Next Article in Journal
Rapamycin Treatment Alleviates Chronic GVHD-Induced Lupus Nephritis in Mice by Recovering IL-2 Production and Regulatory T Cells While Inhibiting Effector T Cells Activation
Previous Article in Journal
Therapeutic Drug Monitoring in Oncohematological Patients: A Fast and Accurate HPLC-UV Method for the Quantification of Nilotinib in Human Plasma and Its Clinical Application
Previous Article in Special Issue
A Mitochondrial Perspective on Noncommunicable Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Preservation of Mitochondrial Health in Liver Ischemia/Reperfusion Injury

by
Ivo F. Machado
1,2,
Carlos M. Palmeira
1,3 and
Anabela P. Rolo
1,3,*
1
CNC—Center for Neuroscience and Cell Biology, University of Coimbra, 3000 Coimbra, Portugal
2
IIIUC—Institute of Interdisciplinary Research, University of Coimbra, 3000 Coimbra, Portugal
3
Department of Life Sciences, University of Coimbra, 3000 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(3), 948; https://doi.org/10.3390/biomedicines11030948
Submission received: 1 February 2023 / Revised: 6 March 2023 / Accepted: 16 March 2023 / Published: 20 March 2023
(This article belongs to the Special Issue A Mitochondrial Perspective on Noncommunicable Diseases)

Abstract

:
Liver ischemia-reperfusion injury (LIRI) is a major cause of the development of complications in different clinical settings such as liver resection and liver transplantation. Damage arising from LIRI is a major risk factor for early graft rejection and is associated with higher morbidity and mortality after surgery. Although the mechanisms leading to the injury of parenchymal and non-parenchymal liver cells are not yet fully understood, mitochondrial dysfunction is recognized as a hallmark of LIRI that exacerbates cellular injury. Mitochondria play a major role in glucose metabolism, energy production, reactive oxygen species (ROS) signaling, calcium homeostasis and cell death. The diverse roles of mitochondria make it essential to preserve mitochondrial health in order to maintain cellular activity and liver integrity during liver ischemia/reperfusion (I/R). A growing body of studies suggest that protecting mitochondria by regulating mitochondrial biogenesis, fission/fusion and mitophagy during liver I/R ameliorates LIRI. Targeting mitochondria in conditions that exacerbate mitochondrial dysfunction, such as steatosis and aging, has been successful in decreasing their susceptibility to LIRI. Studying mitochondrial dysfunction will help understand the underlying mechanisms of cellular damage during LIRI which is important for the development of new therapeutic strategies aimed at improving patient outcomes. In this review, we highlight the progress made in recent years regarding the role of mitochondria in liver I/R and discuss the impact of liver conditions on LIRI.

1. Introduction

Hepatic malignancies, acute liver failure and end-stage liver disease are commonly treated by liver transplantation. In recent years, the success of liver transplantation has improved, but unfortunately, the number of patients on waiting lists keeps growing [1]. Current efforts are to increase the pool of liver donors, for instance, by including marginal liver donors [2]. Liver ischemia-reperfusion (I/R) injury (LIRI) contributes to organ shortage as healthy livers and livers from marginal donors are susceptible to this type of damage [3], which can eventually result in acute and chronic rejection.
The liver has a remarkable ability to regenerate following toxic or physical damage. Upon liver injury, complex physiological and cellular events take place to fully restore the lost hepatic mass while adequate hepatic function is maintained to preserve body homeostasis. This ability allows the liver to successfully recover from resection and transplantation. Generally, LIRI is comprised by two distinct phases. The initial phase starts with interruption of hepatic circulation. During this phase (ischemia), nutrient and oxygen deprivation, pH changes, and adenosine triphosphate (ATP) depletion together lead to the aberrant formation of reactive oxygen species (ROS) and eventually cellular injury starts to set in. In the following phase, reperfusion of the liver exacerbates the damage initiated during ischemia due to metabolic disturbances and induction of a proinflammatory immune response (reviewed in [4,5]).
Mitochondria are critical players in energy and glucose metabolism, as well as in the regulation of several signaling pathways. Dysfunction of mitochondria is tightly linked with many human pathologies and aging [6,7]. Maintenance of mitochondrial health is thus required to ensure cellular and body homeostasis. Mitochondrial biogenesis, mitochondrial fission/fusion, and mitophagy are the main mechanisms that guarantee proper mitochondrial function in adaption to multiple stresses. Impaired mitochondrial function is one of the main causes for liver damage following I/R, however, the underlying mechanisms are still not completely comprehended. Their full understanding is likely to instigate the development of novel strategies capable of improving the surgical outcome of liver transplantation and resection, as well as leading to the implementation of measures that could decrease the number of patients on waiting lists. The current state of pharmacological and surgical approaches utilized to improve LIRI has been discussed in depth elsewhere and will not be reviewed here [8]. In the current review, we summarize the mitochondrial mechanisms underlying LIRI and discuss how maintenance of mitochondrial quality control contributes to the amelioration of LIRI. The impact of hepatic steatosis and liver aging on I/R injury will also be topics of discussion.

2. Cellular and Molecular Mechanisms of LIRI

The liver is an essential organ in vertebrate animals that is responsible for a complexity of functions, including biotransformation of xenobiotics, regulation of metabolites and nutrients, and maintenance of body homeostasis. Loss of hepatic function can have a dramatic effect on the organism, and it can even be fatal [9,10]. As the main parenchymal liver cells, hepatocytes are responsible for conducting most liver functions, while liver non-parenchymal cells support hepatocytes’ function. These include liver sinusoidal endothelial cells (LSECs), biliary duct cells (cholangiocytes), hepatic stellate cells (HSCs), and Kupffer cells. LIRI arises from a complex and intertwined network events taking placing after an ischemic insult that involve the interaction between the different liver cell types. In broad strokes, I/R results in hepatocyte and LSEC death due to metabolic disturbances and oxidative damage. Aggravation of the initial ischemic injury by reperfusion leads to the development of a strong inflammatory immune response due to release of damage-associated molecular patterns (DAMPs) and proinflammatory cytokines, and activation of the complement system. In turn, activation of Kupffer cells, recruitment and adhesion of neutrophils, and platelet activation sustain the inflammatory immune response which exacerbates LIRI (reviewed in [5]). LIRI compromises liver repair mechanisms and impairs liver function which may result in organ failure.
Liver cells are arranged in functional structural units, lobules, that are repeating hexagonal-shaped histological units subdivided into three separate concentrical zones [11]. Each zone has different metabolic functions. Catabolic processes, such as gluconeogenesis and β-oxidation, are prevalent in the oxygen-enriched periportal zone (zone 1) that receives arterial blood through the portal triads, while anabolic processes, such as glycolysis and lipogenesis, are prevalent in the pericentral zone (zone 3) where non-oxygenated blood that has flown through the liver sinusoids is drained into the central vein. The mid-lobule zone (zone 2) is thought to be a transitional zone with no specific function. Hepatocytes from different zones may not only have distinct metabolic functions but also contribute differently to the repopulation of liver cells during liver homeostasis and liver regeneration after injury [12]. Each liver zone is differently affected by liver I/R injury. Compared with the periportal and mid-lobule zones, the pericentral zone is more sensitive to hepatic I/R injury [13,14,15]. Pericentral hepatocytes are thought to be more vulnerable to anoxia-induced injury than periportal hepatocytes, because zone 3 has a lower oxygen concentration [11,16]. Spatial transcriptomic analysis showed that zones 1 and 3 have different profiles following I/R. While zone 3 is characterized by differentially expressed genes related to the inflammatory response, unfolded protein response, autophagy and metabolic pathways, zone 1 is mainly enriched in metabolic pathways [17]. The authors also showed that macrophages are mainly recruited to zone 3 which is associated with a higher inflammatory response following I/R. The higher presence of KCs, HSCs, and epithelial cells in zone 1 may be related to greater protection against I/R injury [17].
Hepatocytes and LSECs are the most susceptible cells to liver I/R-induced death [18,19], and thus we will mainly focus on them. But their sensitivity to warm (37 °C) and cold (4 °C) ischemia is different. While warm ischemia is observed in clinical settings such as liver transplantation, liver resection, and trauma, cold ischemia is typically found in the setting of liver transplantation, from the moment when the liver is cold stored for transport until it is implanted into the recipient. During warm ischemia, the interruption of blood flow leads to anoxia and nutrient depletion causing the decline of ATP synthesis. Consequent disturbances in intracellular metabolic processes cause acidification of the cytosol due to lactate accumulation. Once the blood flow is restored, reoxygenation leads to an increase in the production of mitochondrial ROS, cells are overloaded with calcium ions (Ca2+), and the pH of the tissue returns to physiological levels [4,20]. These events trigger the opening of the mitochondrial permeability transition (MPT) pore resulting in the loss of mitochondrial membrane integrity and subsequent hepatocyte death. Opening of the MPT pore allows molecules whose molecular mass is inferior to 1.5 kDa to diffuse freely across the inner mitochondrial membrane (IMM) [21]. In physiological conditions, the transient opening of the MPT pore is linked to mitochondrial energy metabolism, regulation of ROS signaling and mitochondrial Ca2+ signaling [22]. However, during I/R, prolonged MPT pore opening results in the dissipation of the proton motive force, uncoupling of oxidative phosphorylation, and mitochondrial swelling, which, in turn, lead to apoptotic and necrotic cell death [22,23]. In parallel, damaged parenchymal liver cells release DAMPs such as nuclear DNA, histones, heat shock proteins, ATP, mitochondrial formyl peptide, and mitochondrial DNA (mtDNA) that activate an inflammatory immune response [4,5]. The proinflammatory response is initiated due to the activation of Kupffer cells by DAMPs. Once they become activated, they produce ROS and release proinflammatory cytokines and chemokines [24]. Consequently, additional increase in ROS production foments cellular damage and apoptosis. The release of cytokines and chemokines augments the inflammatory response after reperfusion due to recruitment of monocytes, neutrophils and T cells, resulting in the exacerbation of hepatocellular injury [24].
On the other hand, studies report that LSECs are most susceptible to injury during cold storage [19,25,26]. Changes to their physiology and function result in the impairment of their viability. The transcription factor Krüppel-like factor 2 (KLF2), typically expressed by LSECs, is implicated in vasodilation, and has anti-thrombotic and anti-inflammatory effects. During cold ischemia, KLF2 is downregulated alongside with endothelial nitric oxide synthase (eNOS), thrombomodulin, and nuclear factor erythroid 2-related factor 2 (Nrf2). Restoration of these genes in cold stored rat livers prevents liver damage [27]. LSECs also release DAMPs that contribute to the development of an immune response and express adhesive molecules that promote neutrophil binding [26].

3. Mitochondrial Function and Dynamics during Liver I/R

Mitochondria efficiently produce chemical energy (ATP) through oxidative phosphorylation in most eukaryotic cells. These organelles have a wide variety of important functions other than energy metabolism, including metabolism of lipids, amino acids and nucleotides, ROS signaling, calcium homeostasis, and apoptosis [28]. Perturbations to mitochondrial health have serious effects on whole-body homeostasis. Mitochondrial dysfunction has been reported in several diseases involving different organs, such as the liver. In fact, impaired mitochondrial function hinders liver function and contributes to many liver diseases, such as steatosis, non-alcoholic steatohepatitis (NASH), and diabetes [6,29]. As a central coordinator of body glucose and energy metabolism, the liver heavily relies on mitochondria to fully satisfy its functional needs. In the setting of liver I/R, excessive production of ROS and mitochondrial membrane permeabilization, which all result in impaired mitochondrial function, are leading causes of hepatocyte death and LIRI. Targeting mitochondria with antioxidants confers protection against oxidative stress arising from ROS and has hepatoprotective effects against LIRI. For example, N-acetylcysteine (NAC) has been reported to reduce liver injury after I/R injury by decreasing ROS levels, proinflammatory cytokines, and cell death [30,31]. Similarly, targeting mitochondria with coenzyme Q10 protects against LIRI by reducing oxidative stress [32,33].
A variety of mitochondrial quality control mechanisms are activated in response to stress for the preservation of a normal mitochondrial environment [34,35]. The adaptation of mitochondrial function to stress depends on the fine-tuned coordination between nuclear and mitochondrial genomes [36]. To fully optimize mitochondrial function and preserve cellular homeostasis, mitonuclear communication acts in parallel with mitochondrial quality control mechanisms, which include mitochondrial biogenesis, mitochondrial dynamics, and mitophagy. For instance, the generation of functional mitochondria through mitochondrial biogenesis implicates the transcription and translation of genes from both nuclear and mitochondrial genomes. The quantity and quality of mitochondria therefore relies upon the equilibrium between mitochondrial proliferation and degradation. Disrupting this balance results in the decline of cellular function and cell death, which is a feature found in several pathologies [37]. Deficient mitochondrial biogenesis results in mitochondrial stress and decreased mitochondrial mass, while impaired mitophagy leads to the build-up of defective mitochondria. The remodeling of the mitochondrial network is also complemented by processes of mitochondrial fission and fusion. While fusion allows functional mitochondria to join the mitochondrial network, fission is important to separate damaged mitochondria from the network consequently leading to their clearance through mitophagy [38].
Defects in mitochondrial quality control impair the renovation of the mitochondrial pool and exacerbate mitochondrial and cellular damage after liver I/R (Figure 1) [39,40]. Nutrient and oxygen deprivation during the ischemic period disrupts cellular respiration interfering with ATP synthesis. Paradoxically, reoxygenation further aggravates mitochondrial function due to accumulation of free radicals [41] and excessive uptake of Ca2+ into mitochondria [42]. Ca2+ participate in vast biological pathways as important signaling molecules [43]. The constant dynamic state of mitochondria makes them important organelles in the maintenance of cellular Ca2+ homeostasis, being able to uptake high quantities of these molecules. However, the excessive accumulation of Ca2+ in mitochondria during I/R results in the collapse of mitochondrial membrane potential. Together with decreased mitochondrial biogenesis and compromised mitochondrial dynamics, these factors trigger the permanent opening of MPT pores that results in mitochondrial depolarization and swelling [23,40,44]. Consequently, the onset of MPT induces cell death by apoptosis and necrosis [45].

