Next Article in Journal
UPF1-Mediated RNA Decay—Danse Macabre in a Cloud
Next Article in Special Issue
Biogenesis of Mitochondrial Metabolite Carriers
Previous Article in Journal
CMH-Small Molecule Docks into SIRT1, Elicits Human IPF-Lung Fibroblast Cell Death, Inhibits Ku70-deacetylation, FLIP and Experimental Pulmonary Fibrosis
Previous Article in Special Issue
Diseases Caused by Mutations in Mitochondrial Carrier Genes SLC25: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Physiopathology of the Permeability Transition Pore: Molecular Mechanisms in Human Pathology

by
Massimo Bonora
1,*,
Simone Patergnani
1,
Daniela Ramaccini
1,
Giampaolo Morciano
1,2,
Gaia Pedriali
1,2,
Asrat Endrias Kahsay
1,
Esmaa Bouhamida
1,
Carlotta Giorgi
1,
Mariusz R. Wieckowski
3,† and
Paolo Pinton
1,2,*,†
1
Department of Medical Sciences, Laboratory for Technologies of Advanced Therapies (LTTA), University of Ferrara, 44121 Ferrara, Italy
2
Maria Cecilia Hospital, GVM Care & Research, Via Corriera 1, Cotignola, 48033 Ravenna, Italy
3
Laboratory of Mitochondrial Biology and Metabolism, Nencki Institute of Experimental Biology of the Polish Academy of Sciences, 3 Pasteur Str., 02-093 Warsaw, Poland
*
Authors to whom correspondence should be addressed.
These authors share senior authorship.
Submission received: 9 June 2020 / Revised: 29 June 2020 / Accepted: 2 July 2020 / Published: 4 July 2020
(This article belongs to the Special Issue Mitochondrial Transport Proteins)

Abstract

:
Mitochondrial permeability transition (MPT) is the sudden loss in the permeability of the inner mitochondrial membrane (IMM) to low-molecular-weight solutes. Due to osmotic forces, MPT is paralleled by a massive influx of water into the mitochondrial matrix, eventually leading to the structural collapse of the organelle. Thus, MPT can initiate outer-mitochondrial-membrane permeabilization (MOMP), promoting the activation of the apoptotic caspase cascade and caspase-independent cell-death mechanisms. The induction of MPT is mostly dependent on mitochondrial reactive oxygen species (ROS) and Ca2+, but is also dependent on the metabolic stage of the affected cell and signaling events. Therefore, since its discovery in the late 1970s, the role of MPT in human pathology has been heavily investigated. Here, we summarize the most significant findings corroborating a role for MPT in the etiology of a spectrum of human diseases, including diseases characterized by acute or chronic loss of adult cells and those characterized by neoplastic initiation.

1. Introduction

Mitochondrial permeability transition (MPT) remains one of the most unusual and poorly characterized aspects of mitochondrial biology. This phenomenon was first reported in the late 1970s and was originally considered an artefact due to the experimental conditions required to investigate isolated mitochondria. However, (as better described below) MPT is triggered by the accumulation of Ca2+ in the mitochondrial matrix that occurs in isolated mitochondria exposed to a [Ca2+] significantly higher than that in the cytoplasm. The development of techniques to measure Ca2+ within different compartments later demonstrated that mitochondria can actively uptake Ca2+ even in living cells, prompting the investigation of MPT in cell pathophysiology.

1.1. Mitochondrial Routes of Cell Death

Mitochondria actively participate in multiple forms of regulated cell death (RCD) through different routes, all involving major alterations in the outer mitochondrial membrane (OMM) and/or inner mitochondrial membrane (IMM). Activation of the mitochondrial pathway of RCA causes the redistribution of mitochondrial proteins into the cytoplasm, which activates cell-death effectors, or the dramatic impairment of cell bioenergetics, which ultimately leads to death of the affected cells. These mechanisms are, in general, categorized into two types that function under different conditions: those involving only outer-mitochondrial-membrane permeabilization (MOMP) and those in which there is also a long-lasting increase in IMM permeability, the MPT. MOMP is due to the formation of a pore composed of the protein B-cell lymphoma (Bcl-2) protein family members Bax and Bak. In response to some apoptotic signals, Bax re-localizes from the cytosol into distinct foci on the OMM. There, Bax oligomerizes into specific structures, such as rings and arc-shaped structures, which can create large-conductance pores in the OMM [1,2,3]. Similarly, Bak (which is mostly constitutively located in the OMM) can homo-oligomerize (Figure 1).
Because of the permeabilization of the OMM, many proteins normally localized within the intermembrane space (IMS) are simultaneously released into the cytosol; these proteins are involved in the effector phase of apoptosis. In particular, (i) cytochrome c (CytC) mediates the organization of the apoptosome and then the activation of the caspase cascade [4]; (ii) apoptosis-inducing factor, (iii) mitochondria-associated, 1 (AIF) induces chromatin condensation; (iv) HtrA serine peptidase 2 (HTRA2) and Diablo IAP-Binding Mitochondrial Protein (SMAC/DIABLO) bind inhibitor of apoptosis (IAP), preventing the inhibition of procaspases; and endonuclease G (ENDOG) mediates DNA fragmentation [5] (Figure 1).
Investigation of the Bcl-2-mediated control of RCD revealed a role of Ca2+-mobilization signals. Indeed, several signals can induce RCD by the selective transfer of Ca2+ from the endoplasmic reticulum (ER, which acts as a store) to mitochondria [6,7,8]. When mitochondria are exposed to a pathological overload of Ca2+, MPT is triggered [9]. MPT is associated with the opening of the mitochondrial permeability transition pore complex (PTPC), a voltage-dependent, high-conductance channel assembled at the interface between the IMM and the OMM [10]. PTPC opening leads to the redistribution of small solutes (<1.5 kDa). The dramatic osmotic influx of water into the mitochondrial matrix during MPT collapses mitochondrial membrane potential (Δψm) and all related activities (including ATP recycling), and is followed by structural collapse (swelling). It results in the release of mitochondrial proteins from IMS which triggers the apoptotic pathway. Alternatively, the incapability of the affected cell to sustain ATP production leads to the irreversible deterioration of ion homeostasis, ultimately resulting in cell death with a necrotic morphology [11] (Figure 1).
It is believed that Ca2+ is the only trigger for PTPC opening, while other factors manipulate the threshold of [Ca2+] required for the occurrence of the event. These include sensitizers (e.g., low Δψm, ROS, high matrix pH, long-chain fatty acids, atractyloside, and carboxyatractyloside) [12,13,14,15,16,17] and desensitizers (Mg2+, ADP, ATP, acidic matrix pH) [11]. Among the many desensitizing factors, the best characterized is cyclosporine A (CsA). This small molecule can inhibit the mitochondrial peptidyl-prolyl isomerase cyclophilin D (CypD) and to date has been considered the gold standard for evaluating the involvement of MPT in pathophysiology.

1.2. Current Hypotheses on PTPC Structure

The structure of the PTPC has been investigated for decades and is still not well characterized. Original studies on isolated proteins and reconstituted liposomes proposed the voltage-dependent anion channel (VDAC) on the OMM and the ADP/ATP translocase (ANT) on the IMM. The inclusion of ANT in the model is important as it possibly represents the inhibition site for ADP/ATP or the activation site for atractyloside. Notably, both proteins could be isolated in a complex containing CypD. Murine CypD is coded by the gene Ppif. The investigation of Ppif-KO mice (then ablated for CypD) robustly confirmed the involvement of CypD in PTPC composition [18]; in contrast, the genetic deletion of VDAC failed to do so. Indeed, when cells from VDAC1−/−; VDAC3−/− mice were treated with siRNA targeting VDAC2, they displayed comparable sensitivity to MPT [19].
A similar approach was conducted to investigate ANT. The knockout (KO) of the three mouse isoforms of ANT indeed showed that PTPC requires a large amount of Ca2+ to open, and that the inhibitory effect of ADP was lost [20]. Interestingly, in cells from ANT triple-KO animals, MPT was still sensitive to CsA, and in CypD-KO cells, MPT was still responsive to ADP [18]. The combined deletion of ANTs and CypD in a quadruple-KO model conferred resistance of MPT to [Ca2+] as high as 5 mM (which is considerably high), potentially indicating that the PTPC did not manifest. These experiments confirmed a role for ANT in the PTPC, but also implied that some other partners of CypD participate in the formation of the PTPC.
Recently, a new candidate was proposed as a pore-forming member of the PTPC—mitochondrial F1/FO ATP synthase (hereafter referred to as ATP synthase). Indeed, (i) genetic manipulation of ATP synthase subunits markedly affected MPT; (ii) isolated ATP synthase, or its C subunit, reconstituted in artificial bilayers generated PTPC-like currents after Ca2+ stimulation; (iii) ATP synthase interacts with CypD; and (iv) molecules that target ATP synthase impair MPT [21,22,23]. Furthermore, the mutagenesis of the beta subunit (β-subunit) and the oligomycin-sensitivity-conferring protein (OSCP) of ATP synthase impaired the effects of Ca2+ and acidic pH, respectively, on the PTPC [24,25]. ATP synthase is usually arranged in dimers that further cluster with oligomers on the IMM. PTPC-derived currents were obtained from monomers or dimers in different experimental settings, opening a debate on which portions of the complex are required [26,27]. Despite this, it was demonstrated that MPT is mediated by the rupture of ATP synthase dimers and that it was preventable by mutagenesis of the C subunit, altering the C-ring conformation [21]. CRISPR/Cas9-mediated deletion of the ATP synthase C or B subunits was still detectable, suggesting that ATP synthase might not be directly involved in the pore formation [28,29]. Later studies in cells devoid of the C subunit revealed that conductance of PTPC was significantly reduced and that the remaining current could still be inhibited by CsA or compounds targeting ANT (ADP and bongkrekic acid) [30]. Multiple hypothesis are now under evaluation to explain these apparently conflicting results: (i) ATP synthase might only indirectly regulate PTPC, by controlling crista structure and ADP levels, (ii) ATP synthase and ANT might form two independent pores among the IMM and, (iii) ANT and ATP synthase (which can interact in the so-called ATP synthasome) might be synergistically required for the proper formation of the PTPC pore.

1.3. Involvement of MPT in RCD Subroutines

Studies on Ppif-KO mice demonstrated that CypD-dependent MPT is fundamental for the activation of RCD with necrotic features. Indeed, CypD-deficient cells are resistant to necrotic cell death induced by reactive oxygen species (ROS) and Ca2+ overload, while stimuli that activate MOMP are insensitive to a lack of CypD [31,32]. Additionally, neuronal cell lines stably overexpressing CypD in mitochondria were prone to necrotic cell death after MPT induction, and were instead more resistant to apoptosis induced by nitric oxide (NO) or staurosporine [33]. This evidence suggests that MPT ultimately results in necrosis and does not induce other forms of cell death (Figure 1). As apoptotic cell death is an active, energy-demanding process, this conclusion is logical for all those conditions in which a marked PTPC opening is triggered, reaching the non-return point of energy depletion that engages the mechanism described above. Nevertheless, some deviation from this model should be considered. Indeed, different reports have shown that in several experimental models, different stimuli that increase intracellular ROS elicit markers of intrinsic apoptotic pathways that can be inhibited by CsA, including mitochondrial proapoptotic protein release, phosphatidylserine exposure, and DNA fragmentation [34,35,36,37,38,39,40,41,42,43,44,45,46,47,48]. This suggests that, at least under selected experimental conditions, submaximal MPT might represent an alternative path for apoptosis activation.
Furthermore, proteins that regulate MOMP are strictly connected to MPT. Several direct protein–protein interactions between Bcl-2 family members and constitutive mitochondrial proteins involved in MPT, such as ANT, VDAC, and ATP synthase, have been confirmed. Bax and Bak are also required for PTPC-dependent necrotic cell death; in fact, the loss of Bax/Bak resulted in resistance to mitochondrial calcium overload and swelling [49]. The confirmation of MPT control by these interactions was demonstrated by the evidence of Bax- and Bak-induced loss of Δψm, mitochondrial swelling, and CytC release through a Ca2+- and CsA-dependent mechanism [50]. In addition, it was demonstrated that tBID induced transient openings of the PTPC associated with a conspicuous remodeling of mitochondrial cristae through a Bak-independent but CsA-inhibitable process [51]. ANT function is under the control of Bcl-xL expression: growth-factor-deprived cells avoid apoptosis via an efficient exchange of ADP for ATP promoted by Bcl-xL, permitting mitochondria to adapt to changes in metabolic demand [52]. Furthermore, Bcl-2 positively regulates ANT activity, while Bax inhibits it by disrupting its interaction with Bcl-2 [53].
Necroptosis is a form of regulated necrosis recently discovered under conditions in which the apoptotic pathway was inhibited, and it presents morphological features of both apoptosis and necrosis. The key upstream kinases involved in the activation of necroptosis are RIPK1 [54], RIPK3, and the substrate MLKL [55], which can be inhibited either through genetic or pharmacological methods to block this type of programmed cell death [56].
RIPK3, which is essential for TNFa-induced necrosis, can inhibit the ADP/ATP exchange mediated by ANT [57], which coincides with the loss of the CypD–ANT interaction, reduced ATP, and the induction of necrotic cell death, suggesting a role of the ANT–CypD interaction in necroptosis [58]. Bax and Bak have also been defined as mediators of necroptosis [59]; in fact, the elimination of Bax/Bak or the overexpression of Bcl-xL leads to the inhibition of the necroptotic process [60]. Furthermore, necroptosis is associated with mitochondrial CytC release and is partly sensitive to CsA inhibition [61].
The investigation of MPT over the years has revealed its importance in multiple subroutines of RCD, prompting investigation into its involvement in human pathology. In the present manuscript, we review the most recent literature discussing the role of the PTPC in human diseases caused by the dysregulation of cell death.

2. PTPC in Acute Conditions

2.1. MPT during Ischemia

Ischemia and consequent reperfusion injury (RI) are pathological manifestations in which the involvement of MPT has been robustly confirmed. Ischemia is characterized by the reduced oxygenation of a portion of a tissue (hypoxia), which results in the loss of tissue functions and eventually the activation of multiple forms of RCD. Ischemia impacts multiple organs, especially those that are more susceptible to hypoxia, such as the heart, brain, and kidney. Interestingly, not all tissues share the same susceptibility to ischemia/RI. In fact, the severity of the injury largely depends on how different types of cells, and therefore tissues, can survive under hypoxic conditions [62,63].
Mechanistically, a lack of oxygenation translates into reduced activity of the electron transport chain (ETC), and hence results in the blockage of oxidative phosphorylation (OXPHOS) and a consequent reduction in ATP recycling. In this scenario, to meet energy demands, cells upregulate anaerobic glycolysis, producing lactic acid and hydrogen ions, which results in intracellular acidosis. By neutralizing pH via the activation of Na+/H+ antiporter (NHE), the cell undergoes sodium accumulation that is counterbalanced by the reverse activity of Na+/Ca2+ exchanger (NCX). Concomitantly, the reduction in available ATP depresses the activity of Na+/K+ ATPase, plasma membrane calcium ATPase (PMCA), and sarco-/endoplasmic reticulum Ca2+-ATPase (SERCA), which ultimately results in an overload of cytosolic Ca2+ [64]. Additionally, the reduced pO2 causes the accumulation of electrons among different respiratory complexes, leading to the production of ROS. The concomitant increase in intracellular Ca2+ and ROS and the decrease in Δψm favors the induction of MPT. This mechanism has been demonstrated in cultured neonatal rat cardiomyocytes [65], hepatocytes [66], and immortalized cells [67].
Among all tissues, the brain exhibits high sensitivity to ischemia due to its glucose-dependent metabolisms [68]. Rapid and CsA-dependent mitochondrial depolarization is reported to occur early during experimental stroke in the mouse somatosensory cortex in vivo [69]. Additionally, CsA protects the retinal ganglion from cell death during acute intraocular pressure (IOP) elevation, a peculiar inducer of ischemia [70]. Furthermore, CsA administration attenuates hypoxic–ischemic brain injury in newborn rats induced by unilateral carotid artery ligation [71].
Despite this, MPT is commonly believed to occur to a minimal extent during ischemia. Indeed, as described, hypoxia leads to the accumulation of protons and ADP, which are strong inhibitors of PTPC. During ischemia, the manifestation of MPT depends on the subtle equilibrium between PTPC inducers (Ca2+ and ROS) and inhibitors (protons and ADP), which accumulate as a result of impaired mitochondrial respiration and excess glycolysis (Figure 2). Furthermore, the adaptive response to hypoxia-mediated by HIF1a affects PTPC opening by regulating hexokinase II (HKII) levels [72], another reported PTPC regulator. Indeed, when HIF1a is stabilized (e.g., following hypoxia or GSK360A administration), HKII protein expression significantly increases in the mitochondrial fraction, and this is directly involved in the cytoprotective effect started by HIF1a stabilization. Genetic depletion of HKII completely abolished this path even in the presence of an activated HIF1a [72]. Nevertheless, HKII is considered an inhibitor of the PTPC only when bound to the OMM (especially to VDAC), rather than when generally overexpressed in mitochondria.
Nevertheless, under this condition, Ca2+ overload may induce MPT-independent cell death as a result of the excessive activation of Ca2+-dependent enzymes such as phospholipases, proteases, and endonucleases [73,74]. Noticeably, during myocardial infarction, the activation of calpains, Ca2+-dependent cysteine proteases, results in myofibril disruption, thus promoting hypercontracture in the heart, which consists of sustained shortening and stiffening of the myocardium [75,76].

