The dynamic nature of mitochondria, which can fuse, divide and move throughout the cell, allows these critical organelles to adapt their function in response to cellular demands, and is also important for regulating mitochondrial DNA (mtDNA). While it is established that impairments in mitochondrial fusion and fission impact the mitochondrial genome and can lead to mtDNA depletion, abnormal nucleoid organization or accumulation of deletions, it is not entirely clear how or why remodeling mitochondrial network morphology affects mtDNA. Here, we focus on recent advances in our understanding of how mitochondrial dynamics contribute to the regulation of mtDNA and discuss links to human disease.

Mitochondria are dynamic organelles best known for their role in producing energy via oxidative phosphorylation, but which are also important mediators of other cellular functions and processes, such as metabolism, innate immunity and cell death (Nunnari and Suomalainen, 2012). These organelles comprise an outer mitochondrial membrane (OMM) and an inner mitochondrial membrane (IMM), which define two distinct compartments, the intermembrane space (IMS) and the matrix (Fig. 1, diagram 1). Rather than bean-shaped structures depicted in textbooks, human mitochondria form a constantly changing network as they fuse (Fig. 1, diagrams 2 and 3), divide (Fig. 1, diagrams 4 and 5), and move throughout the cell (Sabouny and Shutt, 2020). Additionally, mitochondria can be degraded by mitochondrial autophagy (mitophagy) (Fig. 1, diagram 6) (Onishi et al., 2021), form vesicles (Sugiura et al., 2014), and interact both transiently and stably with a variety of other organelles (Lackner, 2019). The term ‘mitochondrial dynamics’ is often used to encompass many of these processes.

Fig. 1.

Schematic representation of dynamic mitochondrial processes. Top, the reciprocal processes of mitochondrial fission and fusion. (1) Diagram showing the outer mitochondrial membrane (OMM), inner mitochondrial membrane (IMM), intermembrane space (IMS), matrix, and mitochondrial DNA (mtDNA). (2) Mitochondrial fusion begins with tethering and fusion of the OMM, which is mediated by the OMM GTPases mitofusin 1 and 2 (MFN1 and MFN2) that can interact in a homotypic or heterotypic manner. The cytosolic protein MSTO1 mediates fusion through an undefined mechanism. (3) IMM fusion is mediated by the GTPase activity of optic atrophy protein 1 (OPA1), which is present in long and short isoforms (L-OPA1 and S-OPA1, respectively) due to proteolytic processing of multiple splice isoforms. L-OPA1 interacts with the inner membrane lipid cardiolipin (CL) to mediate inner membrane fusion. Notably, the F-box protein FBXL4, proposed to localize to the IMS, also regulates mitochondrial fusion by unknown means. (4) Mitochondrial fission begins with endoplasmic reticulum (ER)-mediated constriction of mitochondria, which is mediated by actin and non-muscle myosins (NMII). (5) The main fission GTPase, dynamin-related protein 1 (DRP1), is recruited to fission sites via the OMM adaptor proteins MFF, MID49, MID51 and FIS1. GTP hydrolysis and conformational change induces final OMM scission and gives rise to mitochondrial fragments. Bottom, following fission events near the ends of mitochondrial tubules, smaller mitochondrial fragments enriched in damaged components can undergo mitophagy. This process allows the eliminate of dysfunctional mitochondria from the cell, which may be due to damaged or mutant mtDNA (depicted in red). (6) Dysfunctional mitochondria accumulate the kinase PINK1 on the OMM, which recruits the E3 ubiquitin ligase Parkin leading to the ubiquitylation (Ub) of OMM proteins, including MFN2, triggering autophagosome formation.

Fig. 1.

Schematic representation of dynamic mitochondrial processes. Top, the reciprocal processes of mitochondrial fission and fusion. (1) Diagram showing the outer mitochondrial membrane (OMM), inner mitochondrial membrane (IMM), intermembrane space (IMS), matrix, and mitochondrial DNA (mtDNA). (2) Mitochondrial fusion begins with tethering and fusion of the OMM, which is mediated by the OMM GTPases mitofusin 1 and 2 (MFN1 and MFN2) that can interact in a homotypic or heterotypic manner. The cytosolic protein MSTO1 mediates fusion through an undefined mechanism. (3) IMM fusion is mediated by the GTPase activity of optic atrophy protein 1 (OPA1), which is present in long and short isoforms (L-OPA1 and S-OPA1, respectively) due to proteolytic processing of multiple splice isoforms. L-OPA1 interacts with the inner membrane lipid cardiolipin (CL) to mediate inner membrane fusion. Notably, the F-box protein FBXL4, proposed to localize to the IMS, also regulates mitochondrial fusion by unknown means. (4) Mitochondrial fission begins with endoplasmic reticulum (ER)-mediated constriction of mitochondria, which is mediated by actin and non-muscle myosins (NMII). (5) The main fission GTPase, dynamin-related protein 1 (DRP1), is recruited to fission sites via the OMM adaptor proteins MFF, MID49, MID51 and FIS1. GTP hydrolysis and conformational change induces final OMM scission and gives rise to mitochondrial fragments. Bottom, following fission events near the ends of mitochondrial tubules, smaller mitochondrial fragments enriched in damaged components can undergo mitophagy. This process allows the eliminate of dysfunctional mitochondria from the cell, which may be due to damaged or mutant mtDNA (depicted in red). (6) Dysfunctional mitochondria accumulate the kinase PINK1 on the OMM, which recruits the E3 ubiquitin ligase Parkin leading to the ubiquitylation (Ub) of OMM proteins, including MFN2, triggering autophagosome formation.

A critical aspect of the dynamic mitochondrial network is how it impacts the mitochondrial genome (mtDNA), which encodes proteins that are essential for oxidative phosphorylation (Chapman et al., 2020). Despite being sequenced over 40 years ago (Anderson et al., 1981), there is still much to learn about mtDNA and how it is regulated. For example, even though each cell typically contains ∼100–1000 copies of the mtDNA, we still do not completely understand how mtDNA copy number is determined and maintained, nor how mtDNA is distributed throughout the everchanging mitochondrial network. In this Review, we examine what is known about how mitochondrial fission and fusion impact mtDNA in these respects, as well as cover the implications for mtDNA mutations and human disease. As previous reviews have extensively covered mitochondrial dynamics (Chan, 2019) and mtDNA replication (Gustafsson et al., 2016), here we will only briefly review these key aspects to set the stage for how these topics are related.

The morphology of the mitochondrial network, which can vary from many small individual puncta to large, reticulated networks, is determined by the balance between fission and fusion events (Fig. 1). Changes in network morphology have important implications for mitochondrial function beyond their impact on the mtDNA. For example, the physical structure of mitochondria can influence biophysical properties (e.g. surface:volume ratio), affecting mitochondrial physiology or interactions with other organelles (Glancy et al., 2020). Meanwhile, fusion and fission events, which are often coordinated (Abrisch et al., 2020; Twig et al., 2008), allow for movement of mitochondria, content mixing and isolation of damaged mitochondria via mitophagy (Chan, 2012). It is also worth noting that defects in fusion and fission perturb many aspects of mitochondrial function, such as oxidative phosphorylation (Chen and Chan, 2010) and apoptosis (Karbowski and Youle, 2003).

Mitochondrial fission is orchestrated by a sequence of events that is initiated at sites where the endoplasmic reticulum (ER) wraps around mitochondrial tubules (Fig. 1, diagram 4) (Friedman et al., 2011). The actomyosin cytoskeleton at these contact sites provides the force to constrict mitochondrial tubules, initiating fission (Moore and Holzbaur, 2018). Next, the dynamin-related GTPase DRP1 (also known as DNM1L) can be recruited from the cytosol by several outer mitochondrial membrane (OMM) adaptor proteins, including MFF, MID49 (also known as MIEF2), MID51 (also known as MIEF1) and FIS1 (Kraus et al., 2021; Loson et al., 2013) (Fig. 1, diagram 5). Once on the OMM, DRP1 assembles into oligomeric rings around mitochondria that mediate scission through GTP hydrolysis and conformational changes.

Mitochondrial fusion begins with tethering and fusion of the OMM, processes mediated by the mitofusin GTPases MFN1 and MFN2 (Fig. 1, diagram 2) (Chen et al., 2003). Next, inner mitochondrial membrane (IMM) fusion is mediated by the GTPase activity of optic atrophy 1 (OPA1) (Fig. 1, diagram 3). Notably, OPA1 interacts with mitochondrial-specific lipid cardiolipin in the IMM to drive inner membrane fusion (Ban et al., 2017, 2018). While the GTPases MFN1, MFN2 and OPA1 constitute the core fusion machinery, recent studies have highlighted additional cytosolic and mitochondrial proteins regulating fusion. Notable among these factors are the cytosolic protein MSTO1 (Donkervoort et al., 2019) and the mitochondrial protein FBXL4 (Sabouny et al., 2019), mutations in which lead to depletion of the mtDNA genome.

Mitophagy, the clearance of damaged, dysfunctional or unwanted mitochondria by a specific form of autophagy, is tightly coordinated with the remodeling of mitochondrial networks (Fig. 1, diagram 6). For example, inhibiting fission results in hyperfused mitochondrial networks that accumulate damage (Parone et al., 2008) and are resistant to mitophagy (Gomes et al., 2011; Rambold et al., 2011). Thus, fission is believed to be required to form the initial mitochondrial fragments that are small enough to be removed by mitophagy. Meanwhile, the mitochondrial fusion protein MFN2 can also play a role in signaling mitophagy (Chen and Dorn, 2013). The molecular mechanisms regulating mitophagy include the well-characterized ubiquitin-dependent pathway mediated by the mitochondrial kinase PINK1 and the cytosolic E3 ubiquitin ligase Parkin (PRKN) (Pickles et al., 2018). While additional mitophagy pathways exist (Onishi et al., 2021), there is still much work required to understand exactly how these pathways work together or independently to maintain mitochondrial quality control and abundance under different contexts.

The importance of these dynamic processes is highlighted by the growing list of diseases caused by their impairment (Box 1), while the impact of fusion and fission on mtDNA are further highlighted by loss-of-function studies (Table 1). Although the link between maintenance of the mitochondrial genome and mitochondrial dynamics is an intriguing facet of fusion and fission dynamics, there remain many unanswered questions. However, before we examine recent advances in our understanding on how mitochondrial dynamics influence mtDNA, we will first review some basics of the mitochondrial genome.

Box 1. Mitochondrial dynamics and disease

The importance of mitochondrial dynamics is evidenced by the fact that pathogenic variants leading to a variety of human disease phenotypes are found in genes encoding several fusion, fission and mitophagy proteins, as well as the fact that mitochondrial dynamics are impaired in many disease models (Chan, 2019). The first indication that mitochondrial fusion is relevant to human disease was the recognition that pathogenic OPA1 variants cause dominant optic atrophy, a neuro-ophthalmic condition characterized by a bilateral degeneration of the optic nerves (Alexander et al., 2000). Shortly thereafter, pathogenic variants in MFN2 were recognized to cause Charcot–Marie–Tooth disease, a progressive motor and sensory neuropathy of the peripheral nervous system (Zuchner et al., 2004). Subsequently, the first report of a pathogenic variant in DRP1 found in a patient with neuronal issues and neonatal lethality also linked fission defects to human disease (Waterham et al., 2007). More recently, additional pathogenic variants have been found in a growing list of proteins involved in both fusion [e.g. MSTO1 (Gal et al., 2017; Nasca et al., 2017) and FBXL4 (Bonnen et al., 2013; Gai et al., 2013)] and fission [e.g. MFF (Koch et al., 2016), MID49 (Bartsakoulia et al., 2018) and NMIIC (Almutawa et al., 2019)]. Similarly, impaired mitophagy is linked to human disease, with the best example being the recognition that pathogenic variants of PINK1 and Parkin cause Parkinson's disease (Onishi et al., 2021). Importantly, as approaches to restore these processes are increasingly recognized to be beneficial in a growing list of human disease models (Whitley et al., 2019), it is crucial to understand how mitochondrial dynamics are regulated, and how their dysfunction leads to disease. Critically, disturbances to mtDNA are a common feature in several diseases arising from impairments in mitochondrial fusion, fission and mitophagy, including accumulation of mtDNA mutations, as well as alterations in mtDNA abundance and nucleoid distribution (Table 1).

Table 1.

Summary of reports linking impairment of mitochondrial fusion and fission proteins to mtDNA abnormalities and patient phenotypes.

Summary of reports linking impairment of mitochondrial fusion and fission proteins to mtDNA abnormalities and patient phenotypes.
Summary of reports linking impairment of mitochondrial fusion and fission proteins to mtDNA abnormalities and patient phenotypes.