3.1. Mitochondrial Biogenesis

The dynamic nature of mitochondria allows the reshaping of the mitochondrial network in adaptation to stresses of different natures. Mitochondrial biogenesis and mitochondrial dynamics are governed by a sophisticated signaling network involving the activation of many transcription factors in response to intracellular signals and environmental stimuli, such as nutrients and oxygen availability, growth factors or toxins [35]. Peroxisome proliferator-activated receptor (PPAR) gamma coactivator 1-alpha (PGC-1α) is the master regulator of energy metabolism being involved in many biological pathways, such as mitochondrial biogenesis, adaptive thermogenesis and glucose and fatty acid metabolism [46,47]. Activation of PGC-1α and its downstream transcription factors, including nuclear respiratory factors (NRF1 and NRF2) and mitochondrial transcription factor A (TFAM), stimulates the transcription and replication of mtDNA. mtDNA encodes multiple subunits of the mitochondrial respiration chain complexes and therefore is essential for the genesis of new mitochondria [48,49].
Figure 1. Mitochondrial quality control during liver ischemia/reperfusion injury (LIRI). Disruption of mitochondrial quality control processes, such as mitochondrial biogenesis, mitochondrial fission/fusion, and mitophagy, leads to mitochondrial dysfunction and is believed to exacerbate of LIRI.
Figure 1. Mitochondrial quality control during liver ischemia/reperfusion injury (LIRI). Disruption of mitochondrial quality control processes, such as mitochondrial biogenesis, mitochondrial fission/fusion, and mitophagy, leads to mitochondrial dysfunction and is believed to exacerbate of LIRI.
Biomedicines 11 00948 g001
Mitochondrial damage is one of the main causes for LIRI [8,50,51,52]. Studies report that both mitochondrial number and mtDNA content decrease during liver I/R [50,53], suggesting that both mitochondrial function and mitochondrial quality control mechanisms are compromised. In fact, mitochondria from rats subjected to liver transplantation were characterized by having poor bioenergetics and collapsed membrane potential [54]. As expected, a compromised oxidative phosphorylation, resulting in decreased energy production and altered hepatocyte morphology, led to the onset of LIRI. Pretreating the animals with berberine was sufficient to preserve mitochondrial function and to prevent tissue injury [54] (Table 1). Berberine is a natural compound with promising mitochondrial protective effects. Berberine prevents and reverts mitochondrial dysfunction by stimulating mitochondrial biogenesis through activation of the sirtuin 1 (SIRT1)-AMP-activated protein kinase (AMPK) axis [55,56]. During I/R, preconditioning with berberine increased the content of mitochondrial biogenesis-related proteins, such as PGC-1α, SIRT1 and SIRT3 [54]. Similarly, activation of heme oxygenase-1 (HO-1) in mice previous to I/R attenuated LIRI through improvement of mitochondrial function and regulation of mitochondrial quality control [57,58]. Maintenance of a mitochondrial healthy pool by activation of HO-1 using either cilostazol or hemin involved the activation of PGC-1α, NRF-1 and TFAM, which was associated with the increase of mtDNA copy number and mitochondrial mass. In fact, proteins relevant to mitochondrial biogenesis were found to be downregulated during I/R [50,53]. Activation of the SIRT1-AMPK pathway by treating mice with genipin or the SIRT1 activator, SRT1720, was effective in protecting mice against LIRI [50,53].
The energy sensing protein AMPK plays an important role in the regulation of mitochondrial homeostasis and metabolism. Under conditions of energy stress, AMPK is activated to, simultaneously, reduce energy consumption and increase ATP synthesis via stimulation of catabolic pathways [75]. To preserve mitochondrial health, AMPK activates the downstream transcription factor PGC-1α, in a process which requires SIRT1 and which stimulates mitochondrial biogenesis, dynamics, and quality [76]. It is thought that inducing the AMPK-PGC-1α signaling pathway protects mice livers against LIRI [59], since AMPK has protective effects against inflammation, oxidative stress, and apoptosis, which are all underlying pathological causes of I/R [59,77]. Evidence suggests that SIRT1 is necessary for the activation of AMPK through a positive feedback cycle [76]. Given its role in stress responses and its relevance in mitochondria [78], it is plausible to think that it has an important role in LIRI. Indeed, prolonged ischemia causes SIRT1 depletion in hepatocytes and its levels are not restored after reperfusion. Consequently, loss of SIRT1 leads to mitochondrial dysfunction and necrosis after I/R [60]. Activation of SIRT1 during I/R improved the clearance of defective mitochondria through mitophagy and led to the suppression of MPT and necrosis thereby attenuating LIRI [60]. Similarly, increasing SIRT1 levels protected fatty livers against LIRI by decreasing oxidative stress, activating AMPK, and suppressing apoptosis [79]. Curiously, mitofusin-2 (MFN2), a key protein for mitochondrial fusion, was found to have an important role after LIRI. Silencing MFN2 in cells overexpressing SIRT1 abolished the cytoprotective effect of SIRT1 [60].

3.2. Mitochondrial Fission and Fusion

Mitochondrial turnover is crucial for cellular wellbeing. To prevent mitochondrial dysfunction and cellular damage, cells rely on the permanent balance between mitochondrial biogenesis, dynamics and mitophagy. Mitochondrial networks adapt to various cytosolic signals by transitioning between different states. Mitochondria become energetically more efficient when in a hyperfused state whereas clearance of dysfunctional organelles is favored when in a microfused state [80,81]. Transition between these two states requires events of mitochondrial fission and fusion. Upon post-translational modification of dynamin-related protein 1 (DRP1) in the cytosol, DRP1 is translocated to the outer mitochondrial membrane (OMM) where it interacts with its receptors mitochondrial fission factor (MFF), mitochondrial dynamic proteins of 49 kDa (MID49), MID51, and mitochondrial fission 1 protein (FIS1) [81,82,83]. The subsequent oligomerization of DRP1 leads to the constriction and scission of mitochondria. Whereas mitochondrial fusion is initiated with the association between molecules of MFN1 that are anchored to the OMM of the two organelles, the interaction between optic atrophy protein 1 (OPA1) and cardiolipin promotes the fusion of the IMM [81,82,83]. It is believed that MFN2 has a similar role to MFN1 in mitochondrial function, however it has not been fully elucidated yet [81,83].
A growing body of evidence supports the suggestion that I/R injury is influenced by mitochondrial dynamics [63,84,85,86]. Excessive mitochondrial fragmentation is thought to increase cell susceptibility to the MPT onset [87] and to trigger cell death by apoptosis [88], which are hallmarks of LIRI. Reducing mitochondrial fission suppressed the MPT pore opening and decreased cell death after cardiac I/R injury [89]. Moreover, preserving mitochondrial function and suppressing apoptosis during I/R protected mice livers against LIRI [61] (Table 1). Studies have reported that the content of mitochondrial fission- and fusion-related proteins including DRP1 and MFN2 is altered with liver I/R, suggesting that mitochondrial dynamics mechanisms are imbalanced during liver I/R [50,58]. Therefore, it is reasonable to assume that manipulation of mitochondrial fragmentation by directly targeting fission and fusion during liver I/R might have protective effects against LIRI. Indeed, decreasing DRP1 protein levels reduced mitochondrial fission and attenuated LIRI [40,50,58]. Similarly, administration of the hormone irisin protected mice against LIRI by inhibiting DRP1 and FIS1, increasing mitochondrial biogenesis and suppressing apoptosis [40]. Overexpression of augmenter of liver regeneration (ALR) in mice granted resistance over LIRI [61,62]. ALR suppressed the phosphorylation and translocation of DRP1 to the OMM in a mechanism involving cyclin-dependent kinase 1 (CDK1) and cyclin B [62]. Normal activation of DRP1 is regulated by post-translational modifications such as phosphorylation and SUMOylation [90,91]. Curiously, SUMOylation of DRP1 was also found to be controlled by ALR [63]. Manipulation of DRP1 SUMOylation through ALR protected hepatocytes from mitochondrial fragmentation and injury after I/R [63]. Furthermore, decreasing the levels of Drp1 methylated mRNA and its translation impaired DRP1-related mitochondrial fission and ameliorated LIRI [64]. Further evidence suggests that regulation of mitochondrial fusion may also have beneficial effects after I/R. Knocking-out toll-like receptor 4 (TLR4), a protein involved in inflammation, resulted in the upregulation of MFN2 and PGC-1α, and thus improving mitochondrial function and reducing LIRI [65]. However, amelioration of LIRI was also verified by treating mice with silibinin which decreased the levels of mitochondrial fusion-related proteins MFN1 and OPA1 [66].

3.3. Mitophagy

Cell survival depends on the quality of the mitochondrial pool. Accumulation of damaged mitochondria have dire consequences for a cell’s fate. Excessive increase in mitochondrial Ca2+ can lead to mitochondrial dysfunction [92,93]. High levels of Ca2+ induce the MPT, leading to mitochondrial swelling and triggering the release of cytochrome c, which result in apoptotic death [94]. Thus, cells must ensure proper elimination of defective mitochondria. The removal and recycling of unwanted organelles is accomplished through autophagic processes that include macroautophagy, microautophagy, and chaperone-mediated autophagy. In macroautophagy, the major class of autophagy, cellular material is engulfed in autophagosomes that then fuse with lysosomes. The enclosed material is promptly degraded by hydrolytic enzymes in autolysosomes [95]. The selective removal of mitochondria through macroautophagy is denominated mitophagy. Mitophagy ensures adequate cellular homeostasis as its inhibition results in the collapse of mitochondrial function contributing to several human diseases such as metabolic disorders, Alzheimer’s disease, cancer and aging [96,97]. Two main regulatory pathways are responsible for the regulation of mitophagy in mammals, ubiquitin-mediated mitophagy and receptor-mediated mitophagy. These pathways will not be covered here in depth as it is beyond the scope of the current review and they have been extensively reviewed elsewhere [96,98]. Briefly, ubiquitin-mediated mitophagy is dependent upon the interaction between phosphatase and tensin homologue-induced putative kinase 1 (PINK1) and Parkin. Under basal conditions, PINK1 is quickly degraded by ubiquitin-proteasome system after being recruited to the IMM [99]. However, depolarization of mitochondria hinders mitochondrial import and so PINK1 remains stabilized in the OMM [100]. Autophosphorylation of PINK1 leads to the recruitment and activation of Parkin in damaged mitochondria [101]. Substrates of Parkin such as MFN, Miro, and voltage-dependent anion channel (VDAC) are poly-ubiquitinated due to its E3 ligase activity being recognized by adaptor proteins that are anchored to microtubule-associated protein 1A/1B light chain 3 (LC3) proteins in the phagophore [98]. Clearance of mitochondria through receptor-mediated mitophagy involves proteins such as BCL2 interacting protein 3 (BNIP3), BCL2 interacting protein 3 like (NIX) and FUN14 domain containing 1 (FUNDC1) [96,98]. These mitophagy receptors are localized in the OMM and interact directly with LC3 to initiate mitochondrial engulfment and degradation.
Multiple types of cell death contribute to the development of liver damage following liver I/R. Necrosis is the most predominant, although programmed cell death pathways such as apoptosis, ferroptosis, and pyroptosis are also implicated [4,102,103]. Autophagy is typically associated with cell death pathways but it can also act as a mediator of apoptosis, or it can even occur independently (autophagy-dependent cell death) [104]. Nutrient starvation is a trigger of autophagy via regulation of canonical AMPK and mammalian target of rapamycin (mTOR) complex 1 (mTORC1) signaling pathways [105]. In brief, activation of AMPK inhibits mTORC1 which results in the activation of Unc-51 like autophagy activating kinase (ULK1), marking the initiation of autophagy [106,107]. The formation of the autophagosome is characterized by the sequential assembly of protein complexes that involve the recruitment and activation of proteins as beclin-1 (BECN1), ATG7 and LC3 [108]. The mechanisms underlying autophagy during and after liver I/R have been a topic of great debate for the past years as they appear to have a duality of effects [109]. Which possibly results from the employment of different I/R models, utilization of different treatments, and different conditions [103,109,110]. Nonetheless, manipulation of autophagy remains an appealing therapeutic that warrants further studies to clarify the complex mechanisms underlying it during LIRI. Moreover, the enhancement of mitophagy during I/R has shown promising effects [50,67,69]. It is believed that despite nutrient depletion during prolonged ischemia being a trigger of autophagy, the energy deficiency that accompanies it is a major factor in its impairment [23,111]. Loss of ATP is likely to halt autophagosome formation and thus to impair autophagy and mitophagy [23,112]. Similarly, in an advanced stage of reperfusion the accumulation of Ca2+ in mitochondria leads to activation of calpains and in turn proteins implicated in the autophagy machinery, such as ATG7 and BECN1, are hydrolyzed [112]. Energy insufficiency in association with the degradation of autophagy-related proteins contributes to a vicious cycle in which failure to remove deficient mitochondria leads to ROS accumulation, uncoupling of oxidative phosphorylation and cell death. Exploring mitophagy as a therapeutic approach holds great promise not only in diverse pathological conditions but also in aging [97]. Indeed, mitophagy modulators such as urolithin A and spermidine have been shown to have therapeutical potential against physiological decline [113,114].
Mitophagy-mediated elimination of unhealthy mitochondria is expected to enhance mitochondrial and cellular health by impairing the onset of the MPT and improving mitochondrial function, thereby preventing hepatocellular death after LIRI. An increasing number of studies report that mitophagy is disturbed during I/R. After I/R both autophagy- and mitophagy-related proteins are dysregulated, which is likely to be correlated with impaired mitochondrial function, cell death, and progression of LIRI. Although the mechanisms of mitophagy during I/R are still not fully understood, modulation of mitophagy seems to have beneficial effects for the amelioration of LIRI (Table 1). Treatment of mice with genipin before I/R is suspected to restore ubiquitin-dependent mitophagy after I/R as shown by the increase in mitochondrial Parkin [50]. SIRT1 and AMPK are likely to be interconnected and to play a relevant role in the regulation of mitophagy during I/R since inhibition of SIRT1 decreased phosphorylation of AMPK and impaired the effects of genipin on LIRI [50]. I/R injury is aggravated in alcoholic fatty liver-induced mice being characterized by impaired autophagy and mitophagy, increased inflammatory response and extensive hepatocellular death. Pretreatment with 2-methoxyestradiol restored the expression of autophagy-related proteins including BECN1, ATG5, ATG7, LC3-II, and Parkin. This was found to attenuate LIRI through a SIRT1-dependent mechanism [67]. Actually, SIRT1 is also thought to induce mitochondrial autophagy through direct deacetylation of MFN2 [111]. In a recent study, stimulation of mitophagy using 5-aminoimidazole-4-carboxamide ribonucleotide (AICAR), an AMPK activator, during I/R in diabetic mice improved an already aggravated I/R injury. Oxidative stress was attenuated and PINK1/Parkin-related mitophagy was stimulated [68]. Besides being deeply involved in the regulation of energy metabolism, AMPK also regulates a variety of aspects crucial for mitochondrial homeostasis. It is known to regulate mitochondrial biogenesis, dynamics and mitophagy [75]. Its participation in these processes was found to be crucial for its role in the regulation of I/R injury [115,116]. However, the signaling pathways by which it contributes to the amelioration of LIRI are still being uncovered. Additional evidence supports the role of AMPKα in the alleviation of LIRI [69]. Transfusion of mesenchymal stem cells into mice after reperfusion attenuated I/R-induced hepatocellular damage through activation of AMPKα-mediated PINK1-dependent mitophagy [69]. Similarly, both 25-hydroxycholesterol and pterostilbene improved LIRI via activation of PINK1-mediated mitophagy [70,71]. Gu and colleagues provided evidence suggesting that PINK1 is translocated to mitochondria upon I/R through mitochondria-associated membranes, and knocking down PINK1 aggravates LIRI by increasing the pool of dysfunctional mitochondria [117]. Moreover, modulation of Parkin also seems to be beneficial for I/R injury. Treatment of mice with resolvin D1 improved mitochondrial health by reducing mitochondrial swelling, decreasing oxidative stress, and modifying mitochondrial-related proteins including PGC1α, NRF1, TFAM, DRP1, PINK1 and Parkin. Additionally, it was found that the observed effects were mediated by thioredoxin 2 (TRX2)-thioredoxin-interacting protein (TXNIP) signaling which is being discovered as a regulator of mitochondrial quality [72]. Parkin deficiency exacerbates LIRI in rats by suppressing mitophagy [118]. Other pathways of mitophagy are also involved in I/R. For instance, Zhou and colleagues demonstrated that CCAAT/enhancer-binding protein homologous protein (CHOP) knockout decreased hepatocellular death during LIRI. The absence of CHOP led to upregulation of DRP1 and BECN1 and thus increased mitophagy following I/R in mice [73].
The synergy between improved mitochondrial biogenesis and mitophagy resulted in an increased cellular tolerance against LIRI. Mitophagy is interconnected with mitochondrial dynamics as evidenced by Kong and colleagues, where silencing Mfn2 impaired the effects of ALR during LIRI primarily by impairing ubiquitin-dependent mitophagy [74]. However, most previous studies share a common limitation. The lack of a methodology to directly analyze mitophagy has led to the utilization of indirect methods, such as measuring the content of autophagy- and mitophagy-related proteins.