2.2. Role of MPT in Reperfusion Injury

The situation dramatically changes during tissue reperfusion. Indeed, the restoration of pO2 recovers respiration, ATP synthesis, and the activity of plasma membrane pumps, which re-equilibrate intracellular pH. As a result, the inhibition of MPT by ADP and protons is removed, lowering the threshold for PTPC opening. In addition, reperfusion stimulates ROS production by multiple sources [77]. At the mitochondrial level, succinate is observed to accumulate in the mitochondrial matrix during ischemia and it was proposed to stimulate ROS production from complex I through reverse electron transport (RET), at the time of reperfusion [78]. Other pieces of evidence suggest that ETC conditions are not favorable for RET at the beginning of reperfusion, and that a burst of ROS production only occurs after a first wave of MPT. Besides mitochondria, reperfusion induces ROS via other enzymes including (but not limited to) xanthine oxidase, NADPH oxidase, and nitric oxide synthase [77]. These are believed to further stimulate ROS generation in the mitochondria, forming a vicious cycle and making mitochondria de facto the largest ROS source during reperfusion [79,80]. The elevated superoxide anion (O2) reacts with nitric oxide (NO), producing the highly reactive peroxynitrite. This leads to a reduction in the availability of NO (which is a potent vasodilator) and causes the accumulation of neutrophils [77]. Furthermore, oxidative damage causes lipid peroxidation, DNA damage, and enzyme denaturation and activates the innate anti-inflammatory response, aggravating reperfusion injury (RI) [81].
The occurrence of MPT during the reperfusion phase was demonstrated by the experiments of Griffith and Halestrap in 1995. They showed that mitochondrial accumulation of radioactive deoxy glucose (hot-dog), which can pass through the IMM only during MPT, did not occur in isolated hearts undergoing ischemia, but was significantly induced (and inhibitable by CsA) during reperfusion [82].
In the past decade, several in vitro and in vivo studies have confirmed the involvement of MPT in ischemia/RI [83]. The administration of CsA (as well as its analog FK506) has been shown to protect against ischemia/RI in multiple animal models and in multiple tissues, including cardiac and skeletal muscle, brain, kidney, liver, lungs, and testis [84,85,86,87,88,89]. Ppif-KO mice are significantly protected from ischemia/RI in both cardiac muscle and the brain, in terms of both tissue function and survival rates [31,32,90]. In addition to CypD manipulation, genetic interference in mechanisms of Ca2+ homeostasis proved the importance of MPT in ischemia/RI. Cardiac-specific ablation of NCX significantly decreased ischemia/RI in isolated hearts [91]. Similarly, mice overexpressing Bcl2 have reduced [Ca2+]m accumulation and are therefore protected from myocardial ischemia/RI [92]. Accordingly, tissue-specific overexpression of the mitochondrial Na+/Ca2+ exchanger (TER-NCLX) in cardiac muscle accentuates the extrusion of Ca2+ from mitochondria to the cytosol and then suppresses [Ca2+]m, decreasing the sensitivity of cardiomyocytes to PTPC opening. Hearts from TER-NCLX mice also display protection from left coronary artery ligation-induced ischemia/RI [93]. Additionally, impairing mitochondrial calcium uniporter (MCU), using the inhibitor Ru360 [94] or in MCU-KO mice [95], lowers the uptake of mitochondrial Ca2+ and is correlated with increased brain and heart function. We recently demonstrated that inhibiting ATP synthase via N,N-dicyclohexylcarbodiimide (DCCD) partially recovered contractility of isolated heart exposure to ischemia/RI by Langhendorff apparatus. Most interestingly, we observed that serum levels of the C subunit in patients with ST-segment elevation myocardial infarction (STEMI) correlated to several surrogate markers of myocardial reperfusion [96].

2.3. Role of MPT in Acute Kidney Injury

Renal tissue is also known to be affected by MPT in multiple conditions. Acute kidney failure (mostly known as acute kidney injury, AKI) is often characterized by extensive necrosis. The protective effect of CsA or CypD inactivation on experimental ischemia/RI has been largely reported in kidneys, as previously stated. Atherosclerotic renal artery stenosis (ARAS) is probably the most frequent cause of ischemia/RI-related AKI. The outcomes of experimental ARAS are significantly improved by exposure to the mitochondrial-targeted peptide Elamipretide. This peptide displays multiple protective functions, including the buffering of ROS and the stabilization of the structural mitochondrial lipid cardiolipin. Specifically, in experimental ARAS, the protective effect of Elamipretide is believed to be mediated by desensitization to PTPC opening [97].
Renal tissue can also undergo, at nominal pO2, conditions resembling manifestation of RI and that appear to be dependent on MPT. For example, an important cause of AKI with extended necrosis is exposure to drugs (e.g., FANS or chemotherapy) and crystal nephropathies. The etiology of these ischemic-like conditions is not yet fully comprehended, although it is proposed to act through excess ROS production, which ultimately triggers PTPC opening. In support of this model, under conditions of oxidative stress, glycogen synthase kinase-3β (GSK3β), an interactor and regulator of the putative PTPC, translocates from the cytosol to mitochondria in a VDAC2-dependent manner, promoting PTPC opening [98]. Inhibition of GSK3β promotes resistance to MPT in mice undergoing paraquat- or diclofenac-induced nephrotoxicity [99,100]. Interestingly, it is now well recognized that cell death during AKI is significantly dependent on necroptosis, a finding largely confirmed by investigations on Mlkl-KO and Ripk3-KO animals [101]. Cisplatin-induced AKI is protected by the inactivation of both Ripk3 and CypD. Ripk3/Ppif-double-KO models are more protected than the models with either KO [102], suggesting that MPT and Ripk3 contribute to RCD by independent but concomitant pathways. A recent study by Mulay et al. using CypD-KO and Mlkl-KO mice showed that experimental conditions mimicking acute oxalosis (crystal nephropathy characterized by sudden increases in serum oxalate levels) resulted in kidney failure with necroptotic features that was strongly dependent on MPT [103]. That study, however, did not report significant differences among the Ppif -KO, Mlkl-KO and Ppif/Mlkl-double-KO mice in terms of protection from cell death.

3. PTPC in Degenerative Conditions

Degenerative disorders are human diseases characterized by the chronic loss of fully differentiated cells, which leads to the progressive impairment of the structure and/or function of the affected tissue/organ. Among the most frequent and probably most investigated degenerative disorders are conditions that affect the CNS or the skeletal or cardiac muscle. As previously mentioned, these organs all require large amounts of energy; therefore, it is not surprising that the mutations most frequently associated with degenerative disorders are linked to mitochondria. Intriguingly, many of these conditions share impaired mitochondrial respiration and ATP production, increased ROS production, and altered PTPC sensitivity. The involvement of MPT in these disorders appears to be more complex than the involvement of MPT in ischemia/RI in terms of molecular and cellular interactions. We therefore summarize the major observations made in different degenerative conditions.

3.1. Role of the PTPC in Protein-Aggregation-Related Neurodegenerative Diseases (NDs)

The primary features of neurodegeneration are the abnormal presence and accumulation of mutant and/or damaged proteins. Protein aggregation is the main cause of changes in the intracellular environment such as oxidative status, impaired protein quality control system, transcriptional alteration, and mitochondrial dysfunction. All these variations critically contribute to the pathogenesis of NDs and culminate in neuronal cell death. Among the diverse types of proteins that aggregate, amyloid-beta (Aβ), Tau, and alpha-synuclein (αSyn) are the most commonly studied and represent the primary cause of sporadic and familial Alzheimer’s disease (AD) and Parkinson’s disease (PD), the most prevalent NDs.
Interestingly, Aβ, Tau protein, and αSyn are found in the mitochondria of patients affected by AD or PD [104,105,106] or their related animal models [107,108,109]. In particular, these protein aggregates have been observed to colocalize or directly interact with multiple partners of the PTPC, such as CypD, VDAC, ANT, and ATP synthase [104,107,109,110].
Additionally, Aβ exposure in cultured cortical neural progenitor cells induces PTPC opening [111]. In particular, short Aβ exposure led to decreased cell proliferation. However, when the exposure and thus PTPC opening were prolonged, CsA-inhibitable necrotic RCD was activated. Consistent with this result, intravital multiphoton imaging of AD mouse models demonstrated that near senile Aβ plaques, mitochondria showed severe structural and functional abnormalities, suggesting that senile plaques are the main source of toxicity in vivo [112]. Aβ plaques and phosphorylated Tau interact with VDAC, leading to mitochondrial dysfunction [105]. Similarly, a specific Tau fragment (NH2-26-44 fragment) affects OXPHOS and mitochondrial dynamics by interacting with ANT and impairing PTPC regulation [113].
Furthermore, αSyn oligomers move into mitochondria and colocalize with ATP synthase, inducing its oxidation concomitantly with increased PTPC opening, mitochondrial swelling, and necrosis activation [114]. Additionally, a PD mouse model characterized as having a mutant human αSyn (Thy1-hαSyn-A53T tg mice) proved that αSyn associates with neuronal mitochondria and interacts with VDAC and CypD in vivo [115]. This work directly linked motor abnormalities and neuropathology to the PTPC.
The study of rare inherited mutations in PD has provided insight into the molecular mechanisms of mitophagy, the regulated delivery of dysfunctional mitochondria to lysosomes via autophagic machinery. Multiple PD-related genes have been identified, among which the mitochondrial kinase PINK1 and the cytosolic E3 ubiquitin ligase Parkin are the most characterized. The PTPC may be involved in mitophagy-mediated quality control processes, which play a critical role in conserving neuronal health and function. Indeed, it has been suggested that the opening of the PTPC regulates mitochondrial depolarization and subsequent mitochondrial degradation in autophagosomes. Furthermore, CsA and its analogs block autophagosome formation, and the alteration of PTPC opening has been unveiled in models lacking PINK1, Parkin, and DJ1, which are components of the best-characterized stress-induced mitophagy pathways [116,117]. Consistently, the downregulation of PINK1 in mouse neurons resulted in altered mitochondrial morphology and function, ROS production, and finally PTPC opening. All these events were accompanied by the induction of mitochondrial autophagy [118].
Similarly, PINK1-deficient neurons showed selective increases in mitochondrial Ca2+, PTPC opening, and defective mitochondrial respiration. In addition, the inhibition of PTPC opening was found to be sufficient to rescue the mitochondrial impairments observed in Pink1/− cells [119]. ROS production and PTPC opening were increased in primary mouse embryonic fibroblasts (MEFs) and brains from Park7−/− (the gene coding for DJ-1) mice compared with wild-type (WT) samples. In contrast, antioxidant molecules decreased ROS levels and PTPC opening. Interestingly, in contrast to Pink1−/− cells, the lack of DJ-1 did not affect mitochondrial respiration and Ca2+ dynamics, suggesting that DJ-1 has a possible antioxidant role. Finally, Parkinsonian toxins (such as 6-hydroxydopamine and neurotoxin1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine) are widely employed as in vivo and in vitro chemical models of PD and have been found to be potent activators of both PTPC and mitophagic processes [120,121]. Notably, both VDAC and ANT were demonstrated to be required for proper mitophagy [122,123]. Finally, recent pieces of evidence demonstrated that human samples obtained by AD-affected patients displayed inhibited damaged mitochondrial clearance [124] and that mitophagy activation diminished insoluble Aβ and Tau hyperphosphorylation to revert cognitive impairments in an AD mouse model [125,126] (Figure 2).
Taken together, these findings suggest that the mitochondrial accumulation of disease-specific protein aggregates might favor MPT via direct interactions or mitophagy impairment, which leads to the accumulation of mitochondria prone to PTPC opening and RCD (Figure 2).

3.2. Amyotrophic Lateral Sclerosis (ALS) and PTPC

ALS is the most common neuromuscular degenerative disease affecting adults. While several works have suggested a protein-aggregation origin, this progressive and severely disabling fatal neurological disease is generally considered to have multifactorial causes. Currently, there are no cures or effective treatments for ALS, and the molecular pathogenesis of ALS is poorly understood.
Recent findings show that mitochondrial perturbations are implicated in the pathogenesis and progression of ALS. Altered fission–fusion dynamics, altered mitochondrial Ca2+ homeostasis, excessive oxidative stress, reduced OXPHOS activity, and decreased proapoptotic factor release have been found in ALS models in vitro and in vivo. Additionally, the PTPC is emerging as a critical player in ALS. In a transgenic model of ALS (G93Ahigh), a profound alteration in mitochondrial structures with increased PTPC activity was observed [127]. Interestingly, dendritic mitochondria from the same ALS animal model displayed increased contact sites between the IMM and OMM. This conformation might favor the formation of the PTPC. Furthermore, the deletion of CypD delayed disease onset and extended the survival of transgenic ALS mice [127]. Consistent with this finding, the exposure of two independent ALS murine models to the novel PTPC inhibitor GNX-4728 protected against motor neuron degeneration and mitochondrial impairment and promoted their survival nearly 2-fold [128].
Preclinical studies have shown that olesoxime, a member of the cholesteroloxime family, improves the survival of neural cells and reduces the effects of oxidative stress by modulating the PTPC. In particular, this compound concentrates in mitochondrial compartments, where it binds the PTPC interactors VDAC and mitochondrial translocator protein (TSPO). Following olesoxime binding, the PTPC was desensitized, leading to neural cell protection both in vitro and in vivo [129,130]. Accordingly, the potent antioxidant and inhibitor of PTPC edaravone (Radicut™) has been approved for the treatment of ALS [131]. Compounds derived from cinnamic anilides such as GNX-4728 and GNX-4975 also represent an MPT-based treatment of ALS. In a murine model of ALS, these compounds delayed the onset of symptoms, increased lifespan, and reduced the inflammatory response [128].

3.3. Multiple Sclerosis (MS) and PTPC

MS is the most common primary demyelinating disease of the brain. MS is an inflammatory T-cell-mediated autoimmune disease characterized by progressive demyelination, gliosis (scarring), and neuronal loss [132]. Recently, mitochondrial dysfunction has been increasingly linked to the pathogenesis of MS. Additionally, impaired mitochondrial enzyme complex activity [133,134], increased oxidative stress [135], altered mitochondrial DNA [136], impaired quality control systems [137,138], and abnormal mitochondrial number and morphology have been described in MS patients [133] and in vivo MS mouse models [139,140]. In this context, it was discovered that the PTPC might also contribute to MS; indeed, CypD-KO mice with experimental autoimmune encephalomyelitis (EAE), a commonly used animal model for MS, recovered from their induced disabilities. Furthermore, axonal damage was decreased, and the mitochondria of cultured CypD-KO neurons accumulated higher levels of Ca2+ and were more resistant to oxidative stress compared to WT [141]. Subsequent studies confirmed this finding and demonstrated that the selective inhibition of PTPC exerted neuroprotective effects on the EAE model by increasing mitochondrial function, reducing oxidative stress, and blocking mitochondrial swelling and Ca2+-mediated PTPC formation [142]. The p66Shc protein may also modulate the mitochondrial dynamics and PTPC opening that occur during neurodegeneration in EAE. P66Shc is a product of the ShcA gene normally localized in the cytoplasm. However, once phosphorylated by protein kinase C-Beta and following interaction with PIN1, P66Shc moves to the mitochondria, where it regulates distinct cellular processes such as apoptosis and autophagy [143,144]. Furthermore, in the mitochondrial compartment, p66Shc works as a ROS amplifier by generating mitochondrial ROS to induce PTP opening. Consistent with this function, when EAE was induced in p66Shc-KO (p66Shc−/−) mice, the clinical symptoms (manifested as limb weakness and paralysis) of these model mice were less severe than those of WT mice [145]. The fact that the onset and development of EAE in p66Shc/Cyc-D-double-KO mice were identical to those observed in p66Shc−/− mice validates the role of the p66Shc-PTPC pathway in neurodegeneration and confirms that once activated, p66Shc interacts with PTPC to promote its opening [146].

