Next Article in Journal
The In Vitro Antioxidant and Immunomodulatory Effects of the Irish Monofloral Ivy and Heather Honey Varieties
Previous Article in Journal
Therapeutic Potential of Mesenchymal Stem Cell and Tenocyte Secretomes for Tendon Repair: Proteomic Profiling and Functional Characterization In Vitro and In Ovo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of N6-Methyladenosine in Mitochondrial Dysfunction and Pathology

School of Life Science, Inner Mongolia University, Hohhot 010020, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2025, 26(8), 3624; https://doi.org/10.3390/ijms26083624
Submission received: 28 February 2025 / Revised: 1 April 2025 / Accepted: 9 April 2025 / Published: 11 April 2025
(This article belongs to the Section Molecular Biology)

Abstract

Mitochondria are indispensable in cells and play crucial roles in maintaining cellular homeostasis, energy production, and regulating cell death. Mitochondrial dysfunction has various manifestations, causing different diseases by affecting the diverse functions of mitochondria in the body. Previous studies have mainly focused on mitochondrial-related diseases caused by nuclear gene mutations or mitochondrial gene mutations, or mitochondrial dysfunction resulting from epigenetic regulation, such as DNA and histone modification. In recent years, as a popular research area, m6A has been involved in a variety of important processes under physiological and pathological conditions. However, there are few summaries on how RNA methylation, especially m6A RNA methylation, affects mitochondrial function. Additionally, the role of m6A in pathology through influencing mitochondrial function may provide us with a new perspective on disease treatment. In this review, we summarize several manifestations of mitochondrial dysfunction and compile examples from recent years of how m6A affects mitochondrial function and its role in some diseases.

1. Introduction

Mitochondria, as dynamic organelles with a double-membrane structure, perform various functions in cells, such as maintaining various cellular homeostases, generating energy through oxidative phosphorylation, and regulating cell death [1]. Mitochondrial dysfunction directly disrupts core cellular processes, such as energy metabolism, redox balance, and apoptosis regulation, serving as a central driver of diverse pathologies. For example, mitochondrial dysfunction is considered one of the major causes of cardiovascular diseases, with factors such as an impaired mitophagy, reduced oxidative phosphorylation capacity, and increased production of reactive oxygen species (ROS) all contributing to their development [2]. Mitochondrial dysfunction, such as damage to the electron transport chain, the increased production of ROS, and alterations in mitochondrial dynamics, is a prominent feature of neurodegenerative diseases [3]. In addition, changes in mitochondrial fusion and fission can also lead to the occurrence of tumors and affect tumor metabolism and proliferation [4]. In summary, it can be found that the phenomenon of mitochondrial dysfunction occurs in a variety of diseases.
As an important part of epigenetics, RNA methylation has so far discovered over 150 types of RNA chemical modifications. Among them, methylation at the N6 position of adenine, namely N6-methyladenosine (m6A), which was first discovered in mRNA in 1974, is the most abundant RNA modification in eukaryotes [5]. Studies have found that m6A is enriched around stop codons, in the 3′ untranslated regions (3′UTRs), and near long internal exons, accounting for approximately 0.1% to 0.4% of the total adenosine residues in cellular RNA [6]. M6A is involved in regulating nearly every aspect of RNA biology, including RNA splicing, export, stability, and translation [7].
As shown in Figure 1, m6A methylation is a dynamic process that can be assembled by the methyltransferase complex (MTC) and removed by demethylases [8]. The regulators involved in this dynamic process can be classified into three categories: methyltransferase complex (writers), demethylases (erasers), and RNA-binding proteins (readers) [9].
METTL3, METTL14, and WTAP, as core components of the methyltransferase complex, play a major role in mediating m6A deposition on RNA through the stable heterodimer formed by METTL3 and METTL14. WTAP affects m6A deposition by interacting with the METTL3 METTL14 complex, but it does not have methylation activity itself [10]. Additionally, METTL16, as a new writer, targets pre-mRNAs, lncRNAs, and ncRNAs and methylates A43 of U6-snRNA [11].
FTO and ALKBH5, as two demethylases, can oxidatively remove m6A modifications on RNA (Table 1). FTO, as a member of the non-heme Fe(II) and α-KG-dependent dioxygenase AlkB family proteins, is located in the nucleus and cytoplasm of different mammalian cells. It has different structural features from other AlkB family proteins, including an additional C-terminal domain. The structural features of FTO enable it to demethylate different RNA substrates by binding to different partner proteins in different cellular environments [12,13]. Unlike FTO, ALKBH5 is only responsible for the removal of m6A on ssRNAs and is mainly located in the nucleus. Due to its localization in nuclear speckles that facilitate the assembly of mRNA-processing factors, it is speculated that nascent nuclear RNA is the main substrate of ALKBH5 [14,15].
Readers can recognize m6A and bind to it, playing a crucial role in determining the fate of target RNAs. These readers include YT521-B homology (YTH) domain-containing protein families, IGF2 mRNA-binding protein (IGF2BP) families, heterogeneous nuclear ribonucleoprotein (HNRNP) protein families, and eukaryotic initiation factor (eIF) 3 [16]. YTHDF2 plays a major role in RNA degradation, which induces transcript degradation by selectively recognizing and binding to m6A-modified mRNA and locating the bound mRNA to mRNA decay sites [17]. YTHDF1 enhances the translation efficiency of m6A-modified mRNA by interacting with initiation factors [18]. YTHDF3 cooperates with YTHDF1 to promote RNA translation and affects the YTHDF2-mediated degradation of methylated mRNA. As a partner of YTHDF1 and YTHDF2, YTHDF3 significantly influences their functions [19]. YTHDC1, as a nuclear reader, participates in alternative splicing [20]. Compared to other family members, YTHDC2 is the only one that contains helicase domains and multiple RNA-binding domains, and it has multiple functions, such as potentially increasing the translation efficiency of its target RNA and playing a crucial role in fertility [21,22]. The YTH-domain-containing proteins are widely involved in post-transcriptional regulation, participating in the regulation of target mRNA splicing, promoting translation, or promoting RNA degradation [23].
IGF2BP promotes the stability of target mRNA by inhibiting the degradation of m6A-containing mRNA and promoting its translation, as well as enhancing mRNA storage [24]. HNRNPC, as a nuclear RNA-binding protein, binds to nascent RNA transcripts by recognizing m6A sites, thereby affecting the stability, alternative splicing, and translation of target pre-mRNA [25]. HNRNPA2B1, together with METTL3, regulates the alternative splicing of target RNAs and the processing of nuclear pri-miRNA [26]. Eukaryotic initiation factor 3 (eIF3) promotes m6A-mediated cap-independent translation [27].
M6A RNA methylation is widely involved in the regulation of physiological and pathological processes, including cancer. Multiple studies have shown that m6A is involved in the regulation of tumors, but there is still no definite conclusion on how m6A affects mitochondrial function.
Table 1. The function of m6A regulators in RNA metabolism.
Table 1. The function of m6A regulators in RNA metabolism.
CategoryFactorFull NameFunctionReference
WritersMETTL3Methyltransferase-like 3As the catalytic core of the m6A methyltransferase complex, it catalyzes m6A modification.[28]
METTL14Methyltransferase-like 14Acts as an RNA binding platform within the m6A methyltransferase complex and forms a heterodimer with METTL3 to catalyze m6A modification.[5,28]
WTAPWilms tumor 1- associated proteinAs a regulatory subunit in the m6A methyltransferase complex, it interacts with METTL3 and METTL14 and directs them to nuclear speckles.[29,30]
METTL16Methyltransferase-like 16A single-component methyltransferase that can methylate U6-snRNA, the MALAT1 long non-coding RNA, and the MAT2A pre-mRNA.[31,32]
WritersVIRMA
(KIAA1429)
Vir-like m6A methyltransferase-associatedRecruits the core components of the methyltransferase METTL3/METTL14/WTAP to specific regions for selective methylation.[33]
RBM15/
RBM15B
RNA binding motif protein 15/ RNA binding motif protein 15BBinds to the m6A methyltransferase complex and guides it to specific RNA sites.[34]
ZC3H13Zinc finger CCCH-type containing 13Controls the nuclear localization of the Zc3h13–WTAP–Virilizer–Hakai complex and promotes m6A methylation.[35]
HAKAI
(CBLL1)
Cbl Proto-Oncogene Like 1Maintains the stability of the m6A methyltransferase complex.[36]
ErasersFTOFat mass and obesity-associatedRemoves m6A modification and promotes mRNA splicing.[37]
ALKBH5AlkB homolog 5Removes m6A modification on nuclear RNA, thereby regulating nuclear mRNA export.[14]
ReadersYTHDF1YTH N6-methyladenosine RNA-binding protein 1Promotes the translation initiation of m6A-modified mRNA.[18]
YTHDF2YTH N6-methyladenosine RNA-binding protein 2Promotes the degradation of m6A-modified mRNA.[17]
YTHDF3YTH N6-methyladenosine RNA-binding protein 3As a partner of YTHDF1 and YTHDF2, it interacts with YTHDF1/YTHDF2 to promote mRNA translation or degradation.[19]
YTHDC1YTH domain containing 1Promotes mRNA splicing.[20]
YTHDC2YTH domain containing 2Increases the translation efficiency of the target mRNA.[21,22]
IGF2BP1/2/3Insulin-like growth factor 2 mRNA-binding protein 1/2/3Promotes the stability of the target mRNA and facilitates its translation.[24]
HNRNPCHeterogeneous nuclear ribonucleoprotein CAffects the stability, alternative splicing, and translation of pre-mRNA.[25]
HNRNPA2B1Heterogeneous nuclear ribonucleoprotein A2/B1Promotes the processing of nuclear pri-miRNA and mRNA splicing.[26]
eIF3Eukaryotic initiation factor 3Promotes m6A-mediated cap-independent translation.[27]

2. Mitochondrial Function and Dysregulation

2.1. Mitochondrial Structure and Function

The α-proteobacterium engulfed by the precursor of modern eukaryotic cells is the ancestor of mitochondria [38]. During the evolution of bacteria into mitochondria, a portion of the genomic material was retained, known as mitochondrial DNA (mtDNA), which replicates in a semi-conservative manner and encodes some essential RNAs and proteins [39]. Mitochondria are organelles with a double-membrane structure consisting of four domains: the outer membrane, the inner membrane, the intermembrane space, and the mitochondrial matrix. The inner membrane folds inward to form cristae, which are the main site for oxidative phosphorylation [40]. Large biomolecules are degraded through the tricarboxylic acid (TCA) cycle and β-oxidation to produce NADH and FADH2. NADH and FADH2 dehydrogenate to form high-energy electrons and protons. The high-energy electrons are transferred along the electron transport chain (ETC) to O2 and gradually release energy, while protons are directionally transferred across the inner mitochondrial membrane (IMM) from the matrix side to the intermembrane space, finally forming a transmembrane potential difference and pH difference, known as the proton motive force, which then drives the rotational catalysis of ATP synthase to generate ATP. This process is called oxidative phosphorylation (OXPHOS) [41,42,43,44]. Mitochondria are multifunctional organelles. Besides OXPHOS, they also participate in amino acid metabolism, lipid synthesis, lipid oxidation, ROS generation, redox homeostasis, signal transduction, and apoptosis. These functions are crucial for maintaining multicellular life [45]. Mitochondrial dysfunction leads to reduced ATP synthesis, oxidative stress, inflammatory responses, and increased ROS levels, and may trigger various diseases, including neurodegenerative diseases, cardiomyopathy, metabolic syndrome, cancer, and obesity [46,47,48,49]. The causes of mitochondrial dysfunction include impaired mitochondrial biogenesis (Figure 2), increased ROS production, defective mitophagy, transport impairment, ETC dysfunction, changes in mitochondrial dynamics, and calcium imbalance or a combination of these factors [50].

2.2. Manifestations of Mitochondrial Dysfunction

2.2.1. Abnormal Mitochondrial Energy Metabolism and Impaired Biogenesis

Mitochondria can generate ATP through OXPHOS, as described in the process of OXPHOS outlined above. Mitochondrial diseases are caused by mitochondrial dysfunction resulting from nuclear gene mutations or mitochondrial gene mutations, and they are some of the most common genetic diseases. Their main feature is mitochondrial OXPHOS deficiency [51]. Only a small number of the proteins in mitochondria are encoded by mtDNA, which are the core components of OXPHOS complexes in the IMM, including seven polypeptides (ND1, ND2, ND3, ND4, ND4L, ND5, and ND6) in complex I, one polypeptide (cytochrome b) in complex III, three polypeptides (COI, COII, and COIII) in complex IV, and two polypeptides (ATP6 and ATP8) in complex V, totaling thirteen mitochondrial proteins. In addition, mtDNA also encodes 22 mt-tRNAs and 2 mt-rRNAs [52,53]. Approximately 99% of mitochondrial proteins are encoded by nuclear genes [53]. Moreover, the absence of or defects in enzymes catalyzing different steps in the ETC can also lead to mitochondrial dysfunction.
The electrochemical gradient (membrane potential) of the IMM that accompanies OXPHOS is a fundamental property of mitochondria and a major driving force for protein transport [54]. The steady state of the mitochondrial membrane potential is an important prerequisite for mitochondrial health and the maintenance of normal cellular and tissue functions [55].
Mitochondrial biogenesis (MB) is a complex and multifaceted regeneration process that controls mitochondrial self-renewal, replacing damaged mitochondria with healthy ones, and maintaining mtDNA to promote cellular homeostasis. The main events include the mitochondrial fusion, reduction in ROS, restoration of mitochondrial membrane potential, and increased expression of OXPHOS proteins. Impaired MB can lead to mitochondrial and cellular dysfunction [56,57]. Many studies have shown that peroxisome proliferator-activated receptor gamma coactivator 1 alpha and 1 beta (PGC-1α and PGC-1β) and PGC-1-related coactivator (PRC) play significant roles in mitochondrial biogenesis (Table 2). As downstream factors of these coactivators, the nuclear transcription factors NRF1 and NRF2 regulate the expression of mitochondrial respiratory complex subunits, and mitochondrial transcription factor A (TFAM) is essential for mtDNA transcription and replication [58]. PGC-1α coordinates the expression of essential biogenesis proteins through NRF1 and NRF2 and drives the production of mitochondrial proteins, such as complex V and cytochrome c (Cyt c). PGC-1α is currently considered the main regulator of MB [59]. Various signaling cascades regulate MB by influencing the PGC-1α-NRF1/2-TFAM pathway. For example, an increase in AMP will activate AMPK, and AMPK increases the expression of PGC-1α and TFAM by phosphorylating PGC-1α. After AMP is converted into cAMP, cAMP can regulate PGC-1α through the cAMP-PKA-CREB pathway. Ca2+ increases the activity and expression of PGC-1α by stimulating calcium/calmodulin-dependent protein kinase (CaMK) to phosphorylate p38 mitogen-activated protein kinase (p38 MAPK). In addition, CaMK can also stimulate PGC-1α through CREB. An increase in the level of NAD+ leads to the deacetylation of PGC-1α by Sirtuin 1 (Sirt1), activating PGC-1α [60].

2.2.2. Alterations in Mitochondrial Dynamics

Mitochondrial dynamics encompass regulated processes (fission, fusion, mitophagy, and trafficking) that maintain mitochondrial morphology, quantity, and functional integrity. These adaptive mechanisms underpin mitochondria’s roles in metabolic and signaling networks. Dysregulation disrupts mitochondrial homeostasis, resulting in disease pathogenesis [61,62].
Fission is essential for providing a certain number of mitochondria during cell growth and division and can conduct the quality control of mitochondria by eliminating damaged or dysfunctional mitochondria and promoting apoptosis in severely stressed cells [63]. In mammals, the process of mitochondrial fission is mediated by the GTPase dynamin-related protein DRP1 (or DLP1), which is mainly located in the cytoplasm and recruited to mitochondria by MID49, MID51, and MFF, forming a helix around the mitochondria, and the helix contracts to sever the inner and outer membranes [64,65]. Mitochondrial fission is regulated by post-translational modifications of DRP1. DRP1 exists in an inactive state in the cytoplasm, and its activation is regulated by phosphorylation, SUMOylation, ubiquitination, and S-nitrosylation modification [66,67]. Defects in mitochondrial fission may lead to mitochondrial dysfunction, further causing related diseases with the accumulation of damaged mitochondria. This accumulation of damaged mitochondria has a negative impact on the components of the ETC and inhibits ATP production, ultimately leading to cell death [68,69].
During the fusion process, the contents of mitochondria are mixed and exchanged; this process can redistribute mtDNA, mRNA, and proteins to compensate for defective mitochondria, promote the ability of OXPHOS, and maintain mitochondrial function [63,70]. Mitochondrial fusion is a two-step process involving three dynamin-related GTPases. The fusion of the outer mitochondrial membrane (OMM) is mediated by mitofusin 1 (MFN1) and mitofusin 2 (MFN2), while the fusion of the IMM is mediated by optic atrophy protein 1 (OPA1) [71,72]. OPA1, together with MFN1, mediates mitochondrial fusion. In addition, OPA1 also plays an important role in regulating the shape of cristae, the arrangement of ETC supercomplexes, and the control of Cyt c release [73]. The absence of OPA1 leads to the collapse of the mitochondrial network and damage to mitochondrial morphology, and it eventually promotes apoptosis [74]. The deficiency of MFN1 has a relatively minor impact on mitochondrial function, while the deficiency of MFN2 leads to an increase in proton leakage, a decrease in the mitochondrial membrane potential (ΔΨm), and the generation of ROS [75]. During the fusion process, the mixing of the matrix and inner membrane enables the entire mitochondrial compartment to participate in respiration, maximizing ATP synthesis. When there are fusion defects in mitochondria, it leads to mitochondrial dysfunction and the loss of respiratory function [76].