4. Impact of Liver Conditions on LIRI

The increasing demand for liver transplants has led to an alarming number of patients on waiting lists [1,2]. In spite of significant improvements in the past few decades that have improved the success of liver transplantation and improve the quality of life of patients post-surgery, many patients with end-stage liver disease die while waiting for transplantation [119]. Inclusion of marginal liver donors in the liver donor pool is a plausible strategy to neutralize these effects [2]. Currently, liver steatosis and advanced donor age are among the main reasons to discard available livers due to their higher susceptibility to I/R injury that contributes to post-surgery complications and lower survival rate. Decreasing the detrimental effects of I/R injury in livers from marginal donors would increase the number of donors and improve the outcome of liver surgery. The development of such strategies is dependent on the understanding of the underlying mechanisms of I/R injury in livers from marginal donors. In the following sections, we will discuss the impact of age and steatosis in liver transplantation and possible strategies to minimize liver susceptibly to I/R injury (Table 2).

4.1. Fatty Liver Disease

Obesity is a major concern for human health, especially in developed countries, as it has been on the rise for many decades and is associated with many comorbidities. Obesity increases the risk of developing several diseases including metabolic syndrome, type 2 diabetes (T2DM), cardiovascular diseases, and non-alcoholic fatty liver disease (NAFLD) [129]. NAFLD is characterized by the accumulation of triglycerides in hepatocytes (hepatic steatosis) that prompts hepatocyte injury and death, as well as liver inflammation [130]. The progression of NAFLD is related to cirrhosis and hepatocellular carcinoma (HCC). Hepatic steatosis is typically associated with obesity, as substantial accumulation of body fat results in the dysfunction and death of adipocytes [131]. In turn, development of insulin resistance due to secretion of cytokines and inflammatory mediators by adipocytes leads to the release and accumulation of free fatty acids in hepatocytes [130]. This triggers the synthesis of triglycerides by hepatocytes which is associated with the accumulation of diacylglycerols intermediates that lead to the development of hyperglycemia [130]. The excessive accumulation of free fatty acids in the liver makes it more susceptible to injury due to mitochondrial uncoupling and overproduction of ROS through the mitochondrial respiratory chain, stimulation of endoplasmic reticulum stress and activation of cell death receptors [130,131]. Interestingly, mitochondria from injured hepatocytes release DAMPs that activate HSCs and stimulate the progression into liver fibrosis [132]. Liver steatosis increases the risk of morbidity and mortality after liver surgery including liver transplantation and resection [133]. NAFLD is a major risk factor for increased HCC recurrence after liver transplantation [134,135]. Liver transplantation is associated with increased recurrence of HCC, which is likely associated with LIRI [136,137]. The release of proinflammatory cytokines, growth factors and ROS, as well as changes in the liver microenvironment, facilitate the formation and development of metastasis after liver transplantation [136,138,139,140]. While the role of mitochondria in HCC recurrence after liver transplantation is not well understood, it can be assumed that mitochondrial function and health are important contributors to HCC progression. Poorly regulated mitochondrial metabolism, dynamics, ROS, and mitophagy exacerbate proliferation and growth of HCC cells and the development of metastasis [141,142].
Steatotic livers are more vulnerable to I/R injury than non-steatotic livers. In fact, the mechanisms underlying LIRI seem to be different for both cases [26,143]. Apoptosis was shown to be the prevalent type of cell death after I/R injury in normal rat livers while necrosis was the main type of death following I/R injury in steatotic rat livers [143]. Moreover, since mitochondrial dysfunction is a major contributor to the development and progression of NAFLD [144], it might be a likely cause for the increased susceptibility of steatotic livers to I/R injury. During NAFLD, loss of hepatic mitochondrial homeostasis leads to increased oxidative stress that impairs mitochondrial respiration, causes lipid peroxidation, increases the secretion of cytokines, and eventually leads to hepatocyte death [145]. Indeed, higher generation of superoxide was observed during reperfusion in fatty rat livers than in lean rat livers, which is thought to be linked to increased lipid peroxidation and higher susceptibility to LIRI [41]. The superoxide anion is a main type of ROS that when produced in excessive quantity is typically associated with damage to mtDNA and mitochondrial-encoded respiratory chain proteins, thus deteriorating mitochondrial function [146]. ROS-mediated damage to mitochondrial respiratory complexes may help explain the impairment of mitochondrial energy metabolism after I/R in steatotic livers. Caraceni and colleagues demonstrated that oxidative phosphorylation is significantly impaired in fatty livers [147]. In this study, complex I and ATPase activities were found to be decreased during I/R and, after reperfusion, ATP levels were not recovered in steatotic livers [147]. These results are consistent with a later study where the authors show that warm I/R leads to mitochondrial complex I dysfunction and consequently to higher susceptibility of steatotic livers to LIRI [148]. Complex I is a mitochondrial site that is associated with significant ROS production [149]. Mitochondrial uncoupling protein 2 (UCP2) regulates metabolism by working as an uncoupler of oxidative phosphorylation from ATP generation [150]. Interestingly, UCP2 deficiency ameliorated LIRI and increased the survival of mice after I/R [120]. UCP2 was found to be responsible for hepatocyte susceptibility to hypoxia/reoxygenation by reducing mitochondrial membrane potential and ATP levels [121]. Moreover, activation of the AMPK-SIRT1 signaling pathway by renalase mitigated oxidative stress and alleviated mitochondrial dysfunction in steatotic livers [122]. Induction of HO-1 in steatotic livers has also shown promise in the amelioration of steatotic liver I/R injury [151]. Stimulation of mitochondrial function to decrease steatosis-induced susceptibility to LIRI may be a promising strategy to increase the use of marginal liver donors in liver transplantation.

4.2. Aging

Organismal function progressively declines during adulthood, eventually resulting in death. For a long time, researchers have been looking for approaches to hinder this functional decline to increase the health span and life expectancy of human beings. Current efforts are focused on understanding the driving causes of aging and in the discovery of anti-aging interventions, including nutrient restriction and natural compounds [152,153,154]. The causes of aging are currently summarized in well accepted hallmarks [155,156]. Mitochondrial and metabolic dysfunction have major roles in aging [7], since age progression is accompanied by metabolic changes, decline of mitochondrial function, and decline of autophagy efficiency [157,158,159]. The liver is a major regulator of systemic metabolism. But despite having a notable resilience to aging its function still declines during aging, leading to an increase of the incidence of liver diseases [160,161,162]. The decrease of its function is correlated with impaired mitochondrial bioenergetics and increased oxidative injury caused by mtDNA mutations, oxidative stress, and defects in oxidative phosphorylation. Genomic instability, telomere attrition, epigenetic alterations, and deregulated nutrient sensing pathways are some of the other causes underlying age-related liver dysfunction [161]. Evidence suggests that surgical interventions including liver transplantation, as well as resection, have worse outcomes in aged patients than in younger patients [163,164]. Using aged liver grafts for liver transplantation is associated with higher mortality of the recipient [162,165,166]. However, this is source of controversy as systematic reviews demonstrate that the impact of age on liver interventions is not significant, ruling out age as an exclusion criterion for liver resection [167,168]. However, more studies are warranted to clarify if this controversy simply arises from careful liver donor selection or from the type of disease that the patients have at the time of liver resection [169,170,171]. Nonetheless, the increased morbidity and mortality associated with the outcome of liver transplantation is in part due to sensitivity to LIRI. Aged livers have been associated with higher susceptibility to I/R injury in the context of liver transplantation and resection [172,173,174]. In comparison to young mice, adult mice had increased liver injury in response to I/R injury, which was shown to be associated with altered inflammatory response and impaired neutrophil function [174]. Similarly, livers from 9 month-old rats that underwent warm I/R had increased hepatocellular injury compared to livers from 2 month-old rats [173]. Following I/R, livers from old animals were characterized by having increased cellular stress, reduced vasodilation and sinusoidal capillarization [123]. Simvastatin inhibits hydroxy-methylglutaryl coenzyme A (HMG-CoA) and is typically used to lower lipid blood levels. Its administration before I/R in old animals protected their livers against damage [123].
Aging is characterized by the progressive loss of mitochondrial function. With aging the mitochondrial morphology is altered [175], the efficiency of oxidative phosphorylation declines [176], and mitochondrial quality control mechanisms are impaired [159]. Since mitochondrial dysfunction has a big impact on liver’s susceptibility to LIRI, it is possible that aged livers might be more vulnerable to I/R injury in part because of age-related mitochondrial dysfunction. Indeed, Selzner and colleagues noticed that hepatic ATP content decreases after ischemia in both young and old mice. However, contrary to young livers, ATP content is not recovered after reperfusion in the livers of old mice [124]. These results hint that the impact of old age in oxidative phosphorylation and consequently on ATP synthesis exacerbates LIRI. Preconditioning the livers of old mice with short periods of I/R and glucose administration before ischemia was enough to increase ATP content after reperfusion and protect against injury [124]. I/R injury was found to particularly aggravate mitochondrial function in old mice as demonstrated by depletion of proteins relevant to maintain mitochondrial quality such as SIRT1 and MFN2 due to increased expression of calpains [125]. Overexpression of SIRT1 and MFN2 reduced I/R injury in old mice by preventing the opening of the MPT pore, stimulating mitophagy and promoting mitochondrial function [125]. During I/R there is a decrease of calpastatin, which is an inhibitor of calpains. This decrease is aggravated in old animals resulting in mitochondrial dysfunction, impaired autophagy, and hepatocyte death. Its overexpression was shown to be protective against LIRI [126]. The accumulation of unhealthy mitochondria during aging is likely associated with a decline in mitophagy [97]. Age-associated defects in mitophagy were found to exacerbate LIRI in mice as proteins that play essential roles in mitophagy and autophagy, such as Parkin and ATG5, were found to be decreased in the livers of old mice [177]. Actually, improving autophagy by administration of rapamycin and ischemic preconditioning successfully protected livers from I/R injury in old mice [127]. Rapamycin-mediated inhibition of mTOR was shown to improve mitophagy [178], and to have positive effects on lifespan extension [179]. There are several strategies that are currently being study to reverse aging. Parabiosis consists in uniting two living animals resulting in the sharing of blood supply between the two [180]. Circulating factors present in the blood of young mice are thought to have rejuvenating effects and has shown promising effects on restoring synaptic plasticity and improving cognitive function in old mice [181]. Interestingly, administration of plasma from young mice to old mice before I/R attenuated LIRI by improving autophagic activity due to activation of the AMPK/ULK1 signaling pathway [128]. With age progression there is dysregulation of the immune system resulting in a higher tendency for the development of inflammation [182]. Mitochondrial dysfunction and activation of the nucleotide-binding domain and leucine-rich repeat containing protein 3 (NLPR3) inflammasome are potential mechanisms for age-associated inflammation. Since one of the causes for LIRI is related to the inflammatory immune response, its suppression might be a promising strategy to ameliorate LIRI in old livers. Recently, it was found that hepatocytes from old mice release more mtDNA than those from young mice after I/R [183]. Circulating mtDNA has been recognized has being a proinflammatory DAMP [4]. Indeed, increased release of mtDNA activated the NLPR3 inflammasome leading to the stimulation of a macrophage-mediated proinflammatory response that exacerbated LIRI [183].