3.4. PTPC in Muscular Dystrophies (MDs)

MDs refer to a clinically and genetically heterogeneous group of degenerative muscle diseases, manifested primarily as the progressive weakness and degeneration of skeletal muscles that control movement, resulting in severe pain, disability, and ultimately death. Some forms of MD also affect cardiac muscle [147,148]. Current therapies to treat muscle degenerative diseases are still limited by poor targeting, although promising new therapeutic directions remain [149].
The most common and severe form of human muscular dystrophy is Duchenne MD (DMD), which is an X-linked recessive genetic disorder associated with respiratory complications and cardiac dysfunction [150]. The disease is caused by a genetic defect—the absence of the cytoskeletal protein dystrophin, the primary function of which is to link the myofiber cytoskeleton to the extracellular matrix, stabilizing the sarcolemma [151,152,153]. Although the gene underlying the disorder was identified in 1987 [154], the pathophysiology leading to disease remains unclear. Numerous mitochondrial alterations are correlated with dystrophic conditions, including impaired ATP production, substrate handling, Ca2+ buffering capacity, and elevated ROS production. The mitochondria in muscular fibers from dystrophic mdx mice (a murine model of DMD) displayed a significantly shorter time to MPT induction in response to Ca2+ than WT mice [155]. In C. elegans and zebrafish models of DMD, mitochondrial fragmentation is detectable before overt signs of muscular degeneration, and CsA feeding delays muscle degeneration [156]. Cyclosporine also inhibits calcineurin, a signaling protein involved in skeletal muscle, and its inhibition was found to worsen muscular dystrophy in an mdx mouse model [157,158]. Nonetheless, the deletion of Ppif and the administration of Debio-025 (a CsA inhibitor with no effect on calcineurin) prevented dystrophic conditions in mdx and d-sarcoglycan-KO (Scgd−/−) animals [159,160]. Additionally, several lines of evidence have shown that Ppif deletion is protective in Col6a1−/− mice [161].
Mutations of collagen VI (ColVI) genes encoding the extracellular matrix protein, which is abundant in skeletal muscle, cause three muscle diseases in humans: Ullrich congenital muscular dystrophy (UCMD), Bethlem myopathy (BM), and the recently identified myosclerosis myopathy (MM) [162,163]. These collagen VI myopathies are inherited muscle diseases that share mitochondrial dysfunction due to altered PTPC opening [164]. Mouse models lacking collagen VI (Col6a1−/−) display an early onset myopathic phenotype correlated with ultrastructure defects in mitochondria and the sarcoplasmic reticulum (SR), altered mitochondria caused by eventual inappropriate PTPC opening, and elevated muscle fiber apoptosis [161,165,166]. In addition, the absence of ColVI led to a marked decrease in the expression of proapoptotic Bcl-2, which may synergize with calcium to enhance the opening of the PTPC and eventually promote the release of mitochondrial proapoptotic factors [167].
In light of this fact, the altered expression of collagen in Col6a1−/− mice and in BM and UCMD patients correlates with enhanced PTPC opening, resulting in the functional and ultrastructural deficiency of mitochondria, followed by impaired autophagy, [168]. As such, autophagy is altered in MDs, and autophagy activation due to low amino acid intake improved the skeletal muscle phenotype of a DMD mouse model (mdx), [169]. Consistently, pharmacological treatment with CsA has been reported to dramatically recover myofiber degeneration in a Col6a1−/− mouse model and UCMD patients [161,166]. Additionally, it has been demonstrated that mitochondria-mediated cell death can be reduced by CsA in Ullrich congenital muscular dystrophy models [167].

4. Mitochondrial Disorders

Mitochondrial diseases are a clinically heterogeneous group of disorders that arise because of mitochondrial respiratory chain dysfunction. These diseases are caused by mutations in nuclear DNA (nDNA) and in mitochondrial DNA (mtDNA). Although mitochondrial diseases can involve any organ or tissue, they characteristically involve multiple systems, typically affect organs that are highly dependent on aerobic metabolism and are often relentlessly progressive with high morbidity and mortality. Like other degenerative diseases, the involvement of MPT in mitochondrial diseases has been proposed.
Mutations in mtDNA are responsible for the etiology of the most frequent mitochondrial diseases, especially Leber’s hereditary optic neuropathy (LHON); neurogenic muscle weakness, ataxia, and retinitis pigmentosa (NARP); mitochondrial encephalomyopathy, lactic acidosis, and stroke-like episodes (MELAS); and myoclonic epilepsy with ragged red fibers (MERRF). These alterations most frequently involve single-base mutations in genes encoding components of respiratory complex I, ATP synthase, transfer RNA, and DNA polymerase, but also may be caused by large deletions of mtDNA. Hybrids carrying mtDNA mutations associated with LHON, NARP, MELAS, and MERRF displayed poor resistance to oxidative stress, which could be prevented by the administration of CsA or the deprivation of extracellular Ca2+ [170]. Accordingly, many of these mutations lower the threshold for PTPC opening in response to Ca2+ and ROS [171,172].
Mutations in nDNA associated with mitochondrial diseases have also been also related to alterations in MPT activity. Mutations in leucine-rich pentatricopeptide repeat containing (LRPPRC), a protein involved in the maturation and stability of mitochondrial RNA [173], cause the French-Canadian variant of Leigh syndrome. Loss of LRPPRC results in defects in the assembly of respiratory complexes IV and V, leading to severe metabolic alterations. Fibroblasts isolated from patients presenting LRPPRC inactivation displayed multiple types of mitochondrial dysfunction, including a reduced threshold for Ca2+-induced MPT [174]. Interestingly, all the mitochondrial alterations mentioned so far are often associated with impaired or unstable assembly of ATP synthase, providing significant clues to its involvement in MPT. Direct investigations of ATP synthase alterations and MPT in the etiology of mitochondrial disease have not been performed. Other significant nuclear genes relating mitochondrial diseases to MPT are optic atrophy 1 (OPA1) and Spastic paraplegia 7 (SPG7). OPA1 is an essential protein involved in the fusion and cristae arrangement of the IMM. Its mutations manifest clinically as optic atrophy, ataxia, and deafness [175]. SPG7 (paraplegin) is an ATP-dependent zinc metalloprotease located in the mitochondrial matrix, and its mutations are associated with chronic progressive ophthalmoplegia [176]. Interestingly, both proteins positively regulate the MPT threshold in response to Ca2+ induction, and their inactivation significantly inhibits PTPC opening [177,178]. Mechanisms by which OPA1 and SPG7 act on PTPC are still to be defined. While the role of OPA1 seems to be dependent on its control of crista morphology [179], SPG7 appears to regulate the amount of Ca2+ available for MPT via regulation of mitochondrial calcium uniporter assembly [180]. Taken together, this evidence indicates that PTPC might be a significant target for mitochondrial disease; however, the mechanism by which MPT influences these diseases is still poorly understood and might differ among syndromes, calling for further investigations in more complex experimental models.

5. PTPC in Nonalcoholic Fatty Liver Disease

Nonalcoholic fatty liver disease (NAFLD) is among the most prevalent chronic liver diseases in both children and adults, and is predicted to be the primary cause for liver transplants by 2020 [181]. NAFLD is characterized by an accumulation of fat (steatosis) in the liver, which can progress to inflammatory NASH and into more severe stages: fibrosis, cirrhosis, and hepatocellular carcinoma [182]. The prevalence of NAFLD has risen rapidly in Western societies, particularly in most of the European countries, due to an increase in mass consumption of highly processed ready-made food, which is rich in fructose and saturated fat. NAFLD has risen rapidly in parallel with the recent surge in metabolic-related diseases such as obesity and type 2 diabetes mellitus (T2DM), which have been indicated as risk factors for the prevalence and progression of NAFLD. Despite the attempts of the liver to recover from fat accumulation, in the long run, mitochondrial adaptation is insufficient to prevent lipotoxicity due to continuous FFA accumulation [183]. At this later time point, mitochondria present alterations in the OXPHOS complexes, mitochondrial membrane potential, reduced ATP synthesis, and induced PTPC opening [183]. Opening of the PTPC may be the basis of the steatosis-induced apoptosis of hepatocytes observed in vitro, and may be related to the steatosis in NAFLD of human beings [184]. It has been postulated that stimulated PTPC opening in a rat model of NAFLD is the result of increased Bax expression and aberrant Bcl-2/Bax ratio. This seems to be an important mechanism of the mitochondrial damage in hepatocytes that occurs in NAFLD [185]. Wang et al. proposed that the overexpression of mitochondrial hepatic CypD induced mitochondrial stress and could be an early event that leads to the liver steatosis [186]. They found that overexpression of CypD is manifested in mitochondrial swelling and increased mitochondrial ROS production. Such mitochondrial perturbations provoke ER stress through Ca2+/p38 MAPK activation, finally resulting in the increase of SREBP1c-mediated synthesis of triglycerides. Interestingly, in mice fed with a high-fat diet, the increased level of CypD was observed earlier than triglyceride accumulation in the liver. Moreover, Wang et al. speculated that CypD knockout or pharmacological inhibition of CypD could ameliorate triglyceride accumulation in HFD-fed mice [186]. On the other hand, the observations of Lazarin et al. indicated that mitochondria isolated from livers of monosodium l-glutamate obese rats were less susceptible to the opening of PTPC by calcium [187]. Regardless of several pieces of evidence for the involvement of the MPT in the NAFLD animal models, there is no direct evidence supporting the role of PTPC and especially CypD in NAFLD in humans.

6. PTPC in Cancer

One of the earliest established hallmarks of cancers is the resistance of transformed cells to RCD. Considering the discussed role of MPT in RCD, it is straightforward to hypothesize that the alteration of the PTPC machinery is involved in the establishment of neoplasia. According to this hypothesis, MPT is predicted to have a tumor-controlling mechanism, and its suppression is required for tumor development. It is harder to prove that a phenomenon does not occur than it is to prove that it does occur. Indeed, strong direct evidence confirming or confuting this hypothesis is lacking. Experiments based on the genetic manipulation of CypD for the other disease types discussed should provide significant evidence, but nothing of this kind has yet to be reported. Still, reduced MPT in transformed cells could be predicted by a large number of findings, as discussed below.
Another significant hallmark of cancer is metabolic rewiring, especially the abnormal increase in the glycolytic rate at almost normal pO2 (Warburg effect). This large glucose consumption causes the significant conversion of pyruvate into lactate, resulting in intracellular acidification and leading to PTPC desensitization similar to that observed in ischemia [6]. Furthermore, many solid tumors develop a hypoxic area that, analogous to ischemic tissue, also promotes Hif1a accumulation and HKII-mediated desensitization of PTPC. The Warburg effect then allows the hypoxic tumor to limit Pi and Ca2+ accumulation, similarly to ischemia, favoring the inhibitory effect of low pH. If, by analogy to ischemia, an increase in [Ca2+] could be expected, it is true that transformed cells have lower intracellular Ca2+ (Figure 2). Indeed, it has been shown that H-RAS-driven transformation is concomitant with a progressive reduction in the amount of intracellular Ca2+ [188]. Bcl-xL has been shown to negatively regulate PTPC opening by directly interacting with VDAC [189]. Additionally, Bcl-xL can interact with the β subunit of ATP synthase to promote its synthase activity and inhibit PTPC [190].
Furthermore, many other oncogenes and oncosuppressor genes regulate Ca2+. Among the most characterized is Bcl-2, which can impair [Ca2+] in the ER lumen and its transfer to mitochondria via multiple proposed mechanisms [191]. Additionally, many tumors show alterations in the PI3K pathway. Members of this pathway, AKT and PTEN, can localize at mitochondrion-ER contact sites, especially with Ip3R, to alter its activity [192,193,194,195,196]. This effect is coordinated by the oncosuppressor PML [197,198,199]. Loss of PML or PTEN and activation of AKT (conditions prototypical of multiple tumor types) lead to the inactivation of IP3R, which also limits the Ca2+ available to mitochondria. Additionally, AKT phosphorylates GSK3β (altered in several cancer types), resulting in its inactivation and, as discussed, PTPC desensitization [200]. Finally, the master oncosuppressor p53 interacts with SERCA to maintain a high level of ER [Ca2+]. Loss or inactivation of p53 (one of the most common alterations in cancer) impairs [Ca2+] and results in reduced sensitivity to MPT-mediated RCD [201,202,203,204,205]. In addition, p53 can localize to mitochondria and interact with CypD to favor MPT and necrosis.
Interestingly, a network of chaperones seems to interact with CypD (which could be considered a chaperone itself) to modulate PTPC. In particular, HSP90, HSP60, and DnaJC15, which are often overexpressed in tumors, have been shown to interact with CypD, leading to their inhibition (then mimicking CsA) and suppressing RCD initiation [206,207,208]. These findings are supportive of the hypothesis, but contradict other strong evidence. First, transformed cells often have increased levels of ROS, not of a significant increase in multiple scavenging systems. It is currently accepted that elevated ROS can act as a mitogenic signal in tumors to support proliferation [209,210]. In addition, studies in cancer cell lines and tumor models have shown that different PTPC members are overexpressed, especially TSPO, VDAC, and ANT, possibly to favor their specific metabolic condition [16,211,212,213,214,215,216,217]. We can, therefore, speculate that tumor cells can survive pressure selection by the highly regulated suppression of MPT, which allows the maintenance of MPT-related features with potential mitogenic effects (Figure 2).

7. Potential of PTPC Targeting and Concluding Remarks

Because of the presented evidence (and much more), MPT has been investigated for the treatment of human disease in multiple clinical trials. CsA entered a trial procedure that lasted 15 years and ended with failure at phase III in cardiac RI. Indeed, the CIRCUS [218] and CYCLE [42] trials consisting of a single intravenous bolus of CsA (2.5 mg/kg) before revascularization had no effect on ST-segment resolution or cardiac enzymes and did not improve clinical outcome. These findings were reconsidered by a study by Piot et al., who saw hope in the use of CsA to treat RI. Similarly, TSPO targeting by 3,5-Seco-4-nor-cholestan-5-one oxime-3-ol (TRO40303) was tested in clinical trials, due to promising cardioprotective effects on a rat model of cardiac ischemia. However, the desensitization of PTPC opening seemed to be secondary to its remarkable antioxidant properties [219]. Indeed, studies on TSPO-KO mice showed that the protein was dispensable in models of ischemia/RI [220]. However, the safety and efficacy of this drug were evaluated a few years later in patients undergoing percutaneous coronary intervention (PCI). This multicenter, double-blinded, phase II study (MITOCARE) showed the inefficacy of the compound in reducing or limiting RI [221]. Additionally, an oral version of CsA was tested for the treatment of LHON but failed to reach the primary endpoint of the study, although it delayed the onset of the disease. [222]. There are multiple reasons that CsA-based trials have failed to reproduce the protection reported in preclinical studies, but a discussion of these failures is not the purpose of this review. The investigation of novel compounds able to target the PTPC is ongoing, and many investigations have already yielded promising results in preclinical studies. Most of these are designed to be more potent and specific inhibitors of CypD, including the small molecules C-9, C-19, and C-31, which have already been proven to be protective in models of AD, acute pancreatitis, and hepatic injury [223,224,225].
Library screening has also identified ML-404 [226] and N-phenylbenzamide [227] as CypD-independent, offering the possibility to combine their use with CsA to limit its side effects. It is of interest that ATP synthase is now the subject of investigation in these terms. Oligomycin and DCCD, which target the C subunit, displayed powerful MPT inhibition in vitro, but they also depleted mitochondrial ATP, causing an additional injury [21].
We recently generated a library of small molecules that target the C subunit by modifying the functional core of oligomycin and obtained new patented compounds able to notably reduce reperfusion damage in animal models of global ischemia without interfering with ATP production [22]. Finally, the natural hormone melatonin is of great interest. This hormone can act as an antioxidant and can modulate the PTPC, although its exact mechanism is currently under investigation [228,229]. Currently, melatonin is considered the safest drug that can be used as a PTPC inhibitor, and we will probably see an increasing number of investigations on this molecule in the future.