The circular 16,569 bp human mitochondrial genome is replicated independently of the nuclear genome, and encodes 37 genes, including 13 proteins that are essential for oxidative phosphorylation. Replication, maintenance and expression of the multicopy mtDNA depends on nuclear-encoded proteins that are imported into the mitochondrial matrix. The core mtDNA replication machinery comprises the mitochondrial polymerase γ (POLG), the mitochondrial DNA helicase (Twinkle; TWNK) and the mitochondrial single-stranded DNA-binding protein (mtSSB; also known as SSBP1) (Gustafsson et al., 2016), which work in concert with several additional proteins. Importantly, components of the mitochondrial replication machinery need to be balanced stoichiometrically, as imbalance can lead to mtDNA depletion and deletions (Phillips et al., 2017). Replication of mitochondrial genomes also requires a balanced pool of intra-mitochondrial deoxyribonucleotide triphosphates (dNTPs) (Gorman et al., 2016).

The mtDNA copy number varies in different cell types, probably to help mitochondria meet differing energetic demands. However, the processes that regulate copy number are not completely understood. Nevertheless, the importance of maintaining sufficient copies of the mtDNA genome is evidenced by a class of mitochondrial diseases termed mitochondrial DNA depletion syndromes (MTDPS), which are characterized by a significant reduction of mtDNA levels in affected tissues (Viscomi and Zeviani, 2017). Notably, MTDPS arise from mutations in nuclear genes encoding factors important for mtDNA replication, maintenance of mtDNA nucleotide pools or mitochondrial dynamics (Table 1). Although it is not surprising that defects in mtDNA replication or imbalances in nucleotides would lead to mtDNA loss, exactly how fusion–fission dynamics affect mtDNA levels is not fully understood.

Another way through which mtDNA impairment can lead to dysfunction is through the acquisition of point mutations, as well as both small- and large-scale deletions or duplications. Critically, mtDNA mutations can and do accumulate to a point where they negatively impact mitochondrial function, and lead to diseases such as mitochondrial encephalomyopathy, lactic acidosis and stroke-like episodes (MELAS), myoclonus epilepsy with ragged-red fibers (MERFF) and Leigh syndrome (Stewart and Chinnery, 2015). A mixture of wild-type and mutant mtDNA is referred to as heteroplasmy, while the term homoplasmy refers to a situation where all mtDNA copies in the cell are identical (either wild-type or mutant). As part of the mitochondrial genome, deleterious mtDNA variants are inherited maternally. Additionally, mtDNA mutations arise de novo in somatic cells, and their accumulation has been proposed to contribute to both neurodegenerative conditions (Keogh and Chinnery, 2015) and cancers (Yuan et al., 2020). The level of heteroplasmy required to meet a pathogenic threshold varies depending on the specific mutation, the affected tissue and the genetic background of an individual. Generally, increased levels of mutant mtDNA correspond to worse disease phenotypes. Whereas heteroplasmy levels can vary significantly between cells in the same individual, the factors and mechanisms leading to selective expansion of mutant mtDNA molecules are not entirely clear, although, as discussed later, mitochondrial dynamics have been implicated. Nevertheless, as heteroplasmy levels can change, approaches such as using targeted nucleases (Zekonyte et al., 2020) or editing enzymes (Mok et al., 2020) to reduce the abundance of mtDNA mutations offer promise for mtDNA disease patients, although they are still not yet applicable clinically (Al Khatib and Shutt, 2019).

The mtDNA is organized into nucleoprotein structures known as nucleoids, which are localized in the mitochondrial matrix (Gustafsson et al., 2016). Nucleoids are tethered to the IMM, which facilitates their movement and distribution within the mitochondrial network (Nicholls and Gustafsson, 2018). Although early work estimated that nucleoids contained approximately six to ten mtDNA copies (Iborra et al., 2004), more recent high-resolution microscopy studies show that nucleoid are ∼100 nm in diameter and estimated to contain 1.4 copies of mtDNA on average, reflecting the ongoing replication of mitochondrial genomes (Kukat et al., 2011). The fact that individual nucleoids are comprised of a single mtDNA genome supports earlier work showing that nucleoids do not exchange genomes (Gilkerson et al., 2008). However, despite this fact, cells with a mixture of distinct mtDNA mutations can have normal mitochondrial function owing to fusion-dependent content mixing and trans-complementation (Schon and Gilkerson, 2010). Although nucleoids are typically distributed evenly throughout the mitochondrial network, this distribution can be disturbed by several perturbations, including impaired mitochondrial dynamics (Box 2) (Fig. 2).

Box 2. mtDNA nucleoids

A number of different approaches have been developed to visualize mtDNA nucleoids (Prole et al., 2020), revealing a typical organization whereby uniformly sized nucleoids are distributed evenly throughout mitochondrial networks, regardless of whether they are elongated or fragmented (Chen et al., 2007). However, altered nucleoid distribution with enlarged nucleoids is reported in response to a variety of mitochondrial stresses. Notably, super-resolution microscopy has confirmed that in some cases these enlarged nucleoids, sometimes called ‘mito-bulbs’, are clusters of individual nucleoids (Ban-Ishihara et al., 2013; Nasonovs et al., 2021). Here, we will use the generic term enlarged nucleoid, as most studies to date only report such enlarged nucleoids under normal microscopy where it has not been confirmed that these are indeed clusters. Various mitochondrial stresses linked to enlarged nucleoids include those induced by drugs that intercalate into mtDNA (Alán et al., 2016; Ashley and Poulton, 2009), inhibit oxidative phosphorylation (Tauber et al., 2013) and uncouple mitochondrial membrane potential (Tauber et al., 2013), as well as upon genetic modifications that target several mitochondrial proteins (Dalla Rosa et al., 2014). Notably, these targets include the mtDNA-binding protein TFAM (transcription factor A, mitochondrial) (Kasashima et al., 2011), which induces a U-turn in the DNA structure and mediates compaction of mtDNA into nucleoids (Kukat and Larsson, 2013). As an aside, TFAM can spontaneously phase separate in vitro, which contributes to the recently described phase separation properties of mtDNA nucleoids (Feric et al., 2021). Highlighting their relevance to human heath, enlarged nucleoids are also reported in response to viral infection (West et al., 2015) and in several disease models, including mtDNA mutations (Newell et al., 2018), fibrotic lung disease (Ryu et al., 2017) and Hutchinson–Guilford progeria (Feric et al., 2021). Although it is not clear why enlarged nucleoids occur in the cases listed above, as discussed in the main text, mitochondrial fusion and fission are recognized to be important mediators of nucleoid size and distribution (Fig. 2).

Fig. 2.

Mitochondrial network remodeling and coordinated mtDNA nucleoid distribution following fusion and fission. (A) Mitochondrial morphology is determined by the balance between fusion and fission events (i.e. hyperfused, intermediate or fragmented). Mitochondria become hyperfused when there is more fusion than fission, or become fragmented when there is excess fission. Regardless of the mitochondrial network morphology, in healthy cells, mtDNA nucleoids are typically evenly distributed and copy number is maintained even upon normal shifting of mitochondrial morphology, which can be modulated in response to various physiological stimuli. (B) When mitochondrial fusion is impaired, this can lead to reduced mtDNA copy number, and can gives rise to small mitochondrial fragments that are either completely devoid of mtDNA nucleoids, or contain enlarged nucleoids that consist of clusters. (C) When mitochondrial fission is impaired, enlarged nucleoids can accumulate in hyperfused networks.

Fig. 2.

Mitochondrial network remodeling and coordinated mtDNA nucleoid distribution following fusion and fission. (A) Mitochondrial morphology is determined by the balance between fusion and fission events (i.e. hyperfused, intermediate or fragmented). Mitochondria become hyperfused when there is more fusion than fission, or become fragmented when there is excess fission. Regardless of the mitochondrial network morphology, in healthy cells, mtDNA nucleoids are typically evenly distributed and copy number is maintained even upon normal shifting of mitochondrial morphology, which can be modulated in response to various physiological stimuli. (B) When mitochondrial fusion is impaired, this can lead to reduced mtDNA copy number, and can gives rise to small mitochondrial fragments that are either completely devoid of mtDNA nucleoids, or contain enlarged nucleoids that consist of clusters. (C) When mitochondrial fission is impaired, enlarged nucleoids can accumulate in hyperfused networks.

Nucleoid distribution

A link between mitochondrial dynamics and nucleoid distribution was first recognized during early studies of mitochondrial dynamics, with the observation that nucleoids were often located near fission and fusion sites (Iborra et al., 2004; Margineantu et al., 2002). Subsequently, abnormally enlarged nucleoid structures were observed upon disruption of the mitochondrial fission machinery (e.g. through perturbing DRP1 or MFF) (Ashley and Poulton, 2009; Ban-Ishihara et al., 2013; Parone et al., 2008). In the context of diseases with impaired mitochondrial fission, larger nucleoids are reported in fibroblasts with reduced fission due to pathogenic variants in MYH14 and DNM1L (Almutawa et al., 2019; Ilamathi et al., 2021 preprint). Similarly, the abundance of nucleoids within extremely hyperfused regions of cells from patients with pathogenic variants in the fission protein DRP1 and non-muscle myosin IIC (NMIIC; heavy chain encoded by MYH14) are reduced (Almutawa et al., 2019; Ilamathi et al., 2021 preprint).

While enlarged nucleoids can be associated with mitochondrial dysfunction, they may not lead to dysfunction per se. For example, A549 human lung carcinoma cells contain enlarged nucleoid clusters that are functional with respect to transcription and replication competence (Nasonovs et al., 2021). Notably, the number of enlarged nucleoids in A549 cells increases in high-glucose conditions and decreases with TGFβ treatment, demonstrating that dynamic changes in nucleoid arrangement can occur in response to physiological cues (Nasonovs et al., 2021). Nevertheless, the question of why enlarged nucleoids form remains elusive. In this regard, recent work found that mtDNA replication is coordinated with fission events, as fission sites occur in close proximity to replicating nucleoids (Lewis et al., 2016). As such, fission has been proposed to be important for the segregation of nascent mtDNA molecules after replication (Fig. 3), which may explain the nucleoid clustering observed when fission is inhibited.

Fig. 3.

Role of mitochondrial fission in the appropriate distribution of newly replicated mtDNA nucleoids. (A) Following mtDNA replication, mitochondrial fission factors assemble and divide mitochondria, segregating nucleoids into separate daughter mitochondria. (B) In cells with a defective fission machinery, mitochondria do not undergo fission after mtDNA replication, impairing the distribution of mtDNA nucleoids and giving rise to enlarged clusters of nucleoids. However, the exact underlying mechanisms are not well understood.

Fig. 3.

Role of mitochondrial fission in the appropriate distribution of newly replicated mtDNA nucleoids. (A) Following mtDNA replication, mitochondrial fission factors assemble and divide mitochondria, segregating nucleoids into separate daughter mitochondria. (B) In cells with a defective fission machinery, mitochondria do not undergo fission after mtDNA replication, impairing the distribution of mtDNA nucleoids and giving rise to enlarged clusters of nucleoids. However, the exact underlying mechanisms are not well understood.

Impairments to mitochondrial fusion also impact the distribution of nucleoids throughout the mitochondrial network in two important ways. First, whereas mitochondria in normal cells typically each contain at least a single nucleoid, even when fragmented into smaller mitochondria, cells with impaired fusion can have mitochondrial fragments that do not contain nucleoids. Such an uneven distribution is reported in mouse embryonic fibroblast (MEF) cells lacking any one of the fusion proteins MFN1, MFN2 or OPA1 (Chen et al., 2007), as well as in fibroblasts from patients with pathogenic mutations in FBXL4 (Sabouny et al., 2019) and MSTO1 (Donkervoort et al., 2019). Second, upon impaired fusion, nucleoids appear larger, likely due to clustering of multiple genomes. Enlarged nucleoids are reported upon double knockout of MFN1 and MFN2 (Silva Ramos et al., 2019), as well as in cells with pathogenic variants of MSTO1 (Donkervoort et al., 2019) and FBXL4 (Sabouny et al., 2019). Intriguingly, there are conflicting reports about the role of OPA1 in enlarged nucleoids. Whereas siRNA-mediated knockdown of an alternatively spliced isoform of OPA1 (OPA1-exon4b) that is reported to bind mtDNA leads to enlarged nucleoids (Elachouri et al., 2011; Vidoni et al., 2013; Yang et al., 2020), OPA1-knockout MEFs lacking all isoforms have normal nucleoids (Ban-Ishihara et al., 2013; Silva Ramos et al., 2019). These contradictory findings suggest that an imbalance of the different OPA1 isoforms may be more detrimental than their complete absence, or the possibility of cell-type-specific responses to loss of OPA1. However, it should be noted that we do not have a complete understanding of the physiological role played by the OPA1-exon4b isoform, whose proposed topology (Elachouri et al., 2011) does not agree with the majority of OPA1 research suggesting the N-terminus resides within the matrix.