5. Concluding Remarks

This review provides current knowledge on mitochondrial quality control during LIRI and proposes that the preservation of mitochondrial health is essential for improving patient outcomes. Our knowledge of liver I/R has increased significantly over the last few decades. Understanding the complex mechanisms underlying liver I/R is crucial for improving the clinical success of liver surgeries such as liver transplantation and resection. The ubiquity of mitochondria, as shown by their involvement in several metabolic and signaling pathways, makes them extremely relevant organelles for the maintenance of whole-body homeostasis. Dysregulation of mitochondrial function contributes to the development and progression of many human diseases. As anticipated, mitochondria are also important for liver I/R injury. The production of mitochondrial ROS, calcium overload, and increased mitochondrial permeability are hallmarks of LIRI. These hallmarks are being targeted by emerging strategies that focus on improving mitochondrial structure and function through modulation of mitochondrial biogenesis and dynamics, as well as mitophagy. Thus, a better understanding of mitochondrial and metabolic functions will allow the development of novel interventions aimed at ameliorating LIRI in livers from healthy and marginal donors, including steatotic and old livers. For instance, potential interventions may include supplementation of preservation solutions with chemical compounds or even administration of chemical compounds during reperfusion of the liver to increase the success of the transplanted grafts.
Interestingly, the release of DAMPs from damaged mitochondria into the circulation of liver donors correlates with early graft dysfunction in liver transplant recipients [184,185]. Dysregulation of mitochondrial function is implicated in graft dysfunction and rejection after liver transplantation [186]. Mitochondrial injury is closely related to the outcome of liver transplantation. Indeed, improving mitochondrial function with machine perfusion leads to improved graft function [8,187]. Machine perfusion allows the monitoring of mitochondrial function and the detection of mitochondrial-related biomarkers [187,188]. Both parameters could help predict liver function before liver implantation and liver morbidity and mortality after transplantation [187,188]. A clinical study found that poorly performing mitochondria from human patients were correlated with worse complications after liver transplantation [189].

Author Contributions

I.F.M. and A.P.R. wrote the manuscript. A.P.R. and C.M.P. reviewed and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