Funding

P.P. is grateful to Camilla degli Scrovegni for continuous support. The Signal Transduction Laboratory is supported by the Italian Association for Cancer Research (AIRC: IG-23670 to P.P. and IG-19803 to C.G.), A-ROSE, Telethon (GGP11139B to P.P.), Progetti di Rilevante Interesse Nazionale (PRIN2017E5L5P3 to P.P and PRIN20177E9EPY to C.G.), the Italian Ministry of Health (GR-2013-02356747 to C.G.), the European Research Council (ERC, 853057-InflaPML to C.G.), local funds from the University of Ferrara to P.P. and M.B. S.P. was supported by Fondazione Umberto Veronesi. MRW was supported by the National Science Centre, Poland (UMO-2018/29/B/NZ1/00589). Moreover, MRW gratefully acknowledge the financial support from the FOIE GRAS and mtFOIE GRAS projects. These projects received funding from the European Union’s Horizon 2020 Research and Innovation program under the Marie Skłodowska-Curie Grant Agreement No. 722619 (FOIE GRAS) and Grant Agreement No. 734719 (mtFOIE GRAS).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dejean, L.M.; Martinez-Caballero, S.; Guo, L.; Hughes, C.; Teijido, O.; Ducret, T.; Ichas, F.; Korsmeyer, S.J.; Antonsson, B.; Jonas, E.A.; et al. Oligomeric Bax is a component of the putative cytochrome c release channel MAC, mitochondrial apoptosis-induced channel. Mol. Biol. Cell 2005, 16, 2424–2432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Dejean, L.M.; Martinez-Caballero, S.; Manon, S.; Kinnally, K.W. Regulation of the mitochondrial apoptosis-induced channel, MAC, by BCL-2 family proteins. Biochim. Biophys. Acta 2006, 1762, 191–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Salvador-Gallego, R.; Mund, M.; Cosentino, K.; Schneider, J.; Unsay, J.; Schraermeyer, U.; Engelhardt, J.; Ries, J.; Garcia-Saez, A.J. Bax assembly into rings and arcs in apoptotic mitochondria is linked to membrane pores. EMBO J. 2016, 35, 389–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kantrow, S.P.; Piantadosi, C.A. Release of cytochrome c from liver mitochondria during permeability transition. Biochem. Biophys. Res. Commun. 1997, 232, 669–671. [Google Scholar] [CrossRef]
  5. Galluzzi, L.; Vitale, I.; Aaronson, S.A.; Abrams, J.M.; Adam, D.; Agostinis, P.; Alnemri, E.S.; Altucci, L.; Amelio, I.; Andrews, D.W.; et al. Molecular mechanisms of cell death: Recommendations of the Nomenclature Committee on Cell Death 2018. Cell Death Differ. 2018, 25, 486–541. [Google Scholar] [CrossRef] [PubMed]
  6. Bonora, M.; Pinton, P. The mitochondrial permeability transition pore and cancer: Molecular mechanisms involved in cell death. Front. Oncol. 2014, 4, 302. [Google Scholar] [CrossRef] [Green Version]
  7. Giorgi, C.; Danese, A.; Missiroli, S.; Patergnani, S.; Pinton, P. Calcium Dynamics as a Machine for Decoding Signals. Trends Cell Biol. 2018, 28, 258–273. [Google Scholar] [CrossRef] [PubMed]
  8. Giorgi, C.; Marchi, S.; Pinton, P. The machineries, regulation and cellular functions of mitochondrial calcium. Nat. Rev. Mol. Cell Biol. 2018, 19, 713–730. [Google Scholar] [CrossRef]
  9. Gunter, T.E.; Pfeiffer, D.R. Mechanisms by which mitochondria transport calcium. Am. J. Physiol. 1990, 258, C755–C786. [Google Scholar] [CrossRef]
  10. Kroemer, G.; Galluzzi, L.; Brenner, C. Mitochondrial membrane permeabilization in cell death. Physiol. Rev. 2007, 87, 99–163. [Google Scholar] [CrossRef]
  11. Bonora, M.; Wieckowski, M.R.; Chinopoulos, C.; Kepp, O.; Kroemer, G.; Galluzzi, L.; Pinton, P. Molecular mechanisms of cell death: Central implication of ATP synthase in mitochondrial permeability transition. Oncogene 2015, 34, 1475–1486. [Google Scholar] [CrossRef] [PubMed]
  12. Schroers, A.; Kramer, R.; Wohlrab, H. The reversible antiport-uniport conversion of the phosphate carrier from yeast mitochondria depends on the presence of a single cysteine. J. Biol. Chem. 1997, 272, 10558–10564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Petronilli, V.; Cola, C.; Bernardi, P. Modulation of the mitochondrial cyclosporin A-sensitive permeability transition pore. II. The minimal requirements for pore induction underscore a key role for transmembrane electrical potential, matrix pH, and matrix Ca2+. J. Biol. Chem. 1993, 268, 1011–1016. [Google Scholar] [PubMed]
  14. Takeyama, N.; Matsuo, N.; Tanaka, T. Oxidative damage to mitochondria is mediated by the Ca(2+)-dependent inner-membrane permeability transition. Biochem. J. 1993, 294, 719–725. [Google Scholar] [CrossRef] [Green Version]
  15. Wieckowski, M.R.; Brdiczka, D.; Wojtczak, L. Long-chain fatty acids promote opening of the reconstituted mitochondrial permeability transition pore. FEBS Lett. 2000, 484, 61–64. [Google Scholar] [CrossRef] [Green Version]
  16. Brenner, C.; Grimm, S. The permeability transition pore complex in cancer cell death. Oncogene 2006, 25, 4744–4756. [Google Scholar] [CrossRef] [Green Version]
  17. Kowaltowski, A.J.; Castilho, R.F.; Vercesi, A.E. 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, 150–152. [Google Scholar] [CrossRef] [Green Version]
  18. 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. [Google Scholar] [CrossRef] [Green Version]
  19. 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. [Google Scholar] [CrossRef]
  20. Karch, J.; Bround, M.J.; Khalil, H.; Sargent, M.A.; Latchman, N.; Terada, N.; Peixoto, P.M.; Molkentin, J.D. Inhibition of mitochondrial permeability transition by deletion of the ANT family and CypD. Sci. Adv. 2019, 5, eaaw4597. [Google Scholar] [CrossRef] [Green Version]
  21. Bonora, M.; Morganti, C.; Morciano, G.; Pedriali, G.; Lebiedzinska-Arciszewska, M.; Aquila, G.; Giorgi, C.; Rizzo, P.; Campo, G.; Ferrari, R.; et al. Mitochondrial permeability transition involves dissociation of F1FO ATP synthase dimers and C-ring conformation. EMBO Rep. 2017, 18, 1077–1089. [Google Scholar] [CrossRef] [PubMed]
  22. Morciano, G.; Preti, D.; Pedriali, G.; Aquila, G.; Missiroli, S.; Fantinati, A.; Caroccia, N.; Pacifico, S.; Bonora, M.; Talarico, A.; et al. Discovery of Novel 1,3,8-Triazaspiro[4.5]decane Derivatives That Target the c Subunit of F1/FO-Adenosine Triphosphate (ATP) Synthase for the Treatment of Reperfusion Damage in Myocardial Infarction. J. Med. Chem. 2018, 61, 7131–7143. [Google Scholar] [CrossRef] [PubMed]
  23. Giorgio, V.; von Stockum, S.; Antoniel, M.; Fabbro, A.; Fogolari, F.; Forte, M.; Glick, G.D.; Petronilli, V.; Zoratti, M.; Szabo, I.; et al. Dimers of mitochondrial ATP synthase form the permeability transition pore. Proc. Natl. Acad. Sci. USA 2013, 110, 5887–5892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Giorgio, V.; Burchell, V.; Schiavone, M.; Bassot, C.; Minervini, G.; Petronilli, V.; Argenton, F.; Forte, M.; Tosatto, S.; Lippe, G.; et al. Ca(2+) binding to F-ATP synthase beta subunit triggers the mitochondrial permeability transition. EMBO Rep. 2017, 18, 1065–1076. [Google Scholar] [CrossRef]
  25. Antoniel, M.; Jones, K.; Antonucci, S.; Spolaore, B.; Fogolari, F.; Petronilli, V.; Giorgio, V.; Carraro, M.; Di Lisa, F.; Forte, M.; et al. The unique histidine in OSCP subunit of F-ATP synthase mediates inhibition of the permeability transition pore by acidic pH. EMBO Rep. 2018, 19, 257–268. [Google Scholar] [CrossRef]
  26. Mnatsakanyan, N.; Llaguno, M.C.; Yang, Y.; Yan, Y.; Weber, J.; Sigworth, F.J.; Jonas, E.A. A mitochondrial megachannel resides in monomeric F1FO ATP synthase. Nat. Commun. 2019, 10, 5823. [Google Scholar] [CrossRef]
  27. Urbani, A.; Giorgio, V.; Carrer, A.; Franchin, C.; Arrigoni, G.; Jiko, C.; Abe, K.; Maeda, S.; Shinzawa-Itoh, K.; Bogers, J.F.M.; et al. Purified F-ATP synthase forms a Ca(2+)-dependent high-conductance channel matching the mitochondrial permeability transition pore. Nat. Commun. 2019, 10, 4341. [Google Scholar] [CrossRef] [Green Version]
  28. He, J.; Carroll, J.; Ding, S.; Fearnley, I.M.; Walker, J.E. Permeability transition in human mitochondria persists in the absence of peripheral stalk subunits of ATP synthase. Proc. Natl. Acad. Sci. USA 2017, 114, 9086–9091. [Google Scholar] [CrossRef] [Green Version]
  29. Carroll, J.; He, J.; Ding, S.; Fearnley, I.M.; Walker, J.E. Persistence of the permeability transition pore in human mitochondria devoid of an assembled ATP synthase. Proc. Natl. Acad. Sci. USA 2019, 116, 12816–12821. [Google Scholar] [CrossRef] [Green Version]
  30. Neginskaya, M.A.; Solesio, M.E.; Berezhnaya, E.V.; Amodeo, G.F.; Mnatsakanyan, N.; Jonas, E.A.; Pavlov, E.V. ATP Synthase C-Subunit-Deficient Mitochondria Have a Small Cyclosporine A-Sensitive Channel, but Lack the Permeability Transition Pore. Cell Rep. 2019, 26, 11–17. [Google Scholar] [CrossRef] [Green Version]
  31. Baines, C.P.; Kaiser, R.A.; Purcell, N.H.; Blair, N.S.; Osinska, H.; Hambleton, M.A.; Brunskill, E.W.; Sayen, M.R.; Gottlieb, R.A.; Dorn, G.W.; et al. Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature 2005, 434, 658–662. [Google Scholar] [CrossRef] [PubMed]
  32. Nakagawa, T.; Shimizu, S.; Watanabe, T.; Yamaguchi, O.; Otsu, K.; Yamagata, H.; Inohara, H.; Kubo, T.; Tsujimoto, Y. Cyclophilin D-dependent mitochondrial permeability transition regulates some necrotic but not apoptotic cell death. Nature 2005, 434, 652–658. [Google Scholar] [CrossRef] [PubMed]
  33. Li, Y.; Johnson, N.; Capano, M.; Edwards, M.; Crompton, M. Cyclophilin-D promotes the mitochondrial permeability transition but has opposite effects on apoptosis and necrosis. Biochem. J. 2004, 383, 101–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Zhang, J.; Han, Y.; Shi, H.; Chen, J.; Zhang, X.; Wang, X.; Zhou, L.; Liu, J.; Zhang, J.; Ji, Z.; et al. Swine acute diarrhea syndrome coronavirus-induced apoptosis is caspase- and cyclophilin D- dependent. Emerg. Microbes Infect. 2020, 9, 439–456. [Google Scholar] [CrossRef]
  35. Garcia-Alvarado, F.; Govoni, G.; de Pascual, R.; Ruiz-Ruiz, C.; Munoz-Montero, A.; Gandia, L.; de Diego, A.M.G.; Garcia, A.G. Otilonium and pinaverium trigger mitochondrial-mediated apoptosis in rat embryo cortical neurons in vitro. Neurotoxicology 2019, 70, 99–111. [Google Scholar] [CrossRef]
  36. Lee, Y.J.; Lee, C. Porcine deltacoronavirus induces caspase-dependent apoptosis through activation of the cytochrome c-mediated intrinsic mitochondrial pathway. Virus Res. 2018, 253, 112–123. [Google Scholar] [CrossRef]
  37. Wang, H.; Chen, Y.; Zhai, N.; Chen, X.; Gan, F.; Li, H.; Huang, K. Ochratoxin A-Induced Apoptosis of IPEC-J2 Cells through ROS-Mediated Mitochondrial Permeability Transition Pore Opening Pathway. J. Agric. Food. Chem. 2017, 65, 10630–10637. [Google Scholar] [CrossRef]
  38. Naserzadeh, P.; Ansari Esfeh, F.; Kaviani, M.; Ashtari, K.; Kheirbakhsh, R.; Salimi, A.; Pourahmad, J. Single-walled carbon nanotube, multi-walled carbon nanotube and Fe2O3 nanoparticles induced mitochondria mediated apoptosis in melanoma cells. Cutan. Ocul. Toxicol. 2018, 37, 157–166. [Google Scholar] [CrossRef]
  39. Xie, Z.; Wang, J.; Liu, M.; Chen, D.; Qiu, C.; Sun, K. CC-223 blocks mTORC1/C2 activation and inhibits human hepatocellular carcinoma cells in vitro and in vivo. PLoS ONE 2017, 12, e0173252. [Google Scholar] [CrossRef] [Green Version]
  40. Yang, M.; Wang, B.; Gao, J.; Zhang, Y.; Xu, W.; Tao, L. Spinosad induces programmed cell death involves mitochondrial dysfunction and cytochrome C release in Spodoptera frugiperda Sf9 cells. Chemosphere 2017, 169, 155–161. [Google Scholar] [CrossRef]
  41. Liu, G.; Zou, H.; Luo, T.; Long, M.; Bian, J.; Liu, X.; Gu, J.; Yuan, Y.; Song, R.; Wang, Y.; et al. Caspase-Dependent and Caspase-Independent Pathways Are Involved in Cadmium-Induced Apoptosis in Primary Rat Proximal Tubular Cell Culture. PLoS ONE 2016, 11, e0166823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Ottani, F.; Latini, R.; Staszewsky, L.; La Vecchia, L.; Locuratolo, N.; Sicuro, M.; Masson, S.; Barlera, S.; Milani, V.; Lombardi, M.; et al. Cyclosporine A in Reperfused Myocardial Infarction: The Multicenter, Controlled, Open-Label CYCLE Trial. J. Am. Coll. Cardiol. 2016, 67, 365–374. [Google Scholar] [CrossRef] [Green Version]
  43. Fakharnia, F.; Khodagholi, F.; Dargahi, L.; Ahmadiani, A. Prevention of Cyclophilin D-Mediated mPTP Opening Using Cyclosporine-A Alleviates the Elevation of Necroptosis, Autophagy and Apoptosis-Related Markers Following Global Cerebral Ischemia-Reperfusion. J. Mol. Neurosci. 2017, 61, 52–60. [Google Scholar] [CrossRef] [PubMed]
  44. Ou, Z.; Jiang, T.; Gao, Q.; Tian, Y.Y.; Zhou, J.S.; Wu, L.; Shi, J.Q.; Zhang, Y.D. Mitochondrial-dependent mechanisms are involved in angiotensin II-induced apoptosis in dopaminergic neurons. J. Renin. Angiotensin Aldosterone Syst. 2016, 17, 1470320316672349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Dong, Y.Y.; Zhuang, Y.H.; Cai, W.J.; Liu, Y.; Zou, W.B. The mitochondrion interfering compound NPC-26 exerts potent anti-pancreatic cancer cell activity in vitro and in vivo. Tumour Biol. 2016, 37, 15053–15063. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, Y.; Li, M.; Li, Z.; Gao, P.; Zhou, X.; Zhang, J. Bufalin induces apoptosis in the U2OS human osteosarcoma cell line via triggering the mitochondrial pathway. Mol. Med. Rep. 2016, 13, 817–822. [Google Scholar] [CrossRef] [Green Version]