2.2.3. Abnormal Mitophagy

Cells remove dysfunctional or redundant mitochondria through mitophagy, which is essential for maintaining a healthy mitochondrial network [77]. In mammals, mitophagy has a complex and precise mechanism and regulatory mode, which can generally be divided into two categories: ubiquitin (Ub)-dependent pathways and Ub-independent pathways.
Ub-dependent pathways promote mitophagy through the ubiquitination of mitochondrial surface proteins. The phosphatase and tensin homolog (PTEN)-induced putative kinase 1 (PINK1)–Parkin pathway regulates mitophagy in mammals. In this process, the depolarization of the IMM of damaged mitochondria prevents PINK1 from being degraded and allows it to accumulate in the OMM. PINK1 recruits the E3 ubiquitin ligase Parkin and phosphorylates ubiquitin at the Ser65 site, activating Parkin ubiquitin ligase activity. The ubiquitination of multiple OMM proteins (such as MFN1, MFN2, and VDAC1) mediated by Parkin enables them to be recognized by autophagy adaptor proteins or degraded by the proteasome. At the same time, it attracts autophagy receptors (such as p62 and optineurin), and LC3 autophagy-related proteins bind to damaged mitochondria through these receptors, promoting the formation of autophagosomes. Autophagosomes fuse with lysosomes to form autolysosomes, thereby eliminating damaged mitochondria [77,78].
Ub-independent pathways refer to receptor proteins on the OMM containing the LC3-interacting region (LIR), which can directly bind to LC3 to initiate mitophagy without ubiquitination. These receptors include Nip3-like protein X (NIX), also known as BCL2-interacting protein 3 like (BNIP3L) receptor, BCL2-interacting protein 3 (BNIP3) receptor, FUN14 domain containing 1 (FUNDC1) receptor, etc. [79]. Both NIX and BNIP3 can directly bind to LC3 through the BH3 domain to induce mitophagy, and phosphorylation enhances their binding to LC3; FUNDC1 induces Parkin-independent mitophagy by interacting with LC3 under hypoxic conditions. Under normal physiological conditions, the Tyr18 and Ser13 sites of FUNDC1 are phosphorylated by Src kinase and CK2 kinase, respectively, preventing mitophagy. During the process of mitophagy, Ser13 is dephosphorylated, and Ser17 is phosphorylated by ULK1 to activate FUNDC1. In addition, mitophagy receptors can also promote the fission of damaged organelles by recruiting mitochondrial-dynamics-related proteins, such as DRP1 and OPA1, on the mitochondrial surface [80,81].
Mitophagy is meticulously coordinated by adaptive molecular circuits, which calibrate organelle turnover to metabolic flux. The breakdown of these circuits, through aberrant phosphorylation or defective cargo recognition, unleashes irreversible mitochondrial damage, promoting aging-associated and oncogenic trajectories. Various diseases, such as tumorigenesis, cardiovascular diseases, neurodegenerative diseases, and chronic inflammation, are associated with mitochondrial dysfunction and abnormal mitochondrial content [82,83]. When mitochondrial autophagy is insufficient, it leads to mitochondrial dysfunction and the accumulation of a large number of damaged mitochondria within the cell, further damaging the cell; conversely, highly activated mitochondrial autophagy can cause the excessive degradation of normal mitochondria, thereby affecting the cell’s energy supply and leading to cell death [60].

2.2.4. Imbalance Between Generation and Clearance of Mitochondrial ROS

ROS are generated when a single electron is transferred from a redox donor to molecular oxygen (O2), and this process produces the anionic free radical superoxide, which can also be converted into hydrogen peroxide by superoxide dismutase (SOD). Electrons are transferred sequentially through the ETC to complex IV, where O2 acts as the final electron acceptor. This process drives the reduction of O2 to H2O via coupling with H+ in the mitochondrial matrix. However, electrons can also prematurely react with oxygen at ETC sites to form superoxide or hydrogen peroxide [84]. So far, up to 16 sources of ROS have been found in mitochondria, among which 12 ROS-producing sites are related to nutrient oxidation, electron transfer, and OXPHOS. Complex I and complex III are the main sites of ROS production in mitochondria [85].
Low concentrations of mitochondrial H2O2 (one of the types of ROS) are involved in regulating various cellular functions, including cell differentiation, apoptosis, cell proliferation, T-cell activation, and stress signal transduction. However, high concentrations of H2O2 are destructive, so controlling its generation and overall concentration is crucial. The antioxidant system plays an important role in removing ROS [86]. H2O2 can be degraded by mitochondria through three systems: reduced glutathione (GSH), thioredoxin-2 (TRX2), and catalase (CAT) [87]. The GSH system degrades peroxides through glutathione peroxidase (GPX) and simultaneously converts GSH into glutathione disulfide (GSSG) [88]. The TRX2 system degrades hydrogen peroxide through peroxiredoxin (PRX), which reduces hydrogen peroxide to water while being oxidized and inactivated, and the inactivated PRX can be reactivated by mitochondrial TRX2 [89]. Studies have found that mitochondria in the liver and heart also contain catalase, which plays a major role in removing hydrogen peroxide when its concentration is higher than normal. The GSH system and PRX system may act as buffering systems to maintain a low level of hydrogen peroxide for mitochondrial redox signaling [90].
ROS are produced during normal cell metabolism. Low concentrations of ROS are necessary for normal physiological functions, participating in immune responses, and maintaining cellular homeostasis [91]. However, if the level of ROS in mitochondria is too high, it is harmful. ROS can damage DNA, proteins, and lipids, leading to mitochondrial dysfunction and various diseases [92]. For example, ROS can induce mtDNA mutations, causing mitochondrial dysfunction; attack mitochondrial membrane lipids, disrupting the integrity of the membrane; and oxidatively modify mitochondrial proteins, affecting their function and stability. Conversely, mitochondrial dysfunction can lead to the abnormal function of the ETC, increased electron leakage, excessive ROS production, and oxidative stress, further damaging mitochondria and other cellular components [93].

2.2.5. Imbalance of Calcium Homeostasis

Ca2+ is an important signaling molecule in cells. Through the mitochondrial calcium uniporter (MCU) complex, cytoplasmic calcium in the cytoplasm enters the mitochondrial matrix and functions as a signal to regulate ATP production and metabolic fuel selection [94,95]. Ca2+ enters the mitochondrial intermembrane space (IMS) through the voltage-dependent anion channel (VDAC) located on the OMM and then enters the mitochondrial matrix through the MCU complex [96]. The MCU complex is located on the IMM and is composed of the mitochondrial calcium uniporter (MCU) and accessory proteins, such as mitochondrial calcium uptake 1/2 (MICU1/2), MCU regulator 1 (MCUR1), MCU dominant-negative β-subunit (MCUb), solute carrier 25A23 (SLC25A23), and essential MCU regulator (EMRE). The mitochondrial membrane potential, cytosolic Ca2+ gradient, oxidants, pH, and ions are involved in regulating the activity of the MCU complex [97]. The extrusion of Ca2+ from mitochondria is mainly mediated by the mitochondrial Na+/Ca2+/Li+ exchanger (NCLX). One cycle involves the import of three Na+ and the extrusion of one Ca2+ from the mitochondria. PTEN-induced putative kinase 1 (PINK1) acts as a regulator of mitochondrial Ca2+ and regulates the NCLX by mediating the phosphorylation of the NCLX through protein kinase A (PKA) [98]. Mitochondrial Ca2+ homeostasis plays a crucial role in mitochondrial energy metabolism, apoptosis, and ROS production [99]. A low level of matrix-free calcium can improve OXPHOS and activate matrix dehydrogenases, while calcium overload inhibits OXPHOS and reduces ATP synthesis through multiple mechanisms, and high concentrations of calcium phosphate precipitates can disrupt the stability of the mitochondrial cristae network, leading to mitochondrial rupture and fragmentation [100]. In addition, calcium can affect the activities of metabolic enzymes in the TCA cycle, such as pyruvate dehydrogenase, isocitrate dehydrogenase, and α-ketoglutarate dehydrogenase [101]. If an excessive amount of Ca2+ flows into mitochondria, it will lead to mitochondrial outer membrane permeabilization, and pro-apoptotic factors, such as Cyt c and apoptosis-inducing factor, will be released into the cytoplasm, causing cell death [102].

2.2.6. Cell Death

Mitochondria execute important functions in the process of intrinsic apoptosis; during this process, the B-cell lymphoma gene 2 (BCL-2) protein family exerts significant regulatory effects. The oligomerization pro-apoptotic factors Bak and Bax in the BCL-2 family interact with VDAC on the OMM, regulate the permeability of the OMM, and release Cyt c into the cytoplasm, and subsequently trigger the caspase cascade, leading to apoptosis [103,104]. Under normal circumstances, BAX and BAK are in an inactive state, with BAX located in the cytoplasm and BAK in the mitochondria. Both can freely shuttle between the mitochondria and cytoplasm. During apoptosis, BAX and BAK are directly activated by binding to the BH3 protein subfamily, causing them to anchor to the OMM and dimerize. These dimers then further oligomerize, creating lipid pores on the OMM and changing the permeability of the OMM [105]. The mitochondrial permeability transition pore (mPTP) is a conductive channel located on the IMM and plays a significant role in cell death. The mPTP will be activated when there is calcium overload and an increase in ROS, disrupting the mitochondrial membrane potential and leading to cell death [106].
The progressive mitochondrial dysfunction that occurs during aging may predispose cells to necroptosis. For instance, mitochondrial ROS can increase the autophosphorylation of receptor-interacting protein kinase 1 (RIPK1), recruit RIPK3, and promote the assembly of the necrosome complex, thus facilitating the occurrence of necroptosis [107]. Recent studies have shown that RIPK3 promotes aerobic respiration and mitochondrial ROS production by phosphorylating and activating the E3 subunit of the pyruvate dehydrogenase complex, indicating an association between mitochondrial instability and necroptosis [108].
Ferroptosis is closely associated with mitochondrial dysfunction. The excessive activation of the ETC leads to an increased release of mitochondrial ROS, enhancing the oxidation of polyunsaturated fatty acids (PUFAs) and increasing PUFA synthesis by inhibiting the AMPK pathway, resulting in increased lipid peroxidation. Additionally, damaged mitochondria can also indirectly downregulate glutathione-dependent peroxidase (GPX4) through the inhibition of the integrated stress response mediated by EIF2α phosphorylation. The breakdown of glutamine and the TCA cycle in mitochondria can further promote ferroptosis by driving the activity of the ETC. As mentioned above, mitochondria can promote ferroptosis by playing various roles in bioenergetics, biosynthesis, and ROS regulation [109,110,111]. In the cells undergoing ferroptosis, there are huge morphological changes in mitochondria, such as contracted mitochondria with increased membrane density, enlarged cristae, and a ruptured OMM [112]. The role of mitochondria in ferroptosis is conditional: it plays a key role in ferroptosis induced by cysteine deprivation. If ferroptosis is induced by the inhibition of GPX4 activity, cells will adopt a form of ferroptosis independent of mitochondrial function [113].
ROS produced in mitochondria activate the NLRP3 inflammasome, and caspase-1 is activated to form mature caspase-1, which leads to the maturation of interleukin (IL)-18 and IL-1β and the cleavage of GSDMD to generate N-GSDMD, inducing pyroptosis. Mitophagy can reduce the activation of the NLRP3 inflammasome by eliminating damaged mitochondria and reducing ROS production. However, the activation of caspase-1 inhibits mitophagy and increases mitochondrial damage [114,115]. Moreover, mitochondrial damage caused by Ca2+ overload generates a large amount of ROS, which can also lead to the activation of the NLRP3 inflammasome and pyroptosis [107].

2.2.7. Mitochondria-Targeted Therapy in Diseases

Mitochondria-based disease treatments have become an emerging trend. A wide range of studies have developed mitochondria-targeted delivery systems for disease therapy. For instance, mitochondriotropic particulate carriers, on the basis of liposomes, biodegradable polymers, and metals, target mitochondria through the incorporation of mitochondriotropic agents onto the surface and delivering therapeutic molecules to mitochondria to induce mitochondria dysfunction [116]. Derivatives of triphenylphosphonium (TPP) are lipophilic and have an affinity for mitochondria; by leveraging mitochondrial localization to enhance the therapeutic effect of diseases, they serve as effective cancer-targeting ligands [117]. Research based on amphipathic peptoids discovered two peptoids that showed highly efficient cell penetration and mitochondrial localization and might serve as potential transporters for the delivery of bioactive compounds, such as drugs, antioxidants, etc. [118]. Recently, a mitochondria-targeted RNAi nanoparticle (NP) platform, which targeted mitochondrial metabolism, was developed and employed in the treatment of breast cancer, and this system could successfully transport siRNA into the mitochondria to regulate mtDNA-encoded protein expression and inhibit tumor growth [119]. Researchers have also devised more novel methods to target mitochondria for ischemic stroke (IS) treatment; one study adopted nanomedicine-based brain targeting strategies and developed a magnetic-field-driven, mitochondria-targeted ceria (MMTCe) nanosystem. By using this system, the targeting of the damaged mitochondria was achieved, and the ischemic microenvironment was improved [120]. In addition, a platinum-based terminal-sensitive projectile (TSB) was developed to precisely target the mitochondria of tumor cells, significantly enhancing the sensitivity of tumor cells to platinum-based chemotherapeutic drugs [119]. Given the indispensable function of mitochondria in cell metabolism and growth, a growing body of evidence suggests that targeting mitochondria may ultimately be accepted as a promising strategy for cancer medication. Most studies focus on the nanoparticle-based delivery of mitochondria for cancer therapy [121,122,123,124,125].
Table 2. The role of factors in mitochondrial function.
Table 2. The role of factors in mitochondrial function.
CategoryFactorFull NameFunctionReference
Mitochondrial BiogenesisPGC-1αPeroxisome proliferator-activated receptor gamma coactivator 1 alphaRegulates processes such as mitochondrial biogenesis, fission, fusion, and mitophagy by modulating coactivators and downstream effector factors.[126]
PGC-1βPeroxisome proliferator-activated receptor gamma coactivator 1 betaPromotes mitochondrial biogenesis and is essential for normal OXPHOS and mitochondrial function.[126]
PRCPGC-1-related coactivatorActivates the transcription factors NRF1 and NRF2, which are associated with the expression of the respiratory chain.[127]
TFAMMitochondrial transcription factor AEssential for the transcription and replication of mtDNA.[58]
Mitochondrial DynamicsDRP1
(DLP1/
DNM1L)
The GTPase dynamin-related protein 1Recruited from the cytoplasm to the outer mitochondrial membrane and mediate the process of mitochondrial fission.[64,65]
MID49Mitochondrial dynamics proteins of 49Recruits DRP1 and promotes mitochondrial fission.[128]
MID51Mitochondrial dynamics proteins of 51Recruits DRP1 and promotes mitochondrial fission.[128]
MFFMitochondrial fission factorActs as a DRP1 receptor on the mitochondrial membrane, recruits DRP1, and promotes mitochondrial fission.[129]
FIS1Fission protein 1Recruits DRP1 and facilitates mitochondrial fission.[128]
MTP18
(MTFP1)
Mitochondrial fission process 1Maintains the mitochondrial morphology by regulating mitochondrial fission.[130]
MFN1Mitofusin 1Together with MFN2, mediates the fusion of the OMM in a GTP-dependent manner, coordinating the sequential fusion of the OMM and the IMM with the IMM fusion regulator OPA1.[129]
MFN2Mitofusin 2Together with MFN1, it mediates the fusion of the OMM in a GTP-dependent manner, mediates mitochondria–ER tethering, and transfers phosphatidylserine from ER to mitochondria.[129]
OPA1Optic atrophy protein 1Controls the fusion of the IMM and participates in processes such as regulating the shape of cristae, the arrangement of ETC supercomplexes, and the control of Cyt c release.[73]
MitophagyPINK1Phosphatase and tensin homolog (PTEN)-induced putative kinase 1Recruits Parkin to initiate mitophagy and eliminate damaged mitochondria.[130]
ParkinParkinE3 ubiquitin ligase, ubiquitinates multiple OMM proteins, recruits autophagy receptors to damaged mitochondria, and promotes mitophagy.[130]
BNIP3L
(NIX)
BCL2-interacting protein 3 likeParticipates in the process of mitophagy as a mitophagy receptor and promotes the formation of autophagosomes.[131]
BNIP3BCL2-interacting protein 3Induces mitophagy by binding to LC3 through the BH3 domain.[132]
FUNDC1FUN14 domain containing 1A ubiquitin-independent mitophagy receptor that can directly bind to LC3 to initiate mitophagy.[133]
Mitochondrial Oxidative StressSODSuperoxide dismutaseAn antioxidant enzyme that converts superoxide anion radicals into hydrogen peroxide and oxygen.[84]
GPXGlutathione peroxidaseAn antioxidant enzyme family that utilizes reduced glutathione as an electron donor to catalyze the reduction of hydrogen peroxide or organic hydroperoxides to water or the corresponding alcohols.[134]
TRX2Thioredoxin-2Clears ROS in the cell through the TRX2/PRX system and regulates the apoptotic signaling pathway by inhibiting oxidative stress.[135]
PRXPeroxiredoxinA multifunctional enzyme that reduces peroxides through the cysteine residues at the active center and also acts as a redox signaling regulator, chaperone, and pro-inflammatory factor.[136]
Calcium
Homeostasis
MCUMitochondrial calcium uniporterForms a pore through which calcium ions enter the mitochondria and regulates the concentration of Ca2+ in the mitochondria.[137,138]
MICU1Mitochondrial calcium uptake 1A key regulatory factor for mitochondrial Ca2+ uptake. When the intracellular Ca2+ level is high, it promotes the influx of calcium ions into the mitochondria.[139,140]
MICU2Mitochondrial calcium uptake 2When the level of calcium ions outside the mitochondria is low, it turns off the activity of the MCU.[138]
MCUR1MCU regulator 1Acts as a scaffolding factor to bind the MCU and EMRE.[141]
MCUbMCU dominant-negative β-subunitAn inhibitory subunit of the MCU complex that forms a multimer with the MCU to inhibit the influx of Ca2+.[142]
SCL25A23Solute carrier 25A23Participates in mitochondrial Ca2+ uptake and interacts with the MCU and MICU1 to enhance the activity of the MCU channel.[143]
EMREEssential MCU regulatorActivates the function of the MCU, increases the uptake of Ca2+, and maintains the MICU regulation of the MCU pore.[137]
NCLXMitochondrial Na+/Ca2+/Li+ exchangerMediates the efflux of mitochondrial Ca2+ using the entry of sodium ions into the mitochondria along their concentration gradient as the driving force; transports calcium ions out of the mitochondria to maintain mitochondrial calcium homeostasis.[98]
Cell DeathBAKBCL2-antagonist/killerTogether with BAX, mediates the permeabilization of the OMM in the mitochondrial pathway and promotes apoptosis.[144]
BAXBCL2 Associated X ProteinTogether with BAK, mediates the permeabilization of the OMM in the mitochondrial pathway and promotes apoptosis.[144]
RIPK1Receptor-interacting protein kinase 1A key mediator of the apoptotic, necroptotic, and inflammatory pathways that mediate necroptosis.[145]
RIPK3Receptor-interacting protein kinase 3Acts as a downstream mediator of RIPK1 to trigger necroptosis.[146]
GPX4Glutathione-dependent peroxidase 4A form of glutathione peroxidase that specifically catalyzes the conversion of lipid hydroperoxides into non-toxic lipid alcohols to alleviate ferroptosis.[147]