I.F.M. is recipient of a PhD scholarship from FCT (DFA/BD/8529/2020).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bodzin, A.S.; Baker, T.B. Liver Transplantation Today: Where We Are Now and Where We Are Going. Liver Transpl. 2018, 24, 1470–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Trapero-Marugán, M.; Little, E.C.; Berenguer, M. Stretching the Boundaries for Liver Transplant in the 21st Century. Lancet Gastroenterol. Hepatol. 2018, 3, 803–811. [Google Scholar] [CrossRef]
  3. Busuttil, R.W.; Tanaka, K. The Utility of Marginal Donors in Liver Transplantation. Liver Transpl. 2003, 9, 651–663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hirao, H.; Nakamura, K.; Kupiec-Weglinski, J.W. Liver Ischaemia–Reperfusion Injury: A New Understanding of the Role of Innate Immunity. Nat. Rev. Gastroenterol. Hepatol. 2022, 19, 239–256. [Google Scholar] [CrossRef]
  5. Zhai, Y.; Petrowsky, H.; Hong, J.C.; Busuttil, R.W.; Kupiec-Weglinski, J.W. Ischaemia–Reperfusion Injury in Liver Transplantation—From Bench to Bedside. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 79–89. [Google Scholar] [CrossRef]
  6. Morio, B.; Panthu, B.; Bassot, A.; Rieusset, J. Role of Mitochondria in Liver Metabolic Health and Diseases. Cell Calcium 2021, 94, 102336. [Google Scholar] [CrossRef]
  7. Amorim, J.A.; Coppotelli, G.; Rolo, A.P.; Palmeira, C.M.; Ross, J.M.; Sinclair, D.A. Mitochondrial and Metabolic Dysfunction in Ageing and Age-Related Diseases. Nat. Rev. Endocrinol. 2022, 18, 243–258. [Google Scholar] [CrossRef]
  8. Teodoro, J.S.; Silva, R.T.D.; Machado, I.F.; Panisello-Roselló, A.; Roselló-Catafau, J.; Rolo, A.P.; Palmeira, C.M. Shaping of Hepatic Ischemia/Reperfusion Events: The Crucial Role of Mitochondria. Cells 2022, 11, 688. [Google Scholar] [CrossRef]
  9. Alexandrino, H.; Rolo, A.; Tralhão, J.G.; Castro e Sousa, F.; Palmeira, C. Mitochondria in Liver Regeneration: Energy Metabolism and Posthepatectomy Liver Dysfunction. In Mitochondrial Biology and Experimental Therapeutics; Oliveira, P.J., Ed.; Springer International Publishing: Berlin/Heidelberg, Germany, 2018; pp. 127–152. [Google Scholar]
  10. Bernal, W.; Lee, W.M.; Wendon, J.; Larsen, F.S.; Williams, R. Acute Liver Failure: A Curable Disease by 2024? J. Hepatol. 2015, 62, S112–S120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Ben-Moshe, S.; Itzkovitz, S. Spatial Heterogeneity in the Mammalian Liver. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 395–410. [Google Scholar] [CrossRef]
  12. Wei, Y.; Wang, Y.G.; Jia, Y.; Li, L.; Yoon, J.; Zhang, S.; Wang, Z.; Zhang, Y.; Zhu, M.; Sharma, T.; et al. Liver Homeostasis Is Maintained by Midlobular Zone 2 Hepatocytes. Science 2021, 371, eabb1625. [Google Scholar] [CrossRef]
  13. Branum, G.D.; Selim, N.; Liu, X.; Whalen, R.; Boyer, T.D. Ischaemia and Reperfusion Injury of Rat Liver Increases Expression of Glutathione S-Transferase A1/A2 in Zone 3 of the Hepatic Lobule. Biochem. J. 1998, 330, 73–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kern, H.; Bald, C.; Brill, T.; Fend, F.; Von Weihern, C.H.; Kriner, M.; Hüser, N.; Thorban, S.; Stangl, M.; Matevossian, E. The Influence of Retrograde Reperfusion on the Ischaemia-/ Reperfusion Injury after Liver Transplantation in the Rat. Int. J. Exp. Path. 2008, 89, 433–437. [Google Scholar] [CrossRef] [PubMed]
  15. Nakanuma, S.; Tajima, H.; Takamura, H.; Sakai, S.; Gabata, R.; Okazaki, M.; Shinbashi, H.; Ohbatake, Y.; Makino, I.; Hayashi, H.; et al. Pretreatment with a Phosphodiesterase-3 Inhibitor, Milrinone, Reduces Hepatic Ischemia-Reperfusion Injury, Minimizing Pericentral Zone-Based Liver and Small Intestinal Injury in Rats. Ann. Transplant. 2020, 25, e922306. [Google Scholar] [CrossRef]
  16. Broughan, T.A.; Naukam, R.; Tan, C.; Van De Wiele, C.J.; Refai, H.; Teague, T.K. Effects of Hepatic Zonal Oxygen Levels on Hepatocyte Stress Responses. J. Surg. Res. 2008, 145, 150–160. [Google Scholar] [CrossRef] [PubMed]
  17. Xin, J.; Yang, T.; Wu, X.; Wu, Y.; Liu, Y.; Liu, X.; Jiang, M.; Gao, W. Spatial Transcriptomics Analysis of Zone-Dependent Hepatic Ischemia-Reperfusion Injury Murine Model. Commun. Biol. 2023, 6, 194. [Google Scholar] [CrossRef] [PubMed]
  18. Gujral, J.S.; Bucci, T.J.; Farhood, A.; Jaeschke, H. Mechanism of Cell Death during Warm Hepatic Ischemia-Reperfusion in Rats: Apoptosis or Necrosis? Hepatology 2001, 33, 397–405. [Google Scholar] [CrossRef]
  19. Rauen, U.; Elling, B.; Gizewski, E.R.; Korth, H.-G.; Sustmann, R.; de Groot, H. Involvement of Reactive Oxygen Species in the Preservation Injury to Cultured Liver Endothelial Cells. Free Radic. Biol. Med. 1997, 22, 17–24. [Google Scholar] [CrossRef]
  20. Montalvo-Jave, E.E.; Escalante-Tattersfield, T.; Ortega-Salgado, J.A.; Piña, E.; Geller, D.A. Factors in the Pathophysiology of the Liver Ischemia-Reperfusion Injury. J. Surg. Res. 2008, 147, 153–159. [Google Scholar] [CrossRef] [Green Version]
  21. Halestrap, A.P. What Is the Mitochondrial Permeability Transition Pore? J. Mol. Cell. Cardiol. 2009, 46, 821–831. [Google Scholar] [CrossRef]
  22. Kwong, J.Q.; Molkentin, J.D. Physiological and Pathological Roles of the Mitochondrial Permeability Transition Pore in the Heart. Cell Metab. 2015, 21, 206–214. [Google Scholar] [CrossRef] [Green Version]
  23. Go, K.L.; Lee, S.; Zendejas, I.; Behrns, K.E.; Kim, J.-S. Mitochondrial Dysfunction and Autophagy in Hepatic Ischemia/Reperfusion Injury. BioMed Res. Int. 2015, 2015, 183469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Li, P.; He, K.; Li, J.; Liu, Z.; Gong, J. The Role of Kupffer Cells in Hepatic Diseases. Mol. Immunol. 2017, 85, 222–229. [Google Scholar] [CrossRef] [PubMed]
  25. Caldwell-Kenkel, J.C.; Thurman, R.G.; Lemasters, J.J. Selective Loss of Nonparenchymal Cell Viability after Cold Ischemic Storage of Rat Livers. Transplantation 1988, 45, 834–836. [Google Scholar] [CrossRef] [PubMed]
  26. Peralta, C.; Jiménez-Castro, M.B.; Gracia-Sancho, J. Hepatic Ischemia and Reperfusion Injury: Effects on the Liver Sinusoidal Milieu. J. Hepatol. 2013, 59, 1094–1106. [Google Scholar] [CrossRef] [Green Version]
  27. Russo, L.; Gracia-Sancho, J.; García-Calderó, H.; Marrone, G.; García-Pagán, J.C.; García-Cardeña, G.; Bosch, J. Addition of Simvastatin to Cold Storage Solution Prevents Endothelial Dysfunction in Explanted Rat Livers. Hepatology 2012, 55, 921–930. [Google Scholar] [CrossRef]
  28. Pfanner, N.; Warscheid, B.; Wiedemann, N. Mitochondrial Proteins: From Biogenesis to Functional Networks. Nat. Rev. Mol. Cell Biol. 2019, 20, 267–284. [Google Scholar] [CrossRef] [PubMed]
  29. Mansouri, A.; Gattolliat, C.-H.; Asselah, T. Mitochondrial Dysfunction and Signaling in Chronic Liver Diseases. Gastroenterology 2018, 155, 629–647. [Google Scholar] [CrossRef] [Green Version]
  30. Chaves Cayuela, N.; Kiyomi Koike, M.; Jacysyn, J.; Rasslan, R.; Azevedo Cerqueira, A.; Pereira Costa, S.; Picanço Diniz-Júnior, J.; Massazo Utiyama, E.; Frasson de Souza Montero, E. N-Acetylcysteine Reduced Ischemia and Reperfusion Damage Associated with Steatohepatitis in Mice. Int. J. Mol. Sci. 2020, 21, 4106. [Google Scholar] [CrossRef]
  31. Sun, Y.; Pu, L.-Y.; Lu, L.; Wang, X.-H.; Zhang, F.; Rao, J.-H. N-Acetylcysteine Attenuates Reactive-Oxygen-Species-Mediated Endoplasmic Reticulum Stress during Liver Ischemia-Reperfusion Injury. World J. Gastroenterol. 2014, 20, 15289. [Google Scholar] [CrossRef] [Green Version]
  32. Portakal, O.; İnal-Erden, M. Effects of Pentoxifylline and Coenzyme Q10 in Hepatic Ischemia/Reperfusion Injury. Clin. Biochem. 1999, 32, 461–466. [Google Scholar] [CrossRef]
  33. Genova, M.L.; Bonacorsi, E.; D’Aurelio, M.; Formiggini, G.; Nardo, B.; Cuccomarino, S.; Turi, P.; Pich, M.M.; Lenaz, G.; Bovina, C. Protective Effect of Exogenous Coenzyme Q in Rats Subjected to Partial Hepatic Ischemia and Reperfusion. BioFactors 1999, 9, 345–349. [Google Scholar] [CrossRef]
  34. Annesley, S.J.; Fisher, P.R. Mitochondria in Health and Disease. Cells 2019, 8, 680. [Google Scholar] [CrossRef] [Green Version]
  35. Palikaras, K.; Lionaki, E.; Tavernarakis, N. Balancing Mitochondrial Biogenesis and Mitophagy to Maintain Energy Metabolism Homeostasis. Cell Death Differ. 2015, 22, 1399–1401. [Google Scholar] [CrossRef] [Green Version]
  36. Quirós, P.M.; Mottis, A.; Auwerx, J. Mitonuclear Communication in Homeostasis and Stress. Nat. Rev. Mol. Cell Biol. 2016, 17, 213–226. [Google Scholar] [CrossRef] [PubMed]
  37. Palikaras, K.; Tavernarakis, N. Mitochondrial Homeostasis: The Interplay between Mitophagy and Mitochondrial Biogenesis. Exp. Gerontol. 2014, 56, 182–188. [Google Scholar] [CrossRef] [PubMed]
  38. Youle, R.J.; van der Bliek, A.M. Mitochondrial Fission, Fusion, and Stress. Science 2012, 337, 1062–1065. [Google Scholar] [CrossRef] [Green Version]
  39. Zhang, H.; Yan, Q.; Wang, X.; Chen, X.; Chen, Y.; Du, J.; Chen, L. The Role of Mitochondria in Liver Ischemia-Reperfusion Injury: From Aspects of Mitochondrial Oxidative Stress, Mitochondrial Fission, Mitochondrial Membrane Permeable Transport Pore Formation, Mitophagy, and Mitochondria-Related Protective Measures. Oxid. Med. Cell. Longev. 2021, 2021, 6670579. [Google Scholar] [CrossRef] [PubMed]
  40. Bi, J.; Zhang, J.; Ren, Y.; Du, Z.; Li, Q.; Wang, Y.; Wei, S.; Yang, L.; Zhang, J.; Liu, C.; et al. Irisin Alleviates Liver Ischemia-Reperfusion Injury by Inhibiting Excessive Mitochondrial Fission, Promoting Mitochondrial Biogenesis and Decreasing Oxidative Stress. Redox Biol. 2019, 20, 296–306. [Google Scholar] [CrossRef]
  41. Nardo, B.; Caraceni, P.; Pasini, P.; Domenicali, M.; Catena, F.; Cavallari, G.; Santoni, B.; Maiolini, E.; Grattagliano, I.; Vendemiale, G.; et al. Increased Generation of Reactive Oxygen Species in Isolated Rat Fatty Liver during Postischemic Reoxygenation. Transplantation 2001, 71, 1816–1820. [Google Scholar] [CrossRef]
  42. Chang, W.J.; Chehab, M.; Kink, S.; Toledo-Pereyra, L.H. Intracellular Calcium Signaling Pathways during Liver Ischemia and Reperfusion. J. Investig. Surg. 2010, 23, 228–238. [Google Scholar] [CrossRef] [PubMed]
  43. 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]
  44. Wang, J.; Toan, S.; Zhou, H. Mitochondrial Quality Control in Cardiac Microvascular Ischemia-Reperfusion Injury: New Insights into the Mechanisms and Therapeutic Potentials. Pharmacol. Res. 2020, 156, 104771. [Google Scholar] [CrossRef] [PubMed]
  45. Kim, J.-S.; He, L.; Lemasters, J.J. Mitochondrial Permeability Transition: A Common Pathway to Necrosis and Apoptosis. Biochem. Biophys. Res. Commun. 2003, 304, 463–470. [Google Scholar] [CrossRef] [PubMed]
  46. Puigserver, P.; Spiegelman, B.M. Peroxisome Proliferator-Activated Receptor-γ Coactivator 1α (PGC-1α): Transcriptional Coactivator and Metabolic Regulator. Endocr. Rev. 2003, 24, 78–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Liang, H.; Ward, W.F. PGC-1α: A Key Regulator of Energy Metabolism. Adv. Physiol. Educ. 2006, 30, 145–151. [Google Scholar] [CrossRef]
  48. Larsson, N.-G. Somatic Mitochondrial DNA Mutations in Mammalian Aging. Annu. Rev. Biochem. 2010, 79, 683–706. [Google Scholar] [CrossRef]
  49. Taylor, R.W.; Turnbull, D.M. Mitochondrial DNA Mutations in Human Disease. Nat. Rev. Genet. 2005, 6, 389–402. [Google Scholar] [CrossRef] [Green Version]
  50. Shin, J.-K.; Lee, S.-M. Genipin Protects the Liver from Ischemia/Reperfusion Injury by Modulating Mitochondrial Quality Control. Toxicol. Appl. Pharmacol. 2017, 328, 25–33. [Google Scholar] [CrossRef]
  51. Halestrap, A.P. Calcium, Mitochondria and Reperfusion Injury: A Pore Way to Die. Biochem. Soc. Trans. 2006, 34, 232–237. [Google Scholar] [CrossRef]
  52. Jassem, W.; Heaton, N.D. The Role of Mitochondria in Ischemia/Reperfusion Injury in Organ Transplantation. Kidney Int. 2004, 66, 514–517. [Google Scholar] [CrossRef] [Green Version]
  53. Khader, A.; Yang, W.-L.; Godwin, A.; Prince, J.M.; Nicastro, J.M.; Coppa, G.F.; Wang, P. Sirtuin 1 Stimulation Attenuates Ischemic Liver Injury and Enhances Mitochondrial Recovery and Autophagy. Crit. Care Med. 2016, 44, e651–e663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Martins, R.; Pinto Rolo, A.; Soeiro Teodoro, J.; Furtado, E.; Caetano Oliveira, R.; Tralhão, J.; Marques Palmeira, C. Addition of Berberine to Preservation Solution in an Animal Model of Ex Vivo Liver Transplant Preserves Mitochondrial Function and Bioenergetics from the Damage Induced by Ischemia/Reperfusion. Int. J. Mol. Sci. 2018, 19, 284. [Google Scholar] [CrossRef] [Green Version]
  55. Teodoro, J.S.; Duarte, F.V.; Gomes, A.P.; Varela, A.T.; Peixoto, F.M.; Rolo, A.P.; Palmeira, C.M. Berberine Reverts Hepatic Mitochondrial Dysfunction in High-Fat Fed Rats: A Possible Role for SirT3 Activation. Mitochondrion 2013, 13, 637–646. [Google Scholar] [CrossRef] [Green Version]