  47. Borutaite, V.; Jekabsone, A.; Morkuniene, R.; Brown, G.C. Inhibition of mitochondrial permeability transition prevents mitochondrial dysfunction, cytochrome c release and apoptosis induced by heart ischemia. J. Mol. Cell. Cardiol. 2003, 35, 357–366. [Google Scholar] [CrossRef]
  48. Precht, T.A.; Phelps, R.A.; Linseman, D.A.; Butts, B.D.; Le, S.S.; Laessig, T.A.; Bouchard, R.J.; Heidenreich, K.A. The permeability transition pore triggers Bax translocation to mitochondria during neuronal apoptosis. Cell Death Differ. 2005, 12, 255–265. [Google Scholar] [CrossRef]
  49. Karch, J.; Kwong, J.Q.; Burr, A.R.; Sargent, M.A.; Elrod, J.W.; Peixoto, P.M.; Martinez-Caballero, S.; Osinska, H.; Cheng, E.H.; Robbins, J.; et al. 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. [Google Scholar] [CrossRef]
  50. Narita, M.; Shimizu, S.; Ito, T.; Chittenden, T.; Lutz, R.J.; Matsuda, H.; Tsujimoto, Y. Bax interacts with the permeability transition pore to induce permeability transition and cytochrome c release in isolated mitochondria. Proc. Natl. Acad. Sci. USA 1998, 95, 14681–14686. [Google Scholar] [CrossRef] [Green Version]
  51. Scorrano, L.; Ashiya, M.; Buttle, K.; Weiler, S.; Oakes, S.A.; Mannella, C.A.; Korsmeyer, S.J. A distinct pathway remodels mitochondrial cristae and mobilizes cytochrome c during apoptosis. Dev. Cell 2002, 2, 55–67. [Google Scholar] [CrossRef] [Green Version]
  52. Vander Heiden, M.G.; Chandel, N.S.; Schumacker, P.T.; Thompson, C.B. Bcl-xL prevents cell death following growth factor withdrawal by facilitating mitochondrial ATP/ADP exchange. Mol. Cell 1999, 3, 159–167. [Google Scholar] [CrossRef]
  53. Belzacq, A.S.; Vieira, H.L.; Verrier, F.; Vandecasteele, G.; Cohen, I.; Prevost, M.C.; Larquet, E.; Pariselli, F.; Petit, P.X.; Kahn, A.; et al. Bcl-2 and Bax modulate adenine nucleotide translocase activity. Cancer Res. 2003, 63, 541–546. [Google Scholar] [PubMed]
  54. Degterev, A.; Hitomi, J.; Germscheid, M.; Ch’en, I.L.; Korkina, O.; Teng, X.; Abbott, D.; Cuny, G.D.; Yuan, C.; Wagner, G.; et al. Identification of RIP1 kinase as a specific cellular target of necrostatins. Nat. Chem. Biol. 2008, 4, 313–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Sun, X.; Lee, J.; Navas, T.; Baldwin, D.T.; Stewart, T.A.; Dixit, V.M. RIP3, a novel apoptosis-inducing kinase. J. Biol. Chem. 1999, 274, 16871–16875. [Google Scholar] [CrossRef] [Green Version]
  56. Degterev, A.; Huang, Z.; Boyce, M.; Li, Y.; Jagtap, P.; Mizushima, N.; Cuny, G.D.; Mitchison, T.J.; Moskowitz, M.A.; Yuan, J. Chemical inhibitor of nonapoptotic cell death with therapeutic potential for ischemic brain injury. Nat. Chem. Biol. 2005, 1, 112–119. [Google Scholar] [CrossRef]
  57. Kokoszka, J.E.; Waymire, K.G.; Levy, S.E.; Sligh, J.E.; Cai, J.; Jones, D.P.; MacGregor, G.R.; Wallace, D.C. The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore. Nature 2004, 427, 461–465. [Google Scholar] [CrossRef]
  58. Temkin, V.; Huang, Q.; Liu, H.; Osada, H.; Pope, R.M. Inhibition of ADP/ATP exchange in receptor-interacting protein-mediated necrosis. Mol. Cell. Biol. 2006, 26, 2215–2225. [Google Scholar] [CrossRef] [Green Version]
  59. Kroemer, G.; Galluzzi, L.; Vandenabeele, P.; Abrams, J.; Alnemri, E.S.; Baehrecke, E.H.; Blagosklonny, M.V.; El-Deiry, W.S.; Golstein, P.; Green, D.R.; et al. Classification of cell death: Recommendations of the Nomenclature Committee on Cell Death 2009. Cell Death Differ. 2009, 16, 3–11. [Google Scholar] [CrossRef]
  60. Irrinki, K.M.; Mallilankaraman, K.; Thapa, R.J.; Chandramoorthy, H.C.; Smith, F.J.; Jog, N.R.; Gandhirajan, R.K.; Kelsen, S.G.; Houser, S.R.; May, M.J.; et al. Requirement of FADD, NEMO, and BAX/BAK for aberrant mitochondrial function in tumor necrosis factor alpha-induced necrosis. Mol. Cell. Biol. 2011, 31, 3745–3758. [Google Scholar] [CrossRef] [Green Version]
  61. Tischner, D.; Manzl, C.; Soratroi, C.; Villunger, A.; Krumschnabel, G. Necrosis-like death can engage multiple pro-apoptotic Bcl-2 protein family members. Apoptosis 2012, 17, 1197–1209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Weiner, H. A medicine of human relationships. Pharos. Alpha Omega Alpha Honor Med. Soc. 1989, 52, 2–6. [Google Scholar] [PubMed]
  63. Krawczak, M.; Bockel, B. The formal analysis of multilocus DNA fingerprints. EXS 1993, 67, 249–255. [Google Scholar] [PubMed]
  64. Bonora, M.; Wieckowski, M.R.; Sinclair, D.A.; Kroemer, G.; Pinton, P.; Galluzzi, L. Targeting mitochondria for cardiovascular disorders: Therapeutic potential and obstacles. Nat. Rev. Cardiol. 2019, 16, 33–55. [Google Scholar] [CrossRef] [PubMed]
  65. Wei, C.; Li, H.; Wang, Y.; Peng, X.; Shao, H.; Li, H.; Bai, S.; Xu, C. Exogenous spermine inhibits hypoxia/ischemia-induced myocardial apoptosis via regulation of mitochondrial permeability transition pore and associated pathways. Exp. Biol. Med. (Maywood) 2016, 241, 1505–1515. [Google Scholar] [CrossRef] [Green Version]
  66. Zahrebelski, G.; Nieminen, A.L.; al-Ghoul, K.; Qian, T.; Herman, B.; Lemasters, J.J. Progression of subcellular changes during chemical hypoxia to cultured rat hepatocytes: A laser scanning confocal microscopic study. Hepatology 1995, 21, 1361–1372. [Google Scholar]
  67. Fang, Y.D.; Xu, X.; Dang, Y.M.; Zhang, Y.M.; Zhang, J.P.; Hu, J.Y.; Zhang, Q.; Dai, X.; Teng, M.; Zhang, D.X.; et al. MAP4 mechanism that stabilizes mitochondrial permeability transition in hypoxia: Microtubule enhancement and DYNLT1 interaction with VDAC1. PLoS ONE 2011, 6, e28052. [Google Scholar] [CrossRef]
  68. Soares, R.O.S.; Losada, D.M.; Jordani, M.C.; Evora, P.; Castro, E.S.O. Ischemia/Reperfusion Injury Revisited: An Overview of the Latest Pharmacological Strategies. Int. J. Mol. Sci. 2019, 20, 5034. [Google Scholar] [CrossRef] [Green Version]
  69. Liu, R.R.; Murphy, T.H. Reversible cyclosporin A-sensitive mitochondrial depolarization occurs within minutes of stroke onset in mouse somatosensory cortex in vivo: A two-photon imaging study. J. Biol. Chem. 2009, 284, 36109–36117. [Google Scholar] [CrossRef] [Green Version]
  70. Adachi, M.; Takahashi, K.; Nishikawa, M.; Miki, H.; Uyama, M. High intraocular pressure-induced ischemia and reperfusion injury in the optic nerve and retina in rats. Graefes Arch. Clin. Exp. Ophthalmol. 1996, 234, 445–451. [Google Scholar] [CrossRef]
  71. Hwang, J.H.; Lee, J.H.; Lee, K.H.; Bae, E.J.; Sung, D.K.; Chang, Y.S.; Park, W.S. Cyclosporine A attenuates hypoxic-ischemic brain injury in newborn rats. Brain Res. 2010, 1359, 208–215. [Google Scholar] [CrossRef] [PubMed]
  72. Ong, S.G.; Lee, W.H.; Theodorou, L.; Kodo, K.; Lim, S.Y.; Shukla, D.H.; Briston, T.; Kiriakidis, S.; Ashcroft, M.; Davidson, S.M.; et al. HIF-1 reduces ischaemia-reperfusion injury in the heart by targeting the mitochondrial permeability transition pore. Cardiovasc. Res. 2014, 104, 24–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Chatterjee, P.K.; Brown, P.A.; Cuzzocrea, S.; Zacharowski, K.; Stewart, K.N.; Mota-Filipe, H.; McDonald, M.C.; Thiemermann, C. Calpain inhibitor-1 reduces renal ischemia/reperfusion injury in the rat. Kidney Int. 2001, 59, 2073–2083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kalogeris, T.; Baines, C.P.; Krenz, M.; Korthuis, R.J. Ischemia/Reperfusion. Compr. Physiol. 2016, 7, 113–170. [Google Scholar] [PubMed] [Green Version]
  75. Shintani-Ishida, K.; Yoshida, K. Mitochondrial m-calpain opens the mitochondrial permeability transition pore in ischemia-reperfusion. Int. J. Cardiol. 2015, 197, 26–32. [Google Scholar] [CrossRef] [PubMed]
  76. Piper, H.M.; Meuter, K.; Schafer, C. Cellular mechanisms of ischemia-reperfusion injury. Ann. Thorac. Surg. 2003, 75, S644–S648. [Google Scholar] [CrossRef]
  77. Granger, D.N.; Kvietys, P.R. Reperfusion injury and reactive oxygen species: The evolution of a concept. Redox Biol. 2015, 6, 524–551. [Google Scholar] [CrossRef] [Green Version]
  78. Chouchani, E.T.; Pell, V.R.; Gaude, E.; Aksentijevic, D.; Sundier, S.Y.; Robb, E.L.; Logan, A.; Nadtochiy, S.M.; Ord, E.N.J.; Smith, A.C.; et al. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature 2014, 515, 431–435. [Google Scholar] [CrossRef] [Green Version]
  79. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef] [Green Version]
  80. Chen, Y.R.; Zweier, J.L. Cardiac mitochondria and reactive oxygen species generation. Circ. Res. 2014, 114, 524–537. [Google Scholar] [CrossRef] [Green Version]
  81. Barzyc, A.; Lysik, W.; Slyk, J.; Kuszewski, M.; Zarebinski, M.; Wojciechowska, M.; Cudnoch-Jedrzejewska, A. Reperfusion injury as a target for diminishing infarct size. Med. Hypotheses 2020, 137, 109558. [Google Scholar] [CrossRef] [PubMed]
  82. Griffiths, E.J.; Halestrap, A.P. Mitochondrial non-specific pores remain closed during cardiac ischaemia, but open upon reperfusion. Biochem. J. 1995, 307, 93–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Morciano, G.; Bonora, M.; Campo, G.; Aquila, G.; Rizzo, P.; Giorgi, C.; Wieckowski, M.R.; Pinton, P. Mechanistic Role of mPTP in Ischemia-Reperfusion Injury. Adv. Exp. Med. Biol. 2017, 982, 169–189. [Google Scholar] [PubMed]
  84. Upadhaya, S.; Madala, S.; Baniya, R.; Subedi, S.K.; Saginala, K.; Bachuwa, G. Impact of cyclosporine A use in the prevention of reperfusion injury in acute myocardial infarction: A meta-analysis. Cardiol. J. 2017, 24, 43–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Pottecher, J.; Kindo, M.; Chamaraux-Tran, T.N.; Charles, A.L.; Lejay, A.; Kemmel, V.; Vogel, T.; Chakfe, N.; Zoll, J.; Diemunsch, P.; et al. Skeletal muscle ischemia-reperfusion injury and cyclosporine A in the aging rat. Fundam. Clin. Pharmacol. 2016, 30, 216–225. [Google Scholar] [CrossRef] [PubMed]
  86. Uchino, H.; Minamikawa-Tachino, R.; Kristian, T.; Perkins, G.; Narazaki, M.; Siesjo, B.K.; Shibasaki, F. Differential neuroprotection by cyclosporin A and FK506 following ischemia corresponds with differing abilities to inhibit calcineurin and the mitochondrial permeability transition. Neurobiol. Dis. 2002, 10, 219–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Konukoglu, D.; Tasci, I.; Cetinkale, O. Effects of cyclosporin A and ibuprofen on liver ischemia-reperfusion injury in the rat. Clin. Chim. Acta 1998, 275, 1–8. [Google Scholar] [CrossRef]
  88. Li, J.; Yan, Z.; Fang, Q. A Mechanism Study Underlying the Protective Effects of Cyclosporine-A on Lung Ischemia-Reperfusion Injury. Pharmacology 2017, 100, 83–90. [Google Scholar] [CrossRef]
  89. Yazdani, I.; Majdani, R.; Ghasemnejad-Berenji, M.; Dehpour, A.R. Comparison of multiple doses of cyclosporine A on germ cell apoptosis and epididymal sperm parameters after testicular ischemia/reperfusion in rats. Exp. Mol. Pathol. 2019, 110, 104271. [Google Scholar] [CrossRef]
  90. Ruiz-Meana, M.; Inserte, J.; Fernandez-Sanz, C.; Hernando, V.; Miro-Casas, E.; Barba, I.; Garcia-Dorado, D. The role of mitochondrial permeability transition in reperfusion-induced cardiomyocyte death depends on the duration of ischemia. Basic Res. Cardiol. 2011, 106, 1259–1268. [Google Scholar] [CrossRef]
  91. Imahashi, K.; Pott, C.; Goldhaber, J.I.; Steenbergen, C.; Philipson, K.D.; Murphy, E. Cardiac-specific ablation of the Na+-Ca2+ exchanger confers protection against ischemia/reperfusion injury. Circ. Res. 2005, 97, 916–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Imahashi, K.; Schneider, M.D.; Steenbergen, C.; Murphy, E. Transgenic expression of Bcl-2 modulates energy metabolism, prevents cytosolic acidification during ischemia, and reduces ischemia/reperfusion injury. Circ. Res. 2004, 95, 734–741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Luongo, T.S.; Lambert, J.P.; Gross, P.; Nwokedi, M.; Lombardi, A.A.; Shanmughapriya, S.; Carpenter, A.C.; Kolmetzky, D.; Gao, E.; van Berlo, J.H.; et al. The mitochondrial Na+/Ca2+ exchanger is essential for Ca2+ homeostasis and viability. Nature 2017, 545, 93–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhao, Q.; Wang, S.; Li, Y.; Wang, P.; Li, S.; Guo, Y.; Yao, R. The role of the mitochondrial calcium uniporter in cerebral ischemia/reperfusion injury in rats involves regulation of mitochondrial energy metabolism. Mol. Med. Rep. 2013, 7, 1073–1080. [Google Scholar] [CrossRef] [Green Version]
  95. Luongo, T.S.; Lambert, J.P.; Yuan, A.; Zhang, X.; Gross, P.; Song, J.; Shanmughapriya, S.; Gao, E.; Jain, M.; Houser, S.R.; et al. The Mitochondrial Calcium Uniporter Matches Energetic Supply with Cardiac Workload during Stress and Modulates Permeability Transition. Cell Rep. 2015, 12, 23–34. [Google Scholar] [CrossRef] [Green Version]
  96. Campo, G.; Morciano, G.; Pavasini, R.; Bonora, M.; Sbano, L.; Biscaglia, S.; Bovolenta, M.; Pinotti, M.; Punzetti, S.; Rizzo, P.; et al. Fo ATP synthase C subunit serum levels in patients with ST-segment Elevation Myocardial Infarction: Preliminary findings. Int. J. Cardiol. 2016, 221, 993–997. [Google Scholar] [CrossRef]
  97. Eirin, A.; Li, Z.; Zhang, X.; Krier, J.D.; Woollard, J.R.; Zhu, X.Y.; Tang, H.; Herrmann, S.M.; Lerman, A.; Textor, S.C.; et al. A mitochondrial permeability transition pore inhibitor improves renal outcomes after revascularization in experimental atherosclerotic renal artery stenosis. Hypertension 2012, 60, 1242–1249. [Google Scholar] [CrossRef] [Green Version]