3. M6A and Mitochondrial Dysfunction

3.1. Role of m6A in Influencing Mitochondrial Function in Disease

Nuclear-encoded genes, when regulated by epigenetics, can affect the expression of mitochondrial proteins, such as histone acetylation, as well as histone and DNA methylation [148]. However, the relationship between RNA methylation and mitochondrial function has not been systematically explained. This review will explore and summarize how m6A RNA methylation in epigenetics affects mitochondrial function.
Studies have found that m6A RNA methylation regulates mitochondrial function by promoting the translation of nuclear-encoded mitochondrial ETC subunit RNAs. After knocking out the RNA methyltransferase METTL14, the absence of m6A significantly downregulated metabolites related to energy metabolism and led to a significant decrease in the mitochondrial respiratory capacity and membrane potential. These functional defects are consistent with the reduced expression of mitochondrial ETC complexes, as well as the decreased assembly and activity of mitochondrial supercomplexes. Mechanistically, the absence of m6A reduces the translation efficiency by decreasing the binding of methylated RNAs encoding mitochondrial complex subunits to polysomes, but it has no effect on RNA stability. Moreover, it has been found that the regulation of mitochondrial complex RNA translation by m6A may be mediated by the YTHDF family of readers [149].
As an m6A demethylase, FTO regulates mitochondrial content by mediating the expression of genes related to mitochondrial fusion, fission, and biogenesis. Experimental results show that when FTO is overexpressed, genes involved in mitochondrial fusion (MFN1/2 and OPA1) are upregulated, genes involved in mitochondrial fission (FIS1, DRP1, and MTP18) are downregulated, and the expression of genes involved in mitochondrial biogenesis (PGC-1α and TFAM) is inhibited. In summary, FTO may reduce mitochondrial content and ATP levels by promoting mitochondrial fusion and inhibiting mitochondrial fission [150,151,152]. As a positive regulator of mitophagy, FTO can promote mitophagy by regulating the level of the mitophagy receptor BNIP3 in an m6A-dependent manner. Mechanistically, the absence of FTO leads to an increase in the m6A modification level of BNIP3. YTHDF2 binds to BNIP3 by recognizing m6A sites, promoting the degradation of its mRNA and a reduction in BNIP3 protein levels. FTO promotes mitophagy by increasing the stability of BNIP3 mRNA [153]. After cerebral ischemia/reperfusion (I/R), a low expression of FTO was found in the brain tissue, the level of m6A modification on NRF2 mRNA increased, and the expression of NRF2 was downregulated. This is because YTHDF2 binds to m6A-methylated NRF2, promoting the degradation of its mRNA. The antioxidant effect mediated by NRF2 can alleviate cerebral I/R injury [154].
IGF2BP2 plays an important role in maintaining the function of hematopoietic stem cells (HSCs). The absence of IGF2BP2 accelerates the decay of Bmi1 polycomb ring finger oncogene (Bmi1) mRNA, thus increasing mitochondrial activity in HSCs. High mitochondrial activity indicates a lower reconstitution ability of HSCs. Furthermore, the absence of METTL3 increases the expression of mitochondrial-related genes and reduces the expression of genes involved in HSC maintenance. All of these indicate that IGF2BP2 plays a role in inhibiting mitochondrial activity in HSCs [155].
METTL3 can improve mitochondrial dysfunction by increasing the m6A modification level of MFN2 mRNA and enhancing the expression of MFN2. The overexpression of METTL3 or MFN2 can improve mitochondrial dysfunction in Alzheimer’s disease (AD) by, for instance, reducing mitochondrial damage and increasing ATP levels [156]. TRAF6 is a multifunctional signaling molecule belonging to the TRAF family, which performs physiological functions through the Toll-like receptor 4 (TLR4) signaling pathway. The evolutionarily conserved signaling intermediate in Toll pathways (ECSIT) is also a multifunctional protein partially located in the IMM and participates in the assembly of OXPHOS complex I. It has been reported that TRAF6, after rapid activation, translocates to the mitochondria and increases the production of mitochondrial reactive oxygen species (mROS) in macrophages. ECSIT can interact with TRAF6 to participate in the production of mROS and the formation of complex I. The METTL3-mediated m6A modification of TRAF6 activates TRAF6 and causes it to translocate to the mitochondria, activating the TRAF6/ECSIT pathway and leading to the production of mROS. When METTL3 is inhibited, the translocation of TRAF6 to the mitochondria, the expression of ECSIT, and the production of mROS are all weakened, while the levels of ATP and ΔΨm in the mitochondria increase [157]. Mitochondrial lon peptidase 1 (LONP1) is a protein quality control protease that plays an important role in regulating mitochondrial protein homeostasis and maintains the integrity of mtDNA by selectively degrading abnormal and oxidatively damaged proteins [158]. It has been found that as the kidney ages, the expression of LONP1 and TFAM is significantly reduced, and the levels of mtDNA and ATP are also significantly reduced. Moreover, the mitochondrial structure changes, with mitochondrial swelling and disordered cristae. Studies have found that after overexpressing METTL3, the expression of LONP1 and TFAM significantly increased and the expression of DRP1 reduced, the levels of mtDNA and ATP increased, and the integrity of the mitochondrial structure was maintained; above all, mitochondrial dysfunction was alleviated [159]. A high expression of METTL3 was found in sepsis-induced acute lung injury, which increased the level of m6A modification. IGF2BP3 recognizes these modifications and enhances the stability of the mRNA of certain genes, such as HIF-α, leading to the abnormal activation of the mitochondrial metabolic pathway, downregulation of GPX4, and exacerbation of ferroptosis, thus aggravating lung tissue injury [160]. Both METTL3 and YTHDF2 are highly expressed in fibrotic cardiac tissues. Growth arrest-specific 5 (GAS5) is a long non-coding RNA (lncRNA) located in the mitochondria and regulates the mitochondrial metabolism. GAS5 can bind to DRP1 and inhibit the expression of DRP1. In addition, the expression of GAS5 is regulated by m6A methylation. A high expression of METTL3 increases the m6A modification of GAS5. YTHDF2 induces the degradation of GAS5 mRNA by recognizing the m6A sites on GAS5, resulting in increased mitochondrial fission and leading to cardiac fibrosis [161].
The absence of YTHDF1 leads to a decrease in the level of mitochondrial ATP production, a reduction in the expression levels of PGC-1α, TFAM, and MFN2, and an increase in the expression levels of DRP1 and FIS1. This results in a decrease in mitochondrial biogenesis and mitochondria damage, such as fragmentation, swelling, and loss of cristae. Overall, it leads to mitochondrial dysfunction, promotes oxidative stress, and finally causes cell apoptosis [162]. In metabolic-dysfunction-associated steatotic liver disease (MASLD), the expression level of METTL3 increases while the expression level of YTHDF1 decreases. The increased expression of METTL3 promotes the function of the mitochondrial ETC, and the level of energy metabolism also increases significantly. However, the decreased expression of YTHDF1 reverses the effect of METTL3 and inhibits mitochondrial OXPHOS. This is mainly because the m6A modification of mitochondrial subunits mediated by METTL3 requires YTHDF1 as a “reader” to stabilize the expression of these RNAs, thus promoting the expression of mitochondrial complexes [163].
Studies have found that the Notch signaling pathway can restore mitochondrial dynamics [164]. ALKBH5 increases the stability of notch receptor 1 (NOTCH1) mRNA, leading to the upregulation of NOTCH1 expression and downregulation of DRP1 expression, thereby reducing mitochondrial fission. However, when the m6A modification level of NOTCH1 mRNA increases, it enhances the recognition of m6A modification on NOTCH1 mRNA by YTHDF2, inducing the degradation of NOTCH1 and promoting mitochondrial fission [165]. ALKBH5 is significantly downregulated in fibrotic liver tissues, while YTHDF1 is significantly upregulated. ALKBH5 demethylates the m6A modification on the 3′ UTR of DRP1. The downregulation of ALKBH5 leads to the upregulation of m6A modification. YTHDF1 promotes the expression of DRP1 by recognizing the m6A modification on the 3′ UTR of DRP1, thus promoting mitochondrial fission and leading to liver fibrosis [166].
During myocardial ischemia/reperfusion injury (MIRI), various mitochondrial dysfunctions occur, including an impaired mitochondrial ATP production, decreased mitochondrial membrane potential, and excessive production of ROS, resulting in conditions such as myocardial dysfunction, DNA damage, and apoptosis. LncRNA Snhg1 can upregulate the expression of OPA1 by sponging miR-361-5p, thereby improving the progression of MIRI. However, in MIRI, the expression of lncRNA Snhg1 is found to be downregulated. This is because WTAP induces the m6A modification of lncRNA Snhg1, and YTHDF2 recognizes and binds to lncRNA Snhg1 through its m6A modification, thus promoting its degradation [167].
In conclusion, we can observe that although methyltransferases and demethylases play crucial roles in the regulation of diseases, whether m6A modification can exert its functions mainly depends on the roles of readers. Different readers mediate different processes, and whether an m6A-modified mRNA is upregulated in expression or degraded depends on which reader mediates this process.

3.2. Role of m6A in Influencing Mitochondrial Function in Cancer

Many studies have found that the growth of tumor cells can be regulated by mitochondrial metabolic reprogramming. Next, we will summarize how m6A methylation affects cancer through mitochondria function. Uncontrolled cell proliferation is one of the hallmarks of cancer. To maintain the ability of rapid proliferation, cancer cells make certain adjustments to energy metabolism to obtain more energy—that is, enhancing or activating metabolic pathways [168]. Metabolic reprogramming has become one of the hallmarks of cancer and is also one of the abilities acquired during malignant transformation [169]. Warburg discovered that in the presence of oxygen, cancer cells limit energy metabolism by activating glycolysis and converting glucose into lactic acid, which is known as the Warburg effect [170,171]. Although mitochondrial dysfunction in cancer cells can lead to this shift in energy metabolism, studies have shown that many cancer cells have fully functional mitochondria and are able to oxidize glucose through OXPHOS [172]. The efficiency of ATP production via glycolysis is much lower than that of OXPHOS. To eliminate this efficiency difference, the glucose uptake is increased by upregulating glucose transporters, and glycolysis is enhanced, leading to the accumulation of glycolytic intermediates. These glycolytic intermediates are transferred to different biosynthetic pathways, including pathways for nucleotide synthesis and pathways for generating amino acids and NADPH, to meet the requirements of cell proliferation [173,174]. During the development of cancer, mitochondria promote typical characteristics such as metabolic reprogramming, continuous proliferation, evasion of cell death, invasion, and the induction of angiogenesis (Table 3). Therefore, mitochondria are key mediators of tumorigenesis [175].
Methylenetetrahydrofolate dehydrogenase 2 (MTHFD2), a mitochondrial enzyme, is a key enzyme in the one-carbon metabolic pathway; research has found that the expression of MTHFD2 is significantly upregulated in renal cell carcinoma (RCC). It specifically increases the methylation level of METTL3-dependent hypoxia-inducible factor 2α (HIF-2α), promotes the translation of HIF-2α mRNA, and increases the protein level of HIF-2α, thus promoting the glycolysis of tumor cells and tumor progression [217].
Von Hippel–Lindau (VHL) deficiency or mutation is frequently observed in clear cell renal cell carcinoma (ccRCC). The loss of VHL function promotes tumorigenesis. Peroxisome proliferator-activated receptor gamma coactivator (PGC), as a family of transcriptional coactivators, mediates mitochondrial biogenesis and OXPHOS. Its family members include PGC-1α, PGC-1β, and PRC. PGC-1α, as a core regulator of mitochondrial function, plays a crucial role. Its expression inhibits tumor growth. FTO has an anti-cancer effect in cancer development and is downregulated in ccRCC. Studies have found that Von Hippel–Lindau-deficient cells expressing FTO reduced the m6A level of PGC-1α mRNA, stabilizing PGC-1α mRNA, restoring mitochondrial activity, promoting oxidative stress and ROS production, and thus inhibiting tumor growth [218].
Gene NADH dehydrogenase (ubiquinone) 1 alpha subcomplex 4 (NDUFA4) encodes a subunit in the ETC complex of the mitochondrial respiratory chain and is highly expressed in gastric cancer (GC). NDUFA4 can promote the glycolysis and oxidative metabolism of GC cells, inhibit ROS levels in GC cells, and promote mitochondrial membrane potential (MMP) levels. Moreover, it has been found that METTL3 can increase the m6A level of NDUFA4 mRNA through IGF2BP1 to promote the expression of NDUFA4 in GC cells, thus promoting the development of GC [219].
Studies have found that in cervical cancer and liver cancer cells, when Mettl3 is absent or ALKBH5 is overexpressed, glucose consumption, lactate production, and ATP generation are all inhibited, but there is no significant effect on mitochondrial DNA content. Mechanistically, m6A positively regulates glycolysis and ATP generation in cancer cells through pyruvate dehydrogenase kinase 4 (PDK4). The 5′UTR of m6A-modified PDK4 binds to the YTHDF1/eEF-2 complex and IGF2BP3, positively regulating its mRNA stability and translation elongation and thereby affecting the energy metabolism process of cells and indirectly influencing tumor growth and progression [220].
M6A can promote the development of colorectal cancer (CRC) by inducing the synthesis of GSH and stabilizing OPA1 mRNA to promote mitochondrial fusion. Mechanistically, METTL3 modifies ribonucleotide reductase regulatory TP53 inducible subunit M2B (RRM2B) and OPA1 mRNA with m6A, while IGF2BP2 increases the stability of mRNA by binding to RRM2B and OPA1, increasing the expression of their proteins and thereby promoting GSH synthesis and mitochondrial fusion [221].
Caveolin 1 (CAV1) is a membrane protein that is essential for maintaining mitochondrial structure and function and is associated with mitochondrial quantity and bioenergetic function. Its role in tumors is controversial. Studies have found that FTO is highly expressed in gastric cancer tissues. The knockdown of FTO inhibits the proliferation, migration, and invasion of gastric cancer cells. Mechanistically, FTO promotes the degradation of CAV1 mRNA by reducing its m6A modification level, thereby affecting mitochondrial fission, fusion, and metabolism, promoting the development of GC [222].
Mitochondrial fission promotes the progression of glioma. Mechanistically, METTL3 promotes the alternative splicing (AS) of LINC00475 to generate LINC00475-S by increasing the binding of m6A recognition protein HNRNPH1 to LINC00475. By inhibiting the expression of macrophage migration inhibitory factor (MIF), it enhances the expression of DRP1 and p-DRP1 while simultaneously inhibiting the expression of OPA1 and MFN2 to promote mitochondrial fission, thereby promoting the progression of glioma [223].
Research has found that FTO is upregulated in breast cancer and promotes the development of breast cancer. Mechanistically, FTO downregulates BNIP3 through demethylating m6A in the 3′ UTR of BINP3 mRNA and inhibits the apoptosis induced by it, thus promoting the proliferation of breast cancer cells, and this process is independent of YTHDF2 [224].
In papillary thyroid cancer (PTC), the expression of FTO is significantly downregulated. Apolipoprotein E (APOE) is an m6A modification target downstream of FTO; its m6A modification is recognized and mediated by IGF2BP2. The downregulation of FTO promotes the expression of APOE and enhances glycolysis in PTC, thus promoting tumor growth. Mechanistically, FTO acts as a negative regulator of PTC and inhibits the expression of APOE. These changes inhibit glycolysis in PTC by regulating the IL-6/JAK/STAT3 signaling pathway, thereby affecting tumor growth [225].
A high expression of METTL3 is found in chemoresistant small cell lung cancer (SCLC) cell lines and promotes chemoresistance in SCLC. DCP2 is a downstream target of METTL3; the highly m6A-modified DCP2 leads to the degradation of its mRNA through the selective recognition of YTHDF2. The decreased expression of DCP2 protein enhances the stability of PINK1 and Parkin mRNA, resulting in increased mitophagy and reduced mitochondrial damage. Eventually, it leads to the chemoresistance of SCLC to chemotherapeutic drugs [226].
In CRC cell lines, IGF2BP2 is highly expressed and lncRNA ZFAS1 is significantly highly expressed. IGF2BP2 promotes the stability of ZFAS1 through the m6A sites on ZFAS1, thus promoting the development of CRC. Mechanistically, obg-like ATPase 1 (OLA1) serves as a key downstream target. ZFAS1 binds to the OBG-type domain of OLA1 (Table 4), thereby exposing the ATP binding site on OLA1, enhancing its ATP hydrolysis ability, and activating the glycolytic pathway, thus promoting the development of CRC [227].
In GC tissues, an upregulated expression of METTL3 and IGF2BP3 is found, leading to an increase in the m6A modification on DRP1. IGF2BP3 recognizes the m6A sites and increases the expression of DRP1 mRNA, thus exacerbating mitochondrial fission, dysfunction, and the increase in mROS, inducing the activation of the NLRP3 inflammasome and promoting the development of gastric cancer [228].
IGF2BP2 is highly expressed in acute myeloid leukemia (AML). It mainly participates in glutamine (Gln) metabolism in AML by enhancing the expression of GPT2 and SLC1A5 in the Gln metabolic pathway. Among them, GPT2 encodes an enzyme that catalyzes the reversible conversion of pyruvate and glutamate (Glu) into alanine and α-ketoglutarate (αKG) in mitochondria during Gln metabolism, and SLC1A5 encodes the main transporter of Gln in cancer cells. IGF2BP2 recognizes and increases the mRNA stability and expression of MYC, GPT2, and SLC1A5 through m6A modification, providing fuel for the TCA cycle, thus promoting the development of AML [229].
From the above examples, we can see that the development of cancer is complex. There may be two or even multiple regulatory processes in the same type of cancer. The same demethylase, such as FTO, plays different functions in different cancers. It is a positive regulator in breast cancer but a negative regulator in papillary thyroid cancer. M6A methylation plays a dual role in cancer treatment. Both the excessive modification and lack of modification of m6A may lead to the development of cancer (Figure 3).
Table 4. A summary of the factors that change in diseases.
Table 4. A summary of the factors that change in diseases.
FactorFull NameFunctionReference
TRAF6Tumor necrosis factor receptor-associated factor 6Belonging to the TRAFs family, an adaptor protein is recruited to the intracellular region when activated and executes a variety of physiological functions through the TLR4 signaling pathway.[157]
ECSITEvolutionarily conserved signaling intermediate in Toll pathwaysA multifunctional protein partially located in the IMM, which participates in the assembly of oxidative phosphorylation complex I.[157]
LONP1Lon peptidase 1A protein quality control protease that plays an important role in regulating mitochondrial protein homeostasis and maintains the integrity of mtDNA by selectively degrading abnormal and oxidatively damaged proteins.[158]
MTHFD2Methylenetetrahydrofolate dehydrogenase 2A mitochondrial enzyme encoded by the nucleus, which participates in folate metabolism and one-carbon metabolism in mitochondria and maintains intracellular redox balance.[230]
NDUFA4NADH dehydrogenase (ubiquinone) 1 alpha subcomplex 4Encodes a subunit in the electron transport chain complex of the mitochondrial respiratory chain to generate ATP.[219]
PDK4Pyruvate dehydrogenase kinase 4Promotes the transition from mitochondrial oxidative phosphorylation to glycolysis and regulates glucose metabolism by phosphorylating pyruvate dehydrogenase.[231]
RRM2BRibonucleotide reductase regulatory TP53 inducible subunit M2BA key subunit of ribonucleotide reductase (RR) plays important roles in DNA repair, replication, oxidative stress, and mtDNA synthesis.[232]
RRRibonucleotide reductaseCatalyzes ribonucleoside diphosphates to deoxyribonucleoside diphosphates and plays an important role in DNA synthesis and repair.[232]
Caveolin-1Caveolin-1A membrane protein that is essential for maintaining the structure and function of mitochondria and is associated with the number of mitochondria and their bioenergetic functions.[222]
APOEApolipoprotein EA glycoprotein that functions as a lipid transport protein.[233]
DCP2Decapping MRNA 2A major decapping enzyme during 5′ to 3′ mRNA decay, controlling the expression of PINK1 and Parkin to regulate mitophagy and the level of mitochondrial damage.[226]
OLA1Obg-like ATPase 1An ATP hydrolase that mediates mitochondrial energy metabolism, including ATP hydrolysis and glycolysis.[227]