Another potential link between fusion and nucleoid remodeling is the lipid cardiolipin, which can bind nucleoids (Luévano-Martínez et al., 2015), and has a direct role in mediating IMM fusion (Ban et al., 2017). Importantly, manipulation of enzymes regulating cardiolipin impacts mitochondrial morphology and mtDNA nucleoids (Ban et al., 2018; Huang et al., 2020; Li et al., 2012). One interesting observation is that knockdown of the Barth syndrome protein Tafazzin (Hauff and Hatch, 2006), an enzyme that regulates remodeling of cardiolipin, causes accumulation of large nucleoids and mtDNA depletion, but does not impair mitochondrial fusion (Ban et al., 2018). This finding suggests that subtle changes in cardiolipin might impact nucleoids independently of its effects on mitochondrial fusion.

It remains unclear how impaired fusion results in enlarged nucleoids. One possibility, given that fission and fusion events are often coordinated (Abrisch et al., 2020; Twig et al., 2008), is that cells with reduced rates of fusion also have reduced rates of fission, and that it is a secondary reduction in fission accounting for nucleoid clustering. Alternatively, it could be that the balance between fission and fusion events is critical for the proper distribution of nucleoids. With respect to this latter possibility, it is worth noting that enlarged nucleoids observed in DRP1-knockdown cells were suppressed when either OPA1 or both of MFN1 and MFN2 were also depleted, whereas MFN1 overexpression induced nucleoid enlargement (Ban-Ishihara et al., 2013), suggesting that balanced fission and fusion is important. Finally, whereas enlarged nucleoids and mtDNA depletion can occur in cells with both impaired fission or fusion, they do not always occur together (e.g. OPA1 and MYH14). Another example of this separation of phenotypes is the fact that whereas treatment with the fission inhibitor M-divi1 rescues mtDNA copy number in cells of patients with pathogenic FBXL4 variants, some nucleoids remain enlarged (Sabouny et al., 2019). Thus, it does not appear that mtDNA depletion and enlarged nucleoids are necessarily caused by the same underlying mechanisms.

Fission and mtDNA replication

Although fission is clearly important for segregating mtDNA genomes to avoid enlarged nucleoids, it has also been proposed to play a role in initiating mtDNA replication. One idea that remains to be tested is that ER-mediated mitochondrial constriction initiates mtDNA replication (Lewis et al., 2016). However, it is not clear how the fission machinery, which is located outside mitochondria, would influence mtDNA in the matrix. Although a two-membrane spanning structure linked to replicating nucleoids has been described in yeast (Meeusen and Nunnari, 2003), similar machinery in mammalian cells is yet to be identified. In this regard, several factors worth discussing in the context of mtDNA replication and segregation include the mitochondrial contact site and cristae organizing system (MICOS) complex, which may provide a physical tether from across the OMM and IMM, IMM lipids, which physically interact with mtDNA, and the mitochondrial transport machinery, which is also implicated in nucleoid segregation.

The MICOS complex is involved in many functions, including lipid metabolism, protein import and maintaining mitochondrial architecture (Khosravi and Harner, 2020). Notably, this large multi-subunit structure links the IMM and the OMM, and is suggested to be a tether that could connect nucleoids to the exterior of the mitochondria (van der Laan et al., 2012). Disruption of the MICOS subunit Mic60 (also known as mitofilin, encoded by IMMT) leads to enlarged nucleoids, which appears to be partially independent of DRP1 (Li et al., 2016). Also supporting the notion of a role for the MICOS complex in mtDNA maintenance is the identification of pathogenic variants in the MICOS subunit MICOS13 (also known as QIl1) that lead to ∼90% mtDNA depletion (Kishita et al., 2020). Furthermore, pathogenic variants in CHCHD10, which regulates MICOS but is not a core subunit, lead to fewer but larger nucleoids (Genin et al., 2016), as well as mtDNA deletions (Bannwarth et al., 2014). Finally, loss of SAMM50, an OMM protein that interacts with MICOS, leads to an accumulation of large nucleoids (Jian et al., 2018). Thus, several lines of evidence demonstrate that the proper function of the MICOS complex is important for mtDNA maintenance.

The fact that IMM lipids are relevant to mtDNA maintenance should come as no surprise given that nucleoids interact with the IMM (Nicholls and Gustafsson, 2018). Two lipids in particular, cholesterol and cardiolipin, potentially link fission and mtDNA. Cardiolipin, which interacts physically with mtDNA and the MICOS complex (Friedman et al., 2015), can also localize to the OMM, where it interacts with DRP1 to promote mitochondrial fission (Mahajan et al., 2021). Meanwhile, despite its low mitochondrial abundance, cholesterol has also been implicated in mediating nucleoid interactions with the IMM (Gerhold et al., 2015). A role for cholesterol in mtDNA maintenance is supported by the fact that pharmacological disruption of cholesterol homeostasis leads to the accumulation of large nucleoids (Desai et al., 2017). Furthermore, defects in the intracellular cholesterol transporter NPC1, which cause Niemann–Pick disease type C, result in enlarged nucleoids (Desai et al., 2017). Similarly, ATAD3A, a regulator of mitochondrial cholesterol import, is also important for mtDNA maintenance, as its deletion results in mtDNA depletion and accumulation of mtDNA deletions (Peralta et al., 2018). Moreover, pathogenic variants in ATAD3A lead to enlarged nucleoids (Desai et al., 2017). However, there are several ways in which ATAD3A could impact mtDNA in addition to affecting cholesterol, as ATAD3A is proposed to be a nucleoid-associated protein (Bogenhagen et al., 2008; He et al., 2007), and its loss also destabilizes the MICOS complex (Peralta et al., 2018). Finally, as the import of cholesterol into mitochondria is mediated by interactions between ER and mitochondria (Martin et al., 2016), it is tempting to speculate that this process is coordinated with fission initiation to promote cholesterol import. Such a coordination could favor a localized cholesterol accumulation in the IMM, perhaps creating a platform for replication that signals the initiation of mtDNA replication.

Another consideration for nucleoid segregation is the dedicated transport machinery that moves mitochondria about the cell via the cytoskeleton (Kruppa and Buss, 2021), components of which are associated with nucleoids. While early work showed that nucleoids localize near the KIF5B kinesin motor, which moves mitochondria along microtubules (Iborra et al., 2004), recent work also showed an enrichment of MIRO1, a mitochondrial OMM adaptor protein involved in transport, with the MICOS subunit MIC60 and nucleoids at mitochondria–ER contacts (Qin et al., 2020). Thus, it is possible that the mitochondrial transport machinery is tethered to mtDNA nucleoids via the MICOS complex, which might help to distribute mitochondria with newly synthesized nucleoids within the mitochondrial network following fission by physically separation.

Further arguing for a role of fission in initiating mtDNA replication, impaired fission can also cause reduced mtDNA copy number. For example, blocking fission leads to mtDNA depletion in several experimental models (e.g. genetic ablation of DRP1, MFF and MYH10) (Parone et al., 2008; Reyes et al., 2011). However, it is notable that a transient reduction in fission (e.g. siRNA-mediated knockdown of DRP1 or MFF) does not lead to mtDNA depletion, despite leading to nucleoid clustering (Ban-Ishihara et al., 2013). These discrepancies could be explained if a more severe inhibition of fission is required to induce loss of mtDNA, or if nucleoid clustering precedes mtDNA depletion. Nevertheless, it seems that nucleoid clustering and mtDNA depletion are not necessarily linked.

Fusion and mtDNA replication

One of the earliest connections between mitochondrial network remodeling and mtDNA copy number dynamics was noted in yeast, where disruption of the OPA1 and MFN homologs, Mgm1 and Fzo1, respectively, led to a complete loss of mtDNA (Jones and Fangman, 1992; Rapaport et al., 1998). This observation has since been extended to mammalian models of impaired mitochondrial fusion, as MEFs lacking either OPA1 or both MFN1 and MFN2 have a lower mtDNA copy number (Chen et al., 2007; Silva Ramos et al., 2019). Likewise, mice in which these mitochondrial fusion proteins are conditionally knocked out also exhibit reduced mtDNA copy number in targeted tissues (Chen et al., 2010; Silva Ramos et al., 2019).

Insights from mitochondrial diseases has strengthened the connection between fusion and mtDNA. For example, some (but not all) pathogenic variants of both MFN2 and OPA1 can cause mtDNA depletion (Elachouri et al., 2011; Spiegel et al., 2016; Vielhaber et al., 2013). An explanation for this discrepancy could be that not all pathogenic variants impair mitochondrial fusion to the same degree. Alternatively, it is worth noting that MFN2 and OPA1 also have additional cellular functions that are independent of fusion (e.g. mitochondria–ER tethering and stabilizing cristae junctions, respectively), which could cause mitochondrial dysfunction if they are impaired but not affect mtDNA. In addition, pathogenic variants of the fusion regulators FBXL4 and MSTO1 also cause mtDNA depletion (Donkervoort et al., 2019; Sabouny et al., 2019). Of particular interest, restoring the mitochondrial network morphology by treatment with the fission inhibitor m-Divi1 was sufficient to restore mtDNA copy number in cells harboring pathogenic variants in FBXL4 (Sabouny et al., 2019). Similarly, mtDNA copy number rescue was observed when crossing mice harboring a fission defect (MFF deletion) with mice that have a fusion defect (MFN1 deletion) (Chen et al., 2015). Thus, there is hope that therapeutic strategies to restore mitochondrial morphology by reducing fission or increasing fusion could provide clinical benefits for patients with mtDNA depletion owing to impaired mitochondrial fission or fusion.

Despite the evidence that impaired fusion leads to mtDNA depletion, the underlying mechanism is not completely understood. There are two, non-mutually exclusive, explanations for how impaired mitochondrial fusion leads to mtDNA depletion – reduced replication and/or increased mtDNA turnover. Supporting the reduced mtDNA replication hypothesis, the relative levels of mitochondrial replisome enzymes have been shown to be significantly disrupted in models of impaired fusion (Silva Ramos et al., 2019). Whereas the protein levels of mtSSB and POLG are reduced in MEFs and conditional MFN-knockout mice, the helicase Twinkle was upregulated in these animals (Silva Ramos et al., 2019). Thus, the authors posit that mitochondrial fusion and content mixing are important for maintaining the stoichiometric balance in mitochondrial replisome components (Silva Ramos et al., 2019). In support of the increased turnover notion, loss of the fusion regulator FBXL4 increases mitochondrial removal via mitophagy (Alsina et al., 2020). In this case, it is tempting to speculate that reduced fusion leads to smaller mitochondria that are simply more likely to be degraded by mitophagy. Whether this increase in mitophagy leads to increased turnover of mtDNA and reduced copy number is unknown.

Another intriguing connection between mtDNA and mitochondrial fusion is the recent finding that the anti-inflammation drug leflunomide increases hyperfusion of the mitochondrial network by increasing the expression of both MFN1 and MFN2 (Miret-Casals et al., 2018). Interestingly, leflunomide inhibits dihydroorotate dehydrogenoase (DHODH), a mitochondrial enzyme required for the synthesis of pyrimidine nucleotides, suggesting that cells may increase expression of MFN1 and MFN2 in response to reduced levels of pyrimidines. Whether this previously unappreciated connection between mitochondrial fusion and dNTPs is also relevant to mtDNA depletion disorders remains unknown. Although loss of mitochondrial fusion does not appear to affect cellular dNTP levels (Silva Ramos et al., 2019), the effects on mitochondrial dNTP pools have not been investigated. Given the role of fusion in content mixing to distribute the mtDNA replication machinery, it is easy to envision fusion playing a similar role in delivering dNTPs for mtDNA replication. Regardless, as reduced DHODH activity also leads to reduced mtDNA copy number (Fang et al., 2013), leflunomide may not be an appropriate choice as a therapeutic for mtDNA depletion syndromes due to impaired mitochondrial fusion.

