod J.W., Wong R., Mishra S., Vagnozzi R.J., Sakthievel B., Goonasekera S.A. Cyclophilin D controls mitochondrial pore-dependent Ca2 + exchange, metabolic flexibility, and propensity for heart failure in mice. J Clin Invest. 2010;120:3680–3687. doi: 10.1172/JCI43171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Woodfield K.Y., Price N.T., Halestrap A.P. cDNA cloning of rat mitochondrial cyclophilin. Biochim Biophys Acta. 1997;1351:27–30. doi: 10.1016/s0167-4781(97)00017-1. [DOI] [PubMed] [Google Scholar]
  • 131.Rasola A., Sciacovelli M., Chiara F., Pantic B., Brusilow W.S., Bernardi P. Activation of mitochondrial ERK protects cancer cells from death through inhibition of the permeability transition. Proc Natl Acad Sci U S A. 2010;107:726–731. doi: 10.1073/pnas.0912742107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Masgras I., Rasola A., Bernardi P. Induction of the permeability transition pore in cells depleted of mitochondrial DNA. Biochim Biophys Acta. 1817;2012:1860–1866. doi: 10.1016/j.bbabio.2012.02.022. [DOI] [PubMed] [Google Scholar]
  • 133.Kang B.H., Plescia J., Dohi T., Rosa J., Doxsey S.J., Altieri D.C. Regulation of tumor cell mitochondrial homeostasis by an organelle-specific Hsp90 chaperone network. Cell. 2007;131:257–270. doi: 10.1016/j.cell.2007.08.028. [DOI] [PubMed] [Google Scholar]
  • 134.Vaseva A.V., Marchenko N.D., Ji K., Tsirka S.E., Holzmann S., Moll U.M. p53 opens the mitochondrial permeability transition pore to trigger necrosis. Cell. 2012;149:1536–1548. doi: 10.1016/j.cell.2012.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Karch J., Molkentin J.D. Is p53 the long-sought molecular trigger for cyclophilin D-regulated mitochondrial permeability transition pore formation and necrosis? Circ Res. 2012;111:1258–1260. doi: 10.1161/CIRCRESAHA.112.280990. [DOI] [PubMed] [Google Scholar]
  • 136.Belaidi E., Decorps J., Augeul L., Durand A., Ovize M. Endoplasmic reticulum stress contributes to heart protection induced by cyclophilin D inhibition. Basic Res Cardiol. 2013;108:363. doi: 10.1007/s00395-013-0363-z. [DOI] [PubMed] [Google Scholar]
  • 137.Paillard M., Tubbs E., Thiebaut P.A., Gomez L., Fauconnier J., Da Silva C.C. Depressing mitochondria–reticulum interactions protects cardiomyocytes from lethal hypoxia–reoxygenation injury. Circulation. 2013;128:1555–1565. doi: 10.1161/CIRCULATIONAHA.113.001225. [DOI] [PubMed] [Google Scholar]
  • 138.Szabadkai G., Bianchi K., Varnai P., De Stefani D., Wieckowski M.R., Cavagna D. Chaperone-mediated coupling of endoplasmic reticulum and mitochondrial Ca2 + channels. J Cell Biol. 2006;175:901–911. doi: 10.1083/jcb.200608073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Baughman J.M., Perocchi F., Girgis H.S., Plovanich M., Belcher-Timme C.A., Sancak Y. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature. 2011;476:341–345. doi: 10.1038/nature10234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.De Stefani D., Raffaello A., Teardo E., Szabó I., Rizzuto R. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature. 2011;476:336–340. doi: 10.1038/nature10230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Altschuld R.A., Hohl C.M., Castillo L.C., Garleb A.A., Starling R.C., Brierley G.P. Cyclosporin inhibits mitochondrial calcium efflux in isolated adult rat ventricular cardiomyocytes. Am J Physiol. 1992;262:H1699–H1704. doi: 10.1152/ajpheart.1992.262.6.H1699. [DOI] [PubMed] [Google Scholar]