3.3. Small Molecule Drug Therapy

STM2457 acts as an inhibitor of METTL3. The high expression of METTL3 promotes the degradation of DCP2 and increases the expression of PINK1 and Parkin, thereby enhancing mitophagy and contributing to chemoresistance in SCLC. STM2457 increases the expression level of DCP2 and decreases the expression levels of PINK1 and Parkin by inhibiting the expression of METTL3 in SCLC cells, thus reversing the chemoresistance of SCLC cells [226].
CWI1-2 serves as an inhibitor of IGF2BP2, binding to the RNA binding site of IGF2BP2 and competitively inhibiting its binding to other RNAs. Studies have found that CWI1-2 exhibits antileukemic efficacy both in vivo and in vitro. It mainly inhibits the role of IGF2BP2 in AML, reduces the uptake of Gln, impairs mitochondrial function, and decreases ATP production, thus inhibiting the progression of AML [229].
Quercetin (QUE) is a natural flavonoid with properties such as regulating mitochondrial biogenesis and anti-cancer and anti-inflammatory effects. Research has shown that QUE interacts with METTL3. QUE significantly inhibits the expression level of METTL3. In lipopolysaccharide (LPS)-induced liver injury, the use of QUE reduces the m6A methylation level of PTEN, promotes the activation of the PI3K/AKT signaling pathway, decreases the expression of DRP1, and increases the expression of MFN2, MFN1, and OPA1 as well as key enzymes related to glycolysis and the TCA cycle, effectively preventing the imbalance of mitochondrial dynamics and energy metabolism induced by LPS [234]. In addition, another study found that QUE reduces the m6A modification of serine–threonine kinase protein kinase D2 (PRKD2) by inhibiting the expression of METTL3 and improves insulin resistance (IR) by promoting glucose uptake and inhibiting oxidative stress [235].
From the above, we can see that small molecule inhibitors targeting m6A regulators also hold potential in treating diseases by targeting mitochondrial function. However, when using small molecule inhibitors, attention should be paid to the dosage and whether they are specific. It is also important to develop targeted delivery systems.

4. Summary and Prospects

In recent years, it has been found that m6A plays an important role in maintaining the normal function of mitochondria, and its abnormality often leads to mitochondrial dysfunction. It plays its role mainly through several aspects: regulating the expression of mitochondria-related genes, influencing mitochondrial biogenesis, regulating mitochondrial dynamics, participating in the oxidative stress reaction, and affecting mitophagy. Readers determine the fate of mitochondria-related genes by recognizing the m6A sites on them. For example, some readers will promote the degradation of the mRNA of mitochondria-related genes with m6A modification, while other readers will increase the stability of the mRNA and promote translation. M6A affects the quality control of mitochondria by participating in the expression of genes related to mitochondrial biogenesis and mitophagy. M6A may affect the processes of mitochondrial fission and fusion by regulating the expression of proteins related to mitochondrial dynamics, thereby influencing mitochondrial function. Oxidative stress is a significant cause of mitochondrial dysfunction. M6A modification can regulate the expression of genes related to the antioxidant defense system in cells and affect the response ability of cells to oxidative stress. When mitochondria are seriously damaged, and mitophagy fails to remove the damaged mitochondria in a timely manner, it will trigger cell death. M6A is involved in the regulation of mitochondrial function at multiple levels, and its abnormality is closely related to mitochondrial dysfunction.
Mitochondrial dysfunction plays a certain role in promoting the development of various diseases. Each disease will create a specific physiological environment. It is crucial to identify specific therapeutic targets according to the physiological characteristics of the disease and different m6A-mitochondrial function processes. Secondly, tumor cells meet the requirements of their rapid proliferation through metabolic reprogramming. M6A modification can affect the growth and proliferation of tumor cells by regulating the mRNA stability and translation efficiency of key enzymes involved in OXPHOS and glycolysis. M6A can also affect mitochondrial fission, fusion, and quality, thereby influencing the progression of cancer by affecting the expression of proteins related to mitochondrial dynamics, biogenesis, and mitophagy. M6A plays a complex but important role in cancer by influencing mitochondrial function. Further research on the relationship between m6A and mitochondrial function is expected to provide new targets and strategies for cancer treatment.

Author Contributions

All authors substantially contributed to this work. Conceptualization, W.Y., S., K.L. (Ke Li), K.L. (Kexin Li), and C.W.; writing—original draft preparation, W.Y., S., K.L. (Ke Li), K.L. (Kexin Li), and C.W.; writing—review and editing, W.Y., S., K.L. (Ke Li), K.L. (Kexin Li), and C.W.; visualization, W.Y.; project administration, C.W., K.L. (Kexin Li), W.Y., and S.; funding acquisition, C.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (CN) (grant Nos. 81660024 and 32460149) and the Natural Science Foundation of Inner Mongolia Autonomous Region (No. 2020MS08096) to C.W.