Mitochondrial dynamics and mtDNA heteroplasmy

Another link between impaired mitochondrial dynamics and the mitochondrial genome is the accumulation of mtDNA mutations. For example, blocking fission in cultured cells leads to increased levels of mutant mtDNA (Malena et al., 2009). Another interesting study supporting a role for fission in mediating heteroplasmy shifting involved crossing a mouse model of mtDNA heteroplasmy comprising a mix of two non-pathogenic mtDNA haplotypes (NZB and BALB) with the Python mouse harboring a pathogenic mutation in DRP1 (Jokinen et al., 2016). This work showed selection for mtDNA haplotypes, but only in liver, kidney and hematopoietic tissues, emphasizing the importance of tissue specificity. With respect to fusion, disruption of MFN1 and MFN2 in mouse skeletal muscle leads to elevated levels of mtDNA point mutations and deletions (Chen et al., 2010). However, these effects are likely tissue specific, as no differences in mtDNA deletions or point mutations were observed in a cardiac-specific double MFN1 and MFN2 knockout mice (Silva Ramos et al., 2019). Relevant to disease, accumulation of mtDNA deletions has been reported in patient cells harboring certain pathogenic variants in MFN2 (Vielhaber et al., 2013) and OPA1 (Amati-Bonneau et al., 2008; Hudson et al., 2008). As fusion-mediated content mixing is important for the distribution of enzymes that regulate mtDNA, the accumulation of point mutations when fusion is impaired may be due to a reduced capacity to repair mtDNA. Supporting this notion, fibroblasts from patients with pathogenic MFN2 variants showed reduced mtDNA repair (Rouzier et al., 2012). Additionally, given that human mtDNA can be degraded by components of the mtDNA replication machinery (Nissanka et al., 2018; Peeva et al., 2018; Wiehe et al., 2018), reduced content mixing of enzymes that can degrade mtDNA could also lead to reduced turnover of mutant mtDNA. With respect to mtDNA deletions, which cannot be repaired by the limited mtDNA repair machinery, their accumulation must be solely due to increased mutational rates. Given that stalled mtDNA replication complexes are believed to be a contributing factor to the production of mtDNA deletions (Nissanka et al., 2019), it is possible that reduced content mixing and uneven distribution of the mtDNA replication machinery causes stalling, promoting the generation of mtDNA deletions. Supporting the importance of fusion-dependent content mixing to protect against mutant mtDNA accumulation is the finding that crossing MFN1−/− mice with the POLGA mutator mice, which accumulate mutations due to a proofreading deficiency in the polymerase, is embryonic lethal, despite the individual lines being viable (Chen et al., 2010). Nevertheless, it is curious that mtDNA deletions are only linked to a few pathogenic variants in MFN2 or OPA1, and are not observed in cells harboring variants in FBXL4 (Sabouny et al., 2019). These observations suggests that a threshold of impaired fusion is required for deletions, that deletions occur in a tissue-specific manner or that other factors may also be involved in eliminating mtDNA mutations.

Mitophagy and mtDNA

Mitophagy is another dynamic process relevant to mtDNA regulation in the context of quality control and potential mtDNA elimination. Exciting recent work highlights the fate of mitochondria following fission and describes a distinct type of fission event that occurs near the ends of mitochondrial tubules, with smaller mitochondrial fragments that contain damaged cargo recruiting Parkin and subsequently undergoing mitophagy (Kleele et al., 2021). Given that mitophagy requires mitochondrial fission, initial observations that blocking fission increased mutant mtDNA levels (Malena et al., 2009), implicated mitophagy as a process important for the removal of mutant mtDNA. Similarly, mitophagy is required for the degradation of damaged mtDNA (Bess et al., 2013; Shu et al., 2021). Moreover, the induction of DNA damage leads to increased prevalence of dysfunctional mtDNA in mitochondrial fragments destined for mitophagy (Kleele et al., 2021). These findings suggest a mechanism by which mitochondria harboring mutant or damaged mtDNA genomes are preferentially recognized for mitophagy. In this regard, cells harboring mutant mtDNA genomes are reported to have reduced IMM fusion (Mishra et al., 2014), which may lead to smaller mitochondrial fragments that are more susceptible to mitophagy.

Several lines of evidence demonstrate that mitophagy selects against deleterious mtDNA mutations (Hahn and Zuryn, 2019). For example, loss of the mitophagy protein Parkin leads to an increase in the predicted pathogenicity of mtDNA mutations in the striatum of POLGA mutant mice (Pickrell et al., 2015). Similarly, Parkin overexpression reduces heteroplasmy levels of mutant mtDNA genomes in both cultured cells (Suen et al., 2010) and in Drosophila (Kandul et al., 2016). Conversely, mutations in PINK1 and Parkin that reduce mitophagy favor the accumulation of mutant mtDNA in several C. elegans models (Valenci et al., 2015). Beyond the well-studied PINK1–Parkin-mediated form of mitophagy, Parkin-independent mitophagy mediated by ATAD3B also promotes the removal of pathogenic mtDNA variants (Shu et al., 2021). However, despite evidence that mitophagy can eliminate mutant mtDNA in certain circumstances, the fact that mtDNA mutations can and do accumulate to pathogenic levels suggests that normal rates of mitophagy are insufficient to completely prevent mtDNA disease (de Vries et al., 2012). In this respect, it is notable that Parkin expression is downregulated in response to certain mtDNA mutations (Gilkerson et al., 2012). Such a response may be a compensatory mechanism to prevent mitophagy from eliminating too many mitochondria, which would be detrimental to the cell. Nevertheless, approaches to upregulate mitophagy offer a potential approach to reduce mtDNA heteroplasmy, and thus may have therapeutic benefits (Diot et al., 2015). In this regard, it is notable that promoting general autophagy by rapamycin treatment reduces mtDNA heteroplasmy in cultured cells (Dai et al., 2014).

Another emerging connection between mitophagy and mtDNA relates to the release of mtDNA into the cytosol (McArthur et al., 2018; Riley et al., 2018), where it can activate innate immune pathways such as the cGAS-STING axis (West et al., 2015). While it remains to be seen how mtDNA release is influenced by fusion or fission, and whether it can also contribute to mtDNA depletion, two recent reports link the PINK1–Parkin mitophagy pathway to mtDNA release and activation of the cGAS-STING pathway. The first study showed that following exhaustive exercise, mice lacking either PINK1 or Parkin had a strong inflammatory response dependent on STING (Sliter et al., 2018). A second study has directly linked PINK1 deficiency to mtDNA release into the cytosol (Bueno et al., 2019). Intriguingly, activation of the cGAS-STING pathway, which can be mediated by release of mtDNA into the cytosol, induces autophagy (Gui et al., 2019). Although this increase in autophagy has not yet been linked to mitophagy specifically, it is tempting to speculate that cytosolic mtDNA acts as a signal of mitochondrial distress that can induce mitophagy to target damaged mitochondria.

Mitochondrial dynamics are important for regulating many aspects of mitochondrial function, including mtDNA maintenance. As highlighted above, both fission and fusion of the mitochondrial network are important for mediating the distribution of nucleoids throughout the mitochondrial network, maintaining mtDNA copy number, and regulating mtDNA heteroplasmy. In addition, mitophagy, which is linked to mitochondrial fission and fusion, is also important for mtDNA turnover, which could impact both the copy number and heteroplasmy levels of mutant mtDNA. Nonetheless, many critical questions remain to be answered about how exactly mitochondrial dynamics regulate mtDNA. The fact that impairment of the opposing forces of fusion and fission has similar outcomes with respect to mtDNA is somewhat surprising as our current understanding suggests that the underlying mechanisms for how both fusion and fission affect mtDNA may not be the same. For example, fission appears to be directly involved in initiating the replication of mtDNA, whereas fusion appears to be required for an even distribution of the mtDNA replication machinery. Future work will need to address whether the mechanisms leading to loss of mtDNA are indeed distinct in different models of impaired fission and fusion. The fact that different pathogenic variants in both fusion and fission proteins can affect mtDNA differently suggests that the underlying mechanisms are more complicated than currently appreciated. Another question that will need to be answered is how fission events, which are mediated by cytosolic and OMM proteins, are coordinated with the replication of the mtDNA in the matrix. With respect to the distribution of nucleoids throughout the mitochondrial network, the functional consequences of their clustering remain to be elucidated. There are also still questions with regard to whether nucleoid distribution is linked to copy number maintenance, although it does not appear that this is always the case. Moreover, although fission appears to be required to segregate nucleoids, exactly why nucleoids cluster upon inhibition of fusion remains unknown. Overall, although there is certainly more to learn mechanistically about how fusion, fission and mitophagy impact mtDNA, the lessons gleaned from both basic mechanistic models and human diseases clearly highlight the importance of this link. A better understanding of how and why mitochondrial dynamics impact mtDNA will be important in developing novel therapeutic approaches to treat diseases of mitochondrial dysfunction by restoring mtDNA copy number or eliminating mutant mtDNA.

Funding

Our work in this area is supported by a Natural Sciences and Engineering Research Council of Canada Discovery Grant to T.E.S.