  56. Gomes, A.P.; Duarte, F.V.; Nunes, P.; Hubbard, B.P.; Teodoro, J.S.; Varela, A.T.; Jones, J.G.; Sinclair, D.A.; Palmeira, C.M.; Rolo, A.P. Berberine Protects against High Fat Diet-Induced Dysfunction in Muscle Mitochondria by Inducing SIRT1-Dependent Mitochondrial Biogenesis. Biochim. Biophys. Acta 2012, 1822, 185–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Joe, Y.; Zheng, M.; Kim, H.J.; Uddin, M.J.; Kim, S.-K.; Chen, Y.; Park, J.; Cho, G.J.; Ryter, S.W.; Chung, H.T. Cilostazol Attenuates Murine Hepatic Ischemia and Reperfusion Injury via Heme Oxygenase-Dependent Activation of Mitochondrial Biogenesis. Am. J. Physiol. Gastrointest. Liver Physiol. 2015, 309, G21–G29. [Google Scholar] [CrossRef] [Green Version]
  58. Hong, J.-M.; Lee, S.-M. Heme Oxygenase-1 Protects Liver against Ischemia/Reperfusion Injury via Phosphoglycerate Mutase Family Member 5-Mediated Mitochondrial Quality Control. Life Sci. 2018, 200, 94–104. [Google Scholar] [CrossRef] [PubMed]
  59. Li, J.; Li, J.; Fang, H.; Yang, H.; Wu, T.; Shi, X.; Pang, C. Isolongifolene Alleviates Liver Ischemia/Reperfusion Injury by Regulating AMPK-PGC1α Signaling Pathway-Mediated Inflammation, Apoptosis, and Oxidative Stress. Int. Immunopharmacol. 2022, 113, 109185. [Google Scholar] [CrossRef]
  60. Biel, T.G.; Lee, S.; Flores-Toro, J.A.; Dean, J.W.; Go, K.L.; Lee, M.-H.; Law, B.K.; Law, M.E.; Dunn, W.A.; Zendejas, I.; et al. Sirtuin 1 Suppresses Mitochondrial Dysfunction of Ischemic Mouse Livers in a Mitofusin 2-Dependent Manner. Cell Death Differ. 2016, 23, 279–290. [Google Scholar] [CrossRef] [Green Version]
  61. Jiang, S.-J.; Li, W.; An, W. Adenoviral Gene Transfer of Hepatic Stimulator Substance Confers Resistance against Hepatic Ischemia–Reperfusion Injury by Improving Mitochondrial Function. Hum. Gene Ther. 2013, 24, 443–456. [Google Scholar] [CrossRef] [Green Version]
  62. Zhang, C.; Huang, J.; An, W. Hepatic Stimulator Substance Resists Hepatic Ischemia/Reperfusion Injury by Regulating Drp1 Translocation and Activation. Hepatology 2017, 66, 1989–2001. [Google Scholar] [CrossRef] [Green Version]
  63. Huang, J.; Xie, P.; Dong, Y.; An, W. Inhibition of Drp1 SUMOylation by ALR Protects the Liver from Ischemia-Reperfusion Injury. Cell Death Differ. 2021, 28, 1174–1192. [Google Scholar] [CrossRef] [PubMed]
  64. Du, Y.D.; Guo, W.Y.; Han, C.H.; Wang, Y.; Chen, X.S.; Li, D.W.; Liu, J.L.; Zhang, M.; Zhu, N.; Wang, X. N6-Methyladenosine Demethylase FTO Impairs Hepatic Ischemia–Reperfusion Injury via Inhibiting Drp1-Mediated Mitochondrial Fragmentation. Cell Death Dis. 2021, 12, 442. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, C.; Jia, Y.; Liu, B.; Wang, G.; Zhang, Y. TLR4 Knockout Upregulates the Expression of Mfn2 and PGC-1α in a High-Fat Diet and Ischemia-Reperfusion Mice Model of Liver Injury. Life Sci. 2020, 254, 117762. [Google Scholar] [CrossRef] [PubMed]
  66. Qajari, N.M.; Shafaroudi, M.M.; Gholami, M.; Khonakdar-Tarsi, A. Silibinin Treatment Results in Reducing OPA1&MFN1 Genes Expression in a Rat Model Hepatic Ischemia–Reperfusion. Mol. Biol. Rep. 2020, 47, 3271–3280. [Google Scholar] [CrossRef]
  67. Cho, H.-I.; Seo, M.-J.; Lee, S.-M. 2-Methoxyestradiol Protects against Ischemia/Reperfusion Injury in Alcoholic Fatty Liver by Enhancing Sirtuin 1-Mediated Autophagy. Biochem. Pharmacol. 2017, 131, 40–51. [Google Scholar] [CrossRef]
  68. Zhijun, K.; Xudong, Z.; Baoqiang, W.; Chunfu, Z.; Qiang, Y.; Yuan, G.; Xihu, Q. Increased Oxidative Stress Caused by Impaired Mitophagy Aggravated Liver Ischemia and Reperfusion Injury in Diabetic Mice. J. Diabetes Investig. 2023, 14, 28–36. [Google Scholar] [CrossRef]
  69. Zheng, J.; Chen, L.; Lu, T.; Zhang, Y.; Sui, X.; Li, Y.; Huang, X.; He, L.; Cai, J.; Zhou, C.; et al. MSCs Ameliorate Hepatocellular Apoptosis Mediated by PINK1-Dependent Mitophagy in Liver Ischemia/Reperfusion Injury through AMPKα Activation. Cell Death Dis. 2020, 11, 256. [Google Scholar] [CrossRef] [Green Version]
  70. Shi, Q.; Zhao, G.; Wei, S.; Guo, C.; Wu, X.; Zhao, R.C.; Di, G. Pterostilbene Alleviates Liver Ischemia/Reperfusion Injury via PINK1-Mediated Mitophagy. J. Pharmacol. Sci. 2022, 148, 19–30. [Google Scholar] [CrossRef]
  71. Cao, Q.; Luo, J.; Xiong, Y.; Liu, Z.; Ye, Q. 25-Hydroxycholesterol Mitigates Hepatic Ischemia Reperfusion Injury via Mediating Mitophagy. Int. Immunopharmacol. 2021, 96, 107643. [Google Scholar] [CrossRef]
  72. Kang, J.-W.; Choi, H.-S.; Lee, S.-M. Resolvin D1 Attenuates Liver Ischaemia/Reperfusion Injury through Modulating Thioredoxin 2-Mediated Mitochondrial Quality Control: Role of Resolvin D1 in Mitochondrial Quality Control. Br. J. Pharmacol. 2018, 175, 2441–2453. [Google Scholar] [CrossRef] [Green Version]
  73. Zhou, S.; Rao, Z.; Xia, Y.; Wang, Q.; Liu, Z.; Wang, P.; Cheng, F.; Zhou, H. CCAAT/Enhancer-Binding Protein Homologous Protein Promotes ROS-Mediated Liver Ischemia and Reperfusion Injury by Inhibiting Mitophagy in Hepatocytes. Transplantation 2023, 107, 129–139. [Google Scholar] [CrossRef]
  74. Kong, W.; Li, W.; Bai, C.; Dong, Y.; Wu, Y.; An, W. Augmenter of Liver Regeneration-Mediated Mitophagy Protects against Hepatic Ischemia/Reperfusion Injury. Am. J. Transplant. 2022, 22, 130–143. [Google Scholar] [CrossRef] [PubMed]
  75. Herzig, S.; Shaw, R.J. AMPK: Guardian of Metabolism and Mitochondrial Homeostasis. Nat. Rev. Mol. Cell Biol. 2017, 19, 121–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Price, N.L.; Gomes, A.P.; Ling, A.J.Y.; Duarte, F.V.; Martin-Montalvo, A.; North, B.J.; Agarwal, B.; Ye, L.; Ramadori, G.; Teodoro, J.S.; et al. SIRT1 Is Required for AMPK Activation and the Beneficial Effects of Resveratrol on Mitochondrial Function. Cell Metab. 2012, 15, 675–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Ding, R.; Wu, W.; Sun, Z.; Li, Z. AMP-Activated Protein Kinase: An Attractive Therapeutic Target for Ischemia-Reperfusion Injury. Eur. J. Pharmacol. 2020, 888, 173484. [Google Scholar] [CrossRef]
  78. Gomes, A.P.; Price, N.L.; Ling, A.J.Y.; Moslehi, J.J.; Montgomery, M.K.; Rajman, L.; White, J.P.; Teodoro, J.S.; Wrann, C.D.; Hubbard, B.P.; et al. Declining NAD+ Induces a Pseudohypoxic State Disrupting Nuclear-Mitochondrial Communication during Aging. Cell 2013, 155, 1624–1638. [Google Scholar] [CrossRef] [Green Version]
  79. Pantazi, E.; Zaouali, M.A.; Bejaoui, M.; Serafin, A.; Folch-Puy, E.; Petegnief, V.; De Vera, N.; Abdennebi, H.B.; Rimola, A.; Roselló-Catafau, J. Silent Information Regulator 1 Protects the Liver against Ischemia-Reperfusion Injury: Implications in Steatotic Liver Ischemic Preconditioning. Transpl. Int. 2014, 27, 493–503. [Google Scholar] [CrossRef]
  80. Hoitzing, H.; Johnston, I.G.; Jones, N.S. What Is the Function of Mitochondrial Networks? A Theoretical Assessment of Hypotheses and Proposal for Future Research. BioEssays 2015, 37, 687–700. [Google Scholar] [CrossRef] [Green Version]
  81. Giacomello, M.; Pyakurel, A.; Glytsou, C.; Scorrano, L. The Cell Biology of Mitochondrial Membrane Dynamics. Nat. Rev. Mol. Cell Biol. 2020, 21, 204–224. [Google Scholar] [CrossRef]
  82. Tilokani, L.; Nagashima, S.; Paupe, V.; Prudent, J. Mitochondrial Dynamics: Overview of Molecular Mechanisms. Essays Biochem. 2018, 62, 341–360. [Google Scholar] [CrossRef] [Green Version]
  83. Filadi, R.; Pendin, D.; Pizzo, P. Mitofusin 2: From Functions to Disease. Cell Death Dis. 2018, 9, 330. [Google Scholar] [CrossRef] [Green Version]
  84. Kim, H.; Scimia, M.C.; Wilkinson, D.; Trelles, R.D.; Wood, M.R.; Bowtell, D.; Dillin, A.; Mercola, M.; Ronai, Z.A. Fine-Tuning of Drp1/Fis1 Availability by AKAP121/Siah2 Regulates Mitochondrial Adaptation to Hypoxia. Mol. Cell 2011, 44, 532–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Tian, X.; Zhao, Y.; Yang, Z.; Lu, Q.; Zhou, L.; Zheng, S. USP15 Regulates P66Shc Stability Associated with Drp1 Activation in Liver Ischemia/Reperfusion. Cell Death Dis. 2022, 13, 823. [Google Scholar] [CrossRef]
  86. Tang, J.; Hu, Z.; Tan, J.; Yang, S.; Zeng, L. Parkin Protects against Oxygen-Glucose Deprivation/Reperfusion Insult by Promoting Drp1 Degradation. Oxid. Med. Cell. Longev. 2016, 2016, 8474303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Kong, D.; Xu, L.; Yu, Y.; Zhu, W.; Andrews, D.W.; Yoon, Y.; Kuo, T.H. Regulation of Ca2+-Induced Permeability Transition by Bcl-2 Is Antagonized by Drp1 and HFis1. Mol. Cell. Biochem. 2005, 272, 187–199. [Google Scholar] [CrossRef] [PubMed]
  88. Youle, R.J.; Karbowski, M. Mitochondrial Fission in Apoptosis. Nat. Rev. Mol. Cell Biol. 2005, 6, 657–663. [Google Scholar] [CrossRef]
  89. Ong, S.-B.; Subrayan, S.; Lim, S.Y.; Yellon, D.M.; Davidson, S.M.; Hausenloy, D.J. Inhibiting Mitochondrial Fission Protects the Heart against Ischemia/Reperfusion Injury. Circulation 2010, 121, 2012–2022. [Google Scholar] [CrossRef] [Green Version]
  90. Wasiak, S.; Zunino, R.; McBride, H.M. Bax/Bak Promote Sumoylation of DRP1 and Its Stable Association with Mitochondria during Apoptotic Cell Death. J. Cell Biol. 2007, 177, 439–450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Chang, C.-R.; Blackstone, C. Cyclic AMP-Dependent Protein Kinase Phosphorylation of Drp1 Regulates Its GTPase Activity and Mitochondrial Morphology. J. Biol. Chem. 2007, 282, 21583–21587. [Google Scholar] [CrossRef] [Green Version]
  92. Strubbe-Rivera, J.O.; Schrad, J.R.; Pavlov, E.V.; Conway, J.F.; Parent, K.N.; Bazil, J.N. The Mitochondrial Permeability Transition Phenomenon Elucidated by Cryo-EM Reveals the Genuine Impact of Calcium Overload on Mitochondrial Structure and Function. Sci. Rep. 2021, 11, 1037. [Google Scholar] [CrossRef] [PubMed]
  93. Santulli, G.; Xie, W.; Reiken, S.R.; Marks, A.R. Mitochondrial Calcium Overload Is a Key Determinant in Heart Failure. Proc. Natl. Acad. Sci. USA 2015, 112, 11389–11394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Giorgi, C.; Baldassari, F.; Bononi, A.; Bonora, M.; De Marchi, E.; Marchi, S.; Missiroli, S.; Patergnani, S.; Rimessi, A.; Suski, J.M.; et al. Mitochondrial Ca2+ and Apoptosis. Cell Calcium 2012, 52, 36–43. [Google Scholar] [CrossRef] [Green Version]
  95. Mehrpour, M.; Esclatine, A.; Beau, I.; Codogno, P. Overview of Macroautophagy Regulation in Mammalian Cells. Cell Res. 2010, 20, 748–762. [Google Scholar] [CrossRef] [PubMed]
  96. Palikaras, K.; Lionaki, E.; Tavernarakis, N. Mechanisms of Mitophagy in Cellular Homeostasis, Physiology and Pathology. Nat. Cell Biol. 2018, 20, 1013–1022. [Google Scholar] [CrossRef]
  97. Bakula, D.; Scheibye-Knudsen, M. MitophAging: Mitophagy in Aging and Disease. Front. Cell Dev. Biol. 2020, 8, 239. [Google Scholar] [CrossRef] [Green Version]
  98. Onishi, M.; Yamano, K.; Sato, M.; Matsuda, N.; Okamoto, K. Molecular Mechanisms and Physiological Functions of Mitophagy. EMBO J. 2021, 40, e104705. [Google Scholar] [CrossRef]
  99. Sekine, S.; Youle, R.J. PINK1 Import Regulation; a Fine System to Convey Mitochondrial Stress to the Cytosol. BMC Biol. 2018, 16, 2. [Google Scholar] [CrossRef] [Green Version]
  100. Narendra, D.; Tanaka, A.; Suen, D.-F.; Youle, R.J. Parkin Is Recruited Selectively to Impaired Mitochondria and Promotes Their Autophagy. J. Cell Biol. 2008, 183, 795–803. [Google Scholar] [CrossRef] [Green Version]
  101. Okatsu, K.; Oka, T.; Iguchi, M.; Imamura, K.; Kosako, H.; Tani, N.; Kimura, M.; Go, E.; Koyano, F.; Funayama, M.; et al. PINK1 Autophosphorylation upon Membrane Potential Dissipation Is Essential for Parkin Recruitment to Damaged Mitochondria. Nat. Commun. 2012, 3, 1016. [Google Scholar] [CrossRef] [Green Version]
  102. Jaeschke, H.; Lemasters, J.J. Apoptosis versus Oncotic Necrosis in Hepatic Ischemia/Reperfusion Injury. Gastroenterology 2003, 125, 1246–1257. [Google Scholar] [CrossRef]
  103. Hu, C.; Zhao, L.; Zhang, F.; Li, L. Regulation of Autophagy Protects against Liver Injury in Liver Surgery-induced Ischaemia/Reperfusion. J. Cell. Mol. Med. 2021, 25, 9905–9917. [Google Scholar] [CrossRef] [PubMed]
  104. Denton, D.; Kumar, S. Autophagy-Dependent Cell Death. Cell Death Differ. 2019, 26, 605–616. [Google Scholar] [CrossRef] [Green Version]
  105. He, L.; Zhang, J.; Zhao, J.; Ma, N.; Kim, S.W.; Qiao, S.; Ma, X. Autophagy: The Last Defense against Cellular Nutritional Stress. Adv. Nutr. 2018, 9, 493–504. [Google Scholar] [CrossRef] [Green Version]
  106. Ge, Y.; Zhou, M.; Chen, C.; Wu, X.; Wang, X. Role of AMPK Mediated Pathways in Autophagy and Aging. Biochimie 2022, 195, 100–113. [Google Scholar] [CrossRef] [PubMed]
  107. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.-L. AMPK and MTOR Regulate Autophagy through Direct Phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Aman, Y.; Schmauck-Medina, T.; Hansen, M.; Morimoto, R.I.; Simon, A.K.; Bjedov, I.; Palikaras, K.; Simonsen, A.; Johansen, T.; Tavernarakis, N.; et al. Autophagy in Healthy Aging and Disease. Nat. Aging 2021, 1, 634–650. [Google Scholar] [CrossRef] [PubMed]
  109. Sheng, R.; Qin, Z. The Divergent Roles of Autophagy in Ischemia and Preconditioning. Acta Pharmacol. Sin. 2015, 36, 411–420. [Google Scholar] [CrossRef]
  110. Cursio, R.; Colosetti, P.; Gugenheim, J. Autophagy and Liver Ischemia-Reperfusion Injury. BioMed Res. Int. 2015, 2015, 417590. [Google Scholar] [CrossRef]