Acknowledgments

We wish to convey great appreciation to the colleagues who have contributed constructive work in this rapidly expanding field.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Adebayo, M.; Singh, S.; Singh, A.P.; Dasgupta, S. Mitochondrial fusion and fission: The fine-tune balance for cellular homeostasis. FASEB J. 2021, 35. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, Y.; Huang, Y.; Xu, C.; An, P.; Luo, Y.; Jiao, L.; Luo, J.; Li, Y. Mitochondrial Dysfunction and Therapeutic Perspectives in Cardiovascular Diseases. IJMS 2022, 23, 16053. [Google Scholar] [CrossRef] [PubMed]
  3. Hung, C.H.-L.; Cheng, S.S.-Y.; Cheung, Y.-T.; Wuwongse, S.; Zhang, N.Q.; Ho, Y.-S.; Lee, S.M.-Y.; Chang, R.C.-C. A reciprocal relationship between reactive oxygen species and mitochondrial dynamics in neurodegeneration. Redox Biol. 2018, 14, 7–19. [Google Scholar] [CrossRef] [PubMed]
  4. Kashatus, D.F. The regulation of tumor cell physiology by mitochondrial dynamics. Biochem. Biophys. Res. Commun. 2018, 500, 9–16. [Google Scholar] [CrossRef]
  5. An, Y. The Role of m6A RNA methylation in cancer metabolism. Mol. Cancer 2022, 21, 14. [Google Scholar] [CrossRef]
  6. Meyer, K.D.; Saletore, Y.; Zumbo, P.; Elemento, O.; Mason, C.E.; Jaffrey, S.R. Comprehensive analysis of mRNA methylation reveals enrichment in 30 UTRs and near Stop Codons. Cell 2012, 149, 1635–1646. [Google Scholar] [CrossRef]
  7. Liu, Q.; Gregory, R.I. RNAmod: An Integrated System for the Annotation of mRNA Modifications. Nucleic Acids Res. 2019, 47, W548–W555. [Google Scholar] [CrossRef]
  8. He, L.; Li, H.; Wu, A.; Peng, Y.; Shu, G.; Yin, G. Functions of N6-methyladenosine and its role in cancer. Mol. Cancer 2019, 18, 176. [Google Scholar] [CrossRef]
  9. Roundtree, I.A.; Evans, M.E.; Pan, T.; He, C. Dynamic RNA modifications in gene expression regulation. Cell 2017, 169, 1187–1200. [Google Scholar] [CrossRef]
  10. Liu, J.; Yue, Y.; Han, D.; Wang, X.; Fu, Y.; Zhang, L.; Jia, G.; Yu, M.; Lu, Z.; Deng, X.; et al. A METTL3–METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat. Chem. Biol. 2014, 10, 93–95. [Google Scholar] [CrossRef]
  11. Warda, A.S.; Kretschmer, J.; Hackert, P.; Lenz, C.; Urlaub, H.; Höbartner, C.; Sloan, K.E.; Bohnsack, M.T. Human METTL16 is a n6-methyladenosine (m6a) methyltransferase that targets pre-mRNAs and various non-coding RNAs. EMBO Rep. 2017, 18, 2004–2014. [Google Scholar] [CrossRef]
  12. Jia, G.; Fu, Y.; Zhao, X.; Dai, Q.; Zheng, G.; Yang, Y.; Yi, C.; Lindahl, T.; Pan, T.; Yang, Y.-G.; et al. N6-Methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat. Chem. Biol. 2011, 7, 885–887. [Google Scholar] [CrossRef]
  13. Wei, J.; Liu, F.; Lu, Z.; Fei, Q.; Ai, Y.; He, P.C.; Shi, H.; Cui, X.; Su, R.; Klungland, A.; et al. Differential m6A, m6Am, and m1A Demethylation Mediated by FTO in the Cell Nucleus and Cytoplasm. Mol. Cell 2018, 71, 973–985.e5. [Google Scholar] [CrossRef]
  14. Zheng, G. ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol. Cell 2013, 49, 18–29. [Google Scholar] [CrossRef]
  15. Qu, J.; Yan, H.; Hou, Y.; Cao, W.; Liu, Y.; Zhang, E.; He, J.; Cai, Z. RNA Demethylase ALKBH5 in cancer: From mechanisms to therapeutic potential. J. Hematol. Oncol. 2022, 15, 8. [Google Scholar] [CrossRef]
  16. Chen, X.-Y. The Role of m6A RNA Methylation in Human Cancer. Mol. Cancer 2019, 18, 1–9. [Google Scholar] [CrossRef]
  17. Wang, X. N6-methyladenosine-dependent regulation of messenger RNA stability. Nature 2014, 505, 117–120. [Google Scholar] [CrossRef]
  18. Wang, X. N6-methyladenosine modulates messenger RNA translation efficiency. Cell 2015, 161, 1388–1399. [Google Scholar] [CrossRef]
  19. Shi, H. YTHDF3 facilitates translation and decay of N6-methyladenosine-Modified RNA. Cell Res. 2017, 27. [Google Scholar] [CrossRef]
  20. Xiang, Y.; Laurent, B.; Hsu, C.-H.; Nachtergaele, S.; Lu, Z.; Sheng, W.; Xu, C.; Chen, H.; Ouyang, J.; Wang, S.; et al. RNA m6A Methylation Regulates the Ultraviolet-Induced DNA Damage Response. Nature 2017, 543, 573–576. [Google Scholar] [CrossRef]
  21. Hsu, P.J. Ythdc2 Is an N6-methyladenosine binding protein that regulates mammalian spermatogenesis. Cell Res. 2017, 27, 1115–1127. [Google Scholar] [CrossRef]
  22. Jain, D.; Puno, M.R.; Meydan, C.; Lailler, N.; Mason, C.E.; Lima, C.D.; Anderson, K.V.; Keeney, S. Ketu mutant mice uncover an essential meiotic function for the ancient RNA helicase YTHDC2. eLife 2018, 7, e30919. [Google Scholar] [CrossRef]
  23. Liao, S.; Sun, H.; Xu, C. YTH domain: A family of N. 6 -methyladenosine (m 6 A) readers. Genom. Proteom. Bioinform. 2018, 16, 99–107. [Google Scholar] [CrossRef]
  24. Huang, H.; Weng, H.; Sun, W.; Qin, X.; Shi, H.; Wu, H.; Zhao, B.S.; Mesquita, A.; Liu, C.; Yuan, C.L.; et al. Recognition of RNA N6-methyladenosine by IGF2BP proteins enhances mRNA stability and translation. Nat. Cell Biol. 2018, 20, 285–295. [Google Scholar] [CrossRef]
  25. Liu, N.; Dai, Q.; Zheng, G.; He, C.; Parisien, M.; Pan, T. N6-methyladenosine-dependent RNA structural switches regulate RNA–protein interactions. Nature 2015, 518, 560–564. [Google Scholar] [CrossRef]
  26. Alarcón, C.R.; Goodarzi, H.; Lee, H.; Liu, X.; Tavazoie, S.; Tavazoie, S.F. HNRNPA2B1 Is a mediator of m6A-dependent nuclear RNA processing events. Cell 2015, 162, 1299–1308. [Google Scholar] [CrossRef]
  27. Meyer, K.D.; Patil, D.P.; Zhou, J.; Zinoviev, A.; Skabkin, M.A.; Elemento, O.; Pestova, T.V.; Qian, S.-B.; Jaffrey, S.R. 5′ UTR m6A promotes cap-independent translation. Cell 2015, 163, 999–1010. [Google Scholar] [CrossRef]
  28. Lu, F.; Zhang, Y. Cell totipotency: Molecular features, induction, and maintenance. Natl. Sci. Rev. 2015, 2, 217–225. [Google Scholar] [CrossRef]
  29. Schöller, E.; Weichmann, F.; Treiber, T.; Ringle, S.; Treiber, N.; Flatley, A.; Feederle, R.; Bruckmann, A.; Meister, G. Interactions, localization, and phosphorylation of the m6 a generating METTL3–METTL14–WTAP complex. RNA 2018, 24, 499–512. [Google Scholar] [CrossRef]
  30. Ping, X.-L.; Sun, B.-F.; Wang, L.; Xiao, W.; Yang, X.; Wang, W.-J.; Adhikari, S.; Shi, Y.; Lv, Y.; Chen, Y.-S.; et al. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res. 2014, 24, 177–189. [Google Scholar] [CrossRef]
  31. Wang, F.; Zhang, J.; Lin, X.; Yang, L.; Zhou, Q.; Mi, X.; Li, Q.; Wang, S.; Li, D.; Liu, X.-M.; et al. METTL16 Promotes translation and lung tumorigenesis by sequestering cytoplasmic eIF4E2. Cell Rep. 2023, 42, 112150. [Google Scholar] [CrossRef]
  32. Satterwhite, E.R.; Mansfield, K.D. RNA methyltransferase METTL16: Targets and function. WIREs RNA 2022, 13, e1681. [Google Scholar] [CrossRef]
  33. Yue, Y.; Liu, J.; Cui, X.; Cao, J.; Luo, G.; Zhang, Z.; Cheng, T.; Gao, M.; Shu, X.; Ma, H.; et al. VIRMA mediates preferential m6A mRNA methylation in 3′utr and near stop codon and associates with alternative polyadenylation. Cell Discov. 2018, 4, 10. [Google Scholar] [CrossRef] [PubMed]
  34. Patil, D.P.; Chen, C.-K.; Pickering, B.F.; Chow, A.; Jackson, C.; Guttman, M.; Jaffrey, S.R. m6A RNA methylation promotes XIST-mediated transcriptional repression. Nature 2016, 537, 369–373. [Google Scholar] [CrossRef] [PubMed]
  35. Wen, J.; Lv, R.; Ma, H.; Shen, H.; He, C.; Wang, J.; Jiao, F.; Liu, H.; Yang, P.; Tan, L.; et al. Zc3h13 regulates nuclear RNA m6A methylation and mouse embryonic stem cell self-renewal. Mol. Cell 2018, 69, 1028–1038.e6. [Google Scholar] [CrossRef] [PubMed]
  36. Bawankar, P.; Lence, T.; Paolantoni, C.; Haussmann, I.U.; Kazlauskiene, M.; Jacob, D.; Heidelberger, J.B.; Richter, F.M.; Nallasivan, M.P.; Morin, V.; et al. Hakai is required for stabilization of core components of the m6A mRNA methylation machinery. Nat. Commun. 2021, 12, 3778. [Google Scholar] [CrossRef]
  37. Zhao, X.; Yang, Y.; Sun, B.-F.; Shi, Y.; Yang, X.; Xiao, W.; Hao, Y.-J.; Ping, X.-L.; Chen, Y.-S.; Wang, W.-J.; et al. FTO-dependent demethylation of n6-methyladenosine regulates mRNA splicing and is required for adipogenesis. Cell Res. 2014, 24, 1403–1419. [Google Scholar] [CrossRef]
  38. Lane, N.; Martin, W. The energetics of genome complexity. Nature 2010, 467, 929–934. [Google Scholar] [CrossRef]
  39. Friedman, J.R.; Nunnari, J. Mitochondrial form and function. Nature 2014, 505, 335–343. [Google Scholar] [CrossRef]
  40. McBride, H.M.; Neuspiel, M.; Wasiak, S. Mitochondria: More than just a powerhouse. Curr. Biol. 2006, 16, R551–R560. [Google Scholar] [CrossRef]
  41. Walsh, C.T.; Tu, B.P.; Tang, Y. Eight kinetically stable but thermodynamically activated molecules that power cell metabolism. Chem. Rev. 2018, 118, 1460–1494. [Google Scholar] [CrossRef]
  42. Watt, I.N.; Montgomery, M.G.; Runswick, M.J.; Leslie, A.G.W.; Walker, J.E. Bioenergetic cost of making an adenosine triphosphate molecule in animal mitochondria. Proc. Natl. Acad. Sci. USA. 2010, 107, 16823–16827. [Google Scholar] [CrossRef]
  43. Sun, F.; Zhou, Q.; Pang, X.; Xu, Y.; Rao, Z. Revealing various coupling of electron transfer and proton pumping in mitochondrial respiratory chain. Curr. Opin. Struct. Biol. 2013, 23, 526–538. [Google Scholar] [CrossRef] [PubMed]
  44. Prasun, P. Mitochondrial dysfunction in metabolic syndrome. Biochim. Et. Biophys. Acta (BBA)—Mol. Basis Dis. 2020, 1866, 165838. [Google Scholar] [CrossRef] [PubMed]
  45. Monzel, A.S.; Enríquez, J.A.; Picard, M. Multifaceted mitochondria: Moving mitochondrial science beyond function and dysfunction. Nat. Metab. 2023, 5, 546–562. [Google Scholar] [CrossRef] [PubMed]
  46. Nunnari, J.; Suomalainen, A. Mitochondria: In sickness and in health. Cell 2012, 148, 1145–1159. [Google Scholar] [CrossRef]
  47. Doke, T.; Susztak, K. The multifaceted role of kidney tubule mitochondrial dysfunction in kidney disease development. Trends Cell Biol. 2022, 32, 841–853. [Google Scholar] [CrossRef]
  48. Wang, H.; Sun, Y.; Pi, C.; Yu, X.; Gao, X.; Zhang, C.; Sun, H.; Zhang, H.; Shi, Y.; He, X. Nicotinamide mononucleotide supplementation improves mitochondrial dysfunction and rescues cellular senescence by NAD+/Sirt3 pathway in mesenchymal stem cells. IJMS 2022, 23, 14739. [Google Scholar] [CrossRef]
  49. Mito, T.; Vincent, A.E.; Faitg, J.; Taylor, R.W.; Khan, N.A.; McWilliams, T.G.; Suomalainen, A. Mosaic dysfunction of mitophagy in mitochondrial muscle disease. Cell Metab. 2022, 34, 197–208.e5. [Google Scholar] [CrossRef]
  50. Gu, Y.; Zhao, X.; Zhang, N.; Yang, Y.; Yi, Y.; Shao, Q.; Liu, M.; Zhang, X. Mitochondrial dysfunction as a therapeutic strategy for neurodegenerative diseases: Current insights and future directions. Ageing Res. Rev. 2024, 102, 102577. [Google Scholar] [CrossRef]
  51. Russell, O.M.; Gorman, G.S.; Lightowlers, R.N.; Turnbull, D.M. Mitochondrial diseases: Hope for the future. Cell 2020, 181, 168–188. [Google Scholar] [CrossRef] [PubMed]
  52. Wallace, D.C. Mitochondrial genetic medicine. Nat. Genet. 2018, 50, 1642–1649. [Google Scholar] [CrossRef]
  53. Pfanner, N.; Warscheid, B.; Wiedemann, N. Mitochondrial proteins: From biogenesis to functional networks. Nat. Rev. Mol. Cell Biol. 2019, 20, 267–284. [Google Scholar] [CrossRef]
  54. Hoogenraad, N.J.; Ryan, M.T. Translocation of proteins into mitochondria. IUBMB Life 2001, 51, 345–350. [Google Scholar] [CrossRef]
  55. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef] [PubMed]
  56. Simmons, E.C.; Scholpa, N.E.; Schnellmann, R.G. Mitochondrial biogenesis as a therapeutic target for traumatic and neurodegenerative CNS diseases. Exp. Neurol. 2020, 329, 113309. [Google Scholar] [CrossRef] [PubMed]
  57. Popov, L. Mitochondrial biogenesis: An update. J. Cell. Mol. Medi 2020, 24, 4892–4899. [Google Scholar] [CrossRef]
  58. Kozhukhar, N.; Alexeyev, M.F. Limited predictive value of TFAM in mitochondrial biogenesis. Mitochondrion 2019, 49, 156–165. [Google Scholar] [CrossRef]
  59. McQuate, A.; Raible, D.W. Finding the balance: The elusive mechanisms underlying auditory hair cell mitochondrial biogenesis and mitophagy. Hear. Res. 2023, 428, 108664. [Google Scholar] [CrossRef]
  60. Wang, S.; Long, H.; Hou, L.; Feng, B.; Ma, Z.; Wu, Y.; Zeng, Y.; Cai, J.; Zhang, D.; Zhao, G. The mitophagy pathway and its implications in human diseases. Sig. Transduct. Target. Ther. 2023, 8, 304. [Google Scholar] [CrossRef]
  61. Chan, D.C. Mitochondrial dynamics and its involvement in disease. Annu. Rev. Pathol. Mech. Dis. 2020, 15, 235–259. [Google Scholar] [CrossRef]
  62. Chen, W.; Zhao, H.; Li, Y. Mitochondrial dynamics in health and disease: Mechanisms and potential targets. Sig. Transduct. Target. Ther. 2023, 8, 333. [Google Scholar] [CrossRef] [PubMed]
  63. Youle, R.J.; Van Der Bliek, A.M. Mitochondrial fission, fusion, and stress. Science 2012, 337, 1062–1065. [Google Scholar] [CrossRef] [PubMed]
  64. Otera, H.; Ishihara, N.; Mihara, K. New insights into the function and regulation of mitochondrial fission. Biochim. Et. Biophys. Acta (BBA)—Mol. Cell Res. 2013, 1833, 1256–1268. [Google Scholar] [CrossRef] [PubMed]
  65. Fonseca, T.B.; Sánchez-Guerrero, Á.; Milosevic, I.; Raimundo, N. Mitochondrial fission requires DRP1 but not dynamins. Nature 2019, 570, E34–E42. [Google Scholar] [CrossRef]
  66. Chang, C.; Blackstone, C. Dynamic regulation of mitochondrial fission through modification of the dynamin-related protein drp1. Ann. New York Acad. Sci. 2010, 1201, 34–39. [Google Scholar] [CrossRef]
  67. Adaniya, S.M.; O-Uchi, J.; Cypress, M.W.; Kusakari, Y.; Jhun, B.S. Posttranslational modifications of mitochondrial fission and fusion proteins in cardiac physiology and pathophysiology. Am. J. Physiol. -Cell Physiol. 2019, 316, C583–C604. [Google Scholar] [CrossRef]
  68. Burté, F.; Carelli, V.; Chinnery, P.F.; Yu-Wai-Man, P. Disturbed mitochondrial dynamics and neurodegenerative disorders. Nat. Rev. Neurol. 2015, 11, 11–24. [Google Scholar] [CrossRef]
  69. Galloway, C.A.; Lee, H.; Yoon, Y. Mitochondrial morphology—emerging role in bioenergetics. Free Radic. Biol. Med. 2012, 53, 2218–2228. [Google Scholar] [CrossRef]
  70. Labbé, K.; Murley, A.; Nunnari, J. Determinants and functions of mitochondrial behavior. Annu. Rev. Cell Dev. Biol. 2014, 30, 357–391. [Google Scholar] [CrossRef]
  71. Cipolat, S.; De Brito, O.M.; Dal Zilio, B.; Scorrano, L. OPA1 Requires mitofusin 1 to promote mitochondrial fusion. Proc. Natl. Acad. Sci. USA 2004, 101, 15927–15932. [Google Scholar] [CrossRef] [PubMed]
  72. Song, Z.; Ghochani, M.; McCaffery, J.M.; Frey, T.G.; Chan, D.C. Mitofusins and OPA1 mediate sequential steps in mitochondrial membrane fusion. Mol. Biol. Cell 2009, 20, 3525–3532. [Google Scholar] [CrossRef]
  73. Pyakurel, A.; Savoia, C.; Hess, D.; Scorrano, L. Extracellular regulated kinase phosphorylates mitofusin 1 to control mitochondrial morphology and apoptosis. Mol. Cell 2015, 58, 244–254. [Google Scholar] [CrossRef]
  74. Zheng, J.; Croteau, D.L.; Bohr, V.A.; Akbari, M. Diminished OPA1 expression and impaired mitochondrial morphology and homeostasis in aprataxin-deficient cells. Nucleic Acids Res. 2019, 47, 4086–4110. [Google Scholar] [CrossRef]
  75. Quintana-Cabrera, R.; Scorrano, L. Determinants and outcomes of mitochondrial dynamics. Mol. Cell 2023, 83, 857–876. [Google Scholar] [CrossRef]
  76. Westermann, B. Bioenergetic role of mitochondrial fusion and fission. Biochim. Et. Biophys. Acta (BBA)—Bioenerg. 2012, 1817, 1833–1838. [Google Scholar] [CrossRef]