Abrisch
,
R. G.
,
Gumbin
,
S. C.
,
Wisniewski
,
B. T.
,
Lackner
,
L. L.
and
Voeltz
,
G. K.
(
2020
).
Fission and fusion machineries converge at ER contact sites to regulate mitochondrial morphology
.
J. Cell Biol.
219
,
e201911122
.
Alán
,
L.
,
Špaček
,
T.
,
Pajuelo Reguera
,
D.
,
Jabůrek
,
M.
and
Ježek
,
P.
(
2016
).
Mitochondrial nucleoid clusters protect newly synthesized mtDNA during Doxorubicin- and Ethidium Bromide-induced mitochondrial stress
.
Toxicol. Appl. Pharmacol.
302
,
31
-
40
.
Al Khatib
,
I.
and
Shutt
,
T. E.
(
2019
).
Advances towards therapeutic approaches for mtDNA disease
.
Adv. Exp. Med. Biol.
1158
,
217
-
246
.
Alexander
,
C.
,
Votruba
,
M.
,
Pesch
,
U. E.
,
Thiselton
,
D. L.
,
Mayer
,
S.
,
Moore
,
A.
,
Rodriguez
,
M.
,
Kellner
,
U.
,
Leo-Kottler
,
B.
,
Auburger
,
G.
et al. 
(
2000
).
OPA1, encoding a dynamin-related GTPase, is mutated in autosomal dominant optic atrophy linked to chromosome 3q28
.
Nat. Genet.
26
,
211
-
215
.
Almutawa
,
W.
,
Smith
,
C.
,
Sabouny
,
R.
,
Smit
,
R. B.
,
Zhao
,
T.
,
Wong
,
R.
,
Lee-Glover
,
L.
,
Desrochers-Goyette
,
J.
,
Ilamathi
,
H. S.
,
Suchowersky
,
O.
et al. 
(
2019
).
The R941L mutation in MYH14 disrupts mitochondrial fission and associates with peripheral neuropathy
.
EBioMedicine
45
,
379
-
392
.
Alsina
,
D.
,
Lytovchenko
,
O.
,
Schab
,
A.
,
Atanassov
,
I.
,
Schober
,
F. A.
,
Jiang
,
M.
,
Koolmeister
,
C.
,
Wedell
,
A.
,
Taylor
,
R. W.
,
Wredenberg
,
A.
et al. 
(
2020
).
FBXL4 deficiency increases mitochondrial removal by autophagy
.
EMBO Mol. Med.
12
,
e11659
.
Amati-Bonneau
,
P.
,
Valentino
,
M. L.
,
Reynier
,
P.
,
Gallardo
,
M. E.
,
Bornstein
,
B.
,
Boissiere
,
A.
,
Campos
,
Y.
,
Rivera
,
H.
,
de la Aleja
,
J. G.
,
Carroccia
,
R.
et al. 
(
2008
).
OPA1 mutations induce mitochondrial DNA instability and optic atrophy ‘plus’ phenotypes
.
Brain
131
,
338
-
351
.
Anderson
,
S.
,
Bankier
,
A. T.
,
Barrell
,
B. G.
,
de Bruijn
,
M. H.
,
Coulson
,
A. R.
,
Drouin
,
J.
,
Eperon
,
I. C.
,
Nierlich
,
D. P.
,
Roe
,
B. A.
,
Sanger
,
F.
et al. 
(
1981
).
Sequence and organization of the human mitochondrial genome
.
Nature
290
,
457
-
465
.
Antoun
,
G.
,
McBride
,
S.
,
Vanstone
,
J. R.
,
Naas
,
T.
,
Michaud
,
J.
,
Redpath
,
S.
,
McMillan
,
H. J.
,
Brophy
,
J.
,
Daoud
,
H.
,
Chakraborty
,
P.
et al. 
(
2015
).
Detailed biochemical and bioenergetic characterization of FBXL4-related encephalomyopathic mitochondrial DNA depletion
.
JIMD Rep
27
,
1
-
9
.
Ashley
,
N.
and
Poulton
,
J.
(
2009
).
Anticancer DNA intercalators cause p53-dependent mitochondrial DNA nucleoid re-modelling
.
Oncogene
28
,
3880
-
3891
.
Ballout
,
R. A.
,
Al Alam
,
C.
,
Bonnen
,
P. E.
,
Huemer
,
M.
,
El-Hattab
,
A. W.
and
Shbarou
,
R.
(
2019
).
FBXL4-Related mitochondrial DNA depletion syndrome 13 (MTDPS13): a case report with a comprehensive mutation review
.
Frontiers in Genetics
10
,
39
.
Ban-Ishihara
,
R.
,
Ishihara
,
T.
,
Sasaki
,
N.
,
Mihara
,
K.
and
Ishihara
,
N.
(
2013
).
Dynamics of nucleoid structure regulated by mitochondrial fission contributes to cristae reformation and release of cytochrome c
.
Proc. Natl. Acad. Sci. USA
110
,
11863
-
11868
.
Ban
,
T.
,
Ishihara
,
T.
,
Kohno
,
H.
,
Saita
,
S.
,
Ichimura
,
A.
,
Maenaka
,
K.
,
Oka
,
T.
,
Mihara
,
K.
and
Ishihara
,
N.
(
2017
).
Molecular basis of selective mitochondrial fusion by heterotypic action between OPA1 and cardiolipin
.
Nat. Cell Biol.
19
,
856
-
863
.
Ban
,
T.
,
Kohno
,
H.
,
Ishihara
,
T.
and
Ishihara
,
N.
(
2018
).
Relationship between OPA1 and cardiolipin in mitochondrial inner-membrane fusion
.
Biochim. Biophys. Acta Bioenerg.
1859
,
951
-
957
.
Bannwarth
,
S.
,
Ait-El-Mkadem
,
S.
,
Chaussenot
,
A.
,
Genin
,
E. C.
,
Lacas-Gervais
,
S.
,
Fragaki
,
K.
,
Berg-Alonso
,
L.
,
Kageyama
,
Y.
,
Serre
,
V.
,
Moore
,
D. G.
et al. 
(
2014
).
A mitochondrial origin for frontotemporal dementia and amyotrophic lateral sclerosis through CHCHD10 involvement
.
Brain
137
,
2329
-
2345
.
Baroy
,
T.
,
Pedurupillay
,
C. R.
,
Bliksrud
,
Y. T.
,
Rasmussen
,
M.
,
Holmgren
,
A.
,
Vigeland
,
M. D.
,
Hughes
,
T.
,
Brink
,
M.
,
Rodenburg
,
R.
,
Nedregaard
,
B.
et al. 
(
2016
).
A novel mutation in FBXL4 in a Norwegian child with encephalomyopathic mitochondrial DNA depletion syndrome 13
.
Eur. J. Med. Genet.
59
,
342
-
346
.
Bartsakoulia
,
M.
,
Pyle
,
A.
,
Troncoso-Chandía
,
D.
,
Vial-Brizzi
,
J.
,
Paz-Fiblas
,
M. V.
,
Duff
,
J.
,
Griffin
,
H.
,
Boczonadi
,
V.
,
Lochmüller
,
H.
,
Kleinle
,
S.
et al. 
(
2018
).
A novel mechanism causing imbalance of mitochondrial fusion and fission in human myopathies
.
Hum. Mol. Genet.
27
,
1186
-
1195
.
Bess
,
A. S.
,
Ryde
,
I. T.
,
Hinton
,
D. E.
and
Meyer
,
J. N.
(
2013
).
UVC-Induced Mitochondrial Degradation via Autophagy Correlates with mtDNA Damage Removal in Primary Human Fibroblasts
.
J. Biochem. Mol. Toxicol.
27
,
28
-
41
.
Bogenhagen
,
D. F.
,
Rousseau
,
D.
and
Burke
,
S.
(
2008
).
The layered structure of human mitochondrial DNA nucleoids
.
J. Biol. Chem.
283
,
3665
-
3675
.
Bonnen
,
P. E.
,
Yarham
,
J. W.
,
Besse
,
A.
,
Wu
,
P.
,
Faqeih
,
E. A.
,
Al-Asmari
,
A. M.
,
Saleh
,
M. A.
,
Eyaid
,
W.
,
Hadeel
,
A.
,
He
,
L.
et al. 
(
2013
).
Mutations in FBXL4 cause mitochondrial encephalopathy and a disorder of mitochondrial DNA maintenance
.
Am. J. Hum. Genet.
93
,
471
-
481
.
Bueno
,
M.
,
Zank
,
D.
,
Buendia-Roldán
,
I.
,
Fiedler
,
K.
,
Mays
,
B. G.
,
Alvarez
,
D.
,
Sembrat
,
J.
,
Kimball
,
B.
,
Bullock
,
J. K.
,
Martin
,
J. L.
et al. 
(
2019
).
PINK1 attenuates mtDNA release in alveolar epithelial cells and TLR9 mediated profibrotic responses
.
PLoS One
14
,
e0218003
.
Chan
,
D. C.
(
2019
).
Mitochondrial dynamics and its involvement in disease
.
Annu. Rev. Pathol.
15
,
235
-
259
.
Chan
,
D. C.
(
2012
).
Fusion and fission: interlinked processes critical for mitochondrial health
.
Annu. Rev. Genet.
46
,
265
-
287
.
Chapman
,
J.
,
Ng
,
Y. S.
and
Nicholls
,
T. J.
(
2020
).
The maintenance of mitochondrial DNA integrity and dynamics by mitochondrial membranes
.
Life (Basel)
10
,
164
.
Chen
,
H.
and
Chan
,
D. C.
(
2010
).
Physiological functions of mitochondrial fusion
.
Ann. N. Y. Acad. Sci.
1201
,
21
-
25
.
Chen
,
Y.
and
Dorn
,
G. W.
II
. (
2013
).
PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria
.
Science
340
,
471
-
475
.
Chen
,
H.
,
Detmer
,
S. A.
,
Ewald
,
A. J.
,
Griffin
,
E. E.
,
Fraser
,
S. E.
and
Chan
,
D. C.
(
2003
).
Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development
.
J. Cell Biol.
160
,
189
-
200
.
Chen
,
H.
,
McCaffery
,
J. M.
and
Chan
,
D. C.
(
2007
).
Mitochondrial fusion protects against neurodegeneration in the cerebellum
.
Cell
130
,
548
-
562
.
Chen
,
H.
,
Vermulst
,
M.
,
Wang
,
Y. E.
,
Chomyn
,
A.
,
Prolla
,
T. A.
,
McCaffery
,
J. M.
and
Chan
,
D. C.
(
2010
).
Mitochondrial fusion is required for mtDNA stability in skeletal muscle and tolerance of mtDNA mutations
.
Cell
141
,
280
-
289
.
Chen
,
H.
,
Ren
,
S.
,
Clish
,
C.
,
Jain
,
M.
,
Mootha
,
V.
,
McCaffery
,
J. M.
and
Chan
,
D. C.
(
2015
).
Titration of mitochondrial fusion rescues Mff-deficient cardiomyopathy
.
J. Cell Biol.
211
,
795
-
805
.
Choi
,
B. O.
,
Kang
,
S. H.
,
Hyun
,
Y. S.
,
Kanwal
,
S.
,
Park
,
S. W.
,
Koo
,
H.
,
Kim
,
S. B.
,
Choi
,
Y. C.
,
Yoo
,
J. H.
,
Kim
,
J. W.
et al. 
(
2011
).
A complex phenotype of peripheral neuropathy, myopathy, hoarseness, and hearing loss is linked to an autosomal dominant mutation in MYH14
.
Hum. Mutat.
32
,
669
-
677
.
Dai
,
Y.
,
Zheng
,
K.
,
Clark
,
J.
,
Swerdlow
,
R. H.
,
Pulst
,
S. M.
,
Sutton
,
J. P.
,
Shinobu
,
L. A.
and
Simon
,
D. K.
(
2014
).
Rapamycin drives selection against a pathogenic heteroplasmic mitochondrial DNA mutation
.
Hum. Mol. Genet.
23
,
637
-
647
.
Dai
,
H.
,
Zhang
,
V. W.
,
El-Hattab
,
A. W.
,
Ficicioglu
,
C.
,
Shinawi
,
M.
,
Lines
,
M.
,
Schulze
,
A.
,
McNutt
,
M.
,
Gotway
,
G.
,
Tian
,
X.
et al. 
(
2017
).
FBXL4 defects are common in patients with congenital lactic acidemia and encephalomyopathic mitochondrial DNA depletion syndrome
.
Clin. Genet.
91
,
634
-
639
.
Dalla Rosa
,
I.
,
Durigon
,
R.
,
Pearce
,
S. F.
,
Rorbach
,
J.
,
Hirst
,
E. M.
,
Vidoni
,
S.
,
Reyes
,
A.
,
Brea-Calvo
,
G.
,
Minczuk
,
M.
,
Woellhaf
,
M. W.
et al. 
(
2014
).
MPV17L2 is required for ribosome assembly in mitochondria
.
Nucleic Acids Res.
42
,
8500
-
8515
.
Desai
,
R.
,
Frazier
,
A. E.
,
Durigon
,
R.
,
Patel
,
H.
,
Jones
,
A. W.
,
Dalla Rosa
,
I.
,
Lake
,
N. J.
,
Compton
,
A. G.
,
Mountford
,
H. S.
,
Tucker
,
E. J.
et al. 
(
2017
).
ATAD3 gene cluster deletions cause cerebellar dysfunction associated with altered mitochondrial DNA and cholesterol metabolism
.
Brain
140
,
1595
-
1610
.
de Vries
,
R. L.
,
Gilkerson
,
R. W.
,
Przedborski
,
S.
and
Schon
,
E. A.
(
2012
).
Mitophagy in cells with mtDNA mutations: being sick is not enough
.
Autophagy
8
,
699
-
700
.
Diot
,
A.
,
Hinks-Roberts
,
A.
,
Lodge
,
T.
,
Liao
,
C.
,
Dombi
,
E.
,
Morten
,
K.
,
Brady
,
S.
,
Fratter
,
C.
,
Carver
,
J.
,
Muir
,
R.
et al. 
(
2015
).