  111. Chun, S.K.; Go, K.; Yang, M.-J.; Zendejas, I.; Behrns, K.E.; Kim, J.-S. Autophagy in Ischemic Livers: A Critical Role of Sirtuin 1/Mitofusin 2 Axis in Autophagy Induction. Toxicol. Res. 2016, 32, 35–46. [Google Scholar] [CrossRef] [Green Version]
  112. Kim, J.-S.; Nitta, T.; Mohuczy, D.; O’Malley, K.A.; Moldawer, L.L.; Dunn, W.A.; Behrns, K.E. Impaired Autophagy: A Mechanism of Mitochondrial Dysfunction in Anoxic Rat Hepatocytes. Hepatology 2008, 47, 1725–1736. [Google Scholar] [CrossRef] [Green Version]
  113. Madeo, F.; Eisenberg, T.; Pietrocola, F.; Kroemer, G. Spermidine in Health and Disease. Science 2018, 359, eaan2788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. D’Amico, D.; Olmer, M.; Fouassier, A.M.; Valdés, P.; Andreux, P.A.; Rinsch, C.; Lotz, M. Urolithin A Improves Mitochondrial Health, Reduces Cartilage Degeneration, and Alleviates Pain in Osteoarthritis. Aging Cell 2022, 21, e13662. [Google Scholar] [CrossRef] [PubMed]
  115. Cai, J.; Chen, X.; Liu, X.; Li, Z.; Shi, A.; Tang, X.; Xia, P.; Zhang, J.; Yu, P. AMPK: The Key to Ischemia-reperfusion Injury. J. Cell. Physiol. 2022, 237, 4079–4096. [Google Scholar] [CrossRef] [PubMed]
  116. Iorio, R.; Celenza, G.; Petricca, S. Mitophagy: Molecular Mechanisms, New Concepts on Parkin Activation and the Emerging Role of AMPK/ULK1 Axis. Cells 2021, 11, 30. [Google Scholar] [CrossRef]
  117. Gu, J.; Zhang, T.; Guo, J.; Chen, K.; Li, H.; Wang, J. PINK1 Activation and Translocation to Mitochondria-Associated Membranes Mediates Mitophagy and Protects against Hepatic Ischemia/Reperfusion Injury. Shock 2020, 54, 783–793. [Google Scholar] [CrossRef]
  118. Ning, X.; Yan, X.; Wang, Y.; Wang, R.; Fan, X.; Zhong, Z.; Ye, Q. Parkin Deficiency Elevates Hepatic Ischemia/Reperfusion Injury Accompanying Decreased Mitochondrial Autophagy, Increased Apoptosis, Impaired DNA Damage Repair and Altered Cell Cycle Distribution. Mol. Med. Rep. 2018, 18, 5663–5668. [Google Scholar] [CrossRef] [Green Version]
  119. Northup, P.G.; Intagliata, N.M.; Shah, N.L.; Pelletier, S.J.; Berg, C.L.; Argo, C.K. Excess Mortality on the Liver Transplant Waiting List: Unintended Policy Consequences and Model for End-Stage Liver Disease (MELD) Inflation. Hepatology 2015, 61, 285–291. [Google Scholar] [CrossRef]
  120. Evans, Z.P.; Ellett, J.D.; Schmidt, M.G.; Schnellmann, R.G.; Chavin, K.D. Mitochondrial Uncoupling Protein-2 Mediates Steatotic Liver Injury Following Ischemia/Reperfusion. J. Biol. Chem. 2008, 283, 8573–8579. [Google Scholar] [CrossRef] [Green Version]
  121. Evans, Z.P.; Palanisamy, A.P.; Sutter, A.G.; Ellett, J.D.; Ramshesh, V.K.; Attaway, H.; Schmidt, M.G.; Schnellmann, R.G.; Chavin, K.D. Mitochondrial Uncoupling Protein-2 Deficiency Protects Steatotic Mouse Hepatocytes from Hypoxia/Reoxygenation. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 302, G336–G342. [Google Scholar] [CrossRef] [PubMed]
  122. Zhang, T.; Gu, J.; Guo, J.; Chen, K.; Li, H.; Wang, J. Renalase Attenuates Mouse Fatty Liver Ischemia/Reperfusion Injury through Mitigating Oxidative Stress and Mitochondrial Damage via Activating SIRT1. Oxid. Med. Cell. Longev. 2019, 2019, 7534285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Hide, D.; Warren, A.; Fernández-Iglesias, A.; Maeso-Díaz, R.; Peralta, C.; Couteur, D.G.L.; Bosch, J.; Cogger, V.C.; Gracia-Sancho, J. Ischemia/Reperfusion Injury in the Aged Liver: The Importance of the Sinusoidal Endothelium in Developing Therapeutic Strategies for the Elderly. J. Gerontol. A Biol. Sci. Med. Sci. 2019, 75, 268–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Selzner, M.; Selzner, N.; Jochum, W.; Graf, R.; Clavien, P.-A. Increased Ischemic Injury in Old Mouse Liver: An ATP-Dependent Mechanism. Liver Transpl. 2007, 13, 382–390. [Google Scholar] [CrossRef] [PubMed]
  125. Chun, S.K.; Lee, S.; Flores-Toro, J.; U, R.Y.; Yang, M.-J.; Go, K.L.; Biel, T.G.; Miney, C.E.; Pierre Louis, S.; Law, B.K.; et al. Loss of Sirtuin 1 and Mitofusin 2 Contributes to Enhanced Ischemia/Reperfusion Injury in Aged Livers. Aging Cell 2018, 17, e12761. [Google Scholar] [CrossRef]
  126. Flores-Toro, J.; Chun, S.-K.; Shin, J.-K.; Campbell, J.; Lichtenberger, M.; Chapman, W.; Zendejas, I.; Behrns, K.; Leeuwenburgh, C.; Kim, J.-S. Critical Roles of Calpastatin in Ischemia/Reperfusion Injury in Aged Livers. Cells 2021, 10, 1863. [Google Scholar] [CrossRef] [PubMed]
  127. Jiang, T.; Zhan, F.; Rao, Z.; Pan, X.; Zhong, W.; Sun, Y.; Wang, P.; Lu, L.; Zhou, H.; Wang, X. Combined Ischemic and Rapamycin Preconditioning Alleviated Liver Ischemia and Reperfusion Injury by Restoring Autophagy in Aged Mice. Int. Immunopharmacol. 2019, 74, 105711. [Google Scholar] [CrossRef]
  128. Liu, A.; Yang, J.; Hu, Q.; Dirsch, O.; Dahmen, U.; Zhang, C.; Gewirtz, D.A.; Fang, H.; Sun, J. Young Plasma Attenuates Age-dependent Liver Ischemia Reperfusion Injury. FASEB J. 2019, 33, 3063–3073. [Google Scholar] [CrossRef]
  129. Pi-Sunyer, X. The Medical Risks of Obesity. Postgrad. Med. 2009, 121, 21–33. [Google Scholar] [CrossRef]
  130. Brunt, E.M.; Wong, V.W.-S.; Nobili, V.; Day, C.P.; Sookoian, S.; Maher, J.J.; Bugianesi, E.; Sirlin, C.B.; Neuschwander-Tetri, B.A.; Rinella, M.E. Nonalcoholic Fatty Liver Disease. Nat. Rev. Dis. Primers 2015, 1, 15080. [Google Scholar] [CrossRef]
  131. Basaranoglu, M.; Neuschwander-Tetri, B.A. Nonalcoholic Fatty Liver Disease: Clinical Features and Pathogenesis. Gastroenterol. Hepatol. 2006, 2, 282–291. [Google Scholar]
  132. An, P.; Wei, L.-L.; Zhao, S.; Sverdlov, D.Y.; Vaid, K.A.; Miyamoto, M.; Kuramitsu, K.; Lai, M.; Popov, Y.V. Hepatocyte Mitochondria-Derived Danger Signals Directly Activate Hepatic Stellate Cells and Drive Progression of Liver Fibrosis. Nat. Commun. 2020, 11, 2362. [Google Scholar] [CrossRef]
  133. Kaufmann, B.; Reca, A.; Wang, B.; Friess, H.; Feldstein, A.E.; Hartmann, D. Mechanisms of Nonalcoholic Fatty Liver Disease and Implications for Surgery. Langenbecks Arch. Surg. 2021, 406, 1–17. [Google Scholar] [CrossRef]
  134. Orci, L.A.; Lacotte, S.; Oldani, G.; Slits, F.; De Vito, C.; Crowe, L.A.; Rubbia-Brandt, L.; Vallée, J.-P.; Morel, P.; Toso, C. Effect of Ischaemic Preconditioning on Recurrence of Hepatocellular Carcinoma in an Experimental Model of Liver Steatosis. Br. J. Surg. 2016, 103, 417–426. [Google Scholar] [CrossRef] [PubMed]
  135. Yang, F.; Zhang, Y.; Ren, H.; Wang, J.; Shang, L.; Liu, Y.; Zhu, W.; Shi, X. Ischemia Reperfusion Injury Promotes Recurrence of Hepatocellular Carcinoma in Fatty Liver via ALOX12-12HETE-GPR31 Signaling Axis. J. Exp. Clin. Cancer. Res. 2019, 38, 489. [Google Scholar] [CrossRef] [Green Version]
  136. Grąt, M.; Krawczyk, M.; Wronka, K.M.; Stypułkowski, J.; Lewandowski, Z.; Wasilewicz, M.; Krawczyk, P.; Grąt, K.; Patkowski, W.; Zieniewicz, K. Ischemia-Reperfusion Injury and the Risk of Hepatocellular Carcinoma Recurrence after Deceased Donor Liver Transplantation. Sci. Rep. 2018, 8, 8935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Tang, Y.; Wang, T.; Ju, W.; Li, F.; Zhang, Q.; Chen, Z.; Gong, J.; Zhao, Q.; Wang, D.; Chen, M.; et al. Ischemic-Free Liver Transplantation Reduces the Recurrence of Hepatocellular Carcinoma after Liver Transplantation. Front. Oncol. 2021, 11, 773535. [Google Scholar] [CrossRef]
  138. Man, K.; Ng, K.T.; Lo, C.M.; Ho, J.W.; Sun, B.S.; Sun, C.K.; Lee, T.K.; Poon, R.T.P.; Fan, S.T. Ischemia-Reperfusion of Small Liver Remnant Promotes Liver Tumor Growth and Metastases—Activation of Cell Invasion and Migration Pathways. Liver. Transpl. 2007, 13, 1669–1677. [Google Scholar] [CrossRef] [PubMed]
  139. Hamaguchi, Y.; Mori, A.; Fujimoto, Y.; Ito, T.; Iida, T.; Yagi, S.; Okajima, H.; Kaido, T.; Uemoto, S. Longer Warm Ischemia Can Accelerate Tumor Growth through the Induction of HIF-1α and the IL-6-JAK-STAT3 Signaling Pathway in a Rat Hepatocellular Carcinoma Model: Longer Warm Ischemia Can Accelerate Tumor Growth through the Induction of HIF-1α and the IL-6-JAK-STAT3 Signaling Pathway in a Rat HCC Model. J. Hepatobiliary. Pancreat. Sci. 2016, 23, 771–779. [Google Scholar] [CrossRef] [PubMed]
  140. Ng, K.T.-P.; Yeung, O.W.-H.; Lam, Y.F.; Liu, J.; Liu, H.; Pang, L.; Yang, X.X.; Zhu, J.; Zhang, W.; Lau, M.Y.H.; et al. Glutathione S-Transferase A2 Promotes Hepatocellular Carcinoma Recurrence after Liver Transplantation through Modulating Reactive Oxygen Species Metabolism. Cell Death Discov. 2021, 7, 188. [Google Scholar] [CrossRef]
  141. Bian, J.; Zhang, D.; Wang, Y.; Qin, H.; Yang, W.; Cui, R.; Sheng, J. Mitochondrial Quality Control in Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 713721. [Google Scholar] [CrossRef]
  142. Zhou, J.; Feng, J.; Wu, Y.; Dai, H.-Q.; Zhu, G.-Z.; Chen, P.-H.; Wang, L.-M.; Lu, G.; Liao, X.-W.; Lu, P.-Z.; et al. Simultaneous Treatment with Sorafenib and Glucose Restriction Inhibits Hepatocellular Carcinoma in Vitro and in Vivo by Impairing SIAH1-Mediated Mitophagy. Exp. Mol. Med. 2022, 54, 2007–2021. [Google Scholar] [CrossRef] [PubMed]
  143. Selzner, M.; RüDiger, H.A.; Sindram, D.; Madden, J.; Clavien, P.-A. Mechanisms of Ischemic Injury Are Different in the Steatotic and Normal Rat Liver. Hepatology 2000, 32, 1280–1288. [Google Scholar] [CrossRef]
  144. Nassir, F.; Ibdah, J. Role of Mitochondria in Nonalcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2014, 15, 8713–8742. [Google Scholar] [CrossRef] [Green Version]
  145. Xu, J.; Shen, J.; Yuan, R.; Jia, B.; Zhang, Y.; Wang, S.; Zhang, Y.; Liu, M.; Wang, T. Mitochondrial Targeting Therapeutics: Promising Role of Natural Products in Non-Alcoholic Fatty Liver Disease. Front. Pharmacol. 2021, 12, 796207. [Google Scholar] [CrossRef]
  146. Indo, H.P.; Yen, H.-C.; Nakanishi, I.; Matsumoto, K.; Tamura, M.; Nagano, Y.; Matsui, H.; Gusev, O.; Cornette, R.; Okuda, T.; et al. A Mitochondrial Superoxide Theory for Oxidative Stress Diseases and Aging. J. Clin. Biochem. Nutr. 2015, 56, 1–7. [Google Scholar] [CrossRef] [Green Version]
  147. Caraceni, P.; Bianchi, C.; Domenicali, M.; Maria Pertosa, A.; Maiolini, E.; Parenti Castelli, G.; Nardo, B.; Trevisani, F.; Lenaz, G.; Bernardi, M. Impairment of Mitochondrial Oxidative Phosphorylation in Rat Fatty Liver Exposed to Preservation-Reperfusion Injury. J. Hepatol. 2004, 41, 82–88. [Google Scholar] [CrossRef] [PubMed]
  148. Chu, M.J.; Premkumar, R.; Hickey, A.J.; Jiang, Y.; Delahunt, B.; Phillips, A.R.; Bartlett, A.S. Steatotic Livers Are Susceptible to Normothermic Ischemia-Reperfusion Injury from Mitochondrial Complex-I Dysfunction. World J. Gastroenterol. 2016, 22, 4673. [Google Scholar] [CrossRef] [PubMed]
  149. Murphy, M.P. How Mitochondria Produce Reactive Oxygen Species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [Green Version]
  150. Toda, C.; Diano, S. Mitochondrial UCP2 in the Central Regulation of Metabolism. Best Pract. Res. Clin. Endocrinol. Metab. 2014, 28, 757–764. [Google Scholar] [CrossRef]
  151. Li, S.; Fujino, M.; Takahara, T.; Li, X.-K. Protective Role of Heme Oxygenase-1 in Fatty Liver Ischemia–Reperfusion Injury. Med. Mol. Morphol. 2019, 52, 61–72. [Google Scholar] [CrossRef] [Green Version]
  152. Longo, V.D.; Anderson, R.M. Nutrition, Longevity and Disease: From Molecular Mechanisms to Interventions. Cell 2022, 185, 1455–1470. [Google Scholar] [CrossRef] [PubMed]
  153. Green, C.L.; Lamming, D.W.; Fontana, L. Molecular Mechanisms of Dietary Restriction Promoting Health and Longevity. Nat. Rev. Mol. Cell Biol. 2022, 23, 56–73. [Google Scholar] [CrossRef] [PubMed]
  154. Partridge, L.; Fuentealba, M.; Kennedy, B.K. The Quest to Slow Ageing through Drug Discovery. Nat. Rev. Drug. Discov. 2020, 19, 513–532. [Google Scholar] [CrossRef] [PubMed]
  155. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The Hallmarks of Aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. Hallmarks of Aging: An Expanding Universe. Cell 2023, 186, S0092867422013770. [Google Scholar] [CrossRef]
  157. Kitada, M.; Koya, D. Autophagy in Metabolic Disease and Ageing. Nat. Rev. Endocrinol. 2021, 17, 647–661. [Google Scholar] [CrossRef]
  158. Finkel, T. The Metabolic Regulation of Aging. Nat. Med. 2015, 21, 1416–1423. [Google Scholar] [CrossRef]
  159. Lima, T.; Li, T.Y.; Mottis, A.; Auwerx, J. Pleiotropic Effects of Mitochondria in Aging. Nat. Aging 2022, 2, 199–213. [Google Scholar] [CrossRef]
  160. Hoare, M.; Das, T.; Alexander, G. Ageing, Telomeres, Senescence, and Liver Injury. J. Hepatol. 2010, 53, 950–961. [Google Scholar] [CrossRef] [Green Version]
  161. Hunt, N.J.; Kang, S.W.; Lockwood, G.P.; Le Couteur, D.G.; Cogger, V.C. Hallmarks of Aging in the Liver. Comput. Struct. Biotechnol. J. 2019, 17, 1151–1161. [Google Scholar] [CrossRef]
  162. Kan, C.; Ungelenk, L.; Lupp, A.; Dirsch, O.; Dahmen, U. Ischemia-Reperfusion Injury in Aged Livers—The Energy Metabolism, Inflammatory Response, and Autophagy. Transplantation 2018, 102, 368–377. [Google Scholar] [CrossRef]
  163. Fortner, J.G.; Lincer, R.M. Hepatic Resection in the Elderly. Ann. Surg. 1990, 211, 141–145. [Google Scholar] [CrossRef] [PubMed]