  • 142.Eliseev R.A., Malecki J., Lester T., Zhang Y., Humphrey J., Gunter T.E. Cyclophilin D interacts with Bcl2 and exerts an anti-apoptotic effect. J Biol Chem. 2009;284:9692–9699. doi: 10.1074/jbc.M808750200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Rasola A., Sciacovelli M., Pantic B., Bernardi P. Signal transduction to the permeability transition pore. FEBS Lett. 2010;584:1989–1996. doi: 10.1016/j.febslet.2010.02.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Shulga N., Pastorino J.G. Ethanol sensitizes mitochondria to the permeability transition by inhibiting deacetylation of cyclophilin-D mediated by sirtuin-3. J Cell Sci. 2010;123:4117–4127. doi: 10.1242/jcs.073502. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 145.Kohr M.J., Aponte A.M., Sun J., Wang G., Murphy E., Gucek M. Characterization of potential S-nitrosylation sites in the myocardium. Am J Physiol Heart Circ Physiol. 2011;300:H1327–H1335. doi: 10.1152/ajpheart.00997.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Menazza S., Wong R., Nguyen T., Wang G., Gucek M., Murphy E. CypD−/− hearts have altered levels of proteins involved in Krebs cycle, branch chain amino acid degradation and pyruvate metabolism. J Mol Cell Cardiol. 2013;56:81–90. doi: 10.1016/j.yjmcc.2012.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Nguyen T.T., Wong R., Menazza S., Sun J., Chen Y., Wang G. Cyclophilin D modulates mitochondrial acetylome. Circ Res. 2013;113:1308–1319. doi: 10.1161/CIRCRESAHA.113.301867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Cereghetti G.M., Stangherlin A., Martins de Brito O., Chang C.R., Blackstone C, Bernardi P. Dephosphorylation by calcineurin regulates translocation of Drp1 to mitochondria. Proc Natl Acad Sci U S A. 2008;105:15803–15808. doi: 10.1073/pnas.0808249105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Cereghetti G.M., Costa V., Scorrano L. Inhibition of Drp1-dependent mitochondrial fragmentation and apoptosis by a polypeptide antagonist of calcineurin. Cell Death Differ. 2010;17:1785–1794. doi: 10.1038/cdd.2010.61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Parra V., Eisner V., Chiong M., Criollo A., Moraga F., Garcia A. Changes in mitochondrial dynamics during ceramide-induced cardiomyocyte early apoptosis. Cardiovasc Res. 2008;77:387–397. doi: 10.1093/cvr/cvm029. [DOI] [PubMed] [Google Scholar]
  • 151.Yu T., Sheu S.S., Robotham J.L., Yoon Y. Mitochondrial fission mediates high glucose-induced cell death through elevated production of reactive oxygen species. Cardiovasc Res. 2008;79:341–351. doi: 10.1093/cvr/cvn104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Ong S.B., Subrayan S., Lim S.Y., Yellon D.M., Davidson S.M., Hausenloy D.J. Inhibiting mitochondrial fission protects the heart against ischemia/reperfusion injury. Circulation. 2010;121:2012–2022. doi: 10.1161/CIRCULATIONAHA.109.906610. [DOI] [PubMed] [Google Scholar]
  • 153.Qi X., Disatnik M.H., Shen N., Sobel R.A., Mochly-Rosen D. Aberrant mitochondrial fission in neurons induced by protein kinase Cδ under oxidative stress conditions in vivo. Mol Biol Cell. 2011;22:256–265. doi: 10.1091/mbc.E10-06-0551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.D'Angelo G., Duplan E., Vigne P., Frelin C. Cyclosporin A prevents the hypoxic adaptation by activating hypoxia-inducible factor-1alpha pro-564 hydroxylation. J Biol Chem. 2003;278:15406–15411. doi: 10.1074/jbc.M211293200. [DOI] [PubMed] [Google Scholar]
  • 155.Daum S., Schumann M., Mathea S., Aumuller T., Balsley M.A., Constant S.L. Isoform-specific inhibition of cyclophilins. Biochemistry. 2009;48:6268–6277. doi: 10.1021/bi9007287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Fancelli D., Abate A., Amici R., Bernardi P., Ballarini M., Cappa A. Cinnamic anilides as new mitochondrial permeability transition pore inhibitors endowed with ischemia–reperfusion injury protective effect in vivo. J Med Chem. 2014;57:5333–5347. doi: 10.1021/jm500547c. [DOI] [PubMed] [Google Scholar]

RESOURCES