A novel quantitative assay of mitophagy: Combining high content fluorescence microscopy and mitochondrial DNA load to quantify mitophagy and identify novel pharmacological tools against pathogenic heteroplasmic mtDNA
.
Pharmacol. Res.
100
,
24
-
35
.
Donkervoort
,
S.
,
Sabouny
,
R.
,
Yun
,
P.
,
Gauquelin
,
L.
,
Chao
,
K. R.
,
Hu
,
Y.
,
Al Khatib
,
I.
,
Topf
,
A.
,
Mohassel
,
P.
,
Cummings
,
B. B.
et al. 
(
2019
).
MSTO1 mutations cause mtDNA depletion, manifesting as muscular dystrophy with cerebellar involvement
.
Acta Neuropathol.
138
,
1013
-
1031
.
Ebrahimi-Fakhari
,
D.
,
Seitz
,
A.
,
Kolker
,
S.
and
Hoffmann
,
G. F.
(
2015
).
Recurrent stroke-like episodes in FBXL4-associated early-onset mitochondrial encephalomyopathy
.
Pediatr. Neurol.
53
,
549
-
550
.
Elachouri
,
G.
,
Vidoni
,
S.
,
Zanna
,
C.
,
Pattyn
,
A.
,
Boukhaddaoui
,
H.
,
Gaget
,
K.
,
Yu-Wai-Man
,
P.
,
Gasparre
,
G.
,
Sarzi
,
E.
,
Delettre
,
C.
et al. 
(
2011
).
OPA1 links human mitochondrial genome maintenance to mtDNA replication and distribution
.
Genome Res.
21
,
12
-
20
.
El-Hattab
,
A. W.
,
Dai
,
H.
,
Almannai
,
M.
,
Wang
,
J.
,
Faqeih
,
E. A.
,
Al Asmari
,
A.
,
Saleh
,
M. A. M.
,
Elamin
,
M. A. O.
,
Alfadhel
,
M.
,
Alkuraya
,
F. S.
et al. 
(
2017
).
Molecular and clinical spectra of FBXL4 deficiency
.
Hum. Mutat.
38
,
1649
-
1659
.
Fahrner
,
J. A.
,
Liu
,
R.
,
Perry
,
M. S.
,
Klein
,
J.
and
Chan
,
D. C.
(
2016
).
A novel de novo dominant negative mutation in DNM1L impairs mitochondrial fission and presents as childhood epileptic encephalopathy
.
Am. J. Med. Genet. A
170
,
2002
-
2011
.
Fang
,
J.
,
Uchiumi
,
T.
,
Yagi
,
M.
,
Matsumoto
,
S.
,
Amamoto
,
R.
,
Takazaki
,
S.
,
Yamaza
,
H.
,
Nonaka
,
K.
and
Kang
,
D.
(
2013
).
Dihydro-orotate dehydrogenase is physically associated with the respiratory complex and its loss leads to mitochondrial dysfunction
.
Biosci. Rep.
33
,
e00021
-
e00021
.
Feric
,
M.
,
Demarest
,
T. G.
,
Tian
,
J.
,
Croteau
,
D. L.
,
Bohr
,
V. A.
and
Misteli
,
T.
(
2021
).
Self-assembly of multi-component mitochondrial nucleoids via phase separation
.
EMBO J.
40
,
e107165
.
Friedman
,
J. R.
,
Lackner
,
L. L.
,
West
,
M.
,
DiBenedetto
,
J. R.
,
Nunnari
,
J.
and
Voeltz
,
G. K.
(
2011
).
ER tubules mark sites of mitochondrial division
.
Science
334
,
358
-
362
.
Friedman
,
J. R.
,
Mourier
,
A.
,
Yamada
,
J.
,
McCaffery
,
J. M.
and
Nunnari
,
J.
(
2015
).
MICOS coordinates with respiratory complexes and lipids to establish mitochondrial inner membrane architecture
.
Elife
4
,
e07739
.
Gai
,
X.
,
Ghezzi
,
D.
,
Johnson
,
M. A.
,
Biagosch
,
C. A.
,
Shamseldin
,
H. E.
,
Haack
,
T. B.
,
Reyes
,
A.
,
Tsukikawa
,
M.
,
Sheldon
,
C. A.
,
Srinivasan
,
S.
et al. 
(
2013
).
Mutations in FBXL4, encoding a mitochondrial protein, cause early-onset mitochondrial encephalomyopathy
.
Am. J. Hum. Genet.
93
,
482
-
495
.
Gal
,
A.
,
Balicza
,
P.
,
Weaver
,
D.
,
Naghdi
,
S.
,
Joseph
,
S. K.
,
Varnai
,
P.
,
Gyuris
,
T.
,
Horvath
,
A.
,
Nagy
,
L.
,
Seifert
,
E. L.
et al. 
(
2017
).
MSTO1 is a cytoplasmic pro-mitochondrial fusion protein, whose mutation induces myopathy and ataxia in humans
.
EMBO Mol. Med.
9
,
967
-
984
.
Genin
,
E. C.
,
Plutino
,
M.
,
Bannwarth
,
S.
,
Villa
,
E.
,
Cisneros-Barroso
,
E.
,
Roy
,
M.
,
Ortega-Vila
,
B.
,
Fragaki
,
K.
,
Lespinasse
,
F.
,
Pinero-Martos
,
E.
et al. 
(
2016
).
CHCHD10 mutations promote loss of mitochondrial cristae junctions with impaired mitochondrial genome maintenance and inhibition of apoptosis
.
EMBO Mol. Med.
8
,
58
-
72
.
Gerhold
,
J. M.
,
Cansiz-Arda
,
Ş.
,
Lõhmus
,
M.
,
Engberg
,
O.
,
Reyes
,
A.
,
van Rennes
,
H.
,
Sanz
,
A.
,
Holt
,
I. J.
,
Cooper
,
H. M.
and
Spelbrink
,
J. N.
(
2015
).
Human mitochondrial DNA-protein complexes attach to a cholesterol-rich membrane structure
.
Sci. Rep.
5
,
15292
.
Gilkerson
,
R. W.
,
Schon
,
E. A.
,
Hernandez
,
E.
and
Davidson
,
M. M.
(
2008
).
Mitochondrial nucleoids maintain genetic autonomy but allow for functional complementation
.
J. Cell Biol.
181
,
1117
-
1128
.
Gilkerson
,
R. W.
,
De Vries
,
R. L.
,
Lebot
,
P.
,
Wikstrom
,
J. D.
,
Torgyekes
,
E.
,
Shirihai
,
O. S.
,
Przedborski
,
S.
and
Schon
,
E. A.
(
2012
).
Mitochondrial autophagy in cells with mtDNA mutations results from synergistic loss of transmembrane potential and mTORC1 inhibition
.
Hum. Mol. Genet.
21
,
978
-
990
.
Glancy
,
B.
,
Kim
,
Y.
,
Katti
,
P.
and
Willingham
,
T. B.
(
2020
).
The functional impact of mitochondrial structure across subcellular scales
.
Front. Physiol.
11
,
541040
.
Gomes
,
L. C.
,
Di Benedetto
,
G.
and
Scorrano
,
L.
(
2011
).
During autophagy mitochondria elongate, are spared from degradation and sustain cell viability
.
Nat. Cell Biol.
13
,
589
-
598
.
Gorman
,
G. S.
,
Chinnery
,
P. F.
,
DiMauro
,
S.
,
Hirano
,
M.
,
Koga
,
Y.
,
McFarland
,
R.
,
Suomalainen
,
A.
,
Thorburn
,
D. R.
,
Zeviani
,
M.
and
Turnbull
,
D. M.
(
2016
).
Mitochondrial diseases
.
Nat. Rev. Dis. Primers
2
,
16080
.
Gui
,
X.
,
Yang
,
H.
,
Li
,
T.
,
Tan
,
X.
,
Shi
,
P.
,
Li
,
M.
,
Du
,
F.
and
Chen
,
Z. J.
(
2019
).
Autophagy induction via STING trafficking is a primordial function of the cGAS pathway
.
Nature
567
,
262
-
266
.
Gustafsson
,
C. M.
,
Falkenberg
,
M.
and
Larsson
,
N. G.
(
2016
).
Maintenance and expression of mammalian mitochondrial DNA
.
Annu. Rev. Biochem.
85
,
133
-
160
.
Hahn
,
A.
and
Zuryn
,
S.
(
2019
).
The cellular mitochondrial genome landscape in disease
.
Trends Cell Biol.
29
,
227
-
240
.
Hauff
,
K. D.
and
Hatch
,
G. M.
(
2006
).
Cardiolipin metabolism and Barth Syndrome
.
Prog. Lipid Res.
45
,
91
-
101
.
He
,
J.
,
Mao
,
C. C.
,
Reyes
,
A.
,
Sembongi
,
H.
,
Di Re
,
M.
,
Granycome
,
C.
,
Clippingdale
,
A. B.
,
Fearnley
,
I. M.
,
Harbour
,
M.
,
Robinson
,
A. J.
et al. 
(
2007
).
The AAA+ protein ATAD3 has displacement loop binding properties and is involved in mitochondrial nucleoid organization
.
J. Cell Biol.
176
,
141
-
146
.
Huang
,
L. S.
,
Kotha
,
S. R.
,
Avasarala
,
S.
,
VanScoyk
,
M.
,
Winn
,
R. A.
,
Pennathur
,
A.
,
Yashaswini
,
P. S.
,
Bandela
,
M.
,
Salgia
,
R.
,
Tyurina
,
Y. Y.
et al. 
(
2020
).
Lysocardiolipin acyltransferase regulates NSCLC cell proliferation and migration by modulating mitochondrial dynamics
.
J. Biol. Chem.
295
,
13393
-
13406
.
Hudson
,
G.
,
Amati-Bonneau
,
P.
,
Blakely
,
E. L.
,
Stewart
,
J. D.
,
He
,
L.
,
Schaefer
,
A. M.
,
Griffiths
,
P. G.
,
Ahlqvist
,
K.
,
Suomalainen
,
A.
,
Reynier
,
P.
et al. 
(
2008
).
Mutation of OPA1 causes dominant optic atrophy with external ophthalmoplegia, ataxia, deafness and multiple mitochondrial DNA deletions: a novel disorder of mtDNA maintenance
.
Brain
131
,
329
-
337
.
Huemer
,
M.
,
Karall
,
D.
,
Schossig
,
A.
,
Abdenur
,
J. E.
,
Al Jasmi
,
F.
,
Biagosch
,
C.
,
Distelmaier
,
F.
,
Freisinger
,
P.
,
Graham
,
B. H.
,
Haack
,
T. B.
et al. 
(
2015
).
Clinical, morphological, biochemical, imaging and outcome parameters in 21 individuals with mitochondrial maintenance defect related to FBXL4 mutations
.
J. Inherit. Metab. Dis.
38
,
905
-
914
.
Iborra
,
F. J.
,
Kimura
,
H.
and
Cook
,
P. R.
(
2004
).
The functional organization of mitochondrial genomes in human cells
.
BMC Biol.
2
,
9
.
Ilamathi
,
H. S.
,
Ouellet
,
M.
,
Sabouny
,
R.
,
Desrochers-Goyette
,
J.
,
Lines
,
M. A.
,
Pfeffer
,
G.
,
Shutt
,
T. E.
and
Germain
,
M.
(
2021
).
Mitochondrial fission is required for proper nucleoid distribution within mitochondrial networks
.
bioRxiv
,
2021.03.17.435804
.
Iwama
,
K.
,
Takaori
,
T.
,
Fukushima
,
A.
,
Tohyama
,
J.
,
Ishiyama
,
A.
,
Ohba
,
C.
,
Mitsuhashi
,
S.
,
Miyatake
,
S.
,
Takata
,
A.
,
Miyake
,
N.
et al. 
(
2018
).
Novel recessive mutations in MSTO1 cause cerebellar atrophy with pigmentary retinopathy
.
J. Hum. Genet.
63
,
263
-
270
.
Jian
,
F.
,
Chen
,
D.
,
Chen
,
L.
,
Yan
,
C.
,
Lu
,
B.
,
Zhu
,
Y.
,
Chen
,
S.
,
Shi
,
A.
,
Chan
,
D. C.
and
Song
,
Z.
(
2018
).
Sam50 regulates PINK1-parkin-mediated mitophagy by controlling PINK1 stability and mitochondrial morphology
.
Cell Rep.
23
,
2989
-
3005
.
Jokinen
,
R.
,
Marttinen
,
P.
,
Stewart
,
J. B.
,
Neil Dear
,
T.
and
Battersby
,
B. J.
(
2016
).
Tissue-specific modulation of mitochondrial DNA segregation by a defect in mitochondrial division
.
Hum. Mol. Genet.
25
,
706
-
714
.
Jones
,
B. A.
and
Fangman
,
W. L.
(
1992
).
Mitochondrial DNA maintenance in yeast requires a protein containing a region related to the GTP-binding domain of dynamin
.
Genes Dev.
6
,
380
-
389
.
Kandul
,
N. P.
,
Zhang
,
T.
,
Hay
,
B. A.
and
Guo
,
M.
(
2016
).
Selective removal of deletion-bearing mitochondrial DNA in heteroplasmic Drosophila
.
Nat. Commun.
7
,
13100
.
Karbowski
,
M.
and
Youle
,
R. J.
(
2003
).
Dynamics of mitochondrial morphology in healthy cells and during apoptosis
.
Cell Death Differ.
10
,
870
-
880
.
Kasashima
,
K.
,
Sumitani
,
M.
and
Endo
,
H.
(
2011
).
Human mitochondrial transcription factor A is required for the segregation of mitochondrial DNA in cultured cells
.
Exp. Cell Res.
317
,
210
-
220
.
Keogh
,
M. J.
and
Chinnery
,
P. F.
(
2015
).
Mitochondrial DNA mutations in neurodegeneration
.
Biochim. Biophys. Acta
1847
,
1401
-
1411
.
Khosravi
,
S.
and
Harner
,
M. E.
(
2020
).