  164. Yanaga, K.; Kanematsu, T.; Takenaka, K.; Matsumata, T.; Yoshida, Y.; Sugimachi, K. Hepatic Resection for Hepatocellular Carcinoma in Elderly Patients. Am. J. Surg. 1988, 155, 238–241. [Google Scholar] [CrossRef] [PubMed]
  165. Busquets, J.; Xiol, X.; Figueras, J.; Jaurrieta, E.; Torras, J.; Ramos, E.; Rafecas, A.; Fabregat, J.; Lama, C.; Ibañez, L.; et al. The Impact of Donor Age on Liver Transplantation: Influence of Donor Age on Early Liver Function and on Subsequent Patient and Graft Survival. Transplantation 2001, 71, 1765–1771. [Google Scholar] [CrossRef]
  166. Lai, Q.; Melandro, F.; Sandri, G.B.L.; Mennini, G.; Corradini, S.G.; Merli, M.; Berloco, P.B.; Rossi, M. Use of Elderly Donors for Liver Transplantation: Has the Limit Been Reached? J. Gastrointest. Liver Dis. 2011, 20, 383–387. [Google Scholar]
  167. Ghanie, A.; Formica, M.K.; Dhir, M. Systematic Review and Meta-Analysis of 90-Day and 30-Day Mortality after Liver Resection in the Elderly. Surgery 2022, 172, 1164–1173. [Google Scholar] [CrossRef]
  168. Andert, A.; Lodewick, T.; Ulmer, T.F.; Schmeding, M.; Schöning, W.; Neumann, U.; Dejong, K.; Heidenhain, C. Liver Resection in the Elderly: A Retrospective Cohort Study of 460 Patients—Feasible and Safe. Int. J. Surg. 2016, 28, 126–130. [Google Scholar] [CrossRef]
  169. Kim, J.-S.; Chapman, W.C.; Lin, Y. Mitochondrial Autophagy in Ischemic Aged Livers. Cells 2022, 11, 4083. [Google Scholar] [CrossRef]
  170. Mizuguchi, T.; Kawamoto, M.; Meguro, M.; Okita, K.; Ota, S.; Ishii, M.; Ueki, T.; Nishidate, T.; Kimura, Y.; Furuhata, T.; et al. The Impact of Aging on Morbidity and Mortality after Liver Resection: A Systematic Review and Meta-Analysis. Surg. Today 2015, 45, 259–270. [Google Scholar] [CrossRef]
  171. Tzeng, C.-W.D.; Cooper, A.B.; Vauthey, J.-N.; Curley, S.A.; Aloia, T.A. Predictors of Morbidity and Mortality after Hepatectomy in Elderly Patients: Analysis of 7621 NSQIP Patients. HPB 2014, 16, 459–468. [Google Scholar] [CrossRef] [Green Version]
  172. Dickson, K.M.; Martins, P.N. Implications of Liver Donor Age on Ischemia Reperfusion Injury and Clinical Outcomes. Transplant. Rev. 2020, 34, 100549. [Google Scholar] [CrossRef] [PubMed]
  173. Park, Y.; Hirose, R.; Coatney, J.L.; Ferrell, L.; Behrends, M.; Roberts, J.P.; Serkova, N.J.; Niemann, C.U. Ischemia-Reperfusion Injury Is More Severe in Older versus Young Rat Livers. J. Surg. Res. 2007, 137, 96–102. [Google Scholar] [CrossRef]
  174. Okaya, T.; Blanchard, J.; Schuster, R.; Kuboki, S.; Husted, T.; Caldwell, C.C.; Zingarelli, B.; Wong, H.; Solomkin, J.S.; Lentsch, A.B. Age-Dependent Responses to Hepatic Ischemia/Reperfusion Injury. Shock 2005, 24, 421–427. [Google Scholar] [CrossRef]
  175. Brandt, T.; Mourier, A.; Tain, L.S.; Partridge, L.; Larsson, N.-G.; Kühlbrandt, W. Changes of Mitochondrial Ultrastructure and Function during Ageing in Mice and Drosophila. eLife 2017, 6, e24662. [Google Scholar] [CrossRef] [PubMed]
  176. Lesnefsky, E.J.; Hoppel, C.L. Oxidative Phosphorylation and Aging. Ageing Res. Rev. 2006, 5, 402–433. [Google Scholar] [CrossRef] [PubMed]
  177. Li, Y.; Ruan, D.; Jia, C.; Zheng, J.; Wang, G.; Zhao, H.; Yang, Q.; Liu, W.; Yi, S.; Li, H.; et al. Aging Aggravates Hepatic Ischemia-Reperfusion Injury in Mice by Impairing Mitophagy with the Involvement of the EIF2α-Parkin Pathway. Aging 2018, 10, 1902–1920. [Google Scholar] [CrossRef]
  178. Li, Q.; Gao, S.; Kang, Z.; Zhang, M.; Zhao, X.; Zhai, Y.; Huang, J.; Yang, G.-Y.; Sun, W.; Wang, J. Rapamycin Enhances Mitophagy and Attenuates Apoptosis after Spinal Ischemia-Reperfusion Injury. Front. Neurosci. 2018, 12, 865. [Google Scholar] [CrossRef]
  179. Selvarani, R.; Mohammed, S.; Richardson, A. Effect of Rapamycin on Aging and Age-Related Diseases—Past and Future. GeroScience 2021, 43, 1135–1158. [Google Scholar] [CrossRef]
  180. Bitto, A.; Kaeberlein, M. Rejuvenation: It’s in Our Blood. Cell Metab. 2014, 20, 2–4. [Google Scholar] [CrossRef] [Green Version]
  181. Villeda, S.A.; Plambeck, K.E.; Middeldorp, J.; Castellano, J.M.; Mosher, K.I.; Luo, J.; Smith, L.K.; Bieri, G.; Lin, K.; Berdnik, D.; et al. Young Blood Reverses Age-Related Impairments in Cognitive Function and Synaptic Plasticity in Mice. Nat. Med. 2014, 20, 659–663. [Google Scholar] [CrossRef] [Green Version]
  182. Ferrucci, L.; Fabbri, E. Inflammageing: Chronic Inflammation in Ageing, Cardiovascular Disease, and Frailty. Nat. Rev. Cardiol. 2018, 15, 505–522. [Google Scholar] [CrossRef] [PubMed]
  183. Zhong, W.; Rao, Z.; Rao, J.; Han, G.; Wang, P.; Jiang, T.; Pan, X.; Zhou, S.; Zhou, H.; Wang, X. Aging Aggravated Liver Ischemia and Reperfusion Injury by Promoting STING-mediated NLRP3 Activation in Macrophages. Aging Cell 2020, 19, e13186. [Google Scholar] [CrossRef] [PubMed]
  184. Lin, L.; Xu, H.; Bishawi, M.; Feng, F.; Samy, K.; Truskey, G.; Barbas, A.S.; Kirk, A.D.; Brennan, T.V. Circulating Mitochondria in Organ Donors Promote Allograft Rejection. Am. J. Transplant. 2019, 19, 1917–1929. [Google Scholar] [CrossRef] [PubMed]
  185. Pollara, J.; Edwards, R.W.; Lin, L.; Bendersky, V.A.; Brennan, T.V. Circulating Mitochondria in Deceased Organ Donors Are Associated with Immune Activation and Early Allograft Dysfunction. JCI Insight 2018, 3, e121622. [Google Scholar] [CrossRef]
  186. Saeb-Parsy, K.; Martin, J.L.; Summers, D.M.; Watson, C.J.E.; Krieg, T.; Murphy, M.P. Mitochondria as Therapeutic Targets in Transplantation. Trends Mol. Med. 2021, 27, 185–198. [Google Scholar] [CrossRef]
  187. Schlegel, A.; Muller, X.; Mueller, M.; Stepanova, A.; Kron, P.; de Rougemont, O.; Muiesan, P.; Clavien, P.-A.; Galkin, A.; Meierhofer, D.; et al. Hypothermic Oxygenated Perfusion Protects from Mitochondrial Injury before Liver Transplantation. EBioMedicine 2020, 60, 103014. [Google Scholar] [CrossRef]
  188. Meszaros, A.T.; Hofmann, J.; Buch, M.L.; Cardini, B.; Dunzendorfer-Matt, T.; Nardin, F.; Blumer, M.J.; Fodor, M.; Hermann, M.; Zelger, B.; et al. Mitochondrial Respiration during Normothermic Liver Machine Perfusion Predicts Clinical Outcome. eBioMedicine 2022, 85, 104311. [Google Scholar] [CrossRef]
  189. Martins, R.M.; Teodoro, J.S.; Furtado, E.; Rolo, A.P.; Palmeira, C.M.; Tralhão, J.G. Evaluation of Bioenergetic and Mitochondrial Function in Liver Transplantation. Clin. Mol. Hepatol. 2019, 25, 190–198. [Google Scholar] [CrossRef] [Green Version]
Table 1. Interventions that ameliorate or protect against liver ischemia/reperfusion injury (LIRI) by interfering with mitochondrial quality control mechanisms.
Table 1. Interventions that ameliorate or protect against liver ischemia/reperfusion injury (LIRI) by interfering with mitochondrial quality control mechanisms.
InterventionSpecies/ModelMechanismEffect on MitochondriaReference
Mitochondrial biogenesis
BerberineWistar rats↑ ATP content
↓ ROS
↑ Mitochondrial biogenesis markers
Martins et al. [54]
CilostazolMice
HepG2 cells
Activation of HO-1 and Nrf2↑ Mitochondrial biogenesis markers
↑ Mitochondrial function
Joe et al. [57]
HeminMiceActivation of HO-1↑ Mitochondrial biogenesis markers
↑ Fission/fusion markers
↑ Mitophagy
Hong et al. [58]
GenipinMiceActivation of AMPK and SIRT1↑ Mitochondrial biogenesis markers
↑ Fission/fusion markers
↑ Mitophagy
Shin et al. [50]
SRT1720MiceActivation of SIRT1↑ Mitochondrial biogenesis markers
↑ Mitochondrial function
↑ Mitochondrial mass
Khader et al. [53]
IsolongifoleneMiceActivation of AMPK and PGC1α↓ Oxidative stressLi et al. [59]
Sirt1 overexpressionMice (Ad-SIRT1)
Primary hepatocytes
Interaction between SIRT1 and MFN2↓ MPT
↓ Mitochondrial dysfunction
Biel et al. [60]
Mitochondrial fission/fusion
Alr overexpressionMice (Ad-HSS)
BEL-7402 cells (Ad-HSS)
↑ Mitochondrial function
↓ Mitochondrial ROS
↓ Mitochondrial-related apoptosis
Jiang et al. [61]
Alr overexpressionHSS+/– mice
HepG2 cells
Translocation and activation of DRP1↑ Fission markersZhang et al. [62]
IrisinMice
HL-7702 cells
Inhibition of excessive fission (through inhibition of DRP1 and FIS1)↑ Mitochondrial biogenesis markers
↑ Mitochondrial content
↓ Oxidative stress
Bi et al. [40]
Alr silencingALR+/– miceDRP1 SUMOylation and recruitment to mitochondria↓ Mitochondrial fissionHuang et al. [63]
Fto overexpressionMice (Ad-FTO)Inhibition of DRP1↓ Mitochondrial fragmentation
↓ Oxidative stress
Du et al. [64]
Tlr4 silencingTLR4-KO miceActivation of IL6 and TNFα pathways↑ Mitochondrial biogenesis markers
↑ Mitochondrial fusion markers
↓ ROS
Zhang et al. [65]
SilibininWistar rats↑ Mitochondrial fusion markersQajari et al. [66]
Mitophagy
2-MethoxyestradiolAFL miceActivation of SIRT1↑ MitophagyCho et al. [67]
AICARdb/db mice ↑ MitophagyZhijun et al. [68]
UC-MSC transfusionMice
L02 hepatocytes
Activation of AMPKα↓ Mitochondrial ROS
↑ Mitophagy
Zheng et al. [69]
PterostilbeneMice
L02 hepatocytes
Activation of PINK1↓ Mitochondrial dysfunction
↓ Mitochondrial ROS
↑ Mitophagy
Shi et al. [70]
25-HydroxycholesterolSprague Dawley ratsActivation of PINK1/Parkin pathway↑ MitophagyCao et al. [71]
Resolvin D1MiceActivation of TRX2↑ Mitophagy markers
↓ Mitochondrial swelling
↓ Oxidative stress
↑ Mitochondrial biogenesis markers
↑ Mitochondrial fission markers
Kang et al. [72]
CHOP silencingCHOP-KO miceActivation of DRP1-Beclin1 pathway↓ ROS
↑ Mitophagy
Zhou et al. [73]
Alr overexpressionBrown-Norway rats (Ad-ALR)Activation of MFN2↑ Mitochondrial function
↑ Mitophagy
Kong et al. [74]
↑, increase; ↓, decrease; Ad, adenovirus; AFL, alcoholic fatty liver; AICAR, 5-aminoimidazole-4-carboxamide ribonucleotide; ALR, HSS, augmenter of liver regeneration; AMPK, AMP-activated protein kinase; ATP, adenosine triphosphate; CHOP, CCAAT/enhancer-binding protein homologous protein; DRP1, dynamin-related protein 1; FIS1, mitochondrial fission 1 protein; FTO, fat mass and obesity associated; HO-1, heme oxygenase-1; IL6, interleukin 6; KO, knockout; MEF, mouse embryonic fibroblasts; MFN2, mitofusins 2; MPT, mitochondrial permeability transition; Nrf2, nuclear factor erythroid 2-related factor 2; PGC-1α, Peroxisome proliferator-activated receptor gamma coactivator 1-alpha; PINK1, phosphatase and tensin homologue-induced putative kinase 1; ROS, reactive oxygen species; SIRT1, sirtuin 1; TLR4, toll-like receptor 4; TNFα, tumor necrosis factor α; TRX2, thioredoxin 2; UC-MSC, umbilical cord-derived mesenchymal stem cell.
Table 2. Interventions that ameliorate or protect against liver ischemia/reperfusion injury (LIRI) during non-alcoholic fatty liver disease (NAFLD) and aging.
Table 2. Interventions that ameliorate or protect against liver ischemia/reperfusion injury (LIRI) during non-alcoholic fatty liver disease (NAFLD) and aging.
InterventionSpecies/ModelMechanism of ActionReferences
NAFLD
Ucp2 silencingUCP2-KO ob/ob miceIncreased ATP levels
Increased mice survival following I/R
Evans et al. [120]
Ucp2 silencingUCP2-KO primary hepatocytesIncreased cellular viability
Increased ATP levels
Increased mitochondrial membrane potential
Evans et al. [121]
RenalaseHFD mice
HepG2
Increased NAD+ levels and activation of SIRT1
Decreased ROS production
Increased mitochondrial function
Zhang et al. [122]
Aging
SimvastatinWistar ratsInhibition of HMG-CoA
Decreased hepatocellular damage Decreased oxidative stress
Hide et al. [123]
Ischemic and glucose preconditioningMiceIncreased ATP levelsSelzner et al. [124]
Sirt1 and Mfn2 overexpressionMice
Primary hepatocytes
Prevention of mitochondrial dysfunction
Prevention of MPT pore opening
Increased mitophagy
Prevention of cell death
Chun et al. [125]
Calpastatin overexpressionMice
Primary hepatocytes
Mitochondrial elongation
Prevention of MPT pore opening
Prevention of mitochondrial depolarization
Prevention of necrosis
Flores-Toro et al. [126]
Ischemic and rapamycin preconditioningMiceIncreased autophagyJiang et al. [127]
Plasma from young miceSprague Dawley rats
Primary hepatocytes
Activation of AMPK/ULK1 pathway
Increased autophagy
Liu et al. [128]
AMPK, AMP-activated protein kinase; ATP, adenosine triphosphate; HFD, high fat diet; HMG-CoA, hydroxy-methylglutaryl coenzyme A; I/R, ischemia/reperfusion; KO, knockout; MFN2, mitofusin-2; MPT, mitochondrial permeability transition; NAD, nicotinamide adenine dinucleotide; NAFLD, non-alcoholic fatty liver disease; ROS, reactive oxygen species; SIRT1, sirtuin 1; UCP2, uncoupling protein 2; ULK1, Unc-51 like autophagy activating kinase.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Machado, I.F.; Palmeira, C.M.; Rolo, A.P. Preservation of Mitochondrial Health in Liver Ischemia/Reperfusion Injury. Biomedicines 2023, 11, 948. https://doi.org/10.3390/biomedicines11030948

AMA Style

Machado IF, Palmeira CM, Rolo AP. Preservation of Mitochondrial Health in Liver Ischemia/Reperfusion Injury. Biomedicines. 2023; 11(3):948. https://doi.org/10.3390/biomedicines11030948

Chicago/Turabian Style

Machado, Ivo F., Carlos M. Palmeira, and Anabela P. Rolo. 2023. "Preservation of Mitochondrial Health in Liver Ischemia/Reperfusion Injury" Biomedicines 11, no. 3: 948. https://doi.org/10.3390/biomedicines11030948

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