Skip to main content

Mitochondrial quality control in human health and disease

Abstract

Mitochondria, the most crucial energy-generating organelles in eukaryotic cells, play a pivotal role in regulating energy metabolism. However, their significance extends beyond this, as they are also indispensable in vital life processes such as cell proliferation, differentiation, immune responses, and redox balance. In response to various physiological signals or external stimuli, a sophisticated mitochondrial quality control (MQC) mechanism has evolved, encompassing key processes like mitochondrial biogenesis, mitochondrial dynamics, and mitophagy, which have garnered increasing attention from researchers to unveil their specific molecular mechanisms. In this review, we present a comprehensive summary of the primary mechanisms and functions of key regulators involved in major components of MQC. Furthermore, the critical physiological functions regulated by MQC and its diverse roles in the progression of various systemic diseases have been described in detail. We also discuss agonists or antagonists targeting MQC, aiming to explore potential therapeutic and research prospects by enhancing MQC to stabilize mitochondrial function.

Background

Mitochondria are organelles found within eukaryotic cells that play a crucial role in energy production via oxidative phosphorylation (OXPHOS), generating adenosine triphosphate (ATP) [1]. Beyond their pivotal function as “the powerhouse of the cell”, mitochondria also contribute to the regulation of several cellular processes, including fatty acid oxidation (FAO), calcium buffering, phospholipid synthesis, iron-sulfur cluster biosynthesis, innate immune signaling, and cell death [2, 3]. Despite their importance in cell homeostasis, the process of OXPHOS generates reactive oxygen species (ROS) as a by-product, which can damage mitochondrial proteins, lipids, and DNA. Compounding this inherent challenge, mitochondria are also continually exposed to various environmental stressors [4], which augment their vulnerability to dysregulation. To combat this threat, eukaryotic cells have developed an intricate set of mitochondrial quality control (MQC) mechanisms to diligently monitor and maintain the integrity and function of the intracellular mitochondrial network.

MQC is a multifaceted group of processes that serve to safeguard the mitochondria against damage and prevent the accumulation of defective mitochondria. Those processes mainly involve three distinct mechanisms, namely mitochondrial biogenesis, mitochondrial dynamics (fusion and fission), and mitophagy [5]. Mitochondrial biogenesis is a tightly regulated process that involves the coordinated expression of nuclear and mitochondrial genes to bolster the size and quantity of mitochondria. Meanwhile, mitochondrial dynamics maintain mitochondrial health by continuously shifting between fusion and fission, enabling the elimination of unhealthy mitochondria to prevent their accumulation [6]. In cases where mild to moderate damage occurs, mitochondria can compensate for their loss of function by fusing with healthy mitochondria or undergoing fission to remove harmful components. However, when damage is too severe, mitophagy is required to selectively remove hypofunctional or damaged mitochondria [7]. The components of degraded mitochondria are subsequently renewed by protein and mitochondrial biogenesis. When the primary damage surpasses the capacity of MQC, the mitochondria unleash danger signals that can prompt cell-death decisions, ultimately resulting in tissue injury and potential organ failure. Therefore, the preservation of mitochondrial homeostasis is of paramount importance in ensuring cellular survival and overall organismal well-being.

In this review, we aim to clarify the current state of knowledge about MQC in human health and disease. We will delve into the intricate molecular mechanisms underlying mitochondrial biogenesis, mitochondrial dynamics, and mitophagy, along with their regulation. Moreover, we will thoroughly investigate the crucial role of MQC in a wide range of human diseases, including cancer, cardiovascular disease, metabolic disorders, neurological disorders, respiratory diseases, renal diseases, and digestive diseases. By deepening our comprehension of the complex mechanisms involved in MQC, we can devise innovative strategies to mitigate mitochondrial dysfunction and associated pathologies.

Molecular regulations of MQC

Mitochondrial morphology, structures, mass, and even quantity are highly plastic in response to signals generated by normal physiological activities or pathological stimuli. To sustain cellular function and homeostasis, complex and highly dynamic MQC mechanisms have evolved under the regulation of the nuclear and mitochondrial genomes [8]. In general, the homeostasis of mitochondrial function and quality is achieved through two processes: the elimination of damaged mitochondria and the synthesis of new intact mitochondria [9]. Following, we review several important mechanisms of MQC and recent advances including mitochondrial biogenesis, mitochondrial dynamics, and mitophagy (Fig. 1).

Fig. 1
figure 1

Molecular regulation of Mitochondrial quality control. a PGC-1α plays a central role in mitochondrial biogenesis. Several regulators, including AMPK, Sirts, and Ca2+, are involved in the regulation of PCG-1α expression and activity. In addition, PGC-1β is also involved in the mitochondrial biogenesis process. b Mitochondrial dynamics consists of fission and fusion. The fission-associated proteins (DRP1, FIS1, MFF, et al.) mediate the fission of mitochondria, a process receiving complex regulation by various factors such as endoplasmic reticulum and multiple kinases. The fusion event consists of mitochondrial outer membrane fusion mediated by MFNs and mitochondrial inner membrane fusion mediated by OPA1, similarly, they are subject to complex regulation at different stages of fusion. c The role of mitophagy is to remove damaged mitochondria promptly. There are two pathways of mitophagy, respectively, the PINK1/Parkin-dependent pathway and the PINK1/Parkin-independent pathway. Their common feature is the formation of autophagosomes that enclose damaged mitochondria and the complex regulation by multiple intracellular signals. Moreover, the presence of protein quality control systems in mitochondria removes misfolded mitochondrial proteins, and the accumulated unfolded proteins will promote mitophagy. AMPK AMP-activated protein kinaseBNIP3 BCL2 interacting protein 3, CaMK calcium/calmodulin-dependent protein kinase, DeAc deactylation, DRP1 dynamin-related protein 1, ETC electron transport chain, FAO fatty acid oxidation, FIS1 fission protein 1, FUNDC1 FUN14 domain containing 1, INF2 inverted formin 2, JNK Jun N-terminal kinase, LC3 microtubule associated protein 1 light chain 3, LONP1 lon protease 1, MAPK mitogen-activated protein kinases, MARCH5 membrane associated ring-CH-type finger 5, MDV mitochondria derived vesicle, Met MET proto-oncogene, MFF mitochondrial fission factor, MFN mitofusin, MiD mitochondrial dynamics, NDP52 nuclear dots protein 52, Nix NIP3-like protein X, Nrf2 nuclear factor E2-related factor 2, OMA1 OMA1 zinc metallopeptidase, OPA1 optic atrophy 1, OPTN optineurin, PGC-1α PPAR-γ coactivator-1α, PINK1 PTEN-induced kinase 1, Sirt1 sirtuin 1, TFAM mitochondrial transcription factor A, TFB1M mitochondrial transcription factors B1, TFB2M mitochondrial transcription factors B1, TNF-α tumor necrosis factor-α, Tom20 translocase of outer mitochondrial membrane 20, TORC transducer of regulated CREB (cAMP response element-binding protein), ULK1 unc-51 like autophagy activating kinase 1, UPS ubiquitin–proteasome system, VDAC voltage dependent anion channel, YME1L YME1 like 1 ATPase

Mitochondrial biogenesis

Mitochondrial biogenesis was first found when the level of intracellular mitochondrial content was significantly increased during endurance training in the study of Holloszy et al. [10]. It can be understood as the growth and division of existing mitochondria in cells, which mainly requires the production of the mitochondrial inner and outer membrane, the replication of the mitochondrial genome, and the synthesis and assembly of mitochondrial proteins [11]. As stated, mitochondria are unique in that they have a separate extranuclear genome. Most of the mitochondrial proteins are encoded by nuclear genes, and mitochondrial DNA (mtDNA) is responsible for encoding 13 basic components of the electron transport chain (ETC), transfer RNA, and ribosomal RNA [12]. Therefore, a complex mechanism is necessary to coordinate the transcription and replication of the nuclear and mitochondrial genomes to guarantee dynamic and correct mitochondrial biogenesis, and a targeted delivery and assembly system is required for importing nuclear-encoded proteins to ensure proper mitochondrial morphology and function. The following text mainly elaborates on the regulatory factors and the latest upstream regulatory factors involved in the process of mitochondrial biogenesis.

Master transcriptional regulators of mitochondrial biogenesis

Peroxisome proliferator-activated receptor (PPAR)-γ coactivator-1α (PGC-1α)

It is the most important transcriptional cascade regulator in mitochondrial biogenesis, facilitating the transcription of mitochondrial proteins and the replication of mtDNA, which are key steps in mitochondrial biogenesis mentioned above [13]. PPARs are a superfamily of nuclear receptors that control the expression of related genes through the stimulation of different ligands to maintain intracellular energy homeostasis and manage oxidative stress levels [14]. PGC-1α was initially identified as a protein that interacts with PPARγ and enhances its transcriptional activity. It is noteworthy that upon binding to PPARs, particularly PPARα, PGC-1α further stimulates the generation of intracellular fatty acid transporters and the activation of mitochondrial FAO enzymes, leading to an upregulation of mitochondrial FAO process [15]. In recent years, many studies have shown that PGC-1α is not limited to activating PPARs but is the most crucial regulator in mitochondrial biogenesis, participating in the regulation of almost all processes related to mitochondrial biogenesis [16,17,18].

PGC-1α, as a potent transcriptional coactivator, interacts with various transcription factors without sequence-specific DNA binding to promote the expression of specific proteins, which is indispensable for mitochondrial biogenesis and regulation of cellular metabolism [19]. In tissues with abundant mitochondria and active oxidative metabolism, PGC-1α is expressed at high levels, and its expression is significantly increased by physiological signals such as exercise and fasting [20]. Upon signal stimulation, PGC-1α is activated and translocated into the nucleus, where it acts as a transcriptional coactivator to activate nuclear factor E2-related factor 2 (Nrf2) [21]. It not only directly promotes the expression of ROS scavenging enzymes, but also promotes the expression of mitochondrial-related proteins encoded by the nuclear genome induced by Nrf2. Nrf2, as well as its family member Nrf1, plays important roles in regulating mitochondrial gene replication and protein synthesis [22]. PGC-1α triggers the activation of Nrf1/2, which in turn directly boosts the expression of almost all components of the mitochondrial ETC. The PGC-1α-Nrf1 axis also stimulates the transcription of mitochondrial gene-encoded factors, including mitochondrial transcription factor B1 (TFB1M), TFB2M, and mitochondrial transcription factor A (TFAM), which mediate mtDNA replication and transcription [23]. Moreover, the expression of OXPHOS proteins also depends on PGC-1α promoting the binding of Nrf1/2 to the corresponding gene promoters, thus enhancing their transcription [24]. Additionally, the co-activation of PGC-1α and estrogen-related receptors (ERRs) is crucial for the process of mitochondrial biogenesis. The nuclear receptor family ERRs mainly includes ERRα, ERRβ, and ERRγ, which have significant homology with classic hormone receptors in the ligand-binding domain, but exert biological functions without requiring ligand. Currently, ERRα and ERRγ are known to play important regulatory roles in maintaining mitochondrial function [25]. Numerous studies have shown that PGC-1α activates ERRα to regulate the expression of key genes that are responsible for managing almost all energy transduction and ATP synthesis pathways in mitochondria, including FAO, the tricarboxylic acid (TCA) cycle, and OXPHOS [26, 27]. Furthermore, ERRγ is more likely to control ion channel proteins to affect the process of mitochondrial energy metabolism [28]. However, the function of ERRβ in mitochondrial biogenesis is still unknown. In summary, PGC-1α emerges as the pivotal regulator in orchestrating the transcriptional cascades of mitochondrial biogenesis and modulating the expression level or activation degree of PGC-1α represents a potent strategy to finely regulate the intricate process of mitochondrial biogenesis.

PGC-1β

PGC-1β and PGC-1α are part of the PGC-1 family and have very similar structures and functions [29]. However, PGC-1β is distributed differently among different cell types in the body. It acts as a coactivator for PPARs and ERRs, binding to nuclear receptors and transcription factors related to mitochondrial biogenesis. This process helps increase mitochondrial biogenesis and sustain the baseline function of mitochondrial [30]. For example, both PGC-1β and PGC-1α play a direct role in controlling mitochondrial biogenesis through Nrf1, which allows mitochondria to fulfill the energy needs of cells [31]. Additionally, PGC-1β maintains mitochondrial function and promotes mitochondrial biogenesis during proliferation, differentiation, and activation processes, particularly in M2 macrophages and osteoclasts that can proliferate [32]. Nevertheless, PGC-1β does not increase in brown adipose tissue in response to cold exposure or muscles during exercise. Overall, the role of PGC-1β in mitochondrial biogenesis is also crucial, separate from PGC-1α in this process.

Upstream regulatory targets for mitochondrial biogenesis

AMP-activated protein kinase (AMPK)

Changes in energy levels are an effective factor in stimulating mitochondrial biogenesis at an early stage. AMPK, a key player in connecting cellular metabolism and immune response, serves as a vital sensor for energy changes [33]. When a sudden drop in energy levels causes an increase in AMP, AMPK is triggered and activated to exert its phosphatase function, thereby directly stimulating PGC-1α and subsequently promoting the downstream Nrf1/2-TFAM pathway, leading to the upregulation of mitochondrial biogenesis [34, 35]. Additionally, upon activation, AMPK can also directly stimulate Nrf2 to initiate mitochondrial protein synthesis independently of PGC-1α [36]. Moreover, due to its wide range of functions, AMPK can activate some upstream factors of PGC-1α, such as calcium/calmodulin-dependent protein kinase (CaMK), cytochrome C, and the Sirt family [37,38,39,40]. A study has shown that AMPK can form a complex with PPARδ, even without exercise, enhancing mitochondrial energy metabolism and improving mass [41]. Currently, the use of drugs possessing AMPK-activating properties or upregulation of AMPK upstream-activated proteins offers a novel clinical strategy for enhancing mitochondrial biogenesis in the treatment and prevention of diseases.

CaMK

Ca2+ plays a crucial role in regulating mitochondrial biogenesis. When Ca2+ levels increase, it triggers a pathway involving CaMK activation that leads to the expression of PGC-1α and TFAM. This, along with enhanced binding of Nrfs to DNA, positively regulates mitochondrial biogenesis [42, 43]. The CaMK family can be divided into four categories, among which CaMK I, CaMKII, and CaMKIV have been widely studied concerning mitochondrial biogenesis. Their lower affinity and specificity for substrates enable them to exhibit diverse functions in vivo and act as enzymes that promote the phosphorylation of multiple intracellular proteins [44]. Studies have shown that CaMK upregulates and activates PGC-1α by promoting the phosphorylation of p38 mitogen-activated protein kinases (MAPK) and activating extracellular regulated protein kinase (ERK)/cAMP-response element-binding protein (CREB) pathways [43, 45, 46]. CaMKII is particularly important for promoting mitochondrial biogenesis in the heart and skeletal muscle, while CaMKIV, although significant, is not essential for this process [47]. Further research is needed to fully understand the role of CaMK in mitochondrial biogenesis.

Sirtuin 1 (Sirt1)

The NAD+/NADH ratio is seen as a key factor in modulating mitochondrial function. Sirt1, a well-researched “longevity gene”, is an NAD+-dependent histone deacetylase that is important for deacetylating mitochondrial proteins [48]. Research has shown that Sirt1 upregulates the activity of PGC-1α by promoting its deacetylation, thereby enhancing mitochondrial quality and biogenesis [48]. Additionally, the activity of Sirt1 can be influenced by AMPK as well [49].

Others

In addition to the factors mentioned above, there are still many other factors that can impact mitochondrial biogenesis by controlling the expression levels and activation of PGC-1α. Histone crotonylation, a type of post-translational modification, is one such factor that may regulate the expression of PGC-1α [50]. Therefore, enhancing the levels of the precursor of the substrate for histone crotonylation has been reported to boost mitochondrial biogenesis, particularly in cases of acute kidney injury [50]. The transcription of the PGC-1α gene is strongly activated by the coactivator transducer of Creb-related binding protein 1 (TORC1). Additionally, TORC2 and TORC3, other members of the TORC family, also play a role in activating PGC-1α transcription [51]. Furthermore, TORCs induce mitochondrial respiratory chain and TCA cycle processes. Similarly, there are several negative regulatory factors in the body as well. Members of the tumor necrosis factor (TNF) superfamily, such as TNF-like weak inducer of apoptosis (TWEAK) and TNF-α, reduce the expression of PGC-1α and mitochondrial biogenesis by activating nuclear factor-κB (NF-κB) [52]. Interestingly, research has indicated that the absence of receptor-interacting protein 140 significantly increases mitochondrial biogenesis and intracellular FAO [53]. The exploration of key targets regulating PGC-1α is becoming increasingly important in the study of mitochondrial biogenesis.

Mitochondrial fusion and fission

The regulation of mitochondrial fusion and fission balance, also known as mitochondrial dynamics, is a pivotal process in MQC that allows cells to meet their metabolic demands, ensures the clearance of damaged organelles, and accomplishes self-renewal of mitochondria with minimal resources and energy requirements [54]. Mitochondrial fusion refers to the process in which two mitochondria merge to form a single mitochondrion [55]. This process is not limited to the relative positions of the two mitochondria and can occur at either the ends or the sides. Due to the double-membrane structure of mitochondria, outer membrane fusion, and inner membrane fusion are the two major events involved in mitochondrial fusion, and there may be a time difference between them [55]. Following the fusion of mitochondrial membranes, the mitochondrial matrix fusion occurs. However, it should be noted that the mitochondrial genome is located within the matrix, and its fusion is somewhat restricted [56]. Mitochondrial fission is when the mitochondrial membrane contracts triggered by the dynamin-related protein 1 (DRP1), leading to the division of the original mitochondrion into two independent organelles [57]. The dynamic changes in mitochondrial fusion and fission control the morphology of mitochondria, exchange of contents, and maintain the normal function of mitochondria at physiological or pathological levels [58]. In the regulation of mitochondrial dynamics, the following key molecules serve as the primary agents and regulatory targets for fusion or fission.

Regulators of mitochondrial fusion

Mitofusins (MFNs)

It is well-recognized that outer mitochondrial membrane (OMM) fusion is entirely reliant on MFN1 and MFN2 [59], which are highly homologous GTPases in mammals. They possess conserved GTPase domains that regulate GTP binding and hydrolysis, leading to conformational changes [60]. Consequently, both MFN1 and MFN2 exhibit remarkable similarities in terms of function and structure. Through their α-helical regions, MFNs form homodimers or heterodimers, thereby acting as bridges to facilitate OMM fusion. It has been observed that the knockout of MFNs followed by overexpression of either MFN1 or MFN2 through plasmid transfer reveals the greater significance of MFN1 in OMM fusion [61]. Apart from its core role in regulating mitochondrial fusion, MFN2 plays a more prominent role in mediating the contact between the endoplasmic reticulum (ER) and mitochondria [62]. Within the ER, MFN2 is accumulated in regions where the ER interacts with mitochondria, known as the mitochondria-associated ER membranes (MAMs) [63, 64]. By interacting with MFN1 or MFN2 located in the OMM, MFN2 facilitates fusion between mitochondria and the ER, which is essential for regulating Ca2+ levels and mitochondrial function through inter-organelle communication [65].

The MFNs undergo modifications at different sites in response to cellular stimuli or signals, resulting in distinct patterns of OMM fusion. Both MFN1 and MFN2 can be ubiquitinated by Parkin, a key protein involved in mitophagy [66, 67]. Following ubiquitination, these two proteins are degraded by the proteasomes, leading to the inhibition of mitochondrial fusion while promoting mitochondrial fission and mitophagy. Similarly, MARCH5 can ubiquitinate and inhibit the function of MFN1/2 [68, 69]. Additionally, phosphorylation modifications exert diverse effects on MFNs. Different members of the MAPK family promote the phosphorylation of MFN1/2. For instance, ERK phosphorylates Thr562 of MFN1, whereas Jun N-terminal kinase (JNK) upregulates the phosphorylation level of Ser27 in MFN2 [70, 71]. After being regulated by MAPK, MFN1/2 plays a negative regulatory role in mitochondrial fusion. Interestingly, it has been demonstrated that site-specific phosphorylation is a prerequisite for the ubiquitination and recognition of MFN2 by the proteasomes [72]. However, AMPK, as a cellular sensor for energy and metabolism, exhibits opposite effects on the function of phosphorylated MFN1/2 [73]. It upregulates the process of mitochondrial fusion, and the direct interaction between AMPK and MFN2 significantly impacts ER-mitochondria fusion [64]. In comparison to phosphorylation modifications, the role of deacetylation modifications of MFN1/2 in promoting mitochondrial fusion is more precisely defined [74, 75]. In summary, the fusion of OMM is a central initial step in mitochondrial fusion, and post-translational modifications of MFNs play an important role in maintaining mitochondrial morphology and function.

Optic atrophy 1 (OPA1)

The process of mitochondrial fusion is intricate and not always guaranteed to occur normally, particularly when it comes to the inner mitochondrial membrane (IMM) fusion following the OMM fusion, which adds to the complexity [55]. If the OMM fuses while the IMM fusion fails, the resulting complex will undergo fission and fragmentation. The primary regulator of IMM fusion is OPA1, a member of the GTPase family similar to MFNs. OPA1 is located on the inner side of the mitochondrial inner membrane and is anchored to the IMM through its unique N-terminal transmembrane domain [76]. During inner membrane fusion, GTPase activity of OPA1 is modulated and binds to GTP, facilitating conformational changes that enable polymerization and promote inner membrane fusion [77]. When performing its core function, OPA1 often requires MFN1 rather than MFN2 [78]. In cells, OPA1 exists in various forms distinguished by length. Among them, the long isoform of OPA1, known as L-OPA1, plays a pivotal role in promoting IMM fusion and maintaining mitochondrial network integrity. The cleavage of L-OPA1 is primarily carried out by two proteases on the IMM, i-AAA protease YME1L, and metallopeptidase OMA1, in response to stress or metabolic changes [79]. Excessive accumulation of cleaved fragments, namely S-OPA1, results in mitochondrial fission or even fragmentation. Therefore, maintaining a balanced level between L-OPA1 and S-OPA1 is essential for regulating mitochondrial dynamics [80].

OPA1 is a key factor in maintaining the morphology and function of mitochondria. Its deficiency not only inhibits mitochondrial fusion but also significantly impacts the remodeling of mitochondrial cristae. It has been reported that upon activation of the apoptotic B-cell leukemia/lymphoma 2 (BCL2) family members Bax/Bak, there is an elevation in OPA1 phosphorylation levels, leading to its dissociation and interference with the formation and remodeling of mitochondrial inner membrane cristae, thereby affecting mitochondrial energy metabolism processes [81]. The regulation of OPA1 in cells is complex and involves its transcriptional level, activation state, and proteolytic cleavage. Mitochondrial factors like PGC-1α and TFAM can control the transcriptional level of OPA1 [82]. Furthermore, transcription factors related to immune regulation, such as NF-κB and signal transducer and activator of transcription 3 (STAT3), can also participate in OPA1 regulation in response to cellular environmental changes or inflammatory stimulation [83]. Concerning OPA1 modifications, research suggests that phosphorylation of OPA1 negatively impacts mitochondrial fusion, with upstream regulatory factors including Bax. Additionally, Sirt3 has been found to promote OPA1 deacetylation, enhancing its GTPase activity, which in turn facilitates OPA1 oligomerization and promotes mitochondrial fusion in response to cellular stimuli [84]. Sirt3 also aids in the deacetylation of YME1L, inhibiting its degradation activity and reducing the accumulation of S-OPA1, thereby promoting mitochondrial fusion through an alternative pathway [85].

Regulators of mitochondrial fission

DRP1

The dynamin-like GTPase DRP1 plays a crucial role in regulating mitochondrial fission. While primarily located in the cytoplasm during its inactive state, DRP1 is translocated to the OMM upon receiving signals to initiate mitochondrial fission. Here, it utilizes its GTPase activity by hydrolyzing GTP to obtain the necessary energy [86]. Upon activation, DRP1 undergoes conformational changes facilitated by interactions with other DRP1 molecules within the same mitochondrion, forming ring-shaped or helical aggregates that divide the intermembrane space between the inner and outer mitochondrial membranes [56]. Subsequent contraction of these DRP1 aggregates results in the division of mitochondria into two independent daughter mitochondria, which is a vital process for maintaining mitochondrial quantity and morphology [87].

Due to the lack of a domain directly binding to membrane phospholipids, the localization and activation of DRP1 in mitochondria rely on receptor proteins and recruitment factors on the mitochondrial surface [88]. Fission protein 1 (FIS1) interacts closely with the OMM through its C-terminal transmembrane domain, positioning itself on the OMM and recruiting DRP1 [89]. In yeast, the absence of FIS1 leads to the loss of DRP1’s function in promoting mitochondrial fission [89]. In mammals, although FIS1 is involved in promoting mitochondrial fission, there are also other proteins such as mitochondrial fission factor (MFF), mitochondrial dynamics protein of 49 kD (MiD49), and MiD51 that partially overlap with FIS1 in function and form complexes with DRP1 to regulate its function [90, 91]. MFF shares similar localization and function with FIS1, recruiting high-oligomeric forms of DRP1 and stimulating its GTPase activity [92]. Experimental evidence shows that overexpression of MFF exacerbates mitochondrial fission, while its absence does not significantly affect mitochondrial dynamics. The MiDs (MiD49 and MiD51) possess nucleotide transferase domains, with MiD51 requiring ADP as a cofactor to stimulate DRP1 oligomerization and GTPase activity [93]. The functions of MiDs are distinct from MFF, and in cells lacking MiD49/51, there is limited recruitment of DRP1, with minimal release of mitochondrial contents like cytochrome C [93]. Notably, MiDs exhibit a pro-fission role at low expression levels, but upon overexpression, mitochondria tend to elongate rather than undergo fission in response to stimuli [94, 95], possibly due to limited space for DRP1 contraction and aggregation, necessitating further investigation into the underlying mechanisms.

Additionally, the level of DRP1 modification significantly influences its function. During the early stages of cellular proliferation, Cyclin B/cyclin-dependent kinase 1 (CDK1) enhances the phosphorylation levels of DRP1 at Ser585 and Ser616, thereby increasing the GTPase activity of DRP1 and promoting mitochondrial fission [96, 97]. To interact with mitochondria, DRP1 must bind to specific receptors. Research indicates that phosphorylation of the Ser637 site of DRP1 inhibits its binding to mitochondrial receptors [96]. CaMK has been identified as a drive of mitochondrial fission through the promotion of dephosphorylation at the Ser637 site, underscoring the involvement of Ca2+ in mitochondrial dynamics [98]. However, protein kinase A (PKA) acts in opposition by elevating phosphorylation at the same site, thereby inhibiting the GTPase activity of DRP1 [99]. Moreover, the mitochondrial phosphatase phosphoglycerate mutase 5 (PGAM5) promotes mitochondrial fission and exacerbates cell death by recruiting and regulating dephosphorylation levels of DRP1 at Ser637 [100]. In various cell types, phosphorylation modifications of DRP1 exhibit distinct functional roles. For example, in neuronal cells, phosphatase and tensin homolog (PTEN)-induced kinase 1 (PINK1) enhances Ser616 phosphorylation of DRP1 to facilitate synaptic development, while in macrophages, STAT2 increases mitochondrial quality and modulates the pro-inflammatory phenotype by promoting phosphorylation of DRP1 at the same site [101, 102]. SUMOylation refers to a post-translational modification in which small ubiquitin-like modifier (SUMO) proteins are attached to specific sites of a target protein, altering its conformation and activity without inducing degradation [103]. The key enzyme responsible for SUMOylation modification of DRP1 is the mitochondrial anchoring protein ligase MAPL (mitochondrial E3 ubiquitin protein ligase 1, MUL1) [104]. Overexpression of MAPL promotes SUMOylation modification of DRP1, increasing its stability and activity, and positively regulating the process of mitochondrial fission, which is an important process in programmed cell death (PCD) [104]. Conversely, SUMO-specific peptidase 5 (SENP5), acting as a SUMO protease, inhibits OMM fission by cleaving SUMO1 from DRP1 [105]. Loss of SENP5 leads to uncontrolled fragmentation of mitochondria and even cell cycle arrest. However, a recent study has shown that two isoforms of SENP5, SENP5L, and SENP5S, counteract each other in the SUMOylation modification of DRP1 [105]. The balance between the two isoforms influences mitochondrial dynamics directionally; however, further investigation is required to elucidate this specific regulatory mechanism.

The role of ER in the initial stage of mitochondrial fission is a crucial aspect that warrants attention. DRP1 is not yet fully recruited and activated during the preparatory phase of mitochondrial fission. In the presence of mtDNA and Ca2+, the ER envelops the mitochondria, inducing preconstruction and reducing their average diameter to approximately 150 nm [106]. This process involves the interaction between inverted formin protein 2 (INF2) associated with the ER and SPIRE1C anchored to the mitochondria, which facilitates actin polymerization [107]. Simultaneously, at the interface where the ER contacts the mitochondria, known as the mitochondrial-associated membrane, oligomeric forms of DRP1 are recruited by MFF or MiDs to form a ring-like structure that initiates mitochondrial membrane fission [108]. Furthermore, the activation of DRP1 by certain cytoskeletal proteins has been confirmed, highlighting the significant role of the ER in regulating mitochondrial dynamics.

Adapters of DRP1

As previously discussed, although the adapter proteins of DRP1 exhibit overlapping functions, their post-translational modifications also play a significant role in mitochondrial fission. For instance, the tyrosine kinase Met promotes the phosphorylation of FIS1, facilitating its binding to DRP1 and enhancing mitochondrial fission [109]. Conversely, MARCH5 targets FIS1 for ubiquitination, resulting in its degradation by proteases and inhibiting its overall levels, consequently attenuating mitochondrial fission [110]. Similarly, in response to cellular changes, JNK can elevate the phosphorylation levels of MFF, thereby augmenting the cellular response to mitochondrial fission [111].

S-OPA1

The role of L-OPA1, located in the mitochondrial inner membrane, in mitochondrial fusion is well known. However, the fragmented form of OPA1, termed S-OPA1, has recently been found to exert significant effects on mitochondrial fission. S-OPA1 is generated through the cleavage of L-OPA1 by mitochondrial enzymes like OMA1, YME1L, or MPP, and remains soluble in the intermembrane space of the mitochondria [112]. Although different cleavage sites result in varying lengths of S-OPA1 structures, the accumulation of any isoform of S-OPA1 positively influences mitochondrial fission. Elevated cellular stress signals or alterations in metabolic conditions, particularly abnormalities in OXPHOS, enhance the activity of OMA1 or YME1L, accelerating the accumulation of S-OPA1 [113]. In normal physiological conditions, S-OPA1 synergistically interacts with L-OPA1, but excessive accumulation of S-OPA1 yields contrary effects [79, 114]. The specific mechanisms underlying the limitation of mitochondrial fusion and the promotion of mitochondrial fission due to the accumulation of S-OPA1 remain incompletely elucidated. Nonetheless, a prevailing view among researchers is that S-OPA1 disrupts the stability of the mitochondrial contact sites and cristae junction organizing system complex, as well as the function of OMM-IMM contact, thereby facilitating IMM fission [115]. Furthermore, the decline in the proportion of L-OPA1 significantly impedes mitochondrial fusion. Consequently, there exists substantial scope for further exploration into the ramifications of distinct OPA1 forms on mitochondrial dynamics.

Mitophagy

Autophagy is a highly conserved cellular process in eukaryotic cells. It involves the recycling of cellular components through the formation of autophagosomes, which engulf and degrade cellular contents in lysosomes [116]. This process maintains a balance between biosynthesis and degradation of cellular organelles and facilitates the selective turnover of surplus or dysfunctional cellular components to meet energy and macromolecular demands during cellular renewal [117]. Autophagy can be classified into three main types based on the pathways through which the degraded components reach lysosomes: macroautophagy, microautophagy, and chaperone-mediated autophagy [118]. Among these, macroautophagy is the most important and prevalent type and is characterized by the non-specific degradation of engulfed materials. When cells are subjected to external stimuli such as damage, nutrient deprivation, or oxidative stress, mitochondria may be damaged, depolarized, and lose their membrane potential, which triggers a selective macroautophagy mechanism that degrades mitochondria, known as mitophagy [9, 119]. The damaged mitochondria are engulfed and sequestered into autophagosomes, forming mitochondrial autophagosomes, which then fuse with lysosomes for degradation, ensuring the balance of mitochondrial content and energy [7]. Mitophagy can occur in various cell types. In addition to clearing dysfunctional mitochondria and reducing cellular damage signals, excessive mitophagy can result in the accumulation of cytotoxic substances and metabolic disturbances. Mitophagy can be divided into two categories based on initiating factors and mitophagosome formation mechanism: ubiquitin-dependent pathway and ubiquitin-independent pathway. In the ubiquitin-independent pathway, specific receptor proteins such as BCL2 interacting protein 3 (BNIP3)/ BCL2 interacting protein 3 like (BNIP3L), FUNDC1, and AMBRA1 recruit LC3 to facilitate mitochondria degradation [7]. The remainder of this section provides a comprehensive overview of the main pathways and regulatory factors involved in mitophagy.

Main pathways of mitophagy

PINK1/Parkin

Ubiquitin-dependent process of mitophagy is mediated by Parkin, an E3 ubiquitin ligase that regulates the ubiquitination of mitochondrial proteins. Similar to other RING-between-RING E3 ligases, Parkin forms a ubiquitin-thioester intermediate at the active site Cys431 and transfers ubiquitin to substrates [120]. However, Parkin exhibits low specificity in substrate selection, allowing it to modify multiple proteins on the OMM to drive mitophagy. Notably, numerous OMM proteins ubiquitinated by Parkin are extensively degraded by the ubiquitin–proteasome system (UPS) before mitophagy, including MFN1, MFN2, FIS1, and Tom20, which are essential for subsequent PINK1/Parkin-mediated mitophagy [121,122,123]. These ubiquitinated and degraded OMM proteins, such as MFN2, can drive mitophagy by disrupting mitochondria-ER coupling [124]. The degradation of specific OMM proteins can also result in OMM rupture and the subsequent degradation of IMM and mitochondrial matrix proteins, which may also be in part removed by the lysosome [125, 126]. The acquisition of Parkin enzymatic activity relies on its ubiquitin-like domain [127]. In the inactive state, the inhibitory conformation resulting from interactions among Parkin’s unique domains prevents downstream target ubiquitination. Upon activation by PINK1, the ubiquitin-like domain undergoes phosphorylation at the Ser65 site, altering Parkin’s conformation and exposing the active site Cys431, thereby enabling its ubiquitin ligase activity, along with the phosphorylated ubiquitin by PINK1 activation at Ser65 [127]. PINK1’s activation of Parkin extends beyond this role. PINK1 phosphorylates the ubiquitin moieties on mitochondrial surface proteins, generating phosphorylated ubiquitin, which recruits Parkin to the mitochondria [127]. On depolarized mitochondria, Parkin facilitates further substrate ubiquitination, which promotes the phosphorylation activity of PINK1 and forms a positive feedback loop [128]. These two phosphorylation events establish the essential role of PINK1 in Parkin-mediated mitophagy.

Once Parkin is fully functional, the presence of autophagy-related adaptor proteins ensures the encapsulation of mitochondria by autophagosomes, with LC3 coating their surfaces [129]. Among them, the most important is Sequestosome-1 (p62/SQSTM1), which contains a ubiquitin-binding domain to recognize ubiquitin chains on mitochondria [130]. It also possesses an LC3-interacting region (LIR) to recruit and bind to the autophagosome, thus selectively bridging ubiquitinated proteins to the autophagosome. Research has shown that the absence of p62 does not affect the activation of PINK1/Parkin but does impair the final clearance of targeted mitochondria [67]. Similar autophagy receptor proteins include optineurin (OPTN), NDP52, NBR1, and TAX1BP1 [131]. Among them, OPTN and NDP52 can directly bind ubiquitin-tagged mitochondria to the autophagosome through their LC3-interacting regions, even without the involvement of Parkin [132].

The fusion of autophagosomes with lysosomes is the final step, in which autolysosomes are formed to degrade damaged mitochondria [133]. Pleckstrin homology domain-containing protein family member 1 (PLEKHM1) contributes to the initiation of autophagosome-lysosome fusion by attaching to the LC3, Rab7, and HOPS complex [134]. Furthermore, Atg8 family members, especially the GABARAP subfamily, can recruit PLEKHM1 to promote autophagosome-lysosome fusion during PINK1/Parkin-mediated mitophagy [7, 129].

BNIP3 and BNIP3L (Nix)

Some receptor proteins on the mitochondria can directly bind to LC3 through their own LIR without the need for Parkin-mediated ubiquitination to initiate mitophagy. In mammals, BNIP3/BNIP3L share a certain degree of homology and belong to a subfamily of the BCL2 family [135]. BNIP3 was initially identified as a pro-apoptotic factor that regulates the function of BCL2, with its C-terminus anchoring to the OMM through a transmembrane domain, and its N-terminus being the region where it mainly exerts its pro-mitophagy function [136]. The N-terminus of BNIP3 contains a BH3 domain and a LIR domain surrounded by two serine residues (Ser17 and Ser24), of which phosphorylation of Ser17 is crucial for the binding of BNIP3 to LC3B [137]. BNIP3L, also known as Nix, has a highly similar structure to BNIP3, including the distribution of the C- and N-terminus, and both require self-dimerization for pro-mitophagy function. BNIP3L was initially reported to be involved in mitochondrial clearance during erythrocyte maturation, but subsequent studies have revealed its role in various cell types [138]. Phosphorylation of Ser34 and Ser35 near the LIR domain of BNIP3L stabilizes its interaction with LC3 [139]. Under hypoxic conditions, Ser81 of BNIP3L is activated by hypoxia-inducible factor-1α (HIF-1α), which promotes its binding to LC3 [140]. In this case, ubiquitination of specific OMM proteins and degradation by UPS affect mitochondria fragmentation, representing the upstream of BNIP3-dependent mitophagy [123]. In abnormal cellular metabolic environments with increased energy demands, BNIP3L promptly clears dysfunctional mitochondria to accelerate mitochondrial recovery and enhance OXPHOS efficiency, which requires the recruitment of the small GTPase RHEB to the OMM for upregulating BNIP3 function [141, 142].

Due to the sequence similarity between BNIP3 and BNIP3L, the expression of BNIP3 can restore the mitochondrial clearance rate in reticulocytes lacking BNIP3L, and the knockout of BNIP3 can induce upregulation of BNIP3L expression, but this upregulation cannot compensate for the reduction in mitophagy caused by BNIP3 knockout [143]. Recent findings seem to have indicated the roles of BNIP3 and BNIP3L in PINK1/Parkin-mediated mitophagy. BNIP3L is ubiquitinated by Parkin, which promotes the targeting of the selective autophagy receptor NBR1 [144]. BNIP3 interacts with PINK1, promoting the accumulation of PINK1 on the OMM, leading to the translocation of Parkin to the mitochondria [145].

FUN14 domain containing 1 (FUNDC1)

FUNDC1 is exclusively localized to the OMM and serves as a receptor mediating mitophagy in response to hypoxic stress in mammals [146]. FUNDC1 has three transmembrane domains and a LIR domain at its N-terminus [146]. The pro-mitophagy function of FUNDC1 is regulated by phosphorylation and dephosphorylation of residues Ser13 and Tyr18 near its LIR domain [147]. Under resting conditions, FUNDC1 is phosphorylated at Tyr18 by SRC (SRC proto-oncogene, non-receptor tyrosine kinase) and at Ser13 by casein kinase II, inhibiting its interaction with LC3 under normoxic conditions [148, 149]. However, during hypoxic conditions, the deactivation of SRC weakens Tyr phosphorylation, PGAM5 dephosphorylates Ser13, and Unc-51-like autophagy activating kinase 1(ULK1) activates Ser17 [149, 150]. These events collectively enhance the interaction between FUNDC1 and LC3, thus promoting mitophagy.

Others

There are several other mitophagy receptor proteins present on mitochondria that participate in binding with autophagosomes to promote mitophagy. Similar to BNIP3 and BNIP3L, AMBRA1 contains a BH3 domain that facilitates its interaction with BCL2 family proteins [151]. Early in mitophagy initiation, AMBRA1 is released from mitochondrial BCL2 and interacts with BECN1 to participate in the autophagy process [152]. Moreover, AMBRA1 synergistically interacts with the E3 ubiquitin ligase HUWE1 to recruit LC3B indirectly and induce mitophagy in response to mitochondrial depolarization [151]. BCL2 like 13 (BCL2L13), originally described as a pro-apoptotic member of the BCL2 protein family, contains all conserved BH domains: BH1, BH2, BH3, and BH4, as well as two LIR domains [153]. While the homolog of yeast Atg32 has not been identified in mammals, BCL2L13 can induce mitophagy in cells lacking Atg32, leading some researchers to propose that BCL2L13 is a functional counterpart of mammalian Atg32 [153]. BCL2L13 functions independently of Parkin-mediated mitophagy, and phosphorylation at Ser272 increases the binding of BCL2L13 to LC3 [154]. Additionally, BCL2L13 can regulate mitochondrial morphology by mediating mitochondrial fission through its four BH domains [155].

Under mild stress conditions, mitochondria-derived vesicle (MDV) formation serves as an alternative mechanism, allowing cells to degrade aberrant mitochondrial proteins and delicately regulates the mitochondrial proteome [156, 157]. Different MDV subpopulations are separated by single or double membranes composed of OMM or both OMM and IMM, which contain different cargoes and are selectively transported to multivesicular bodies or lysosomes for degradation [158,159,160]. PINK1 and Parkin appear to be involved in a subset of MDV biogenesis induced by oxidative stress [159]. In this process, syntaxin-17 is recruited to Parkin-dependent mitochondria-derived vesicles (MDVs) and subsequently forms a soluble NSF attachment protein receptor (SNARE) complex with SNAP29 and VAMP7 to facilitate the late endosomes/lysosomes fusion [161, 162]. The formation and degradation of Parkin-dependent MDVs occur independently of autophagy-related proteins and are supported by the silence of Atg5 and beclin-1, separating this process from mitophagy [159]. Conversely, PINK1 and Parkin have been shown to play a negative role in inflammation-induced MDV formation [163]. Previous studies have demonstrated that neither PINK1 nor Parkin are essential for the formation of steady-state MDVs [164]. This subset of MDVs is generated through the protrusion of the mitochondrial membrane facilitated by microtubule-associated motor proteins MIRO1 and MIRO2 (MIRO1/2) and subsequently released via DRP1 recruitment mediated by MID49, MID51, and MFF [164, 165].

Emerging findings on mitochondrial remodeling and MQC mechanisms manifest as a distinct mitochondrial structural transformation, called mitochondrial spheroids [166, 167]. Reportedly, oxidative stress can induce the formation of mitochondrial spheroids, and those with ring or cup-like morphology can surround cytosolic components, including ER and other mitochondria [167,168,169]. MFN1 and MFN2 are required for the generation of mitochondrial spheroids [170]. Parkin lies at the crossroads of mitochondrial spheroids and mitophagy. Parkin suppresses mitochondrial spheroid formation by triggering MFN1 and MFN2 proteasomal degradation. Mitochondrial spheroids are positive for lysosome proteins, implying a possible capacity to degrade their enwrapped contents, while definitive evidence is yet lacking [171]. Given the current findings, mitochondrial spheroids are considered to represent a protective alternative strategy for maintaining mitochondrial homeostasis when PINK1/Parkin-related mitophagy is hindered [169, 170].

Regulators of mitophagy

ULK1

As a functional homolog of yeast Atg1 in mammalian cells, it forms a complex with Atg13 and RB1CC1 to initiate autophagy [172]. ULK1 functions as a phosphatase in mitophagy by promoting FUNDC1 Ser17 phosphorylation to increase its LC3 binding ability, and upregulating ACT domain Ser108 phosphorylation in Parkin to activate it in an alternative pathway to PINK1 [173]. BCL2L13-mediated mitophagy depends on the interaction with ULK1 to localize the autophagy initiation complex to the mitochondria [174].

AMPK

It plays a positive regulatory role in autophagy by inhibiting the phosphorylation of the negative regulator mTOR and directly upregulating the activity of ULK1 [172]. Additionally, the activation of AMPK promotes the translocation of PINK1 from the cytoplasm to the mitochondria, thereby facilitating mitophagy [175]. AMPK can also activate mitochondrial fission protein FIS1 to upregulate the process of mitophagy [176].

MARCH5

As an E3 ubiquitin ligase, MARCH5 inhibits mitochondrial fusion by increasing the ubiquitination of MFN1/2. Research has revealed that MARCH5-mediated ubiquitination of FUNDC1 at Lys119 decreases the levels of FUNDC1 protein in OMM [177]. Specifically, FUNDC1 expression is reduced via ubiquitin-proteasome-dependent degradation during hypoxia, thereby suppressing mitophagy.

Sirts

Sirt family is a subclass of the histone deacetylase family that relies on NAD+ levels to maintain activity. Among them, Sirt1 promotes mitophagy by upregulating the transcription factor forkhead box O3 (FOXO3), which in turn promotes BNIP3 expression [178]. Additionally, Sirt1 has been reported to directly upregulate the protein levels of PINK1 and Parkin [179]. Sirt3 regulates BNIP3 activity through the ERK-CREB signaling pathway and promotes mitophagy by FOXO3a-mediated PINK1/Parkin pathway [180,181,182]. Sirt5 also upregulates the expression levels of mitophagy markers PINK1, Parkin, and even BNIP3 [183]. However, the specific mechanisms still require further exploration.

Human Lon protease 1 (LONP1)

The protein LONP1 localizes within the mitochondrial matrix, and plays a crucial role in degrading misfolded proteins, thereby preventing protein aggregation. This function is essential for maintaining the mitochondrial proteostasis [1, 184]. Research has demonstrated that LONP1 specifically targets unfolded proteins in the mitochondrial matrix, and their accumulation promotes PINK1/Parkin-mediated mitophagy in a manner independent of mitochondrial depolarization [185].

Physiological roles of MQC

Mitochondria play a vital role not only in regulating energy metabolism during physiological activities but also in regulating calcium homeostasis, redox homeostasis, PCD, and other essential processes. Numerous studies have highlighted their irreplaceable significance in these cellular functions. The subsequent section discusses the importance of MQC in ensuring and regulating these critical life activities (Fig. 2).

Fig. 2
figure 2

Physiological roles of mitochondrial quality control. a Mitochondria, as the core of energy metabolism, regulate their quality to adapt to the cellular environment under different bioenergy conditions. When nutrients are in excess, mitochondrial dynamics switch to division dominance and mitophagy is blocked. Under starvation conditions, the mitochondria fuse, and the mitochondrial autophagic flux is enhanced. Together, these regulations promote the balance of mitochondrial energy metabolism. b Mitochondria can act as calcium pools in cells, and there are extensive interactions between calcium ions and mitochondrial quality control to jointly regulate mitochondrial mass and calcium ion homeostasis. c Depolarizing mitochondria leads to oxidative stress. Mildly damaged mitochondria can fuse with healthy mitochondria to neutralize damage, and severely damaged mitochondria are cleared by mitophagy, but excessive mitophagy can also lead to oxidative stress. d Excessive mitochondrial fission is an early event in apoptosis, pyroptosis, and ferroptosis. Mitochondrial fragmentation promotes cytochrome and mtDNA release and causes oxidative stress. Furthermore, the regulation of ferroptosis correlates with mitophagy flux, and moderate mitophagy promotes the clearance of ROS, while excess mitophagy results in excess iron ions. e Mitochondria initiate fine kinetic processes to accommodate the mitochondrial genetic process in progeny during the cell cycle. Cyclin regulator CDK1 promotes protein import into the mitochondria to ensure mitochondrial energetic support during the cell cycle. Mitophagy reduces the pool of mitochondria in the cell, which would further limit the cell cycle. ATP adenosine triphosphate, BAX BCL2 associated X, CDK1 cyclin B/cyclin-dependent kinase 1, DRP1 dynamin-related protein 1, ETC electron transport chain, FIS1 fission protein 1, Fzo1 fuzzy onion 1, INF2 inverted formin 2, GTPase guanosine triphosphatase, MCU mitochondrial calcium uniporter, MFN mitofusin, Mgm1 mitochondria genome maintenance 1, mtDNA mitochondrial DNA, NCLX Na+/Ca2+/Liexchanger, NLRP3 NLR family pyrin domain containing 3, Nrf2 nuclear factor E2-related factor 2, OPA1 optic atrophy 1, PINK1 PTEN-induced kinase 1, RALBP1 ralA binding protein 1, ROS reactive oxygen species, STING stimulator of interferon response cGAMP interactor, TCA tricarboxylic acid, VDAC voltage-dependent anion channel

Energy metabolism regulation

Mitochondria are the energy-generating component of cells, accounting for approximately 90% of cellular energy produced. MQC is involved in the regulation of energy metabolism, which is crucial to maintaining the energy supply of the body in the normal physiological state, most notably for organs and tissues with high energy demands such as the heart, liver, kidney, brain, and skeletal muscle. As the main energy donor, glucose is oxidized into pyruvate by glycolysis in the cytoplasm under normal physiological conditions, and then enters the mitochondria via the pyruvate transporter and is converted to acetyl-CoA, which further drives TCA cycle capacity. Theoretically, 1 mol of glucose produces 36 – 38 ATP molecules, of which 34 – 36 ATP come from OXPHOS and the rest from glycolysis [186].