  98. Tanno, M.; Kuno, A.; Ishikawa, S.; Miki, T.; Kouzu, H.; Yano, T.; Murase, H.; Tobisawa, T.; Ogasawara, M.; Horio, Y.; et al. Translocation of glycogen synthase kinase-3beta (GSK-3beta), a trigger of permeability transition, is kinase activity-dependent and mediated by interaction with voltage-dependent anion channel 2 (VDAC2). J. Biol. Chem. 2014, 289, 29285–29296. [Google Scholar] [CrossRef] [Green Version]
  99. Bao, H.; Ge, Y.; Zhuang, S.; Dworkin, L.D.; Liu, Z.; Gong, R. Inhibition of glycogen synthase kinase-3beta prevents NSAID-induced acute kidney injury. Kidney Int. 2012, 81, 662–673. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, Z.; Ge, Y.; Bao, H.; Dworkin, L.; Peng, A.; Gong, R. Redox-sensitive glycogen synthase kinase 3beta-directed control of mitochondrial permeability transition: Rheostatic regulation of acute kidney injury. Free Radic. Biol. Med. 2013, 65, 849–858. [Google Scholar] [CrossRef] [Green Version]
  101. Linkermann, A.; De Zen, F.; Weinberg, J.; Kunzendorf, U.; Krautwald, S. Programmed necrosis in acute kidney injury. Nephrol. Dial. Transplant. Off. Publ. Eur. Dial. Transpl. Assoc. Eur. Ren. Assoc. 2012, 27, 3412–3419. [Google Scholar] [CrossRef] [PubMed]
  102. Linkermann, A.; Brasen, J.H.; Darding, M.; Jin, M.K.; Sanz, A.B.; Heller, J.O.; De Zen, F.; Weinlich, R.; Ortiz, A.; Walczak, H.; et al. Two independent pathways of regulated necrosis mediate ischemia-reperfusion injury. Proc. Natl. Acad. Sci. USA 2013, 110, 12024–12029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Mulay, S.R.; Honarpisheh, M.M.; Foresto-Neto, O.; Shi, C.; Desai, J.; Zhao, Z.B.; Marschner, J.A.; Popper, B.; Buhl, E.M.; Boor, P.; et al. Mitochondria Permeability Transition versus Necroptosis in Oxalate-Induced AKI. J. Am. Soc. Nephrol. 2019, 30, 1857–1869. [Google Scholar] [CrossRef] [PubMed]
  104. Du, H.; Guo, L.; Fang, F.; Chen, D.; Sosunov, A.A.; McKhann, G.M.; Yan, Y.; Wang, C.; Zhang, H.; Molkentin, J.D.; et al. Cyclophilin D deficiency attenuates mitochondrial and neuronal perturbation and ameliorates learning and memory in Alzheimer’s disease. Nat. Med. 2008, 14, 1097–1105. [Google Scholar] [CrossRef]
  105. Manczak, M.; Reddy, P.H. Abnormal interaction of VDAC1 with amyloid beta and phosphorylated tau causes mitochondrial dysfunction in Alzheimer’s disease. Hum. Mol. Genet. 2012, 21, 5131–5146. [Google Scholar] [CrossRef]
  106. Devi, L.; Raghavendran, V.; Prabhu, B.M.; Avadhani, N.G.; Anandatheerthavarada, H.K. Mitochondrial import and accumulation of alpha-synuclein impair complex I in human dopaminergic neuronal cultures and Parkinson disease brain. J. Biol. Chem. 2008, 283, 9089–9100. [Google Scholar] [CrossRef] [Green Version]
  107. Yao, J.; Du, H.; Yan, S.; Fang, F.; Wang, C.; Lue, L.F.; Guo, L.; Chen, D.; Stern, D.M.; Gunn Moore, F.J.; et al. Inhibition of amyloid-beta (Abeta) peptide-binding alcohol dehydrogenase-Abeta interaction reduces Abeta accumulation and improves mitochondrial function in a mouse model of Alzheimer’s disease. J. Neurosci. 2011, 31, 2313–2320. [Google Scholar] [CrossRef]
  108. Liu, G.; Zhang, C.; Yin, J.; Li, X.; Cheng, F.; Li, Y.; Yang, H.; Ueda, K.; Chan, P.; Yu, S. alpha-Synuclein is differentially expressed in mitochondria from different rat brain regions and dose-dependently down-regulates complex I activity. Neurosci. Lett. 2009, 454, 187–192. [Google Scholar] [CrossRef]
  109. Zhu, Y.; Duan, C.; Lu, L.; Gao, H.; Zhao, C.; Yu, S.; Ueda, K.; Chan, P.; Yang, H. alpha-Synuclein overexpression impairs mitochondrial function by associating with adenylate translocator. Int. J. Biochem. Cell Biol. 2011, 43, 732–741. [Google Scholar] [CrossRef]
  110. Tillement, L.; Lecanu, L.; Yao, W.; Greeson, J.; Papadopoulos, V. The spirostenol (22R, 25R)-20alpha-spirost-5-en-3beta-yl hexanoate blocks mitochondrial uptake of Abeta in neuronal cells and prevents Abeta-induced impairment of mitochondrial function. Steroids 2006, 71, 725–735. [Google Scholar] [CrossRef]
  111. Hou, Y.; Ghosh, P.; Wan, R.; Ouyang, X.; Cheng, H.; Mattson, M.P.; Cheng, A. Permeability transition pore-mediated mitochondrial superoxide flashes mediate an early inhibitory effect of amyloid beta1-42 on neural progenitor cell proliferation. Neurobiol. Aging 2014, 35, 975–989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Xie, H.; Guan, J.; Borrelli, L.A.; Xu, J.; Serrano-Pozo, A.; Bacskai, B.J. Mitochondrial alterations near amyloid plaques in an Alzheimer’s disease mouse model. J. Neurosci. 2013, 33, 17042–17051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Atlante, A.; Amadoro, G.; Bobba, A.; de Bari, L.; Corsetti, V.; Pappalardo, G.; Marra, E.; Calissano, P.; Passarella, S. A peptide containing residues 26–44 of tau protein impairs mitochondrial oxidative phosphorylation acting at the level of the adenine nucleotide translocator. Biochim. Biophys. Acta 2008, 1777, 1289–1300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Ludtmann, M.H.R.; Angelova, P.R.; Horrocks, M.H.; Choi, M.L.; Rodrigues, M.; Baev, A.Y.; Berezhnov, A.V.; Yao, Z.; Little, D.; Banushi, B.; et al. Alpha-synuclein oligomers interact with ATP synthase and open the permeability transition pore in Parkinson’s disease. Nat. Commun. 2018, 9, 2293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Martin, L.J.; Semenkow, S.; Hanaford, A.; Wong, M. Mitochondrial permeability transition pore regulates Parkinson’s disease development in mutant alpha-synuclein transgenic mice. Neurobiol. Aging 2014, 35, 1132–1152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Giaime, E.; Yamaguchi, H.; Gautier, C.A.; Kitada, T.; Shen, J. Loss of DJ-1 does not affect mitochondrial respiration but increases ROS production and mitochondrial permeability transition pore opening. PLoS ONE 2012, 7, e40501. [Google Scholar] [CrossRef] [Green Version]
  117. Gandhi, S.; Wood-Kaczmar, A.; Yao, Z.; Plun-Favreau, H.; Deas, E.; Klupsch, K.; Downward, J.; Latchman, D.S.; Tabrizi, S.J.; Wood, N.W.; et al. PINK1-associated Parkinson’s disease is caused by neuronal vulnerability to calcium-induced cell death. Mol. Cell 2009, 33, 627–638. [Google Scholar] [CrossRef] [Green Version]
  118. Cui, T.; Fan, C.; Gu, L.; Gao, H.; Liu, Q.; Zhang, T.; Qi, Z.; Zhao, C.; Zhao, H.; Cai, Q.; et al. Silencing of PINK1 induces mitophagy via mitochondrial permeability transition in dopaminergic MN9D cells. Brain Res. 2011, 1394, 1–13. [Google Scholar] [CrossRef]
  119. Gautier, C.A.; Giaime, E.; Caballero, E.; Nunez, L.; Song, Z.; Chan, D.; Villalobos, C.; Shen, J. Regulation of mitochondrial permeability transition pore by PINK1. Mol. Neurodegener. 2012, 7, 22. [Google Scholar] [CrossRef] [Green Version]
  120. Dagda, R.K.; Das Banerjee, T.; Janda, E. How Parkinsonian toxins dysregulate the autophagy machinery. Int. J. Mol. Sci. 2013, 14, 22163–22189. [Google Scholar] [CrossRef]
  121. Rasheed, M.Z.; Tabassum, H.; Parvez, S. Mitochondrial permeability transition pore: A promising target for the treatment of Parkinson’s disease. Protoplasma 2017, 254, 33–42. [Google Scholar] [CrossRef] [PubMed]
  122. Geisler, S.; Holmstrom, K.M.; Skujat, D.; Fiesel, F.C.; Rothfuss, O.C.; Kahle, P.J.; Springer, W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat. Cell Biol. 2010, 12, 119–131. [Google Scholar] [CrossRef] [PubMed]
  123. Hoshino, A.; Wang, W.J.; Wada, S.; McDermott-Roe, C.; Evans, C.S.; Gosis, B.; Morley, M.P.; Rathi, K.S.; Li, J.; Li, K.; et al. The ADP/ATP translocase drives mitophagy independent of nucleotide exchange. Nature 2019, 575, 375–379. [Google Scholar] [CrossRef] [PubMed]
  124. Castellazzi, M.; Patergnani, S.; Donadio, M.; Giorgi, C.; Bonora, M.; Bosi, C.; Brombo, G.; Pugliatti, M.; Seripa, D.; Zuliani, G.; et al. Autophagy and mitophagy biomarkers are reduced in sera of patients with Alzheimer’s disease and mild cognitive impairment. Sci. Rep. 2019, 9, 20009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Du, F.; Yu, Q.; Yan, S.; Hu, G.; Lue, L.F.; Walker, D.G.; Wu, L.; Yan, S.F.; Tieu, K.; Yan, S.S. PINK1 signalling rescues amyloid pathology and mitochondrial dysfunction in Alzheimer’s disease. Brain 2017, 140, 3233–3251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Fang, E.F.; Hou, Y.; Palikaras, K.; Adriaanse, B.A.; Kerr, J.S.; Yang, B.; Lautrup, S.; Hasan-Olive, M.M.; Caponio, D.; Dan, X.; et al. Mitophagy inhibits amyloid-beta and tau pathology and reverses cognitive deficits in models of Alzheimer’s disease. Nat. Neurosci. 2019, 22, 401–412. [Google Scholar] [CrossRef] [PubMed]
  127. Martin, L.J.; Gertz, B.; Pan, Y.; Price, A.C.; Molkentin, J.D.; Chang, Q. The mitochondrial permeability transition pore in motor neurons: Involvement in the pathobiology of ALS mice. Exp. Neurol. 2009, 218, 333–346. [Google Scholar] [CrossRef] [Green Version]
  128. Martin, L.J.; Fancelli, D.; Wong, M.; Niedzwiecki, M.; Ballarini, M.; Plyte, S.; Chang, Q. GNX-4728, a novel small molecule drug inhibitor of mitochondrial permeability transition, is therapeutic in a mouse model of amyotrophic lateral sclerosis. Front. Cell Neurosci. 2014, 8, 433. [Google Scholar] [CrossRef] [Green Version]
  129. Weber, J.J.; Clemensson, L.E.; Schioth, H.B.; Nguyen, H.P. Olesoxime in neurodegenerative diseases: Scrutinising a promising drug candidate. Biochem. Pharmacol. 2019, 168, 305–318. [Google Scholar] [CrossRef]
  130. Bordet, T.; Berna, P.; Abitbol, J.L.; Pruss, R.M. Olesoxime (TRO19622): A Novel Mitochondrial-Targeted Neuroprotective Compound. Pharmaceuticals. 2010, 3, 345–368. [Google Scholar] [CrossRef] [Green Version]
  131. Okada, M.; Yamashita, S.; Ueyama, H.; Ishizaki, M.; Maeda, Y.; Ando, Y. Long-term effects of edaravone on survival of patients with amyotrophic lateral sclerosis. eNeurologicalSci 2018, 11, 11–14. [Google Scholar] [CrossRef] [PubMed]
  132. Patergnani, S.; Fossati, V.; Bonora, M.; Giorgi, C.; Marchi, S.; Missiroli, S.; Rusielewicz, T.; Wieckowski, M.R.; Pinton, P. Mitochondria in Multiple Sclerosis: Molecular Mechanisms of Pathogenesis. Int. Rev. Cell Mol. Biol. 2017, 328, 49–103. [Google Scholar] [PubMed]
  133. Mahad, D.; Ziabreva, I.; Lassmann, H.; Turnbull, D. Mitochondrial defects in acute multiple sclerosis lesions. Brain 2008, 131, 1722–1735. [Google Scholar] [CrossRef] [PubMed]
  134. Inarrea, P.; Alarcia, R.; Alava, M.A.; Capablo, J.L.; Casanova, A.; Iniguez, C.; Iturralde, M.; Larrode, P.; Martin, J.; Mostacero, E.; et al. Mitochondrial complex enzyme activities and cytochrome C expression changes in multiple sclerosis. Mol. Neurobiol. 2014, 49, 1–9. [Google Scholar] [CrossRef] [PubMed]
  135. Bonora, M.; De Marchi, E.; Patergnani, S.; Suski, J.M.; Celsi, F.; Bononi, A.; Giorgi, C.; Marchi, S.; Rimessi, A.; Duszynski, J.; et al. Tumor necrosis factor-alpha impairs oligodendroglial differentiation through a mitochondria-dependent process. Cell Death Differ. 2014, 21, 1198–1208. [Google Scholar] [CrossRef] [Green Version]
  136. Tranah, G.J.; Santaniello, A.; Caillier, S.J.; D’Alfonso, S.; Martinelli Boneschi, F.; Hauser, S.L.; Oksenberg, J.R. Mitochondrial DNA sequence variation in multiple sclerosis. Neurology 2015, 85, 325–330. [Google Scholar] [CrossRef] [Green Version]
  137. Castellazzi, M.; Patergnani, S.; Donadio, M.; Giorgi, C.; Bonora, M.; Fainardi, E.; Casetta, I.; Granieri, E.; Pugliatti, M.; Pinton, P. Correlation between auto/mitophagic processes and magnetic resonance imaging activity in multiple sclerosis patients. J. Neuroinflammation 2019, 16, 131. [Google Scholar] [CrossRef]
  138. Patergnani, S.; Castellazzi, M.; Bonora, M.; Marchi, S.; Casetta, I.; Pugliatti, M.; Giorgi, C.; Granieri, E.; Pinton, P. Autophagy and mitophagy elements are increased in body fluids of multiple sclerosis-affected individuals. J. Neurol. Neurosurg. Psychiatry 2018, 89, 439–441. [Google Scholar] [CrossRef]
  139. Barcelos, I.P.; Troxell, R.M.; Graves, J.S. Mitochondrial Dysfunction and Multiple Sclerosis. Biology 2019, 8, 37. [Google Scholar] [CrossRef] [Green Version]
  140. Su, K.; Bourdette, D.; Forte, M. Mitochondrial dysfunction and neurodegeneration in multiple sclerosis. Front. Physiol. 2013, 4, 169. [Google Scholar] [CrossRef] [Green Version]
  141. Forte, M.; Gold, B.G.; Marracci, G.; Chaudhary, P.; Basso, E.; Johnsen, D.; Yu, X.; Fowlkes, J.; Rahder, M.; Stem, K.; et al. Cyclophilin D inactivation protects axons in experimental autoimmune encephalomyelitis, an animal model of multiple sclerosis. Proc. Natl. Acad. Sci. USA 2007, 104, 7558–7563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Warne, J.; Pryce, G.; Hill, J.M.; Shi, X.; Lenneras, F.; Puentes, F.; Kip, M.; Hilditch, L.; Walker, P.; Simone, M.I.; et al. Selective Inhibition of the Mitochondrial Permeability Transition Pore Protects against Neurodegeneration in Experimental Multiple Sclerosis. J. Biol. Chem. 2016, 291, 4356–4373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Pinton, P.; Rimessi, A.; Marchi, S.; Orsini, F.; Migliaccio, E.; Giorgio, M.; Contursi, C.; Minucci, S.; Mantovani, F.; Wieckowski, M.R.; et al. Protein kinase C beta and prolyl isomerase 1 regulate mitochondrial effects of the life-span determinant p66Shc. Science 2007, 315, 659–663. [Google Scholar] [CrossRef] [PubMed]
  144. Patergnani, S.; Marchi, S.; Rimessi, A.; Bonora, M.; Giorgi, C.; Mehta, K.D.; Pinton, P. PRKCB/protein kinase C, beta and the mitochondrial axis as key regulators of autophagy. Autophagy 2013, 9, 1367–1385. [Google Scholar] [CrossRef] [Green Version]