The MICOS complex, a structural element of mitochondria with versatile functions
.
Biol. Chem.
401
,
765
-
778
.
Kishita
,
Y.
,
Shimura
,
M.
,
Kohda
,
M.
,
Akita
,
M.
,
Imai-Okazaki
,
A.
,
Yatsuka
,
Y.
,
Nakajima
,
Y.
,
Ito
,
T.
,
Ohtake
,
A.
,
Murayama
,
K.
et al. 
(
2020
).
A novel homozygous variant in MICOS13/QIL1 causes hepato-encephalopathy with mitochondrial DNA depletion syndrome
.
Mol. Genet. Genomic. Med
.
8
,
e1427
.
Kleele
,
T.
,
Rey
,
T.
,
Winter
,
J.
,
Zaganelli
,
S.
,
Mahecic
,
D.
,
Perreten Lambert
,
H.
,
Ruberto
,
F. P.
,
Nemir
,
M.
,
Wai
,
T.
,
Pedrazzini
,
T.
et al. 
(
2021
).
Distinct fission signatures predict mitochondrial degradation or biogenesis
.
Nature
593
,
435
-
439
.
Koch
,
J.
,
Feichtinger
,
R. G.
,
Freisinger
,
P.
,
Pies
,
M.
,
Schrödl
,
F.
,
Iuso
,
A.
,
Sperl
,
W.
,
Mayr
,
J. A.
,
Prokisch
,
H.
and
Haack
,
T. B.
(
2016
).
Disturbed mitochondrial and peroxisomal dynamics due to loss of MFF causes Leigh-like encephalopathy, optic atrophy and peripheral neuropathy
.
J. Med. Genet.
53
,
270
-
278
.
Kraus
,
F.
,
Roy
,
K.
,
Pucadyil
,
T. J.
and
Ryan
,
M. T.
(
2021
).
Function and regulation of the divisome for mitochondrial fission
.
Nature
590
,
57
-
66
.
Kruppa
,
A. J.
and
Buss
,
F.
(
2021
).
Motor proteins at the mitochondria-cytoskeleton interface
.
J. Cell Sci.
134
,
jcs226084
.
Kukat
,
C.
and
Larsson
,
N. G.
(
2013
).
mtDNA makes a U-turn for the mitochondrial nucleoid
.
Trends Cell Biol.
23
,
457
-
463
.
Kukat
,
C.
,
Wurm
,
C. A.
,
Spåhr
,
H.
,
Falkenberg
,
M.
,
Larsson
,
N.-G.
and
Jakobs
,
S.
(
2011
).
Super-resolution microscopy reveals that mammalian mitochondrial nucleoids have a uniform size and frequently contain a single copy of mtDNA
.
Proc. Natl Acad. Sci. USA
108
,
13534
.
Lackner
,
L. L.
(
2019
).
The expanding and unexpected functions of mitochondria contact sites
.
Trends Cell Biol.
29
,
580
-
590
.
Lewis
,
S. C.
,
Uchiyama
,
L. F.
and
Nunnari
,
J.
(
2016
).
ER-mitochondria contacts couple mtDNA synthesis with mitochondrial division in human cells
.
Science
353
,
aaf5549
.
Li
,
J.
,
Liu
,
X.
,
Wang
,
H.
,
Zhang
,
W.
,
Chan
,
D. C.
and
Shi
,
Y.
(
2012
).
Lysocardiolipin acyltransferase 1 (ALCAT1) controls mitochondrial DNA fidelity and biogenesis through modulation of MFN2 expression
.
Proc. Natl. Acad. Sci. USA
109
,
6975
-
6980
.
Li
,
H.
,
Ruan
,
Y.
,
Zhang
,
K.
,
Jian
,
F.
,
Hu
,
C.
,
Miao
,
L.
,
Gong
,
L.
,
Sun
,
L.
,
Zhang
,
X.
,
Chen
,
S.
et al. 
(
2016
).
Mic60/Mitofilin determines MICOS assembly essential for mitochondrial dynamics and mtDNA nucleoid organization
.
Cell Death Differ.
23
,
380
-
392
.
Loson
,
O. C.
,
Song
,
Z.
,
Chen
,
H.
and
Chan
,
D. C.
(
2013
).
Fis1, Mff, MiD49, and MiD51 mediate Drp1 recruitment in mitochondrial fission
.
Mol. Biol. Cell
24
,
659
-
667
.
Luévano-Martínez
,
L. A.
,
Forni
,
M. F.
,
dos Santos
,
V. T.
,
Souza-Pinto
,
N. C.
and
Kowaltowski
,
A. J.
(
2015
).
Cardiolipin is a key determinant for mtDNA stability and segregation during mitochondrial stress
.
Biochim. Biophys. Acta
1847
,
587
-
598
.
Mahajan
,
M.
,
Bharambe
,
N.
,
Shang
,
Y.
,
Lu
,
B.
,
Mandal
,
A.
,
Madan Mohan
,
P.
,
Wang
,
R.
,
Boatz
,
J. C.
,
Manuel Martinez Galvez
,
J.
,
Shnyrova
,
A. V.
et al. 
(
2021
).
NMR identification of a conserved Drp1 cardiolipin-binding motif essential for stress-induced mitochondrial fission
.
Proc. Natl. Acad. Sci. USA
118
,
e2023079118
.
Malena
,
A.
,
Loro
,
E.
,
Di Re
,
M.
,
Holt
,
I. J.
and
Vergani
,
L.
(
2009
).
Inhibition of mitochondrial fission favours mutant over wild-type mitochondrial DNA
.
Hum. Mol. Genet.
18
,
3407
-
3416
.
Margineantu
,
D. H.
,
Gregory Cox
,
W.
,
Sundell
,
L.
,
Sherwood
,
S. W.
,
Beechem
,
J. M.
and
Capaldi
,
R. A.
(
2002
).
Cell cycle dependent morphology changes and associated mitochondrial DNA redistribution in mitochondria of human cell lines
.
Mitochondrion
1
,
425
-
435
.
Martin
,
L. A.
,
Kennedy
,
B. E.
and
Karten
,
B.
(
2016
).
Mitochondrial cholesterol: mechanisms of import and effects on mitochondrial function
.
J. Bioenerg. Biomembr.
48
,
137
-
151
.
McArthur
,
K.
,
Whitehead
,
L. W.
,
Heddleston
,
J. M.
,
Li
,
L.
,
Padman
,
B. S.
,
Oorschot
,
V.
,
Geoghegan
,
N. D.
,
Chappaz
,
S.
,
Davidson
,
S.
,
San Chin
,
H.
et al. 
(
2018
).
BAK/BAX macropores facilitate mitochondrial herniation and mtDNA efflux during apoptosis
.
Science
359
,
eaao6047
.
Meeusen
,
S.
and
Nunnari
,
J.
(
2003
).
Evidence for a two membrane-spanning autonomous mitochondrial DNA replisome
.
J. Cell Biol.
163
,
503
-
510
.
Miret-Casals
,
L.
,
Sebastián
,
D.
,
Brea
,
J.
,
Rico-Leo
,
E. M.
,
Palacín
,
M.
,
Fernández-Salguero
,
P. M.
,
Loza
,
M. I.
,
Albericio
,
F.
and
Zorzano
,
A.
(
2018
).
Identification of new activators of mitochondrial fusion reveals a link between mitochondrial morphology and pyrimidine metabolism
.
Cell Chem. Biol.
25
,
268
-
278.e4
.
Mishra
,
P.
,
Carelli
,
V.
,
Manfredi
,
G.
and
Chan
,
D. C.
(
2014
).
Proteolytic cleavage of Opa1 stimulates mitochondrial inner membrane fusion and couples fusion to oxidative phosphorylation
.
Cell Metab.
19
,
630
-
641
.
Mok
,
B. Y.
,
de Moraes
,
M. H.
,
Zeng
,
J.
,
Bosch
,
D. E.
,
Kotrys
,
A. V.
,
Raguram
,
A.
,
Hsu
,
F.
,
Radey
,
M. C.
,
Peterson
,
S. B.
,
Mootha
,
V. K.
et al. 
(
2020
).
A bacterial cytidine deaminase toxin enables CRISPR-free mitochondrial base editing
.
Nature
583
,
631
-
637
.
Moore
,
A. S.
and
Holzbaur
,
E. L. F.
(
2018
).
Mitochondrial-cytoskeletal interactions: dynamic associations that facilitate network function and remodeling
.
Curr. Opin. Physiol.
3
,
94
-
100
.
Morton
,
S. U.
,
Neilan
,
E. G.
,
Peake
,
R. W. A.
,
Shi
,
J.
,
Schmitz-Abe
,
K.
,
Towne
,
M.
,
Markianos
,
K.
,
Prabhu
,
S. P.
and
Agrawal
,
P. B.
(
2017
).
Hyperammonemia as a presenting feature in two siblings with FBXL4 variants
.
JIMD Rep.
35
,
7
-
15
.
Nasca
,
A.
,
Legati
,
A.
,
Baruffini
,
E.
,
Nolli
,
C.
,
Moroni
,
I.
,
Ardissone
,
A.
,
Goffrini
,
P.
and
Ghezzi
,
D.
(
2016
).
Biallelic mutations in DNM1L are associated with a slowly progressive infantile encephalopathy
.
Hum. Mutat.
37
,
898
-
903
.
Nasca
,
A.
,
Nardecchia
,
F.
,
Commone
,
A.
,
Semeraro
,
M.
,
Legati
,
A.
,
Garavaglia
,
B.
,
Ghezzi
,
D.
and
Leuzzi
,
V.
(
2018
).
Clinical and biochemical features in a patient with mitochondrial fission factor gene alteration
.
Frontiers in Genetics
9
,
625
-
625
.
Nasca
,
A.
,
Scotton
,
C.
,
Zaharieva
,
I.
,
Neri
,
M.
,
Selvatici
,
R.
,
Magnusson
,
O. T.
,
Gal
,
A.
,
Weaver
,
D.
,
Rossi
,
R.
,
Armaroli
,
A.
et al. 
(
2017
).
Recessive mutations in MSTO1 cause mitochondrial dynamics impairment, leading to myopathy and ataxia
.
Hum. Mutat.
38
,
970
-
977
.
Nasonovs
,
A.
,
Garcia-Diaz
,
M.
and
Bogenhagen
,
D. F.
(
2021
).
A549 cells contain enlarged mitochondria with independently functional clustered mtDNA nucleoids
.
PLoS ONE
16
,
e0249047
.
Newell
,
C.
,
Sabouny
,
R.
,
Hittel
,
D. S.
,
Shutt
,
T. E.
,
Khan
,
A.
,
Klein
,
M. S.
and
Shearer
,
J.
(
2018
).
Mesenchymal stem cells shift mitochondrial dynamics and enhance oxidative phosphorylation in recipient cells
.
Front. Physiol.
9
,
1572
.
Nicholls
,
T. J.
and
Gustafsson
,
C. M.
(
2018
).
Separating and segregating the human mitochondrial genome
.
Trends Biochem. Sci.
43
,
869
-
881
.
Nissanka
,
N.
,
Bacman
,
S. R.
,
Plastini
,
M. J.
and
Moraes
,
C. T.
(
2018
).
The mitochondrial DNA polymerase gamma degrades linear DNA fragments precluding the formation of deletions
.
Nat. Commun.
9
,
2491
.
Nissanka
,
N.
,
Minczuk
,
M.
and
Moraes
,
C. T.
(
2019
).
Mechanisms of mitochondrial DNA deletion formation
.
Trends Genet.
35
,
235
-
244
.
Nunnari
,
J.
and
Suomalainen
,
A.
(
2012
).
Mitochondria: in sickness and in health
.
Cell
148
,
1145
-
1159
.
Onishi
,
M.
,
Yamano
,
K.
,
Sato
,
M.
,
Matsuda
,
N.
and
Okamoto
,
K.
(
2021
).
Molecular mechanisms and physiological functions of mitophagy
.
EMBO J.
40
,
e104705
.
Parone
,
P. A.
,
Da Cruz
,
S.
,
Tondera
,
D.
,
Mattenberger
,
Y.
,
James
,
D. I.
,
Maechler
,
P.
,
Barja
,
F.
and
Martinou
,
J. C.
(
2008
).
Preventing mitochondrial fission impairs mitochondrial function and leads to loss of mitochondrial DNA
.
PLoS ONE
3
,
e3257
.
Peeva
,
V.
,
Blei
,
D.
,
Trombly
,
G.
,
Corsi
,
S.
,
Szukszto
,
M. J.
,
Rebelo-Guiomar
,
P.
,
Gammage
,
P. A.
,
Kudin
,
A. P.
,
Becker
,
C.
,
Altmuller
,
J.
et al. 
(
2018
).
Linear mitochondrial DNA is rapidly degraded by components of the replication machinery
.
Nat. Commun.
9
,
1727
.
Peralta
,
S.
,
Goffart
,
S.
,
Williams
,
S. L.
,
Diaz
,
F.
,
Garcia
,
S.
,
Nissanka
,
N.
,
Area-Gomez
,
E.
,
Pohjoismäki
,
J.
and
Moraes
,
C. T.
(
2018
).
ATAD3 controls mitochondrial cristae structure in mouse muscle, influencing mtDNA replication and cholesterol levels
.
J. Cell Sci.
131
,
jcs217075
.
Phillips
,
A. F.
,
Millet
,
A. R.
,
Tigano
,
M.
,
Dubois
,
S. M.
,
Crimmins
,
H.
,
Babin
,
L.
,
Charpentier
,
M.
,
Piganeau
,
M.
,
Brunet
,
E.
and
Sfeir
,
A.
(
2017
).
Single-molecule analysis of mtDNA replication uncovers the basis of the common deletion
.
Mol. Cell
65
,
527
-
538.e6
.
Pickles
,
S.
,
Vigié
,
P.
and
Youle
,
R. J.