  77. Palikaras, K.; Lionaki, E.; Tavernarakis, N. Mechanisms of mitophagy in cellular homeostasis, physiology and pathology. Nat. Cell Biol. 2018, 20, 1013–1022. [Google Scholar] [CrossRef] [PubMed]
  78. Lazarou, M.; Sliter, D.A.; Kane, L.A.; Sarraf, S.A.; Wang, C.; Burman, J.L.; Sideris, D.P.; Fogel, A.I.; Youle, R.J. The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature 2015, 524, 309–314. [Google Scholar] [CrossRef] [PubMed]
  79. Lu, Y.; Li, Z.; Zhang, S.; Zhang, T.; Liu, Y.; Zhang, L. Cellular mitophagy: Mechanism, roles in diseases and small molecule pharmacological regulation. Theranostics 2023, 13, 736–766. [Google Scholar] [CrossRef]
  80. Pradeepkiran, J.A.; Reddy, P.H. Defective mitophagy in Alzheimer’s disease. Ageing Res. Rev. 2020, 64, 101191. [Google Scholar] [CrossRef]
  81. Uoselis, L.; Nguyen, T.N.; Lazarou, M. Mitochondrial degradation: Mitophagy and beyond. Mol. Cell 2023, 83, 3404–3420. [Google Scholar] [CrossRef] [PubMed]
  82. Palikaras, K.; Daskalaki, I.; Markaki, M.; Tavernarakis, N. Mitophagy and age-related pathologies: Development of new therapeutics by targeting mitochondrial turnover. Pharmacol. Ther. 2017, 178, 157–174. [Google Scholar] [CrossRef]
  83. Lizama, B.N.; Chu, C.T. Neuronal autophagy and mitophagy in Parkinson’s disease. Mol. Asp. Med. 2021, 82, 100972. [Google Scholar] [CrossRef]
  84. Shadel, G.S.; Horvath, T.L. Mitochondrial ROS signaling in organismal homeostasis. Cell 2015, 163, 560–569. [Google Scholar] [CrossRef]
  85. Brand, M.D. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radic. Biol. Med. 2016, 100, 14–31. [Google Scholar] [CrossRef]
  86. Kuksal, N.; Chalker, J.; Mailloux, R.J. Progress in understanding the molecular oxygen paradox—Function of mitochondrial reactive oxygen species in cell signaling. Biol. Chem. 2017, 398, 1209–1227. [Google Scholar] [CrossRef] [PubMed]
  87. Young, A.; Gill, R.; Mailloux, R.J. Protein S-Glutathionylation: The linchpin for the transmission of regulatory information on redox buffering capacity in mitochondria. Chem. -Biol. Interact. 2019, 299, 151–162. [Google Scholar] [CrossRef]
  88. Murphy, M.P. Mitochondrial thiols in antioxidant protection and redox signaling: Distinct roles for glutathionylation and other thiol modifications. Antioxid. Redox Signal. 2012, 16, 476–495. [Google Scholar] [CrossRef] [PubMed]
  89. Wood, Z.A.; Schröder, E.; Robin Harris, J.; Poole, L.B. Structure, mechanism and regulation of peroxiredoxins. Trends Biochem. Sci. 2003, 28, 32–40. [Google Scholar] [CrossRef]
  90. Slade, L.; Chalker, J.; Kuksal, N.; Young, A.; Gardiner, D.; Mailloux, R.J. Examination of the superoxide/hydrogen peroxide forming and quenching potential of mouse liver mitochondria. Biochim. Et. Biophys. Acta (BBA) - Gen. Subj. 2017, 1861, 1960–1969. [Google Scholar] [CrossRef]
  91. Wu, J.Q.; Kosten, T.R.; Zhang, X.Y. Free radicals, antioxidant defense systems, and schizophrenia. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2013, 46, 200–206. [Google Scholar] [CrossRef] [PubMed]
  92. Singh, A.; Kukreti, R.; Saso, L.; Kukreti, S. Oxidative stress: A key modulator in neurodegenerative diseases. Molecules 2019, 24, 1583. [Google Scholar] [CrossRef]
  93. Kirkinezos, I.G.; Moraes, C.T. Reactive oxygen species and mitochondrial diseases. Semin. Cell Dev. Biol. 2001, 12, 449–457. [Google Scholar] [CrossRef]
  94. Wang, R.; Wang, M.; He, S.; Sun, G.; Sun, X. Targeting calcium homeostasis in myocardial ischemia/reperfusion injury: An overview of regulatory mechanisms and therapeutic reagents. Front. Pharmacol. 2020, 11, 872. [Google Scholar] [CrossRef] [PubMed]
  95. Boyman, L.; Lederer, W.J. How the mitochondrial calcium uniporter complex (MCUcx) works. proc. Natl. Acad. Sci. USA 2020, 117, 22634–22636. [Google Scholar] [CrossRef]
  96. Chen, J.; Sitsel, A.; Benoy, V.; Sepúlveda, M.R.; Vangheluwe, P. Primary active Ca2+ transport systems in health and disease. Cold Spring Harb. Perspect. Biol. 2020, 12, a035113. [Google Scholar] [CrossRef] [PubMed]
  97. Alevriadou, B.R.; Patel, A.; Noble, M.; Ghosh, S.; Gohil, V.M.; Stathopulos, P.B.; Madesh, M. Molecular nature and physiological role of the mitochondrial calcium uniporter channel. Am. J. Physiol. -Cell Physiol. 2021, 320, C465–C482. [Google Scholar] [CrossRef]
  98. Pathak, T.; Trebak, M. Mitochondrial Ca2+ signaling. Pharmacol. Ther. 2018, 192, 112–123. [Google Scholar] [CrossRef]
  99. Zhang, D.; Wang, F.; Li, P.; Gao, Y. Mitochondrial Ca2+ homeostasis: Emerging roles and clinical significance in cardiac remodeling. IJMS 2022, 23, 3025. [Google Scholar] [CrossRef]
  100. Strubbe-Rivera, J.O.; Chen, J.; West, B.A.; Parent, K.N.; Wei, G.-W.; Bazil, J.N. Modeling the effects of calcium overload on mitochondrial ultrastructural remodeling. Appl. Sci. 2021, 11, 2071. [Google Scholar] [CrossRef]
  101. Walkon, L.L.; Strubbe-Rivera, J.O.; Bazil, J.N. Calcium overload and mitochondrial metabolism. Biomolecules 2022, 12, 1891. [Google Scholar] [CrossRef] [PubMed]
  102. Cascella, R.; Cecchi, C. Calcium dyshomeostasis in Alzheimer’s disease pathogenesis. IJMS 2021, 22, 4914. [Google Scholar] [CrossRef] [PubMed]
  103. Saqirile; Deng, Y.; Li, K.; Yan, W.; Li, K.; Wang, C. Gene expression regulation and the signal transduction of programmed cell death. CIMB 2024, 46, 10264–10298. [Google Scholar] [CrossRef] [PubMed]
  104. Glover, H.L.; Schreiner, A.; Dewson, G.; Tait, S.W.G. Mitochondria and cell death. Nat. Cell Biol. 2024, 26, 1434–1446. [Google Scholar] [CrossRef]
  105. Bock, F.J.; Tait, S.W.G. Mitochondria as multifaceted regulators of cell death. Nat. Rev. Mol. Cell Biol. 2020, 21, 85–100. [Google Scholar] [CrossRef]
  106. Bauer, T.M.; Murphy, E. Role of mitochondrial calcium and the permeability transition pore in regulating cell death. Circ. Res. 2020, 126, 280–293. [Google Scholar] [CrossRef]
  107. Nguyen, T.T.; Wei, S.; Nguyen, T.H.; Jo, Y.; Zhang, Y.; Park, W.; Gariani, K.; Oh, C.-M.; Kim, H.H.; Ha, K.-T.; et al. Mitochondria-associated programmed cell death as a therapeutic target for age-related disease. Exp. Mol. Med. 2023, 55, 1595–1619. [Google Scholar] [CrossRef]
  108. Bertheloot, D.; Latz, E.; Franklin, B.S. Necroptosis, pyroptosis and apoptosis: An intricate game of cell death. Cell Mol. Immunol. 2021, 18, 1106–1121. [Google Scholar] [CrossRef]
  109. Yang, C.; Yang, X.; Harrington, A.; Potts, C.; Kaija, A.; Ryzhova, L.; Liaw, L. Notch signaling regulates mouse perivascular adipose tissue function via mitochondrial pathways. Genes. 2023, 14, 1964. [Google Scholar] [CrossRef]
  110. Lee, H.; Zandkarimi, F.; Zhang, Y.; Meena, J.K.; Kim, J.; Zhuang, L.; Tyagi, S.; Ma, L.; Westbrook, T.F.; Steinberg, G.R.; et al. Energy-stress-mediated AMPK activation inhibits ferroptosis. Nat. Cell Biol. 2020, 22, 225–234. [Google Scholar] [CrossRef]
  111. Gan, B. Mitochondrial Regulation of Ferroptosis. J. Cell Biol. 2021, 220, e202105043. [Google Scholar] [CrossRef]
  112. Chipuk, J.E.; Mohammed, J.N.; Gelles, J.D.; Chen, Y. Mechanistic connections between mitochondrial biology and regulated cell death. Dev. Cell 2021, 56, 1221–1233. [Google Scholar] [CrossRef] [PubMed]
  113. Gao, M.; Yi, J.; Zhu, J.; Minikes, A.M.; Monian, P.; Thompson, C.B.; Jiang, X. Role of mitochondria in ferroptosis. Mol. Cell 2019, 73, 354–363.e3. [Google Scholar] [CrossRef]
  114. Ding, H.; Li, Y.; Li, X.; Liu, X.; Wang, K.; Wen, M.; Jiang, W.; Zeng, H. Hypercapnia promotes microglial pyroptosis via inhibiting mitophagy in hypoxemic adult rats. CNS Neurosci. Ther. 2020, 26, 1134–1146. [Google Scholar] [CrossRef]
  115. Liu, L.; Li, Y.; Chen, G.; Chen, Q. Crosstalk between mitochondrial biogenesis and mitophagy to maintain mitochondrial homeostasis. J. Biomed. Sci. 2023, 30, 86. [Google Scholar] [CrossRef] [PubMed]
  116. Wongrakpanich, A.; Geary, S.M.; Joiner, M.A.; Anderson, M.E.; Salem, A.K. Mitochondria-targeting particles. Nanomedicine 2014, 9, 2531–2543. [Google Scholar] [CrossRef] [PubMed]
  117. Batheja, S.; Gupta, S.; Tejavath, K.K.; Gupta, U. TPP-based conjugates: Potential targeting ligands. Drug Discov. Today 2024, 29, 103983. [Google Scholar] [CrossRef]
  118. Nam, H.Y.; Hong, J.-A.; Choi, J.; Shin, S.; Cho, S.K.; Seo, J.; Lee, J. Mitochondria-targeting peptoids. Bioconjugate Chem. 2018, 29, 1669–1676. [Google Scholar] [CrossRef]
  119. Xu, R.; Huang, L.; Liu, J.; Zhang, Y.; Xu, Y.; Li, R.; Su, S.; Xu, X. Remodeling of mitochondrial metabolism by a mitochondria-targeted RNAi nanoplatform for effective cancer therapy. Small 2024, 20. [Google Scholar] [CrossRef]
  120. Liao, J.; He, W.; Zhang, S.; Wang, J.; Cheng, H.; Gong, L.; Li, Y.; Sun, Y.; Lu, D.; Zhang, C.; et al. Magnetic field driven ceria nanosystems for mitochondria targeted therapy of ischemic stroke. Adv. Funct. Mater. 2025, 2423291. [Google Scholar] [CrossRef]
  121. Wen, C.; Tang, J.; Wu, M.; Liu, H.; Lin, X.; Fan, M.; Liu, G.; Zhang, J.; Liang, L.; Liu, X.; et al. Preparation, characterization, and stability of pectin-whey protein isolate-based nanoparticles with mitochondrial targeting ability. Int. J. Biol. Macromol. 2025, 301, 140383. [Google Scholar] [CrossRef]
  122. Duan, Z.; Li, L.; Zhan, Q.; Chen, J.; Li, Q.; Liu, R.; Tu, Y. Mitochondria-targeting type-i photodynamic therapy based on phenothiazine for realizing enhanced immunogenic cancer cell death via mitochondrial oxidative stress. Int. J. Nurs. 2025, 20, 125–139. [Google Scholar] [CrossRef]
  123. Wang, W.; Yao, S.-Y.; Luo, J.; Ding, C.; Huang, Q.; Yang, Y.; Shi, Z.; Lin, J.; Pan, Y.-C.; Zeng, X.; et al. Engineered hypoxia-responsive albumin nanoparticles mediating mitophagy regulation for cancer therapy. Nat. Commun. 2025, 16, 596. [Google Scholar] [CrossRef] [PubMed]
  124. Lv, Y.; Song, B.; Yang, G.; Wang, Y.; Wu, Z.; Si, M.; Yang, Z.; Chen, H.; Liu, C.; Li, M.; et al. In situ transformable nanoparticle effectively suppresses bladder cancer by damaging mitochondria and blocking mitochondrial autophagy flux. Adv. Sci. 2025, 12, 2409425. [Google Scholar] [CrossRef] [PubMed]
  125. Chen, S.; Sun, Y.; Xie, Y.; Liu, Y.; Hu, H.; Xie, C.; Xu, S.; Zhang, Z.; Zhang, J.; Shen, Y.; et al. Mitochondria-targeted icaritin nanoparticles induce immunogenic cell death in hepatocellular carcinoma. ACS Appl. Mater. Interfaces 2025, 17, 2899–2910. [Google Scholar] [CrossRef]
  126. Qian, L.; Zhu, Y.; Deng, C.; Liang, Z.; Chen, J.; Chen, Y.; Wang, X.; Liu, Y.; Tian, Y.; Yang, Y. Peroxisome Proliferator-activated receptor gamma coactivator-1 (PGC-1) family in physiological and pathophysiological process and diseases. Sig. Transduct. Target. Ther. 2024, 9, 50. [Google Scholar] [CrossRef]
  127. Scarpulla, R.C. Nuclear control of respiratory chain expression by nuclear respiratory factors and PGC-1-related coactivator. Ann. N. Y. Acad. Sci. 2008, 1147, 321–334. [Google Scholar] [CrossRef]
  128. Losón, O.C.; Song, Z.; Chen, H.; Chan, D.C. Fis1, Mff, MiD49, and MiD51 Mediate Drp1 recruitment in mitochondrial fission. Mol. Biol. Cell 2013, 24, 659–667. [Google Scholar] [CrossRef]
  129. Tábara, L.-C.; Segawa, M.; Prudent, J. Molecular mechanisms of mitochondrial dynamics. Nat. Rev. Mol. Cell Biol. 2025, 26, 123–146. [Google Scholar] [CrossRef]
  130. Panigrahi, D.P.; Praharaj, P.P.; Behera, B.P.; Patra, S.; Patil, S.; Patro, B.S.; Bhutia, S.K. The inner mitochondrial membrane fission protein MTP18 Serves as a mitophagy receptor to prevent apoptosis in oral cancer. J. Cell Sci. 2023, 136, jcs259986. [Google Scholar] [CrossRef]
  131. Li, Y.; Zheng, W.; Lu, Y.; Zheng, Y.; Pan, L.; Wu, X.; Yuan, Y.; Shen, Z.; Ma, S.; Zhang, X.; et al. BNIP3L/NIX-mediated mitophagy: Molecular mechanisms and implications for human disease. Cell Death Dis. 2021, 13, 14. [Google Scholar] [CrossRef] [PubMed]
  132. Niemi, N.M.; Friedman, J.R. Coordinating BNIP3/NIX-mediated mitophagy in space and time. Biochem. Soc. Trans. 2024, 52, 1969–1979. [Google Scholar] [CrossRef]
  133. Bi, Y.; Liu, S.; Qin, X.; Abudureyimu, M.; Wang, L.; Zou, R.; Ajoolabady, A.; Zhang, W.; Peng, H.; Ren, J.; et al. FUNDC1 interacts with GPx4 to govern hepatic ferroptosis and fibrotic injury through a mitophagy-dependent manner. J. Adv. Res. 2024, 55, 45–60. [Google Scholar] [CrossRef]
  134. Pei, J.; Pan, X.; Wei, G.; Hua, Y. Research progress of glutathione peroxidase family (GPX) in redoxidation. Front. Pharmacol. 2023, 14, 1147414. [Google Scholar] [CrossRef]
  135. Tanaka, T. Thioredoxin-2 (TRX-2) is an essential gene regulating mitochondria-dependent apoptosis. EMBO J. 2002, 21, 1695–1703. [Google Scholar] [CrossRef]
  136. Wu, M.; Deng, C.; Lo, T.-H.; Chan, K.-Y.; Li, X.; Wong, C.-M. Peroxiredoxin, senescence, and cancer. Cells 2022, 11, 1772. [Google Scholar] [CrossRef] [PubMed]
  137. Baughman, J.M.; Perocchi, F.; Girgis, H.S.; Plovanich, M.; Belcher-Timme, C.A.; Sancak, Y.; Bao, X.R.; Strittmatter, L.; Goldberger, O.; Bogorad, R.L.; et al. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature 2011, 476, 341–345. [Google Scholar] [CrossRef] [PubMed]
  138. Wang, J.; Jiang, J.; Hu, H.; Chen, L. MCU complex: Exploring emerging targets and mechanisms of mitochondrial physiology and pathology. J. Adv. Res. 2025, 68, 271–298. [Google Scholar] [CrossRef]
  139. Perocchi, F.; Gohil, V.M.; Girgis, H.S.; Bao, X.R.; McCombs, J.E.; Palmer, A.E.; Mootha, V.K. MICU1 encodes a mitochondrial Ef hand protein required for Ca2+ uptake. Nature 2010, 467, 291–296. [Google Scholar] [CrossRef]
  140. Patron, M.; Checchetto, V.; Raffaello, A.; Teardo, E.; Vecellio Reane, D.; Mantoan, M.; Granatiero, V.; Szabò, I.; De Stefani, D.; Rizzuto, R. MICU1 and MICU2 finely tune the mitochondrial Ca2+ uniporter by exerting opposite effects on MCU activity. Mol. Cell 2014, 53, 726–737. [Google Scholar] [CrossRef]
  141. Tomar, D.; Dong, Z.; Shanmughapriya, S.; Koch, D.A.; Thomas, T.; Hoffman, N.E.; Timbalia, S.A.; Goldman, S.J.; Breves, S.L.; Corbally, D.P.; et al. MCUR1 is a scaffold factor for the MCU complex function and promotes mitochondrial bioenergetics. Cell Rep. 2016, 15, 1673–1685. [Google Scholar] [CrossRef] [PubMed]
  142. Huo, J.; Prasad, V.; Grimes, K.M.; Vanhoutte, D.; Blair, N.S.; Lin, S.-C.; Bround, M.J.; Bers, D.M.; Molkentin, J.D. MCUb is an inducible regulator of calcium-dependent mitochondrial metabolism and substrate utilization in muscle. Cell Reports 2023, 42, 113465. [Google Scholar] [CrossRef] [PubMed]
  143. Hoffman, N.E.; Chandramoorthy, H.C.; Shanmughapriya, S.; Zhang, X.Q.; Vallem, S.; Doonan, P.J.; Malliankaraman, K.; Guo, S.; Rajan, S.; Elrod, J.W.; et al. SLC25A23 augments mitochondrial Ca2+ uptake, interacts with MCU, and induces oxidative stress–mediated cell death. Mol. Biol. Cell 2014, 25, 936–947. [Google Scholar] [CrossRef]
  144. Cosentino, K.; Hertlein, V.; Jenner, A.; Dellmann, T.; Gojkovic, M.; Peña-Blanco, A.; Dadsena, S.; Wajngarten, N.; Danial, J.S.H.; Thevathasan, J.V.; et al. The interplay between BAX and BAK tunes apoptotic pore growth to control mitochondrial-DNA-mediated inflammation. Mol. Cell 2022, 82, 933–949.e9. [Google Scholar] [CrossRef]
  145. Mifflin, L.; Ofengeim, D.; Yuan, J. Receptor-interacting protein kinase 1 (RIPK1) as a therapeutic target. Nat. Rev. Drug Discov. 2020, 19, 553–571. [Google Scholar] [CrossRef]