Under the condition of excess nutrition or starvation, the MQC system is initiated to precisely regulate the dynamic balance of mitochondria at the subcellular level to ensure the balance of energy demand and supply. The mitochondria tend to maintain their isolated state in the presence of a nutrient-rich environment, while they undergo elongation under conditions of cellular starvation [187, 188]. Previous studies have shown that mitochondrial fragmentation can be detected when the β cell line INS1 is simply exposed to high fat or combined with high glucose after 4 and 24 h [189]. Furthermore, reduced mitochondrial size in muscle and decreased MFN2 expression was observed in mouse models of type 2 diabetes mellitus (T2DM) and obesity [190,191,192]. In contrast to overnutrition, the recruitment of DRP1 to mitochondria is inhibited during starvation, resulting in suppressed mitochondrial fission and elongation, along with increased amounts of mitochondrial cristae and ATP synthesis capacity to maintain the required ATP demand when nutrient supply is limited [188, 193]. These studies suggest that starvation-induced mitochondrial elongation may be a positive mechanism for mitochondrial bioenergetic efficiency.

An increasing number of studies also suggest that mitophagy is involved in regulating energy metabolism [194,195,196]. Under glucose-rich conditions, the polymerase-associated factor 1 complex (Paf1C) can maintain low levels of mitophagy by binding to the promoter of the Atg32 gene, while upon starvation, Paf1C dissociates from Atg32, resulting in increased expression of the Atg32 gene and induction of mitophagy [197]. Deletion of Fundc1 induces defective mitophagy and impaired MQC, leading to more severe obesity and insulin resistance (IR) in mice fed a high-fat diet (HFD) [198]. Furthermore, mitochondria undergo different topological changes during starvation, either becoming swollen or developing a donut shape, and are further removed by mitophagy, while mitophagy-resistant mitochondria reintegrate via mitochondrial fission and fusion [199].

As a supplementary pathway to mitophagy, MDVs envelop and release damaged mitochondrial material in brown adipose tissue, leading to the obstruction of PPARγ signaling and ultimately inhibiting the TCA cycle in brown adipose tissue [200]. This phenomenon can be alleviated by the engulfment of extracellular vehicles (EVs) generated by these MDVs.

In conclusion, the MQC system can accurately regulate energy metabolism at the subcellular level, maintain the dynamic balance of mitochondria, and ensure the energy demand of the body.

Maintenance of calcium homeostasis

As a second cellular messenger, calcium is involved in regulating gene expression, proliferation, differentiation, metabolism, cell death, and survival, and plays an irreplaceable role in cellular pathophysiology [201]. Mitochondria buffer intracellular calcium levels by absorbing, storing, and releasing Ca2+, while MQC as an important mechanism for regulating cellular fate, also plays an important role in maintaining calcium homeostasis. The OMM regulates the voltage-dependent anion channel (VDAC) to facilitate the absorption of calcium ions from the cytoplasm into cells. Subsequently, these calcium ions traverse through the mitochondrial calcium uniporter (MCU) complex located in the IMM, thereby participating in the regulation of energy production and metabolism [186, 202, 203]. However, the calcium ions of the mitochondrial matrix mainly flow out to the cytoplasm through the Na+/Ca2+ exchanger (NCLX) and H+/Ca2+ antiporter [204]. In mitochondria, Ca2+ can trigger apoptosis by opening the mitochondrial permeability transition pore (mPTP) [205, 206]. The decrease of the adenine nucleotide pool decreases and the increase of phosphate, matrix calcium, and mitochondrial ROS (mtROS) will promote the massive outflow of particles less than 1.5 kD, such as protons and calcium ions, from the mitochondrial matrix [207, 208]. In addition, both mitochondria and ER are the main storage areas of intracellular calcium ions. Mitochondria interact with the ER through the mitochondrial ER contact site (MERCS) to regulate cellular and mitochondrial calcium homeostasis [209,210,211,212].

In the process of responding to changes in energy metabolism and cellular activity, there exists a mutual influence and regulation between MQC and calcium homeostasis. Early findings demonstrated that extracellular or intracellular calcium can inhibit mitochondrial disruption by inhibiting intracellular C2-ceramide [213]. Subsequently, it was confirmed that intracellular calcium regulates mitochondrial fission by controlling DRP1 localization, where calcineurin mediates DRP1 dephosphorylation, which then induces DRP1 transfer to mitochondria and accumulates on the OMM, thereby increasing the rate of mitochondrial fission [214, 215]. Moreover, in MERCS, INF2 was found to mediate actin polymerization, and then control MCU-mediated calcium uptake and DRP1-dependent mitochondrial assembly, thus promoting mitochondrial fission [216, 217]. Mitochondrial division inhibitor 1 (mdivi-1) effectively inhibits DRP1-mediated mitochondrial fission and also induces mitochondrial depolarization, ER Ca2+ depletion, and Ca2+ signaling regulation [218]. Ablation of constitutive DRP1 leads to enlarged mitochondrial morphology, increased mitochondrial calcium uptake, and abnormal mitochondrial function [219]. Moreover, Ca2+ is also related to the regulation of mitophagy. Calcium ion sensors RHOT1 and RHOT2 mediate Parkintrans location, which induces mitophagy after binding to mitochondria [220, 221]. Calcium flows from the ER to MERCS and into the mitochondria, uncoupling the mitochondria from the ER by disrupting MFN2, then regulating PINK1/Parkin-mediated phosphorylation-ubiquitination and inducing mitophagy [124, 222]. MUL1 deficiency increases MFN2 activity and decreases ER-mitochondria coupling, which further leads to increased cytoplasmic Ca2+ load, activating of calcineurin, and induction of DRP1-dependent mitochondrial fragmentation and mitophagy [223]. The mutual regulation between mitophagy and Ca2+ uptake and release is significant. The latest research findings indicate that transmembrane BAX inhibitor motif containing 6 (TMBIM6), an ER protein with calcium leakage activity regulates mitochondrial Ca2+ uptake by preventing the oligomerization of VDAC1, thereby causing impaired mitophagy and disrupted mitochondrial biosynthesis, further leading to compromised mitochondrial respiration and ATP production [224]. At the same time, damaged mitochondria induce the intracellular migration of calcium via the MCU, while mitophagy significantly inhibits this trend after clearing damaged mitochondria [225].

In summary, calcium imbalance leads to MQC dysregulation, resulting in mitochondrial dysfunction. Simultaneously, alterations in MQC can significantly influence the content and distribution of calcium ions within the cell. Further research on VDAC, MCU, and MERCS is expected to elucidate the connection between MQC and calcium homeostasis regulation.

Maintenance of the redox balance

mtROS is the main component of intracellular ROS, and is an important signaling molecule involved in the regulation of physiological functions, playing a significant role in maintaining the redox balance [186]. The production of mtROS is primarily influenced by OXPHOS capacity, the oxidation of NADPH/NADH, and the biosynthesis of heme and iron-sulfur centers [226,227,228]. During OXPHOS, electrons leak from mitochondrial complexes 1 and 3 and interact with oxygen molecules to produce the most dangerous mtROS: superoxide anion [229, 230]. Moreover, mitochondrial complex 1 can also deliver electrons from NADPH/NADH produced in the TCA cycle to oxygen molecules, promote the oxidation of NADPH/NADH and the generation of superoxide radicals, and then induce a large amount of mtROS production [231, 232]. Excessive production of mtROS can induce oxidative stress of lipids and protein and DNA damage. To avoid further cell damage, mitochondria eliminate excess ROS, thus maintaining the redox homeostasis via a system of antioxidants including superoxide oxidase, catalase, glutathione peroxidase, and antioxidant factors such as glutathione, vitamins [233, 234]. An increasing number of studies suggest that MQC plays a crucial role in the redox imbalance mediated by mtROS.

Mitophagy limits mtROS overproduction and maintains the redox balance by sequestering and engulfing aging and damaged mitochondria [235, 236]. Studies have shown that long-term exposure to PM2.5 can induce the massive production of mtROS and impair mitophagy, leading to redox imbalance. However, the addition of mitochondria-targeted antioxidant Coenzyme Q can improve mitophagy and oxidative damage by reducing the accumulation of ROS [237, 238]. Moreover, mitophagy inhibitors have been shown to induce significant ROS accumulation and redox imbalance, and eventually aggravate the disease state [239,240,241]. Among these inhibitors, endogenous and exogenous substances such as α-ketoglutarate and rapamycin, as well as signaling pathways such as AMPK, MAPK, and Nrf2, can maintain the redox balance of mitochondria by increasing mitophagy to clear ROS and oxidative stress [242,243,244,245,246]. However, increased mitophagy is often an early response to promote cell survival, and a chronic state of excessive mitochondrial damage can induce pathological increased mitophagy, resulting in cell death and tissue damage [246]. Thus, excessive mitophagy may also lead to an increase in mtROS, which could further induce redox imbalance. Additionally, under hypoxic conditions, mitochondria undergo widespread ubiquitination facilitated by UPS to mitigate mtROS accumulation and oxidative stress levels. The ubiquitination process activates receptor-dependent mitophagy, specifically emphasizing BNIP3/Nix, rather than the classical pathway mediated by Parkin [123]. Recent research has reported a close association between the biogenesis of peroxisomes and MPVs. The MPVs-encapsulated Pex3 and Pex14 are crucial components of peroxisomal precursors, while the lipids and peroxides carried by MPVs serve as metabolic substrates and tools of the peroxisomes [247]. MPVs primarily regulate the cellular redox balance through the encapsulation and release of mitochondrial contents [165].

Mitochondrial fission and fusion also play an important role in regulating oxidative stress. Under the pressure of oxidative stress, mitochondrial fusion increases to reduce ROS production [248, 249]. Increased MFN2 expression promotes mitochondrial fusion and autophagy, and reduces ROS, thereby maintaining the redox balance [250, 251]. In vitro experiments confirmed that increasing the ratio of GSSG/GSH leads to cis-oligomerization of MFN disulfide bonds and promotes mitochondrial fusion [252, 253]. A study has confirmed that C684 residue is required for MFN2 disulfide bond and fusion activity. When C684 is absent, MFN2 is more prone to redox changes, which affects the energy export of mitochondria [254]. In the newly proposed topological modification of MFN, it was also indicated that residue C684 can directly sense the redox environment of mitochondria [255]. In addition, studies have confirmed that MFN2 is a prerequisite for mitochondrial respiration to stress and ROS production, and the knockout of MFN2 in macrophages leads to ROS production defects and impairs immune response function [256, 257]. Furthermore, when oxidative stress escalates, the activation of the Nrf2 pathway not only aids in upregulating antioxidant defenses and restoring redox balance but also promotes the proteasomal degradation of the mitochondrial fission protein DRP1 [258]. It was also confirmed that the degradation of DRP1 is contingent upon non-ubiquitinated degradation facilitated by the 20S proteasome, while MFN degradation relies on ubiquitin-mediated degradation of the 26S proteasome. The oxidative stress can promote the breakdown of 26S proteasome into 20S and 19S subunits, which in turn promotes the disassembly of DRP1 and stabilizes MFN, thus reducing mitochondrial fission [259]. Overall, under the stimulation of oxidative stress, cells can protectively reduce the accumulation of ROS and restore the redox balance by regulating the MQC system.

Cell cycle regulation

The cell cycle comprises a series of ordered lifecycle stages during which a cell undergoes division from one mother cell into two daughter cells, facilitating cellular renewal, growth, and replication of genetic material [260]. The cell cycle can be divided into two main stages: the mitotic phase (M phase) and interphase. The interphase, consisting of the G1 phase, S phase, and G2 phase, serves as the preparatory stage of the cell cycle that provides conditions for subsequent division, while the M phase is the primary stage of cell division. To meet the material and energy requirements of various stages in the cell cycle, the functionality and morphology of mitochondria undergo dynamic changes [261]. During the G1 and G2 phases, mitochondria form an interconnected network, which subsequently undergoes division in mitosis and the S phase [262]. In addition to exerting influence on the levels of materials and energy required for cellular activities, mitochondria can also passively participate in regulating cell cycle progression by affecting specific checkpoints.

The translocase at the outer membrane of mitochondria (TOM complex) is a protein translocase essential for the import of raw materials into mitochondria during mitochondrial biogenesis [164]. In the M phase, the expression of TOM6, a key component of the TOM complex, significantly increases. Simultaneously, the cell cycle regulatory factor CDK1 facilitates the phosphorylation of TOM6 Ser16 [263]. The phosphorylation of TOM6 enhances the steady-state level and activity of the TOM complex, thereby promoting protein entry into mitochondria. This process involves an increase in mitochondrial M-phase-specific regulatory GTPases, Fzo1 and Mgm1, which ultimately enhance energy production to meet the heightened energy demands during the M phase [55]. Similarly, during the early stages of the M phase, RALBP1, the effector protein of RALA, interacts with CDK1/Cyclin B to phosphorylate DRP1 at Ser616, thus promoting mitochondrial fission [264]. As part of a feedback regulatory mechanism, acute loss of DRP1 enhances the mitochondrial recruitment of Cyclin E, a crucial protein regulating the G1/S transition of the cell cycle, which prevents its degradation and promotes the cellular transition from G1 to S phase [265]. However, chronic inhibition of DRP1 leads to mitochondrial hyperfusion, which in turn results in a p53-dependent blockade of S-phase entry [265]. Subsequently, DRP1 deficiency-induced mitochondrial hyperfusion also leads to DNA replication stress-induced ATM-dependent G2/M arrest through the excessive accumulation of Cyclin E [266]. Depletion of OPA1 can ameliorate the aforementioned phenomena [266]. It highlights the necessity of mitochondrial fission during the S, G2, and M phases following the completion of the G1/S transition to ensure that daughter cells inherit consistently healthy mitochondria. In cells containing damaged mitochondria, PINK1/Parkin-mediated mitophagy is activated, promoting optineurin (TBK1) mitochondrial translocation and phosphorylation [267]. Consequently, this leads to G2/M cell cycle arrest and inhibition of mitosis [267]. In summary, MQC actively participates in regulating the cell cycle to ensure the proper transmission of mitochondrial genetic material and the stability of intracellular energy supply during cell growth.

Regulation of PCD

Different types of cell death are the inevitable fate of all life to maintain healthy cell renewal and the homeostasis of the organism [268, 269]. PCD is the primary means of cell renewal and is controlled by numerous genes. Strict MQC is essential for mitochondrial function and is closely associated with physiological PCD.

Apoptosis is a type of PCD executed by caspases and is characterized by specific morphological changes, including cell membrane blebbing, nucleus condensation, condensation and fragmentation of genetic materials, and formation of apoptotic bodies [268, 270]. Two distinct apoptosis pathways include intrinsic and extrinsic pathways, known as the mitochondrial and death receptor pathways, respectively [271]. Intrinsic apoptosis is caused by DNA damage, hypoxia, metabolic stress, and other stimulations, inducing the expression of pro-apoptotic BH3-only proteins (such as BIM, PUMA, BAD), binding and neutralizing pro-survival BCL2 proteins (such as BCL2, BCL-XL, and BFL-1), thus releasing apoptosis effectors BAK and BAX. It leads to mitochondrial outer membrane permeabilization and subsequent release of mitochondrial proteins, including cytochrome C, the second mitochondrial activator of caspases [272, 273]. Mitochondrial fission is considered an early event during apoptosis [274]. During apoptosis, DRP1 redistributes to mitochondrial membranes and induces mitochondria fragmentation [275]. DRP1 is irreversibly locked on the membrane in a BAX/BAK-dependent manner after mitochondrial fragmentation, accompanied by stable SUMOylation [276]. Deficiency of DRP1 delays the release of cytochrome C and apoptosis, which is related to the enclosure of cytochrome C into the highly stacked cristae [275, 277]. Loss of FIS1 has been found to reduce mitochondrial fission and apoptosis, partly because it inhibits BAX translocation [278]. Correspondingly, insufficient mitochondrial fusion may also lead to mitochondrial fragmentation during apoptosis. OPA1 participates in controlling cristae remodeling and cytochrome C release, whereas deletion of OPA1 leads to the disruption of the mitochondrial network and apoptosis [76, 279]. Interestingly, BAX and BAK physiologically regulate MFN2 activity and normal morphogenesis of mitochondria; yet, during apoptosis, BAK dissociates from MFN2 and interacts with MFN1, which may reduce mitochondrial fusion and induce mitochondrial fragmentation [280, 281]. In addition, MFN1 phosphorylation by ERK also influences BAK oligomerization under apoptotic stimuli [70]. Dynamic PGAM5 multimers have been reported to act as a molecular switch to coordinate mitophagy and apoptosis. In response to distinct stresses, PGAM5 dephosphorylates BCL-XL to inhibit apoptosis or FUNDC1 to enhance mitophagy by switching between dimeric and multimeric states [282]. Surprisingly, in the absence of mitophagy, PINK1/Parkin can promote cell apoptosis through a non-BAX pathway [283]. The ubiquitination effect of UPS on OMM promotes the release of cytochrome C, directly activates autophagy receptors, upregulates mitophagy, and inhibits apoptosis [283].

Pyroptosis is an immunogenic PCD driven by inflammasome [273]. Pathogen- or danger-associated molecular patterns trigger canonical caspase-1 inflammasome pathway. In this process, activated caspase-1 directly cleaves GSDMD and releases N-terminal fragments to bind to phosphatidylinositol on the plasma membrane, forming membrane pores [268, 284]. Meanwhile, caspase-1 cleaves pro-IL-1β and pro-IL-18 to produce mature IL-1β and IL-18, which are released from the GSDMD pores. It has been reported that DRP1-mediated excessive mitochondrial fission contributes to the activation of caspase-1 and NLRP3 inflammasome, which may be related to the increased ROS induced by mitochondrial dynamics dysfunction [285, 286]. Analogously, ROS also acts as a bridge between defective mitophagy and pyroptosis [287, 288]. Defective mitophagy can result in mitochondrial dysfunction, increased mPTP, and mtDNA leakage, leading to a vicious cycle between MQC deficiency and excessive inflammation [289].

Ferroptosis is an iron-dependent form of PCD, primarily driven by the toxic accumulation of cell membrane lipoperoxides [290]. Mitochondria, serving as the primary sources of intracellular ROS and ATP, are also actively involved in various processes, including the TCA cycle, lipid biosynthesis, and glutaminolysis [291]. These biological components play a key regulatory role in the regulation of the ferroptosis process. Within cellular activities, the interplay between MQC and ferroptosis is relatively complex and can be dissected at several key junctures. As described above, Nrf2 regulates proteasome genes that contribute to DRP1 degradation, upregulates MFN2 protein levels to promote mitochondrial fusion and inhibit fission, and functions as a promoter protein for mitochondrial biogenesis, the absence of Nrf2 greatly upregulates ferroptosis [258]. Additionally, STING accumulates significantly during its transportation from the ER to mitochondria, where it binds to MFN1/2, thereby triggering mitochondrial fusion [292]. During this process, the increase in ROS and fatty acid levels leads to lipid peroxidation, consequently promoting ferroptosis. In contrast, mitophagy exhibits a dual role in the regulation of ferroptosis. The release of ROS and certain pro-apoptotic factors from damaged or dysfunctional mitochondria promotes ferroptosis. This positive regulation is notably inhibited when mitophagy clears abnormal mitochondria [293]. Furthermore, during mild stress or in the early stages of iron overload, mitophagy can reduce the source material of ROS in ferroptosis by chelating iron. However, further expansion of mitophagy may ultimately provide additional iron, thereby amplifying lipid peroxidation and ferroptosis [294]. Current perspectives suggest that the direction of mitophagy regulation in ferroptosis primarily depends on the magnitude of mitophagy flux. In conclusion, MQC plays a pivotal role in regulating ferroptosis, and in-depth exploration of their interconnectedness and mechanisms will offer new insights into understanding cellular life processes and disease progression.

MQC in human diseases

MQC and cancer

Tumor initiation and progression

Cancer cells can redirect metabolites towards biosynthetic pathways to support their rapid proliferation and accumulate the cellular building blocks required for tumor growth. Enhanced mitochondrial biogenesis contributes to enhanced OXPHOS and fosters cancer growth [295]. Migratory/invasive cancer cells utilize PGC-1α to boost mitochondrial biogenesis, OXPHOS, and oxygen consumption rates, promoting cancer cell invasion and metastasis [296]. Mitochondrial biogenesis mediated by mitochondrial Ca2+ signaling promotes the growth of colorectal cancer cells in vitro and in vivo [297]. Inhibiting mitochondrial biogenesis through genetic or pharmacological means can suppress tumor cell metabolism and activity, and inhibit tumor progression [298, 299]. However, it is interesting to note that low levels of mitochondrial biogenesis have been linked to sorafenib resistance in hepatocellular carcinoma. The study found that mitochondria in sorafenib-resistant cells maintain greater functional and morphological integrity under sorafenib treatment, but their number was lower. This can be attributed to the decrease in mitochondrial biogenesis, which is caused by the accelerated degradation of PGC-1β [300]. Cancer stem cells (CSCs) are distinct subpopulations of cancer cells with stem cell-like abilities that are more resistant to chemotherapy, leading to tumor recurrence. By enhancing mitochondrial biogenesis, CSCs maintain mitochondrial content and metabolic homeostasis to support cancer growth. Targeting mitochondrial biogenesis may be a potential therapeutic option to eliminate CSCs [301].

Imbalanced mitochondrial dynamics have a significant impact on cancer development and metastasis. Typically, mitochondria are fragmented and distributed within different types of tumor cells through fission, such as lung cancer [302], colon cancer [303], breast cancer [304], neuroblastoma [305], melanoma [306], ovarian cancer [307], prostate cancer [308], pancreatic cancer [309]. Carcinogenic mutations are a major cause of excessive mitochondrial fission, with BRAFV600E or RASG12V mutations driving MAPK pathway activation to induce mitochondrial fission and promote tumor growth. Inhibiting DRP1 or promoting fusion can induce tumor cell death [306, 310,311,312]. The tendency of tumor cells to undergo mitochondrial fission is related to metabolic reprogramming, cell cycle progression, and increased migration, invasion, and metastatic potential [6, 313]. During mitosis in tumor cells, mitochondrial fission is crucial to ensure equal segregation of mitochondrial contents between daughter cells, which is important for the rapid proliferation of tumor cells. Lack of mitochondrial fission induces cell death and cellular dysfunction due to replication pressure and compromised genomic integrity [312, 314]. Additionally, excessive mitochondrial fission resulting in mitochondrial fragmentation redirects metabolism to the peripheral cytoskeleton’s lamellipodia, which serves as a concentrated energy source for tumor cell movement and invasion, promoting tumor cell invasion and metastasis [315]. Therefore, targeting mitochondrial fission or promoting fusion is a potential therapeutic strategy for many cancer treatments [6]. Yu et al. [316] proposed that the main tumor suppression mechanism of promoting mitochondrial fusion is to enhance mitophagy, which proportionally reduces mitochondrial mass and ATP production. However, mitochondrial dynamics exhibit heterogeneity in cancer. Overactive mitochondrial fusion has been observed in tumor tissues of hepatocellular carcinoma patients and tumor-like organoids of cholangiocarcinoma in vitro. Knockdown of OPA1 or MFN1 to inhibit mitochondrial fusion weakened oxygen consumption and cellular ATP production of tumor cells, suppressed cell growth in vitro, and inhibited tumor formation in vivo. This inhibitory effect was related to inducing cell apoptosis but not to cell cycle arrest [317]. Similarly, researchers found that enhancing mitochondrial fission inhibited signal transduction and metastasis of triple-negative breast cancer while enhancing mitochondrial fusion overcame the inhibitory effect of fission on migration, signal transduction, and metastasis. Further exploration of existing datasets on breast cancer revealed that increased expression of mitochondrial fission-related genes was associated with improved survival in human breast cancer [318]. Therefore, more mechanistic studies are needed to comprehensively understand how mitochondrial dynamics imbalance leads to cancer occurrence and progression. Additionally, OPA1-mediated mitochondrial fusion in endothelial cells promotes tumor angiogenesis, affecting tumor growth and metastasis [319]. In addition, activation of the CDK1/Cyclin B-DRP1 pathway induces mitochondrial fission, facilitating cell cycle progression and increasing the sensitivity of tumors to radiotherapy [320]. In CSCs, mitochondrial fission is often associated with increased rates of mitophagy, epithelial-to-mesenchymal transition, and increased rates of glycolysis. Conversely, mitochondrial fusion is typically associated with increased oxidative phosphorylation and resistance to therapy [321].

Mitophagy suppresses tumors by eliminating dysfunctional mitochondria that disrupt cellular metabolism and promote cellular transformation and tumorigenesis. Conversely, mitophagy-mediated clearance of apoptotic mitochondria in cancer cells may have a cytoprotective effect [322, 323]. Mitophagy is closely related to the metabolic reprogramming of tumor cells. Cancer cells generate energy through aerobic glycolysis even under normal oxygen conditions, and the unique metabolic profile of reduced oxidative phosphorylation and enhanced aerobic glycolysis in cancer cells is known as the Warburg effect [324]. Defective mitophagy leads to the Warburg effect, mitochondrial metabolic changes, and associated enhancement of glycolysis. Mitophagy deficiency increases the expression of exokinase 2, resulting in enhanced glycolysis in liver cancer and promoting tumor growth [325]. P53/BNIP3-dependent mitophagy suppresses tumor growth by inhibiting glycolysis in radioresistant cancer cells of head and neck squamous cell carcinoma [326]. High expression of CD44 is linked to enhanced malignant potential of esophageal squamous cell carcinoma, and Parkin-mediated mitophagy promotes CD44+ activity, while inhibition of mitophagy leads to oxidative stress and cell death [327]. NIX-mediated mitophagy promotes the clearance of ROS induced by hypoxia in highly invasive glioblastoma, maintains cancer stem cells, and promotes tumor cell survival [328]. In CSCs, mitophagy is utilized as a positive regulatory mechanism to alter their metabolic state for better adaptation to various metabolic stresses [301]. Mitophagy maintains the stemness of liver cancer CSCs by removing the tumor suppressor p53 [329]. Additionally, Towers et al. [157] have revealed that autophagy-dependent cancer cells can utilize MDVs to compensate for the removal of damaged mitochondria and preserve mitochondrial homeostasis in the presence of mitophagy inactivation.

Tumor immunity

Immune cells participate in the regulation of tumor biology, wherein appropriate mitochondrial function is of crucial importance for the phenotype, proliferation, and differentiation of immune cells [330]. However, driven by tumor progression, nutrient depletion, hypoxia and production of immunosuppressive metabolite in the tumor microenvironment (TME) can alter anti-tumor activities of immune cells, ultimately contributing to immune escape [331]. Therefore, understanding the MQC changes of immune cells in the TME is a promising avenue in cancer research.

As the principal effectors of anti-tumor immunity, functional cytotoxic CD8+ T cells are closely related to the survival of cancer patients [332, 333]. Naïve CD8+ T cells maintain their low metabolic demands by relying on OXPHOS; while the T cell receptor (TCR) is activated, metabolic reprogramming occurs in effector T cells, including aerobic glycolysis, OXPHOS, and glutaminolysis to meet the requirements of proliferative outbreak and effector functions [332, 334]. Both metabolic stresses of T cells driven by TME fuel depletion and increased expression of inhibitory receptors like programmed cell death-1 (PD-1), cytotoxic T lymphocyte-associated protein 4 are the essential factors that result in tumor-infiltrating CD8+ T cells exhaustion [332, 335]. Experiments have demonstrated that the TME can impair PGC-1α-mediated mitochondrial biogenesis and function in tumor-infiltrating CD8+ T cells, in part by chronic activation of Akt, and these processes can be reversed by PGC-1α overexpression [336]. In addition, nuclear receptor coactivator 2 has also been proven to be an upstream factor in the promotion of PGC-1α expression, which is involved in the regulation of CD8+ T cell-mediated anti-tumor immune responses [333]. Malinee et al. [337] reported that EnPGC-1, an epigenetic activator of PGC-1α/β, enhances mitochondrial biogenesis and increases OXPHOS and FAO in effector T cells, improving the synergistic effect of PD-1 blockade therapy. Notably, enforced mitochondrial biogenesis can promote differentiation of CD8+ T cells into memory T cells with better recall capacity [338]. Mitochondrial dynamics also play a crucial role in controlling T cell phenotype. During T cell activation, DRP1-mediated mitochondrial fragmentation can be observed and migrate to the immune synapses, inducing metabolic gene transcription through modulating the balance between AMPK and mTOR/cMyc under an appropriate calcium current [88, 339, 340]. Here, loosely attached mitochondrial cristae in effector T cells lead to lower ETC efficiency and promote aerobic glycolysis [339, 341]. Correspondingly, OPA1-mediated mitochondrial fusion or DRP1 deletion seems to induce the generation of memory T cells. As a result, abnormal mitochondrial dynamics may result in T-cell dysfunction and impaired anti-tumor response. It has been reported that PD‐1 signaling prevents DRP1-mediated mitochondrial fission via regulating mTOR and ERK pathways to impair CD8+ T cell function in TME [342]. Mitophagy in CD8+ T cells is also influenced by the TME. The combined challenge of TCR stimulation, microenvironmental stressors, and PD-1 signals inhibit mitophagy in tumor-infiltrating CD8+ T cells, which leads to the accumulation of depolarized mitochondria and ultimately T cell exhaustion [343]. Persistent activation of ER stress in TME causes T cell dysfunction, whereas interestingly, transient activation of ER stress by carbon monoxide promotes T cell anti-tumor function through driving mitochondrial biogenesis and protective mitophagy [344]. Moreover, Denk et al. [345] recently revealed that the pharmacological inducer of Pink1-dependent mitophagy in CD8+ T cells promotes T memory stem cell formation and anti-tumor effects in colorectal cancer, showing a beneficial effect in chimeric antigen receptor T-cell therapy.

Natural killer (NK) cells are the major effectors of innate immunity against cancer [346]. Increasing evidence suggests that changes in mitochondrial dynamics are connected to the anti-tumor capacity of NK cells. In comparison to the normal NK cells, small and fragmented mitochondria are observed in the tumor-infiltrating NK cells from patients with liver cancer, accompanied by lower mitochondrial mass [347]. To be more specific, the hypoxic state of TME causes mitochondrial fragmentation by increasing mTOR-DRP1 signaling in NK cells, resulting in abnormal mitochondrial respiration and tumor evasion of immune detection, which can be reversed by DRP1 inhibitor [347].

Tumor-associated macrophages (TAMs) are the most abundant components in the TME. After being recruited to TME by tumor-related cytokines, TAMs promote cancer angiogenesis, metastasis, and immunosuppression through the further secretion of cytokines, chemokines, growth factors, and matrix metalloproteinases [348]. In other words, the extent of TAM infiltration is closely associated with a poor prognosis of cancer. In tumor cells, mitochondrial fission mediated by DRP1 enhances the recruitment of TAMs by releasing mtDNA, which promotes the release of the chemokine CCL2, thus facilitating tumor progression [349]. ROS plays a crucial role in the inflammatory state of the TME and immune escape mediated by PD-L1 expression. However, TAMs that T cell immunoglobulin and mucin domain containing-4-positive exhibit elevated levels of oxidative phosphorylation and upregulate mitophagy to eliminate ROS and other hazardous factors, thus promoting tumor metastasis [350]. Furthermore, the activation of the IL-33/ST2 signaling pathway significantly upregulates mitophagy processes, thereby modulating the oxidative phosphorylation and polarization levels of M2 macrophages in the TME to inhibit tumor growth [351].

Overall, the mitochondrial function of immune cells, particularly the MQC system, is affected by complex TME, which may result in decreased tumor susceptibility and immune escape. Therapeutic strategies aimed at maintaining MQC stability in TME cannot be overlooked for their role in inhibiting tumor development (Fig. 3).

Fig. 3
figure 3

Mitochondrial quality control and cancer. a The role of mitochondrial quality control varies in different tumors. Mitophagy could remove dysfunctional mitochondria and hinder the proliferation of tumor cells and, on the other hand, the clearance of apoptotic mitochondria can remove the oxidative stress adaptation of tumor cells and subsequently promote tumor cell apoptosis. Tumor cells tend to have highly active mitochondrial biogenesis, and the high density of mitochondria provides both energy and metabolic intermediates for tumor proliferation. b In tumor-associated macrophages, mitophagy promotes oxidative phosphorylation and the clearance of ROS, thereby promoting the M2 phenotype of tumor macrophages. c In T cells, the insufficiency of mitochondrial fission, mitochondrial biogenesis, and mitophagy leads to the depletion of effector T cells. d In NK cells, the hypoxic microenvironment promotes mitochondrial fragmentation, which leads to the failure of immune surveillance. ATP adenosine triphosphate, DRP1 dynamin-related protein 1, CCL2 C–C motif chemokine ligand 2, TAM tumor-associated macrophage, mTOC1 mammalian target of rapamycin complex 1, Akt v-akt murine thymoma viral oncogene homolog, PD-1 programmed cell death protein 1, NK natural killer, IL interleukin, mtDNA mitochondrial DNA, OXPHOS oxidative phosphorylation, PGC-1α PPAR-γ coactivator-1α, ROS reactive oxygen species, TME tumor microenvironment

MQC and cardiovascular disease

The heart has a high energy demand, and mitochondria maintain the physiological function of cardiomyocytes by supplying large amounts of energy constantly and rapidly [352]. Thus, mitochondria account for 30% of cardiomyocyte volume [353]. Mitochondrial dysfunction is always accompanied by impaired MQC and is significantly linked to the development of various cardiovascular diseases (Fig. 4).

Fig. 4
figure 4

Mitochondrial quality control and cardiovascular disease. The common features of mitochondrial quality control in cardiovascular disease are the downregulation of mitochondrial biogenesis, a shift of mitochondrial dynamics to fission phenotypes, and the downregulation of mitophagy. In the early stage of most cardiovascular diseases, mitophagy is properly upregulated to compensate for mitochondrial quality disorders, but in the stage of disease progression, mitophagy is downregulated. Created by Biorender.com, accessed on 25 Aug 2023. DRP1 dynamin-related protein 1, MFN mitofusin, OPA optic atrophy, PGC-1α PPAR-γ coactivator-1α

Heart failure (HF)

HF is a complex clinical syndrome, which is characterized by insufficient cardiac output due to abnormalities of cardiac structure/function [354]. HF is typically the terminal state of various cardiovascular diseases. Recent evidence has demonstrated that MQC plays a crucial role in the process of HF.

In a clinical trial of HF, microarray analysis of myocardial tissues from cardiomyopathy HF patients and healthy controls demonstrated that the expression level of PGC-1α, the major regulator of mitochondrial biogenesis, was significantly reduced in the myocardium of HF patients [355]. Similarly, the protein expression of PGC-1α was found to be downregulated in both patients with HF and a rat model of HF, relating to the left ventricular ejection fraction (LVEF) [356]. Heart-specific PGC-1α deletion in mice induced failing heart phenotypes, manifested by reduced LVEF, enlarged left ventricular, and increased natriuretic peptide type B, accompanied by the suppression of energy metabolism and mitochondrial biogenesis [357,358,359]. However, some studies have found the opposite. Hu et al. [360] discovered that PGC-1α expression remained unchanged in both patients with HF and in mouse samples. Besides, overexpression of PGC-1α may induce uncontrolled mitochondrial proliferation in the cardiac ventricle and lead to cardiac dysfunction, contingent upon the restoration of mitochondrial biogenesis [361]. As a double-edged sword, the balance between excessive expression treatment of PGC-1α in maintaining mitochondrial biogenesis and cardiac function remains unclear.

Mitochondria are tightly packed between myofibrils in the adult cardiac myocytes, and therefore, mitochondrial dynamics were thought to be insignificant in this subpopulation [362]. However, recent studies have shown that mitochondrial fusion events exist in cardiomyocytes, as evidenced by rapid content mixing events between adjacent organelles and slower events between both neighboring and distant mitochondria, and are associated with cardiac contractility [363]. Mitochondrial dynamics is impaired in HF. Ablation of the OPA1 gene is embryonically lethal, and heterozygous mice exhibit impaired mitochondrial function, susceptibility to hemodynamic stress, and decreased function in the heart, although it can survive [364, 365]. In humans and rats with HF, the reduced expression of OPA1 in the cardiomyocytes is characterized by small and fragmented mitochondria, which may be due to an imbalance between L-OPA1 and S-OPA1, the release of cytochrome C and deacetylation levels [80, 364, 366, 367]. Decreased mitochondrial fusion proteins MFN1/2 have been in the myocardium of mice with HF [368]. Mid-gestational deletion of MFN1/2 resulted in dilated cardiomyopathy and HF in mice on day 7 postnatal, accompanied by a decrease in mitochondrial biogenesis and mitophagy [369]. Yue and colleagues [370] found that prolonged activated Yes-associated protein 1 disrupted mitochondrial dynamics and triggered mitochondrial dysfunction by targeting Dnm1l and MFN1 in a mouse model of abdominal aortic constriction-induced HF, resulting in cardiac hypertrophy, and this process was also related to impaired mitochondrial biogenesis. However, the effects of knockout of MFN1/2 alone on cardiac function in mouse cardiac myocytes remain controversial [371,372,373]. Similarly, DRP1 was upregulated in HF patients and mammal models [374,375,376]. With the treatment of hypertrophic agonist norepinephrine in neonatal rat cardiomyocytes, mitochondrial fission was enhanced followed by a decrease in mitochondrial function. Mechanistically, norepinephrine increased cytoplasmic Ca2+ to activate calcineurin through targeting α1-adrenergic receptors, promoting DRP1 migration to mitochondria, involving dephosphorylation of DRP1 at Ser637 [377]. Interestingly, Xu et al. [378] demonstrated that chronic stimulation of β-adrenergic receptors by isoproterenol-induced mPTP openings and mitochondrial damage in failing hearts through activating CaMKII followed by phosphorylation of DRP1 at Ser616, independent of phosphorylation at Ser637. Pharmacological inhibition of DRP1 can reverse disordered mitochondrial dynamics and cardiac hypertrophy [378, 379]. However, DRP1 deficiency still does not favor mitochondrial health and cardiac function, due to damaged mitophagy and PCD changes, which seem to be associated with the duration of DRP1 deletion [380, 381]. Recently, Donnarumma et al. [382] found that mitochondrial inner membrane protein mitochondrial fission process 1 affected cardiac pathological events by targeting mPTP and mitochondria uncoupling of the inner membrane, with no effect on mitochondrial division.

There is growing evidence suggests that maladaptive mitophagy exacerbates mitochondrial dysfunction and HF [85, 383, 384]. Given the core role of PINK1-mediated MFN2 phosphorylation, which induces the localization of Parkin on the mitochondrion, the PINK1/Parkin pathway is essential for proper MQC. Decreased protein levels of PINK1 and Parkin have been observed in HF samples from patients and mouse models [383, 385]. Supporting these data, PINK1−/− mice exhibit left ventricular dysfunction and cardiac hypertrophy at 2 months of age, accompanied by increased oxidative stress and damaged mitochondrial function [385]. Mechanistically, the isoform switch from AMPKα2 to AMPKα1 during HF resulted in the inhibition of PINK1/Parkin/SQSTM1-mediated mitophagy, aggravating mitochondrial dysfunction and severe cardiomyocyte apoptosis via targeting PINK1 dephosphorylation [383]. Correspondingly, specific activators of AMPKα2 may be a potential clinical translational therapy to attenuate HF. Interestingly, mitophagy was transiently increased during early pressure overload as a protective role, depending on the temporary activation of conventional mitophagy and ULK1-dependent alternative mitophagy [386]. Pharmacological inducer of PINK1/Parkin-mediated mitophagy has also been proven to mitigate cardiac function [384, 387]. However, Chaanine et al. [388] found that in the case of pressure overload, JNK signaling induced activation of BNIP3-dependent mitophagy and impaired cardiomyocytes, exacerbating HF. Therefore, it is crucial to identify the proper timing and method for targeting mitophagy in the treatment of HF.

Myocardial ischemia/reperfusion (I/R) injury and myocardial infarction (MI)

Acute MI is a major cause of death and disability worldwide. The absence of oxygen and nutrient supply at acute myocardial ischemia, excess oxidative stress, and mPTP opening at myocardial reperfusion result in severe biochemical and metabolic disorders in the cardiomyocytes, finally contributing to mitochondrial dysfunction and cardiomyocyte death [389, 390]. During these processes, mitochondrial morphology, and function change dramatically, implying that the MQC system may play important roles in MI.

The continuous renewal of mitochondria through mitochondrial biogenesis enables cardiomyocytes to rapidly adapt to new energy demands during I/R injury and MI. Recently, a full-length transcriptomic analysis in mice revealed that cardiac PGC-1α coding transcripts were decreased during I/R injury, which was associated with the infarcted area [391]. Several studies have suggested that cardiac PGC-1α expression and mitochondrial biogenesis were downregulated in mammal models of MI [392, 393]. In hypoxia/reoxygenation (H/R)-treated cardiomyocytes, the activity of AMPK was decreased with subsequent inhibition of TFAM and Nrf2, while melatonin treatment reactivated the AMPK/PGC-1α pathway to restore mitochondrial biogenesis [394]. PGC-1α-induced mitochondrial biogenesis also participated in cardiac H/R injury by regulating mitochondrial fusion [394]. Sirt1 enhances PGC-1α activity through deacetylation. It has been reported that the disruption of the Sirt1/PGC-1α pathway and increased apoptosis in cardiomyocytes was caused by I/R incubation [395]. Furthermore, due to the relationship between PGC-1α and energy metabolism, activation of the Sirt1/PGC-1α pathway also increased glucose uptake and pyruvate metabolism in cardiomyocytes [395]. It is certain that increased PGC-1α activity is conducive to maintaining mitochondrial homeostasis and alleviating cardiac dysfunction during I/R injury [394,395,396,397].

Under I/R stimulation, cardiomyocytes exhibit a shift toward mitochondrial fission and a loss of mitochondrial fusion, which in turn induces mPTP opening, ROS production, and eventually cardiomyocyte death [389]. The expression and activity of DRP1 increase significantly during MI or I/R, whereas pharmacological inhibition of DRP1 protects the heart from I/R injury and reduces infarct size [398,399,400]. Phosphorylation of Ser616 and dephosphorylation of Ser637 on DRP1 is related to mitochondrial fission. I/R injury induces dephosphorylation of DRP1 at Ser637 through upregulating PGAM5 and calcineurin [401, 402], while the activation of Rho-associated coiled-coil containing protein kinase 1 (ROCK1), cyclin-dependent kinase 1 (CDK1), protein kinase C isoform delta (PKCδ) and glycogen synthase kinase-3β (GSK-3β) or inhibition of Sirt3-AMPK pathway can simultaneously increase phosphorylation at Ser616 [402,403,404,405]. Besides, decreased miR-499 during MI enhances calcineurin-mediated dephosphorylation of DRP1 at Ser656, which induces mitochondrial fragmented and apoptosis [406]. More recently, Sun et al. [407] found that TBC domain family member 15 (TBC1D15) declined after myocardial I/R injury. It was shown to maintain mitochondria-lysosome contacts to regulate asymmetrical mitochondrial fission via TBC1D15-DRP1 interaction, promote autophagy flux, and preserve mitochondrial homeostasis. The DRP1 receptor MFF also plays a critical role in the regulation of mitochondrial fission following cardiac I/R injury [111]. Cardiac-specific DRP1 knockout mice exhibited mitochondrial elongation and were more susceptible to I/R injury [380]. Therefore, moderate control of mitochondrial fission balance is critical for protecting the heart against I/R injury. In contrast to mitochondrial fission, mitochondrial fusion usually participates in the maintenance of mitochondrial health and physiological function as a protective mechanism. After I/R injury, mitochondrial fusion regulator OPA1 is expressed abnormally [408, 409]. It has also been reported that the expression of OPA1 is decreased in the patient’s hearts with ischemic cardiomyopathy [366]. MCU is a unidirectional channel that controls Ca2+ inflow into mitochondria [408]. Guan et al. [408] found that MCU was upregulated during myocardial I/R, leading to mitochondrial calcium overload and activation of calpain, which subsequently inhibited OPA1-mediated mitochondrial fusion and mitophagy. The imbalance of L-OPA1 and S-OPA1 caused by incremental OMA1 also plays an important role during I/R injury [410]. The AMPK-OPA1-mitochondrial fusion/mitophagy axis is disrupted during cardiac I/R injury and contributes to mitochondrial stress and caspase-9-involved mitochondrial apoptosis, while this process can be reversed with the treatment of melatonin [411]. Unlike the effects of OPA1, the role of MFNs in cardiac I/R injury remains controversial. On the one hand, increasing the expression of MFN1/2 has been shown to protect cardiomyocytes against I/R injury, by facilitating mitochondrial fusion and suppressing mPTP opening [398]. The proteasome is responsible for MFN2 degradation while inhibiting proteasome can partly preserve MFN2 to retain mitochondrial integrity and protect against I/R injury [412]. Zhou et al. [413] recently demonstrated that MORN repeat-containing protein 4 exerts endogenous protection in ischemic cardiomyocytes by binding to MFN2 and promoting MFN2 phosphorylation at Ser422 in the ROCK2 complex. On the other hand, unexpectedly, dual ablation of MFN1 and MFN2 in acute hearts leads to MI resistance, even with mitochondrial dysfunction [414]. Similar phenomena have also been observed in mouse cardiomyocytes with single MFN2 deletion [415]. These observations may be related to the extra-mitochondrial pleiotropic actions of MFNs. MFN2 acts as a rope between mitochondria and ER [416]. In the heart, loss of MFNs under I/R stimulation disrupts the interactions between mitochondria and ER, thus attenuating mitochondrial Ca2+ overload and suppressing mPTP opening [414]. However, long-term deletion of MFNs is detrimental to cardiomyocytes, and may rapidly progress to dilated cardiomyopathy and cardiac death [417]. Given the diverse functions of proteins involved in mitochondrial dynamics, more experiments are needed to explicate the roles of these proteins during MI.