(
2018
).
Mitophagy and quality control mechanisms in mitochondrial maintenance
.
Curr. Biol.
28
,
R170
-
R185
.
Pickrell
,
A. M.
,
Huang
,
C. H.
,
Kennedy
,
S. R.
,
Ordureau
,
A.
,
Sideris
,
D. P.
,
Hoekstra
,
J. G.
,
Harper
,
J. W.
and
Youle
,
R. J.
(
2015
).
Endogenous parkin preserves dopaminergic substantia nigral neurons following mitochondrial DNA mutagenic stress
.
Neuron
87
,
371
-
381
.
Prole
,
D. L.
,
Chinnery
,
P. F.
and
Jones
,
N. S.
(
2020
).
Visualizing, quantifying and manipulating mitochondrial DNA in vivo
.
J. Biol. Chem.
295
,
17588
-
17601
.
Qin
,
J.
,
Guo
,
Y.
,
Xue
,
B.
,
Shi
,
P.
,
Chen
,
Y.
,
Su
,
Q. P.
,
Hao
,
H.
,
Zhao
,
S.
,
Wu
,
C.
,
Yu
,
L.
et al. 
(
2020
).
ER-mitochondria contacts promote mtDNA nucleoids active transportation via mitochondrial dynamic tubulation
.
Nat. Commun.
11
,
4471
.
Rambold
,
A. S.
,
Kostelecky
,
B.
,
Elia
,
N.
and
Lippincott-Schwartz
,
J.
(
2011
).
Tubular network formation protects mitochondria from autophagosomal degradation during nutrient starvation
.
Proc. Natl. Acad. Sci. USA
108
,
10190
-
10195
.
Rapaport
,
D.
,
Brunner
,
M.
,
Neupert
,
W.
and
Westermann
,
B.
(
1998
).
Fzo1p is a mitochondrial outer membrane protein essential for the biogenesis of functional mitochondria in Saccharomyces cerevisiae
.
J. Biol. Chem.
273
,
20150
-
20155
.
Renaldo
,
F.
,
Amati-Bonneau
,
P.
,
Slama
,
A.
,
Romana
,
C.
,
Forin
,
V.
,
Doummar
,
D.
,
Barnerias
,
C.
,
Bursztyn
,
J.
,
Mayer
,
M.
,
Khouri
,
N.
et al. 
(
2012
).
MFN2, a new gene responsible for mitochondrial DNA depletion
.
Brain
135
,
e223
.
Reyes
,
A.
,
He
,
J.
,
Mao
,
C. C.
,
Bailey
,
L. J.
,
Di Re
,
M.
,
Sembongi
,
H.
,
Kazak
,
L.
,
Dzionek
,
K.
,
Holmes
,
J. B.
,
Cluett
,
T. J.
et al. 
(
2011
).
Actin and myosin contribute to mammalian mitochondrial DNA maintenance
.
Nucleic Acids Res.
39
,
5098
-
5108
.
Riley
,
J. S.
,
Quarato
,
G.
,
Cloix
,
C.
,
Lopez
,
J.
,
O'Prey
,
J.
,
Pearson
,
M.
,
Chapman
,
J.
,
Sesaki
,
H.
,
Carlin
,
L. M.
,
Passos
,
J. F.
et al. 
(
2018
).
Mitochondrial inner membrane permeabilisation enables mtDNA release during apoptosis
.
EMBO J.
37
,
e99238
.
Rouzier
,
C.
,
Bannwarth
,
S.
,
Chaussenot
,
A.
,
Chevrollier
,
A.
,
Verschueren
,
A.
,
Bonello-Palot
,
N.
,
Fragaki
,
K.
,
Cano
,
A.
,
Pouget
,
J.
,
Pellissier
,
J. F.
et al. 
(
2012
).
The MFN2 gene is responsible for mitochondrial DNA instability and optic atrophy ‘plus’ phenotype
.
Brain
135
,
23
-
34
.
Ryu
,
C.
,
Sun
,
H.
,
Gulati
,
M.
,
Herazo-Maya
,
J. D.
,
Chen
,
Y.
,
Osafo-Addo
,
A.
,
Brandsdorfer
,
C.
,
Winkler
,
J.
,
Blaul
,
C.
,
Faunce
,
J.
et al. 
(
2017
).
Extracellular mitochondrial DNA is generated by fibroblasts and predicts death in idiopathic pulmonary fibrosis
.
Am. J. Respir. Crit. Care. Med.
196
,
1571
-
1581
.
Sabouny
,
R.
and
Shutt
,
T. E.
(
2020
).
Reciprocal regulation of mitochondrial fission and fusion
.
Trends Biochem. Sci.
45
,
564
-
577
.
Sabouny
,
R.
,
Wong
,
R.
,
Lee-Glover
,
L.
,
Greenway
,
S. C.
,
Sinasac
,
D. S.
,
Khan
,
A.
and
Shutt
,
T. E.
(
2019
).
Characterization of the C584R variant in the mtDNA depletion syndrome gene FBXL4, reveals a novel role for FBXL4 as a regulator of mitochondrial fusion
.
Biochim. Biophys. Acta Mol. Basis Dis.
1865
,
165536
.
Schon
,
E. A.
and
Gilkerson
,
R. W.
(
2010
).
Functional complementation of mitochondrial DNAs: mobilizing mitochondrial genetics against dysfunction
.
Biochim. Biophys. Acta
1800
,
245
-
249
.
Shu
,
L.
,
Hu
,
C.
,
Xu
,
M.
,
Yu
,
J.
,
He
,
H.
,
Lin
,
J.
,
Sha
,
H.
,
Lu
,
B.
,
Engelender
,
S.
,
Guan
,
M.
et al. 
(
2021
).
ATAD3B is a mitophagy receptor mediating clearance of oxidative stress-induced damaged mitochondrial DNA
.
EMBO J.
40
,
e106283
.
Silva Ramos
,
E.
,
Motori
,
E.
,
Bruser
,
C.
,
Kuhl
,
I.
,
Yeroslaviz
,
A.
,
Ruzzenente
,
B.
,
Kauppila
,
J. H. K.
,
Busch
,
J. D.
,
Hultenby
,
K.
,
Habermann
,
B. H.
et al. 
(
2019
).
Mitochondrial fusion is required for regulation of mitochondrial DNA replication
.
PLoS Genet.
15
,
e1008085
.
Sliter
,
D. A.
,
Martinez
,
J.
,
Hao
,
L.
,
Chen
,
X.
,
Sun
,
N.
,
Fischer
,
T. D.
,
Burman
,
J. L.
,
Li
,
Y.
,
Zhang
,
Z.
,
Narendra
,
D. P.
et al. 
(
2018
).
Parkin and PINK1 mitigate STING-induced inflammation
.
Nature
561
,
258
-
262
.
Spiegel
,
R.
,
Saada
,
A.
,
Flannery
,
P. J.
,
Burte
,
F.
,
Soiferman
,
D.
,
Khayat
,
M.
,
Eisner
,
V.
,
Vladovski
,
E.
,
Taylor
,
R. W.
,
Bindoff
,
L. A.
et al. 
(
2016
).
Fatal infantile mitochondrial encephalomyopathy, hypertrophic cardiomyopathy and optic atrophy associated with a homozygous OPA1 mutation
.
J. Med. Genet.
53
,
127
-
131
.
Stewart
,
J. B.
and
Chinnery
,
P. F.
(
2015
).
The dynamics of mitochondrial DNA heteroplasmy: implications for human health and disease
.
Nat. Rev. Genet.
16
,
530
-
542
.
Suen
,
D.-F.
,
Narendra
,
D. P.
,
Tanaka
,
A.
,
Manfredi
,
G.
and
Youle
,
R. J.
(
2010
).
Parkin overexpression selects against a deleterious mtDNA mutation in heteroplasmic cybrid cells
.
Proc. Natl Acad. Sci. USA
107
,
11835
.
Sugiura
,
A.
,
McLelland
,
G. L.
,
Fon
,
E. A.
and
McBride
,
H. M.
(
2014
).
A new pathway for mitochondrial quality control: mitochondrial-derived vesicles
.
EMBO J.
33
,
2142
-
2156
.
Tauber
,
J.
,
Dlasková
,
A.
,
Šantorová
,
J.
,
Smolková
,
K.
,
Alán
,
L.
,
Špaček
,
T.
,
Plecitá-Hlavatá
,
L.
,
Jabůrek
,
M.
and
Ježek
,
P.
(
2013
).
Distribution of mitochondrial nucleoids upon mitochondrial network fragmentation and network reintegration in HEPG2 cells
.
Int. J. Biochem. Cell Biol.
45
,
593
-
603
.
Twig
,
G.
,
Elorza
,
A.
,
Molina
,
A. J.
,
Mohamed
,
H.
,
Wikstrom
,
J. D.
,
Walzer
,
G.
,
Stiles
,
L.
,
Haigh
,
S. E.
,
Katz
,
S.
,
Las
,
G.
et al. 
(
2008
).
Fission and selective fusion govern mitochondrial segregation and elimination by autophagy
.
EMBO J.
27
,
433
-
446
.
Valenci
,
I.
,
Yonai
,
L.
,
Bar-Yaacov
,
D.
,
Mishmar
,
D.
and
Ben-Zvi
,
A.
(
2015
).
Parkin modulates heteroplasmy of truncated mtDNA in Caenorhabditis elegans
.
Mitochondrion
20
,
64
-
70
.
van der Laan
,
M.
,
Bohnert
,
M.
,
Wiedemann
,
N.
and
Pfanner
,
N.
(
2012
).
Role of MINOS in mitochondrial membrane architecture and biogenesis
.
Trends Cell Biol.
22
,
185
-
192
.
Vanstone
,
J. R.
,
Smith
,
A. M.
,
McBride
,
S.
,
Naas
,
T.
,
Holcik
,
M.
,
Antoun
,
G.
,
Harper
,
M. E.
,
Michaud
,
J.
,
Sell
,
E.
,
Chakraborty
,
P.
et al. 
(
2016
).
DNM1L-related mitochondrial fission defect presenting as refractory epilepsy
.
Eur. J. Hum. Genet.
24
,
1084
-
1088
.
Vidoni
,
S.
,
Zanna
,
C.
,
Rugolo
,
M.
,
Sarzi
,
E.
and
Lenaers
,
G.
(
2013
).
Why mitochondria must fuse to maintain their genome integrity
.
Antioxid Redox Signal.
19
,
379
-
388
.
Vielhaber
,
S.
,
Debska-Vielhaber
,
G.
,
Peeva
,
V.
,
Schoeler
,
S.
,
Kudin
,
A. P.
,
Minin
,
I.
,
Schreiber
,
S.
,
Dengler
,
R.
,
Kollewe
,
K.
,
Zuschratter
,
W.
et al. 
(
2013
).
Mitofusin 2 mutations affect mitochondrial function by mitochondrial DNA depletion
.
Acta Neuropathol.
125
,
245
-
256
.
Viscomi
,
C.
and
Zeviani
,
M.
(
2017
).
MtDNA-maintenance defects: syndromes and genes
.
J. Inherit. Metab. Dis.
40
,
587
-
599
.
Waterham
,
H. R.
,
Koster
,
J.
,
van Roermund
,
C. W.
,
Mooyer
,
P. A.
,
Wanders
,
R. J.
and
Leonard
,
J. V.
(
2007
).
A lethal defect of mitochondrial and peroxisomal fission
.
N. Engl. J. Med.
356
,
1736
-
1741
.
West
,
A. P.
,
Khoury-Hanold
,
W.
,
Staron
,
M.
,
Tal
,
M. C.
,
Pineda
,
C. M.
,
Lang
,
S. M.
,
Bestwick
,
M.
,
Duguay
,
B. A.
,
Raimundo
,
N.
,
MacDuff
,
D. A.
et al. 
(
2015
).
Mitochondrial DNA stress primes the antiviral innate immune response
.
Nature
520
,
553
-
557
.
Whitley
,
B. N.
,
Engelhart
,
E. A.
and
Hoppins
,
S.
(
2019
).
Mitochondrial dynamics and their potential as a therapeutic target
.
Mitochondrion
49
,
269
-
283
.
Wiehe
,
R. S.
,
Gole
,
B.
,
Chatre
,
L.
,
Walther
,
P.
,
Calzia
,
E.
,
Ricchetti
,
M.
and
Wiesmüller
,
L.
(
2018
).
Endonuclease G promotes mitochondrial genome cleavage and replication
.
Oncotarget
9
,
18309
-
18326
.
Yang
,
L.
,
Tang
,
H.
,
Lin
,
X.
,
Wu
,
Y.
,
Zeng
,
S.
,
Pan
,
Y.
,
Li
,
Y.
,
Xiang
,
G.
,
Lin
,
Y. F.
,
Zhuang
,
S. M.
et al. 
(
2020
).
OPA1-exon4b binds to mtDNA D-loop for transcriptional and metabolic modulation, independent of mitochondrial fusion
.
Front. Cell Dev. Biol.
8
,
180
.
Yuan
,
Y.
,
Ju
,
Y. S.
,
Kim
,
Y.
,
Li
,
J.
,
Wang
,
Y.
,
Yoon
,
C. J.
,
Yang
,
Y.
,
Martincorena
,
I.
,
Creighton
,
C. J.
,
Weinstein
,
J. N.
et al. 
(
2020
).
Comprehensive molecular characterization of mitochondrial genomes in human cancers
.
Nat. Genet.
52
,
342
-
352
.
Zekonyte
,
U.
,
Bacman
,
S. R.
and
Moraes
,
C. T.
(
2020
).
DNA-editing enzymes as potential treatments for heteroplasmic mtDNA diseases
.
J. Intern. Med.
287
,
685
-
697
.
Zuchner
,
S.
,
Mersiyanova
,
I. V.
,
Muglia
,
M.
,
Bissar-Tadmouri
,
N.
,
Rochelle
,
J.
,
Dadali
,
E. L.
,
Zappia
,
M.
,
Nelis
,
E.
,
Patitucci
,
A.
,
Senderek
,
J.
et al. 
(
2004
).
Mutations in the mitochondrial GTPase mitofusin 2 cause charcot-marie-tooth neuropathy type 2A
.
Nat. Genet.
36
,
449
-
451
.

Competing interests

The authors declare no competing or financial interests.