  146. Liu, S.; Joshi, K.; Denning, M.F.; Zhang, J. RIPK3 signaling and its role in the pathogenesis of cancers. Cell. Mol. Life Sci. 2021, 78, 7199–7217. [Google Scholar] [CrossRef]
  147. Bersuker, K.; Hendricks, J.M.; Li, Z.; Magtanong, L.; Ford, B.; Tang, P.H.; Roberts, M.A.; Tong, B.; Maimone, T.J.; Zoncu, R.; et al. The CoQ oxidoreductase FSP1 acts parallel to GPX4 to inhibit ferroptosis. Nature 2019, 575, 688–692. [Google Scholar] [CrossRef]
  148. Matilainen, O.; Quirós, P.M.; Auwerx, J. Mitochondria and epigenetics—crosstalk in homeostasis and stress. Trends Cell Biol. 2017, 27, 453–463. [Google Scholar] [CrossRef] [PubMed]
  149. Kahl, M.; Xu, Z.; Arumugam, S.; Edens, B.M.; Fischietti, M.; Zhu, A.C.; Platanias, L.C.; He, C.; Zhuang, X.; Ma, Y.C. m6A RNA methylation regulates mitochondrial function. Hum. Mol. Genet. 2024, 33, 969–980. [Google Scholar] [CrossRef]
  150. Wang, X.; Huang, N.; Yang, M.; Wei, D.; Tai, H.; Han, X.; Gong, H.; Zhou, J.; Qin, J.; Wei, X.; et al. FTO is required for myogenesis by positively regulating mTOR-PGC-1α pathway-mediated mitochondria biogenesis. Cell Death Dis. 2017, 8, e2702. [Google Scholar] [CrossRef]
  151. Kang, H.; Zhang, Z.; Yu, L.; Li, Y.; Liang, M.; Zhou, L. FTO Reduces mitochondria and promotes hepatic fat accumulation through rna demethylation. J. Cell. Biochem. 2018, 119, 5676–5685. [Google Scholar] [CrossRef] [PubMed]
  152. Du, Y.D.; Guo, W.Y.; Han, C.H.; Wang, Y.; Chen, X.S.; Li, D.W.; Liu, J.L.; Zhang, M.; Zhu, N.; Wang, X. N6-Methyladenosine demethylase FTO impairs hepatic ischemia–reperfusion injury via inhibiting Drp1-mediated mitochondrial fragmentation. Cell Death Dis. 2021, 12, 442. [Google Scholar] [CrossRef] [PubMed]
  153. Qi, P.; Zhang, W.; Gao, Y.; Chen, S.; Jiang, M.; He, R.; Chen, W.; Wei, X.; Hu, B.; Xu, H.; et al. N6-methyladenosine demethyltransferase FTO alleviates sepsis by upregulating BNIP3 to induce mitophagy. J. Cell. Physiol. 2024, 239, e31448. [Google Scholar] [CrossRef]
  154. Hou, L.; Li, S.; Li, S.; Wang, R.; Zhao, M.; Liu, X. FTO inhibits oxidative stress by mediating m6A demethylation of Nrf2 to alleviate cerebral ischemia/reperfusion injury. J. Physiol. Biochem. 2023, 79, 133–146. [Google Scholar] [CrossRef] [PubMed]
  155. Yin, R.; Chang, J.; Li, Y.; Gao, Z.; Qiu, Q.; Wang, Q.; Han, G.; Chai, J.; Feng, M.; Wang, P.; et al. Differential m6A RNA landscapes across hematopoiesis reveal a role for IGF2BP2 in preserving hematopoietic stem cell function. Cell Stem Cell 2022, 29, 149–159.e7. [Google Scholar] [CrossRef]
  156. Chen, H.; Xing, H.; Zhong, C.; Lin, X.; Chen, R.; Luo, N.; Chen, L.; Huang, Y. METTL3 confers protection against mitochondrial dysfunction and cognitive impairment in an alzheimer disease mouse model by upregulating Mfn2 via N6-methyladenosine modification. J. Neuropathol. Exp. Neurol. 2024, 83, 606–614. [Google Scholar] [CrossRef]
  157. Yang, P.; Wang, Y.; Ge, W.; Jing, Y.; Hu, H.; Yin, J.; Xue, M.; Wang, Y.; Li, X.; Li, X.; et al. m6A Methyltransferase METTL3 contributes to sympathetic hyperactivity post-Mi via promoting TRAF6-dependent mitochondrial ROS production. Free Radic. Biol. Med. 2023, 209, 342–354. [Google Scholar] [CrossRef]
  158. Shetty, R.; Noland, R.; Nandi, G.; Suzuki, C.K. Powering down the mitochondrial lonp1 protease: A novel strategy for anticancer therapeutics. Expert. Opin. Ther. Targets 2024, 28, 9–15. [Google Scholar] [CrossRef]
  159. Zhang, C.; Shen, S.; Xu, L.; Li, M.; Tian, B.; Yao, L.; Zhu, X. LONP1 Alleviates ageing-related renal fibrosis by maintaining mitochondrial homeostasis. J. Cell. Mol. Medi. 2024, 28, e70090. [Google Scholar] [CrossRef]
  160. Zhang, H.; Wu, D.; Wang, Y.; Guo, K.; Spencer, C.B.; Ortoga, L.; Qu, M.; Shi, Y.; Shao, Y.; Wang, Z.; et al. METTL3-mediated N6-methyladenosine exacerbates ferroptosis via m6A-IGF2BP2-dependent mitochondrial metabolic reprogramming in sepsis-induced acute lung injury. Clin. Transl. Med. 2023, 13, e1389. [Google Scholar] [CrossRef]
  161. Tu, B.; Song, K.; Zhou, Y.; Sun, H.; Liu, Z.-Y.; Lin, L.-C.; Ding, J.-F.; Sha, J.-M.; Shi, Y.; Yang, J.-J.; et al. METTL3 boosts mitochondrial fission and induces cardiac fibrosis by enhancing LncRNA GAS5 methylation. Pharmacol. Res. 2023, 194, 106840. [Google Scholar] [CrossRef] [PubMed]
  162. Ke, M.-Y.; Fang, Y.; Cai, H.; Lu, J.-W.; Yang, L.; Wang, Y.; Wu, R.-Q.; Zhang, X.-F.; Lv, Y.; Dong, J. The m6 a reader YTHDF1 attenuates fulminant hepatitis via MFG-E8 translation in an m6 a dependent manner. Int. J. Biol. Sci. 2023, 19, 3987–4003. [Google Scholar] [CrossRef]
  163. Wang, S.; Zhang, W.; Wang, Z.; Liu, Z.; Yi, X.; Wu, J. Mettl3-m6A-YTHDF1 axis promotion of mitochondrial dysfunction in metabolic dysfunction–associated steatotic liver disease. Cell. Signal. 2024, 121, 111303. [Google Scholar] [CrossRef]
  164. Moqbel, S.A.A.; Zeng, R.; Ma, D.; Xu, L.; Lin, C.; He, Y.; Ma, C.; Xu, K.; Ran, J.; Jiang, L.; et al. The effect of mitochondrial fusion on chondrogenic differentiation of cartilage progenitor/stem cells via Notch2 signal pathway. Stem Cell Res. Ther. 2022, 13, 127. [Google Scholar] [CrossRef] [PubMed]
  165. Liu, Z.-Y.; Lin, L.-C.; Liu, Z.-Y.; Song, K.; Tu, B.; Sun, H.; Zhou, Y.; Mao, S.; Zhang, Y.; Li, R.; et al. N6-methyladenosine-mediated phase separation suppresses NOTCH1 expression and promotes mitochondrial fission in diabetic cardiac fibrosis. Cardiovasc. Diabetol. 2024, 23, 347. [Google Scholar] [CrossRef] [PubMed]
  166. Wang, J.; Yang, Y.; Sun, F.; Luo, Y.; Yang, Y.; Li, J.; Hu, W.; Tao, H.; Lu, C.; Yang, J.-J. ALKBH5 Attenuates mitochondrial fission and ameliorates liver fibrosis by reducing Drp1 methylation. Pharmacol. Res. 2023, 187, 106608. [Google Scholar] [CrossRef]
  167. Liu, L.; Wu, J.; Lu, C.; Ma, Y.; Wang, J.; Xu, J.; Yang, X.; Zhang, X.; Wang, H.; Xu, J.; et al. WTAP-mediated m6A modification of lncRNA Snhg1 improves myocardial ischemia-reperfusion injury via miR-361-5p/OPA1-dependent mitochondrial fusion. J. Transl. Med. 2024, 22, 499. [Google Scholar] [CrossRef]
  168. Lunt, S.Y.; Vander Heiden, M.G. Aerobic glycolysis: Meeting the metabolic requirements of cell proliferation. Annu. Rev. Cell Dev. Biol. 2011, 27, 441–464. [Google Scholar] [CrossRef]
  169. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  170. Warburg, O. On the origin of cancer cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef]
  171. Warburg, O. On respiratory impairment in cancer cells. Science 1956, 124, 269–270. [Google Scholar] [CrossRef] [PubMed]
  172. Sainero-Alcolado, L.; Liaño-Pons, J.; Ruiz-Pérez, M.V.; Arsenian-Henriksson, M. Targeting mitochondrial metabolism for precision medicine in cancer. Cell Death Differ. 2022, 29, 1304–1317. [Google Scholar] [CrossRef] [PubMed]
  173. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef]
  174. Yu, L.; Chen, X.; Sun, X.; Wang, L.; Chen, S. The glycolytic switch in tumors: How many players are involved? J. Cancer 2017, 8, 3430–3440. [Google Scholar] [CrossRef]
  175. Bueno, M.J.; Ruiz-Sepulveda, J.L.; Quintela-Fandino, M. Mitochondrial inhibition: A treatment strategy in cancer? Curr. Oncol. Rep. 2021, 23, 49. [Google Scholar] [CrossRef]
  176. Miller, J.J.; Gonzalez Castro, L.N.; McBrayer, S.; Weller, M.; Cloughesy, T.; Portnow, J.; Andronesi, O.; Barnholtz-Sloan, J.S.; Baumert, B.G.; Berger, M.S.; et al. Isocitrate dehydrogenase (IDH) mutant gliomas: A society for neuro-oncology (SNO) consensus review on diagnosis, management, and future directions. Neuro-Oncol. 2023, 25, 4–25. [Google Scholar] [CrossRef] [PubMed]
  177. Pirozzi, C.J.; Yan, H. The implications of IDH mutations for cancer development and therapy. Nat. Rev. Clin. Oncol. 2021, 18, 645–661. [Google Scholar] [CrossRef]
  178. Li, J.; Yu, T.; Zeng, P.; Tian, J.; Liu, P.; Qiao, S.; Wen, S.; Hu, Y.; Liu, Q.; Lu, W.; et al. Wild-type IDH2 Is a therapeutic target for triple-negative breast cancer. Nat. Commun. 2024, 15, 3445. [Google Scholar] [CrossRef]
  179. Li, J.; He, Y.; Tan, Z.; Lu, J.; Li, L.; Song, X.; Shi, F.; Xie, L.; You, S.; Luo, X.; et al. Wild-Type IDH2 promotes the warburg effect and tumor growth through HIF1α in lung cancer. Theranostics 2018, 8, 4050–4061. [Google Scholar] [CrossRef]
  180. Chen, X.; Xu, W.; Wang, C.; Guan, S.; Sun, Y.; Wang, X.; An, D.; Wen, Z.; Chen, P.; Cheng, Y. The clinical significance of isocitrate dehydrogenase 2 in esophageal squamous cell carcinoma. Science 2017, 7, 700. [Google Scholar]
  181. Fan, S.; Yan, X.; Hu, X.; Liu, X.; Zhao, S.; Zhang, Y.; Zhou, X.; Shen, X.; Qi, Q.; Chen, Y. Shikonin Blocks CAF-induced TNBC metastasis by suppressing mitochondrial biogenesis through GSK-3β/NEDD4-1 mediated phosphorylation-dependent degradation of PGC-1α. J. Exp. Clin. Cancer Res. 2024, 43, 180. [Google Scholar] [CrossRef] [PubMed]
  182. Raggi, C.; Taddei, M.L.; Sacco, E.; Navari, N.; Correnti, M.; Piombanti, B.; Pastore, M.; Campani, C.; Pranzini, E.; Iorio, J.; et al. Mitochondrial oxidative metabolism contributes to a cancer stem cell phenotype in cholangiocarcinoma. J. Hepatol. 2021, 74, 1373–1385. [Google Scholar] [CrossRef] [PubMed]
  183. Liu, Y.; Jin, M.; Wang, Y.; Zhu, J.; Tan, R.; Zhao, J.; Ji, X.; Jin, C.; Jia, Y.; Ren, T.; et al. MCU-induced mitochondrial calcium uptake promotes mitochondrial biogenesis and colorectal cancer growth. Sig. Transduct. Target. Ther. 2020, 5, 59. [Google Scholar] [CrossRef]
  184. Ren, T.; Zhang, H.; Wang, J.; Zhu, J.; Jin, M.; Wu, Y.; Guo, X.; Ji, L.; Huang, Q.; Zhang, H.; et al. MCU-dependent mitochondrial Ca2+ Inhibits NAD+/SIRT3/SOD2 pathway to promote ROS production and metastasis of HCC Cells. Oncogene 2017, 36, 5897–5909. [Google Scholar] [CrossRef]
  185. Vultur, A.; Gibhardt, C.S.; Stanisz, H.; Bogeski, I. The Role of the Mitochondrial Calcium Uniporter (MCU) Complex in Cancer. Pflug. Arch.—Eur. J. Physiol. 2018, 470, 1149–1163. [Google Scholar] [CrossRef] [PubMed]
  186. Jin, M.; Wang, J.; Ji, X.; Cao, H.; Zhu, J.; Chen, Y.; Yang, J.; Zhao, Z.; Ren, T.; Xing, J. MCUR1 facilitates epithelial-mesenchymal transition and metastasis via the mitochondrial calcium dependent ROS/Nrf2/Notch pathway in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2019, 38, 136. [Google Scholar] [CrossRef]
  187. Marchi, S.; Lupini, L.; Patergnani, S.; Rimessi, A.; Missiroli, S.; Bonora, M.; Bononi, A.; Corrà, F.; Giorgi, C.; De Marchi, E.; et al. Downregulation of the mitochondrial calcium uniporter by cancer-related miR-25. Curr. Biol. 2013, 23, 58–63. [Google Scholar] [CrossRef]
  188. Gil-Hernández, A.; Arroyo-Campuzano, M.; Simoni-Nieves, A.; Zazueta, C.; Gomez-Quiroz, L.E.; Silva-Palacios, A. Relevance of membrane contact sites in cancer progression. Front. Cell Dev. Biol. 2021, 8, 622215. [Google Scholar] [CrossRef]
  189. Yang, H.; Wang, Z.; Hu, S.; Chen, L.; Li, W.; Yang, Z. miRNA-874-3p inhibits the migration, invasion and proliferation of breast cancer cells by targeting VDAC1. Aging 2023, 15, 705–717. [Google Scholar] [CrossRef]
  190. Li, Y.; Chen, H.; Yang, Q.; Wan, L.; Zhao, J.; Wu, Y.; Wang, J.; Yang, Y.; Niu, M.; Liu, H.; et al. Increased Drp1 Promotes autophagy and ESCC progression by mtDNA stress mediated cGAS-STING pathway. J. Exp. Clin. Cancer Res. 2022, 41, 76. [Google Scholar] [CrossRef]
  191. Liang, J.; Yang, Y.; Bai, L.; Li, F.; Li, E. DRP1 Upregulation promotes pancreatic cancer growth and metastasis through increased aerobic glycolysis. J. Gastro Hepatol. 2020, 35, 885–895. [Google Scholar] [CrossRef] [PubMed]
  192. Huang, T.; Chang, C.; Chien, C.; Huang, G.; Chen, Y.; Su, L.; Tsai, H.; Lin, Y.; Fang, F.; Chen, C. DRP1 contributes to head and neck cancer progression and induces glycolysis through modulated FOXM1/MMP12 Axis. Mol. Oncol. 2022, 16, 2585–2606. [Google Scholar] [CrossRef] [PubMed]
  193. Kuang, J.; Liu, H.; Feng, L.; Xue, Y.; Tang, H.; Xu, P. How mitochondrial dynamics imbalance affects the progression of breast cancer:A mini review. Med. Oncol. 2024, 41, 238. [Google Scholar] [CrossRef]
  194. Zhan, L.; Cao, H.; Wang, G.; Lyu, Y.; Sun, X.; An, J.; Wu, Z.; Huang, Q.; Liu, B.; Xing, J. Drp1-Mediated mitochondrial fission promotes cell proliferation through crosstalk of P53 and NF-κB pathways in hepatocellular carcinoma. Oncotarget 2016, 7, 65001–65011. [Google Scholar] [CrossRef]
  195. Huang, Q.; Zhan, L.; Cao, H.; Li, J.; Lyu, Y.; Guo, X.; Zhang, J.; Ji, L.; Ren, T.; An, J.; et al. Increased mitochondrial fission promotes autophagy and hepatocellular carcinoma cell survival through the ROS-modulated coordinated regulation of the NFKB and TP53 pathways. Autophagy 2016, 12, 999–1014. [Google Scholar] [CrossRef] [PubMed]
  196. Zhao, J.; Zhang, J.; Yu, M.; Xie, Y.; Huang, Y.; Wolff, D.W.; Abel, P.W.; Tu, Y. Mitochondrial dynamics regulates migration and invasion of breast cancer cells. Oncogene 2013, 32, 4814–4824. [Google Scholar] [CrossRef]
  197. Rehman, J.; Zhang, H.J.; Toth, P.T.; Zhang, Y.; Marsboom, G.; Hong, Z.; Salgia, R.; Husain, A.N.; Wietholt, C.; Archer, S.L. Inhibition of mitochondrial fission prevents cell cycle progression in lung cancer. FASEB J. 2012, 26, 2175–2186. [Google Scholar] [CrossRef]
  198. Li, S.; Han, S.; Zhang, Q.; Zhu, Y.; Zhang, H.; Wang, J.; Zhao, Y.; Zhao, J.; Su, L.; Li, L.; et al. FUNDC2 Promotes Liver tumorigenesis by inhibiting MFN1-mediated mitochondrial fusion. Nat. Commun. 2022, 13, 3486. [Google Scholar] [CrossRef]
  199. Yeon, S.Y.; Jo, Y.S.; Choi, E.J.; Kim, M.S.; Yoo, N.J.; Lee, S.H. Frameshift mutations in repeat sequences of ANK3, HACD4, TCP10L, TP53BP1, MFN1, LCMT2, RNMT, TRMT6, METTL8 and METTL16 genes in colon cancers. Pathol. Oncol. Res. 2018, 24, 617–622. [Google Scholar] [CrossRef]
  200. Xu, K.; Chen, G.; Li, X.; Wu, X.; Chang, Z.; Xu, J.; Zhu, Y.; Yin, P.; Liang, X.; Dong, L. MFN2 suppresses cancer progression through inhibition of mTORC2/Akt signaling. Sci. Rep. 2017, 7, 41718. [Google Scholar] [CrossRef]
  201. Pang, G.; Xie, Q.; Yao, J. Mitofusin 2 inhibits bladder cancer cell proliferation and invasion via the Wnt/Β-catenin pathway. Oncol. Lett. 2019, 18, 2434–2442. [Google Scholar] [CrossRef] [PubMed]
  202. Wang, Y.; Li, Y.; Jiang, X.; Gu, Y.; Zheng, H.; Wang, X.; Zhang, H.; Wu, J.; Cheng, Y. OPA1 supports mitochondrial dynamics and immune evasion to CD8+ T cell in lung adenocarcinoma. PeerJ 2022, 10, e14543. [Google Scholar] [CrossRef]
  203. Tak, H.; Kang, H.; Ji, E.; Hong, Y.; Kim, W.; Lee, E.K. Potential use of TIA-1, MFF, microRNA-200a-3p, and microRNA-27 as a novel marker for hepatocellular carcinoma. Biochem. Biophys. Res. Commun. 2018, 497, 1117–1122. [Google Scholar] [CrossRef]
  204. Soares, C.D.; Morais, T.M.D.L.; Carlos, R.; De Almeida, O.P.; Mariano, F.V.; Altemani, A.; De Carvalho, M.G.F.; Corrêa, M.B.; Dos Reis, R.R.D.; Amorim, L.S.; et al. Prognostic importance of mitochondrial markers in mucosal and cutaneous head and neck melanomas. Hum. Pathol. 2019, 85, 279–289. [Google Scholar] [CrossRef]
  205. Abo Elwafa, R.; Gamaleldin, M.; Ghallab, O. The clinical and prognostic significance of FIS1, SPI1, PDCD7 and Ang2 expression levels in acute myeloid leukemia. Cancer Genet. 2019, 233–234, 84–95. [Google Scholar] [CrossRef]