Recent evidence has demonstrated that cardiac I/R injury resulted in dysregulation of mitophagy. Cardiac Parkin deficiency makes mice much more sensitive to MI surgery but does not affect mitochondrial or cardiac function under physiological conditions [418]. In response to ischemic stress, PINK1/Parkin-mediated mitophagy is induced in both vivo and in vitro, although other studies have reported the opposite [419]. Increasing Parkin-related mitophagy during ischemia appears to be a cardioprotective event, partly because the Parkin-mitophagy pathway inhibits the opining of mPTP and cardiomyocyte necroptosis [418, 419]. However, the role of PINK1/Parkin-mediated mitophagy in cardiac I/R remains complicated and not yet fully understood. Several studies have revealed that mitophagy played a protective role in response to I/R stimulation. Mice lacking PINK1 are more vulnerable to I/R injury due to worsening mitochondrial function [420]. mPTP opens during reperfusion, while PINK1/Parkin‐mediated mitophagy inhibits this opening [420, 421]. Sun et al. [422] found that Parkin catalyzed the ubiquitination of cyclophilin-D (CypD) to block mPTP opining, thereby preventing cardiomyocytes from programmed necrosis and I/R injury, in addition to regulating mitophagy. Several drugs or compounds that pharmacologically activate PINK1/Parkin-mediated mitophagy have also been shown to have cardioprotective effects against I/R injury [423,424,425]. Conversely, excessive PINK1/Parkin-mediated mitophagy is detrimental during cardiac I/R. The PINK1/Parkin pathway is activated by I/R injury in vivo and in vitro, whereas knockdown or inhibition of Parkin protects cardiomyocytes from mitophagy and apoptosis [426, 427]. Notch1 signaling physiologically regulates cardiac development and cardiomyocyte proliferation, alleviates mitophagy and mitochondrial fragmentation by suppressing the PTEN/PINK1 pathway, and protects the heart from I/R injury [428]. Furthermore, it has been reported that melatonin downregulated the expression of mitophagy-related proteins (Parkin, Beclin1, and NIX) and diminished excessive mitophagy through the MT2/Sirt3/FoxO3a signaling pathway, thus attenuating H/R injury in H9c2 cells [429]. Notably, there is crosstalk between each MQC, and mitochondrial fission is often regarded as to be upstream of mitophagy. As a result, targeting mitochondrial dynamics-related proteins seems to attenuate myocardial I/R injury by controlling mitophagy [411, 430]. FUNDC1-mediated mitophagy plays a beneficial role in cardiac I/R injury. FUNDC1-mediated mitophagy in cardiomyocytes is induced by hypoxia challenge but inhibited during reperfusion [431, 432]. At the molecular level, I/R injury upregulates the expression of CK2α, which contributes to FUNDC1 phosphorylation at Ser13 and subsequent inhibition of FUNDC1-mediated mitophagy [432]. Similarly, elevated RIPK3 also inhibits mitophagy following I/R injury via post-transcriptional modification of the FUNDC1 phosphorylation site [433]. Mao et al. [432] recently found that cardiac I/R stimulation suppressed the expression of Polo-like kinase 1 (PLK1), and thus counteracted the induction of PLK1 and FUNDC1-dependent mitophagy [434]. Impaired FUNDC1-mediated mitophagy hinders the clearance of damaged mitochondria induced by I/R injury, thus facilitating mitochondrial apoptosis and impairing cardiac function. Furthermore, FUNDC1-dependent mitophagy in platelets is associated with platelet activation and cardioprotective effect during I/R injury [431, 435]. A recent study revealed that FUNDC1-mediated mitophagy can maintain MQC and alleviate myocardial I/R damage by activating the mitochondrial unfolded protein response (UPRmt) [180]. It has been reported that mitophagy is enhanced through the HIF-1α/BNIP3 signaling pathway, and berberine could further induce this process to protect against myocardial I/R injury [436, 437]. However, Jin et al. [111] found that downregulation of dual-specificity protein phosphatase1 after cardiac I/R injury promotes excessive BNIP3-mediated mitophagy via the JNK pathway and resulted in cardiomyocyte death. Additionally, vitamin D-mediated cardio protection against I/R injury is related to the inhibition of BNIP3-mediated mitophagy and apoptosis [438]. Due to its additional involvement in apoptosis, the effects of BNIP3 and its dependent mitophagy on cardiac I/R injury remain uncertain. Proper regulation of mitophagy may efficiently remove damaged mitochondria, but excessive or insufficient mitophagy appears to aggravate I/R injury. Therefore, potential methods for maintaining baseline mitophagy following I/R injury warrant further research.

Atherosclerosis

Atherosclerosis is a chronic and progressive vascular disease based on inflammation that predisposes patients to MI, ischemic cardiomyopathy, strokes, and peripheral arterial disease [439]. Convincing evidence has revealed the essential effects of MQC acting in the pathogenesis of atherosclerosis.

Endothelial cell damage induced by ROS is the primary event of atherosclerosis [440]. After oxidized low-density lipoprotein (ox-LDL) treatment in human aortic endothelial cells (HAECs), the expression of PGC-1α is inhibited, accompanied by mitochondrial energy metabolism disorder, increased mtROS and apoptosis [440]. Consistent with this, PGC-1α protein levels were also downregulated in Ang-II-induced atherogenesis in ApoE−/− mice [441]. Karnewar et al. [441] reported that mitochondria-targeted esculetin reduced oxidative stress and increased mitochondrial biogenesis, exhibiting an anti-atherogenic effect in endothelial cells as well as in mouse models.

Endothelial dysfunction is conducive to the development of atherosclerosis in patients with diabetes due to altered mitochondrial dynamics and subsequent increased ROS, as evidenced by higher FIS1 expression and lower mitochondrial network extent [442]. Consistently, mitochondrial fragmentations were observed by electron microscopy in diabetic mouse aortic endothelial cells [443]. Liu et al. [444] found that DRP1 knockdown in human umbilical vein endothelial cells (HUVECs) alleviated ox-LDL-induced mitochondrial fission and apoptosis. Retinol binding protein 4, an IR-related adipokine highly expressed in the serum of patients with metabolic syndrome, disrupts the homeostasis of mitochondrial fusion and fission in endothelial cells, resulting in higher levels of DRP1 and FIS1 as well as lower levels of MFN1 [445]. Treatment with metformin can improve endothelial function and reduce atherosclerotic lesions in diabetic ApoE−/− mice by regulating DRP1-mediated mitochondrial fission [443]. Mitophagy is also implicated in the development of atherosclerosis. Under high-glucose conditions, PINK1/Parkin-mediated mitophagy was impaired in endothelial cells, leading to mitochondrial dysfunction and apoptosis [446]. Furthermore, Xia et al. [350] demonstrated that ox-LDL challenge upregulated the expression of PTEN, which suppressed mitophagic flux through the AMPK-CREB-MFN2 pathway, leading to apoptosis in HUVECs. On the contrary, higher levels of nuclear receptor subfamily 4 group A member 1 (NR4A1) induced by ox-LDL in endothelial cells triggered excessive Parkin-dependent mitophagy, whereas NR4A1 deletion was shown to protect endothelial cells from energy metabolism disorder and apoptosis [447]. In addition to endothelial cells, the phenotypic transformation of vascular smooth muscle cells (VSMC) plays a critical role in the pathogenesis of atherosclerotic plaque formation. Ox-LDL may lead to lipid deposition and foam cell formation in VSMC. During this process, VSMC exhibits mitochondrial over-fission and decreased mitochondrial branch length, which can be reversed by midiv-1 [448]. Similarly, platelet-derived growth factor (PDGF), involved in recruiting VSMCs to the neointima and promoting atherosclerosis development, promotes mitochondrial fragmentation and reduction of MFN2 in VSMCs [449]. Wang et al. [450] also reported that silencing DRP1 could inhibit VSMC migration induced by PDGF and decrease pathological intimal hyperplasia in mice. In an atherosclerosis mouse model, apelin-13 promoted PINK1/Parkin-mediated mitophagy of VSMCs via activating AMPKα thereby aggravating the progression of atherosclerotic lesions [451]. The authors also found that apelin-13 alters the balance of mitochondrial dynamics (increasing DRP1, decreasing MFN1, MFN2, and OPA1), inducing a proliferation phenotype in VSMCs [451].

Ox-LDL stimulation promotes mitochondrial dysfunction and inflammatory response in monocyte-macrophage, which participates in chronic inflammatory diseases, including atherosclerosis. Mechanically, increased methyltransferase-like 3 induced by ox-LDL coordinates with YTHDF2 to suppress the expression of PGC-1α [452]. A recent study reported that enhanced DRP1-mediated mitochondrial fission induced macrophage M1 polarization and foam cell formation through the mito-ROS/NLRP3 inflammasome singling pathway, hence accelerating atherogenesis [453]. Macrophage apoptosis is considered to be a critical step in the formation of a pro-thrombotic necrotic core. Enhancing CD137 (a kind of T cell co-stimulatory molecule) signaling facilitates the progression of atherosclerotic plaque in the ApoE−/− mice, where elevated TUNEL co-staining with the CD68 marker is observed. At the molecular level, CD137 signaling can induce mitochondrial fission through the p38 MAPK pathway in mouse peritoneal macrophages, resulting in mitochondria dysfunction and macrophage apoptosis [454].

Mitophagy is considered to protect the stability of macrophage function. It has been reported that inhibiting caspase 1 antagonizes NLRP3 inflammasome assembly, prevents foam cell formation and pyroptosis in macrophages, partly by boosting Parkin-mediated mitophagy and efferocytosis, and thereby ameliorates vascular inflammation and atherosclerosis [455]. This study demonstrated that targeting the interplay between NLRP3 inflammasome activation and dysfunction of MQC could be a potential therapeutic strategy for atherosclerosis. Apolipoprotein A-I binding protein, recently identified as an autophagy regulator in macrophages, regulates macrophage M1/M2 polarization and performs an anti-atherosclerotic role via PINK1-dependent mitophagy [456, 457]. Notably, despite the associated beneficial metabolic effects, high-protein diets induce the formation of atherosclerotic plaque in mice. Mechanically, elevated levels of amino acid in the blood and atherosclerotic plaque activated mTORC1 in macrophages, leading to repression of downstream mitophagy and consequently exacerbating mitochondrial dysfunction and apoptosis [458].

Cardiac aging

Due to the significant increase in human life expectancy, aging-associated cardiovascular diseases are becoming more prevalent and raising concerns due to their substantial impact on health, quality of life, and socioeconomic burdens [459]. The aging hearts exhibit unique histological and morphological features, which are linked to the development of cardiac dysfunction [460]. Specifically, aged hearts experience impaired mitochondrial function characterized by changes in mitochondrial morphology, mPTP opening, MQC dysfunction, and ROS formation [461].

Mitochondrial biogenesis in cardiac senescence-associated mitochondrial dysfunction has attracted extensive attention. In various aging animal models, a decrease in the expression of PGC-1α in cardiac tissue was observed [462,463,464]. Contrastively, overexpression of PGC-1αin PGC-1α muscle-specific transgenic mice counteracted age-associated pathological changes in the heart [465]. Wang et al. [462] demonstrated that spermidine supplementation enhanced mitochondrial biogenesis and function via the Sirt1/PGC-1α signaling pathway and thus ameliorated cardiomyocyte aging in rats. Alterations in mitochondrial dynamics have also been associated with cardiac aging. Reduced levels of MFN1 and MFN2 were found in the hearts of 25-month-old rats [464]. Furthermore, Fernández-Ortiz and colleagues [466] discovered decreased protein levels of MFN2, OPA1, and DRP1 both in mature mice (12 months old) and old mice (24 months old), compared with young mice (3 months old). In contrast, upregulated expression of OPA1 and DRP1 was reported in the hearts of 36-month-old rats by Ljubicic et al. [467]. Despite the discrepancies in mitochondrial dynamics-related proteins observed across different studies, which can be attributed to dynamic changes occurring at different ages, these findings collectively indicate that there are alterations in mitochondrial dynamics during cardiac aging. Transmission electron microscopy has revealed that aging hearts exhibit a diminished capacity to regulate mitochondrial dynamics, and are more susceptible to I/R injury, which is related to the age‐related deficiency of Sirt1 and Sirt3 [468]. Treatment with melatonin has been shown to enhance the response of mitochondria dynamics during cardiac aging by upregulating mitochondria fusion and fission proteins [466]. Overall, maintaining a synthetic balance between mitochondrial fusion and fission may be a realistic strategy for preserving cardiac health throughout late life rather than targeting specific phases of mitochondrial dynamics.

Mitophagy progressively declines with age in the heart. Deletion of Parkin results in premature cardiac aging in mice, as evidenced by the accumulation of mtDNA deletion mutations, the decline in cardiac functional reserve, and increased SA-β-gal activity [469]. The foregoing process can be reversed by inducing Parkin overexpression, implying that Parkin-mediated mitophagy serves a cardioprotective role in cardiac aging. P53, a transcription factor associated with senescence signal, inhibits Parkin translocation to mitochondria by binding to the RING0 domain, hence suppressing mitophagy and enhancing cardiac aging [469]. It has been reported that Shank3 expression is increased in aging cardiac tissue, and Shank3 knockout alleviates age-induced cardiac dysfunction [470]. Mechanically, Shank3 binds with CaMKII to impede its translocation to the mitochondria, which suppresses CaMKII activation and Parkin-mediated mitophagy, further triggering cardiomyocyte apoptosis. Additionally, both RhoA and lncRNA LOC105378097 also have been found to modify mitophagy as upstream regulators affecting heart aging and cardiac dysfunction in mice [471, 472]. Gao et al. [473] recently revealed that advanced aging dampens mitophagy by reducing the expression of Parkin, microtubule-associated protein light chain 3 II, phosphorylation of p62 and TBK1, whereas Parkin overexpression could rescue cardiac aging by promoting K63-linked polyubiquitination of TBK1 to facilitate mitophagy. Together, these data suggest that mitophagy plays a beneficial role during cardiac aging.

Hypertension

Hypertension is one of the most common cardiovascular disorders worldwide [474]. Endothelial dysfunction of the vascular structure and activation of the sympathetic nervous system are major pathogenic factors contributing to hypertension [475]. The paraventricular nucleus (PVN) plays a crucial role in the regulation of sympathetic output and salt appetite [476], making it of pivotal importance in hypertension. Sun et al. [476] found that PGC-1α and DRP1 expression were decreased while MFN2 expression was elevated in the PVN of spontaneously hypertensive rats, alongside increased oxidative stress. This observation indicates that MQC and mitochondrial function are altered during hypertension. Certain compounds such as oleuropein and punicalagin have been shown to attenuate hypertension partially by improving mitochondrial biogenesis and dynamics [476, 477]. Microglial neuroinflammation in the rostral ventrolateral medulla is also associated with stress-induced hypertension [478]. In a rat model of stress-induced hypertension, decreased Sigma-1R (an ER chaperone protein) levels result in decreased ER-mitochondria contact and mitochondrial hyperfusion, hence inducing microglial M1 polarization [478]. Enhanced mitochondrial fission was also observed in the arterial media of angiotensin II (Ang II)-induced hypertensive mice and midiv-1 treatment prevented Ang II-induced VSMC phenotypic switching and hypertension by inhibiting mitochondrial fission and oxidative stress [479].

MQC and metabolic disease

IR and diabetes mellitus (DM)

The prevalence of DM, a chronic metabolic disorder, is rapidly increasing worldwide [480]. T2DM, accounting for 90 – 95% of cases, represents the major subtype of DM, IR, often accompanied by defects in insulin secretion, is considered to be a primary contributor to the development of T2DM [481]. Given its role in energy metabolism, maintaining mitochondrial homeostasis is of crucial importance to the pathophysiology of DM.

Accumulating evidence suggests that mitochondrial biogenesis is impaired in individuals with DM. Specifically, the expression of PGC-1α is significantly reduced in the liver and skeletal muscle of diabetic mice [481,482,483]. The liver and muscles serve key functions in regulating blood glucose homeostasis. The imbalance between glucose release from the liver and uptake from muscle and adipose tissue contributes to the development of IR and DM [484]. In diabetic hepatocytes, suppressed PGC-1α leads to hepatic mitochondrial dysfunction and increased gluconeogenesis. However, restoring redox balance through a liver mitochondrial-targeting antioxidant nano-mitoPBN can promote mitochondrial biogenesis and enhance glucose catabolism via the AMPK/Sirt3/PGC-1α axis, effectively preventing diabetes in diabetic mice [484]. Additionally, Zhu et al. [482] demonstrated that activating the Sirt1/PGC-1α/MFN2 signaling pathway could enhance mitochondrial biogenesis and alleviate IR in the liver. Nevertheless, various evidence also suggests that PGC-1α drives gluconeogenesis in the liver, leading to increased blood glucose [485]. It has been reported that PGC-1α and TFAM protein levels are decreased in skeletal muscle of db/db mice, indicating a reduction in mitochondrial biogenesis. Catalpol can promote mitochondrial biogenesis in skeletal muscle and increase glucose uptake and ATP production, thereby ameliorating IR primarily through activation of AMPK/PGC-1α/TFAM signaling [481, 483].

DM is related to an imbalance in mitochondrial dynamics. Myotubes derived from patients with T2DM exhibited increased mitochondrial fragmentation and decreased mitochondrial content, accompanied by impaired mitochondrial lipid oxidation and respiratory capacity [486]. Further, Jheng et al. [487] found that protein expressions of DRP1 and FIS1 were significantly increased in the skeletal muscle of ob/ob mice, and inhibiting mitochondrial fission with mdivi-1 improved insulin signaling and insulin sensitivity. Muscles from obese or T2DM patients showed repression of MFN2 expression. It has been reported that liver-specific deletion of MFN2 in mice resulted in glucose intolerance and impaired insulin response [71]. Furthermore, MFN2 deficiency impairs insulin signaling and insulin sensitivity in muscle and liver tissues through mechanisms involving ROS production and ER stress [71]. A recent study demonstrated that high-intensity interval training preserved fasting blood glucose and glucose homeostasis in T2DM mice through remodeling the balance of mitochondrial dynamics (increasing MFN2, DRP1, and FIS1) as well as improving glycolipid metabolism [488]. Enhanced and sustained mitochondrial fragmentation was observed in the white adipose tissue (WAT) from ob/ob mice, with higher expression of DRP1 but lower expression of MFN2 and OPA1 [489]. Midiv-1 therapy could attenuate mitochondrial dysfunction and induce white-to-beige adipocyte transdifferentiation by promoting mitochondrial biogenesis and fusion-to-fission balance [489].

As an adaptive mechanism, increased mitophagy in individuals with pre-diabetes is considered to eliminate dysfunctional mitochondria and modulate mitochondrial oxidative stress, thereby preventing the development of T2DM. However, mitophagy is impaired and higher levels of ROS further exacerbate mitochondrial dysfunction during T2DM [490]. Compared to lean individuals, myotubes derived from T2DM patients were found to exhibit significantly lower expression of Parkin [486]. Heat shock protein (HSP) 72, a stress-inducible chaperone protein, exhibited decreased levels in the muscles of obese and T2DM patients. Drew et al. [491] discovered that deletion of HSP72 resulted in decreased respiratory capacity and IR in muscle, leading to an IR-obesity phenotype in mice. Mechanically, during mitochondrial stress, HSP72 rapidly translocates to mitochondria where it interacts with MFN2 and subsequently forms a complex with Parkin, thereby maintaining mitochondrial morphology and autophagic signaling [491]. Liver-specific deficiency of Parkin also impairs mitochondrial respiratory capacity, and results in hepatic steatosis and IR in HFD-fed mice, implying a link between mitophagy and liver IR [492]. Chronic low-grade inflammation is closely associated with T2DM. A recent clinical study found a significant decrease in mitophagy-related proteins PINK1 and Parkin within peripheral blood mononuclear cells from patients with T2DM [493]. Furthermore, Gupta et al. [494] observed that under palmitate and diabetic conditions, disturbances in mitochondrial homeostasis were accompanied by elevated inflammatory IL-1β response in macrophages. At the molecular level, palmitate increases FOXO3a acetylation and prevents its binding to the PINK1 promoter, consequently downregulating PINK1-mediated mitophagy while enhancing NLRP3 inflammasome activation [494]. Similarly, FUNDC1-dependent mitophagy is implicated in the mitochondrial function of WAT. It has been reported that FUNDC1 deletion mice fed an HFD exhibited obesity and IR, partly due to decreased mitophagy and mitochondrial quality in WAT, resulting in WAT remodeling and inflammation [198].

Obesity

Obesity is a high-risk factor and trigger for many diseases, including T2DM, cancer, cardiovascular disease, dyslipidemia, metabolic syndrome, liver disease, chronic kidney disease (CKD), and other diseases [495]. Epidemiological studies have shown a significant global rise in the prevalence of obesity patients, imposing an immense burden on the healthcare system in terms of managing obesity and its associated disorders [496]. In recent years, an increasing body of studies has confirmed that MQC is one of the main molecular mechanisms underlying obesity [198, 497]. This suggests that ameliorating obesity may be facilitated by reversing mitochondrial dysfunction and regulating mitochondrial quality. As described below, MQC-mediated regulation of obesity primarily stems from its influence on skeletal muscle cells and adipocytes.

Obesity is strongly related to low-grade chronic inflammation, and dysfunction of adipocyte mitochondrial is a major contributor to inflammation in adipose tissue [498]. The role of mitochondria in WAT has long been overlooked due to their limited presence. A variety of proteolytic enzymes in mitochondria, such as ClpP and Lonp1 located in the matrix, and Yme1L and m-AAA located in the inner membrane, are involved in the regulation of MQC [499]. Previous studies have confirmed that OMA1, the metalloprotease, directly regulates and deactivates the motility-related GTPase OPA1 [500, 501]. In OMA1 mutant mice, inhibition of OPA1 hydrolysis disrupts the balance between mitochondrial fusion and fission, thereby aggravating obesity and thermogenic disorders [500]. In adipocytes specifically lacking Parkin protein expression through knockout techniques, mitophagy is slightly reduced while the stability of PGC-1α is enhanced by increased NQO1 protein levels. This promotes the biogenesis of mitochondria and ultimately resists HFD and aging-induced obesity [497]. The deficiency of the FUNDC1 receptor for mitophagy in WAT leads to defective mitophagy and impaired MQC, thereby exacerbating HFD-induced obesity [198]. Therefore, active regulation of MQC in adipocytes has been found to promote metabolic homeostasis and reduce the risk of obesity, thus highlighting the important role of adipose mitochondria.

Skeletal muscle, which constitutes a significant portion of the body, plays a crucial role in regulating energy balance by efficiently utilizing fat and glucose. Mitochondria are highly abundant in skeletal muscle, and any impairment in their function directly affects skeletal muscle performance, potentially leading to chronic diseases such as diabetes, obesity, and aging [194]. Among the elderly population, there is a considerable number of obese individuals who also suffer from sarcopenia, known as sarcopenic obesity, further exacerbating physiological decline [502]. Previous research has shown that the administration of the mitochondrial uncoupler BMA15 can enhance mitochondrial function and attenuate age-related loss of muscle mass and function by activating MQC. This protective effect against sarcopenic obesity was observed in obese mice at 80 weeks [503]. Furthermore, an intervention study involving 81 elderly obese patients showed that high-intensity interval training combined with l-citrulline, a potential pharmacological agent for enhancing mitochondrial and muscle function, as well as adipose tissue metabolism based on findings from aged rodent model, may increase mitophagy, mitochondrial fusion, and mitochondrial biogenesis, resulting in increased muscle strength and reduced adipose tissue accumulation [504]. Maintaining the balance of MQC in skeletal muscle also contributes to the reduction of dietary obesity. The skeletal muscle-specific knockout of FUNDC1, a mediator of mitophagy, can lead to defective mitophagy and impaired mitochondrial energy production. However, it induces a retrograde response that upregulates the expression of FGF21, promoting thermogenic remodeling of adipose tissue, protecting against HFD-induced obesity, and enhancing glucose tolerance and insulin sensitivity [505]. Muscle-specific brain-derived neurotrophic factor increases the mitochondrial content of skeletal muscle, enhances mitochondrial fission and mitophagy, and maintains the balance of MQC in skeletal muscle, thereby reducing dietary obesity in mice [506]. In summary, targeting MQC can reduce dietary obesity and age-related sarcopenic obesity by regulating the communication between skeletal muscle and fat tissues.

MQC and nervous system disease

Neural cells require high concentrations of ATP to maintain their normal function, implying a need for a robust rate of mitochondrial turnover. Dysregulation of mitochondrial homeostasis is involved in various neurological disorders, in which the disruption of MQC is a central factor. Below, we discuss the relationships between MQC and common neurological disorders (Fig. 5).

Fig. 5
figure 5

Mitochondrial quality control and nervous disease. a In amyotrophic lateral sclerosis, a downregulation of PGC-1α mediates a decrease in mitochondrial biogenesis, excessive mitochondrial fission, and impaired mitophagy, which are key features of the disease. b Ischemic brain injury exhibits similar mitochondrial phenotypic characteristics, and mitigating brain tissue damage and oxidative stress levels can be achieved through upregulation of mitochondrial biogenesis mediated by PGC-1α and mitophagy mediated by Parkin-2. c Accumulated Aβ in Alzheimer’s disease activates DRP1, leading to impaired mitochondrial dynamics. Additionally, Tau can interact with Parkin, affecting the clearance of damaged mitochondria. Dysfunctional mitophagy further accelerates Aβ accumulation, forming a positive feedback loop of mitochondrial damage. Mutations in PRKN and PINK1 have been recognized as important genetic factors in Parkinson’s disease, and impaired mitophagy can further activate NLRP3, inducing neuroinflammation. e The key feature of Huntington's disease is the activation of DRP1 by mHtt, triggering severe mitophagy. Created by Biorender.com, accessed on 25 Aug 2023. DRP1 dynamin-related protein 1, mHtt mutant Huntingtin, PGC-1α PPAR-γ coactivator-1α, PINK1 PTEN-induced kinase 1, ROS reactive oxygen species

Alzheimer’s disease (AD)

AD is the most prevalent neurodegenerative disorder characterized by deposition of Tau and Aβ [507]. Previous evidence showed that mitochondrial disruption in neurons is concentrated near Aβ [508], indicating the involvement of mitochondrial dysfunction in the progression of AD. More studies have revealed that impaired mitochondrial function in neurons precedes Tau and Aβ deposition, while Tau and Aβ in turn promote mitochondrial dysfunction, thus constituting a vicious cycle [509,510,511]. Tau can bind to Parkin and block its translocation to depolarized mitochondria, thereby inhibiting the clearance of damaged mitochondria [512]. In early synaptic distribution defects observed in AD, Parkin-mediated mitophagy is widely activated in neurons with Tau lesions, accelerating the turnover of mitochondrial Rho GTPase1 and preventing mitochondrial flow to the synapses for replenishment [513]. Therefore, it is crucial to investigate specific mechanisms underlying Tau-influenced mitophagy at different stages or sites of AD pathology for more precise therapeutic targets. Moreover, maintaining mitochondrial protein homeostasis through mitophagy can effectively delay the deposition of Aβ [514]. As a source of Aβ, elevated APP-CTF within neurons is also associated with impaired mitophagy in early AD [515]. Nitric oxide released from Aβ induces S-nitrosylation of DRP1, leading to excessive mitochondrial fission and neuronal synaptic damage, and aggravating AD progression [516]. Further, Aβ interacts with LONP1 inhibiting protease activity, leading to disruption of mitochondrial proteostasis and dysfunction [517]. These findings suggest a broad interaction between Aβ/Tau and impaired MQC in AD, and targeting the MQC may be a promising therapeutic intervention. In AD, there is impaired mitochondrial biogenesis due to decreased levels of PCG-1α, a transcriptional regulator essential for mitochondrial biogenesis in AD [518]. Moreover, genetic factors such as APOE 4 carrier population, may also contribute to the dysregulation of MQC, leading to impaired mitochondrial biogenesis and dynamics, and an increased susceptibility to AD [519].

Parkinson’s disease (PD)

PD is a neurodegenerative disease characterized by a unique motor phenotype and pathological features, including the presence of Lewy bodies and loss of nigra neurons [520]. Growing evidence emphasizes the significance of MQC in PD pathogenesis. Loss of function mutations in PRKN and PINK1 are the most common genetic factors associated with early-onset hereditary PD, as they disrupt mitophagy and lead to the accumulation of damaged mitochondria in neurons [521, 522]. Of note, genetic forms of PD extend beyond Parkin and PINK1 mutations; for example, mutations in VPS35 can enhance its interaction with DLP1, thereby promoting turnover of the mitochondrial DLP1 complex through transport via mitochondria-derived vesicles to lysosomal degradation, ultimately promoting mitochondrial breakage [523]. Moreover, Parkin inactivation results in the accumulation of PARIS, a Parkin-interacting substrate, which represses PGC-1α promoter activity, and may further inhibit mitochondrial biogenesis [524]. Impaired mitophagy accompanied by subsequent release of mtDNA triggered by PINK1/Parkin deficiency may be involved in the process of neuroinflammation. This has been demonstrated in a cohort study by Borsche et al. [223], who found that both IL-6 and mtDNA were significantly upregulated in the plasma of PINK1/Parkin-associated PD patients. Moreover, a recent study revealed that Parkin normally inhibits inflammasome priming through ubiquitination-mediated targeting and proteasomal degradation of NLRP3; however, under conditions associated with PD, inactivated Parkin leads to inflammasome activation and promotes neuroinflammation [525]. These findings suggest that impaired MQC not only causes cellular energy defects in neurological diseases but may also be involved in neuroinflammation and broader pathological processes.

Huntington’s disease (HD)

HD is an autosomal dominant inherited progressive neurodegenerative disorder characterized by chorea and dystonia, incoordination, cognitive decline, and behavioral challenges [526]. Mutant Huntingtin (mHtt) disrupts the function of PINK1, leading to a decrease in the targeting of mitochondria to autophagosomes and promoting the fusion between damaged mitochondria [527]. A recent study discovered that abnormal binding of mHtt to ULK1 and BECN1 is a consequence of failed autophagy. In mice and humans with HD, mHtt abnormally interacts with the mitochondrial DRP1, stimulating its enzymatic activity and causing excessive mitochondrial fission [528]. Another study reported that p53 plays a role in this process by binding to DRP1 and promoting subsequent mitochondrial fission [529]. Moreover, excessive mitochondrial fission mediated by DRP1 increases the susceptibility of HD cells to apoptosis [530]. Protective mechanisms against excessive mitochondrial fission also exist in HD patients; for example, upregulated Sirt3 may mitigate it by downregulating the expression of DRP1 and FIS1 [531]. In addition, multiple studies suggest PPAR-δ and PCG-1α are repressed in HD, which may further suppress mitochondrial biogenesis [532,533,534]. Conversely, increasing the activity of PCG-1 and PPAR-δ is associated with an increment of mitochondrial mass and neuroprotection [535, 536]. Moreover, defective mitophagy within MQC leads to insufficient clearance of the damaged mitochondria as well as release of mtDNA from these damaged mitochondrial will lead to innate immune activation, thus aggravating neuroinflammation in HD patients [537].

Amyotrophic lateral sclerosis (ALS)

ALS, also known as motor neuron disease, causes progressive denervation of autonomic muscles due to degeneration of upper motor neurons in the motor cortex and lower motor neurons in the brain stem and spinal cord. Multiple genetic alterations are involved in the disease progression of ALS, such as SOD1. In a mouse model associated with SOD1 mutation-induced ALS, elevated levels of PCG-1α in muscle stimulate mitochondrial biogenesis, which helps counterbalance some of the observed mitochondrial dysfunction in the disease. However, despite maintained mitochondrial activity and muscle function at end-stage of ALS, survival is not extended [538]. Moreover, damaged mitochondrial fusion can further lead to defective mitochondrial in SOD1 mutation motor neurons [539, 540]. In neurons from early ALS patients, damaged mitochondria are cleared by Parkin-mediated mitophagy, however, unabated mitophagy along with degradation of mitochondrial proteins and decreased mitochondrial biogenesis can accelerate neurodegeneration by causing mislocalization and depletion of mitochondria [541]. This process is subject to complex regulation under ALS conditions involving multiple molecules, including mutational factors and aberrant activation of signaling pathways. A recent study indicated that mutant TBK1 in ALS also contributes to disrupting mitophagic flux, inhibiting the clearance of damaged mitochondria [542]. The overactivation of the ERK1/2 pathway and subsequent increased levels of translocator protein also contribute to impaired mitophagy [543]. Ubiquitinated transactive response DNA-binding protein-43 kD (TDP-43) is a main pathological hallmark of ALS, primarily localized in the cytoplasm of the brain and spinal cord, while it normally resides in the nucleus under physiological conditions. TDP-43 accumulation induces mitochondrial dysfunction through multiple processes, including disruption of MQC, and binding to Parkin pre-mRNA, inhibiting Parkin translation levels [544]. TDP-43 deficiency can also lead to the downregulation of Parkin [545], suggesting that these different mechanisms may coexist rather than exclude each other during mitophagy dysregulation at different stages of ALS progression. Axonal accumulation of TDP-43 can inhibit the local translation of nuclear-encoded mitochondrial genes, further impairing mitochondrial biogenesis in axon [546]. Abnormal mitochondrial dynamics contribute to disease progression in ALS, hyperactivated DRP1 leads to increased mitochondrial fragmentation and exacerbates the progression of ALS, while negative modulation or suppression of DRP1 expression could rescue cell viability [547].

Ischemic brain injury

In the context of I/R brain injury, upregulated PGC-1α in microglia promotes mitophagy, thereby alleviating ROS-mediated NLRP3 inflammasome activation, and in turn exerting neuroprotective effects [548]. Parkin-mediated mitophagy also alleviates excessive ROS production and may underlie the neuroprotective effects exerted by ER stress [549, 550]. In addition, BNIP3/NIX and FUNDC1-mediated mitophagy are involved in neuroprotection during I/R brain injury and can function independently of PARK2 [140, 551]. A previous study pointed out that excessive mitophagy activated by BNIP3 in ischemic brain injury without reperfusion could lead to neuron death [143]. Mitochondrial fission is an upstream process tightly linked with mitophagy, thus maintaining mitochondrial fission can promote mitophagy and protect against nerve damage caused by ischemic stroke [552]. Additionally, hyperglycemia may exacerbate mitochondrial fission under cerebral I/R conditions, further aggravating neural injury [553]. Activation of the N-Methyl-D-Aspartate (NMDA) receptor triggers Ca2+ influx, which is called excitotoxicity in stroke, acting as a signaling hub of neuronal pro-death events. A previous study revealed that excitotoxicity inhibits the expression of MFN2, leading to mitochondrial fragmentation and cell death [554]. This suggests that the regulation of mitochondrial dynamics from division to fusion plays a protective role in ischemic brain injury, which was confirmed by a previous study showing that inhibition of DRP1 relieved ischemic brain injury [555].

MQC and pulmonary disease

Mitochondria play a crucial role in the regulation of redox and immune responses, exerting influence on processes such as ROS production, activation of inflammasome, cellular proliferation, and prevention of fibrosis. These processes are implicated in the pathogenesis of various lung diseases. Here, we delve into the vital role of MQC in lung diseases (Fig. 6).

Fig. 6
figure 6

Mitochondrial quality control and pulmonary disease. a In idiopathic pulmonary fibrosis, mitochondrial quality control dysfunction manifests as reduced mitochondrial biogenesis and downregulation of mitophagy. b Similarly, in chronic obstructive pulmonary disease, diminished mitochondrial biogenesis and impaired mitochondrial dynamics are observed. c Conversely, in bronchial asthma, mitochondrial biogenesis is upregulated, which is closely associated with the high demand for smooth muscle cell proliferation. d Notably, excessive mitochondrial fission is a key feature of acute lung injury, where increased mitochondrial biogenesis is beneficial. e Likewise, excessive mitochondrial fission is also a notable feature of pulmonary arterial hypertension. Created by Biorender.com, accessed on 25 Aug. 2023

Idiopathic pulmonary fibrosis (IPF)

In myofibroblasts, TGF-β1 can induce transcriptional inhibition for PINK1, and deficient mitochondrial targeting for mitophagy, contributing to the progression of pulmonary fibrosis [556]. Impaired mitochondrial in alveolar type II cells are associated with the downregulation of PINK1 and defective mitophagy, making them susceptible to apoptosis and the development of lung fibrosis [557]. Another study suggests that there is a link between ATF3 and PINK1 in the alveolar epithelium; specifically, increasing ATF3 associated with aging inhibits the transcript level of PINK1 [558]. In alveolar macrophages, ROS-induced mitophagy and apoptosis resistance contribute to the long-term production of TGF-β, promoting the progression of pulmonary fibrosis [559]. However, although mitophagy is often considered to be a protective mechanism against damage mitochondria in pulmonary fibrosis, a recent study showed that PGAM5-mediated mitophagy in turn led to a self-perpetuating escalation of mitochondrial membrane potential depolarization [560]. Although the TGF-β induced Smad1/2 pathway is well recognized as regulating mitochondrial fragments, a previous study demonstrated inactive Smad2 could bind to MFN2 and further mediate mitochondrial fusion [561]. However, further studies are needed to understand how Smad2 influences mitochondrial function in IPF. A recent study proposed that the stiff matrix in fibrotic lung promotes mitochondrial fission in fibroblasts to meet the higher energy demand [562]. Protective mechanisms also exist against lung fibrosis in vivo, for example, thyroid hormone can promote the expression of PINK1 and PGC-1α, resulting in the restoration of mitochondrial function [563]. Chung et al. [564] revealed that MFN1 and MFN2 are associated with regulating the synthesis of lipids, while loss of MFN1 and MFN2 in ACE2 exacerbates bleomycin-induced lung fibrosis.

Chronic obstructive pulmonary disease (COPD)

Oxidative stress-mediated mitochondrial damage is an important pathological manifestation of COPD. Depolarized mitochondria activate mitophagy through the PINK1/Parkin2 pathway to clear the damaged fraction. However, the Parkin expression is downregulated in the lungs of COPD patients, causing mitophagy deficiency, which in turn promotes cell senescence and PCD [565]. Interestingly, overexpression of Parkin is sufficient to induce mitophagy when PINK1 protein levels are reduced. Conversely, PINK1 overexpression fails to rescue Parkin downregulation-mediated mitophagy and subsequent cell senescence [566], indicating that Parkin plays a pivotal role in regulating mitophagy during COPD. Moreover, mitophagy also contributes to necroptosis in COPD [567]. In addition, impaired mitochondrial dynamics characterized by defective fission and fusion processes are observed in the lung tissues of patients with emphysema, suggesting an association between impaired mitochondrial dynamics and emphysema progression [568]. In COPD patients exposed to cigarette smoke, alveolar epithelial cells or fibroblasts partially modulate MQC in response to external stimuli by regulating the coexistence ratio of different forms of OPA1, namely L-OPA1, and S-OPA1, and under the intervention of OPA1-interacting protein SLP-2, thereby mitigating damage [569].

Bronchial asthma

In young children with severe wheezing, increased mitochondria biogenesis is involved in the remodeling of bronchial smooth muscle [570]. The proliferation of asthmatic bronchial smooth muscle cells mainly requires mitochondrial-depend oxidative phosphorylation, thus increased mitochondrial biogenesis is essential in BSM cells [571]. Allergen-induced ROS production modulates mitophagy by regulating OPTN, which participates in the development of allergic airway inflammation [572]. Moreover, the MQC cross-talk between different cell types plays a role in asthma. For example, epithelial-derived thymic stromal lymphopoietin may induce excessive ROS production and subsequently regulate mitophagy and mitochondrial biogenesis, further influencing M1/M2 polarization chemokine expression in monocytes [573]. IL-4 mediated accumulation of asymmetric dimethylarginine causes oxo-nitrative stress, leading to mitochondrial dysfunction and HIF-1α activation. Activation of HIF-1αdecreases mitochondrial biogenesis, resulting in an alteration of mitochondrial turnover that can lead to bronchial epithelial cell death [574].

Acute lung injury (ALI)

Sepsis-induced ALI is associated with mitochondrial damage, whereas activated mitophagy, removes damaged mitochondria, promoting repair of lung tissues and cell survival [575]. In rat models, heme oxygenase-1 and carbon monoxide play a protective role in lung injury by regulating mitochondrial dynamic equilibrium [576]. Importantly, Staphylococcus aureus infection promotes PINK1 activity, which then phosphorylation of cardiolipin synthase on mitochondria (CLS1), leading to the excessive release of cardiolipin and further aggravating ALI [577]. Many studies have shown that the impaired mitochondrial dynamic is characterized by upregulated mitochondrial fission, and resultant oxidative stress is an essential mediator of ALI, as evidenced by the upregulation of OPA1 and MFN1. Conversely, increased mitochondrial fusion can alleviate ALI [578,579,580]. Under conditions of ALI, lung cells initiate Nrf2-mediated protective response, such as promoting mitochondrial biogenesis to reduce oxidative stress and inflammatory response for alleviating the progression of ALI [581]. Mitophagy seems to be an aggravating factor for ALI based on previous research findings. Therefore, PINK1 interaction with the ubiquitin apparatus to enhance mitochondrial quality could limit inflammatory injury [582]. Our team has been dedicated to exploring the pathogenesis of ALI and potential therapeutic strategies [579, 583, 584]. Our latest research findings indicate that mitochondrial dysfunction in type II alveolar epithelial cells contributes to epithelial barrier disruption during ALI [583]. Mechanistically, HDAC3 exerts its regulatory function by promoting the deacetylation and nuclear translocation of FOXO1, leading to upregulation of ROCK1 transcription, which exacerbates mitochondrial damage [583]. In conclusion, targeting MQC holds significant research potential for the prevention and treatment of ALI.

Pulmonary arterial hypertension (PAH)

The pathological process of PAH is characterized by the proliferation of smooth muscle cells and the metabolic transition of mitochondria from oxidative phosphorylation to aerobic glycolysis [585]. Excessive proliferation of pulmonary artery smooth muscle cells (PAMSCs) is associated with mitochondrial fragmentation in human and experimental PAH models. Decreased expression of MFN2 causes excessive mitochondrial content in PAMSCs, contributing to a proliferation-apoptosis imbalance of PAMSCs [586]. The upregulation and activation of DRP1 are also required for the proliferation of smooth muscle cells to accommodate mitosis [587]. Moreover, an in vitro study found that phosphorylation of DRP1 rather than upregulation alone promotes mitochondrial fission and subsequent PAMSC proliferation [588], suggesting the involvement of additional signaling cascades in vivo. Notably, mitochondrial biogenesis does not contribute to the elevated mitochondrial content in PAMSCs, instead, downregulation of PCG-1α in PAH leads to reduced mitochondrial biogenesis as well as further downregulation of MFN2 expression [586]. There is substantial evidence indicating that the post-translational activation of DRP1 or the downregulation of MFNs is the primary driving factor contributing to aberrant cell division in the pathogenesis of PAH [589].

MQC and kidney disease

Kidneys play essential roles in maintaining homeostasis. by actively reabsorbing large amounts of solutes in the renal tubules and collecting ducts, resulting in high energy demands. The proximal tubules are responsible for reabsorbing more than 99% of glucose, while this high energy requirement cannot be met by anaerobic glycolysis. Instead, mitochondrial oxidative phosphorylation or the production of large amounts of ATP from fatty acid oxidation is required [590]. Therefore, MQC directly determines the balance of water and electrolytes in the body.

Acute kidney injury (AKI)

The etiology of AKI encompasses a range of factors, such as sepsis, I/R, and exposure to toxins. The pathological process includes endothelial activation, increased microvascular permeability, and altered regional blood flow distribution resulting in areas of inadequate perfusion and local hypoxemia. Ultimately, this cascade leads to the death of certain renal tubular cells.

Numerous experimental studies have shown that mitochondrial biogenesis plays a beneficial role in AKI and its subsequent renal repair. PGC-1α, the major regulator of mitochondrial biogenesis, has been confirmed to be localized in the proximal tubule with a high energy requirement [591]. In AKI induced by various factors including LPS, cisplatin, and I/R, the expression level of PGC-1α is significantly suppressed [591, 592]. The transcription factor PPARα, downstream of PGC-1α, exhibits reduced binding activity to the retinoid X receptor (RXRα) in AKI, which hampers mitochondrial biogenesis, suggesting impaired mitochondrial function in a mouse model of cisplatin-induced AKI [591]. Renal function remains unaffected under normal conditions using PGC-1α knockout mice, but knockout mice exhibit prolonged renal recovery time after injury under septic inflammatory stress conditions [593]. Overexpression of PGC-1α in renal tubular cells promoted renal recovery from LPS-induced AKI, indicating a positive correlation between PGC-1α expression and the extent of renal repair after injury [593]. Moreover, pharmacological activation of PGC-1α accelerated renal recovery from I/R injury in mice [594].

Mitochondrial fragmentation, caused by excessive fission and/or inhibition of fusion, is thought to be a key event in mitochondrial damage and renal tubular injury during AKI [595, 596]. Silencing or pharmacological inhibition of the DRP1 expression attenuated apoptosis and mitochondrial fragmentation, suggesting that DRP1 mediates excessive mitochondrial fission and ultimately leads to mitochondrial fragmentation during AKI. It is an important player in tubular cell apoptosis [595]. Similarly, specific knockdown of DRP1 in mouse kidney proximal tubules ameliorates I/R-induced inflammation and programmed cell death in AKI, while promoting the recovery process of renal epithelial cells after injury [597]. Consistently, Bif-1 regulates the mitochondrial inner membrane by its interaction with prohibitin-2 to disrupt prohibiting complexes. This induces OPA1 proteolysis and inactivation, leading to enhanced apoptosis, and mitochondrial fragmentation in ischemic AKI [598]. These studies suggest that OPA1-mediated mitochondrial fusion plays a negative regulatory role in ischemic AKI. While MFN2 is responsible for holding mitochondria firmly together during mitochondrial fusion, and specific deficiency of MFN2 in mouse kidneys aggravates ATP depletion-mediated renal tubular apoptosis, mitochondrial outer membrane damage, and mitochondrial fragmentation [599]. However, in ischemia-related AKI, mice with MFN2-specific knockout in the proximal tubule (PT) had higher survival rates and a 5-fold increase in kidney cell proliferation [600]. In contrast to its pro-apoptotic effect observed in vitro, PT-MFN2-CKO enhances the Ras-ERK pathway to promote kidney cell proliferation and thus accelerate the repair process after kidney injury [600], this suggests that the detrimental effects of MFN2 knockdown-mediated inhibition of mitochondrial fusion are counteracted or outweighed by the ERK signaling pathway, which manifests as increased proliferation of kidney cells.

There is growing evidence supporting the important role of mitophagy in AKI and kidney repair. The flux of mitophagy increases over time following kidney injury induced by I/R, contrast agents, and cisplatin-induced AKI mouse models [241, 601, 602]. In mice with cisplatin-associated AKI, PGC-1α was found to coordinate with mitophagy in the kidney, enhancing mitophagy and thus improving AKI [603]. Increased expression of BNIP3, a mediator of mitophagy, was observed in I/R-related AKI. Reduced mitophagy through BNIP3 knockdown in vivo and in vitro resulted in the promoted accumulation of damaged mitochondrial, increased oxidative stress, and increased renal tubular cell death [601]. In contrast-associated AKI, the kidneys of PINK1 or Parkin knockout mice suffered more severe renal injury than wild-type mice, and mitophagy mediated by PINK1 or Parkin was largely eliminated, suggesting that the PINK1/Parkin pathway plays a dominant role in mitophagy [241]. Also, it was discovered that mitophagy reduces the activation of mtROS and NLRP3 inflammatory vesicles to ameliorate kidney injury in mice [241]. These discoveries underscore the beneficial role of mitophagy in AKI by reducing damaged mitochondria accumulation and mtDNA release while minimizing the disturbances to renal homeostasis under inflammatory stress through multiple pathways.

CKD

CKD is the result of severe kidney damage or persistent irritation, characterized by impaired regeneration of renal tubular cells and tubulointerstitial fibrosis. Decreased PGC-1α expression has been observed in the kidneys of patients with diabetic kidney disease (DKD), accompanied by a decrease in mitochondrial proteins and exosomal mtDNA [604]. Similarly, decreased PGC-1α expression has been found in human patients with DKD as well as mouse models featuring renal podocytes [605]. The long non-coding RNA taurine-upregulated gene 1 (Tug1) was shown to interact with PGC-1α, facilitating its binding to the promoter region and enhancing transcriptional activity through direct interaction with the gene encoding PGC-1α [606]. Overexpression of Tug1 transgene specifically in podocytes improved pathological damage and mitochondrial stability in a DKD mouse model, suggesting a beneficial role for PGC-1α-mediated mitochondrial biogenesis in the progression of DKD [606]. As mentioned earlier, AMPK acts as a positive regulator of PGC-1α, and pharmacological activation of AMPK followed by restoration of PGC-1α expression has shown promising results in alleviating DKD [607]. However, the overexpression of PGC-1α leads to abnormal proliferation of podocytes, leading to albuminuria and glomerulosclerosis [605]. CKD differs from acute injury in that cytokine levels are fully activated and persist for a long time. Mitochondrial biogenesis plays a crucial role in maintaining MQC by generating new mitochondria that replace damaged ones. Conversely, the inhibition of mitochondrial formation directs cellular fate towards death.

Increased mitochondrial fragmentation has been demonstrated in renal tubular cells and podocytes of both experimental models and patients with DKD [608, 609]. Phosphorylation of the DRP1 protein serine 600 (S600) promotes mitochondrial fission, while specific mutation of S600 to alanine reduces mitochondrial fission and ameliorates kidney damage in DKD mice [610]. In addition, the podocyte-specific knockdown of DRP1 in the DKD mouse model improved its mitochondrial structure function and adaptability [611], and pharmacological inhibition of DRP1 in podocytes reduced mitochondrial fission to stabilize mitochondria thereby improving DKD [611, 612]. Enhanced mitochondrial fission was detected in renal tubules of unilateral ureteral obstruction (UUO)-induced CKD mice [613], as well as fibroblasts in the kidneys of CKD patients and UUO mice [614]. C/EBP homologous protein (CHOP) may be a positive regulator of mitochondrial fission. In the UUO-induced CKD model using CHOP knockout mice, elevated OPA1 expression was observed along with reduced mitochondrial fragmentation compared with wild-type littermates, indicating reduced renal fibrosis [615]. Increased phosphorylation of DRP1S616 in the kidney of UUO mice stimulates mitochondrial fission and promotes fibroblast activation and proliferation, suggesting a positive role in renal fibrosis [614]. In summary, excessive mitochondrial fragmentation accompanying CKD is mainly caused by an imbalance within the dynamics regulating mitochondria.

Mitochondrial fragmentation, impaired mitophagy, and increased apoptosis were found in the renal tubules of the DKD mouse model as well as in high glucose (HG)-induced HK-2 cells [609, 616]. Similarly, a significant reduction in the expression level of OPTN, an important regulator of phagosome formation, was detected in the kidney and HG-induced tubular epithelial cells of patients with DKD. This decrease showed a negative correlation with the severity of CKD [617]. Cellular aging is an important factor in the progression of DKD, and overexpression of OPTN promotes mitophagy in renal tubular epithelial cells in a DKD model, thereby alleviating cellular aging. Additionally, the pharmacological activation of mitophagy serves a similar function [617]. Furthermore, the damage was exacerbated by the specific absence of the PINK1/Parkin pathway in a mouse model of UUO renal fibrosis [618].

MQC and digestive system disease

I/R-induced liver injury

The occurrence of liver I/R injury is observed in various clinical scenarios, including liver transplantation, trauma, hemorrhagic shock, and liver resections [619]. Despite advancements in techniques for safer liver surgery leading to improved surgical outcomes, liver I/R injury is still the major cause of postoperative liver dysfunction and failure, especially in the context of liver transplantation [620].

Many pieces of evidence confirm the beneficial role of mitochondrial biogenesis in I/R-induced liver injury. The expression of PGC-1α decreased in a mouse model of IR-related liver injury and hypoxia-induced human hepatocytes [621, 622]. Irisin is a positive regulator of mitochondrial biogenesis and can restore PGC-1α expression levels and TFAM protein levels in a mouse model of I/R-induced liver injury, thereby ameliorating liver injury. This effect was also observed in an in vitro hypoxic hepatocyte model [621]. However, Sirt1 inhibition prevented the positive regulation of mitochondrial biogenesis by Nobiletin, suggesting that Sirt1 likely acts as an upstream effector for the increase in PGC-1α and TAFM expression [622]. Cilostazol, a phosphodiesterase inhibitor, has been demonstrated to increase HO-1 expression and enhance mitochondrial biogenesis in a mouse model of I/R-related liver injury and extracted primary hepatocytes, thereby improving mitochondrial function, and reducing liver injury [623]. In contrast, cilostazol’s enhancement of I/R-induced mitochondrial biogenesis in both in vitro and in vivo models was blocked after inhibiting HO-1 or Nrf2 expression [623]. This suggests that mitochondrial biogenesis induced by cilostazol is dependent on HO-1 and Nrf2. The liver is a highly regenerative organ that requires stable mitochondria to maintain its function during early stages of injury as well as meeting high energy demands for cellular regeneration post-injury. Therefore, promoting new mitochondria production is beneficial for liver recovery. Excessive inhibition of mitochondrial fission and/or fusion has detrimental effects on IR-associated liver injury. In I/R-related liver injury, there were significantly higher expression levels observed for DRP1 and FIS1 associated with mitochondrial fission compared with sham group [621], while the expression of MFN1 and MFN2, which are associated with mitochondrial fission, was decreased [624]. Irisin reduces mitochondrial fragmentation by decreasing the expression levels of DRP1 and FIS1 to inhibit mitochondrial fission and thereby ameliorate I/R-related liver injury [621]. In addition, both SUMOylation of DRP1 and phosphorylation at Ser616 were significantly increased in the livers of liver transplant recipients undergoing I/R. These post-translational modifications were designed to recruit DRP1 to mitochondria for mitochondrial fission following I/R injury [625]. Augmenter of liver regeneration (ALR, genetic name, Gfer), formerly called hepatic stimulator substance, was originally identified in the liver of weanling rats in 1975 [626]. ALR knockout mice with liver injury induced by ischemia–reperfusion injury showed a significant increase in SUMOylation of DRP1, which was similarly manifested in I/R-induced hepatocytes, while ALR overexpression prevented the increase in DRP1 SUMOylation [625]. Moreover, inhibition of ALR expression in hepatocytes prevented the translocation of DRP1 to mitochondria and reduced mitochondrial fragmentation [626]. This suggests that ALR is a negative regulator of mitochondrial fission by inhibiting SUMOylation of DRP1, thereby reducing its translocation to mitochondria improving impaired mitochondrial dynamics, and ultimately preventing IR-related liver injury. In addition, the mitochondrial fusion-associated OPA1 protein is hydrolyzed by OMA1 in a mouse model of I/R-associated liver injury, leading to the inhibition of mitochondrial fusion and loss of mitochondrial membrane potential [627]. In contrast, liver injury caused by I/R induced OPA1 hydrolysis and mitochondrial instability was attenuated in a mouse model of hepatocyte-specific overexpression of human cyclooxygenase-2 (COX-2) [627], indicating that human COX-2 may be a positive regulator of mitochondrial fusion in hepatocytes. Mitochondrial dynamics play an important role in IR-induced liver injury, and excessive activation of mitochondrial fission and impaired mitochondrial fusion negatively impact the maintenance of homeostasis and recovery from liver injury.

Mitophagy was inhibited in a mouse model of I/R-related liver injury [628]. During hepatic I/R injury, mitophagy facilitated the degradation and recycling of damaged mitochondria, thereby reducing the release of mtDNA while meeting the survival needs of hepatocytes after liver injury. In I/R-induced liver injury, ALR not only enhanced the mitochondrial translocation of PINK1 and Parkin but also induced MFN2 expression [629]. In this process, depletion of MFN2 blocked ALR-induced mitophagy activation, promoted mitochondrial dysfunction, and ultimately increased apoptosis. In addition, mesenchymal stem cells can reduce mtROS overproduction, reduce mitochondrial debris accumulation, restore ATP production, and upregulate the PINK1/Parkin-mediated mitophagy pathway during liver injury. These findings may provide a theoretical basis for exploring novel strategies for the treatment of IR-related liver injury [628] (Fig. 7).

Fig. 7
figure 7

Mitochondrial quality control and digestive system diseases. Various digestive system diseases demonstrate a downregulation of mitochondrial biogenesis, excessive mitochondrial fission, and downregulation of mitophagy within the mitochondrial quality control system. a In hepatic ischemia–reperfusion injury, downregulation of sirt1 hinders the downstream mitochondrial biogenesis process, with alterations in Cox-2 and ALR expression levels also involved in the regulation of mitochondrial dynamics disruption. Additionally, ALR is implicated in the regulation of the PINK1 and Parkin-mediated mitochondrial translocation process, leading to impaired mitophagy. b In non-alcoholic fatty liver disease, upregulation of P2Y2R affects PGC-1α expression levels, consequently causing inadequate mitochondrial biogenesis. NR4A1 is involved in regulating excessive mitochondrial fission, while mitophagy defect is a notable feature of this disease. c In inflammatory bowel disease, insufficient mitochondrial biogenesis is controlled by SMYD5, and notably, excessive mitochondrial fission contributes to the activation of NLRP3 in intestinal inflammation. ALR augmenter of liver regeneration, BNIP3 BCL2 interacting protein 3, Cox-2 cyclooxygenase 2, DRP1 dynamin-related protein 1, FIS1 fission protein 1, MFN mitofusin, HO-1 heme oxygenase 1, I/R ischemia–reperfusion, NLRP3 NLR family pyrin domain containing 3, NR4A1 nuclear receptor subfamily 4 group A member 1, Nrf2 nuclear factor E2-related factor 2, OPA1 optic atrophy 1, PGC-1α PPAR-γ coactivator-1α, P2Y2R P2Y receptor 2, SMYD5 SET and MYND domain 5, Sirt1 sirtuin 1, SUMO small ubiquitin-like modifier

Non-alcoholic fatty liver disease (NAFLD)

The prevalence of major factors such as obesity and other components of metabolic syndrome is increasing, leading to a rise in the incidence of NAFLD [630, 631]. In both rodent models and patients with NAFLD, there is a decrease in markers associated with mitochondrial renewal [632], and the hepatic expression of PGC-1α showed a negative correlation with liver fat accumulation and disease severity [633]. Pharmacological activation of downregulated PGC-1α in NAFLD using various products can enhance neo-mitochondrial production, alleviate mitochondrial dysfunction, and improve NAFLD [634]. In addition, estrogen positively regulates the protective effect of PGC-1α in the mouse model of NAFLD [633]. P2Y2R may be a negative regulator of PGC-1α associated with mitochondrial biogenesis. Upregulation of PGC-1α-mediated mitochondrial biogenesis observed in the P2Y2R knockout mouse model ameliorates damage and steatosis associated with NAFLD [635]. Therefore, it is essential to induce new mitochondria through PGC-1α-mediated pathways to attenuate NAFLD.

In the NASH model mice fed a Western diet (high in fat, fructose, and cholesterol) for more than 1 month, reduced FIS1 and DRP1 protein levels were observed along with liver inflammation and fibrosis [636]. Reduced mitochondrial fission alleviates hepatic steatosis in a mouse model of NAFLD [637], suggesting the detrimental role played by mitochondrial fission in NAFLD progression and the potential therapeutic target it presents. NR4A1 has been shown to promote mitochondrial fission while inhibiting mitophagy in a mouse model of NAFLD. However, upon melatonin supplementation, the expression of NR4A1 was suppressed, along with reduced downstream mitochondrial fission and restored mitophagy, leading to an improvement of NAFLD [638]. This indicates that pharmacological inhibition of mitochondrial fission may partially contribute to the amelioration of NAFLD along with enhanced mitophagy. In addition, reduced levels of MFN2 expression were detected in steatosis or NASH mouse models [639]. Hepatic-specific knockdown of MFN2 in mice causes inflammation, triglyceride accumulation, fibrosis, and hepatocellular carcinoma; however, re-expression in a mouse model of NASH ameliorates these conditions [639], highlighting the beneficial role played by MFN2 in NAFLD. Similarly, CXCR3-mediated downregulation of the MFN1 protein leads to mitochondrial dysfunction in a mouse model of NASH [640]. These studies confirm that mitochondrial fission is a detrimental factor in the progression of NAFLD, while mitochondrial fusion serves as a beneficial factor by facilitating the exchange of metabolic substrates and substances between mitochondria.

There is mounting evidence indicating the beneficial role of mitophagy in NAFLD. The expression of BNIP3 is significantly upregulated in the livers of fasted mice fed an HFD [641, 642]. Silencing BNIP3 leads to increased lipid biosynthesis in the liver, accompanied by elevated ATP levels, reduced AMPK activity, and enhanced expression of lipogenic enzymes [643]. Liver mitochondria from BNIP3 null mice exhibited reduced mitochondrial membrane potential, structural abnormalities, and reduced oxygen consumption, and are associated with increased ROS, inflammation, and steatohepatitis-like features [643]. These findings suggest that BNIP3 pathway-mediated mitophagy improves mitochondrial function in a mouse model of NAFLD, and is also linked to lipid synthesis and liver inflammation. The protective effects of the PINK1/Parkin mitophagic pathway were initially observed in studies on alcoholic fatty liver. In a model of alcohol-induced hepatic steatosis, Parkin knockout mice showed impaired mitophagy, along with severe swelling and destruction of mitochondria compared with wild-type mice. Additionally, these knockout mice exhibited reduced mitochondrial cristae formation as well as diminished ability for the liver to adapt to alcohol intake [644]. Consistent with this finding, the expression of PINK1 and Parkin was significantly downregulated in an HFD-induced NAFLD mouse model, which was associated with activation of the mitochondria-associated apoptotic pathway and mPTP opening [645]. In addition, pharmacological activation of the PINK1/Parkin mitophagy pathway has been shown to enhance mitochondrial function and alleviate NAFLD in a mouse model of NAFLD [646]. These findings suggest that the PINK1/Parkin and BNIP3 pathways of mitophagy play a beneficial role in the progression of NAFLD by effectively eliminating damaged mitochondria while promoting the generation of new mitochondria through PGC-1α-mediated mechanisms, thereby facilitating hepatocyte mitochondrial renewal and ensuring mitochondrial stability. NAFLD is caused by the persistent accumulation of lipids in hepatocytes and the stabilization of mitochondria can improve the disruption of lipid metabolism and oxidative stress thereby improving NAFLD (Fig. 7).

Inflammatory bowel disease (IBD)

The two most prevalent forms of IBD are Crohn’s Disease (CD) and ulcerative colitis (UC). Mitochondrial function in the gastrointestinal epithelium plays a crucial role in maintaining intestinal health, and emerging research suggests a significant association between mitochondrial dysfunction and IBD [647].

PGC-1α exhibited high expression levels in normal intestinal epithelial cells (IECs) [648], whereas its expression was reduced in IECs from patients with IBD [649]. Similarly, repression of mitochondria-related gene expression was found in UC patients, including mitochondrial biogenesis-related PGC-1α [650]. Post-translational modifications were found to regulate PGC-1α in IBD. In colon tissue of IBD patients and mouse models, as well as in human IECs, the expression of SET and MYND domain-containing protein 5 (SMYD5) (a methyltransferase) was upregulated, while PGC-1α expression was downregulated. Knockdown of SMYD5 blocked this process, resulting in upregulation of PGC-1α and improved mitochondrial function and IBD symptoms [651]. Pharmacological activation of PGC-1α also showed a protective effect against IBD by reducing the expression of inflammatory cytokines such as interleukin (IL)-1β, IL-18, and TNF-α in IBD mice, as well as macrophage infiltration in colon tissue. This activation further alleviated mitochondrial dysfunction and oxidative stress, while protecting the histological structure of the colon in an IBD mouse model [652]. These findings suggest an important protective role of mitochondrial biogenesis in the progression of IBD.

The expression of mitochondrial fission-associated DRP1 is upregulated in the intestinal epithelium of IBD [653]. Further, an increased level of mitochondrial fission was detected, which was attenuated by Atractylenolide I [654]. A DRP1 inhibitor suppressed NLRP3, ASC, and caspase-1 upregulated by LPS/DSS-stimulated mice bone marrow-derived macrophages (BMDMs), and the addition of Atractylenolide I showed similar results, while this anti-inflammatory effect was diminished upon overexpression of DRP1 [654]. In summary, Atractylenolide I ameliorates IBD by inhibiting DRP1-mediated activation of mitochondrial fission-induced NLRP3 inflammatory vesicles and suggests a detrimental effect associated with excessive activation of DRP1 in IBD.

There is increasing evidence for the beneficial role of mitophagy in IBD. In UC patients and mouse models, BNIP3 is upregulated and targeted to mitochondria in the intestinal epithelium. Conversely, UC symptoms are more severe in BNIP3 knockout mouse models, accompanied by mitochondrial accumulation [655]. This suggests a protective role of BNIP3-mediated mitophagy in the UC progression. The CD-associated ATG16L1T300A variant leads to impaired mitophagy, mtROS accumulation, and increased IL-1β production as well as pro-inflammatory macrophage polarization with reduced bacterial killing [656, 657]. Andrographolide induces mitophagy in IBD, thereby reducing mitochondrial dysfunction and reversing mitochondrial membrane potential. It also deactivates NLRP3 inflammatory vesicles in macrophages both in vivo and in vitro to ameliorate IBD [658]. In addition, Prohibitin 1 negatively regulates mitophagy in the human colonic epithelium [659]. It suggests that mitophagy is upregulated in IBD patients and plays a beneficial role by clearing damaged mitochondria. Given that mitochondria are often damaged in IBD patients and experimental models, activating mitophagy to clear damaged mitochondria serves as a defense mechanism against the release of mtDNA and sustained production of mtROS for protecting against IBD (Fig. 7).

Therapies targeting MQC

Due to the critical roles played by mitochondrial function in many biological processes and diseases, targeting the MQC system emerges as an appealing therapeutic strategy. In light of the aforementioned findings, a quantity of MQC-targeted drugs and natural compounds have been widely investigated. This section aims to summarize the actionable mechanisms of these drugs and compounds reported thus far while also envisioning potential future therapeutic directions (Table 1).

Table 1 Representative therapies targeting mitochondrial quality control

Target mitochondrial biogenesis

PGC-1α is widely acknowledged as a key regulator of mitochondrial biogenesis. Upstream, the mTOR, AMPK and Sirtuins pathways play important roles in regulating PGC-1α activity. Downstream, PGC-1α stimulates the expression of Nrf1 and Nrf2, subsequently driving the transcription of the TFAM gene [660, 661]. Most of the drugs targeting mitochondrial biogenesis regulate these signaling pathways or their components. Nitric oxide, an endothelial metabolite with significant effects on mitochondrial biology [662], activates PGC-1α through an AMPK-or p53-dependent mechanism to initiate mitochondrial biogenesis [663, 664]. DETA-NO and S-nitroso-N-acetyl-penicillamine could act as NO providers in vitro and in vivo respectively, both of which led to a proliferation of functional mitochondria [665, 666]. Besides, NAD+ participated in the process of mitochondrial biogenesis by involving the NAD+/Sirt1 axis [667, 668]. In vitro supplementation with the NAD+ precursor Nicotinamide riboside activated Sirt1, Sirt3, and mitochondrial biogenesis and improved exercise endurance and insulin sensitivity [669, 670]. Furthermore, NAD+ has also been demonstrated to alleviate PD and AD by promoting PGC-1α function, reducing neuroinflammation, apoptosis, and DNA damage [667, 670]. Regarding natural products, there are also several compounds found to regulate mitochondrial biogenesis. Resveratrol, the most extensively studied polyphenolic flavonoid, has been shown to improve mitochondrial function and survival rate in models of cardiovascular and neurodegenerative diseases, stroke, epilepsy, aging, depression, and a variety of cancers, by promoting the expression of PGC-1α, Nrf1 and TFAM [671]. Berberine, a purified extract of the traditional Chinese medicine Coptidis rhizome, has exhibited a great impact on metabolic syndrome, cardiovascular disease, and neurodegenerative diseases through the AMPK/Sirt1/PGC-1α pathway [672, 673]. Quercetin, whose chemical structure is similar to that of resveratrol, not only increases mitochondrial copy number but also enhances mitochondrial membrane potential, mitochondrial oxygen consumption, and ATP levels through the AMPK/Sirt1 signaling. It played a protective role in osteoarthritis rats by reducing oxidative stress [674]. Curcumin which has potentially important biological effects on antioxidant and anti-inflammatory properties, has also been shown to potentially trigger mitochondrial biogenesis in vitro and in vivo experimental models, but the exact mechanism remains to be fully understood [675].

As for synthetic drugs, multiple studies have yielded significant findings. Lipoamide, the neutral amide of lipoic acid, functions as an effective antioxidant and mitochondrial nutrient that can stimulate eNOS expression and cGMP production in a dose-dependent manner, and the eNOS/cGMP/PKG signaling pathway was involved in the stimulation of mitochondrial biogenesis [676]. The AMP analog 5-aminoimidazole-4-carboxamide-1-beta-D-ribotide (AICAR) directly activates AMPK, leading to phosphorylation of PGC-1α and activation of Sirt1 through an AMPK-induced increase in NAD+/NADH ratio [677, 678]. However, AICAR’s limited ability to penetrate the restricts its application in the central nervous system [677]. Metformin plays an important role in the regulation of energy metabolism through AMPK activation, which can improve the levels of mitophagy and PGC-1α [493, 679]. While metformin has been reported to improve bleomycin-induced pulmonary fibrosis [680], further clarification is needed regarding its cardioprotective effects due to varying conclusions at different metformin concentrations [681]. Formoterol, a selective β(2)-AR agonist, significantly increased the copy number of mtDNA and upregulated the expression of PGC-1α and several other genes involved in the mitochondrial electron transport chain in an I/R-induced AKI mouse model [594], thereby accelerating the recovery process of podocyte from glomerular injury and improving brain function following traumatic brain injury [682, 683]. Fibrates, small-molecule agonists of the PPAR pathway, have the potential to prevent diabetic peripheral neuropathy by activating the PPARα/AMPK/PGC-1α/eNOS pathway to alleviate nerve and endothelial injury [684]. However, the therapeutic potential of the PPAR hybrid agonist is severely hampered due to hepatotoxicity, with the except for clofibrate [685]. Thiazolidinediones exert a potent effect on improving insulin sensitivity through the PPARγ pathway [686]. Pioglitazone can protect mitochondrial function, defend against oxidative stress promote mitochondrial biogenesis, enhance oligodendrocyte differentiation, and ameliorate the pathological state of demyelinating diseases [687]. Leriglitazone enhances mitochondrial function and increases mitochondrial biogenesis by targeting the PPARγ pathway, offering a potentially effective treatment for Friedreich ataxia [688]. However, its application is limited due to serious safety issues such as life-threatening cardiotoxicity and hepatotoxicity [689]. Rimonabant, a selective pharmacological blocker of cannabinoid type 1 receptor, has also been found to increase mitochondrial biogenesis in WAT by inducing eNOS expression, thereby protecting against HFD-induced adipose accumulation [690].

Target mitochondrial dynamics

Maintaining a balance between mitochondrial fission and fusion is of crucial importance to cellular energy metabolism and homeostasis, and imbalanced mitochondrial dynamics have been implicated in many human diseases [6, 56]. Recent studies have demonstrated the therapeutic potential of pharmacological interventions that restore balanced mitochondrial dynamics. Inhibition of mitochondrial fission seems to reverse the disease process. Mdivi-1, as the selective chemical inhibitor of DRP1, effectively attenuates mitochondrial fission [691]. The protective effect of mdivi-1 in cardiovascular diseases has been reported. Treatment with mdivi-1 can reduce myocardial injury in a mouse model of MI, attenuate doxorubicin-induced cardiotoxicity, and ameliorate hypertension by inhibiting mitochondrial fission [398, 479, 692]. Besides, mdivi-1 treatment has also been shown beneficial effects on many diseases, including AD, subarachnoid hemorrhage-induced brain injury, autoimmune hepatitis, AKI, CKD, sepsis, and metabolic disease, partly due to the protection of mitochondrial dynamics [489, 595, 611, 693,694,695,696]. Furthermore, in cancers such as lung cancer and breast cancer, the expression of DRP1/MFN2 indicates a disruption in mitochondrial function. Treatment with mdivi-1 resulted in reduced proliferation and migration [302, 697]. Lei et al. [698] also demonstrated that mdivi-1 upregulates major histocompatibility complex class I expression in cancer cells by affecting mitochondrial dynamics, thereby diminishing the capability of immune escape and enhancing the efficacy of adoptive T cell therapy. P110 has been reported as another selective inhibitor of excessive mitochondrial fission that blocks the interaction of DRP1 with FIS1 [699]. Inhibiting mitochondrial fragmentation and p53-dependent apoptosis in dopaminergic neuronal cells targeting DRP1, P110 alleviates the PD model both in vivo and in vitro [699, 700]. Moreover, beneficial effects of P110 therapy have been observed in other mouse models such as pulmonary hypertension, IBD, and septic organ injury [701,702,703,704,705].

Pharmacological treatments targeting mitochondrial fusion may be a promising approach for stabilizing mitochondrial dynamics. SAMβA, a novel small peptide that selectively antagonizes the interaction between MFN1 and βIIPKC, demonstrates the ability to reverse excessive mitochondrial fragmentation and ameliorate heart failure in rats [706]. BGP-15 promotes GTPase activity and self-aggregation of OPA1, thereby relieving heart failure and PAH in rats through the maintenance of mitochondrial fusion [707, 708]. Mitochondrial fusion promoter-M1 is considered an inducer of mitochondrial fusion proteins, with its administration separately alleviating prediabetic cardiac I/R injury and diabetic cardiomyopathy in rats [709, 710]. The pharmacological mechanisms of melatonin include modulation of mitophagy and mitochondrial biogenesis. Moreover, melatonin suppresses DRP1‐mediated mitochondrial fission in diabetic cardiomyopathy through Sirt6 as well as the Sirt1/PGC-1α pathway respectively [711, 712]. The function of mitochondrial fusion is destroyed during cerebral I/R injury, but melatonin reverses these conditions via the Yap-OPA1 signaling pathway [713]. In addition, the effects on maintaining mitochondrial dynamics of melatonin have potential benefits in respiratory diseases, including ALI and SARS-CoV-2 infection [579, 714]. Studies have illuminated that mitochondria-targeted antioxidant MitoQ protects against DKD and intervertebral disc degeneration through the regulation of Nrf2-related mitophagy and mitochondrial dynamics balance [616, 715]. In a rat model of myocardial I/R injury, there is a deterioration in the degree of the mitochondria network along with impaired mitophagy; however, this condition can be alleviated by resveratrol treatment [716].

Target mitophagy

Mitophagy, a type of selective autophagy, plays an essential role in maintaining the homeostasis of mitochondria and cells [717]. There has been an increasing interest in the development of compounds that target mitophagy, particularly focusing on the Parkin/PINK1 pathway and the FUNDC1/BNIP3/NIX pathway. Urolithin A, a natural compound derived from the gut microbiome, induces PINK1-dependent mitophagy and exhibits potential for reversing cognitive deficits and alleviating AD in mice [718]. Similarly, Urolithin A activates mitophagy to improve muscle function and exercise capacity in aged mice. Recently, its clinical efficacy has been demonstrated through a randomized clinical trial containing 66 participants [719, 720]. In addition, treatments of Urolithin A have shown beneficial effects in mouse models of obesity-induced metabolic cardiomyopathy and septic myocardial injury [721, 722]. Similar effects have also been observed with spermidine, another natural inducer of autophagy, in rodent models of AD and PD [723]. Furthermore, aging increases the levels of IL-6 and induces mitochondrial dysfunction, thereby enhancing atherosclerosis. This is accompanied by a protective increase in mitophagy. Spermidine plays a protective role in this condition by upregulating the Parkin level [724]. MitoQ, a mitochondria-targeting antioxidant composed of coenzyme Q10 and triphenyl phosphate cations, exhibits dual effects on mitophagy in different disease models. In a mouse model of diabetic mouse kidney tubular injury, hyperglycemic states downregulate the levels of Parkin and PINK1 and lead to increased oxidative stress, apoptosis, and mitochondrial dysfunction. MitoQ induces the transcription of PINK and restores mitophagy in a Nrf2-dependent manner [616]. Correspondingly, intervertebral disc degeneration involves activation of PINK1/Parkin‐mediated mitophagy along with blockade of mitophagic flux. Treatment with MitoQ further promotes PINK1/Parkin‐mediated mitophagy and restores mitophagy flux [715]. Interestingly, exposure to PM2.5 induces excessive vascular fibrosis through enhanced mitophagy; however, this process can be inhibited by MitoQ via ROS/PINK1/Parkin pathway [238]. Whether these discrepancies could be explained by cell specificity requires further investigation. Metformin, a classical hypoglycemic drug, promotes mitophagy of peripheral blood mononuclear cells through AMPK phosphorylation to relieve chronic inflammation and improve β-cell function in vitro and in vivo. This effect has also been confirmed by a randomized controlled trial [493, 725, 726]. Moreover, studies have revealed that metformin can attenuate subarachnoid hemorrhage-induced early brain injury and diabetes-induced renal tubulointerstitial fibrosis through AMPK-dependent mitophagy [727, 728]. Empagliflozin, another novel anti-diabetes drug, activates mitophagy in endothelial cells via the AMPKα1/ULK1/FUNDC1 pathway to alleviate cardiac microvascular I/R injury [173]. Melatonin has been recognized as a regulator of mitophagy and has been shown to ameliorate myocardial I/R injury in mice by improving protective mitophagy and mitochondrial dynamics via the AMPK/OPA1 signaling pathways [411]. Melatonin is beneficial for diabetic cardiomyopathy in mouse models through increased expression of LC3 II, colocalization of mitochondria and lysosomes, as well as translocation of Parkin [729]. Furthermore, the therapeutic effects of melatonin have also been verified in various diseases, including AD, experimental liver fibrosis, traumatic brain injury, and LPS-induced ALI [579, 730,731,732].

Some natural compounds can also contribute to the maintenance of mitophagy. Curcumin, a natural polyphenol extracted from Curcuma longa Linn, can enhance intestine barrier function and mitigate oxidative stress-induced intestinal injury by increasing Parkin/PINK1 expression [733]. Quercetin inhibits oxidative stress, and ER stress and restores PINK1/Parkin-related mitophagy through Sirt1/TMBIM6, and eventually ameliorates hypoxia-caused cardiomyocyte injury [734]. It has been reported that quercetin can improve the neurochemical levels and PD disability score in PD rats, partially owing to the regulation of PINK1/Parkin mitophagy [735]. Furthermore, regulating the interplay between NLRP3 inflammasome and mitophagy in microglia appears to be a potential novel therapeutic target for quercetin in neurological and neurodegenerative diseases [736]. Puerarin limits the inflammation vulnerability of human umbilical vein endothelial cells by improving the mitochondrial antioxidant capacity and promoting mitophagy [737]. Furthermore, puerarin pretreatment can modulate mitophagy, mitochondrial dynamics, and inflammation to ameliorate palmitate-induced IR in skeletal muscle cells [738].

Therapeutic outlook

As a non-pharmacological intervention, lifestyle changes, such as engaging in physical exercise, show promise in improving MQC and maintaining health. Physical exercises can activate protective mechanisms and preserve mitochondrial health [739]. Accumulating evidence suggests that proper physical exercise can induce PINK/Parkin‐mediated mitophagy improve mitochondrial dynamics, and consequently improve cardiac function after MI [740]. However, this protective effect appears to be related to the frequency and duration of exercise, partly due to an increase in ACTH, cortisol, growth hormone levels, and energy expenditure during prolonged exercise sessions [741]. Besides, in the context of exercise, there is a promotion towards a protective phenotype in terms of mitochondrial dynamics and mitophagy observed in the liver and skeletal muscle [645, 742]. A recent study has illuminated that mild ketosis induced by a proper ketone ester diet can improve cardiac function in diabetic mice by limiting oxidative stress and repairing MQC processes [743].

Given the dual origin of mitochondrial proteins from both mitochondrial and nuclear, it is crucial to maintain the complex and adaptable balance of mitochondrial proteome to preserve mitochondrial function [1]. The accumulation of unfolded or misfolded proteins that are over-synthesized within cells can lead to cellular function, which can be mitigated by chaperones (including HSP60, HSP70) and proteases (including ClpP and LON). Targeting mitochondrial proteostasis as an innovative therapeutic regimen is gaining increasing attention. ClpP is upregulated in various cancers, including acute myeloid leukemia, breast cancer, and lung cancer, where its expression correlates with the viability, growth, resistant, and metastasis of malignancies [744]. ONC201 has been identified as one of the ClpP activators that induce degradation of subunits within the respiratory chain complex, therefore leading to the death of tumor cells [745]. Furthermore, treatment with ClpP activator also inhibits inflammatory and ameliorates diet-induced steatohepatitis in mice [746]. However, the mechanisms and underlying substrates of mitochondrial proteostasis have not been fully elucidated, further investigation is required for relevant therapies.

Given the irreversible mitochondrial damage observed in numerous diseases and the limited efficacy of current treatments, mitochondrial transplantation is emerging as a promising therapeutic strategy. This approach involves supplementing impaired mitochondria with healthy autologous mitochondria isolated from normal tissue. These advances have demonstrated positive outcomes in various mammal disease models, including AKI, MI, and ALI, accompanied by enhanced mitochondrial function and minimized adverse reactions [747,748,749,750]. Even so, further evaluation is required to assess the effectiveness of existing therapy strategies and ensure the stability of transplantation in subsequent stages.

Conclusions and perspectives

Over several decades, the relentless efforts by numerous researchers and the rapid development of experimental techniques have led to deep analysis or initial discovery of the diverse functions of mitochondria in biological organisms. Additionally, there has been a more intuitive observation of the dynamic morphology and distribution of mitochondria. Initially recognized as the crucial energy producers in cells, mitochondria now play a pivotal role in regulating cellular functions and phenotypes in response to physiological signals or external stimuli. They not only maintain the basic functions of various organs under physiological conditions but also play a significant role in buffering stimuli and transmitting signals during disease and injury [751]. Therefore, the MQC mechanisms, which ensure the stable function and high plasticity of mitochondria in eukaryotic cells, have gradually gained attention. In this review, we mainly elaborate on the major mechanisms and regulatable targets involved in MQC regulation, including mitochondrial biogenesis, mitochondrial dynamics, and mitophagy, and discuss their roles in physiological activities and various systemic diseases. Furthermore, while summarizing the latest research progress on therapeutic or intervention approaches targeting different aspects of MQC, we also highlight the challenges and gaps in MQC research, providing potential directions for future investigations.

Various key proteins involved in different main aspects of MQC have been identified. Mitochondrial biogenesis ensures the appropriate quantity of mitochondria and flexible expansion in response to increased energy demand. The pivotal regulator for initiating mitochondrial biogenesis is PGC-1α, which controls the synthesis of new mitochondrial proteins and replication of the mitochondrial genome [11]. Additionally, mitochondrial dynamics underscore the highly dynamic nature of mitochondrial function and quantity, involving a delicate balance between mitochondrial fusion and fission. The key regulators responsible for initiating OMM fusion are MFNs, while OPA1 primarily regulates IMM fusion. Conversely, DRP1 and its associated receptors predominantly regulate mitochondrial fission. Currently, these key proteins serve as indicators to monitor the processes of mitochondrial dynamics [58]. During the clearance process of dysfunctional or severely damaged mitochondria, mitophagy is activated. The main pathways include Parkin-mediated mitophagy and non-Parkin-mediated pathways such as BNIP3/BNIP3L and FUNDC1 [131]. Current research on the regulation of MQC primarily focuses on the transcription and post-translational modifications of these key proteins. AMPK, CaMK, and other kinases, as well as deacetylases such as Sirt1 and Sirt3, are among the prominent targets for regulating MQC. Beyond its fundamental roles in metabolism and energy synthesis, MQC is indispensable for regulating calcium release, maintaining redox balance in the cellular environment, and even participating in PCD processes. As a result, MQC has garnered increasing interest among researchers due to its implications in cancer, metabolic diseases, and acute/chronic injuries affecting various parenchymal organs. However, despite these advancements, the dynamic changes of MQC in various diseases have not yet been fully elucidated by existing studies. Moreover, within the same disease models, the direction of mitophagy and dynamics can even exhibit completely opposite effects. Therefore, further exploration is warranted to identify detection indicators or experimental methods that enable objective and dynamic monitoring of MQC alterations during the disease progression. This endeavor will contribute significantly towards comprehensively understanding the physiological and pathological functions of MQC.

With the increasing recognition of the role of MQC in organisms, there is promising research potential for developing therapeutic strategies and drugs to improve or promote specific aspects of MQC. Current approaches to enhance MQC often involve upregulating or inhibiting key regulators to influence the MQC process, but targeted interventions specifically targeting mitochondria have received limited attention. Further development of precise mitochondrial-targeted biomaterials and their application as carriers to modulate MQC processes could provide innovative avenues for research.

In general, the present review emphasizes and summarizes the crucial functions of MQC in human health and disease, highlighting its unshakable position. Additionally, we provide a systematic description of current treatment strategies for MQC, aiming to provide theoretical support and novel perspectives for further elucidating the role of mitochondria in human life activities and disease progression, as well as facilitating the development of targeted drugs to improve MQC in the future.

Availability of data and materials

Not applicable.

Abbreviations

AD:

Alzheimer’s disease

AKI:

Acute kidney injury

ALI:

Acute lung injury

ALR:

Augmenter of liver regeneration

ALS:

Amyotrophic lateral sclerosis

AMPK:

AMP-activated protein kinase

ATP:

Adenosine triphosphate

BCL2:

B-cell leukemia/lymphoma 2

BNIP3:

BCL2 interacting protein 3

BNIP3L:

BCL2 interacting protein 3 like

CaMK:

Calcium/calmodulin-dependent protein kinase

CD:

Crohn’s Disease

CDK1:

Cyclin B/cyclin-dependent kinase 1

CKD:

Chronic kidney disease

COPD:

Chronic obstructive pulmonary disease

CREB:

CAMP-response element-binding protein

CSCs:

Cancer stem cells

DKD:

Diabetic kidney disease

DM:

Diabetes mellitus

DRP1:

Dynamin-related protein 1

ER:

Endoplasmic reticulum

ERK:

Extracellular regulated protein kinase

ERRs:

Estrogen-related receptors

ETC:

Electron transport chain

FAO:

Fatty acid oxidation

FIS1:

Fission protein 1

FOXO3:

Forkhead box O3

FUNDC1:

FUN14 domain containing 1

H/R:

Hypoxia/reoxygenation

HD:

Huntington’s disease

HF:

Heart failure

HFD:

High-fat diet

HIF-1α:

Hypoxia inducible factor-1α

HSP:

Heat shock protein

I/R:

Ischemia/reperfusion

IBD:

Inflammatory bowel disease

IECs:

Intestinal epithelial cells

IL:

Interleukin

IMM:

Inner mitochondrial membrane

IPF:

Idiopathic pulmonary fibrosis

IR:

Insulin resistance

JNK:

Jun N-terminal kinase

LIR:

LC3-interacting region

LONP1:

Lon protease 1

M phase:

Mitotic phase

MAPK:

Mitogen-activated protein kinases

MCU:

Mitochondrial calcium uniporter

mdivi-1:

Mitochondrial division inhibitor 1

MDVs:

Mitochondria derived vesicles

MFF:

Mitochondrial fission factor

MFN:

Mitofusin

mHtt:

Mutant Huntingtin

MI:

Myocardial infarction

MiD49:

Mitochondrial dynamics protein of 49 kD

mPTP:

Mitochondrial permeability transition pore

MQC:

Mitochondrial quality control

mtDNA:

Mitochondrial DNA

NAFLD:

Non-alcoholic fatty liver disease

NF-κB:

Nuclear factor-κB

NK:

Natural killer

NR4A1:

Nuclear receptor subfamily 4 group A member 1

Nrf2:

Nuclear factor E2-related factor 2

OMM:

Outer mitochondrial membrane

OPA1:

Optic atrophy 1

OPTN:

Optineurin

ox-LDL:

Oxidized low-density lipoprotein

OXPHOS:

Oxidative phosphorylation

PAH:

Pulmonary arterial hypertension

PAMSCs:

Pulmonary artery smooth muscle cells

PCD:

Programmed cell death

PD:

Parkinson’s disease

PGAM5:

Phosphoglycerate mutase 5

PGC-1α:

PPAR-γ coactivator-1α

PINK1:

PTEN-induced kinase 1

PPAR:

Peroxisome proliferator-activated receptor

ROS:

Reactive oxygen species

SENP5:

SUMO specific peptidase 5

Sirt1:

Sirtuin 1

SUMO:

Small ubiquitin-like modifier

T2DM:

Type 2 diabetes mellitus

TAMs:

Tumor-associated macrophages

TBK1:

TANK-binding kinase 1

TCA:

Tricarboxylic acid

TDP-43:

Transactive response DNA-binding protein-43 kD

TFAM:

Mitochondrial transcription factor A

TME:

Tumor microenvironment

TNF:

Tumor necrosis factor

TORC1:

Transducer of Creb-related binding protein 1

UC:

Ulcerative Colitis

ULK1:

Unc-51 like autophagy activating kinase 1

UPS:

Ubiquitin-proteasome system

UUO:

Unilateral ureteral obstruction

VDAC:

Voltage-dependent anion channel

VSMC:

Vascular smooth muscle cell

WAT:

White adipose tissue

References

  1. Song J, Herrmann JM, Becker T. Quality control of the mitochondrial proteome. Nat Rev Mol Cell Biol. 2021;22(1):54–70.

    Article  CAS  PubMed  Google Scholar 

  2. Roca-Portoles A, Tait SWG. Mitochondrial quality control: from molecule to organelle. Cell Mol Life Sci. 2021;78(8):3853–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Tang C, Cai J, Yin XM, Weinberg JM, Venkatachalam MA, Dong Z. Mitochondrial quality control in kidney injury and repair. Nat Rev Nephrol. 2021;17(5):299–318.

    Article  CAS  PubMed  Google Scholar 

  4. Ng MYW, Wai T, Simonsen A. Quality control of the mitochondrion. Dev Cell. 2021;56(7):881–905.

    Article  CAS  PubMed  Google Scholar 

  5. Larson-Casey JL, He C, Carter AB. Mitochondrial quality control in pulmonary fibrosis. Redox Biol. 2020;33:101426.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Chan DC. Mitochondrial dynamics and its involvement in disease. Annu Rev Pathol. 2020;15:235–59.

    Article  CAS  PubMed  Google Scholar 

  7. Onishi M, Yamano K, Sato M, Matsuda N, Okamoto K. Molecular mechanisms and physiological functions of mitophagy. EMBO J. 2021;40(3):e104705.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Choong CJ, Okuno T, Ikenaka K, Baba K, Hayakawa H, Koike M, et al. Alternative mitochondrial quality control mediated by extracellular release. Autophagy. 2021;17(10):2962–74.

    Article  CAS  PubMed  Google Scholar 

  9. Ashrafi G, Schwarz TL. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ. 2013;20(1):31–42.

    Article  CAS  PubMed  Google Scholar 

  10. Holloszy JO. Biochemical adaptations in muscle. Effects of exercise on mitochondrial oxygen uptake and respiratory enzyme activity in skeletal muscle. J Biol Chem. 1967;242(9):2278–82.

    Article  CAS  PubMed  Google Scholar 

  11. Pfanner N, Warscheid B, Wiedemann N. Mitochondrial proteins: from biogenesis to functional networks. Nat Rev Mol Cell Biol. 2019;20(5):267–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Jornayvaz FR, Shulman GI. Regulation of mitochondrial biogenesis. Essays Biochem. 2010;47:69–84.

    Article  CAS  PubMed  Google Scholar 

  13. Jannig PR, Dumesic PA, Spiegelman BM, Ruas JL. SnapShot: regulation and biology of PGC-1α. Cell. 2022;185(8):1444–1444.e1.

    Article  CAS  PubMed  Google Scholar 

  14. Christofides A, Konstantinidou E, Jani C, Boussiotis VA. The role of peroxisome proliferator-activated receptors (PPAR) in immune responses. Metabolism. 2021;114:154338.

    Article  CAS  PubMed  Google Scholar 

  15. Chowdhury PS, Chamoto K, Kumar A, Honjo T. PPAR-induced fatty acid oxidation in T cells increases the number of tumor-reactive CD8+ T cells and facilitates anti-PD-1 therapy. Cancer Immunol Res. 2018;6(11):1375–87.

    Article  CAS  PubMed  Google Scholar 

  16. Sun X, Ping Y, Li X, Mao Y, Chen Y, Shi L, et al. Activation of PGC-1α-dependent mitochondrial biogenesis supports therapeutic effects of silibinin against type I diabetic periodontitis. J Clin Periodontol. 2023;50(7):964–79.

    Article  CAS  PubMed  Google Scholar 

  17. Kumar PR, Saad M, Hellmich C, Mistry JJ, Moore JA, Conway S, et al. PGC-1α induced mitochondrial biogenesis in stromal cells underpins mitochondrial transfer to melanoma. Br J Cancer. 2022;127(1):69–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Wang H, Wang X, Ma L, Huang X, Peng Y, Huang H, et al. PGC-1 alpha regulates mitochondrial biogenesis to ameliorate hypoxia-inhibited cementoblast mineralization. Ann N Y Acad Sci. 2022;1516(1):300–11.

    Article  CAS  PubMed  Google Scholar 

  19. Jamwal S, Blackburn JK, Elsworth JD. PPARγ/PGC1α signaling as a potential therapeutic target for mitochondrial biogenesis in neurodegenerative disorders. Pharmacol Ther. 2021;219:107705.

    Article  CAS  PubMed  Google Scholar 

  20. Gentric G, Kieffer Y, Mieulet V, Goundiam O, Bonneau C, Nemati F, et al. PML-regulated mitochondrial metabolism enhances chemosensitivity in human ovarian cancers. Cell Metab. 2019;29(1):156–173.e10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Qian X, Li X, Shi Z, Bai X, Xia Y, Zheng Y, et al. KDM3A senses oxygen availability to regulate PGC-1α-mediated mitochondrial biogenesis. Mol Cell. 2019;76(6):885–895.e7.

    Article  CAS  PubMed  Google Scholar 

  22. Hu S, Feng J, Wang M, Wufuer R, Liu K, Zhang Z, et al. Nrf1 is an indispensable redox-determining factor for mitochondrial homeostasis by integrating multi-hierarchical regulatory networks. Redox Biol. 2022;57:102470.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Itoh Y, Khawaja A, Laptev I, Cipullo M, Atanassov I, Sergiev P, et al. Mechanism of mitoribosomal small subunit biogenesis and preinitiation. Nature. 2022;606(7914):603–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Raggi C, Taddei ML, Sacco E, Navari N, Correnti M, Piombanti B, et al. Mitochondrial oxidative metabolism contributes to a cancer stem cell phenotype in cholangiocarcinoma. J Hepatol. 2021;74(6):1373–85.

    Article  CAS  PubMed  Google Scholar 

  25. Giguère V. To ERR in the estrogen pathway. Trends Endocrinol Metab. 2002;13(5):220–5.

    Article  PubMed  Google Scholar 

  26. Cho Y, Hazen BC, Russell AP, Kralli A. Peroxisome proliferator-activated receptor γ coactivator 1 (PGC-1)- and estrogen-related receptor (ERR)-induced regulator in muscle 1 (Perm1) is a tissue-specific regulator of oxidative capacity in skeletal muscle cells. J Biol Chem. 2013;288(35):25207–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Charest-Marcotte A, Dufour CR, Wilson BJ, Tremblay AM, Eichner LJ, Arlow DH, et al. The homeobox protein Prox1 is a negative modulator of ERR{alpha}/PGC-1{alpha} bioenergetic functions. Genes Dev. 2010;24(6):537–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Sakamoto T, Batmanov K, Wan S, Guo Y, Lai L, Vega RB, et al. The nuclear receptor ERR cooperates with the cardiogenic factor GATA4 to orchestrate cardiomyocyte maturation. Nat Commun. 2022;13(1):1991.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Luo C, Widlund HR, Puigserver P. PGC-1 coactivators: shepherding the mitochondrial biogenesis of tumors. Trends Cancer. 2016;2(10):619–31.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Laurin KM, Coutu-Beaudry K, Salazar A, Méribout N, Audet-Walsh É, Gravel SP. Low expression of PGC-1β and other mitochondrial biogenesis modulators in melanoma is associated with growth arrest and the induction of an immunosuppressive gene expression program dependent on MEK and IRF-1. Cancer Lett. 2022;541:215738.

    Article  CAS  PubMed  Google Scholar 

  31. Dorn GW, Vega RB, Kelly DP. Mitochondrial biogenesis and dynamics in the developing and diseased heart. Genes Dev. 2015;29(19):1981–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Vats D 2nd, Mukundan L, Odegaard JI, Zhang L, Smith KL, Morel CR, et al. Oxidative metabolism and PGC-1beta attenuate macrophage-mediated inflammation. Cell Metab. 2006;4(1):13–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Herzig S, Shaw RJ. AMPK: guardian of metabolism and mitochondrial homeostasis. Nat Rev Mol Cell Biol. 2018;19(2):121–35.

    Article  CAS  PubMed  Google Scholar 

  34. Dong YZ, Li L, Espe M, Lu KL, Rahimnejad S. Hydroxytyrosol attenuates hepatic fat accumulation via activating mitochondrial biogenesis and autophagy through the AMPK pathway. J Agric Food Chem. 2020;68(35):9377–86.

    Article  CAS  PubMed  Google Scholar 

  35. Yan W, Zhang H, Liu P, Wang H, Liu J, Gao C, et al. Impaired mitochondrial biogenesis due to dysfunctional adiponectin-AMPK-PGC-1α signaling contributing to increased vulnerability in diabetic heart. Basic Res Cardiol. 2013;108(3):329.

    Article  PubMed  Google Scholar 

  36. Yang L, Li X, Jiang A, Li X, Chang W, Chen J, et al. Metformin alleviates lead-induced mitochondrial fragmentation via AMPK/Nrf2 activation in SH-SY5Y cells. Redox Biol. 2020;36:101626.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Nakanishi A, Hatano N, Fujiwara Y, Sha’ri A, Takabatake S, Akano H, et al. AMP-activated protein kinase-mediated feedback phosphorylation controls the Ca2+/calmodulin (CaM) dependence of Ca2+/CaM-dependent protein kinase kinase β. J Biol Chem. 2017;292(48):19804–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Jung KA, Lee S, Kwak MK. NFE2L2/NRF2 activity is linked to mitochondria and AMP-activated protein kinase signaling in cancers through miR-181c/mitochondria-encoded cytochrome C oxidase regulation. Antioxid Redox Signal. 2017;27(13):945–61.

    Article  CAS  PubMed  Google Scholar 

  39. Cantó C, Gerhart-Hines Z, Feige JN, Lagouge M, Noriega L, Milne JC, et al. AMPK regulates energy expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature. 2009;458(7241):1056–60.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Chen Y, Wu YY, Si H-B, Lu YR, Shen B. Mechanistic insights into AMPK-SIRT3 positive feedback loop-mediated chondrocyte mitochondrial quality control in osteoarthritis pathogenesis. Pharmacol Res. 2021;166:105497.

    Article  CAS  PubMed  Google Scholar 

  41. Narkar VA, Downes M, Yu RT, Embler E, Wang YX, Banayo E, et al. AMPK and PPARdelta agonists are exercise mimetics. Cell. 2008;134(3):405–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Iwabu M, Yamauchi T, Okada-Iwabu M, Sato K, Nakagawa T, Funata M, et al. Adiponectin and AdipoR1 regulate PGC-1alpha and mitochondria by Ca2+ and AMPK/SIRT1. Nature. 2010;464(7293):1313–9.

    Article  CAS  PubMed  Google Scholar 

  43. Mercy L, de Pauw AD, Payen L, Tejerina S, Houbion A, Demazy C, et al. Mitochondrial biogenesis in mtDNA-depleted cells involves a Ca2+-dependent pathway and a reduced mitochondrial protein import. FEBS J. 2005;272(19):5031–55.

    Article  CAS  PubMed  Google Scholar 

  44. Jhun BS, Mishra J, Monaco S, Fu D, Jiang W, Sheu SS, et al. The mitochondrial Ca2+ uniporter: regulation by auxiliary subunits and signal transduction pathways. Am J Physiol Cell Physiol. 2016;311(1):C67–80.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Wright DC, Geiger PC, Han DH, Jones TE, Holloszy JO. Calcium induces increases in peroxisome proliferator-activated receptor gamma coactivator-1alpha and mitochondrial biogenesis by a pathway leading to p38 mitogen-activated protein kinase activation. J Biol Chem. 2007;282(26):18793–9.

    Article  CAS  PubMed  Google Scholar 

  46. Handschin C, Rhee J, Lin J, Tarr PT, Spiegelman BM. An autoregulatory loop controls peroxisome proliferator-activated receptor gamma coactivator 1alpha expression in muscle. Proc Natl Acad Sci U S A. 2003;100(12):7111–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Liu X, Wang S, Guo X, Li Y, Ogurlu R, Lu F, et al. Increased reactive oxygen species-mediated Ca2+/calmodulin-dependent protein kinase II activation contributes to calcium handling abnormalities and impaired contraction in Barth syndrome. Circulation. 2021;143(19):1894–911.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Wang S, Wan T, Ye M, Qiu Y, Pei L, Jiang R, et al. Nicotinamide riboside attenuates alcohol induced liver injuries via activation of SirT1/PGC-1α/mitochondrial biosynthesis pathway. Redox Biol. 2018;17:89–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Xu Y, Yu T, Ma G, Zheng L, Jiang X, Yang F, et al. Berberine modulates deacetylation of PPARγ to promote adipose tissue remodeling and thermogenesis via AMPK/SIRT1 pathway. Int J Biol Sci. 2021;17(12):3173–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Ruiz-Andres O, Sanchez-Niño MD, Cannata-Ortiz P, Ruiz-Ortega M, Egido J, Ortiz A, et al. Histone lysine crotonylation during acute kidney injury in mice. Dis Model Mech. 2016;9(6):633–45.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Chaturvedi RK, M FB. Mitochondrial diseases of the brain. Free Radic Biol Med. 2013;63:1–29.

    Article  CAS  PubMed  Google Scholar 

  52. Ruiz-Andres O, Suarez-Alvarez B, Sánchez-Ramos C, Monsalve M, Sanchez-Niño MD, Ruiz-Ortega M, et al. The inflammatory cytokine TWEAK decreases PGC-1α expression and mitochondrial function in acute kidney injury. Kidney Int. 2016;89(2):399–410.

    Article  CAS  PubMed  Google Scholar 

  53. Ranea-Robles P, Galino J, Espinosa L, Schlüter A, Ruiz M, Calingasan NY, et al. Modulation of mitochondrial and inflammatory homeostasis through RIP140 is neuroprotective in an adrenoleukodystrophy mouse model. Neuropathol Appl Neurobiol. 2022;48(1):e12747.

    Article  CAS  PubMed  Google Scholar 

  54. Wai T, Langer T. Mitochondrial dynamics and metabolic regulation. Trends Endocrinol Metab. 2016;27(2):105–17.

    Article  CAS  PubMed  Google Scholar 

  55. Gao S, Hu J. Mitochondrial fusion: the machineries in and out. Trends Cell Biol. 2021;31(1):62–74.

    Article  CAS  PubMed  Google Scholar 

  56. Giacomello M, Pyakurel A, Glytsou C, Scorrano L. The cell biology of mitochondrial membrane dynamics. Nat Rev Mol Cell Biol. 2020;21(4):204–24.

    Article  CAS  PubMed  Google Scholar 

  57. Jin JY, Wei XX, Zhi XL, Wang XH, Meng D. Drp1-dependent mitochondrial fission in cardiovascular disease. Acta Pharmacol Sin. 2021;42(5):655–64.

    Article  CAS  PubMed  Google Scholar 

  58. Tilokani L, Nagashima S, Paupe V, Prudent J. Mitochondrial dynamics: overview of molecular mechanisms. Essays Biochem. 2018;62(3):341–60.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Cao YL, Meng S, Chen Y, Feng JX, Gu DD, Yu B, et al. MFN1 structures reveal nucleotide-triggered dimerization critical for mitochondrial fusion. Nature. 2017;542(7641):372–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Rojo M, Legros F, Chateau D, Lombès A. Membrane topology and mitochondrial targeting of mitofusins, ubiquitous mammalian homologs of the transmembrane GTPase Fzo. J Cell Sci. 2002;115(Pt 8):1663–74.

    Article  CAS  PubMed  Google Scholar 

  61. Li S, Han S, Zhang Q, Zhu Y, Zhang H, Wang J, et al. FUNDC2 promotes liver tumorigenesis by inhibiting MFN1-mediated mitochondrial fusion. Nat Commun. 2022;13(1):3486.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Muñoz JP, Ivanova S, Sánchez-Wandelmer J, Martínez-Cristóbal P, Noguera E, Sancho A, et al. Mfn2 modulates the UPR and mitochondrial function via repression of PERK. EMBO J. 2013;32(17):2348–61.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Basso V, Marchesan E, Peggion C, Chakraborty J, von Stockum S, Giacomello M, et al. Regulation of ER-mitochondria contacts by Parkin via Mfn2. Pharmacol Res. 2018;138:43–56.

    Article  CAS  PubMed  Google Scholar 

  64. Hu Y, Chen H, Zhang L, Lin X, Li X, Zhuang H, et al. The AMPK-MFN2 axis regulates MAM dynamics and autophagy induced by energy stresses. Autophagy. 2021;17(5):1142–56.

    Article  CAS  PubMed  Google Scholar 

  65. Song Z, Song H, Liu D, Yan B, Wang D, Zhang Y, et al. Overexpression of MFN2 alleviates sorafenib-induced cardiomyocyte necroptosis via the MAM-CaMKIIδ pathway in vitro and in vivo. Theranostics. 2022;12(3):1267–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Chen Y, Dorn GW 2nd. PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria. Science. 2013;340(6131):471–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Yamada T, Dawson TM, Yanagawa T, Iijima M, Sesaki H. SQSTM1/p62 promotes mitochondrial ubiquitination independently of PINK1 and PRKN/parkin in mitophagy. Autophagy. 2019;15(11):2012–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Park YY, Nguyen OTK, Kang H, Cho H. MARCH5-mediated quality control on acetylated Mfn1 facilitates mitochondrial homeostasis and cell survival. Cell Death Dis. 2014;5(4):e1172.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Wang H, Yi X, Guo S, Wang S, Ma J, Zhao T, et al. The XBP1-MARCH5-MFN2 axis confers endoplasmic reticulum stress resistance by coordinating mitochondrial fission and mitophagy in melanoma. J Invest Dermatol. 2021;141(12):2932–2943.e12.

    Article  CAS  PubMed  Google Scholar 

  70. Pyakurel A, Savoia C, Hess D, Scorrano L. Extracellular regulated kinase phosphorylates mitofusin 1 to control mitochondrial morphology and apoptosis. Mol Cell. 2015;58(2):244–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Sebastián D, Hernández-Alvarez MI, Segalés J, Sorianello E, Muñoz JP, Sala D, et al. Mitofusin 2 (Mfn2) links mitochondrial and endoplasmic reticulum function with insulin signaling and is essential for normal glucose homeostasis. Proc Natl Acad Sci U S A. 2012;109(14):5523–8.

    Article  PubMed  PubMed Central  Google Scholar 

  72. Leboucher GP, Tsai YC, Yang M, Shaw KC, Zhou M, Veenstra TD, et al. Stress-induced phosphorylation and proteasomal degradation of mitofusin 2 facilitates mitochondrial fragmentation and apoptosis. Mol Cell. 2012;47(4):547–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Zhang X, Qin Y, Ruan W, Wan X, Lv C, He L, et al. Targeting inflammation-associated AMPK//Mfn-2/MAPKs signaling pathways by baicalein exerts anti-atherosclerotic action. Phytother Res. 2021;35(8):4442–55.

    Article  PubMed  Google Scholar 

  74. Biel TG, Lee S, Flores-Toro JA, Dean JW, Go KL, Lee MH, et al. Sirtuin 1 suppresses mitochondrial dysfunction of ischemic mouse livers in a mitofusin 2-dependent manner. Cell Death Differ. 2016;23(2):279–90.

    Article  CAS  PubMed  Google Scholar 

  75. Lee JY, Kapur M, Li M, Choi M-C, Choi S, Kim HJ, et al. MFN1 deacetylation activates adaptive mitochondrial fusion and protects metabolically challenged mitochondria. J Cell Sci. 2014;127(Pt 22):4954–63.

    PubMed  PubMed Central  Google Scholar 

  76. Frezza C, Cipolat S, Martins de Brito O, Micaroni M, Beznoussenko GV, Rudka T, et al. OPA1 controls apoptotic cristae remodeling independently from mitochondrial fusion. Cell. 2006;126(1):177–89.

    Article  CAS  PubMed  Google Scholar 

  77. Noone J, O’Gorman DJ, Kenny HC. OPA1 regulation of mitochondrial dynamics in skeletal and cardiac muscle. Trends Endocrinol Metab. 2022;33(10):710–21.

    Article  CAS  PubMed  Google Scholar 

  78. Cipolat S, Martins de Brito O, Dal Zilio B, Scorrano L. OPA1 requires mitofusin 1 to promote mitochondrial fusion. Proc Natl Acad Sci U S A. 2004;101(45):15927–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Anand R, Wai T, Baker MJ, Kladt N, Schauss AC, Rugarli E, et al. The i-AAA protease YME1L and OMA1 cleave OPA1 to balance mitochondrial fusion and fission. J Cell Biol. 2014;204(6):919–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Wai T, García-Prieto J, Baker MJ, Merkwirth C, Benit P, Rustin P, et al. Imbalanced OPA1 processing and mitochondrial fragmentation cause heart failure in mice. Science. 2015;350(6265):aad0116.

    Article  PubMed  Google Scholar 

  81. Yamaguchi R, Lartigue L, Perkins G, Scott RT, Dixit A, Kushnareva Y, et al. Opa1-mediated cristae opening is Bax/Bak and BH3 dependent, required for apoptosis, and independent of Bak oligomerization. Mol Cell. 2008;31(4):557–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Martin OJ, Lai L, Soundarapandian MM, Leone TC, Zorzano A, Keller MP, et al. A role for peroxisome proliferator-activated receptor γ coactivator-1 in the control of mitochondrial dynamics during postnatal cardiac growth. Circ Res. 2014;114(4):626–36.

    Article  CAS  PubMed  Google Scholar 

  83. Nan J, Hu H, Sun Y, Zhu L, Wang Y, Zhong Z, et al. TNFR2 stimulation promotes mitochondrial fusion via Stat3- and NF-kB-dependent activation of OPA1 expression. Circ Res. 2017;121(4):392–410.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Wang R, Xu H, Tan B, Yi Q, Sun Y, Xiang H, et al. SIRT3 promotes metabolic maturation of human iPSC-derived cardiomyocytes via OPA1-controlled mitochondrial dynamics. Free Radic Biol Med. 2023;195:270–82.

    Article  CAS  PubMed  Google Scholar 

  85. He J, Shangguan X, Zhou W, Cao Y, Zheng Q, Tu J, et al. Glucose limitation activates AMPK coupled SENP1-Sirt3 signalling in mitochondria for T cell memory development. Nat Commun. 2021;12(1):4371.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Kraus F, Roy K, Pucadyil TJ, Ryan MT. Function and regulation of the divisome for mitochondrial fission. Nature. 2021;590(7844):57–66.

    Article  CAS  PubMed  Google Scholar 

  87. Quiles JM, Gustafsson ÅB. The role of mitochondrial fission in cardiovascular health and disease. Nat Rev Cardiol. 2022;19(11):723–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Simula L, Campanella M, Campello S. Targeting Drp1 and mitochondrial fission for therapeutic immune modulation. Pharmacol Res. 2019;146:104317.

    Article  CAS  PubMed  Google Scholar 

  89. Zhao J, Lendahl U, Nistér M. Regulation of mitochondrial dynamics: convergences and divergences between yeast and vertebrates. Cell Mol Life Sci. 2013;70(6):951–76.

    Article  CAS  PubMed  Google Scholar 

  90. Atkins K, Dasgupta A, Chen KH, Mewburn J, Archer SL. The role of Drp1 adaptor proteins MiD49 and MiD51 in mitochondrial fission: implications for human disease. Clin Sci (Lond). 2016;130(21):1861–74.

    Article  CAS  PubMed  Google Scholar 

  91. Qin L, Xi S. The role of mitochondrial fission proteins in mitochondrial dynamics in kidney disease. Int J Mol Sci. 2022;23(23):14725.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Lewis TL, Kwon SK, Lee A, Shaw R, Polleux F. MFF-dependent mitochondrial fission regulates presynaptic release and axon branching by limiting axonal mitochondria size. Nat Commun. 2018;9(1):5008.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Palmer CS, Elgass KD, Parton RG, Osellame LD, Stojanovski D, Ryan MT. Adaptor proteins MiD49 and MiD51 can act independently of Mff and Fis1 in Drp1 recruitment and are specific for mitochondrial fission. J Biol Chem. 2013;288(38):27584–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Zhao J, Liu T, Jin S, Wang X, Qu M, Uhlén P, et al. Human MIEF1 recruits Drp1 to mitochondrial outer membranes and promotes mitochondrial fusion rather than fission. EMBO J. 2011;30(14):2762–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Losón OC, Liu R, Rome ME, Meng S, Kaiser JT, Shan S-O, et al. The mitochondrial fission receptor MiD51 requires ADP as a cofactor. Structure. 2014;22(3):367–77.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Ko HJ, Tsai CY, Chiou SJ, Lai YL, Wang CH, Cheng JT, et al. The phosphorylation status of Drp1-Ser637 by PKA in mitochondrial fission modulates mitophagy via PINK1/parkin to exert multipolar spindles assembly during mitosis. Biomolecules. 2021;11(3):424.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Taguchi N, Ishihara N, Jofuku A, Oka T, Mihara K. Mitotic phosphorylation of dynamin-related GTPase Drp1 participates in mitochondrial fission. J Biol Chem. 2007;282(15):11521–9.

    Article  CAS  PubMed  Google Scholar 

  98. Zhan L, Lu Z, Zhu X, Xu W, Li L, Li X, et al. Hypoxic preconditioning attenuates necroptotic neuronal death induced by global cerebral ischemia via Drp1-dependent signaling pathway mediated by CaMKIIα inactivation in adult rats. FASEB J. 2019;33(1):1313–29.

    Article  CAS  PubMed  Google Scholar 

  99. Zhao X, Xu H, Li Y, Liu Y, Li X, Zhou W, et al. Silica nanoparticles perturbed mitochondrial dynamics and induced myocardial apoptosis via PKA-DRP1-mitochondrial fission signaling. Sci Total Environ. 2022;842:156854.

    Article  CAS  PubMed  Google Scholar 

  100. Wang Z, Jiang H, Chen S, Du F, Wang X. The mitochondrial phosphatase PGAM5 functions at the convergence point of multiple necrotic death pathways. Cell. 2012;148(1–2):228–43.

    Article  CAS  PubMed  Google Scholar 

  101. Gao Q, Tian R, Han H, Slone J, Wang C, Ke X, et al. PINK1-mediated Drp1S616 phosphorylation modulates synaptic development and plasticity via promoting mitochondrial fission. Signal Transduct Target Ther. 2022;7(1):103.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Yu W, Wang X, Zhao J, Liu R, Liu J, Wang Z, et al. Stat2-Drp1 mediated mitochondrial mass increase is necessary for pro-inflammatory differentiation of macrophages. Redox Biol. 2020;37:101761.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Eifler K, Vertegaal ACO. SUMOylation-mediated regulation of cell cycle progression and cancer. Trends Biochem Sci. 2015;40(12):779–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Prudent J, Zunino R, Sugiura A, Mattie S, Shore GC, McBride HM. MAPL SUMOylation of Drp1 stabilizes an ER/mitochondrial platform required for cell death. Mol Cell. 2015;59(6):941–55.

    Article  CAS  PubMed  Google Scholar 

  105. Yamada S, Sato A, Ishihara N, Akiyama H, Sakakibara SI. Drp1 SUMO/deSUMOylation by Senp5 isoforms influences ER tubulation and mitochondrial dynamics to regulate brain development. iScience. 2021;24(12):103484.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Lewis SC, Uchiyama LF, Nunnari J. ER-mitochondria contacts couple mtDNA synthesis with mitochondrial division in human cells. Science. 2016;353(6296):aaf5549.

    Article  PubMed  PubMed Central  Google Scholar 

  107. Manor U, Bartholomew S, Golani G, Christenson E, Kozlov M, Higgs H, et al. A mitochondria-anchored isoform of the actin-nucleating spire protein regulates mitochondrial division. Elife. 2015;4:e08828.

    Article  PubMed  PubMed Central  Google Scholar 

  108. Adachi Y, Kato T, Yamada T, Murata D, Arai K, Stahelin RV, et al. Drp1 tubulates the ER in a GTPase-independent manner. Mol Cell. 2020;80(4):621–32.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Yu Y, Peng XD, Qian XJ, Zhang KM, Huang X, Chen YH, et al. Fis1 phosphorylation by Met promotes mitochondrial fission and hepatocellular carcinoma metastasis. Signal Transduct Target Ther. 2021;6(1):401.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Yonashiro R, Ishido S, Kyo S, Fukuda T, Goto E, Matsuki Y, et al. A novel mitochondrial ubiquitin ligase plays a critical role in mitochondrial dynamics. EMBO J. 2006;25(15):3618–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Jin Q, Li R, Hu N, Xin T, Zhu P, Hu S, et al. DUSP1 alleviates cardiac ischemia/reperfusion injury by suppressing the Mff-required mitochondrial fission and Bnip3-related mitophagy via the JNK pathways. Redox Biol. 2018;14:576–87.

    Article  CAS  PubMed  Google Scholar 

  112. Ban T, Ishihara T, Kohno H, Saita S, Ichimura A, Maenaka K, et al. Molecular basis of selective mitochondrial fusion by heterotypic action between OPA1 and cardiolipin. Nat Cell Biol. 2017;19(7):856–63.

    Article  CAS  PubMed  Google Scholar 

  113. Gilkerson R, De La Torre P, St VS. Mitochondrial OMA1 and OPA1 as gatekeepers of organellar structure/function and cellular stress response. Front Cell Dev Biol. 2021;9:626117.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Ge Y, Shi X, Boopathy S, McDonald J, Smith AW, Chao LH. Two forms of Opa1 cooperate to complete fusion of the mitochondrial inner-membrane. Elife. 2020;9:e50973.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Cogliati S, Enriquez JA, Scorrano L. Mitochondrial cristae: where beauty meets functionality. Trends Biochem Sci. 2016;41(3):261–73.

    Article  CAS  PubMed  Google Scholar 

  116. Kim KH, Lee MS. Autophagy–a key player in cellular and body metabolism. Nat Rev Endocrinol. 2014;10(6):322–37.

    Article  CAS  PubMed  Google Scholar 

  117. Gao W, Wang X, Zhou Y, Wang X, Yu Y. Autophagy, ferroptosis, pyroptosis, and necroptosis in tumor immunotherapy. Signal Transduct Target Ther. 2022;7(1):196.

    Article  PubMed  PubMed Central  Google Scholar 

  118. Klionsky DJ, Petroni G, Amaravadi RK, Baehrecke EH, Ballabio A, Boya P, et al. Autophagy in major human diseases. EMBO J. 2021;40(19):e108863.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Johansen T, Lamark T. Selective autophagy: ATG8 family proteins, LIR motifs and cargo receptors. J Mol Biol. 2020;432(1):80–103.

    Article  CAS  PubMed  Google Scholar 

  120. Wauer T, Komander D. Structure of the human Parkin ligase domain in an autoinhibited state. EMBO J. 2013;32(15):2099–112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Chan NC, Salazar AM, Pham AH, Sweredoski MJ, Kolawa NJ, Graham RL, et al. Broad activation of the ubiquitin-proteasome system by Parkin is critical for mitophagy. Hum Mol Genet. 2011;20(9):1726–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Xian H, Liou YC. Loss of MIEF1/MiD51 confers susceptibility to BAX-mediated cell death and PINK1-PRKN-dependent mitophagy. Autophagy. 2019;15(12):2107–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Sulkshane P, Ram J, Thakur A, Reis N, Kleifeld O, Glickman MH. Ubiquitination and receptor-mediated mitophagy converge to eliminate oxidation-damaged mitochondria during hypoxia. Redox Biol. 2021;45:102047.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. McLelland GL, Goiran T, Yi W, Dorval G, Chen CX, Lauinger ND, et al. Mfn2 ubiquitination by PINK1/parkin gates the p97-dependent release of ER from mitochondria to drive mitophagy. Elife. 2018;7:e32866.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Yoshii SR, Kishi C, Ishihara N, Mizushima N. Parkin mediates proteasome-dependent protein degradation and rupture of the outer mitochondrial membrane. J Biol Chem. 2011;286(22):19630–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Rakovic A, Ziegler J, Mårtensson CU, Prasuhn J, Shurkewitsch K, König P, et al. PINK1-dependent mitophagy is driven by the UPS and can occur independently of LC3 conversion. Cell Death Differ. 2019;26(8):1428–41.

    Article  CAS  PubMed  Google Scholar 

  127. Gladkova C, Maslen SL, Skehel JM, Komander D. Mechanism of parkin activation by PINK1. Nature. 2018;559(7714):410–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Schubert AF, Gladkova C, Pardon E, Wagstaff JL, Freund SMV, Steyaert J, et al. Structure of PINK1 in complex with its substrate ubiquitin. Nature. 2017;552(7683):51–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Nguyen TN, Padman BS, Usher J, Oorschot V, Ramm G, Lazarou M. Atg8 family LC3/GABARAP proteins are crucial for autophagosome-lysosome fusion but not autophagosome formation during PINK1/Parkin mitophagy and starvation. J Cell Biol. 2016;215(6):857–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Geisler S, Holmström KM, Skujat D, Fiesel FC, Rothfuss OC, Kahle PJ, et al. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat Cell Biol. 2010;12(2):119–31.

    Article  CAS  PubMed  Google Scholar 

  131. Vargas JNS, Hamasaki M, Kawabata T, Youle RJ, Yoshimori T. The mechanisms and roles of selective autophagy in mammals. Nat Rev Mol Cell Biol. 2023;24(3):167–85.

    Article  CAS  PubMed  Google Scholar 

  132. Lazarou M, Sliter DA, Kane LA, Sarraf SA, Wang C, Burman JL, et al. The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature. 2015;524(7565):309–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Wong YC, Kim S, Peng W, Krainc D. Regulation and function of mitochondria-lysosome membrane contact sites in cellular homeostasis. Trends Cell Biol. 2019;29(6):500–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. McEwan DG, Popovic D, Gubas A, Terawaki S, Suzuki H, Stadel D, et al. PLEKHM1 regulates autophagosome-lysosome fusion through HOPS complex and LC3/GABARAP proteins. Mol Cell. 2015;57(1):39–54.

    Article  CAS  PubMed  Google Scholar 

  135. Zhang J, Ney PA. Role of BNIP3 and NIX in cell death, autophagy, and mitophagy. Cell Death Differ. 2009;16(7):939–46.

    Article  CAS  PubMed  Google Scholar 

  136. Chinnadurai G, Vijayalingam S, Gibson SB. BNIP3 subfamily BH3-only proteins: mitochondrial stress sensors in normal and pathological functions. Oncogene. 2008;27(Suppl 1):S114–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Zhu Y, Massen S, Terenzio M, Lang V, Chen-Lindner S, Eils R, et al. Modulation of serines 17 and 24 in the LC3-interacting region of Bnip3 determines pro-survival mitophagy versus apoptosis. J Biol Chem. 2013;288(2):1099–113.

    Article  CAS  PubMed  Google Scholar 

  138. Zhang J, Loyd MR, Randall MS, Waddell MB, Kriwacki RW, Ney PA. A short linear motif in BNIP3L (NIX) mediates mitochondrial clearance in reticulocytes. Autophagy. 2012;8(9):1325–32.

    Article  PubMed  PubMed Central  Google Scholar 

  139. Rogov VV, Suzuki H, Marinković M, Lang V, Kato R, Kawasaki M, et al. Phosphorylation of the mitochondrial autophagy receptor Nix enhances its interaction with LC3 proteins. Sci Rep. 2017;7(1):1131.

    Article  PubMed  PubMed Central  Google Scholar 

  140. Yuan Y, Zheng Y, Zhang X, Chen Y, Wu X, Wu J, et al. BNIP3L/NIX-mediated mitophagy protects against ischemic brain injury independent of PARK2. Autophagy. 2017;13(10):1754–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. da Silva Rosa SC, Martens MD, Field JT, Nguyen L, Kereliuk SM, Hai Y, et al. BNIP3L/Nix-induced mitochondrial fission, mitophagy, and impaired myocyte glucose uptake are abrogated by PRKA/PKA phosphorylation. Autophagy. 2021;17(9):2257–72.

    Article  PubMed  Google Scholar 

  142. Melser S, Chatelain EH, Lavie J, Mahfouf W, Jose C, Obre E, et al. Rheb regulates mitophagy induced by mitochondrial energetic status. Cell Metab. 2013;17(5):719–30.

    Article  CAS  PubMed  Google Scholar 

  143. Shi RY, Zhu SH, Li V, Gibson SB, Xu XS, Kong JM. BNIP3 interacting with LC3 triggers excessive mitophagy in delayed neuronal death in stroke. CNS Neurosci Ther. 2014;20(12):1045–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Gao F, Chen D, Si J, Hu Q, Qin Z, Fang M, et al. The mitochondrial protein BNIP3L is the substrate of PARK2 and mediates mitophagy in PINK1/PARK2 pathway. Hum Mol Genet. 2015;24(9):2528–38.

    Article  CAS  PubMed  Google Scholar 

  145. Zhang T, Xue L, Li L, Tang C, Wan Z, Wang R, et al. BNIP3 protein suppresses PINK1 kinase proteolytic cleavage to promote mitophagy. J Biol Chem. 2016;291(41):21616–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Liu H, Zang C, Yuan F, Ju C, Shang M, Ning J, et al. The role of FUNDC1 in mitophagy, mitochondrial dynamics and human diseases. Biochem Pharmacol. 2022;197:114891.

    Article  CAS  PubMed  Google Scholar 

  147. Lv M, Wang C, Li F, Peng J, Wen B, Gong Q, et al. Structural insights into the recognition of phosphorylated FUNDC1 by LC3B in mitophagy. Protein Cell. 2017;8(1):25–38.

    Article  CAS  PubMed  Google Scholar 

  148. Zheng T, Wang HY, Chen Y, Chen X, Wu ZL, Hu QY, et al. Src activation aggravates podocyte injury in diabetic nephropathy via suppression of FUNDC1-mediated mitophagy. Front Pharmacol. 2022;13:897046.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Chen G, Han Z, Feng D, Chen Y, Chen L, Wu H, et al. A regulatory signaling loop comprising the PGAM5 phosphatase and CK2 controls receptor-mediated mitophagy. Mol Cell. 2014;54(3):362–77.

    Article  CAS  PubMed  Google Scholar 

  150. Wu W, Tian W, Hu Z, Chen G, Huang L, Li W, et al. ULK1 translocates to mitochondria and phosphorylates FUNDC1 to regulate mitophagy. EMBO Rep. 2014;15(5):566–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Strappazzon F, Di Rita A, Peschiaroli A, Leoncini PP, Locatelli F, Melino G, et al. HUWE1 controls MCL1 stability to unleash AMBRA1-induced mitophagy. Cell Death Differ. 2020;27(4):1155–68.

    Article  CAS  PubMed  Google Scholar 

  152. Tooze SA, Codogno P. Compartmentalized regulation of autophagy regulators: fine-tuning AMBRA1 by Bcl-2. EMBO J. 2011;30(7):1185–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Otsu K, Murakawa T, Yamaguchi O. BCL2L13 is a mammalian homolog of the yeast mitophagy receptor Atg32. Autophagy. 2015;11(10):1932–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Murakawa T, Yamaguchi O, Hashimoto A, Hikoso S, Takeda T, Oka T, et al. Bcl-2-like protein 13 is a mammalian Atg32 homologue that mediates mitophagy and mitochondrial fragmentation. Nat Commun. 2015;6:7527.

    Article  PubMed  Google Scholar 

  155. Fang Q, Zheng S, Chen Q, Chen L, Yang Y, Wang Y, et al. The protective effect of inhibiting mitochondrial fission on the juvenile rat brain following PTZ kindling through inhibiting the BCL2L13/LC3 mitophagy pathway. Metab Brain Dis. 2023;38(2):453–66.

    Article  CAS  PubMed  Google Scholar 

  156. Sugiura A, McLelland GL, Fon EA, McBride HM. A new pathway for mitochondrial quality control: mitochondrial-derived vesicles. EMBO J. 2014;33(19):2142–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Towers CG, Wodetzki DK, Thorburn J, Smith KR, Caino MC, Thorburn A. Mitochondrial-derived vesicles compensate for loss of LC3-mediated mitophagy. Dev Cell. 2021;56(14):2029–42.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Soubannier V, McLelland GL, Zunino R, Braschi E, Rippstein P, Fon EA, et al. A vesicular transport pathway shuttles cargo from mitochondria to lysosomes. Curr Biol. 2012;22(2):135–41.

    Article  CAS  PubMed  Google Scholar 

  159. McLelland GL, Soubannier V, Chen CX, McBride HM, Fon EA. Parkin and PINK1 function in a vesicular trafficking pathway regulating mitochondrial quality control. EMBO J. 2014;33(4):282–95.

    CAS  PubMed  PubMed Central  Google Scholar 

  160. den Brave F, Gupta A, Becker T. Protein quality control at the mitochondrial surface. Front Cell Dev Biol. 2021;9:795685.

    Article  Google Scholar 

  161. McLelland GL, Lee SA, McBride HM, Fon EA. Syntaxin-17 delivers PINK1/parkin-dependent mitochondrial vesicles to the endolysosomal system. J Cell Biol. 2016;214(3):275–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Ryan TA, Phillips EO, Collier CL, Jb Robinson A, Routledge D, Wood RE, et al. Tollip coordinates Parkin-dependent trafficking of mitochondrial-derived vesicles. EMBO J. 2020;39(11):e102539.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Matheoud D, Sugiura A, Bellemare-Pelletier A, Laplante A, Rondeau C, Chemali M, et al. Parkinson’s disease-related proteins PINK1 and parkin repress mitochondrial antigen presentation. Cell. 2016;166(2):314–27.

    Article  CAS  PubMed  Google Scholar 

  164. König T, Nolte H, Aaltonen MJ, Tatsuta T, Krols M, Stroh T, et al. MIROs and DRP1 drive mitochondrial-derived vesicle biogenesis and promote quality control. Nat Cell Biol. 2021;23(12):1271–86.

    Article  PubMed  Google Scholar 

  165. Peng T, Xie Y, Sheng H, Wang C, Lian Y, Xie N. Mitochondrial-derived vesicles: Gatekeepers of mitochondrial response to oxidative stress. Free Radic Biol Med. 2022;188:185–93.

    Article  CAS  PubMed  Google Scholar 

  166. Ding WX, Li M, Biazik JM, Morgan DG, Guo F, Ni HM, et al. Electron microscopic analysis of a spherical mitochondrial structure. J Biol Chem. 2012;287(50):42373–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Ding WX, Guo F, Ni HM, Bockus A, Manley S, Stolz DB, et al. Parkin and mitofusins reciprocally regulate mitophagy and mitochondrial spheroid formation. J Biol Chem. 2012;287(50):42379–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Yin XM, Ding WX. The reciprocal roles of PARK2 and mitofusins in mitophagy and mitochondrial spheroid formation. Autophagy. 2013;9(11):1687–92.

    Article  CAS  PubMed  Google Scholar 

  169. Ni HM, Williams JA, Ding WX. Mitochondrial dynamics and mitochondrial quality control. Redox Biol. 2015;4:6–13.

    Article  CAS  PubMed  Google Scholar 

  170. Ni HM, Williams JA, Jaeschke H, Ding WX. Zonated induction of autophagy and mitochondrial spheroids limits acetaminophen-induced necrosis in the liver. Redox Biol. 2013;1(1):427–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Williams JA, Ding WX. A mechanistic review of mitophagy and its role in protection against alcoholic liver disease. Biomolecules. 2015;5(4):2619–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Kim J, Kundu M, Viollet B, Guan KL. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat Cell Biol. 2011;13(2):132–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Cai C, Guo Z, Chang X, Li Z, Wu F, He J, et al. Empagliflozin attenuates cardiac microvascular ischemia/reperfusion through activating the AMPKα1/ULK1/FUNDC1/mitophagy pathway. Redox Biol. 2022;52:102288.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Murakawa T, Okamoto K, Omiya S, Taneike M, Yamaguchi O, Otsu K. A mammalian mitophagy receptor, Bcl2-L-13, recruits the ULK1 complex to induce mitophagy. Cell Rep. 2019;26(2):338–45.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Jin Z, Chang B, Wei Y, Yang Y, Zhang H, Liu J, et al. Curcumin exerts chondroprotective effects against osteoarthritis by promoting AMPK/PINK1/Parkin-mediated mitophagy. Biomed Pharmacother. 2022;151:113092.

    Article  CAS  PubMed  Google Scholar 

  176. Pei S, Minhajuddin M, Adane B, Khan N, Stevens BM, Mack SC, et al. AMPK/FIS1-mediated mitophagy is required for self-renewal of human AML stem cells. Cell Stem Cell. 2018;23(1):86–100.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Chen Z, Siraj S, Liu L, Chen Q. MARCH5-FUNDC1 axis fine-tunes hypoxia-induced mitophagy. Autophagy. 2017;13(7):1244–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Yao J, Wang J, Xu Y, Guo Q, Sun Y, Liu J, et al. CDK9 inhibition blocks the initiation of PINK1-PRKN-mediated mitophagy by regulating the SIRT1-FOXO3-BNIP3 axis and enhances the therapeutic effects involving mitochondrial dysfunction in hepatocellular carcinoma. Autophagy. 2022;18(8):1879–97.

    Article  CAS  PubMed  Google Scholar 

  179. Zhao N, Xia J, Xu B. Physical exercise may exert its therapeutic influence on Alzheimer’s disease through the reversal of mitochondrial dysfunction via SIRT1-FOXO1/3-PINK1-Parkin-mediated mitophagy. J Sport Health Sci. 2021;10(1):1–3.

    Article  CAS  PubMed  Google Scholar 

  180. Gong Y, Tang N, Liu P, Sun Y, Lu S, Liu W, et al. Newcastle disease virus degrades SIRT3 via PINK1-PRKN-dependent mitophagy to reprogram energy metabolism in infected cells. Autophagy. 2022;18(7):1503–21.

    Article  CAS  PubMed  Google Scholar 

  181. Hu J, Liu T, Fu F, Cui Z, Lai Q, Zhang Y, et al. Omentin1 ameliorates myocardial ischemia-induced heart failure via SIRT3/FOXO3a-dependent mitochondrial dynamical homeostasis and mitophagy. J Transl Med. 2022;20(1):447.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Li R, Xin T, Li D, Wang C, Zhu H, Zhou H. Therapeutic effect of Sirtuin 3 on ameliorating nonalcoholic fatty liver disease: the role of the ERK-CREB pathway and Bnip3-mediated mitophagy. Redox Biol. 2018;18:229–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Polletta L, Vernucci E, Carnevale I, Arcangeli T, Rotili D, Palmerio S, et al. SIRT5 regulation of ammonia-induced autophagy and mitophagy. Autophagy. 2015;11(2):253–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Inigo JR, Chandra D. The mitochondrial unfolded protein response (UPRmt): shielding against toxicity to mitochondria in cancer. J Hematol Oncol. 2022;15(1):98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Jin SM, Youle RJ. The accumulation of misfolded proteins in the mitochondrial matrix is sensed by PINK1 to induce PARK2/Parkin-mediated mitophagy of polarized mitochondria. Autophagy. 2013;9(11):1750–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Harrington JS, Ryter SW, Plataki M, Price DR, Choi AMK. Mitochondria in health, disease, and ageing. Physiol Rev. 2023;103(4):2349–422.

    Article  CAS  PubMed  Google Scholar 

  187. Liesa M, Shirihai OS. Mitochondrial dynamics in the regulation of nutrient utilization and energy expenditure. Cell Metab. 2013;17(4):491–506.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Gomes LC, Di Benedetto G, Scorrano L. During autophagy mitochondria elongate, are spared from degradation and sustain cell viability. Nat Cell Biol. 2011;13(5):589–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Molina AJ, Wikstrom JD, Stiles L, Las G, Mohamed H, Elorza A, et al. Mitochondrial networking protects beta-cells from nutrient-induced apoptosis. Diabetes. 2009;58(10):2303–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Bach D, Pich S, Soriano FX, Vega N, Baumgartner B, Oriola J, et al. Mitofusin-2 determines mitochondrial network architecture and mitochondrial metabolism. A novel regulatory mechanism altered in obesity. J Biol Chem. 2003;278(19):17190–7.

    Article  CAS  PubMed  Google Scholar 

  191. Kelley DE, He J, Menshikova EV, Ritov VB. Dysfunction of mitochondria in human skeletal muscle in type 2 diabetes. Diabetes. 2002;51(10):2944–50.

    Article  CAS  PubMed  Google Scholar 

  192. Bach D, Naon D, Pich S, Soriano FX, Vega N, Rieusset J, et al. Expression of Mfn2, the Charcot-Marie-Tooth neuropathy type 2A gene, in human skeletal muscle: effects of type 2 diabetes, obesity, weight loss, and the regulatory role of tumor necrosis factor alpha and interleukin-6. Diabetes. 2005;54(9):2685–93.

    Article  CAS  PubMed  Google Scholar 

  193. Rambold AS, Kostelecky B, Elia N, Lippincott-Schwartz J. Tubular network formation protects mitochondria from autophagosomal degradation during nutrient starvation. Proc Natl Acad Sci U S A. 2011;108(25):10190–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Xu Z, Fu T, Guo Q, Sun W, Gan Z. Mitochondrial quality orchestrates muscle-adipose dialog to alleviate dietary obesity. Pharmacol Res. 2019;141:176–80.

    Article  CAS  PubMed  Google Scholar 

  195. Cho YK, Son Y, Saha A, Kim D, Choi C, Kim M, et al. STK3/STK4 signalling in adipocytes regulates mitophagy and energy expenditure. Nat Metab. 2021;3(3):428–41.

    Article  CAS  PubMed  Google Scholar 

  196. Ko MS, Yun JY, Baek IJ, Jang JE, Hwang JJ, Lee SE, et al. Mitophagy deficiency increases NLRP3 to induce brown fat dysfunction in mice. Autophagy. 2021;17(5):1205–21.

    Article  CAS  PubMed  Google Scholar 

  197. Zheng L, Shu WJ, Li YM, Mari M, Yan C, Wang D, et al. The Paf1 complex transcriptionally regulates the mitochondrial-anchored protein Atg32 leading to activation of mitophagy. Autophagy. 2020;16(8):1366–79.

    Article  CAS  PubMed  Google Scholar 

  198. Wu H, Wang Y, Li W, Chen H, Du L, Liu D, et al. Deficiency of mitophagy receptor FUNDC1 impairs mitochondrial quality and aggravates dietary-induced obesity and metabolic syndrome. Autophagy. 2019;15(11):1882–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Zhou Y, Long Q, Wu H, Li W, Qi J, Wu Y, et al. Topology-dependent, bifurcated mitochondrial quality control under starvation. Autophagy. 2020;16(3):562–74.

    Article  PubMed  Google Scholar 

  200. Rosina M, Ceci V, Turchi R, Chuan L, Borcherding N, Sciarretta F, et al. Ejection of damaged mitochondria and their removal by macrophages ensure efficient thermogenesis in brown adipose tissue. Cell Metab. 2022;34(4):533–548.e12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Perrone M, Patergnani S, Di Mambro T, Palumbo L, Wieckowski MR, Giorgi C, et al. Calcium homeostasis in the control of mitophagy. Antioxid Redox Signal. 2023;38(7–9):581–98.

    CAS  PubMed  Google Scholar 

  202. Báthori G, Csordás G, Garcia-Perez C, Davies E, Hajnóczky G. Ca2+-dependent control of the permeability properties of the mitochondrial outer membrane and voltage-dependent anion-selective channel (VDAC). J Biol Chem. 2006;281(25):17347–58.

    Article  PubMed  Google Scholar 

  203. Sander P, Gudermann T, Schredelseker J. A Calcium guard in the outer membrane: is VDAC a regulated gatekeeper of mitochondrial calcium uptake? Int J Mol Sci. 2021;22(2):946.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. De Stefani D, Rizzuto R, Pozzan T. Enjoy the trip: calcium in mitochondria back and forth. Annu Rev Biochem. 2016;85:161–92.

    Article  PubMed  Google Scholar 

  205. Marchi S, Bittremieux M, Missiroli S, Morganti C, Patergnani S, Sbano L, et al. Endoplasmic reticulum-mitochondria communication through Ca2+ signaling: the importance of mitochondria-associated membranes (MAMs). Adv Exp Med Biol. 2017;997:49–67.

    Article  CAS  PubMed  Google Scholar 

  206. Godoy JA, Rios JA, Picón-Pagès P, Herrera-Fernández V, Swaby B, Crepin G, et al. Mitostasis, calcium and free radicals in health, aging and neurodegeneration. Biomolecules. 2021;11(7):1012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Bonora M, Giorgi C, Pinton P. Molecular mechanisms and consequences of mitochondrial permeability transition. Nat Rev Mol Cell Biol. 2022;23(4):266–85.

    Article  CAS  PubMed  Google Scholar 

  208. Nesci S. What happens when the mitochondrial H+-translocating F1FO-ATP(hydrol)ase becomes a molecular target of calcium? The pore opens. Biochimie. 2022;198:92–5.

    Article  CAS  PubMed  Google Scholar 

  209. Wilson EL, Metzakopian E. ER-mitochondria contact sites in neurodegeneration: genetic screening approaches to investigate novel disease mechanisms. Cell Death Differ. 2021;28(6):1804–21.

    Article  CAS  PubMed  Google Scholar 

  210. Gao P, Yang W, Sun L. Mitochondria-associated endoplasmic reticulum membranes (MAMs) and their prospective roles in kidney disease. Oxid Med Cell Longev. 2020;2020:3120539.

    Article  PubMed  PubMed Central  Google Scholar 

  211. Bravo-Sagua R, Lopez-Crisosto C, Criollo A, Inagi R, Lavandero S. Organelle communication: joined in sickness and in health. Physiology (Bethesda). 2023;38(3):0. https://doi.org/10.1152/physiol.00024.2022.

  212. Boyman L, Karbowski M, Lederer WJ. Regulation of mitochondrial ATP production: Ca2+ signaling and quality control. Trends Mol Med. 2020;26(1):21–39.

    Article  CAS  PubMed  Google Scholar 

  213. Pinton P, Ferrari D, Rapizzi E, Di Virgilio F, Pozzan T, Rizzuto R. The Ca2+ concentration of the endoplasmic reticulum is a key determinant of ceramide-induced apoptosis: significance for the molecular mechanism of Bcl-2 action. EMBO J. 2001;20(11):2690–701.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Smirnova E, Griparic L, Shurland DL, van der Bliek AM. Dynamin-related protein Drp1 is required for mitochondrial division in mammalian cells. Mol Biol Cell. 2001;12(8):2245–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Cereghetti GM, Stangherlin A, Martins de Brito O, Chang CR, Blackstone C, Bernardi P, et al. Dephosphorylation by calcineurin regulates translocation of Drp1 to mitochondria. Proc Natl Acad Sci U S A. 2008;105(41):15803–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Korobova F, Ramabhadran V, Higgs HN. An actin-dependent step in mitochondrial fission mediated by the ER-associated formin INF2. Science. 2013;339(6118):464–7.

    Article  CAS  PubMed  Google Scholar 

  217. Chakrabarti R, Ji WK, Stan RV, de Juan SJ, Ryan TA, Higgs HN. INF2-mediated actin polymerization at the ER stimulates mitochondrial calcium uptake, inner membrane constriction, and division. J Cell Biol. 2018;217(1):251–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Ruiz A, Quintela-López T, Sánchez-Gómez MV, Gaminde-Blasco A, Alberdi E, Matute C. Mitochondrial division inhibitor 1 disrupts oligodendrocyte Ca2+ homeostasis and mitochondrial function. Glia. 2020;68(9):1743–56.

    Article  PubMed  Google Scholar 

  219. Favaro G, Romanello V, Varanita T, Andrea Desbats M, Morbidoni V, Tezze C, et al. DRP1-mediated mitochondrial shape controls calcium homeostasis and muscle mass. Nat Commun. 2019;10(1):2576.

    Article  PubMed  PubMed Central  Google Scholar 

  220. Zhang T, Liu Q, Gao W, Sehgal SA, Wu H. The multifaceted regulation of mitophagy by endogenous metabolites. Autophagy. 2022;18(6):1216–39.

    Article  CAS  PubMed  Google Scholar 

  221. Wang X, Schwarz TL. The mechanism of Ca2+-dependent regulation of kinesin-mediated mitochondrial motility. Cell. 2009;136(1):163–74.

  222. MacVicar TD, Mannack LV, Lees RM, Lane JD. Targeted siRNA screens identify ER-to-mitochondrial calcium exchange in autophagy and mitophagy responses in RPE1 cells. Int J Mol Sci. 2015;16(6):13356–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Puri R, Cheng XT, Lin MY, Huang N, Sheng ZH. Mul1 restrains Parkin-mediated mitophagy in mature neurons by maintaining ER-mitochondrial contacts. Nat Commun. 2019;10(1):3645.

    Article  PubMed  PubMed Central  Google Scholar 

  224. Zhou H, Dai Z, Li J, Wang J, Zhu H, Chang X, et al. TMBIM6 prevents VDAC1 multimerization and improves mitochondrial quality control to reduce sepsis-related myocardial injury. Metabolism. 2023;140:155383.

    Article  CAS  PubMed  Google Scholar 

  225. Dan X, Babbar M, Moore A, Wechter N, Tian J, Mohanty JG, et al. DNA damage invokes mitophagy through a pathway involving Spata18. Nucleic Acids Res. 2020;48(12):6611–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Read AD, Bentley RE, Archer SL, Dunham-Snary KJ. Mitochondrial iron-sulfur clusters: structure, function, and an emerging role in vascular biology. Redox Biol. 2021;47:102164.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Nogueira NP, Saraiva FMS, Oliveira MP, Mendonça APM, Inacio JDF, Almeida-Amaral EE, et al. Heme modulates Trypanosoma cruzi bioenergetics inducing mitochondrial ROS production. Free Radic Biol Med. 2017;108:183–91.

    Article  CAS  PubMed  Google Scholar 

  228. Su L, Zhang J, Gomez H, Kellum JA, Peng Z. Mitochondria ROS and mitophagy in acute kidney injury. Autophagy. 2023;19(2):401–14.

    Article  CAS  PubMed  Google Scholar 

  229. Sumegi K, Fekete K, Antus C, Debreceni B, Hocsak E, Gallyas F Jr, et al. BGP-15 protects against oxidative stress- or lipopolysaccharide-induced mitochondrial destabilization and reduces mitochondrial production of reactive oxygen species. PLoS ONE. 2017;12(1):e0169372.

    Article  PubMed  PubMed Central  Google Scholar 

  230. Iwasaki Y, Takeshima Y, Fujio K. Basic mechanism of immune system activation by mitochondria. Immunol Med. 2020;43(4):142–7.

    Article  PubMed  Google Scholar 

  231. Albracht SP, Meijer AJ, Rydström J. Mammalian NADH: ubiquinone oxidoreductase (Complex I) and nicotinamide nucleotide transhydrogenase (Nnt) together regulate the mitochondrial production of H2O2–implications for their role in disease, especially cancer. J Bioenerg Biomembr. 2011;43(5):541–64.

  232. Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87(1):245–313.

    Article  CAS  PubMed  Google Scholar 

  233. Ismail T, Kim Y, Lee H, Lee DS, Lee HS. Interplay between mitochondrial peroxiredoxins and ROS in cancer development and progression. Int J Mol Sci. 2019;20(18):4407.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  234. Suski JM, Lebiedzinska M, Bonora M, Pinton P, Duszynski J, Wieckowski MR. Relation between mitochondrial membrane potential and ROS formation. Methods Mol Biol. 2012;810:183–205.

    Article  CAS  PubMed  Google Scholar 

  235. Evans CS, Holzbaur EL. Degradation of engulfed mitochondria is rate-limiting in Optineurin-mediated mitophagy in neurons. Elife. 2020;9:e50260.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Ma K, Chen G, Li W, Kepp O, Zhu Y, Chen Q. Mitophagy, mitochondrial homeostasis, and cell fate. Front Cell Dev Biol. 2020;8:467.

    Article  PubMed  PubMed Central  Google Scholar 

  237. Fan X, Dong T, Yan K, Ci X, Peng L. PM2.5 increases susceptibility to acute exacerbation of COPD via NOX4/Nrf2 redox imbalance-mediated mitophagy. Redox Biol. 2023;59:102587.

    Article  CAS  PubMed  Google Scholar 

  238. Ning R, Li Y, Du Z, Li T, Sun Q, Lin L, et al. The mitochondria-targeted antioxidant MitoQ attenuated PM(2.5)-induced vascular fibrosis via regulating mitophagy. Redox Biol. 2021;46:102113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  239. Zhou B, Fang L, Dong Y, Yang J, Chen X, Zhang N, et al. Mitochondrial quality control protects photoreceptors against oxidative stress in the H2O2-induced models of retinal degeneration diseases. Cell Death Dis. 2021;12(5):413.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  240. Jiang Y, Krantz S, Qin X, Li S, Gunasekara H, Kim YM, et al. Caveolin-1 controls mitochondrial damage and ROS production by regulating fission - fusion dynamics and mitophagy. Redox Biol. 2022;52:102304.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  241. Lin Q, Li S, Jiang N, Shao X, Zhang M, Jin H, et al. PINK1-parkin pathway of mitophagy protects against contrast-induced acute kidney injury via decreasing mitochondrial ROS and NLRP3 inflammasome activation. Redox Biol. 2019;26:101254.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Liu L, Zhang W, Liu T, Tan Y, Chen C, Zhao J, et al. The physiological metabolite α-ketoglutarate ameliorates osteoarthritis by regulating mitophagy and oxidative stress. Redox Biol. 2023;62:102663.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  243. Li W, Jiang WS, Su YR, Tu KW, Zou L, Liao CR, et al. PINK1/Parkin-mediated mitophagy inhibits osteoblast apoptosis induced by advanced oxidation protein products. Cell Death Dis. 2023;14(2):88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Lu X, Xuan W, Li J, Yao H, Huang C, Li J. AMPK protects against alcohol-induced liver injury through UQCRC2 to up-regulate mitophagy. Autophagy. 2021;17(11):3622–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Franci L, Tubita A, Bertolino FM, Palma A, Cannino G, Settembre C, et al. MAPK15 protects from oxidative stress-dependent cellular senescence by inducing the mitophagic process. Aging Cell. 2022;21(7):e13620.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Esteras N, Abramov AY. Nrf2 as a regulator of mitochondrial function: Energy metabolism and beyond. Free Radic Biol Med. 2022;189:136–53.

    Article  CAS  PubMed  Google Scholar 

  247. Schrader M, Pellegrini L. The making of a mammalian peroxisome, version 2.0: mitochondria get into the mix. Cell Death Differ. 2017;24(7):1148–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Sabouny R, Shutt TE. Reciprocal regulation of mitochondrial fission and fusion. Trends Biochem Sci. 2020;45(7):564–77.

    Article  CAS  PubMed  Google Scholar 

  249. Twig G, Elorza A, Molina AJ, Mohamed H, Wikstrom JD, Walzer G, et al. Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J. 2008;27(2):433–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  250. Ashraf R, Kumar S. Mfn2-mediated mitochondrial fusion promotes autophagy and suppresses ovarian cancer progression by reducing ROS through AMPK/mTOR/ERK signaling. Cell Mol Life Sci. 2022;79(11):573.

    Article  CAS  PubMed  Google Scholar 

  251. Li M, Xu B, Li X, Li Y, Qiu S, Chen K, et al. Mitofusin 2 confers the suppression of microglial activation by cannabidiol: insights from in vitro and in vivo models. Brain Behav Immun. 2022;104:155–70.

    Article  CAS  PubMed  Google Scholar 

  252. Tondera D, Grandemange S, Jourdain A, Karbowski M, Mattenberger Y, Herzig S, et al. SLP-2 is required for stress-induced mitochondrial hyperfusion. EMBO J. 2009;28(11):1589–600.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. Shutt T, Geoffrion M, Milne R, McBride HM. The intracellular redox state is a core determinant of mitochondrial fusion. EMBO Rep. 2012;13(10):909–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  254. Thaher O, Wolf C, Dey PN, Pouya A, Wüllner V, Tenzer S, et al. The thiol switch C684 in Mitofusin-2 mediates redox-induced alterations of mitochondrial shape and respiration. Neurochem Int. 2018;117:167–73.

    Article  CAS  PubMed  Google Scholar 

  255. Mattie S, Riemer J, Wideman JG, McBride HM. A new mitofusin topology places the redox-regulated C terminus in the mitochondrial intermembrane space. J Cell Biol. 2018;217(2):507–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  256. Lloberas J, Muñoz JP, Hernández-Álvarez MI, Cardona PJ, Zorzano A, Celada A. Macrophage mitochondrial MFN2 (mitofusin 2) links immune stress and immune response through reactive oxygen species (ROS) production. Autophagy. 2020;16(12):2307–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  257. Tur J, Pereira-Lopes S, Vico T, Marín EA, Muñoz JP, Hernández-Alvarez M, et al. Mitofusin 2 in macrophages links mitochondrial ROS production, cytokine release, phagocytosis, autophagy, and bactericidal activity. Cell Rep. 2020;32(8):108079.

    Article  CAS  PubMed  Google Scholar 

  258. Sabouny R, Fraunberger E, Geoffrion M, Ng AC, Baird SD, Screaton RA, et al. The Keap1-Nrf2 stress response pathway promotes mitochondrial hyperfusion through degradation of the mitochondrial fission protein Drp1. Antioxid Redox Signal. 2017;27(18):1447–59.

    Article  CAS  PubMed  Google Scholar 

  259. Wang X, Yen J, Kaiser P, Huang L. Regulation of the 26S proteasome complex during oxidative stress. Sci Signal. 2010;3(151):ra88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  260. Dalton S. Linking the cell cycle to cell fate decisions. Trends Cell Biol. 2015;25(10):592–600.

    Article  PubMed  PubMed Central  Google Scholar 

  261. Lopez-Mejia IC, Fajas L. Cell cycle regulation of mitochondrial function. Curr Opin Cell Biol. 2015;33:19–25.

    Article  CAS  PubMed  Google Scholar 

  262. Spurlock B, Tullet J, Hartman JL 4th, Mitra K. Interplay of mitochondrial fission-fusion with cell cycle regulation: possible impacts on stem cell and organismal aging. Exp Gerontol. 2020;135:110919.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  263. Harbauer AB, Opalińska M, Gerbeth C, Herman JS, Rao S, Schönfisch B, et al. Mitochondria. Cell cycle-dependent regulation of mitochondrial preprotein translocase. Science. 2014;346(6213):1109–13.

    Article  CAS  PubMed  Google Scholar 

  264. Kashatus DF, Lim KH, Brady DC, Pershing NLK, Cox AD, Counter CM. RALA and RALBP1 regulate mitochondrial fission at mitosis. Nat Cell Biol. 2011;13(9):1108–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  265. Mitra K, Wunder C, Roysam B, Lin G, Lippincott-Schwartz J. A hyperfused mitochondrial state achieved at G1-S regulates cyclin E buildup and entry into S phase. Proc Natl Acad Sci U S A. 2009;106(29):11960–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  266. Qian W, Choi S, Gibson GA, Watkins SC, Bakkenist CJ, Van Houten B. Mitochondrial hyperfusion induced by loss of the fission protein Drp1 causes ATM-dependent G2/M arrest and aneuploidy through DNA replication stress. J Cell Sci. 2012;125(Pt 23):5745–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  267. Sarraf SA, Sideris DP, Giagtzoglou N, Ni L, Kankel MW, Sen A, et al. PINK1/parkin influences cell cycle by sequestering TBK1 at damaged mitochondria, inhibiting mitosis. Cell Rep. 2019;29(1):225–35.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  268. Liu J, Hong M, Li Y, Chen D, Wu Y, Hu Y. Programmed cell death tunes tumor immunity. Front Immunol. 2022;13:847345.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  269. Li M, Wang ZW, Fang LJ, Cheng SQ, Wang X, Liu NF. Programmed cell death in atherosclerosis and vascular calcification. Cell Death Dis. 2022;13(5):467.

    Article  PubMed  PubMed Central  Google Scholar 

  270. Xu X, Lai Y, Hua ZC. Apoptosis and apoptotic body: disease message and therapeutic target potentials. Biosci Rep. 2019;39(1):BSR20180992.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  271. Li P, Dong XR, Zhang B, Zhang XT, Liu JZ, Ma DS, et al. Molecular mechanism and therapeutic targeting of necrosis, apoptosis, pyroptosis, and autophagy in cardiovascular disease. Chin Med J (Engl). 2021;134(22):2647–55.

    Article  CAS  PubMed  Google Scholar 

  272. Bedoui S, Herold MJ, Strasser A. Emerging connectivity of programmed cell death pathways and its physiological implications. Nat Rev Mol Cell Biol. 2020;21(11):678–95.

    Article  CAS  PubMed  Google Scholar 

  273. Tang D, Kang R, Berghe TV, Vandenabeele P, Kroemer G. The molecular machinery of regulated cell death. Cell Res. 2019;29(5):347–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  274. Suen DF, Norris KL, Youle RJ. Mitochondrial dynamics and apoptosis. Genes Dev. 2008;22(12):1577–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  275. Frank S, Gaume B, Bergmann-Leitner ES, Leitner WW, Robert EG, Catez F, et al. The role of dynamin-related protein 1, a mediator of mitochondrial fission, in apoptosis. Dev Cell. 2001;1(4):515–25.

    Article  CAS  PubMed  Google Scholar 

  276. Wasiak S, Zunino R, McBride HM. Bax/Bak promote sumoylation of DRP1 and its stable association with mitochondria during apoptotic cell death. J Cell Biol. 2007;177(3):439–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  277. Ban-Ishihara R, Ishihara T, Sasaki N, Mihara K, Ishihara N. Dynamics of nucleoid structure regulated by mitochondrial fission contributes to cristae reformation and release of cytochrome C. Proc Natl Acad Sci U S A. 2013;110(29):11863–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  278. Lee YJ, Jeong SY, Karbowski M, Smith CL, Youle RJ. Roles of the mammalian mitochondrial fission and fusion mediators Fis1, Drp1, and Opa1 in apoptosis. Mol Biol Cell. 2004;15(11):5001–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  279. Olichon A, Baricault L, Gas N, Guillou E, Valette A, Belenguer P, et al. Loss of OPA1 perturbates the mitochondrial inner membrane structure and integrity, leading to cytochrome C release and apoptosis. J Biol Chem. 2003;278(10):7743–6.

    Article  CAS  PubMed  Google Scholar 

  280. Karbowski M, Norris KL, Cleland MM, Jeong SY, Youle RJ. Role of Bax and Bak in mitochondrial morphogenesis. Nature. 2006;443(7112):658–62.

    Article  CAS  PubMed  Google Scholar 

  281. Brooks C, Wei Q, Feng L, Dong G, Tao Y, Mei L, et al. Bak regulates mitochondrial morphology and pathology during apoptosis by interacting with mitofusins. Proc Natl Acad Sci U S A. 2007;104(28):11649–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  282. Ma K, Zhang Z, Chang R, Cheng H, Mu C, Zhao T, et al. Dynamic PGAM5 multimers dephosphorylate BCL-xL or FUNDC1 to regulate mitochondrial and cellular fate. Cell Death Differ. 2020;27(3):1036–51.

    Article  CAS  PubMed  Google Scholar 

  283. Quarato G, Mari L, Barrows NJ, Yang M, Ruehl S, Chen MJ, et al. Mitophagy restricts BAX/BAK-independent, Parkin-mediated apoptosis. Sci Adv. 2023;9(21):eadg8156.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  284. Li N, Xiong R, He R, Liu B, Wang B, Geng Q. Mangiferin mitigates lipopolysaccharide-induced lung injury by inhibiting NLRP3 inflammasome activation. J Inflamm Res. 2021;14:2289–300.

    Article  PubMed  PubMed Central  Google Scholar 

  285. Zeng C, Duan F, Hu J, Luo B, Huang B, Lou X, et al. NLRP3 inflammasome-mediated pyroptosis contributes to the pathogenesis of non-ischemic dilated cardiomyopathy. Redox Biol. 2020;34:101523.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  286. Wang Y, Xiong L, Yao Y, Ma Y, Liu Q, Pang Y, et al. The involvement of DRP1-mediated caspase-1 activation in inflammatory response by urban particulate matter in EA.hy926 human vascular endothelial cells. Environ Pollut. 2021;287:117369.

    Article  CAS  PubMed  Google Scholar 

  287. Liu M, Lu J, Yang S, Chen Y, Yu J, Guan S. Alliin alleviates LPS-induced pyroptosis via promoting mitophagy in THP-1 macrophages and mice. Food Chem Toxicol. 2022;160:112811.

    Article  CAS  PubMed  Google Scholar 

  288. Liu Z, Wang M, Wang X, Bu Q, Wang Q, Su W, et al. XBP1 deficiency promotes hepatocyte pyroptosis by impairing mitophagy to activate mtDNA-cGAS-STING signaling in macrophages during acute liver injury. Redox Biol. 2022;52:102305.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  289. Yuk JM, Silwal P, Jo EK. Inflammasome and mitophagy connection in health and disease. Int J Mol Sci. 2020;21(13):4714.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  290. Jiang X, Stockwell BR, Conrad M. Ferroptosis: mechanisms, biology and role in disease. Nat Rev Mol Cell Biol. 2021;22(4):266–82.

    Article  PubMed  PubMed Central  Google Scholar 

  291. Stockwell BR, Jiang X, Gu W. Emerging mechanisms and disease relevance of ferroptosis. Trends Cell Biol. 2020;30(6):478–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  292. Li C, Liu J, Hou W, Kang R, Tang D. STING1 promotes ferroptosis through MFN1/2-dependent mitochondrial fusion. Front Cell Dev Biol. 2021;9:698679.

    Article  PubMed  PubMed Central  Google Scholar 

  293. Lin Q, Li S, Jin H, Cai H, Zhu X, Yang Y, et al. Mitophagy alleviates cisplatin-induced renal tubular epithelial cell ferroptosis through ROS/HO-1/GPX4 axis. Int J Biol Sci. 2023;19(4):1192–210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  294. Yu F, Zhang Q, Liu H, Liu J, Yang S, Luo X, et al. Dynamic O-GlcNAcylation coordinates ferritinophagy and mitophagy to activate ferroptosis. Cell Discov. 2022;8(1):40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  295. Fontana F, Limonta P. The multifaceted roles of mitochondria at the crossroads of cell life and death in cancer. Free Radic Biol Med. 2021;176:203–21.

    Article  CAS  PubMed  Google Scholar 

  296. LeBleu VS, O’Connell JT, Gonzalez Herrera KN, Wikman H, Pantel K, Haigis MC, et al. PGC-1α mediates mitochondrial biogenesis and oxidative phosphorylation in cancer cells to promote metastasis. Nat Cell Biol. 2014;16(10):992–1003.1–15.

  297. Liu Y, Jin M, Wang Y, Zhu J, Tan R, Zhao J, et al. MCU-induced mitochondrial calcium uptake promotes mitochondrial biogenesis and colorectal cancer growth. Signal Transduct Target Ther. 2020;5(1):59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  298. Torrens-Mas M, Hernández-López R, Pons DG, Roca P, Oliver J, Sastre-Serra J. Sirtuin 3 silencing impairs mitochondrial biogenesis and metabolism in colon cancer cells. Am J Physiol Cell Physiol. 2019;317(2):C398–404.

    Article  CAS  PubMed  Google Scholar 

  299. Huangyang P, Li F, Lee P, Nissim I, Weljie AM, Mancuso A, et al. Fructose-1,6-bisphosphatase 2 inhibits sarcoma progression by restraining mitochondrial biogenesis. Cell Metab. 2020;31(1):174–88.e7.

    Article  CAS  PubMed  Google Scholar 

  300. Xu J, Ji L, Ruan Y, Wan Z, Lin Z, Xia S, et al. UBQLN1 mediates sorafenib resistance through regulating mitochondrial biogenesis and ROS homeostasis by targeting PGC1β in hepatocellular carcinoma. Signal Transduct Target Ther. 2021;6(1):190.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  301. Praharaj PP, Panigrahi DP, Bhol CS, Patra S, Mishra SR, Mahapatra KK, et al. Mitochondrial rewiring through mitophagy and mitochondrial biogenesis in cancer stem cells: a potential target for anti-CSC cancer therapy. Cancer Lett. 2021;498:217–28.

    Article  CAS  PubMed  Google Scholar 

  302. Rehman J, Zhang HJ, Toth PT, Zhang Y, Marsboom G, Hong Z, et al. Inhibition of mitochondrial fission prevents cell cycle progression in lung cancer. FASEB J. 2012;26(5):2175–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  303. Xiong X, Hasani S, Young LEA, Rivas DR, Skaggs AT, Martinez R, et al. Activation of Drp1 promotes fatty acids-induced metabolic reprograming to potentiate Wnt signaling in colon cancer. Cell Death Differ. 2022;29(10):1913–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  304. Kannan A, Wells RB, Sivakumar S, Komatsu S, Singh KP, Samten B, et al. Mitochondrial reprogramming regulates breast cancer progression. Clin Cancer Res. 2016;22(13):3348–60.

    Article  CAS  PubMed  Google Scholar 

  305. Hagenbuchner J, Kuznetsov AV, Obexer P, Ausserlechner MJ. BIRC5/Survivin enhances aerobic glycolysis and drug resistance by altered regulation of the mitochondrial fusion/fission machinery. Oncogene. 2013;32(40):4748–57.

    Article  CAS  PubMed  Google Scholar 

  306. Serasinghe MN, Wieder SY, Renault TT, Elkholi R, Asciolla JJ, Yao JL, et al. Mitochondrial division is requisite to RAS-induced transformation and targeted by oncogenic MAPK pathway inhibitors. Mol Cell. 2015;57(3):521–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  307. Gao T, Zhang X, Zhao J, Zhou F, Wang Y, Zhao Z, et al. SIK2 promotes reprogramming of glucose metabolism through PI3K/AKT/HIF-1α pathway and Drp1-mediated mitochondrial fission in ovarian cancer. Cancer Lett. 2020;469:89–101.

    Article  CAS  PubMed  Google Scholar 

  308. Lee YG, Nam Y, Shin KJ, Yoon S, Park WS, Joung JY, et al. Androgen-induced expression of DRP1 regulates mitochondrial metabolic reprogramming in prostate cancer. Cancer Lett. 2020;471:72–87.

    Article  CAS  PubMed  Google Scholar 

  309. Nagdas S, Kashatus JA, Nascimento A, Hussain SS, Trainor RE, Pollock SR, et al. Drp1 promotes KRas-driven metabolic changes to drive pancreatic tumor growth. Cell Rep. 2019;28(7):1845–59.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  310. Kashatus JA, Nascimento A, Myers LJ, Sher A, Byrne FL, Hoehn KL, et al. Erk2 phosphorylation of Drp1 promotes mitochondrial fission and MAPK-driven tumor growth. Mol Cell. 2015;57(3):537–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  311. Wieder SY, Serasinghe MN, Sung JC, Choi DC, Birge MB, Yao JL, et al. Activation of the mitochondrial fragmentation protein DRP1 correlates with BRAF(V600E) melanoma. J Invest Dermatol. 2015;135(10):2544–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  312. Adebayo M, Singh S, Singh AP, Dasgupta S. Mitochondrial fusion and fission: the fine-tune balance for cellular homeostasis. FASEB J. 2021;35(6):e21620.

    Article  CAS  PubMed  Google Scholar 

  313. Rodrigues T, Ferraz LS. Therapeutic potential of targeting mitochondrial dynamics in cancer. Biochem Pharmacol. 2020;182:114282.

    Article  CAS  PubMed  Google Scholar 

  314. Qian W, Wang J, Van Houten B. The role of dynamin-related protein 1 in cancer growth: a promising therapeutic target?. Expert Opin Ther Targets. 2013;17(9):997–1001.

  315. Zhao J, Zhang J, Yu M, Xie Y, Huang Y, Wolff DW, et al. Mitochondrial dynamics regulates migration and invasion of breast cancer cells. Oncogene. 2013;32(40):4814–24.

    Article  CAS  PubMed  Google Scholar 

  316. Yu M, Nguyen ND, Huang Y, Lin D, Fujimoto TN, Molkentine JM, et al. Mitochondrial fusion exploits a therapeutic vulnerability of pancreatic cancer. JCI Insight. 2019;5(16):e126915.

    Article  PubMed  Google Scholar 

  317. Li M, Wang L, Wang Y, Zhang S, Zhou G, Lieshout R, et al. Mitochondrial fusion via OPA1 and MFN1 supports liver tumor cell metabolism and growth. Cells. 2020;9(1):121.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  318. Humphries BA, Cutter AC, Buschhaus JM, Chen YC, Qyli T, Palagama DSW, et al. Enhanced mitochondrial fission suppresses signaling and metastasis in triple-negative breast cancer. Breast Cancer Res. 2020;22(1):60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  319. Herkenne S, Ek O, Zamberlan M, Pellattiero A, Chergova M, Chivite I, et al. Developmental and tumor angiogenesis requires the mitochondria-shaping protein Opa1. Cell Metab. 2020;31(5):987–1003.e8.

    Article  CAS  PubMed  Google Scholar 

  320. Huang RX, Zhou PK. DNA damage response signaling pathways and targets for radiotherapy sensitization in cancer. Signal Transduct Target Ther. 2020;5(1):60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  321. Sessions DT, Kashatus DF. Mitochondrial dynamics in cancer stem cells. Cell Mol Life Sci. 2021;78(8):3803–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  322. Panigrahi DP, Praharaj PP, Bhol CS, Mahapatra KK, Patra S, Behera BP, et al. The emerging, multifaceted role of mitophagy in cancer and cancer therapeutics. Semin Cancer Biol. 2020;66:45–58.

    Article  CAS  PubMed  Google Scholar 

  323. Drake LE, Springer MZ, Poole LP, Kim CJ, Macleod KF. Expanding perspectives on the significance of mitophagy in cancer. Semin Cancer Biol. 2017;47:110–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  324. Liberti MV, Locasale JW. The Warburg effect: how does it benefit cancer cells? Trends Biochem Sci. 2016;41(3):211–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  325. Jiao L, Zhang HL, Li DD, Yang KL, Tang J, Li X, et al. Regulation of glycolytic metabolism by autophagy in liver cancer involves selective autophagic degradation of HK2 (hexokinase 2). Autophagy. 2018;14(4):671–84.

    Article  CAS  PubMed  Google Scholar 

  326. Chang HW, Kim MR, Lee HJ, Lee HM, Kim GC, Lee YS, et al. p53/BNIP3-dependent mitophagy limits glycolytic shift in radioresistant cancer. Oncogene. 2019;38(19):3729–42.

    Article  CAS  PubMed  Google Scholar 

  327. Whelan KA, Chandramouleeswaran PM, Tanaka K, Natsuizaka M, Guha M, Srinivasan S, et al. Autophagy supports generation of cells with high CD44 expression via modulation of oxidative stress and Parkin-mediated mitochondrial clearance. Oncogene. 2017;36(34):4843–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  328. Jung J, Zhang Y, Celiku O, Zhang W, Song H, Williams BJ, et al. Mitochondrial NIX promotes tumor survival in the hypoxic niche of glioblastoma. Can Res. 2019;79(20):5218–32.

    Article  CAS  Google Scholar 

  329. Liu K, Lee J, Kim JY, Wang L, Tian Y, Chan ST, et al. Mitophagy controls the activities of tumor suppressor p53 to regulate hepatic cancer stem cells. Mol Cell. 2017;68(2):281–92.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  330. Bai R, Cui J. Mitochondrial immune regulation and anti-tumor immunotherapy strategies targeting mitochondria. Cancer Lett. 2023;564:216223.

    Article  CAS  PubMed  Google Scholar 

  331. Kao KC, Vilbois S, Tsai CH, Ho PC. Metabolic communication in the tumour-immune microenvironment. Nat Cell Biol. 2022;24(11):1574–83.

    Article  CAS  PubMed  Google Scholar 

  332. Zhang L, Romero P. Metabolic control of CD8+ T cell fate decisions and antitumor immunity. Trends Mol Med. 2018;24(1):30–48.

    Article  PubMed  Google Scholar 

  333. Zhong X, Wu H, Ouyang C, Zhang W, Shi Y, Wang YC, et al. Ncoa2 promotes CD8+ T cell-mediated anti-tumor immunity by stimulating T-cell activation via upregulation of PGC-1α critical for mitochondrial function. Cancer Immunol Res. 2023;11(10):1414–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  334. Li W, Zhang L. Rewiring mitochondrial metabolism for CD8+ T cell memory formation and effective cancer immunotherapy. Front Immunol. 2020;11:1834.

  335. Thommen DS, Schumacher TN. T cell dysfunction in cancer. Cancer Cell. 2018;33(4):547–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  336. Scharping NE, Menk AV, Moreci RS, Whetstone RD, Dadey RE, Watkins SC, et al. The tumor microenvironment represses T cell mitochondrial biogenesis to drive intratumoral T cell metabolic insufficiency and dysfunction. Immunity. 2016;45(2):374–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  337. Malinee M, Pandian GN, Sugiyama H. Targeted epigenetic induction of mitochondrial biogenesis enhances antitumor immunity in mouse model. Cell Chem Biol. 2022;29(3):463–75.e6.

    Article  CAS  PubMed  Google Scholar 

  338. Dumauthioz N, Tschumi B, Wenes M, Marti B, Wang H, Franco F, et al. Enforced PGC-1α expression promotes CD8 T cell fitness, memory formation and antitumor immunity. Cell Mol Immunol. 2021;18(7):1761–71.

    Article  CAS  PubMed  Google Scholar 

  339. Buck MD, O’Sullivan D, Klein Geltink RI, Curtis JD, Chang CH, Sanin DE, et al. Mitochondrial dynamics controls T cell fate through metabolic programming. Cell. 2016;166(1):63–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  340. Simula L, Pacella I, Colamatteo A, Procaccini C, Cancila V, Bordi M, et al. Drp1 controls effective T cell immune-surveillance by regulating T cell migration, proliferation, and cMyc-dependent metabolic reprogramming. Cell Rep. 2018;25(11):3059–73.e10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  341. Bird L. T cells: mitochondrial shape shifters. Nat Rev Immunol. 2016;16(7):402–3.

    Article  CAS  PubMed  Google Scholar 

  342. Simula L, Antonucci Y, Scarpelli G, Cancila V, Colamatteo A, Manni S, et al. PD-1-induced T cell exhaustion is controlled by a Drp1-dependent mechanism. Mol Oncol. 2022;16(1):188–205.

    Article  CAS  PubMed  Google Scholar 

  343. Yu YR, Imrichova H, Wang H, Chao T, Xiao Z, Gao M, et al. Disturbed mitochondrial dynamics in CD8+ TILs reinforce T cell exhaustion. Nat Immunol. 2020;21(12):1540–51.

    Article  CAS  PubMed  Google Scholar 

  344. Chakraborty P, Parikh RY, Choi S, Tran D, Gooz M, Hedley ZT, et al. Carbon Monoxide activates PERK-regulated autophagy to induce immunometabolic reprogramming and boost antitumor T-cell function. Can Res. 2022;82(10):1969–90.

    Article  CAS  Google Scholar 

  345. Denk D, Petrocelli V, Conche C, Drachsler M, Ziegler PK, Braun A, et al. Expansion of T memory stem cells with superior anti-tumor immunity by Urolithin A-induced mitophagy. Immunity. 2022;55(11):2059–73.e8.

    Article  CAS  PubMed  Google Scholar 

  346. Wu SY, Fu T, Jiang YZ, Shao ZM. Natural killer cells in cancer biology and therapy. Mol Cancer. 2020;19(1):120.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  347. Zheng X, Qian Y, Fu B, Jiao D, Jiang Y, Chen P, et al. Mitochondrial fragmentation limits NK cell-based tumor immunosurveillance. Nat Immunol. 2019;20(12):1656–67.

    Article  CAS  PubMed  Google Scholar 

  348. Chen S, Saeed AFUH, Liu Q, Jiang Q, Xu H, Xiao GG, et al. Macrophages in immunoregulation and therapeutics. Signal Transduct Target Ther. 2023;8(1):207.

    Article  PubMed  PubMed Central  Google Scholar 

  349. Bao D, Zhao J, Zhou X, Yang Q, Chen Y, Zhu J, et al. Mitochondrial fission-induced mtDNA stress promotes tumor-associated macrophage infiltration and HCC progression. Oncogene. 2019;38(25):5007–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  350. Xia H, Li S, Li X, Wang W, Bian Y, Wei S, et al. Autophagic adaptation to oxidative stress alters peritoneal residential macrophage survival and ovarian cancer metastasis. JCI insight. 2020;5(18):e141115.

    Article  PubMed  PubMed Central  Google Scholar 

  351. Xu H, Li D, Ma J, Zhao Y, Xu L, Tian R, et al. The IL-33/ST2 axis affects tumor growth by regulating mitophagy in macrophages and reprogramming their polarization. Cancer Biol Med. 2021;18(1):172–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  352. Nguyen BY, Ruiz-Velasco A, Bui T, Collins L, Wang X, Liu W. Mitochondrial function in the heart: the insight into mechanisms and therapeutic potentials. Br J Pharmacol. 2019;176(22):4302–18.

    Article  CAS  PubMed  Google Scholar 

  353. Singh H. Mitochondrial ion channels in cardiac function. Am J Physiol Cell Physiol. 2021;321(5):C812–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  354. Savarese G, Becher PM, Lund LH, Seferovic P, Rosano GMC, Coats AJS. Global burden of heart failure: a comprehensive and updated review of epidemiology. Cardiovasc Res. 2023;118(17):3272–87.

    Article  PubMed  Google Scholar 

  355. Sihag S, Cresci S, Li AY, Sucharov CC, Lehman JJ. PGC-1alpha and ERRalpha target gene downregulation is a signature of the failing human heart. J Mol Cell Cardiol. 2009;46(2):201–12.

    Article  CAS  PubMed  Google Scholar 

  356. Garnier A, Zoll J, Fortin D, N’Guessan B, Lefebvre F, Geny B, et al. Control by circulating factors of mitochondrial function and transcription cascade in heart failure: a role for endothelin-1 and angiotensin II. Circ Heart Fail. 2009;2(4):342–50.

    Article  CAS  PubMed  Google Scholar 

  357. Bhat S, Chin A, Shirakabe A, Ikeda Y, Ikeda S, Zhai P, et al. Recruitment of RNA polymerase II to metabolic gene promoters is inhibited in the failing heart possibly through PGC-1α (peroxisome proliferator-activated receptor-γ coactivator-1α) dysregulation. Circ Heart Fail. 2019;12(3):e005529.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  358. Kärkkäinen O, Tuomainen T, Mutikainen M, Lehtonen M, Ruas JL, Hanhineva K, et al. Heart specific PGC-1α deletion identifies metabolome of cardiac restricted metabolic heart failure. Cardiovasc Res. 2019;115(1):107–18.

    Article  PubMed  Google Scholar 

  359. Arany Z, He H, Lin J, Hoyer K, Handschin C, Toka O, et al. Transcriptional coactivator PGC-1 alpha controls the energy state and contractile function of cardiac muscle. Cell Metab. 2005;1(4):259–71.

    Article  CAS  PubMed  Google Scholar 

  360. Hu X, Xu X, Lu Z, Zhang P, Fassett J, Zhang Y, et al. AMP activated protein kinase-α2 regulates expression of estrogen-related receptor-α, a metabolic transcription factor related to heart failure development. Hypertension. 2011;58(4):696–703.

    Article  CAS  PubMed  Google Scholar 

  361. Lehman JJ, Barger PM, Kovacs A, Saffitz JE, Medeiros DM, Kelly DP. Peroxisome proliferator-activated receptor gamma coactivator-1 promotes cardiac mitochondrial biogenesis. J Clin Invest. 2000;106(7):847–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  362. Vásquez-Trincado C, García-Carvajal I, Pennanen C, Parra V, Hill JA, Rothermel BA, et al. Mitochondrial dynamics, mitophagy and cardiovascular disease. J Physiol. 2016;594(3):509–25.

    Article  PubMed  PubMed Central  Google Scholar 

  363. Eisner V, Cupo RR, Gao E, Csordás G, Slovinsky WS, Paillard M, et al. Mitochondrial fusion dynamics is robust in the heart and depends on calcium oscillations and contractile activity. Proc Natl Acad Sci U S A. 2017;114(5):E859–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  364. Chen L, Liu T, Tran A, Lu X, Tomilov AA, Davies V, et al. OPA1 mutation and late-onset cardiomyopathy: mitochondrial dysfunction and mtDNA instability. J Am Heart Assoc. 2012;1(5):e003012.

    Article  PubMed  PubMed Central  Google Scholar 

  365. Piquereau J, Caffin F, Novotova M, Prola A, Garnier A, Mateo P, et al. Down-regulation of OPA1 alters mouse mitochondrial morphology, PTP function, and cardiac adaptation to pressure overload. Cardiovasc Res. 2012;94(3):408–17.

    Article  CAS  PubMed  Google Scholar 

  366. Chen L, Gong Q, Stice JP, Knowlton AA. Mitochondrial OPA1, apoptosis, and heart failure. Cardiovasc Res. 2009;84(1):91–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  367. Benigni A, Cassis P, Conti S, Perico L, Corna D, Cerullo D, et al. Sirt3 deficiency shortens life span and impairs cardiac mitochondrial function rescued by Opa1 gene transfer. Antioxid Redox Signal. 2019;31(17):1255–71.

    Article  CAS  PubMed  Google Scholar 

  368. Kim S, Song J, Ernst P, Latimer MN, Ha CM, Goh KY, et al. MitoQ regulates redox-related noncoding RNAs to preserve mitochondrial network integrity in pressure-overload heart failure. Am J Physiol Heart Circ Physiol. 2020;318(3):H682–95.

    Article  PubMed  PubMed Central  Google Scholar 

  369. Papanicolaou KN, Kikuchi R, Ngoh GA, Coughlan KA, Dominguez I, Stanley WC, et al. Mitofusins 1 and 2 are essential for postnatal metabolic remodeling in heart. Circ Res. 2012;111(8):1012–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  370. Yue P, Zhang Y, Liu L, Zhou K, Xia S, Peng M, et al. Yap1 modulates cardiomyocyte hypertrophy via impaired mitochondrial biogenesis in response to chronic mechanical stress overload. Theranostics. 2022;12(16):7009–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  371. Chen L, Liu B, Qin Y, Li A, Gao M, Liu H, et al. Mitochondrial fusion protein Mfn2 and its role in heart failure. Front Mol Biosci. 2021;8:681237.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  372. Papanicolaou KN, Ngoh GA, Dabkowski ER, O’Connell KA, Ribeiro RF Jr, Stanley WC, et al. Cardiomyocyte deletion of mitofusin-1 leads to mitochondrial fragmentation and improves tolerance to ROS-induced mitochondrial dysfunction and cell death. Am J Physiol Heart Circ Physiol. 2012;302(1):H167–79.

    Article  CAS  PubMed  Google Scholar 

  373. Liu X, Guo C, Zhang Q. Novel insights into the involvement of mitochondrial fission/fusion in heart failure: from molecular mechanisms to targeted therapies. Cell Stress Chaperones. 2023;28(2):133–44.

    Article  PubMed  PubMed Central  Google Scholar 

  374. Chaanine AH, Joyce LD, Stulak JM, Maltais S, Joyce DL, Dearani JA, et al. Mitochondrial morphology, dynamics, and function in human pressure overload or ischemic heart disease with preserved or reduced ejection fraction. Circ Heart Fail. 2019;12(2):e005131.

    Article  CAS  PubMed  Google Scholar 

  375. Sabbah HN, Gupta RC, Singh-Gupta V, Zhang K, Lanfear DE. Abnormalities of mitochondrial dynamics in the failing heart: normalization following long-term therapy with elamipretide. Cardiovasc Drugs Ther. 2018;32(4):319–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  376. Chang YW, Chang YT, Wang Q, Lin JJ, Chen YJ, Chen CC. Quantitative phosphoproteomic study of pressure-overloaded mouse heart reveals dynamin-related protein 1 as a modulator of cardiac hypertrophy. Mol Cell Proteomics. 2013;12(11):3094–107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  377. Pennanen C, Parra V, López-Crisosto C, Morales PE, Del Campo A, Gutierrez T, et al. Mitochondrial fission is required for cardiomyocyte hypertrophy mediated by a Ca2+-calcineurin signaling pathway. J Cell Sci. 2014;127(Pt 12):2659–71.

    CAS  PubMed  PubMed Central  Google Scholar 

  378. Xu S, Wang P, Zhang H, Gong G, Gutierrez Cortes N, Zhu W, et al. CaMKII induces permeability transition through Drp1 phosphorylation during chronic β-AR stimulation. Nat Commun. 2016;7:13189.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  379. Lyu Y, Huo J, Jiang W, Yang W, Wang S, Zhang S, et al. Empagliflozin ameliorates cardiac dysfunction in heart failure mice via regulating mitochondrial dynamics. Eur J Pharmacol. 2023;942:175531.

    Article  CAS  PubMed  Google Scholar 

  380. Ikeda Y, Shirakabe A, Maejima Y, Zhai P, Sciarretta S, Toli J, et al. Endogenous Drp1 mediates mitochondrial autophagy and protects the heart against energy stress. Circ Res. 2015;116(2):264–78.

    Article  CAS  PubMed  Google Scholar 

  381. Dorn GW 2nd. Gone fission…: diverse consequences of cardiac Drp1 deficiency. Circ Res. 2015;116(2):225–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  382. Donnarumma E, Kohlhaas M, Vimont E, Kornobis E, Chaze T, Gianetto QG, et al. Mitochondrial fission process 1 controls inner membrane integrity and protects against heart failure. Nat Commun. 2022;13(1):6634.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  383. Wang B, Nie J, Wu L, Hu Y, Wen Z, Dong L, et al. AMPKα2 protects against the development of heart failure by enhancing mitophagy via PINK1 phosphorylation. Circ Res. 2018;122(5):712–29.

    Article  CAS  PubMed  Google Scholar 

  384. Abudureyimu M, Yu W, Cao RY, Zhang Y, Liu H, Zheng H. Berberine promotes cardiac function by upregulating PINK1/parkin-mediated mitophagy in heart failure. Front Physiol. 2020;11:565751.

    Article  PubMed  PubMed Central  Google Scholar 

  385. Billia F, Hauck L, Konecny F, Rao V, Shen J, Mak TW. PTEN-inducible kinase 1 (PINK1)/Park6 is indispensable for normal heart function. Proc Natl Acad Sci U S A. 2011;108(23):9572–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  386. Nah J, Shirakabe A, Mukai R, Zhai P, Sung EA, Ivessa A, et al. Ulk1-dependent alternative mitophagy plays a protective role during pressure overload in the heart. Cardiovasc Res. 2022;118(12):2638–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  387. Guan Z, Chen J, Wang L, Hao M, Dong X, Luo T, et al. Nuanxinkang prevents the development of myocardial infarction-induced chronic heart failure by promoting PINK1/Parkin-mediated mitophagy. Phytomedicine. 2023;108:154494.

    Article  PubMed  Google Scholar 

  388. Chaanine AH, Jeong D, Liang L, Chemaly ER, Fish K, Gordon RE, et al. JNK modulates FOXO3a for the expression of the mitochondrial death and mitophagy marker BNIP3 in pathological hypertrophy and in heart failure. Cell Death Dis. 2012;3(2):265.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  389. Ramachandra CJA, Hernandez-Resendiz S, Crespo-Avilan GE, Lin YH, Hausenloy DJ. Mitochondria in acute myocardial infarction and cardioprotection. EBioMedicine. 2020;57:102884.

    Article  PubMed  PubMed Central  Google Scholar 

  390. Heusch G. Myocardial ischaemia-reperfusion injury and cardioprotection in perspective. Nat Rev Cardiol. 2020;17(12):773–89.

    Article  PubMed  Google Scholar 

  391. Oehler D, Spychala A, Gödecke A, Lang A, Gerdes N, Ruas J, et al. Full-length transcriptomic analysis in murine and human heart reveals diversity of PGC-1α promoters and isoforms regulated distinctly in myocardial ischemia and obesity. BMC Biol. 2022;20(1):169.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  392. Yurista SR, Silljé HHW, Oberdorf-Maass SU, Schouten EM, Pavez Giani MG, Hillebrands JL, et al. Sodium-glucose co-transporter 2 inhibition with empagliflozin improves cardiac function in non-diabetic rats with left ventricular dysfunction after myocardial infarction. Eur J Heart Fail. 2019;21(7):862–73.

    Article  CAS  PubMed  Google Scholar 

  393. Subramani J, Kundumani-Sridharan V, Das KC. Thioredoxin protects mitochondrial structure, function and biogenesis in myocardial ischemia-reperfusion via redox-dependent activation of AKT-CREB- PGC1α pathway in aged mice. Aging (Albany NY). 2020;12(19):19809–27.

    Article  CAS  PubMed  Google Scholar 

  394. Qi X, Wang J. Melatonin improves mitochondrial biogenesis through the AMPK/PGC1α pathway to attenuate ischemia/reperfusion-induced myocardial damage. Aging (Albany NY). 2020;12(8):7299–312.

    Article  CAS  PubMed  Google Scholar 

  395. Huang Q, Su H, Qi B, Wang Y, Yan K, Wang X, et al. A SIRT1 activator, ginsenoside Rc, promotes energy metabolism in cardiomyocytes and neurons. J Am Chem Soc. 2021;143(3):1416–27.

    Article  CAS  PubMed  Google Scholar 

  396. Shi X, Li Y, Wang Y, Ding T, Zhang X, Wu N. Pharmacological postconditioning with sappanone A ameliorates myocardial ischemia reperfusion injury and mitochondrial dysfunction via AMPK-mediated mitochondrial quality control. Toxicol Appl Pharmacol. 2021;427:115668.

    Article  CAS  PubMed  Google Scholar 

  397. Sun L, Zhao M, Yu XJ, Wang H, He X, Liu JK, et al. Cardioprotection by acetylcholine: a novel mechanism via mitochondrial biogenesis and function involving the PGC-1α pathway. J Cell Physiol. 2013;228(6):1238–48.

    Article  CAS  PubMed  Google Scholar 

  398. Ong SB, Subrayan S, Lim SY, Yellon DM, Davidson SM, Hausenloy DJ. Inhibiting mitochondrial fission protects the heart against ischemia/reperfusion injury. Circulation. 2010;121(18):2012–22.

    Article  CAS  PubMed  Google Scholar 

  399. Kalkhoran SB, Kriston-Vizi J, Hernandez-Resendiz S, Crespo-Avilan GE, Rosdah AA, Lees JG, et al. Hydralazine protects the heart against acute ischaemia/reperfusion injury by inhibiting Drp1-mediated mitochondrial fission. Cardiovasc Res. 2022;118(1):282–94.

    Article  CAS  PubMed  Google Scholar 

  400. Maneechote C, Palee S, Kerdphoo S, Jaiwongkam T, Chattipakorn SC, Chattipakorn N. Differential temporal inhibition of mitochondrial fission by Mdivi-1 exerts effective cardioprotection in cardiac ischemia/reperfusion injury. Clin Sci (Lond). 2018;132(15):1669–83.

    Article  CAS  PubMed  Google Scholar 

  401. Zhu H, Tan Y, Du W, Li Y, Toan S, Mui D, et al. Phosphoglycerate mutase 5 exacerbates cardiac ischemia-reperfusion injury through disrupting mitochondrial quality control. Redox Biol. 2021;38:101777.

    Article  CAS  PubMed  Google Scholar 

  402. Zaja I, Bai X, Liu Y, Kikuchi C, Dosenovic S, Yan Y, et al. Cdk1, PKCδ and calcineurin-mediated Drp1 pathway contributes to mitochondrial fission-induced cardiomyocyte death. Biochem Biophys Res Commun. 2014;453(4):710–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  403. Li Y, Xiong Z, Jiang Y, Zhou H, Yi L, Hu Y, et al. Klf4 deficiency exacerbates myocardial ischemia/reperfusion injury in mice via enhancing ROCK1/DRP1 pathway-dependent mitochondrial fission. J Mol Cell Cardiol. 2023;174:115–32.

    Article  CAS  PubMed  Google Scholar 

  404. Chen L, Chen XY, Wang QL, Yang SJ, Zhou H, Ding LS, et al. Astragaloside IV derivative (LS-102) alleviated myocardial ischemia reperfusion injury by inhibiting Drp1(Ser616) phosphorylation-mediated mitochondrial fission. Front Pharmacol. 2020;11:1083.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  405. Liu J, Yan W, Zhao X, Jia Q, Wang J, Zhang H, et al. Sirt3 attenuates post-infarction cardiac injury via inhibiting mitochondrial fission and normalization of AMPK-Drp1 pathways. Cell Signal. 2019;53:1–13.

    Article  CAS  PubMed  Google Scholar 

  406. Wang JX, Jiao JQ, Li Q, Long B, Wang K, Liu JP, et al. miR-499 regulates mitochondrial dynamics by targeting calcineurin and dynamin-related protein-1. Nat Med. 2011;17(1):71–8.

    Article  PubMed  Google Scholar 

  407. Sun S, Yu W, Xu H, Li C, Zou R, Wu NN, et al. TBC1D15-Drp1 interaction-mediated mitochondrial homeostasis confers cardioprotection against myocardial ischemia/reperfusion injury. Metabolism. 2022;134:155239.

    Article  CAS  PubMed  Google Scholar 

  408. Guan L, Che Z, Meng X, Yu Y, Li M, Yu Z, et al. MCU up-regulation contributes to myocardial ischemia-reperfusion Injury through calpain/OPA-1-mediated mitochondrial fusion/mitophagy Inhibition. J Cell Mol Med. 2019;23(11):7830–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  409. Maneechote C, Palee S, Kerdphoo S, Jaiwongkam T, Chattipakorn SC, Chattipakorn N. Balancing mitochondrial dynamics via increasing mitochondrial fusion attenuates infarct size and left ventricular dysfunction in rats with cardiac ischemia/reperfusion injury. Clin Sci (Lond). 2019;133(3):497–513.

    Article  CAS  PubMed  Google Scholar 

  410. Wang Z, Wang SP, Shao Q, Li PF, Sun Y, Luo LZ, et al. Brain-derived neurotrophic factor mimetic, 7,8-dihydroxyflavone, protects against myocardial ischemia by rebalancing optic atrophy 1 processing. Free Radic Biol Med. 2019;145:187–97.

    Article  CAS  PubMed  Google Scholar 

  411. Zhang Y, Wang Y, Xu J, Tian F, Hu S, Chen Y, et al. Melatonin attenuates myocardial ischemia-reperfusion injury via improving mitochondrial fusion/mitophagy and activating the AMPK-OPA1 signaling pathways. J Pineal Res. 2019;66(2):e12542.

    Article  PubMed  Google Scholar 

  412. Olmedo I, Pino G, Riquelme JA, Aranguiz P, Díaz MC, López-Crisosto C, et al. Inhibition of the proteasome preserves Mitofusin-2 and mitochondrial integrity, protecting cardiomyocytes during ischemia-reperfusion injury. Biochim Biophys Acta Mol Basis Dis. 2020;1866(5):165659.

    Article  CAS  PubMed  Google Scholar 

  413. Zhou J, Liu H, Zhang T, Wang Z, Zhang J, Lu Y, et al. MORN4 protects cardiomyocytes against ischemic injury via MFN2-mediated mitochondrial dynamics and mitophagy. Free Radic Biol Med. 2023;196:156–70.

    Article  CAS  PubMed  Google Scholar 

  414. Hall AR, Burke N, Dongworth RK, Kalkhoran SB, Dyson A, Vicencio JM, et al. Hearts deficient in both Mfn1 and Mfn2 are protected against acute myocardial infarction. Cell Death Dis. 2016;7(5):e2238.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  415. Papanicolaou KN, Khairallah RJ, Ngoh GA, Chikando A, Luptak I, O’Shea KM, et al. Mitofusin-2 maintains mitochondrial structure and contributes to stress-induced permeability transition in cardiac myocytes. Mol Cell Biol. 2011;31(6):1309–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  416. de Brito OM, Scorrano L. Mitofusin 2 tethers endoplasmic reticulum to mitochondria. Nature. 2008;456(7222):605–10.

    Article  PubMed  Google Scholar 

  417. Chen Y, Liu Y, Dorn GW 2nd. Mitochondrial fusion is essential for organelle function and cardiac homeostasis. Circ Res. 2011;109(12):1327–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  418. Kubli DA, Zhang X, Lee Y, Hanna RA, Quinsay MN, Nguyen CK, et al. Parkin protein deficiency exacerbates cardiac injury and reduces survival following myocardial infarction. J Biol Chem. 2013;288(2):915–26.

    Article  CAS  PubMed  Google Scholar 

  419. Tu M, Tan VP, Yu JD, Tripathi R, Bigham Z, Barlow M, et al. RhoA signaling increases mitophagy and protects cardiomyocytes against ischemia by stabilizing PINK1 protein and recruiting Parkin to mitochondria. Cell Death Differ. 2022;29(12):2472–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  420. Siddall HK, Yellon DM, Ong SB, Mukherjee UA, Burke N, Hall AR, et al. Loss of PINK1 increases the heart’s vulnerability to ischemia-reperfusion injury. PLoS ONE. 2013;8(4):e62400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  421. Bai Y, Wu J, Yang Z, Wang X, Zhang D, Ma J. Mitochondrial quality control in cardiac ischemia/reperfusion injury: new insights into mechanisms and implications. Cell Biol Toxicol. 2023;39(1):33–51.

    Article  CAS  PubMed  Google Scholar 

  422. Sun T, Ding W, Xu T, Ao X, Yu T, Li M, et al. Parkin regulates programmed necrosis and myocardial ischemia/reperfusion injury by targeting cyclophilin-D. Antioxid Redox Signal. 2019;31(16):1177–93.

    Article  CAS  PubMed  Google Scholar 

  423. Liu W, Chen C, Gu X, Zhang L, Mao X, Chen Z, et al. AM1241 alleviates myocardial ischemia-reperfusion injury in rats by enhancing Pink1/Parkin-mediated autophagy. Life Sci. 2021;272:119228.

    Article  CAS  PubMed  Google Scholar 

  424. Xiang Q, Wu M, Zhang L, Fu W, Yang J, Zhang B, et al. Gerontoxanthone I and macluraxanthone induce mitophagy and attenuate ischemia/reperfusion injury. Front Pharmacol. 2020;11:452.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  425. Tang L, Li YP, Hu J, Chen AH, Mo Y. Dexpramipexole attenuates myocardial ischemia/reperfusion injury through upregulation of mitophagy. Eur J Pharmacol. 2021;899:173962.

    Article  CAS  PubMed  Google Scholar 

  426. Cao S, Sun Y, Wang W, Wang B, Zhang Q, Pan C, et al. Poly (ADP-ribose) polymerase inhibition protects against myocardial ischaemia/reperfusion injury via suppressing mitophagy. J Cell Mol Med. 2019;23(10):6897–906.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  427. Ji W, Wei S, Hao P, Xing J, Yuan Q, Wang J, et al. Aldehyde dehydrogenase 2 has cardioprotective effects on myocardial ischaemia/reperfusion injury via suppressing mitophagy. Front Pharmacol. 2016;7:101.

    Article  PubMed  PubMed Central  Google Scholar 

  428. Xu Q, Liu S, Gong Q, Zhu R, Liu J, Wu Q, et al. Notch1 protects against ischemic-reperfusion injury by suppressing PTEN-pink1-mediated mitochondrial dysfunction and mitophagy. Cells. 2022;12(1):137.

    Article  PubMed  PubMed Central  Google Scholar 

  429. Wu J, Yang Y, Gao Y, Wang Z, Ma J. Melatonin attenuates anoxia/reoxygenation injury by inhibiting excessive mitophagy through the MT2/SIRT3/FoxO3a signaling pathway in H9c2 cells. Drug Des Devel Ther. 2020;14:2047–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  430. Xin T, Lu C. Irisin activates Opa1-induced mitophagy to protect cardiomyocytes against apoptosis following myocardial infarction. Aging (Albany NY). 2020;12(5):4474–88.

    Article  CAS  PubMed  Google Scholar 

  431. Zhang W, Ren H, Xu C, Zhu C, Wu H, Liu D, et al. Hypoxic mitophagy regulates mitochondrial quality and platelet activation and determines severity of I/R heart injury. Elife. 2016;5:e21407.

    Article  PubMed  PubMed Central  Google Scholar 

  432. Zhou H, Zhu P, Wang J, Zhu H, Ren J, Chen Y. Pathogenesis of cardiac ischemia reperfusion injury is associated with CK2α-disturbed mitochondrial homeostasis via suppression of FUNDC1-related mitophagy. Cell Death Differ. 2018;25(6):1080–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  433. Zhou H, Zhu P, Guo J, Hu N, Wang S, Li D, et al. Ripk3 induces mitochondrial apoptosis via inhibition of FUNDC1 mitophagy in cardiac IR injury. Redox Biol. 2017;13:498–507.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  434. Mao S, Tian S, Luo X, Zhou M, Cao Z, Li J. Overexpression of PLK1 relieved the myocardial ischemia-reperfusion injury of rats through inducing the mitophagy and regulating the p-AMPK/FUNDC1 axis. Bioengineered. 2021;12(1):2676–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  435. Zhang W, Siraj S, Zhang R, Chen Q. Mitophagy receptor FUNDC1 regulates mitochondrial homeostasis and protects the heart from I/R injury. Autophagy. 2017;13(6):1080–1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  436. Zhang Y, Liu D, Hu H, Zhang P, Xie R, Cui W. HIF-1α/BNIP3 signaling pathway-induced-autophagy plays protective role during myocardial ischemia-reperfusion injury. Biomed Pharmacother. 2019;120:109464.

    Article  CAS  PubMed  Google Scholar 

  437. Zhu N, Li J, Li Y, Zhang Y, Du Q, Hao P, et al. Berberine protects against simulated ischemia/reperfusion injury-induced H9C2 cardiomyocytes apoptosis in vitro and myocardial ischemia/reperfusion-induced apoptosis in vivo by regulating the mitophagy-mediated HIF-1α/BNIP3 pathway. Front Pharmacol. 2020;11:367.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  438. Lee TL, Lee MH, Chen YC, Lee YC, Lai TC, Lin HY, et al. Vitamin D attenuates ischemia/reperfusion-induced cardiac injury by reducing mitochondrial fission and mitophagy. Front Pharmacol. 2020;11:604700.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  439. Libby P. The changing landscape of atherosclerosis. Nature. 2021;592(7855):524–33.

    Article  CAS  PubMed  Google Scholar 

  440. Wang G, Yang Y, Ma H, Shi L, Jia W, Hao X, et al. LncRNA FENDRR inhibits ox-LDL induced mitochondrial energy metabolism disorder in aortic endothelial cells via miR-18a-5p/PGC-1α signaling pathway. Front Endocrinol (Lausanne). 2021;12:622665.

    Article  PubMed  Google Scholar 

  441. Karnewar S, Vasamsetti SB, Gopoju R, Kanugula AK, Ganji SK, Prabhakar S, et al. Mitochondria-targeted esculetin alleviates mitochondrial dysfunction by AMPK-mediated nitric oxide and SIRT3 regulation in endothelial cells: potential implications in atherosclerosis. Sci Rep. 2016;6:24108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  442. Shenouda SM, Widlansky ME, Chen K, Xu G, Holbrook M, Tabit CE, et al. Altered mitochondrial dynamics contributes to endothelial dysfunction in diabetes mellitus. Circulation. 2011;124(4):444–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  443. Wang Q, Zhang M, Torres G, Wu S, Ouyang C, Xie Z, et al. Metformin suppresses diabetes-accelerated atherosclerosis via the inhibition of Drp1-mediated mitochondrial fission. Diabetes. 2017;66(1):193–205.

    Article  CAS  PubMed  Google Scholar 

  444. Liu S, Zhao Y, Yao H, Zhang L, Chen C, Zheng Z, et al. DRP1 knockdown and atorvastatin alleviate ox-LDL-induced vascular endothelial cells injury: DRP1 is a potential target for preventing atherosclerosis. Exp Cell Res. 2023;429(2):113688.

    Article  CAS  PubMed  Google Scholar 

  445. Wang J, Chen H, Liu Y, Zhou W, Sun R, Xia M. Retinol binding protein 4 induces mitochondrial dysfunction and vascular oxidative damage. Atherosclerosis. 2015;240(2):335–44.

    Article  CAS  PubMed  Google Scholar 

  446. Zhu W, Yuan Y, Liao G, Li L, Liu J, Chen Y, et al. Mesenchymal stem cells ameliorate hyperglycemia-induced endothelial injury through modulation of mitophagy. Cell Death Dis. 2018;9(8):837.

    Article  PubMed  PubMed Central  Google Scholar 

  447. Li P, Bai Y, Zhao X, Tian T, Tang L, Ru J, et al. NR4A1 contributes to high-fat associated endothelial dysfunction by promoting CaMKII-Parkin-mitophagy pathways. Cell Stress Chaperones. 2018;23(4):749–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  448. Fang Y, Zhu Y, Wu Y, Liu L, Wang H. Protective effects of mitochondrial fission inhibition on ox-LDL induced VSMC foaming via metabolic reprogramming. Front Pharmacol. 2022;13:970151.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  449. Salabei JK, Hill BG. Mitochondrial fission induced by platelet-derived growth factor regulates vascular smooth muscle cell bioenergetics and cell proliferation. Redox Biol. 2013;1(1):542–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  450. Wang L, Yu T, Lee H, O’Brien DK, Sesaki H, Yoon Y. Decreasing mitochondrial fission diminishes vascular smooth muscle cell migration and ameliorates intimal hyperplasia. Cardiovasc Res. 2015;106(2):272–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  451. He L, Zhou Q, Huang Z, Xu J, Zhou H, Lv D, et al. PINK1/Parkin-mediated mitophagy promotes apelin-13-induced vascular smooth muscle cell proliferation by AMPKα and exacerbates atherosclerotic lesions. J Cell Physiol. 2019;234(6):8668–82.

    Article  CAS  PubMed  Google Scholar 

  452. Zhang X, Li X, Jia H, An G, Ni J. The m6A methyltransferase METTL3 modifies PGC-1α mRNA promoting mitochondrial dysfunction and oxLDL-induced inflammation in monocytes. J Biol Chem. 2021;297(3):101058.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  453. Su ZDZ, Li CQ, Wang HW, Zheng MM, Chen QW. Inhibition of DRP1-dependent mitochondrial fission by Mdivi-1 alleviates atherosclerosis through the modulation of M1 polarization. J Transl Med. 2023;21(1):427.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  454. Xu Y, Zhang Y, Xu Y, Zang G, Li B, Xia H, et al. Activation of CD137 signaling promotes macrophage apoptosis dependent on p38 MAPK pathway-mediated mitochondrial fission. Int J Biochem Cell Biol. 2021;136:106003.

    Article  CAS  PubMed  Google Scholar 

  455. Jin Y, Liu Y, Xu L, Xu J, Xiong Y, Peng Y, et al. Novel role for caspase 1 inhibitor VX765 in suppressing NLRP3 inflammasome assembly and atherosclerosis via promoting mitophagy and efferocytosis. Cell Death Dis. 2022;13(5):512.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  456. Choi SH, Agatisa-Boyle C, Gonen A, Kim A, Kim J, Alekseeva E, et al. Intracellular AIBP (apolipoprotein A-I binding protein) regulates oxidized LDL (low-density lipoprotein)-induced mitophagy in macrophages. Arterioscler Thromb Vasc Biol. 2021;41(2):e82–96.

    Article  CAS  PubMed  Google Scholar 

  457. Duan M, Chen H, Yin L, Zhu X, Novák P, Lv Y, et al. Mitochondrial apolipoprotein A-I binding protein alleviates atherosclerosis by regulating mitophagy and macrophage polarization. Cell Commun Signal. 2022;20(1):60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  458. Zhang X, Sergin I, Evans TD, Jeong SJ, Rodriguez-Velez A, Kapoor D, et al. High-protein diets increase cardiovascular risk by activating macrophage mTOR to suppress mitophagy. Nat Metab. 2020;2(1):110–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  459. Alfaras I, Di Germanio C, Bernier M, Csiszar A, Ungvari Z, Lakatta EG, et al. Pharmacological strategies to retard cardiovascular aging. Circ Res. 2016;118(10):1626–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  460. Nakou ES, Parthenakis FI, Kallergis EM, Marketou ME, Nakos KS, Vardas PE. Healthy aging and myocardium: a complicated process with various effects in cardiac structure and physiology. Int J Cardiol. 2016;209:167–75.

    Article  CAS  PubMed  Google Scholar 

  461. Boengler K, Kosiol M, Mayr M, Schulz R, Rohrbach S. Mitochondria and ageing: role in heart, skeletal muscle and adipose tissue. J Cachexia Sarcopenia Muscle. 2017;8(3):349–69.

    Article  PubMed  PubMed Central  Google Scholar 

  462. Wang J, Li S, Wang J, Wu F, Chen Y, Zhang H, et al. Spermidine alleviates cardiac aging by improving mitochondrial biogenesis and function. Aging (Albany NY). 2020;12(1):650–71.

    Article  CAS  PubMed  Google Scholar 

  463. Feng W, Liu J, Wang S, Hu Y, Pan H, Hu T, et al. Alginate oligosaccharide alleviates D-galactose-induced cardiac ageing via regulating myocardial mitochondria function and integrity in mice. J Cell Mol Med. 2021;25(15):7157–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  464. Zhao L, Zou X, Feng Z, Luo C, Liu J, Li H, et al. Evidence for association of mitochondrial metabolism alteration with lipid accumulation in aging rats. Exp Gerontol. 2014;56:3–12.

    Article  CAS  PubMed  Google Scholar 

  465. Whitehead N, Gill JF, Brink M, Handschin C. Moderate modulation of cardiac PGC-1α expression partially affects age-associated transcriptional remodeling of the heart. Front Physiol. 2018;9:242.

    Article  PubMed  PubMed Central  Google Scholar 

  466. Fernández-Ortiz M, Sayed RKA, Fernández-Martínez J, Cionfrini A, Aranda-Martínez P, Escames G, et al. Melatonin/Nrf2/NLRP3 connection in mouse heart mitochondria during aging. Antioxidants (Basel). 2020;9(12):1187.

    Article  PubMed  Google Scholar 

  467. Ljubicic V, Menzies KJ, Hood DA. Mitochondrial dysfunction is associated with a pro-apoptotic cellular environment in senescent cardiac muscle. Mech Ageing Dev. 2010;131(2):79–88.

    Article  CAS  PubMed  Google Scholar 

  468. Zhang YY, Ning BT. Signaling pathways and intervention therapies in sepsis. Signal Transduct Target Ther. 2021;6(1):407.

    Article  PubMed  PubMed Central  Google Scholar 

  469. Hoshino A, Mita Y, Okawa Y, Ariyoshi M, Iwai-Kanai E, Ueyama T, et al. Cytosolic p53 inhibits Parkin-mediated mitophagy and promotes mitochondrial dysfunction in the mouse heart. Nat Commun. 2013;4:2308.

    Article  PubMed  Google Scholar 

  470. Wang Y, Xu Y, Guo W, Fang Y, Hu L, Wang R, et al. Ablation of Shank3 alleviates cardiac dysfunction in aging mice by promoting CaMKII activation and Parkin-mediated mitophagy. Redox Biol. 2022;58:102537.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  471. Soh JEC, Shimizu A, Molla MR, Zankov DP, Nguyen LKC, Khan MR, et al. RhoA rescues cardiac senescence by regulating Parkin-mediated mitophagy. J Biol Chem. 2023;299(3):102993.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  472. Liu X, Bai X, Liu H, Hong Y, Cui H, Wang L, et al. LncRNA LOC105378097 inhibits cardiac mitophagy in natural ageing mice. Clin Transl Med. 2022;12(6):e908.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  473. Gao B, Yu W, Lv P, Liang X, Sun S, Zhang Y. Parkin overexpression alleviates cardiac aging through facilitating K63-polyubiquitination of TBK1 to facilitate mitophagy. Biochim Biophys Acta Mol Basis Dis. 2021;1867(1):165997.

    Article  CAS  PubMed  Google Scholar 

  474. Franco C, Sciatti E, Favero G, Bonomini F, Vizzardi E, Rezzani R. Essential hypertension and oxidative stress: novel future perspectives. Int J Mol Sci. 2022;23(22):14489.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  475. Lin KL, Chen SD, Lin KJ, Liou CW, Chuang YC, Wang PW, et al. Quality matters? The involvement of mitochondrial quality control in cardiovascular disease. Front Cell Dev Biol. 2021;9:636295.

    Article  PubMed  PubMed Central  Google Scholar 

  476. Sun W, Wang X, Hou C, Yang L, Li H, Guo J, et al. Oleuropein improves mitochondrial function to attenuate oxidative stress by activating the Nrf2 pathway in the hypothalamic paraventricular nucleus of spontaneously hypertensive rats. Neuropharmacology. 2017;113(Pt A):556–66.

    Article  CAS  PubMed  Google Scholar 

  477. Sun W, Yan C, Frost B, Wang X, Hou C, Zeng M, et al. Pomegranate extract decreases oxidative stress and alleviates mitochondrial impairment by activating AMPK-Nrf2 in hypothalamic paraventricular nucleus of spontaneously hypertensive rats. Sci Rep. 2016;6:34246.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  478. Ooi K, Hu L, Feng Y, Han C, Ren X, Qian X, et al. Sigma-1 receptor activation suppresses microglia M1 polarization via regulating endoplasmic reticulum-mitochondria contact and mitochondrial functions in stress-induced hypertension rats. Mol Neurobiol. 2021;58(12):6625–46.

    Article  CAS  PubMed  Google Scholar 

  479. Deng Y, Li S, Chen Z, Wang W, Geng B, Cai J. Mdivi-1, a mitochondrial fission inhibitor, reduces angiotensin-II- induced hypertension by mediating VSMC phenotypic switch. Biomed Pharmacother. 2021;140:111689.

    Article  CAS  PubMed  Google Scholar 

  480. Cole JB, Florez JC. Genetics of diabetes mellitus and diabetes complications. Nat Rev Nephrol. 2020;16(7):377–90.

    Article  PubMed  PubMed Central  Google Scholar 

  481. Xu DQ, Li CJ, Jiang ZZ, Wang L, Huang HF, Li ZJ, et al. The hypoglycemic mechanism of catalpol involves increased AMPK-mediated mitochondrial biogenesis. Acta Pharmacol Sin. 2020;41(6):791–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  482. Zhu Y, Yang H, Deng J, Fan D. Ginsenoside Rg5 improves insulin resistance and mitochondrial biogenesis of liver via regulation of the Sirt1/PGC-1α signaling pathway in db/db mice. J Agric Food Chem. 2021;69(30):8428–39.

    Article  CAS  PubMed  Google Scholar 

  483. Li X, Xu Z, Jiang Z, Sun L, Ji J, Miao J, et al. Hypoglycemic effect of catalpol on high-fat diet/streptozotocin-induced diabetic mice by increasing skeletal muscle mitochondrial biogenesis. Acta Biochim Biophys Sin (Shanghai). 2014;46(9):738–48.

    Article  CAS  PubMed  Google Scholar 

  484. Wu M, Zhang C, Xie M, Zhen Y, Lai B, Liu J, et al. Compartmentally scavenging hepatic oxidants through AMPK/SIRT3-PGC1α axis improves mitochondrial biogenesis and glucose catabolism. Free Radic Biol Med. 2021;168:117–28.

    Article  CAS  PubMed  Google Scholar 

  485. Besseiche A, Riveline JP, Gautier JF, Bréant B, Blondeau B. Metabolic roles of PGC-1α and its implications for type 2 diabetes. Diabetes Metab. 2015;41(5):347–57.

    Article  CAS  PubMed  Google Scholar 

  486. Gundersen AE, Kugler BA, McDonald PM, Veraksa A, Houmard JA, Zou K. Altered mitochondrial network morphology and regulatory proteins in mitochondrial quality control in myotubes from severely obese humans with or without type 2 diabetes. Appl Physiol Nutr Metab. 2020;45(3):283–93.

    Article  CAS  PubMed  Google Scholar 

  487. Jheng HF, Tsai PJ, Guo SM, Kuo LH, Chang CS, Su IJ, et al. Mitochondrial fission contributes to mitochondrial dysfunction and insulin resistance in skeletal muscle. Mol Cell Biol. 2012;32(2):309–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  488. Zheng L, Rao Z, Guo Y, Chen P, Xiao W. High-intensity interval training restores glycolipid metabolism and mitochondrial function in skeletal muscle of mice with type 2 diabetes. Front Endocrinol (Lausanne). 2020;11:561.

    Article  PubMed  Google Scholar 

  489. Finocchietto P, Perez H, Blanco G, Miksztowicz V, Marotte C, Morales C, et al. Inhibition of mitochondrial fission by Drp-1 blockade by short-term leptin and Mdivi-1 treatment improves white adipose tissue abnormalities in obesity and diabetes. Pharmacol Res. 2022;178:106028.

    Article  CAS  PubMed  Google Scholar 

  490. Apostolova N, Vezza T, Muntane J, Rocha M, Víctor VM. Mitochondrial dysfunction and mitophagy in type 2 diabetes: pathophysiology and therapeutic targets. Antioxid Redox Signal. 2023;39(4–6):278–320.

    Article  CAS  PubMed  Google Scholar 

  491. Drew BG, Ribas V, Le JA, Henstridge DC, Phun J, Zhou Z, et al. HSP72 is a mitochondrial stress sensor critical for Parkin action, oxidative metabolism, and insulin sensitivity in skeletal muscle. Diabetes. 2014;63(5):1488–505.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  492. Edmunds LR, Xie B, Mills AM, Huckestein BR, Undamatla R, Murali A, et al. Liver-specific Prkn knockout mice are more susceptible to diet-induced hepatic steatosis and insulin resistance. Mol Metab. 2020;41:101051.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  493. de Marañón AM, Díaz-Pozo P, Canet F, Díaz-Morales N, Abad-Jiménez Z, López-Domènech S, et al. Metformin modulates mitochondrial function and mitophagy in peripheral blood mononuclear cells from type 2 diabetic patients. Redox Biol. 2022;53:102342.

    Article  PubMed  PubMed Central  Google Scholar 

  494. Gupta P, Sharma G, Lahiri A, Barthwal MK. FOXO3a acetylation regulates PINK1, mitophagy, inflammasome activation in murine palmitate-conditioned and diabetic macrophages. J Leukoc Biol. 2022;111(3):611–27.

    Article  CAS  PubMed  Google Scholar 

  495. Jin X, Qiu T, Li L, Yu R, Chen X, Li C, et al. Pathophysiology of obesity and its associated diseases. Acta Pharm Sin B. 2023;13(6):2403–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  496. Blüher M. Obesity: global epidemiology and pathogenesis. Nat Rev Endocrinol. 2019;15(5):288–98.

    Article  PubMed  Google Scholar 

  497. Moore TM, Cheng L, Wolf DM, Ngo J, Segawa M, Zhu X, et al. Parkin regulates adiposity by coordinating mitophagy with mitochondrial biogenesis in white adipocytes. Nat Commun. 2022;13(1):6661.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  498. Woo CY, Jang JE, Lee SE, Koh EH, Lee KU. Mitochondrial dysfunction in adipocytes as a primary cause of adipose tissue inflammation. Diabetes Metab J. 2019;43(3):247–56.

    Article  PubMed  PubMed Central  Google Scholar 

  499. Koppen M, Langer T. Protein degradation within mitochondria: versatile activities of AAA proteases and other peptidases. Crit Rev Biochem Mol Biol. 2007;42(3):221–42.

    Article  CAS  PubMed  Google Scholar 

  500. Quirós PM, Ramsay AJ, Sala D, Fernández-Vizarra E, Rodríguez F, Peinado JR, et al. Loss of mitochondrial protease OMA1 alters processing of the GTPase OPA1 and causes obesity and defective thermogenesis in mice. EMBO J. 2012;31(9):2117–33.

    Article  PubMed  PubMed Central  Google Scholar 

  501. Korwitz A, Merkwirth C, Richter-Dennerlein R, Tröder SE, Sprenger H-G, Quirós PM, et al. Loss of OMA1 delays neurodegeneration by preventing stress-induced OPA1 processing in mitochondria. J Cell Biol. 2016;212(2):157–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  502. Axelrod CL, Dantas WS, Kirwan JP. Sarcopenic obesity: emerging mechanisms and therapeutic potential. Metabolism. 2023;146:155639.

    Article  CAS  PubMed  Google Scholar 

  503. Dantas WS, Zunica ERM, Heintz EC, Vandanmagsar B, Floyd ZE, Yu Y, et al. Mitochondrial uncoupling attenuates sarcopenic obesity by enhancing skeletal muscle mitophagy and quality control. J Cachexia Sarcopenia Muscle. 2022;13(3):1821–36.

    Article  PubMed  PubMed Central  Google Scholar 

  504. Marcangeli V, Youssef L, Dulac M, Carvalho LP, Hajj-Boutros G, Reynaud O, et al. Impact of high-intensity interval training with or without l-citrulline on physical performance, skeletal muscle, and adipose tissue in obese older adults. J Cachexia Sarcopenia Muscle. 2022;13(3):1526–40.

    Article  PubMed  PubMed Central  Google Scholar 

  505. Fu T, Xu Z, Liu L, Guo Q, Wu H, Liang X, et al. Mitophagy directs muscle-adipose crosstalk to alleviate dietary obesity. Cell Rep. 2018;23(5):1357–72.

    Article  CAS  PubMed  Google Scholar 

  506. Ahuja P, Ng CF, Pang BPS, Chan WS, Tse MCL, Bi X, et al. Muscle-generated BDNF (brain derived neurotrophic factor) maintains mitochondrial quality control in female mice. Autophagy. 2022;18(6):1367–84.

    Article  CAS  PubMed  Google Scholar 

  507. Scheltens P, De Strooper B, Kivipelto M, Holstege H, Chételat G, Teunissen CE, et al. Alzheimer’s disease. Lancet. 2021;397(10284):1577–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  508. Xie H, Guan J, Borrelli LA, Xu J, Serrano-Pozo A, Bacskai BJ. Mitochondrial alterations near amyloid plaques in an Alzheimer’s disease mouse model. J Neurosci. 2013;33(43):17042–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  509. Yao J, Irwin RW, Zhao L, Nilsen J, Hamilton RT, Brinton RD. Mitochondrial bioenergetic deficit precedes Alzheimer’s pathology in female mouse model of Alzheimer’s disease. Proc Natl Acad Sci U S A. 2009;106(34):14670–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  510. Kondadi AK, Wang S, Montagner S, Kladt N, Korwitz A, Martinelli P, et al. Loss of the m-AAA protease subunit AFG3L2 causes mitochondrial transport defects and tau hyperphosphorylation. EMBO J. 2014;33(9):1011–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  511. Mattson MP, Gleichmann M, Cheng A. Mitochondria in neuroplasticity and neurological disorders. Neuron. 2008;60(5):748–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  512. Cummins N, Tweedie A, Zuryn S, Bertran-Gonzalez J, Götz J. Disease-associated tau impairs mitophagy by inhibiting Parkin translocation to mitochondria. EMBO J. 2019;38(3):e99360.

    Article  PubMed  Google Scholar 

  513. Jeong YY, Han S, Jia N, Zhang M, Sheshadri P, Tammineni P, et al. Broad activation of the Parkin pathway induces synaptic mitochondrial deficits in early tauopathy. Brain. 2022;145(1):305–23.

    Article  PubMed  PubMed Central  Google Scholar 

  514. Sorrentino V, Romani M, Mouchiroud L, Beck JS, Zhang H, D’Amico D, et al. Enhancing mitochondrial proteostasis reduces amyloid-β proteotoxicity. Nature. 2017;552(7684):187–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  515. Vaillant-Beuchot L, Mary A, Pardossi-Piquard R, Bourgeois A, Lauritzen I, Eysert F, et al. Accumulation of amyloid precursor protein C-terminal fragments triggers mitochondrial structure, function, and mitophagy defects in Alzheimer’s disease models and human brains. Acta Neuropathol. 2021;141(1):39–65.

    Article  CAS  PubMed  Google Scholar 

  516. Cho DH, Nakamura T, Fang J, Cieplak P, Godzik A, Gu Z, et al. S-nitrosylation of Drp1 mediates beta-amyloid-related mitochondrial fission and neuronal injury. Science. 2009;324(5923):102–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  517. Wang W, Ma X, Bhatta S, Shao C, Zhao F, Fujioka H, et al. Intraneuronal β-amyloid impaired mitochondrial proteostasis through the impact on LONP1. Proc Natl Acad Sci U S A. 2023;120(51):e2316823120.

    Article  CAS  PubMed  Google Scholar 

  518. Katsouri L, Lim YM, Blondrath K, Eleftheriadou I, Lombardero L, Birch AM, et al. PPARγ-coactivator-1α gene transfer reduces neuronal loss and amyloid-β generation by reducing β-secretase in an Alzheimer’s disease model. Proc Natl Acad Sci U S A. 2016;113(43):12292–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  519. Yin J, Reiman EM, Beach TG, Serrano GE, Sabbagh MN, Nielsen M, et al. Effect of ApoE isoforms on mitochondria in Alzheimer disease. Neurology. 2020;94(23):e2404–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  520. Ye H, Robak LA, Yu M, Cykowski M, Shulman JM. Genetics and pathogenesis of Parkinson’s syndrome. Annu Rev Pathol. 2023;18:95–121.

    Article  CAS  PubMed  Google Scholar 

  521. Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey K, Gispert S, et al. Hereditary early-onset Parkinson’s disease caused by mutations in PINK1. Science. 2004;304(5674):1158–60.

    Article  CAS  PubMed  Google Scholar 

  522. Poewe W, Seppi K, Tanner CM, Halliday GM, Brundin P, Volkmann J, et al. Parkinson disease. Nat Rev Dis Primers. 2017;3:17013.

    Article  PubMed  Google Scholar 

  523. Wang W, Wang X, Fujioka H, Hoppel C, Whone AL, Caldwell MA, et al. Parkinson’s disease-associated mutant VPS35 causes mitochondrial dysfunction by recycling DLP1 complexes. Nat Med. 2016;22(1):54–63.

    Article  CAS  PubMed  Google Scholar 

  524. Lee Y, Stevens DA, Kang SU, Jiang H, Lee YI, Ko HS, et al. PINK1 primes Parkin-mediated ubiquitination of PARIS in dopaminergic neuronal survival. Cell Rep. 2017;18(4):918–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  525. Panicker N, Kam TI, Wang H, Neifert S, Chou SC, Kumar M, et al. Neuronal NLRP3 is a parkin substrate that drives neurodegeneration in Parkinson’s disease. Neuron. 2022;110(15):2422–37.e9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  526. Walker FO. Huntington’s disease. Lancet. 2007;369(9557):218–28.

    Article  CAS  PubMed  Google Scholar 

  527. Khalil B, El Fissi N, Aouane A, Cabirol-Pol MJ, Rival T, Liévens JC. PINK1-induced mitophagy promotes neuroprotection in Huntington’s disease. Cell Death Dis. 2015;6(1):e1617.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  528. Song W, Chen J, Petrilli A, Liot G, Klinglmayr E, Zhou Y, et al. Mutant huntingtin binds the mitochondrial fission GTPase dynamin-related protein-1 and increases its enzymatic activity. Nat Med. 2011;17(3):377–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  529. Guo X, Disatnik MH, Monbureau M, Shamloo M, Mochly-Rosen D, Qi X. Inhibition of mitochondrial fragmentation diminishes Huntington’s disease-associated neurodegeneration. J Clin Invest. 2013;123(12):5371–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  530. Costa V, Giacomello M, Hudec R, Lopreiato R, Ermak G, Lim D, et al. Mitochondrial fission and cristae disruption increase the response of cell models of Huntington’s disease to apoptotic stimuli. EMBO Mol Med. 2010;2(12):490–503.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  531. Naia L, Carmo C, Campesan S, Fão L, Cotton VE, Valero J, et al. Mitochondrial SIRT3 confers neuroprotection in Huntington’s disease by regulation of oxidative challenges and mitochondrial dynamics. Free Radic Biol Med. 2021;163:163–79.

    Article  CAS  PubMed  Google Scholar 

  532. Dickey AS, Pineda VV, Tsunemi T, Liu PP, Miranda HC, Gilmore-Hall SK, et al. PPAR-δ is repressed in Huntington’s disease, is required for normal neuronal function and can be targeted therapeutically. Nat Med. 2016;22:37–45.

    Article  CAS  PubMed  Google Scholar 

  533. Cui L, Jeong H, Borovecki F, Parkhurst CN, Tanese N, Krainc D. Transcriptional repression of PGC-1alpha by mutant huntingtin leads to mitochondrial dysfunction and neurodegeneration. Cell. 2006;127(1):59–69.

    Article  CAS  PubMed  Google Scholar 

  534. Weydt P, Pineda VV, Torrence AE, Libby RT, Satterfield TF, Lazarowski ER, et al. Thermoregulatory and metabolic defects in Huntington’s disease transgenic mice implicate PGC-1alpha in Huntington’s disease neurodegeneration. Cell Metab. 2006;4(5):349–62.

    Article  CAS  PubMed  Google Scholar 

  535. Chiang MC, Chern Y, Huang RN. PPARgamma rescue of the mitochondrial dysfunction in Huntington’s disease. Neurobiol Dis. 2012;45(1):322–8.

    Article  CAS  PubMed  Google Scholar 

  536. Di Cristo F, Finicelli M, Digilio FA, Paladino S, Valentino A, Scialò F, et al. Meldonium improves Huntington’s disease mitochondrial dysfunction by restoring peroxisome proliferator-activated receptor γ coactivator 1α expression. J Cell Physiol. 2019;234(6):9233–46.

    Article  PubMed  Google Scholar 

  537. Lee H, Fenster RJ, Pineda SS, Gibbs WS, Mohammadi S, Davila-Velderrain J, et al. Cell type-specific transcriptomics reveals that mutant huntingtin leads to mitochondrial RNA release and neuronal innate immune activation. Neuron. 2020;107(5):891–908.e8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  538. Da Cruz S, Parone PA, Lopes VS, Lillo C, McAlonis-Downes M, Lee SK, et al. Elevated PGC-1α activity sustains mitochondrial biogenesis and muscle function without extending survival in a mouse model of inherited ALS. Cell Metab. 2012;15(5):778–86.

    Article  PubMed  PubMed Central  Google Scholar 

  539. Magrané J, Sahawneh MA, Przedborski S, Estévez ÁG, Manfredi G. Mitochondrial dynamics and bioenergetic dysfunction is associated with synaptic alterations in mutant SOD1 motor neurons. J Neurosci. 2012;32(1):229–42.

    Article  PubMed  PubMed Central  Google Scholar 

  540. Wang L, Gao J, Liu J, Siedlak SL, Torres S, Fujioka H, et al. Mitofusin 2 regulates axonal transport of calpastatin to prevent neuromuscular synaptic elimination in skeletal muscles. Cell Metab. 2018;28(3):400–14.e8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  541. Palomo GM, Granatiero V, Kawamata H, Konrad C, Kim M, Arreguin AJ, et al. Parkin is a disease modifier in the mutant SOD1 mouse model of ALS. EMBO Mol Med. 2018;10(10):e8888.

    Article  PubMed  PubMed Central  Google Scholar 

  542. Harding O, Evans CS, Ye J, Cheung J, Maniatis T, Holzbaur ELF. ALS- and FTD-associated missense mutations in TBK1 differentially disrupt mitophagy. Proc Natl Acad Sci U S A. 2021;118(24):e2025053118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  543. Magrì A, Lipari CLR, Risiglione P, Zimbone S, Guarino F, Caccamo A, et al. ERK1/2-dependent TSPO overactivation associates with the loss of mitophagy and mitochondrial respiration in ALS. Cell Death Dis. 2023;14(2):122.

    Article  PubMed  PubMed Central  Google Scholar 

  544. Sun X, Duan Y, Qin C, Li JC, Duan G, Deng X, et al. Distinct multilevel misregulations of Parkin and PINK1 revealed in cell and animal models of TDP-43 proteinopathy. Cell Death Dis. 2018;9(10):953.

    Article  PubMed  PubMed Central  Google Scholar 

  545. Gaweda-Walerych K, Walerych D, Berdyński M, Buratti E, Zekanowski C. Parkin levels decrease in fibroblasts with progranulin (PGRN) pathogenic variants and in a cellular model of PGRN deficiency. Front Mol Neurosci. 2021;14:676478.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  546. Altman T, Ionescu A, Ibraheem A, Priesmann D, Gradus-Pery T, Farberov L, et al. Axonal TDP-43 condensates drive neuromuscular junction disruption through inhibition of local synthesis of nuclear encoded mitochondrial proteins. Nat Commun. 2021;12(1):6914.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  547. Song W, Song Y, Kincaid B, Bossy B, Bossy-Wetzel E. Mutant SOD1G93A triggers mitochondrial fragmentation in spinal cord motor neurons: neuroprotection by SIRT3 and PGC-1α. Neurobiol Dis. 2013;51:72–81.

    Article  CAS  PubMed  Google Scholar 

  548. Han B, Jiang W, Cui P, Zheng K, Dang C, Wang J, et al. Microglial PGC-1α protects against ischemic brain injury by suppressing neuroinflammation. Genome Med. 2021;13(1):47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  549. Zhang X, Yuan Y, Jiang L, Zhang J, Gao J, Shen Z, et al. Endoplasmic reticulum stress induced by tunicamycin and thapsigargin protects against transient ischemic brain injury: Involvement of PARK2-dependent mitophagy. Autophagy. 2014;10(10):1801–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  550. Zhang X, Yan H, Yuan Y, Gao J, Shen Z, Cheng Y, et al. Cerebral ischemia-reperfusion-induced autophagy protects against neuronal injury by mitochondrial clearance. Autophagy. 2013;9(9):1321–33.

    Article  CAS  PubMed  Google Scholar 

  551. Cai Y, Yang E, Yao X, Zhang X, Wang Q, Wang Y, et al. FUNDC1-dependent mitophagy induced by tPA protects neurons against cerebral ischemia-reperfusion injury. Redox Biol. 2021;38:101792.

    Article  CAS  PubMed  Google Scholar 

  552. Wu Q, Liu J, Mao Z, Tian L, Wang N, Wang G, et al. Ligustilide attenuates ischemic stroke injury by promoting Drp1-mediated mitochondrial fission via activation of AMPK. Phytomedicine. 2022;95:153884.

    Article  CAS  PubMed  Google Scholar 

  553. Kumari S, Anderson L, Farmer S, Mehta SL, Li PA. Hyperglycemia alters mitochondrial fission and fusion proteins in mice subjected to cerebral ischemia and reperfusion. Transl Stroke Res. 2012;3(2):296–304.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  554. Martorell-Riera A, Segarra-Mondejar M, Muñoz JP, Ginet V, Olloquequi J, Pérez-Clausell J, et al. Mfn2 downregulation in excitotoxicity causes mitochondrial dysfunction and delayed neuronal death. EMBO J. 2014;33(20):2388–407.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  555. Grohm J, Kim SW, Mamrak U, Tobaben S, Cassidy-Stone A, Nunnari J, et al. Inhibition of Drp1 provides neuroprotection in vitro and in vivo. Cell Death Differ. 2012;19(9):1446–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  556. Sosulski ML, Gongora R, Danchuk S, Dong C, Luo F, Sanchez CG. Deregulation of selective autophagy during aging and pulmonary fibrosis: the role of TGFβ1. Aging Cell. 2015;14(5):774–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  557. Bueno M, Lai YC, Romero Y, Brands J, St Croix CM, Kamga C, et al. PINK1 deficiency impairs mitochondrial homeostasis and promotes lung fibrosis. J Clin Invest. 2015;125(2):521–38.

    Article  PubMed  Google Scholar 

  558. Bueno M, Brands J, Voltz L, Fiedler K, Mays B, St Croix C, et al. ATF3 represses PINK1 gene transcription in lung epithelial cells to control mitochondrial homeostasis. Aging Cell. 2018;17(2):e12720.

    Article  PubMed  PubMed Central  Google Scholar 

  559. Larson-Casey JL, Deshane JS, Ryan AJ, Thannickal VJ, Carter AB. Macrophage Akt1 kinase-mediated mitophagy modulates apoptosis resistance and pulmonary fibrosis. Immunity. 2016;44(3):582–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  560. Ganzleben I, He GW, Günther C, Prigge ES, Richter K, Rieker RJ, et al. PGAM5 is a key driver of mitochondrial dysfunction in experimental lung fibrosis. Cell Mol Life Sci. 2019;76(23):4783–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  561. Kumar S, Pan CC, Shah N, Wheeler SE, Hoyt KR, Hempel N, et al. Activation of mitofusin2 by Smad2-RIN1 complex during mitochondrial fusion. Mol Cell. 2016;62(4):520–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  562. Guo T, Jiang CS, Yang SZ, Zhu Y, He C, Carter AB, et al. Mitochondrial fission and bioenergetics mediate human lung fibroblast durotaxis. JCI insight. 2023;8(1):e157348.

    Article  PubMed  PubMed Central  Google Scholar 

  563. Yu G, Tzouvelekis A, Wang R, Herazo-Maya JD, Ibarra GH, Srivastava A, et al. Thyroid hormone inhibits lung fibrosis in mice by improving epithelial mitochondrial function. Nat Med. 2018;24(1):39–49.

    Article  CAS  PubMed  Google Scholar 

  564. Chung KP, Hsu CL, Fan LC, Huang Z, Bhatia D, Chen YJ, et al. Mitofusins regulate lipid metabolism to mediate the development of lung fibrosis. Nat Commun. 2019;10(1):3390.

    Article  PubMed  PubMed Central  Google Scholar 

  565. Ito S, Araya J, Kurita Y, Kobayashi K, Takasaka N, Yoshida M, et al. PARK2-mediated mitophagy is involved in regulation of HBEC senescence in COPD pathogenesis. Autophagy. 2015;11(3):547–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  566. Araya J, Tsubouchi K, Sato N, Ito S, Minagawa S, Hara H, et al. PRKN-regulated mitophagy and cellular senescence during COPD pathogenesis. Autophagy. 2019;15(3):510–26.

    Article  CAS  PubMed  Google Scholar 

  567. Mizumura K, Cloonan SM, Nakahira K, Bhashyam AR, Cervo M, Kitada T, et al. Mitophagy-dependent necroptosis contributes to the pathogenesis of COPD. J Clin Invest. 2014;124(9):3987–4003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  568. Kosmider B, Lin CR, Karim L, Tomar D, Vlasenko L, Marchetti N, et al. Mitochondrial dysfunction in human primary alveolar type II cells in emphysema. EBioMedicine. 2019;46:305–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  569. Maremanda KP, Sundar IK, Rahman I. Role of inner mitochondrial protein OPA1 in mitochondrial dysfunction by tobacco smoking and in the pathogenesis of COPD. Redox Biol. 2021;45:102055.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  570. Beaufils F, Esteves P, Enaud R, Germande O, Celle A, Marthan R, et al. Mitochondria are involved in bronchial smooth muscle remodeling in severe preschool wheezers. J Allergy Clin Immunol. 2021;148(2):645–51.e11.

    Article  CAS  PubMed  Google Scholar 

  571. Trian T, Benard G, Begueret H, Rossignol R, Girodet PO, Ghosh D, et al. Bronchial smooth muscle remodeling involves calcium-dependent enhanced mitochondrial biogenesis in asthma. J Exp Med. 2007;204(13):3173–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  572. Zhang Y, Do DC, Hu X, Wang J, Zhao Y, Mishra S, et al. CaMKII oxidation regulates cockroach allergen-induced mitophagy in asthma. J Allergy Clin Immunol. 2021;147(4):1464–77.e11.

    Article  CAS  PubMed  Google Scholar 

  573. Lin YC, Lin YC, Tsai ML, Liao WT, Hung CH. TSLP regulates mitochondrial ROS-induced mitophagy via histone modification in human monocytes. Cell Biosci. 2022;12(1):32.

    Article  PubMed  PubMed Central  Google Scholar 

  574. Pattnaik B, Bodas M, Bhatraju NK, Ahmad T, Pant R, Guleria R, et al. IL-4 promotes asymmetric dimethylarginine accumulation, oxo-nitrative stress, and hypoxic response-induced mitochondrial loss in airway epithelial cells. J Allergy Clin Immunol. 2016;138(1):130–41.e9.

    Article  CAS  PubMed  Google Scholar 

  575. Chang AL, Ulrich A, Suliman HB, Piantadosi CA. Redox regulation of mitophagy in the lung during murine Staphylococcus aureus sepsis. Free Radic Biol Med. 2015;78:179–89.

    Article  CAS  PubMed  Google Scholar 

  576. Yu J, Shi J, Wang D, Dong S, Zhang Y, Wang M, et al. Heme oxygenase-1/carbon monoxide-regulated mitochondrial dynamic equilibrium contributes to the attenuation of endotoxin-induced acute lung injury in rats and in lipopolysaccharide-activated macrophages. Anesthesiology. 2016;125(6):1190–201.

    Article  CAS  PubMed  Google Scholar 

  577. Chen BB, Coon TA, Glasser JR, Zou C, Ellis B, Das T, et al. E3 ligase subunit Fbxo15 and PINK1 kinase regulate cardiolipin synthase 1 stability and mitochondrial function in pneumonia. Cell Rep. 2014;7(2):476–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  578. Shi J, Yu T, Song K, Du S, He S, Hu X, et al. Dexmedetomidine ameliorates endotoxin-induced acute lung injury in vivo and in vitro by preserving mitochondrial dynamic equilibrium through the HIF-1a/HO-1 signaling pathway. Redox Biol. 2021;41:101954.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  579. Ning L, Rui X, Guorui L, Tinglv F, Donghang L, Chenzhen X, et al. A novel mechanism for the protection against acute lung injury by melatonin: mitochondrial quality control of lung epithelial cells is preserved through SIRT3-dependent deacetylation of SOD2. Cell Mol Life Sci. 2022;79(12):610.

    Article  CAS  PubMed  Google Scholar 

  580. Wang Z, White A, Wang X, Ko J, Choudhary G, Lange T, et al. Mitochondrial fission mediated cigarette smoke-induced pulmonary endothelial injury. Am J Respir Cell Mol Biol. 2020;63(5):637–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  581. Athale J, Ulrich A, MacGarvey NC, Bartz RR, Welty-Wolf KE, Suliman HB, et al. Nrf2 promotes alveolar mitochondrial biogenesis and resolution of lung injury in Staphylococcus aureus pneumonia in mice. Free Radic Biol Med. 2012;53(8):1584–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  582. Liu Y, Lear TB, Verma M, Wang KZ, Otero PA, McKelvey AC, et al. Chemical inhibition of FBXO7 reduces inflammation and confers neuroprotection by stabilizing the mitochondrial kinase PINK1. JCI insight. 2020;5(11):e131834.

    Article  PubMed  PubMed Central  Google Scholar 

  583. Li N, Liu B, Xiong R, Li G, Wang B, Geng Q. HDAC3 deficiency protects against acute lung injury by maintaining epithelial barrier integrity through preserving mitochondrial quality control. Redox Biol. 2023;63:102746.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  584. Li G, Fu T, Wang W, Xiong R, Liu B, He R, et al. Pretreatment with Kahweol attenuates sepsis-induced acute lung injury via improving mitochondrial homeostasis in a CaMKKII/AMPK-dependent pathway. Mol Nutr Food Res. 2023;67(19):e2300083.

    Article  PubMed  Google Scholar 

  585. Thenappan T, Ormiston ML, Ryan JJ, Archer SL. Pulmonary arterial hypertension: pathogenesis and clinical management. BMJ. 2018;360:j5492.

    Article  PubMed  PubMed Central  Google Scholar 

  586. Ryan JJ, Marsboom G, Fang YH, Toth PT, Morrow E, Luo N, et al. PGC1α-mediated mitofusin-2 deficiency in female rats and humans with pulmonary arterial hypertension. Am J Respir Crit Care Med. 2013;187(8):865–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  587. Marsboom G, Toth PT, Ryan JJ, Hong Z, Wu X, Fang YH, et al. Dynamin-related protein 1-mediated mitochondrial mitotic fission permits hyperproliferation of vascular smooth muscle cells and offers a novel therapeutic target in pulmonary hypertension. Circ Res. 2012;110(11):1484–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  588. Feng W, Wang J, Yan X, Zhang Q, Chai L, Wang Q, et al. ERK/Drp1-dependent mitochondrial fission contributes to HMGB1-induced autophagy in pulmonary arterial hypertension. Cell Prolif. 2021;54(6):e13048.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  589. Colpman P, Dasgupta A, Archer SL. The role of mitochondrial dynamics and mitotic fission in regulating the cell cycle in cancer and pulmonary arterial hypertension: implications for dynamin-related protein 1 and mitofusin2 in hyperproliferative diseases. Cells. 2023;12(14):1897.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  590. Guder WG, Ross BD. Enzyme distribution along the nephron. Kidney Int. 1984;26(2):101–11.

    Article  CAS  PubMed  Google Scholar 

  591. Portilla D, Dai G, McClure T, Bates L, Kurten R, Megyesi J, et al. Alterations of PPARα and its coactivator PGC-1 in cisplatin-induced acute renal failure. Kidney Int. 2002;62(4):1208–18.

    Article  CAS  PubMed  Google Scholar 

  592. Tran MT, Zsengeller ZK, Berg AH, Khankin EV, Bhasin MK, Kim W, et al. PGC-1α drives NAD biosynthesis linking oxidative metabolism to renal protection. Nature. 2016;531(7595):528–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  593. Tran M, Tam D, Bardia A, Bhasin M, Rowe GC, Kher A, et al. PGC-1α promotes recovery after acute kidney injury during systemic inflammation in mice. J Clin Invest. 2011;121(10):4003–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  594. Jesinkey SR, Funk JA, Stallons LJ, Wills LP, Megyesi JK, Beeson CC, et al. Formoterol restores mitochondrial and renal function after ischemia-reperfusion injury. J Am Soc Nephrol. 2014;25(6):1157–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  595. Brooks C, Wei Q, Cho SG, Dong Z. Regulation of mitochondrial dynamics in acute kidney injury in cell culture and rodent models. J Clin Invest. 2009;119(5):1275–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  596. Wei Q, Sun H, Song S, Liu Y, Liu P, Livingston MJ, et al. MicroRNA-668 represses MTP18 to preserve mitochondrial dynamics in ischemic acute kidney injury. J Clin Invest. 2018;128(12):5448–64.

    Article  PubMed  PubMed Central  Google Scholar 

  597. Perry HM, Huang L, Wilson RJ, Bajwa A, Sesaki H, Yan Z, et al. Dynamin-related protein 1 deficiency promotes recovery from AKI. J Am Soc Nephrol. 2018;29(1):194–206.

    Article  CAS  PubMed  Google Scholar 

  598. Cho SG, Xiao X, Wang S, Gao H, Rafikov R, Black S, et al. Bif-1 interacts with prohibitin-2 to regulate mitochondrial inner membrane during cell stress and apoptosis. J Am Soc Nephrol. 2019;30(7):1174–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  599. Gall JM, Wang Z, Liesa M, Molina A, Havasi A, Schwartz JH, et al. Role of mitofusin 2 in the renal stress response. PLoS ONE. 2012;7(1):e31074.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  600. Gall JM, Wang Z, Bonegio RG, Havasi A, Liesa M, Vemula P, et al. Conditional knockout of proximal tubule mitofusin 2 accelerates recovery and improves survival after renal ischemia. J Am Soc Nephrol. 2015;26(5):1092–102.

    Article  CAS  PubMed  Google Scholar 

  601. Tang C, Han H, Liu Z, Liu Y, Yin L, Cai J, et al. Activation of BNIP3-mediated mitophagy protects against renal ischemia-reperfusion injury. Cell Death Dis. 2019;10(9):677.

    Article  PubMed  PubMed Central  Google Scholar 

  602. Wang Y, Tang C, Cai J, Chen G, Zhang D, Zhang Z, et al. PINK1/Parkin-mediated mitophagy is activated in cisplatin nephrotoxicity to protect against kidney injury. Cell Death Dis. 2018;9(11):113.

    Article  Google Scholar 

  603. Lynch MR, Tran MT, Ralto KM, Zsengeller ZK, Raman V, Bhasin SS, et al. TFEB-driven lysosomal biogenesis is pivotal for PGC1α-dependent renal stress resistance. JCI insight. 2019;5(8):e126749.

    Article  PubMed  Google Scholar 

  604. Sharma K, Karl B, Mathew AV, Gangoiti JA, Wassel CL, Saito R, et al. Metabolomics reveals signature of mitochondrial dysfunction in diabetic kidney disease. J Am Soc Nephrol. 2013;24(11):1901–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  605. Li SY, Park J, Qiu C, Han SH, Palmer MB, Arany Z, et al. Increasing the level of peroxisome proliferator-activated receptor γ coactivator-1α in podocytes results in collapsing glomerulopathy. JCI insight. 2017;2(14):e92930.

    Article  PubMed  PubMed Central  Google Scholar 

  606. Long J, Badal SS, Ye Z, Wang Y, Ayanga BA, Galvan DL, et al. Long noncoding RNA Tug1 regulates mitochondrial bioenergetics in diabetic nephropathy. J Clin Invest. 2016;126(11):4205–18.

    Article  PubMed  PubMed Central  Google Scholar 

  607. Dugan LL, You YH, Ali SS, Diamond-Stanic M, Miyamoto S, DeCleves AE, et al. AMPK dysregulation promotes diabetes-related reduction of superoxide and mitochondrial function. J Clin Invest. 2013;123(11):4888–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  608. Xiao L, Zhu X, Yang S, Liu F, Zhou Z, Zhan M, et al. Rap1 ameliorates renal tubular injury in diabetic nephropathy. Diabetes. 2014;63(4):1366–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  609. Zhan M, Usman IM, Sun L, Kanwar YS. Disruption of renal tubular mitochondrial quality control by Myo-inositol oxygenase in diabetic kidney disease. J Am Soc Nephrol. 2015;26(6):1304–21.

    Article  CAS  PubMed  Google Scholar 

  610. Galvan DL, Long J, Green N, Chang BH, Lin JS, Schumacker P, et al. Drp1S600 phosphorylation regulates mitochondrial fission and progression of nephropathy in diabetic mice. J Clin Invest. 2019;129(7):2807–23.

    Article  PubMed  PubMed Central  Google Scholar 

  611. Ayanga BA, Badal SS, Wang Y, Galvan DL, Chang BH, Schumacker PT, et al. Dynamin-related protein 1 deficiency improves mitochondrial fitness and protects against progression of diabetic nephropathy. J Am Soc Nephrol. 2016;27(9):2733–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  612. Qin X, Zhao Y, Gong J, Huang W, Su H, Yuan F, et al. Berberine protects glomerular podocytes via inhibiting Drp1-mediated mitochondrial fission and dysfunction. Theranostics. 2019;9(6):1698–713.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  613. Quan Y, Park W, Jin J, Kim W, Park SK, Kang KP. Sirtuin 3 activation by honokiol decreases unilateral ureteral obstruction-induced renal inflammation and fibrosis via regulation of mitochondrial dynamics and the renal NF-κBTGF-β1/Smad signaling pathway. Int J Mol Sci. 2020;21(2):402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  614. Wang Y, Lu M, Xiong L, Fan J, Zhou Y, Li H, et al. Drp1-mediated mitochondrial fission promotes renal fibroblast activation and fibrogenesis. Cell Death Dis. 2020;11(1):29.

    Article  PubMed  PubMed Central  Google Scholar 

  615. Noh MR, Woo CH, Park MJ, In Kim J, Park KM. Ablation of C/EBP homologous protein attenuates renal fibrosis after ureteral obstruction by reducing autophagy and microtubule disruption. Biochim Biophys Acta Mol Basis Dis. 2018;1864(5 Pt A):1634–41.

    Article  CAS  PubMed  Google Scholar 

  616. Xiao L, Xu X, Zhang F, Wang M, Xu Y, Tang D, et al. The mitochondria-targeted antioxidant MitoQ ameliorated tubular injury mediated by mitophagy in diabetic kidney disease via Nrf2/PINK1. Redox Biol. 2017;11:297–311.

    Article  CAS  PubMed  Google Scholar 

  617. Chen K, Dai H, Yuan J, Chen J, Lin L, Zhang W, et al. Optineurin-mediated mitophagy protects renal tubular epithelial cells against accelerated senescence in diabetic nephropathy. Cell Death Dis. 2018;9(2):105.

    Article  PubMed  PubMed Central  Google Scholar 

  618. Li S, Lin Q, Shao X, Zhu X, Wu J, Wu B, et al. Drp1-regulated PARK2-dependent mitophagy protects against renal fibrosis in unilateral ureteral obstruction. Free Radic Biol Med. 2020;152:632–49.

    Article  CAS  PubMed  Google Scholar 

  619. Thorgersen EB, Barratt-Due A, Haugaa H, Harboe M, Pischke SE, Nilsson PH, et al. The role of complement in liver injury, regeneration, and transplantation. Hepatology. 2019;70(2):725–36.

    Article  PubMed  Google Scholar 

  620. van Riel WG, van Golen RF, Reiniers MJ, Heger M, van Gulik TM. How much ischemia can the liver tolerate during resection? Hepatobiliary Surg Nutr. 2016;5(1):58–71.

    PubMed  PubMed Central  Google Scholar 

  621. Bi J, Zhang J, Ren Y, Du Z, Li Q, Wang Y, 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.

    Article  CAS  PubMed  Google Scholar 

  622. Dusabimana T, Kim SR, Kim HJ, Park SW, Kim H. Nobiletin ameliorates hepatic ischemia and reperfusion injury through the activation of SIRT-1/FOXO3a-mediated autophagy and mitochondrial biogenesis. Exp Mol Med. 2019;51(4):1–16.

    Article  CAS  PubMed  Google Scholar 

  623. Joe Y, Zheng M, Kim HJ, Uddin MJ, Kim SK, Chen Y, et al. Cilostazol attenuates murine hepatic ischemia and reperfusion injury via heme oxygenase-dependent activation of mitochondrial biogenesis. Am J Physiol Gastrointest Liver Physiol. 2015;309(1):G21–9.

    Article  CAS  PubMed  Google Scholar 

  624. Hou J, Tolbert E, Birkenbach M, Ghonem NS. Treprostinil alleviates hepatic mitochondrial injury during rat renal ischemia-reperfusion injury. Biomed Pharmacother. 2021;143:112172.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  625. 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(4):1174–92.

    Article  CAS  PubMed  Google Scholar 

  626. LaBrecque DR, Pesch LA. Preparation and partial characterization of hepatic regenerative stimulator substance (SS) from rat liver. J Physiol. 1975;248(2):273–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  627. Fuertes-Agudo M, Luque-Tévar M, Cucarella C, Brea R, Boscá L, Quintana-Cabrera R, et al. COX-2 expression in hepatocytes improves mitochondrial function after hepatic ischemia-reperfusion injury. Antioxidants (Basel). 2022;11(9):1724.

    Article  CAS  PubMed  Google Scholar 

  628. Zheng J, Chen L, Lu T, Zhang Y, Sui X, Li Y, et al. MSCs ameliorate hepatocellular apoptosis mediated by PINK1-dependent mitophagy in liver ischemia/reperfusion injury through AMPKα activation. Cell Death Dis. 2020;11(4):256.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  629. Kong WN, 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(1):130–43.

    Article  CAS  PubMed  Google Scholar 

  630. Rinella ME. Nonalcoholic fatty liver disease: a systematic review. JAMA. 2015;313(22):2263–73.

    Article  CAS  PubMed  Google Scholar 

  631. Brunt EM, Wong VW, Nobili V, Day CP, Sookoian S, Maher JJ, et al. Nonalcoholic fatty liver disease. Nat Rev Dis Primers. 2015;1:15080.

    Article  PubMed  Google Scholar 

  632. Moore MP, Cunningham RP, Meers GM, Johnson SA, Wheeler AA, Ganga RR, et al. Compromised hepatic mitochondrial fatty acid oxidation and reduced markers of mitochondrial turnover in human NAFLD. Hepatology. 2022;76(5):1452–65.

    Article  CAS  PubMed  Google Scholar 

  633. Besse-Patin A, Léveillé M, Oropeza D, Nguyen BN, Prat A, Estall JL. Estrogen signals through peroxisome proliferator-activated receptor-γ coactivator 1α to reduce oxidative damage associated with diet-induced fatty liver disease. Gastroenterology. 2017;152(1):243–56.

    Article  CAS  PubMed  Google Scholar 

  634. Wu L, Mo W, Feng J, Li J, Yu Q, Li S, et al. Astaxanthin attenuates hepatic damage and mitochondrial dysfunction in non-alcoholic fatty liver disease by up-regulating the FGF21/PGC-1α pathway. Br J Pharmacol. 2020;177(16):3760–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  635. Dusabimana T, Park EJ, Je J, Jeong K, Yun SP, Kim HJ, et al. P2Y2R deficiency ameliorates hepatic steatosis by reducing lipogenesis and enhancing fatty acid β-oxidation through AMPK and PGC-1α induction in high-fat diet-fed mice. Int J Mol Sci. 2021;22(11):5528.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  636. Krishnasamy Y, Gooz M, Li L, Lemasters JJ, Zhong Z. Role of mitochondrial depolarization and disrupted mitochondrial homeostasis in non-alcoholic steatohepatitis and fibrosis in mice. Int J Physiol Pathophysiol Pharmacol. 2019;11(5):190–204.

    CAS  PubMed  PubMed Central  Google Scholar 

  637. Galloway CA, Lee H, Brookes PS, Yoon Y. Decreasing mitochondrial fission alleviates hepatic steatosis in a murine model of nonalcoholic fatty liver disease. Am J Physiol Gastrointest Liver Physiol. 2014;307(6):G632–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  638. Zhou H, Du W, Li Y, Shi C, Hu N, Ma S, et al. Effects of melatonin on fatty liver disease: the role of NR4A1/DNA-PKcs/p53 pathway, mitochondrial fission, and mitophagy. J Pineal Res. 2018;64(1). https://doi.org/10.1111/jpi.12450.

  639. Hernández-Alvarez MI, Sebastián D, Vives S, Ivanova S, Bartoccioni P, Kakimoto P, et al. Deficient endoplasmic reticulum-mitochondrial phosphatidylserine transfer causes liver disease. Cell. 2019;177(4):881–95.e17.

    Article  PubMed  Google Scholar 

  640. Du J, Zhang X, Han J, Man K, Zhang Y, Chu ES, et al. Pro-inflammatory CXCR3 impairs mitochondrial function in experimental non-alcoholic steatohepatitis. Theranostics. 2017;7(17):4192–203.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  641. Dethlefsen MM, Kristensen CM, Tøndering AS, Lassen SB, Ringholm S, Pilegaard H. Impact of liver PGC-1α on exercise and exercise training-induced regulation of hepatic autophagy and mitophagy in mice on HFF. Physiol Rep. 2018;6(13):e13731.

    Article  PubMed  PubMed Central  Google Scholar 

  642. Xu JL, Li LY, Wang YQ, Li YQ, Shan M, Sun SZ, et al. Hepatocyte-specific deletion of BAP31 promotes SREBP1C activation, promotes hepatic lipid accumulation, and worsens IR in mice. J Lipid Res. 2018;59(1):35–47.

    Article  CAS  PubMed  Google Scholar 

  643. Glick D, Zhang W, Beaton M, Marsboom G, Gruber M, Simon MC, et al. BNip3 regulates mitochondrial function and lipid metabolism in the liver. Mol Cell Biol. 2012;32(13):2570–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  644. Williams JA, Ni HM, Ding Y, Ding WX. Parkin regulates mitophagy and mitochondrial function to protect against alcohol-induced liver injury and steatosis in mice. Am J Physiol Gastrointest Liver Physiol. 2015;309(5):G324–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  645. Gonçalves IO, Passos E, Diogo CV, Rocha-Rodrigues S, Santos-Alves E, Oliveira PJ, et al. Exercise mitigates mitochondrial permeability transition pore and quality control mechanisms alterations in nonalcoholic steatohepatitis. Appl Physiol Nutr Metab. 2016;41(3):298–306.

    Article  PubMed  Google Scholar 

  646. Li X, Shi Z, Zhu Y, Shen T, Wang H, Shui G, et al. Cyanidin-3-O-glucoside improves non-alcoholic fatty liver disease by promoting PINK1-mediated mitophagy in mice. Br J Pharmacol. 2020;177(15):3591–607.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  647. Ho GT, Theiss AL. Mitochondria and inflammatory bowel diseases: toward a stratified therapeutic intervention. Annu Rev Physiol. 2022;84:435–59.

    Article  CAS  PubMed  Google Scholar 

  648. D’Errico I, Salvatore L, Murzilli S, Lo Sasso G, Latorre D, Martelli N, et al. Peroxisome proliferator-activated receptor-γ coactivator 1-α (PGC1α) is a metabolic regulator of intestinal epithelial cell fate. Proc Natl Acad Sci U S A. 2011;108(16):6603–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  649. Cunningham KE, Vincent G, Sodhi CP, Novak EA, Ranganathan S, Egan CE, et al. Peroxisome proliferator-activated receptor-γ coactivator 1-α (PGC1α) protects against experimental murine colitis. J Biol Chem. 2016;291(19):10184–200.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  650. Haberman Y, Karns R, Dexheimer PJ, Schirmer M, Somekh J, Jurickova I, et al. Ulcerative colitis mucosal transcriptomes reveal mitochondriopathy and personalized mechanisms underlying disease severity and treatment response. Nat Commun. 2019;10(1):38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  651. Hou Y, Sun X, Gheinani PT, Guan X, Sharma S, Zhou Y, et al. Epithelial SMYD5 exaggerates IBD by down-regulating mitochondrial functions via post-translational control of PGC-1α stability. Cell Mol Gastroenterol Hepatol. 2022;14(2):375–403.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  652. Tian Q, Xu Z, Sun Q, Iniguez AB, Du M, Zhu MJ. Broccoli-derived glucoraphanin activates AMPK/PGC1α/NRF2 pathway and ameliorates dextran-sulphate-sodium-induced colitis in mice. Antioxidants (Basel). 2022;11(12):2404.

    Article  CAS  PubMed  Google Scholar 

  653. Astorga J, Gasaly N, Dubois-Camacho K, De la Fuente M, Landskron G, Faber KN, et al. The role of cholesterol and mitochondrial bioenergetics in activation of the inflammasome in IBD. Front Immunol. 2022;13:1028953.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  654. Qin Y, Yu Y, Yang C, Wang Z, Yang Y, Wang C, et al. Atractylenolide I inhibits NLRP3 inflammasome activation in colitis-associated colorectal cancer via suppressing Drp1-mediated mitochondrial fission. Front Pharmacol. 2021;12:674340.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  655. Vincent G, Novak EA, Siow VS, Cunningham KE, Griffith BD, Comerford TE, et al. Nix-mediated mitophagy modulates mitochondrial damage during intestinal inflammation. Antioxid Redox Signal. 2020;33(1):1–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  656. Lassen KG, Kuballa P, Conway KL, Patel KK, Becker CE, Peloquin JM, et al. Atg16L1 T300A variant decreases selective autophagy resulting in altered cytokine signaling and decreased antibacterial defense. Proc Natl Acad Sci U S A. 2014;111(21):7741–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  657. Zhang H, Zheng L, McGovern DP, Hamill AM, Ichikawa R, Kanazawa Y, et al. Myeloid ATG16L1 facilitates host-bacteria interactions in maintaining intestinal homeostasis. J Immunol. 2017;198(5):2133–46.

    Article  CAS  PubMed  Google Scholar 

  658. Guo W, Sun Y, Liu W, Wu X, Guo L, Cai P, et al. Small molecule-driven mitophagy-mediated NLRP3 inflammasome inhibition is responsible for the prevention of colitis-associated cancer. Autophagy. 2014;10(6):972–85.

    Article  PubMed  PubMed Central  Google Scholar 

  659. Kathiria AS, Butcher LD, Feagins LA, Souza RF, Boland CR, Theiss AL. Prohibitin 1 modulates mitochondrial stress-related autophagy in human colonic epithelial cells. PLoS ONE. 2012;7(2):e31231.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  660. Sorrentino V, Menzies KJ, Auwerx J. Repairing mitochondrial dysfunction in disease. Annu Rev Pharmacol Toxicol. 2018;58:353–89.

    Article  CAS  PubMed  Google Scholar 

  661. Li Y, Feng YF, Liu XT, Li YC, Zhu HM, Sun MR, et al. Songorine promotes cardiac mitochondrial biogenesis via Nrf2 induction during sepsis. Redox Biol. 2021;38:101771.

    Article  CAS  PubMed  Google Scholar 

  662. Tejero J, Shiva S, Gladwin MT. Sources of vascular nitric oxide and reactive oxygen species and their regulation. Physiol Rev. 2019;99(1):311–79.

    Article  CAS  PubMed  Google Scholar 

  663. Rawat PS, Jaiswal A, Khurana A, Bhatti JS, Navik U. Doxorubicin-induced cardiotoxicity: an update on the molecular mechanism and novel therapeutic strategies for effective management. Biomed Pharmacother. 2021;139:111708.

    Article  CAS  PubMed  Google Scholar 

  664. Beyfuss K, Hood DA. A systematic review of p53 regulation of oxidative stress in skeletal muscle. Redox Rep. 2018;23(1):100–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  665. Trevellin E, Scorzeto M, Olivieri M, Granzotto M, Valerio A, Tedesco L, et al. Exercise training induces mitochondrial biogenesis and glucose uptake in subcutaneous adipose tissue through eNOS-dependent mechanisms. Diabetes. 2014;63(8):2800–11.

    Article  CAS  PubMed  Google Scholar 

  666. Tengan CH, Kiyomoto BH, Godinho RO, Gamba J, Neves AC, Schmidt B, et al. The role of nitric oxide in muscle fibers with oxidative phosphorylation defects. Biochem Biophys Res Commun. 2007;359(3):771–7.

    Article  CAS  PubMed  Google Scholar 

  667. Carrera-Juliá S, Moreno ML, Barrios C, de la Rubia Ortí JE, Drehmer E. Antioxidant alternatives in the treatment of amyotrophic lateral sclerosis: a comprehensive review. Front Physiol. 2020;11:63.

    Article  PubMed  PubMed Central  Google Scholar 

  668. Rajman L, Chwalek K, Sinclair DA. Therapeutic Ppotential of NAD-boosting molecules: the in vivo evidence. Cell Metab. 2018;27(3):529–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  669. Cantó C, Houtkooper RH, Pirinen E, Youn DY, Oosterveer MH, Cen Y, et al. The NAD(+) precursor nicotinamide riboside enhances oxidative metabolism and protects against high-fat diet-induced obesity. Cell Metab. 2012;15(6):838–47.

    Article  PubMed  PubMed Central  Google Scholar 

  670. Schöndorf DC, Ivanyuk D, Baden P, Sanchez-Martinez A, De Cicco S, Yu C, et al. The NAD+ precursor nicotinamide riboside rescues mitochondrial defects and neuronal loss in iPSC and fly models of Parkinson’s disease. Cell Rep. 2018;23(10):2976–88.

    Article  PubMed  Google Scholar 

  671. Chuang YC, Chen SD, Hsu CY, Chen SF, Chen NC, Jou SB. Resveratrol promotes mitochondrial biogenesis and protects against seizure-induced neuronal cell damage in the hippocampus following status epilepticus by activation of the PGC-1α signaling pathway. Int J Mol Sci. 2019;20(4):998.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  672. Hang W, He B, Chen J, Xia L, Wen B, Liang T, et al. Berberine ameliorates high glucose-induced cardiomyocyte injury via AMPK signaling activation to stimulate mitochondrial biogenesis and restore autophagic flux. Front Pharmacol. 2018;9:1121.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  673. Yu Y, Zhao Y, Teng F, Li J, Guan Y, Xu J, et al. Berberine improves cognitive deficiency and muscular dysfunction via activation of the AMPK/SIRT1/PGC-1a pathway in skeletal muscle from naturally aging rats. J Nutr Health Aging. 2018;22(6):710–7.

    Article  CAS  PubMed  Google Scholar 

  674. Koshinaka K, Honda A, Masuda H, Sato A. Effect of quercetin treatment on mitochondrial biogenesis and exercise-induced AMP-activated protein kinase activation in rat skeletal muscle. Nutrients. 2020;12(3):729.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  675. de Oliveira MR, Jardim FR, Setzer WN, Nabavi SM, Nabavi SF. Curcumin, mitochondrial biogenesis, and mitophagy: exploring recent data and indicating future needs. Biotechnol Adv. 2016;34(5):813–26.

    Article  PubMed  Google Scholar 

  676. Shen W, Hao J, Feng Z, Tian C, Chen W, Packer L, et al. Lipoamide or lipoic acid stimulates mitochondrial biogenesis in 3T3-L1 adipocytes via the endothelial NO synthase-cGMP-protein kinase G signalling pathway. Br J Pharmacol. 2011;162(5):1213–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  677. Komen JC, Thorburn DR. Turn up the power - pharmacological activation of mitochondrial biogenesis in mouse models. Br J Pharmacol. 2014;171(8):1818–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  678. Inata Y, Kikuchi S, Samraj RS, Hake PW, O’Connor M, Ledford JR, et al. Autophagy and mitochondrial biogenesis impairment contribute to age-dependent liver injury in experimental sepsis: dysregulation of AMP-activated protein kinase pathway. FASEB J. 2018;32(2):728–41.

    Article  CAS  PubMed  Google Scholar 

  679. Ma WQ, Sun XJ, Wang Y, Zhu Y, Han XQ, Liu NF. Restoring mitochondrial biogenesis with metformin attenuates β-GP-induced phenotypic transformation of VSMCs into an osteogenic phenotype via inhibition of PDK4/oxidative stress-mediated apoptosis. Mol Cell Endocrinol. 2019;479:39–53.

    Article  CAS  PubMed  Google Scholar 

  680. Rangarajan S, Bone NB, Zmijewska AA, Jiang S, Park DW, Bernard K, et al. Metformin reverses established lung fibrosis in a bleomycin model. Nat Med. 2018;24(8):1121–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  681. Emelyanova L, Bai X, Yan Y, Bosnjak ZJ, Kress D, Warner C, et al. Biphasic effect of metformin on human cardiac energetics. Transl Res. 2021;229:5–23.

    Article  CAS  PubMed  Google Scholar 

  682. Arif E, Solanki AK, Srivastava P, Rahman B, Fitzgibbon WR, Deng P, et al. Mitochondrial biogenesis induced by the β2-adrenergic receptor agonist formoterol accelerates podocyte recovery from glomerular injury. Kidney Int. 2019;96(3):656–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  683. Vekaria HJ, Hubbard WB, Scholpa NE, Spry ML, Gooch JL, Prince SJ, et al. Formoterol, a β(2)-adrenoreceptor agonist, induces mitochondrial biogenesis and promotes cognitive recovery after traumatic brain injury. Neurobiol Dis. 2020;140:104866.

    Article  CAS  PubMed  Google Scholar 

  684. Cho YR, Lim JH, Kim MY, Kim TW, Hong BY, Kim YS, et al. Therapeutic effects of fenofibrate on diabetic peripheral neuropathy by improving endothelial and neural survival in db/db mice. PLoS One. 2014;9(1):e83204.

  685. Frambach SJCM, van de Wal MAE, van den Broek PHH, Smeitink JAM, Russel FGM, de Haas R, et al. Effects of clofibrate and KH176 on life span and motor function in mitochondrial complex I-deficient mice. Biochim Biophys Acta Mol Basis Dis. 2020;1866(6):165727.

    Article  CAS  PubMed  Google Scholar 

  686. Nanjan MJ, Mohammed M, Prashantha Kumar BR, Chandrasekar MJN. Thiazolidinediones as antidiabetic agents: a critical review. Bioorg Chem. 2018;77:548–67.

    Article  CAS  PubMed  Google Scholar 

  687. De Nuccio C, Bernardo A, Troiano C, Brignone MS, Falchi M, Greco A, et al. NRF2 and PPAR-γ pathways in oligodendrocyte progenitors: focus on ROS protection, mitochondrial biogenesis and promotion of cell differentiation. Int J Mol Sci. 2020;21(19):7216.

    Article  PubMed  PubMed Central  Google Scholar 

  688. Rodríguez-Pascau L, Britti E, Calap-Quintana P, Dong YN, Vergara C, Delaspre F, et al. PPAR gamma agonist leriglitazone improves frataxin-loss impairments in cellular and animal models of Friedreich Ataxia. Neurobiol Dis. 2021;148:105162.

    Article  PubMed  Google Scholar 

  689. Maréchal L, Laviolette M, Rodrigue-Way A, Sow B, Brochu M, Caron V, et al. The CD36-PPARγ pathway in metabolic disorders. Int J Mol Sci. 2018;19(5):1529.

    Article  PubMed  PubMed Central  Google Scholar 

  690. Tedesco L, Valerio A, Cervino C, Cardile A, Pagano C, Vettor R, et al. Cannabinoid type 1 receptor blockade promotes mitochondrial biogenesis through endothelial nitric oxide synthase expression in white adipocytes. Diabetes. 2008;57(8):2028–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  691. Cassidy-Stone A, Chipuk JE, Ingerman E, Song C, Yoo C, Kuwana T, et al. Chemical inhibition of the mitochondrial division dynamin reveals its role in Bax/Bak-dependent mitochondrial outer membrane permeabilization. Dev Cell. 2008;14(2):193–204.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  692. Gharanei M, Hussain A, Janneh O, Maddock H. Attenuation of doxorubicin-induced cardiotoxicity by mdivi-1: a mitochondrial division/mitophagy inhibitor. PLoS ONE. 2013;8(10):e77713.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  693. Wang W, Yin J, Ma X, Zhao F, Siedlak SL, Wang Z, et al. Inhibition of mitochondrial fragmentation protects against Alzheimer’s disease in rodent model. Hum Mol Genet. 2017;26(21):4118–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  694. He GW, Günther C, Kremer AE, Thonn V, Amann K, Poremba C, et al. PGAM5-mediated programmed necrosis of hepatocytes drives acute liver injury. Gut. 2017;66(4):716–23.

    Article  CAS  PubMed  Google Scholar 

  695. Wu P, Li Y, Zhu S, Wang C, Dai J, Zhang G, et al. Mdivi-1 alleviates early brain injury after experimental subarachnoid hemorrhage in rats, possibly via inhibition of Drp1-activated mitochondrial fission and oxidative stress. Neurochem Res. 2017;42(5):1449–58.

    Article  CAS  PubMed  Google Scholar 

  696. Gonzalez AS, Elguero ME, Finocchietto P, Holod S, Romorini L, Miriuka SG, et al. Abnormal mitochondrial fusion-fission balance contributes to the progression of experimental sepsis. Free Radic Res. 2014;48(7):769–83.

    Article  CAS  PubMed  Google Scholar 

  697. Cheng Z, Jiang X, Pansuria M, Fang P, Mai J, Mallilankaraman K, et al. Hyperhomocysteinemia and hyperglycemia induce and potentiate endothelial dysfunction via μ-calpain activation. Diabetes. 2015;64(3):947–59.

    Article  CAS  PubMed  Google Scholar 

  698. Lei X, Lin H, Wang J, Ou Z, Ruan Y, Sadagopan A, et al. Mitochondrial fission induces immunoescape in solid tumors through decreasing MHC-I surface expression. Nat Commun. 2022;13(1):3882.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  699. Qi X, Qvit N, Su YC, Mochly-Rosen D. A novel Drp1 inhibitor diminishes aberrant mitochondrial fission and neurotoxicity. J Cell Sci. 2013;126(Pt 3):789–802.

    CAS  PubMed  PubMed Central  Google Scholar 

  700. Filichia E, Hoffer B, Qi X, Luo Y. Inhibition of Drp1 mitochondrial translocation provides neural protection in dopaminergic system in a Parkinson’s disease model induced by MPTP. Sci Rep. 2016;6:32656.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  701. Tian L, Neuber-Hess M, Mewburn J, Dasgupta A, Dunham-Snary K, Wu D, et al. Ischemia-induced Drp1 and Fis1-mediated mitochondrial fission and right ventricular dysfunction in pulmonary hypertension. J Mol Med (Berl). 2017;95(4):381–93.

    Article  CAS  PubMed  Google Scholar 

  702. Mancini NL, Goudie L, Xu W, Sabouny R, Rajeev S, Wang A, et al. Perturbed mitochondrial dynamics is a novel feature of colitis that can be targeted to lessen disease. Cell Mol Gastroenterol Hepatol. 2020;10(2):287–307.

    Article  PubMed  PubMed Central  Google Scholar 

  703. Haileselassie B, Mukherjee R, Joshi AU, Napier BA, Massis LM, Ostberg NP, et al. Drp1/Fis1 interaction mediates mitochondrial dysfunction in septic cardiomyopathy. J Mol Cell Cardiol. 2019;130:160–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  704. Haileselassie B, Joshi AU, Minhas PS, Mukherjee R, Andreasson KI, Mochly-Rosen D. Mitochondrial dysfunction mediated through dynamin-related protein 1 (Drp1) propagates impairment in blood brain barrier in septic encephalopathy. J Neuroinflammation. 2020;17(1):36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  705. Mukherjee R, Tompkins CA, Ostberg NP, Joshi AU, Massis LM, Vijayan V, et al. Drp1/Fis1-dependent pathologic fission and associated damaged extracellular mitochondria contribute to macrophage dysfunction in endotoxin tolerance. Crit Care Med. 2022;50(6):e504–15.

    Article  PubMed  PubMed Central  Google Scholar 

  706. Ferreira JCB, Campos JC, Qvit N, Qi X, Bozi LHM, Bechara LRG, et al. A selective inhibitor of mitofusin 1-βIIPKC association improves heart failure outcome in rats. Nat Commun. 2019;10(1):329.

    Article  PubMed  PubMed Central  Google Scholar 

  707. Horvath O, Ordog K, Bruszt K, Kalman N, Kovacs D, Radnai B, et al. Modulation of mitochondrial quality control processes by BGP-15 in oxidative stress scenarios: from cell culture to heart failure. Oxid Med Cell Longev. 2021;2021:6643871.

    Article  PubMed  PubMed Central  Google Scholar 

  708. Szabo A, Sumegi K, Fekete K, Hocsak E, Debreceni B, Setalo G Jr, et al. Activation of mitochondrial fusion provides a new treatment for mitochondria-related diseases. Biochem Pharmacol. 2018;150:86–96.

    Article  CAS  PubMed  Google Scholar 

  709. Ding M, Liu C, Shi R, Yu M, Zeng K, Kang J, et al. Mitochondrial fusion promoter restores mitochondrial dynamics balance and ameliorates diabetic cardiomyopathy in an optic atrophy 1-dependent way. Acta Physiol (Oxf). 2020;229(1):e13428.

    Article  CAS  PubMed  Google Scholar 

  710. Maneechote C, Palee S, Kerdphoo S, Jaiwongkam T, Chattipakorn SC, Chattipakorn N. Modulating mitochondrial dynamics attenuates cardiac ischemia-reperfusion injury in prediabetic rats. Acta Pharmacol Sin. 2022;43(1):26–38.

    Article  CAS  PubMed  Google Scholar 

  711. Yu LM, Dong X, Xue XD, Xu S, Zhang X, Xu YL, et al. Melatonin attenuates diabetic cardiomyopathy and reduces myocardial vulnerability to ischemia-reperfusion injury by improving mitochondrial quality control: role of SIRT6. J Pineal Res. 2021;70(1):e12698.

    Article  CAS  PubMed  Google Scholar 

  712. Ding M, Feng N, Tang D, Feng J, Li Z, Jia M, et al. Melatonin prevents Drp1-mediated mitochondrial fission in diabetic hearts through SIRT1-PGC1α pathway. J Pineal Res. 2018;65(2):e12491.

    Article  PubMed  PubMed Central  Google Scholar 

  713. Wei N, Pu Y, Yang Z, Pan Y, Liu L. Therapeutic effects of melatonin on cerebral ischemia reperfusion injury: Role of Yap-OPA1 signaling pathway and mitochondrial fusion. Biomed Pharmacother. 2019;110:203–12.

    Article  CAS  PubMed  Google Scholar 

  714. Mehrzadi S, Karimi MY, Fatemi A, Reiter RJ, Hosseinzadeh A. SARS-CoV-2 and other coronaviruses negatively influence mitochondrial quality control: beneficial effects of melatonin. Pharmacol Ther. 2021;224:107825.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  715. Kang L, Liu S, Li J, Tian Y, Xue Y, Liu X. The mitochondria-targeted anti-oxidant MitoQ protects against intervertebral disc degeneration by ameliorating mitochondrial dysfunction and redox imbalance. Cell Prolif. 2020;53(3):e12779.

    Article  PubMed  PubMed Central  Google Scholar 

  716. Zheng M, Bai Y, Sun X, Fu R, Liu L, Liu M, et al. Resveratrol reestablishes mitochondrial quality control in myocardial ischemia/reperfusion injury through Sirt1/Sirt3-Mfn2-Parkin-PGC-1α pathway. Molecules. 2022;27(17):5545.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  717. Ajoolabady A, Chiong M, Lavandero S, Klionsky DJ, Ren J. Mitophagy in cardiovascular diseases: molecular mechanisms, pathogenesis, and treatment. Trends Mol Med. 2022;28(10):836–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  718. Fang EF, Hou Y, Palikaras K, Adriaanse BA, Kerr JS, Yang B, et al. Mitophagy inhibits amyloid-β and tau pathology and reverses cognitive deficits in models of Alzheimer’s disease. Nat Neurosci. 2019;22(3):401–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  719. Ryu D, Mouchiroud L, Andreux PA, Katsyuba E, Moullan N, Nicolet-Dit-Félix AA, et al. Urolithin A induces mitophagy and prolongs lifespan in C. elegans and increases muscle function in rodents. Nat Med. 2016;22(8):879–88.

    Article  CAS  PubMed  Google Scholar 

  720. Liu S, D’Amico D, Shankland E, Bhayana S, Garcia JM, Aebischer P, et al. Effect of urolithin A supplementation on muscle endurance and mitochondrial health in older adults: a randomized clinical trial. JAMA Netw Open. 2022;5(1):e2144279.

    Article  PubMed  PubMed Central  Google Scholar 

  721. Huang JR, Zhang MH, Chen YJ, Sun YL, Gao ZM, Li ZJ, et al. Urolithin A ameliorates obesity-induced metabolic cardiomyopathy in mice via mitophagy activation. Acta Pharmacol Sin. 2023;44(2):321–31.

    Article  CAS  PubMed  Google Scholar 

  722. Wang Y, Jasper H, Toan S, Muid D, Chang X, Zhou H. Mitophagy coordinates the mitochondrial unfolded protein response to attenuate inflammation-mediated myocardial injury. Redox Biol. 2021;45:102049.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  723. Yang X, Zhang M, Dai Y, Sun Y, Aman Y, Xu Y, et al. Spermidine inhibits neurodegeneration and delays aging via the PINK1-PDR1-dependent mitophagy pathway in C. elegans. Aging (Albany NY). 2020;12(17):16852–66.

    Article  CAS  PubMed  Google Scholar 

  724. Tyrrell DJ, Blin MG, Song J, Wood SC, Zhang M, Beard DA, et al. Age-associated mitochondrial dysfunction accelerates atherogenesis. Circ Res. 2020;126(3):298–314.

    Article  CAS  PubMed  Google Scholar 

  725. Bhansali S, Bhansali A, Dhawan V. Metformin promotes mitophagy in mononuclear cells: a potential in vitro model for unraveling metformin’s mechanism of action. Ann N Y Acad Sci. 2020;1463(1):23–36.

    Article  CAS  PubMed  Google Scholar 

  726. Bhansali S, Bhansali A, Dutta P, Walia R, Dhawan V. Metformin upregulates mitophagy in patients with T2DM: A randomized placebo-controlled study. J Cell Mol Med. 2020;24(5):2832–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  727. Zhang Y, Zhang T, Li Y, Guo Y, Liu B, Tian Y, et al. Metformin attenuates early brain injury after subarachnoid hemorrhage in rats via AMPK-dependent mitophagy. Exp Neurol. 2022;353:114055.

    Article  CAS  PubMed  Google Scholar 

  728. Han YC, Tang SQ, Liu YT, Li AM, Zhan M, Yang M, et al. AMPK agonist alleviate renal tubulointerstitial fibrosis via activating mitophagy in high fat and streptozotocin induced diabetic mice. Cell Death Dis. 2021;12(10):925.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  729. Wang S, Zhao Z, Feng X, Cheng Z, Xiong Z, Wang T, et al. Melatonin activates Parkin translocation and rescues the impaired mitophagy activity of diabetic cardiomyopathy through Mst1 inhibition. J Cell Mol Med. 2018;22(10):5132–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  730. Chen C, Yang C, Wang J, Huang X, Yu H, Li S, et al. Melatonin ameliorates cognitive deficits through improving mitophagy in a mouse model of Alzheimer’s disease. J Pineal Res. 2021;71(4):e12774.

    Article  CAS  PubMed  Google Scholar 

  731. Kang JW, Hong JM, Lee SM. Melatonin enhances mitophagy and mitochondrial biogenesis in rats with carbon tetrachloride-induced liver fibrosis. J Pineal Res. 2016;60(4):383–93.

    Article  CAS  PubMed  Google Scholar 

  732. Lin C, Chao H, Li Z, Xu X, Liu Y, Hou L, et al. Melatonin attenuates traumatic brain injury-induced inflammation: a possible role for mitophagy. J Pineal Res. 2016;61(2):177–86.

    Article  CAS  PubMed  Google Scholar 

  733. Cao S, Wang C, Yan J, Li X, Wen J, Hu C. Curcumin ameliorates oxidative stress-induced intestinal barrier injury and mitochondrial damage by promoting Parkin dependent mitophagy through AMPK-TFEB signal pathway. Free Radic Biol Med. 2020;147:8–22.

    Article  CAS  PubMed  Google Scholar 

  734. Chang X, Zhang T, Meng Q, ShiyuanWang, Yan P, Wang X, et al. Quercetin improves cardiomyocyte vulnerability to hypoxia by regulating SIRT1/TMBIM6-related mitophagy and endoplasmic reticulum stress. Oxid Med Cell Longev. 2021;2021:5529913.

    Article  PubMed  PubMed Central  Google Scholar 

  735. Wang WW, Han R, He HJ, Li J, Chen SY, Gu Y, et al. Administration of quercetin improves mitochondria quality control and protects the neurons in 6-OHDA-lesioned Parkinson’s disease models. Aging (Albany NY). 2021;13(8):11738–51.

    Article  CAS  PubMed  Google Scholar 

  736. Pan T, Zhou Q, Miao K, Zhang L, Wu G, Yu J, et al. Suppressing Sart1 to modulate macrophage polarization by siRNA-loaded liposomes: a promising therapeutic strategy for pulmonary fibrosis. Theranostics. 2021;11(3):1192–206.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  737. Chang X, Zhang T, Liu D, Meng Q, Yan P, Luo D, et al. Puerarin attenuates LPS-induced inflammatory responses and oxidative stress injury in human umbilical vein endothelial cells through mitochondrial quality control. Oxid Med Cell Longev. 2021;2021:6659240.

    Article  PubMed  PubMed Central  Google Scholar 

  738. Chen X, Yi L, Song S, Wang L, Liang Q, Wang Y, et al. Puerarin attenuates palmitate-induced mitochondrial dysfunction, impaired mitophagy and inflammation in L6 myotubes. Life Sci. 2018;206:84–92.

    Article  CAS  PubMed  Google Scholar 

  739. Memme JM, Erlich AT, Phukan G, Hood DA. Exercise and mitochondrial health. J Physiol. 2021;599(3):803–17.

    Article  CAS  PubMed  Google Scholar 

  740. Campos JC, Queliconi BB, Bozi LHM, Bechara LRG, Dourado PMM, Andres AM, et al. Exercise reestablishes autophagic flux and mitochondrial quality control in heart failure. Autophagy. 2017;13(8):1304–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  741. Zhao D, Sun Y, Tan Y, Zhang Z, Hou Z, Gao C, et al. Short-duration swimming exercise after myocardial infarction attenuates cardiac dysfunction and regulates mitochondrial quality control in aged mice. Oxid Med Cell Longev. 2018;2018:4079041.

    Article  PubMed  PubMed Central  Google Scholar 

  742. Romanello V, Sandri M. The connection between the dynamic remodeling of the mitochondrial network and the regulation of muscle mass. Cell Mol Life Sci. 2021;78(4):1305–28.

    Article  CAS  PubMed  Google Scholar 

  743. Thai PN, Miller CV, King MT, Schaefer S, Veech RL, Chiamvimonvat N, et al. Ketone ester D-β-hydroxybutyrate-(R)-1,3 butanediol prevents decline in cardiac function in type 2 diabetic mice. J Am Heart Assoc. 2021;10(19):e020729.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  744. Nouri K, Feng Y, Schimmer AD. Mitochondrial ClpP serine protease-biological function and emerging target for cancer therapy. Cell Death Dis. 2020;11(10):841.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  745. Prabhu VV, Morrow S, Rahman Kawakibi A, Zhou L, Ralff M, Ray J, et al. ONC201 and imipridones: anti-cancer compounds with clinical efficacy. Neoplasia. 2020;22(12):725–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  746. Choi SE, Hwang Y, Lee SJ, Jung H, Shin TH, Son Y, et al. Mitochondrial protease ClpP supplementation ameliorates diet-induced NASH in mice. J Hepatol. 2022;77(3):735–47.

    Article  CAS  PubMed  Google Scholar 

  747. Jabbari H, Roushandeh AM, Rostami MK, Razavi-Toosi MT, Shokrgozar MA, Jahanian-Najafabadi A, et al. Mitochondrial transplantation ameliorates ischemia/reperfusion-induced kidney injury in rat. Biochim Biophys Acta Mol Basis Dis. 2020;1866(8):165809.

    Article  CAS  PubMed  Google Scholar 

  748. Masuzawa A, Black KM, Pacak CA, Ericsson M, Barnett RJ, Drumm C, et al. Transplantation of autologously derived mitochondria protects the heart from ischemia-reperfusion injury. Am J Physiol Heart Circ Physiol. 2013;304(7):H966–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  749. Moskowitzova K, Orfany A, Liu K, Ramirez-Barbieri G, Thedsanamoorthy JK, Yao R, et al. Mitochondrial transplantation enhances murine lung viability and recovery after ischemia-reperfusion injury. Am J Physiol Lung Cell Mol Physiol. 2020;318(1):L78–l88.

    Article  CAS  PubMed  Google Scholar 

  750. Park A, Oh M, Lee SJ, Oh KJ, Lee EW, Lee SC, et al. Mitochondrial transplantation as a novel therapeutic strategy for mitochondrial diseases. Int J Mol Sci. 2021;22(9):4793.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  751. Nunnari J, Suomalainen A. Mitochondria: in sickness and in health. Cell. 2012;148(6):1145–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the National Natural Science Foundation of China (8210082163, 81800343), the Fundamental Research Funds for the Central Universities (2042021kf0081), the Science Fund for Creative Research Groups of the Natural Science Foundation of Hubei Province (2020CFA027), and the Doctor of excellence program of The First Hospital of Jilin University (JDYY-DEP-2023043).

Author information

Authors and Affiliations

Authors

Contributions

QG and NL contributed to conception and were responsible for the whole work. BH L, CZ X and YL completed the writing and figure making of this review. ZLL, TLF, GRL, YD, GQL and SD provided ideas and assistance. All authors read and approval the final manuscript.

Corresponding authors

Correspondence to Ning Li or Qing Geng.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, BH., Xu, CZ., Liu, Y. et al. Mitochondrial quality control in human health and disease. Military Med Res 11, 32 (2024). https://doi.org/10.1186/s40779-024-00536-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40779-024-00536-5

Keywords