  145. Su, K.G.; Savino, C.; Marracci, G.; Chaudhary, P.; Yu, X.; Morris, B.; Galipeau, D.; Giorgio, M.; Forte, M.; Bourdette, D. Genetic inactivation of the p66 isoform of ShcA is neuroprotective in a murine model of multiple sclerosis. Eur. J. Neurosci. 2012, 35, 562–571. [Google Scholar] [CrossRef] [Green Version]
  146. Savino, C.; Pelicci, P.; Giorgio, M. The P66Shc/mitochondrial permeability transition pore pathway determines neurodegeneration. Oxidative Med. Cell. Longev. 2013, 2013, 719407. [Google Scholar] [CrossRef] [Green Version]
  147. Burr, A.R.; Molkentin, J.D. Genetic evidence in the mouse solidifies the calcium hypothesis of myofiber death in muscular dystrophy. Cell Death Differ. 2015, 22, 1402–1412. [Google Scholar] [CrossRef] [Green Version]
  148. Emery, A.E. The muscular dystrophies. Lancet 2002, 359, 687–695. [Google Scholar] [CrossRef]
  149. Tabebordbar, M.; Wang, E.T.; Wagers, A.J. Skeletal muscle degenerative diseases and strategies for therapeutic muscle repair. Annu. Rev. Pathol. 2013, 8, 441–475. [Google Scholar] [CrossRef] [Green Version]
  150. Emery, A.E. Population frequencies of inherited neuromuscular diseases--a world survey. Neuromuscul. Disord. 1991, 1, 19–29. [Google Scholar] [CrossRef]
  151. Durbeej, M.; Campbell, K.P. Muscular dystrophies involving the dystrophin-glycoprotein complex: An overview of current mouse models. Curr. Opin. Genet. Dev. 2002, 12, 349–361. [Google Scholar] [CrossRef]
  152. Ervasti, J.M.; Campbell, K.P. A role for the dystrophin-glycoprotein complex as a transmembrane linker between laminin and actin. J. Cell Biol. 1993, 122, 809–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Jung, C.; Martins, A.S.; Niggli, E.; Shirokova, N. Dystrophic cardiomyopathy: Amplification of cellular damage by Ca2+ signalling and reactive oxygen species-generating pathways. Cardiovasc. Res. 2008, 77, 766–773. [Google Scholar] [CrossRef] [PubMed]
  154. Hoffman, E.P.; Monaco, A.P.; Feener, C.C.; Kunkel, L.M. Conservation of the Duchenne muscular dystrophy gene in mice and humans. Science 1987, 238, 347–350. [Google Scholar] [CrossRef]
  155. Dubinin, M.V.; Talanov, E.Y.; Tenkov, K.S.; Starinets, V.S.; Mikheeva, I.B.; Sharapov, M.G.; Belosludtsev, K.N. Duchenne muscular dystrophy is associated with the inhibition of calcium uniport in mitochondria and an increased sensitivity of the organelles to the calcium-induced permeability transition. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165674. [Google Scholar] [CrossRef]
  156. Giacomotto, J.; Brouilly, N.; Walter, L.; Mariol, M.C.; Berger, J.; Segalat, L.; Becker, T.S.; Currie, P.D.; Gieseler, K. Chemical genetics unveils a key role of mitochondrial dynamics, cytochrome c release and IP3R activity in muscular dystrophy. Hum. Mol. Genet. 2013, 22, 4562–4578. [Google Scholar] [CrossRef] [Green Version]
  157. Stupka, N.; Gregorevic, P.; Plant, D.R.; Lynch, G.S. The calcineurin signal transduction pathway is essential for successful muscle regeneration in mdx dystrophic mice. Acta Neuropathol. 2004, 107, 299–310. [Google Scholar] [CrossRef]
  158. Chakkalakal, J.V.; Harrison, M.A.; Carbonetto, S.; Chin, E.; Michel, R.N.; Jasmin, B.J. Stimulation of calcineurin signaling attenuates the dystrophic pathology in mdx mice. Hum. Mol. Genet. 2004, 13, 379–388. [Google Scholar] [CrossRef] [Green Version]
  159. Reutenauer, J.; Dorchies, O.M.; Patthey-Vuadens, O.; Vuagniaux, G.; Ruegg, U.T. Investigation of Debio 025, a cyclophilin inhibitor, in the dystrophic mdx mouse, a model for Duchenne muscular dystrophy. Br. J. Pharmacol. 2008, 155, 574–584. [Google Scholar] [CrossRef] [Green Version]
  160. Millay, D.P.; Sargent, M.A.; Osinska, H.; Baines, C.P.; Barton, E.R.; Vuagniaux, G.; Sweeney, H.L.; Robbins, J.; Molkentin, J.D. Genetic and pharmacologic inhibition of mitochondrial-dependent necrosis attenuates muscular dystrophy. Nat. Med. 2008, 14, 442–447. [Google Scholar] [CrossRef] [Green Version]
  161. Palma, E.; Tiepolo, T.; Angelin, A.; Sabatelli, P.; Maraldi, N.M.; Basso, E.; Forte, M.A.; Bernardi, P.; Bonaldo, P. Genetic ablation of cyclophilin D rescues mitochondrial defects and prevents muscle apoptosis in collagen VI myopathic mice. Hum. Mol. Genet. 2009, 18, 2024–2031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Lampe, A.K.; Bushby, K.M. Collagen VI related muscle disorders. J. Med Genet. 2005, 42, 673–685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Merlini, L.; Martoni, E.; Grumati, P.; Sabatelli, P.; Squarzoni, S.; Urciuolo, A.; Ferlini, A.; Gualandi, F.; Bonaldo, P. Autosomal recessive myosclerosis myopathy is a collagen VI disorder. Neurology 2008, 71, 1245–1253. [Google Scholar] [CrossRef] [PubMed]
  164. Bernardi, P.; Krauskopf, A.; Basso, E.; Petronilli, V.; Blachly-Dyson, E.; Di Lisa, F.; Forte, M.A. The mitochondrial permeability transition from in vitro artifact to disease target. FEBS J. 2006, 273, 2077–2099. [Google Scholar] [CrossRef] [PubMed]
  165. Bonaldo, P.; Braghetta, P.; Zanetti, M.; Piccolo, S.; Volpin, D.; Bressan, G.M. Collagen VI deficiency induces early onset myopathy in the mouse: An animal model for Bethlem myopathy. Hum. Mol. Genet. 1998, 7, 2135–2140. [Google Scholar] [CrossRef]
  166. Irwin, W.A.; Bergamin, N.; Sabatelli, P.; Reggiani, C.; Megighian, A.; Merlini, L.; Braghetta, P.; Columbaro, M.; Volpin, D.; Bressan, G.M.; et al. Mitochondrial dysfunction and apoptosis in myopathic mice with collagen VI deficiency. Nat. Genet. 2003, 35, 367–371. [Google Scholar] [CrossRef]
  167. Angelin, A.; Tiepolo, T.; Sabatelli, P.; Grumati, P.; Bergamin, N.; Golfieri, C.; Mattioli, E.; Gualandi, F.; Ferlini, A.; Merlini, L.; et al. Mitochondrial dysfunction in the pathogenesis of Ullrich congenital muscular dystrophy and prospective therapy with cyclosporins. Proc. Natl. Acad. Sci. USA 2007, 104, 991–996. [Google Scholar] [CrossRef] [Green Version]
  168. Tagliavini, F.; Sardone, F.; Squarzoni, S.; Maraldi, N.M.; Merlini, L.; Faldini, C.; Sabatelli, P. Ultrastructural changes in muscle cells of patients with collagen VI-related myopathies. Muscles Ligaments Tendons J. 2013, 3, 281–286. [Google Scholar] [CrossRef] [Green Version]
  169. De Palma, C.; Morisi, F.; Cheli, S.; Pambianco, S.; Cappello, V.; Vezzoli, M.; Rovere-Querini, P.; Moggio, M.; Ripolone, M.; Francolini, M.; et al. Autophagy as a new therapeutic target in Duchenne muscular dystrophy. Cell Death Dis. 2012, 3, e418. [Google Scholar] [CrossRef] [Green Version]
  170. Wong, A.; Cortopassi, G. mtDNA mutations confer cellular sensitivity to oxidant stress that is partially rescued by calcium depletion and cyclosporin A. Biochem. Biophys. Res. Commun. 1997, 239, 139–145. [Google Scholar] [CrossRef]
  171. Haroon, M.F.; Fatima, A.; Scholer, S.; Gieseler, A.; Horn, T.F.; Kirches, E.; Wolf, G.; Kreutzmann, P. Minocycline, a possible neuroprotective agent in Leber’s hereditary optic neuropathy (LHON): Studies of cybrid cells bearing 11,778 mutation. Neurobiol. Dis. 2007, 28, 237–250. [Google Scholar] [CrossRef]
  172. Cotan, D.; Cordero, M.D.; Garrido-Maraver, J.; Oropesa-Avila, M.; Rodriguez-Hernandez, A.; Gomez Izquierdo, L.; De la Mata, M.; De Miguel, M.; Lorite, J.B.; Infante, E.R.; et al. Secondary coenzyme Q10 deficiency triggers mitochondria degradation by mitophagy in MELAS fibroblasts. FASEB J. 2011, 25, 2669–2687. [Google Scholar] [CrossRef] [PubMed]
  173. Cui, J.; Wang, L.; Ren, X.; Zhang, Y.; Zhang, H. LRPPRC: A Multifunctional Protein Involved in Energy Metabolism and Human Disease. Front. Physiol. 2019, 10, 595. [Google Scholar] [CrossRef] [PubMed]
  174. Burelle, Y.; Bemeur, C.; Rivard, M.E.; Thompson Legault, J.; Boucher, G.; Consortium, L.; Morin, C.; Coderre, L.; Des Rosiers, C. Mitochondrial vulnerability and increased susceptibility to nutrient-induced cytotoxicity in fibroblasts from leigh syndrome French canadian patients. PLoS ONE 2015, 10, e0120767. [Google Scholar] [CrossRef] [PubMed]
  175. MacVicar, T.; Langer, T. OPA1 processing in cell death and disease—The long and short of it. J. Cell Sci. 2016, 129, 2297–2306. [Google Scholar] [CrossRef] [Green Version]
  176. Pfeffer, G.; Gorman, G.S.; Griffin, H.; Kurzawa-Akanbi, M.; Blakely, E.L.; Wilson, I.; Sitarz, K.; Moore, D.; Murphy, J.L.; Alston, C.L.; et al. Mutations in the SPG7 gene cause chronic progressive external ophthalmoplegia through disordered mitochondrial DNA maintenance. Brain 2014, 137, 1323–1336. [Google Scholar] [CrossRef] [Green Version]
  177. Piquereau, J.; Caffin, F.; Novotova, M.; Prola, A.; Garnier, A.; Mateo, P.; Fortin, D.; Huynh le, H.; Nicolas, V.; Alavi, M.V.; et al. Down-regulation of OPA1 alters mouse mitochondrial morphology, PTP function, and cardiac adaptation to pressure overload. Cardiovasc. Res. 2012, 94, 408–417. [Google Scholar] [CrossRef] [Green Version]
  178. Shanmughapriya, S.; Rajan, S.; Hoffman, N.E.; Higgins, A.M.; Tomar, D.; Nemani, N.; Hines, K.J.; Smith, D.J.; Eguchi, A.; Vallem, S.; et al. SPG7 Is an Essential and Conserved Component of the Mitochondrial Permeability Transition Pore. Mol. Cell 2015, 60, 47–62. [Google Scholar] [CrossRef] [Green Version]
  179. Zhang, D.; Lu, C.; Whiteman, M.; Chance, B.; Armstrong, J.S. The mitochondrial permeability transition regulates cytochrome c release for apoptosis during endoplasmic reticulum stress by remodeling the cristae junction. J. Biol. Chem. 2008, 283, 3476–3486. [Google Scholar] [CrossRef] [Green Version]
  180. Hurst, S.; Baggett, A.; Csordas, G.; Sheu, S.S. SPG7 targets the m-AAA protease complex to process MCU for uniporter assembly, Ca(2+) influx, and regulation of mitochondrial permeability transition pore opening. J. Biol. Chem. 2019, 294, 10807–10818. [Google Scholar] [CrossRef]
  181. Wong, R.J.; Cheung, R.; Ahmed, A. Nonalcoholic steatohepatitis is the most rapidly growing indication for liver transplantation in patients with hepatocellular carcinoma in the U.S. Hepatology 2014, 59, 2188–2195. [Google Scholar] [CrossRef] [PubMed]
  182. Gastaldelli, A.; Kozakova, M.; Hojlund, K.; Flyvbjerg, A.; Favuzzi, A.; Mitrakou, A.; Balkau, B.; Investigators, R. Fatty liver is associated with insulin resistance, risk of coronary heart disease, and early atherosclerosis in a large European population. Hepatology 2009, 49, 1537–1544. [Google Scholar] [CrossRef] [PubMed]
  183. Teodoro, J.S.; Rolo, A.P.; Duarte, F.V.; Simoes, A.M.; Palmeira, C.M. Differential alterations in mitochondrial function induced by a choline-deficient diet: Understanding fatty liver disease progression. Mitochondrion 2008, 8, 367–376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Chen, D.F.; Wang, C.H. The relationship between the opening of mitochondrial permeability transition pores of cultured hepatocytes with their apoptoses in a non-alcoholic fatty liver disease model. Chin. J. Hepatol. 2007, 15, 837–839. [Google Scholar]
  185. Kang, M.; Li, S.; Zhong, D.; Yang, Z.; Li, P. Hepatocyte apoptosis and mitochondrial permeability transition pore opening in rats with nonalcoholic fatty liver. J. South. Med Univ. 2013, 33, 1062–1066. [Google Scholar]
  186. Wang, X.; Du, H.; Shao, S.; Bo, T.; Yu, C.; Chen, W.; Zhao, L.; Li, Q.; Wang, L.; Liu, X.; et al. Cyclophilin D deficiency attenuates mitochondrial perturbation and ameliorates hepatic steatosis. Hepatology 2018, 68, 62–77. [Google Scholar] [CrossRef]
  187. Lazarin Mde, O.; Ishii-Iwamoto, E.L.; Yamamoto, N.S.; Constantin, R.P.; Garcia, R.F.; da Costa, C.E.; Vitoriano Ade, S.; de Oliveira, M.C.; Salgueiro-Pagadigorria, C.L. Liver mitochondrial function and redox status in an experimental model of non-alcoholic fatty liver disease induced by monosodium L-glutamate in rats. Exp. Mol. Pathol. 2011, 91, 687–694. [Google Scholar] [CrossRef]
  188. Rimessi, A.; Marchi, S.; Patergnani, S.; Pinton, P. H-Ras-driven tumoral maintenance is sustained through caveolin-1-dependent alterations in calcium signaling. Oncogene 2014, 33, 2329–2340. [Google Scholar] [CrossRef]
  189. Arbel, N.; Ben-Hail, D.; Shoshan-Barmatz, V. Mediation of the antiapoptotic activity of Bcl-xL protein upon interaction with VDAC1 protein. J. Biol. Chem. 2012, 287, 23152–23161. [Google Scholar] [CrossRef] [Green Version]
  190. Alavian, K.N.; Li, H.; Collis, L.; Bonanni, L.; Zeng, L.; Sacchetti, S.; Lazrove, E.; Nabili, P.; Flaherty, B.; Graham, M.; et al. Bcl-xL regulates metabolic efficiency of neurons through interaction with the mitochondrial F1FO ATP synthase. Nat. Cell Biol. 2011, 13, 1224–1233. [Google Scholar] [CrossRef]
  191. Bittremieux, M.; Parys, J.B.; Pinton, P.; Bultynck, G. ER functions of oncogenes and tumor suppressors: Modulators of intracellular Ca(2+) signaling. Biochim. Biophys. Acta 2016, 1863, 1364–1378. [Google Scholar] [CrossRef] [PubMed]