  206. Li, Q.; Chu, Y.; Li, S.; Yu, L.; Deng, H.; Liao, C.; Liao, X.; Yang, C.; Qi, M.; Cheng, J.; et al. The oncoprotein MUC1 facilitates breast cancer progression by promoting Pink1-dependent mitophagy via ATAD3A destabilization. Cell Death Dis. 2022, 13, 899. [Google Scholar] [CrossRef]
  207. Abdel-Moety, A.; Baddour, N.; Salem, P.; El-Tobgy, H.; El-Shendidi, A. SQSTM1 expression in hepatocellular carcinoma and relation to tumor recurrence after radiofrequency ablation. J. Clin. Exp. Hepatol. 2022, 12, 774–784. [Google Scholar] [CrossRef] [PubMed]
  208. Yamashita, K.; Miyata, H.; Makino, T.; Masuike, Y.; Furukawa, H.; Tanaka, K.; Miyazaki, Y.; Takahashi, T.; Kurokawa, Y.; Yamasaki, M.; et al. High expression of the mitophagy-related protein Pink1 is associated with a poor response to chemotherapy and a poor prognosis for patients treated with neoadjuvant chemotherapy for esophageal squamous cell carcinoma. Ann. Surg. Oncol. 2017, 24, 4025–4032. [Google Scholar] [CrossRef] [PubMed]
  209. Lu, X.; Liu, Q.-X.; Zhang, J.; Zhou, D.; Yang, G.-X.; Li, M.-Y.; Qiu, Y.; Chen, Q.; Zheng, H.; Dai, J.-G. PINK1 Overexpression promotes cell migration and proliferation via regulation of autophagy and predicts a poor prognosis in lung cancer cases. CMAR 2020, 12, 7703–7714. [Google Scholar] [CrossRef]
  210. Claudia, C.V.D.S.; Rosimeri, K.S.B.; Nayanne, L.C.P.; Amanda, P.A.; Marcia, O.; Lúcia, D.N.; Vanessa, S.S. Parkin protein expression and its impact on survival of patients with advanced colorectal cancer. Cancer Biol. Med. 2018, 15, 61. [Google Scholar] [CrossRef]
  211. Toma, M.I.; Wuttig, D.; Kaiser, S.; Herr, A.; Weber, T.; Zastrow, S.; Koch, R.; Meinhardt, M.; Baretton, G.B.; Wirth, M.P.; et al. PARK2 and PACRG are commonly downregulated in clear-cell renal cell carcinoma and are associated with aggressive disease and poor clinical outcome. Genes. Chromosomes Cancer 2013, 52, 265–273. [Google Scholar] [CrossRef] [PubMed]
  212. Gao, G.; Kasperbauer, J.L.; Tombers, N.M.; Wang, V.; Mayer, K.; Smith, D.I. A selected group of large common fragile site genes have decreased expression in oropharyngeal squamous cell carcinomas. Genes. Chromosomes Cancer 2014, 53, 392–401. [Google Scholar] [CrossRef] [PubMed]
  213. Li, C.; Zhang, Y.; Cheng, X.; Yuan, H.; Zhu, S.; Liu, J.; Wen, Q.; Xie, Y.; Liu, J.; Kroemer, G.; et al. PINK1 and PARK2 suppress pancreatic tumorigenesis through control of mitochondrial iron-mediated immunometabolism. Dev. Cell 2018, 46, 441–455.e8. [Google Scholar] [CrossRef]
  214. Macher-Goeppinger, S.; Keith, M.; Hatiboglu, G.; Hohenfellner, M.; Schirmacher, P.; Roth, W.; Tagscherer, K.E. Expression and functional characterization of the BNIP3 protein in renal cell carcinomas. Transl. Oncol. 2017, 10, 869–875. [Google Scholar] [CrossRef]
  215. Wu, L.; Zhang, D.; Zhou, L.; Pei, Y.; Zhuang, Y.; Cui, W.; Chen, J. FUN14 domain-containing 1 promotes breast cancer proliferation and migration by activating calcium-NFATC1-BMI1 axis. EBioMedicine 2019, 41, 384–394. [Google Scholar] [CrossRef]
  216. Hou, H.; Er, P.; Cheng, J.; Chen, X.; Ding, X.; Wang, Y.; Chen, X.; Yuan, Z.; Pang, Q.; Wang, P.; et al. High expression of FUNDC1 predicts poor prognostic outcomes and is a promising target to improve chemoradiotherapy effects in patients with cervical cancer. Cancer Med. 2017, 6, 1871–1881. [Google Scholar] [CrossRef] [PubMed]
  217. Green, N.H.; Galvan, D.L.; Badal, S.S.; Chang, B.H.; LeBleu, V.S.; Long, J.; Jonasch, E.; Danesh, F.R. MTHFD2 Links RNA Methylation to metabolic reprogramming in renal cell carcinoma. Oncogene 2019, 38, 6211–6225. [Google Scholar] [CrossRef]
  218. Zhuang, C.; Zhuang, C.; Luo, X.; Huang, X.; Yao, L.; Li, J.; Li, Y.; Xiong, T.; Ye, J.; Zhang, F.; et al. N6-methyladenosine demethylase FTO suppresses clear cell renal cell carcinoma through a novel FTO - PGC -1α signalling axis. J. Cell. Mol. Medi. 2019, 23, 2163–2173. [Google Scholar] [CrossRef]
  219. Xu, W.; Lai, Y.; Pan, Y.; Tan, M.; Ma, Y.; Sheng, H.; Wang, J. m6A RNA methylation-mediated NDUFA4 promotes cell proliferation and metabolism in gastric cancer. Cell Death Dis. 2022, 13, 715. [Google Scholar] [CrossRef]
  220. Li, Z.; Peng, Y.; Li, J.; Chen, Z.; Chen, F.; Tu, J.; Lin, S.; Wang, H. N6-methyladenosine regulates glycolysis of cancer cells through PDK4. Nat. Commun. 2020, 11, 2578. [Google Scholar] [CrossRef]
  221. Zhou, J.; Zhang, H.; Zhong, K.; Tao, L.; Lin, Y.; Xie, G.; Tan, Y.; Wu, Y.; Lu, Y.; Chen, Z.; et al. N6-methyladenosine facilitates mitochondrial fusion of colorectal cancer cells via induction of GSH synthesis and stabilization of OPA1 mRNA. Natl. Sci. Rev. 2024, 11, nwae039. [Google Scholar] [CrossRef] [PubMed]
  222. Zhou, Y.; Wang, Q.; Deng, H.; Xu, B.; Zhou, Y.; Liu, J.; Liu, Y.; Shi, Y.; Zheng, X.; Jiang, J. N6-methyladenosine demethylase FTO promotes growth and metastasis of gastric cancer via m6A modification of caveolin-1 and metabolic regulation of mitochondrial dynamics. Cell Death Dis. 2022, 13, 72. [Google Scholar] [CrossRef]
  223. Yan, Y.; Luo, A.; Liu, S.; Cai, M.; Liu, X.; Zhang, X.; Zhang, S.; Liu, Y.; Zeng, J.; Xu, X.; et al. METTL3-mediated LINC00475 alternative splicing promotes glioma progression by inducing mitochondrial fission. Research 2024, 7, 0324. [Google Scholar] [CrossRef] [PubMed]
  224. Niu, Y.; Lin, Z.; Wan, A.; Chen, H.; Liang, H.; Sun, L.; Wang, Y.; Li, X.; Xiong, X.; Wei, B.; et al. RNA N6-methyladenosine demethylase FTO promotes breast tumor progression through inhibiting BNIP3. Mol. Cancer 2019, 18, 46. [Google Scholar] [CrossRef]
  225. Huang, J.; Sun, W.; Wang, Z.; Lv, C.; Zhang, T.; Zhang, D.; Dong, W.; Shao, L.; He, L.; Ji, X.; et al. FTO suppresses glycolysis and growth of papillary thyroid cancer via decreasing stability of APOE mRNA in an N6-methyladenosine-dependent manner. J. Exp. Clin. Cancer Res. 2022, 41, 42. [Google Scholar] [CrossRef]
  226. Sun, Y.; Shen, W.; Hu, S.; Lyu, Q.; Wang, Q.; Wei, T.; Zhu, W.; Zhang, J. METTL3 promotes chemoresistance in small cell lung cancer by inducing mitophagy. J. Exp. Clin. Cancer Res. 2023, 42, 65. [Google Scholar] [CrossRef] [PubMed]
  227. Lu, S.; Han, L.; Hu, X.; Sun, T.; Xu, D.; Li, Y.; Chen, Q.; Yao, W.; He, M.; Wang, Z.; et al. N6-methyladenosine reader IMP2 stabilizes the ZFAS1/OLA1 axis and activates the warburg effect: Implication in colorectal cancer. J. Hematol. Oncol. 2021, 14, 188. [Google Scholar] [CrossRef]
  228. Xiao, Y.; Liu, X.; Xie, K.; Luo, J.; Zhang, Y.; Huang, X.; Luo, J.; Tan, S. Mitochondrial dysfunction induced by HIF-1α under hypoxia contributes to the development of gastric mucosal lesions. Clin. Amp. Transl. Med. 2024, 14, e1653. [Google Scholar] [CrossRef]
  229. Weng, H.; Huang, F.; Yu, Z.; Chen, Z.; Prince, E.; Kang, Y.; Zhou, K.; Li, W.; Hu, J.; Fu, C.; et al. The m6A reader IGF2BP2 regulates glutamine metabolism and represents a therapeutic target in acute myeloid leukemia. Cancer Cell 2022, 40, 1566–1582.e10. [Google Scholar] [CrossRef]
  230. Zhang, H.; Zhu, S.; Zhou, H.; Li, R.; Xia, X.; Xiong, H. Identification of MTHFD2 as a Prognostic biomarker and ferroptosis regulator in triple-negative breast cancer. Front. Oncol. 2023, 13, 1098357. [Google Scholar] [CrossRef]
  231. Dou, X.; Fu, Q.; Long, Q.; Liu, S.; Zou, Y.; Fu, D.; Xu, Q.; Jiang, Z.; Ren, X.; Zhang, G.; et al. PDK4-dependent hypercatabolism and lactate production of senescent cells promotes cancer malignancy. Nat. Metab. 2023, 5, 1887–1910. [Google Scholar] [CrossRef] [PubMed]
  232. Cho, E.-C.; Kuo, M.-L.; Cheng, J.; Cheng, Y.-C.; Hsieh, Y.-C.; Liu, Y.-R.; Hsieh, R.-H.; Yen, Y. RRM2B-mediated regulation of mitochondrial activity and inflammation under oxidative stress. Mediat. Inflamm. 2015, 2015, 287345. [Google Scholar] [CrossRef] [PubMed]
  233. Muñoz, S.S.; Garner, B.; Ooi, L. Understanding the role of ApoE fragments in Alzheimer’s disease. Neurochem. Res. 2019, 44, 1297–1305. [Google Scholar] [CrossRef]
  234. Xia, Y.; Wang, Y.; Chen, K.; Zhang, M.; Jiang, Q.; Xu, T. Quercetin attenuated necroptosis and apoptosis caused by LPS-Induced mitochondrial function dysfunction through the METTL3-mediated PTEN m6A methylation/PI3K/AKT signaling in broiler livers. Phytomedicine 2025, 139, 156551. [Google Scholar] [CrossRef] [PubMed]
  235. Jiao, Y.; Williams, A.; Wei, N. Quercetin ameliorated insulin resistance via regulating METTL3-mediated N6-methyladenosine modification of PRKD2 mRNA in skeletal muscle and C2C12 myocyte cell line. Nutr. Metab. Cardiovasc. Dis. 2022, 32, 2655–2668. [Google Scholar] [CrossRef]
Figure 1. M6A regulators (created with BioRender.com). M6A methylation is a dynamic process, and there are three types of regulators involved in this process: writers, erasers, and readers. Writers mainly include the methyltransferase complex with METTL3, METTL14, and WTAP as the core, as well as METTL16. FTO and ALKBH5 are two demethylases, which are erasers. Readers play an important role in determining the fate of target RNAs, mainly including YT521-B homology (YTH) domain-containing protein families, IGF2 mRNA-binding protein (IGF2BP) families, heterogeneous nuclear ribonucleoprotein (HNRNP) protein families, and eukaryotic initiation factor (eIF) 3.
Figure 1. M6A regulators (created with BioRender.com). M6A methylation is a dynamic process, and there are three types of regulators involved in this process: writers, erasers, and readers. Writers mainly include the methyltransferase complex with METTL3, METTL14, and WTAP as the core, as well as METTL16. FTO and ALKBH5 are two demethylases, which are erasers. Readers play an important role in determining the fate of target RNAs, mainly including YT521-B homology (YTH) domain-containing protein families, IGF2 mRNA-binding protein (IGF2BP) families, heterogeneous nuclear ribonucleoprotein (HNRNP) protein families, and eukaryotic initiation factor (eIF) 3.
Ijms 26 03624 g001
Figure 2. Comparison between normal mitochondrial function and mitochondrial dysfunction (created with BioRender.com). Mitochondria are multifunctional organelles that play crucial roles in maintaining cellular ROS and Ca2+ homeostasis, ATP production, and regulating cell death; when mitochondrial dysfunction occurs, there will be decrease in ATP production, damage to mitochondrial DNA, oxidative stress, imbalance of calcium homeostasis, abnormal mitochondrial dynamics, and occurrence of cell death.
Figure 2. Comparison between normal mitochondrial function and mitochondrial dysfunction (created with BioRender.com). Mitochondria are multifunctional organelles that play crucial roles in maintaining cellular ROS and Ca2+ homeostasis, ATP production, and regulating cell death; when mitochondrial dysfunction occurs, there will be decrease in ATP production, damage to mitochondrial DNA, oxidative stress, imbalance of calcium homeostasis, abnormal mitochondrial dynamics, and occurrence of cell death.
Ijms 26 03624 g002
Figure 3. Role of m6A in influencing mitochondrial function in cancer (created with BioRender.com). Dysregulated expression of m6A regulators (upregulation or downregulation) in tumors modulates cancer progression by affecting mitochondrial functions, including mitochondrial energy metabolism, biogenesis, redox homeostasis, dynamics, mitophagy, and apoptosis regulation. In this figure, the red upward arrow behind the m6A regulators indicates upregulation, and the blue downward arrow indicates downregulation.
Figure 3. Role of m6A in influencing mitochondrial function in cancer (created with BioRender.com). Dysregulated expression of m6A regulators (upregulation or downregulation) in tumors modulates cancer progression by affecting mitochondrial functions, including mitochondrial energy metabolism, biogenesis, redox homeostasis, dynamics, mitophagy, and apoptosis regulation. In this figure, the red upward arrow behind the m6A regulators indicates upregulation, and the blue downward arrow indicates downregulation.
Ijms 26 03624 g003
Table 3. Some functional units of mitochondria become abnormal in cancer.
Table 3. Some functional units of mitochondria become abnormal in cancer.
Mitochondrial FunctionAbnormalities of Functional UnitsCancer TypeReference
Mitochondrial energy metabolismIsocitrate dehydrogenase (IDH) mutantGliomas, acute myeloid leukemia, cholangiocarcinoma, chondrosarcoma[176,177]
Upregulation of wild-type IDH2 Triple-negative breast cancer, lung cancer, esophageal squamous cell carcinoma[178,179,180]
Upregulation of PGC-1αTriple-negative breast cancer, cholangiocarcinoma[181,182]
Mitochondrial calcium homeostasisUpregulation of MCU and downregulation of MICU1Colorectal cancer, hepatocellular carcinoma, breast cancer[183,184,185]
Upregulation of MCUR1Hepatocellular carcinoma[186]
Downregulation of MCUProstate cancer, colon cancer[187]
Upregulation of VDACBreast cancer, head and neck cancer, lung adenocarcinoma[188]
Upregulation of VDAC1Breast cancer[189]
Mitochondrial
dynamics
Upregulation of DRP1Esophageal squamous cell carcinoma, pancreatic cancer, head and neck cancer, breast cancer, hepatocellular carcinoma[190,191,192,193,194]
Upregulation of DRP1 and downregulation of MFN1Hepatocellular carcinoma, breast cancer[195,196]
Upregulation of DRP1 and downregulation of MFN2Lung cancer[197]
The mitochondrial protein FUNDC2 inhibits MFN1Hepatocellular carcinoma[198]
MFN1 frameshift mutationsColorectal cancer[199]
Downregulation of MFN2Breast cancer, lung cancer, bladder cancer[200,201]
Upregulation of OPA1 and MFN1Lung adenocarcinoma[202]
Upregulation of fission factor Hepatocellular carcinoma[203]
Upregulation of FIS1Oral melanoma, acute myeloid leukemia[204,205]
MitophagyHigh expression of mucin 1 (MUC1) protects PINK1 from cleavage, thereby increasing mitochondrial autophagyBreast cancer[206]
Upregulation of sequestosome 1 (SQSTM1/p62) Hepatocellular carcinoma[207]
Upregulation of PINK1Esophageal squamous cell carcinoma, non-small cell lung cancer[208,209]
Downregulation of PankinColorectal cancer, clear-cell renal cell carcinoma, oropharyngeal squamous cell carcinoma, pancreatic ductal adenocarcinoma[210,211,212,213]
Upregulation of BNIP3Renal cell carcinoma[214]
Upregulation of FUNDC1Breast cancer, cervical cancer[215,216]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yan, W.; Saqirile; Li, K.; Li, K.; Wang, C. The Role of N6-Methyladenosine in Mitochondrial Dysfunction and Pathology. Int. J. Mol. Sci. 2025, 26, 3624. https://doi.org/10.3390/ijms26083624

AMA Style

Yan W, Saqirile, Li K, Li K, Wang C. The Role of N6-Methyladenosine in Mitochondrial Dysfunction and Pathology. International Journal of Molecular Sciences. 2025; 26(8):3624. https://doi.org/10.3390/ijms26083624

Chicago/Turabian Style

Yan, Wenxin, Saqirile, Ke Li, Kexin Li, and Changshan Wang. 2025. "The Role of N6-Methyladenosine in Mitochondrial Dysfunction and Pathology" International Journal of Molecular Sciences 26, no. 8: 3624. https://doi.org/10.3390/ijms26083624

APA Style

Yan, W., Saqirile, Li, K., Li, K., & Wang, C. (2025). The Role of N6-Methyladenosine in Mitochondrial Dysfunction and Pathology. International Journal of Molecular Sciences, 26(8), 3624. https://doi.org/10.3390/ijms26083624

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

Article Metrics

Back to TopTop