  192. Marchi, S.; Marinello, M.; Bononi, A.; Bonora, M.; Giorgi, C.; Rimessi, A.; Pinton, P. Selective modulation of subtype III IP(3)R by Akt regulates ER Ca(2)(+) release and apoptosis. Cell Death Dis. 2012, 3, e304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Marchi, S.; Rimessi, A.; Giorgi, C.; Baldini, C.; Ferroni, L.; Rizzuto, R.; Pinton, P. Akt kinase reducing endoplasmic reticulum Ca2+ release protects cells from Ca2+-dependent apoptotic stimuli. Biochem. Biophys. Res. Commun. 2008, 375, 501–505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Papa, A.; Wan, L.; Bonora, M.; Salmena, L.; Song, M.S.; Hobbs, R.M.; Lunardi, A.; Webster, K.; Ng, C.; Newton, R.H.; et al. Cancer-associated PTEN mutants act in a dominant-negative manner to suppress PTEN protein function. Cell 2014, 157, 595–610. [Google Scholar] [CrossRef] [Green Version]
  195. Bononi, A.; Bonora, M.; Marchi, S.; Missiroli, S.; Poletti, F.; Giorgi, C.; Pandolfi, P.P.; Pinton, P. Identification of PTEN at the ER and MAMs and its regulation of Ca(2+) signaling and apoptosis in a protein phosphatase-dependent manner. Cell Death Differ. 2013, 20, 1631–1643. [Google Scholar] [CrossRef] [Green Version]
  196. Bononi, A.; Agnoletto, C.; De Marchi, E.; Marchi, S.; Patergnani, S.; Bonora, M.; Giorgi, C.; Missiroli, S.; Poletti, F.; Rimessi, A.; et al. Protein kinases and phosphatases in the control of cell fate. Enzym. Res. 2011, 2011, 329098. [Google Scholar] [CrossRef] [Green Version]
  197. Missiroli, S.; Bonora, M.; Patergnani, S.; Giorgi, C. Novel function of the tumor suppressor PML at ER-mitochondria sites in the control of autophagy. Oncotarget 2017, 8, 81723–81724. [Google Scholar] [CrossRef]
  198. Missiroli, S.; Bonora, M.; Patergnani, S.; Poletti, F.; Perrone, M.; Gafa, R.; Magri, E.; Raimondi, A.; Lanza, G.; Tacchetti, C.; et al. PML at Mitochondria-Associated Membranes Is Critical for the Repression of Autophagy and Cancer Development. Cell Rep. 2016, 16, 2415–2427. [Google Scholar] [CrossRef] [Green Version]
  199. Giorgi, C.; Ito, K.; Lin, H.K.; Santangelo, C.; Wieckowski, M.R.; Lebiedzinska, M.; Bononi, A.; Bonora, M.; Duszynski, J.; Bernardi, R.; et al. PML regulates apoptosis at endoplasmic reticulum by modulating calcium release. Science 2010, 330, 1247–1251. [Google Scholar] [CrossRef] [Green Version]
  200. Luo, J. Glycogen synthase kinase 3beta (GSK3beta) in tumorigenesis and cancer chemotherapy. Cancer Lett. 2009, 273, 194–200. [Google Scholar] [CrossRef] [Green Version]
  201. Giorgi, C.; Bonora, M.; Missiroli, S.; Morganti, C.; Morciano, G.; Wieckowski, M.R.; Pinton, P. Alterations in Mitochondrial and Endoplasmic Reticulum Signaling by p53 Mutants. Front. Oncol. 2016, 6, 42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Giorgi, C.; Bonora, M.; Sorrentino, G.; Missiroli, S.; Poletti, F.; Suski, J.M.; Galindo Ramirez, F.; Rizzuto, R.; Di Virgilio, F.; Zito, E.; et al. p53 at the endoplasmic reticulum regulates apoptosis in a Ca2+-dependent manner. Proc. Natl. Acad. Sci. USA 2015, 112, 1779–1784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Bonora, M.; Giorgi, C.; Pinton, P. Novel frontiers in calcium signaling: A possible target for chemotherapy. Pharmacol. Res. 2015, 99, 82–85. [Google Scholar] [CrossRef]
  204. Giorgi, C.; Bonora, M.; Pinton, P. Inside the tumor: p53 modulates calcium homeostasis. Cell Cycle 2015, 14, 933–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Giorgi, C.; Bonora, M.; Missiroli, S.; Poletti, F.; Ramirez, F.G.; Morciano, G.; Morganti, C.; Pandolfi, P.P.; Mammano, F.; Pinton, P. Intravital imaging reveals p53-dependent cancer cell death induced by phototherapy via calcium signaling. Oncotarget 2015, 6, 1435–1445. [Google Scholar] [CrossRef] [Green Version]
  206. 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. [Google Scholar] [CrossRef]
  207. Ghosh, J.C.; Siegelin, M.D.; Dohi, T.; Altieri, D.C. Heat shock protein 60 regulation of the mitochondrial permeability transition pore in tumor cells. Cancer Res. 2010, 70, 8988–8993. [Google Scholar] [CrossRef] [Green Version]
  208. Sinha, D.; D’Silva, P. Chaperoning mitochondrial permeability transition: Regulation of transition pore complex by a J-protein, DnaJC15. Cell Death Dis. 2014, 5, e1101. [Google Scholar] [CrossRef]
  209. Aggarwal, V.; Tuli, H.S.; Varol, A.; Thakral, F.; Yerer, M.B.; Sak, K.; Varol, M.; Jain, A.; Khan, M.A.; Sethi, G. Role of Reactive Oxygen Species in Cancer Progression: Molecular Mechanisms and Recent Advancements. Biomolecules 2019, 9, 735. [Google Scholar] [CrossRef] [Green Version]
  210. Giorgi, C.; Marchi, S.; Simoes, I.C.M.; Ren, Z.; Morciano, G.; Perrone, M.; Patalas-Krawczyk, P.; Borchard, S.; Jedrak, P.; Pierzynowska, K.; et al. Mitochondria and Reactive Oxygen Species in Aging and Age-Related Diseases. Int. Rev. Cell Mol. Biol. 2018, 340, 209–344. [Google Scholar]
  211. Fulda, S.; Galluzzi, L.; Kroemer, G. Targeting mitochondria for cancer therapy. Nat. Rev. Drug Discov. 2010, 9, 447–464. [Google Scholar] [CrossRef] [PubMed]
  212. Shinohara, Y.; Ishida, T.; Hino, M.; Yamazaki, N.; Baba, Y.; Terada, H. Characterization of porin isoforms expressed in tumor cells. Eur. J. Biochem. 2000, 267, 6067–6073. [Google Scholar] [CrossRef] [PubMed]
  213. Beinlich, A.; Strohmeier, R.; Kaufmann, M.; Kuhl, H. Relation of cell proliferation to expression of peripheral benzodiazepine receptors in human breast cancer cell lines. Biochem. Pharmacol. 2000, 60, 397–402. [Google Scholar] [CrossRef]
  214. Maaser, K.; Hopfner, M.; Jansen, A.; Weisinger, G.; Gavish, M.; Kozikowski, A.P.; Weizman, A.; Carayon, P.; Riecken, E.O.; Zeitz, M.; et al. Specific ligands of the peripheral benzodiazepine receptor induce apoptosis and cell cycle arrest in human colorectal cancer cells. Br. J. Cancer 2001, 85, 1771–1780. [Google Scholar] [CrossRef] [PubMed]
  215. Faure Vigny, H.; Heddi, A.; Giraud, S.; Chautard, D.; Stepien, G. Expression of oxidative phosphorylation genes in renal tumors and tumoral cell lines. Mol. Carcinog. 1996, 16, 165–172. [Google Scholar] [CrossRef]
  216. Rempel, A.; Mathupala, S.P.; Griffin, C.A.; Hawkins, A.L.; Pedersen, P.L. Glucose catabolism in cancer cells: Amplification of the gene encoding type II hexokinase. Cancer Res. 1996, 56, 2468–2471. [Google Scholar]
  217. Gudnason, V.; Ingvarsson, S.; Jonasdottir, A.; Andresdottir, V.; Egilsson, V. Isoenzyme pattern and subcellular localization of hexokinases in human breast cancer and nonpathological breast tissue. Int. J. Cancer 1984, 34, 63–66. [Google Scholar] [CrossRef] [Green Version]
  218. Cung, T.T.; Morel, O.; Cayla, G.; Rioufol, G.; Garcia-Dorado, D.; Angoulvant, D.; Bonnefoy-Cudraz, E.; Guerin, P.; Elbaz, M.; Delarche, N.; et al. Cyclosporine before PCI in Patients with Acute Myocardial Infarction. New Engl. J. Med. 2015, 373, 1021–1031. [Google Scholar] [CrossRef]
  219. Schaller, S.; Paradis, S.; Ngoh, G.A.; Assaly, R.; Buisson, B.; Drouot, C.; Ostuni, M.A.; Lacapere, J.J.; Bassissi, F.; Bordet, T.; et al. TRO40303, a new cardioprotective compound, inhibits mitochondrial permeability transition. J. Pharmacol. Exp. Ther. 2010, 333, 696–706. [Google Scholar] [CrossRef]
  220. Sileikyte, J.; Blachly-Dyson, E.; Sewell, R.; Carpi, A.; Menabo, R.; Di Lisa, F.; Ricchelli, F.; Bernardi, P.; Forte, M. 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. [Google Scholar] [CrossRef] [Green Version]
  221. Atar, D.; Arheden, H.; Berdeaux, A.; Bonnet, J.L.; Carlsson, M.; Clemmensen, P.; Cuvier, V.; Danchin, N.; Dubois-Rande, J.L.; Engblom, H.; et al. Effect of intravenous TRO40303 as an adjunct to primary percutaneous coronary intervention for acute ST-elevation myocardial infarction: MITOCARE study results. Eur. Heart J. 2015, 36, 112–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Leruez, S.; Verny, C.; Bonneau, D.; Procaccio, V.; Lenaers, G.; Amati-Bonneau, P.; Reynier, P.; Scherer, C.; Prundean, A.; Orssaud, C.; et al. Cyclosporine A does not prevent second-eye involvement in Leber’s hereditary optic neuropathy. Orphanet J. Rare Dis. 2018, 13, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Valasani, K.R.; Vangavaragu, J.R.; Day, V.W.; Yan, S.S. Structure based design, synthesis, pharmacophore modeling, virtual screening, and molecular docking studies for identification of novel cyclophilin D inhibitors. J. Chem. Inf. Modeling 2014, 54, 902–912. [Google Scholar] [CrossRef] [PubMed]
  224. Shore, E.R.; Awais, M.; Kershaw, N.M.; Gibson, R.R.; Pandalaneni, S.; Latawiec, D.; Wen, L.; Javed, M.A.; Criddle, D.N.; Berry, N.; et al. Small Molecule Inhibitors of Cyclophilin D To Protect Mitochondrial Function as a Potential Treatment for Acute Pancreatitis. J. Med. Chem. 2016, 59, 2596–2611. [Google Scholar] [CrossRef]
  225. Panel, M.; Ruiz, I.; Brillet, R.; Lafdil, F.; Teixeira-Clerc, F.; Nguyen, C.T.; Calderaro, J.; Gelin, M.; Allemand, F.; Guichou, J.F.; et al. Small-Molecule Inhibitors of Cyclophilins Block Opening of the Mitochondrial Permeability Transition Pore and Protect Mice From Hepatic Ischemia/Reperfusion Injury. Gastroenterology 2019, 157, 1368–1382. [Google Scholar] [CrossRef]
  226. Roy, S.; Sileikyte, J.; Schiavone, M.; Neuenswander, B.; Argenton, F.; Aube, J.; Hedrick, M.P.; Chung, T.D.; Forte, M.A.; Bernardi, P.; et al. Discovery, Synthesis, and Optimization of Diarylisoxazole-3-carboxamides as Potent Inhibitors of the Mitochondrial Permeability Transition Pore. ChemMedChem 2015, 10, 1655–1671. [Google Scholar] [CrossRef] [Green Version]
  227. Roy, S.; Sileikyte, J.; Neuenswander, B.; Hedrick, M.P.; Chung, T.D.; Aube, J.; Schoenen, F.J.; Forte, M.A.; Bernardi, P. N-Phenylbenzamides as Potent Inhibitors of the Mitochondrial Permeability Transition Pore. ChemMedChem 2016, 11, 283–288. [Google Scholar] [CrossRef] [Green Version]
  228. Tarocco, A.; Caroccia, N.; Morciano, G.; Wieckowski, M.R.; Ancora, G.; Garani, G.; Pinton, P. Melatonin as a master regulator of cell death and inflammation: Molecular mechanisms and clinical implications for newborn care. Cell Death Dis. 2019, 10, 317. [Google Scholar] [CrossRef] [Green Version]
  229. Zhou, H.; Zhang, Y.; Hu, S.; Shi, C.; Zhu, P.; Ma, Q.; Jin, Q.; Cao, F.; Tian, F.; Chen, Y. Melatonin protects cardiac microvasculature against ischemia/reperfusion injury via suppression of mitochondrial fission-VDAC1-HK2-mPTP-mitophagy axis. J. Pineal Res. 2017, 63, e12413. [Google Scholar] [CrossRef]
Figure 1. Major molecular paths in mitochondria-related regulated cell death (RCD). Mitochondrial calcium overload and ROS levels can trigger either the activation of intrinsic apoptotic pathway (left side) through the recruitment of Bcl-2 family proteins at the mitochondria, or permeability transition pore complex (PTPC) formation which could lead to mitochondrial outer membrane permeabilization (MOMP), energetic imbalance, and subsequent release of proapoptotic cofactors from the inter membrane space, such as SMAC/DIABLO, CytC, and ENDOG (right side).
Figure 1. Major molecular paths in mitochondria-related regulated cell death (RCD). Mitochondrial calcium overload and ROS levels can trigger either the activation of intrinsic apoptotic pathway (left side) through the recruitment of Bcl-2 family proteins at the mitochondria, or permeability transition pore complex (PTPC) formation which could lead to mitochondrial outer membrane permeabilization (MOMP), energetic imbalance, and subsequent release of proapoptotic cofactors from the inter membrane space, such as SMAC/DIABLO, CytC, and ENDOG (right side).
Biomolecules 10 00998 g001
Figure 2. MPT alterations in human diseases. (A) Effect of ischemia and reperfusion in levels of MPT-regulating factor in insurgence of RCD (purple cells). (B) Schematic representation of the effect of mutations in mitochondrial or nuclear DNA (represented by circular DNA or chromosome, respectively) in human diseases characterized by degeneration of neuronal or muscular tissue. (C) Representation of major alterations in MPT regulators detected in tumor microenvironment.
Figure 2. MPT alterations in human diseases. (A) Effect of ischemia and reperfusion in levels of MPT-regulating factor in insurgence of RCD (purple cells). (B) Schematic representation of the effect of mutations in mitochondrial or nuclear DNA (represented by circular DNA or chromosome, respectively) in human diseases characterized by degeneration of neuronal or muscular tissue. (C) Representation of major alterations in MPT regulators detected in tumor microenvironment.
Biomolecules 10 00998 g002

Share and Cite

MDPI and ACS Style

Bonora, M.; Patergnani, S.; Ramaccini, D.; Morciano, G.; Pedriali, G.; Kahsay, A.E.; Bouhamida, E.; Giorgi, C.; Wieckowski, M.R.; Pinton, P. Physiopathology of the Permeability Transition Pore: Molecular Mechanisms in Human Pathology. Biomolecules 2020, 10, 998. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10070998

AMA Style

Bonora M, Patergnani S, Ramaccini D, Morciano G, Pedriali G, Kahsay AE, Bouhamida E, Giorgi C, Wieckowski MR, Pinton P. Physiopathology of the Permeability Transition Pore: Molecular Mechanisms in Human Pathology. Biomolecules. 2020; 10(7):998. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10070998

Chicago/Turabian Style

Bonora, Massimo, Simone Patergnani, Daniela Ramaccini, Giampaolo Morciano, Gaia Pedriali, Asrat Endrias Kahsay, Esmaa Bouhamida, Carlotta Giorgi, Mariusz R. Wieckowski, and Paolo Pinton. 2020. "Physiopathology of the Permeability Transition Pore: Molecular Mechanisms in Human Pathology" Biomolecules 10, no. 7: 998. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10070998

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop