Next Article in Journal
Bone Marker Proteins in Women With and Without Polycystic Ovary Syndrome
Previous Article in Journal
Molecular and Clinical Considerations for Anesthesia in the Aging Brain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial DNA Replication and Disease: A Historical Perspective on Molecular Insights and Therapeutic Advances

Genome Integrity and Structural Biology Laboratory, Mitochondrial DNA Replication Group, National Institute of Environmental Health Sciences, National Institutes of Health, Research Triangle Park, Durham, NC 27709, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2025, 26(21), 10275; https://doi.org/10.3390/ijms262110275
Submission received: 30 September 2025 / Revised: 16 October 2025 / Accepted: 17 October 2025 / Published: 22 October 2025
(This article belongs to the Special Issue Eukaryotic DNA Replication—from Bench to Bedside)

Abstract

Mitochondria are vital for cellular energy production, as these organelles generate most of the cellular energy required for various metabolic processes. Mitochondria contain their own circular DNA, which is present in multiple copies and is exclusively maternally inherited. Cellular energy in the form of adenosine 5′-triphosphate is produced via oxidative phosphorylation and involves the coordinated expression of genes encoded by both the nuclear and mitochondrial genomes. Mitochondrial DNA itself is replicated by a dedicated set of nuclear-encoded proteins composed of the DNA polymerase gamma, the Twinkle helicase, the mitochondrial single-stranded DNA binding protein, as well as several accessory factors. Mutations in these genes, as well as in the genes involved in nucleotide metabolism, are associated with a spectrum of mitochondrial disorders that can affect individuals from infancy to old age. Additionally, mitochondrial disease can arise as a result of point mutations, deletions, or depletion in the mitochondrial DNA or in genes involved in mitochondrial transcription, replication, maintenance, and repair. Although a cure for mitochondrial diseases is currently elusive, several treatment options have been explored. In this review, we explore the molecular insights of the core mitochondrial replisome proteins that have aided our understanding of mitochondrial diseases and influenced current therapies.

1. Introduction

Mitochondria are intracellular organelles present in nearly all eukaryotes. Known mainly for their role in cellular energy production, mitochondria harbor the electron transport chain (ETC), which generates more than 90% of adenosine 5′-triphosphate (ATP) through oxidative phosphorylation (OXPHOS) [1,2,3]. Mitochondria are the major source of endogenous reactive oxygen species (ROS) and are the host of many fundamental biochemical pathways, such as the tricarboxylic acid (TCA) cycle, fatty acid oxidation, as well as certain parts of the urea cycle [4]. In addition to cellular energy production, mitochondria fulfill critical roles in many key biological processes, such as apoptosis, ion homeostasis, iron-sulfur cluster biogenesis, cellular signaling, and immune response regulation [5].
A unique feature of mitochondria is that these organelles harbor their own circular, double-stranded genome, which is exclusively maternally inherited and distinct from the nuclear genome. Human mitochondrial DNA (mtDNA) is 16,569 base pairs (bp) in size and encodes 37 genes, all of which are directly or indirectly involved in ATP production (Figure 1A). Of the 37 genes encoded by mtDNA, 13 genes encode polypeptides that make up essential subunits of the electron transport chain (ETC) (complexes I, III, IV, and V) and are required for oxidative phosphorylation (OXPHOS). The remaining 24 genes encode 22 transfer RNAs and 2 ribosomal RNAs necessary for the mitochondrial translation of the 13 polypeptides [1] (Figure 1A). The OXPHOS system itself consists of five multiprotein complexes and involves more than 90 proteins, encoded by both the nuclear and mitochondrial genomes [6]. The nuclear genome is responsible for encoding most mitochondrial proteins, including the subunits of the ETC complexes, as well as proteins involved in mtDNA maintenance, transcription, replication, and copy number regulation [7]. Although mtDNA encodes only a small proportion of proteins involved in OXPHOS compared to the nuclear genome, these genes are critical for cellular energy production [8,9]. Given that mitochondria are under dual genetic control of both the nuclear and mitochondrial genomes, mutations in either genome may impair ETC function and therefore be detrimental to human health [10,11]. Defects in mitochondrial gene expression significantly impair OXPHOS capacity and lead to a wide spectrum of diseases, which will be further discussed in this review [10,12].
Similar to bacterial chromosomal DNA, mtDNA is organized into nucleoprotein complexes, known as nucleoids, which are located in the inner mitochondrial matrix. Within these nucleoids, mtDNA is compacted by different proteins, with the most abundant being transcription factor A, mitochondrial (TFAM) [13,14]. Human mtDNA consists of two strands, termed heavy strand (H-strand) and light strand (L-strand) due to their strand bias in nucleotide composition. The H-strand has a higher content of guanine and thymine nucleotides compared to the L-strand, allowing these strands to be separated by density gradient centrifugation in alkaline cesium chloride [15,16]. MtDNA also contains a noncoding region (NCR) of 1121 nucleotides that harbors the promoter regions for transcription initiation for both the H-strand and the L-strand, as well as the origin of replication for the H-strand (OH) [17] (Figure 1A). Within the NCR, a linear DNA strand exists, referred to as 7S DNA based upon its sedimentation characteristics. The 7S DNA is approximately 650 nucleotides in length and forms a displacement loop (D-loop). Although the exact function of the D-loop is currently unknown, several studies have proposed that the D-loop might play a role in nucleoid organization and may be involved in the regulation of mtDNA replication [18].
Mammalian mtDNA replication differs from nuclear DNA replication in both mechanism and proteins involved [19,20,21]. The mitochondrial replisome consists of a dedicated set of nuclear encoded proteins, with the core components being the mitochondrial RNA polymerase (POLRMT), the replicative mtDNA helicase Twinkle, DNA polymerase γ (pol γ), and the mitochondrial single-stranded DNA-binding protein (mtSSB). Various models of mtDNA replication have been proposed in different cell types and organisms; however, one of the most widely accepted models for mammalian mtDNA replication is the strand-displacement model, which was first proposed in 1972 [22]. According to this model, both strands of the mtDNA are replicated asynchronously and continuously, without the formation of Okazaki fragments [22,23,24,25,26,27] (Figure 1B). Since human mitochondria lack a dedicated DNA primase, POLRMT is responsible for synthesizing short RNA primers required for initiating mtDNA replication. MtDNA replication is first initiated at the H-strand origin of replication (OH) located within the NCR [28,29,30]. To initiate the synthesis of the nascent H-strand, the mitochondrial Twinkle helicase first unwinds the double-stranded mtDNA ahead of pol γ, and the displaced parental H-strand is stabilized by mtSSB. H-strand synthesis is performed by the addition of nucleotides to the 3′ end of an existing RNA primer by pol γ [23,24]. Once two-thirds of the nascent H-strand is synthesized, the L-strand origin of replication (OL) in the parental H-strand becomes exposed. At this site, a stem-loop structure is formed which is used by POLRMT for initiating primer synthesis for nascent L-strand synthesis. Once primer synthesis has proceeded for approximately 25 nucleotides, POLRMT is replaced by pol γ, allowing for the synthesis of the nascent L-strand in the opposite direction [23,25,31]. At this stage of mtDNA replication, both nascent H- and L-strands are synthesized simultaneously, until two new daughter molecules of mtDNA are formed. After completion of mtDNA replication, the newly formed mtDNA daughter molecules are separated by topoisomerase 3α (TOP3A) [32,33]. Although other mechanisms for mtDNA replication have been proposed, the strand-displacement model of mtDNA replication has received tremendous support, and the validity of this model has been highlighted by findings from biochemical analyses, reconstitution experiments, single-molecule analysis, and in vivo occupancy binding patterns of mitochondrial replisome proteins [19,20,21,22,24,26].
Unlike the nuclear genome, the mitochondrial genome exists in multiple copies per cell, ranging from 100 to 10,000 depending on cellular energy demands, cell type, and differentiation status [34]. Moreover, each cell, as well as each mitochondrion, can contain a mixture of wild-type and mutant mtDNA, a condition referred to as heteroplasmy. The level of heteroplasmy can change throughout development and aging and represents a critical determinant of mitochondrial dysfunction and disease severity [35,36,37,38]. The multicopy nature of mtDNA allows the tolerance of many pathological mutations until both a cell- and mutation-specific threshold is exceeded. This threshold generally varies from 60% to 90% heteroplasmy depending on the specific mutation and cellular context. Beyond this threshold, OXPHOS becomes significantly impaired, resulting in an overall disruption of cellular energy production and cellular dysfunction [39,40].
Despite the high copy number of mtDNA, which naturally serves as a buffer against endogenous and exogenous perturbations, disruption of mtDNA integrity results in a broad spectrum of human diseases, commonly known as mitochondrial diseases [41,42,43,44]. Mitochondrial diseases involving mtDNA defects are characterized by mitochondrial DNA depletion syndrome (MDS), or multiple deletions combined with or without multiple point mutations. MDS is a group of autosomal recessive disorders that manifest with a significant reduction in the mtDNA content within affected organs and tissues [45]. These disorders typically present in early childhood and include Alpers–Huttenlocher syndrome, hepatocerebral syndromes, myocerebrohepatopathy spectrum disorders, fatal myopathies, and Leigh syndrome. In contrast, mtDNA deletions and point mutations are associated with disorders such as progressive external ophthalmoplegia (PEO), ataxia–neuropathy syndromes, as well as rare disorders of TCA cycle abnormalities. Treatment of mitochondrial diseases is extremely challenging given the involvement of multiple organs, disease severity, age of onset, and broad range of symptoms [46,47,48,49,50,51]. Although effective treatments for mitochondrial diseases remain elusive, current treatment strategies involve physical therapy to improve ETC activity combined with dietary changes and providing patients with supplements that enhance ETC function and mitochondrial biogenesis [52,53,54]. Over the past decade, novel therapeutic approaches have emerged and will be discussed in detail later in this review.
Mitochondrial diseases are caused by genetic defects in either the mtDNA or in the nuclear genes of proteins that function in the mitochondria. Genetic alterations can also occur in genes encoding proteins involved in maintaining a balanced mitochondrial nucleotide pool, which cumulatively result in compromised mtDNA integrity, leading to mitochondrial dysfunction [51]. Various studies have estimated that approximately one in every 5000 children and adults will develop a mitochondrial disease sometime in their lifetime [55,56,57]. Mitochondrial diseases are clinically heterogeneous and typically affect multiple organs, in particular those with high energy demands such as the brain, heart, extraocular muscles, kidney, and liver. Adequate amounts of mtDNA are required for the synthesis of essential subunits of the ETC complexes, and thus for cellular energy production. Unsurprisingly, mutations, deletions, and/or depletion of the mtDNA result in cellular dysfunction due to insufficient ATP production to meet the metabolic demands of the affected tissues [54,58,59].
Since there is currently no cure for mitochondrial diseases, biochemical and molecular characterization of mitochondrial replisome proteins remains valuable for elucidating the clinical consequences of specific amino acid variations. In this review, we aim to provide a historical perspective on the molecular insights of the key mitochondrial replisome proteins whose genes have been identified as mitochondrial disease alleles. We focus solely on those genes that are directly involved in mtDNA maintenance, more specifically, the genes that encode the minimal replication machinery. Furthermore, we highlight current therapeutic advances that offer promising treatment strategies for mitochondrial diseases.

2. DNA Polymerase γ

Pol γ is the sole replicative polymerase in mitochondria. The subunit structure of eukaryotic pol γ varies across species. In humans and throughout mammals, pol γ exists as a heterotrimer, consisting of a catalytic subunit bound to a dimeric accessory subunit (Figure 2) [60]. The human catalytic subunit is encoded by the POLG gene and is also referred to in the literature as the p140 subunit, based on its molecular weight; as PolG when discussing the protein; and by the aliases POLγA and POLG1 [61]. The human accessory subunit is encoded by the POLG2 gene and is also referred to as the p55 subunit, based on its molecular weight; as PolG2 when discussing the protein; and by the alias POLγB [62]. In Drosophila, pol γ forms a heterodimer, with only one accessory subunit attached to the catalytic core [63]. In unicellular yeasts like Saccharomyces cerevisiae and Schizosaccharomyces pombe, as well as in Caenorhabditis elegans, pol γ functions as a single catalytic subunit. This catalytic subunit ranges from 143 kDa in S. cerevisiae to 125 kDa in S. pombe, the latter being among the smallest known [60].

3. Catalytic Subunit, PolG

The catalytic subunit of human pol γ is encoded by the POLG gene on chromosome 15q25 and comprises 1239 amino acids with a molecular weight of 140 kDa. Deletion of either Polg or Polg2 in mice results in embryonic lethality by day 8.5, coinciding with a complete loss of mtDNA [64,65]. The p140 subunit exhibits three enzymatic activities: DNA polymerase, 3′→5′ exonuclease, and 5′ dRP lyase. The exonuclease domain is located at the N-terminus, whereas the polymerase active site lies at the C-terminus, separated by a linker region that facilitates interaction with the accessory subunit.
The 5′ dRP lyase activity is implicated in base excision repair, specifically removing the 5′-deoxyribose phosphate left behind by AP endonuclease activity [66]. Pol γ belongs to the family A group of DNA polymerases and shares structural and biochemical similarities with bacteriophage T7 DNA polymerase [61]. The crystal structure of the pol γ–DNA complex has been resolved as a heterotrimer at 3.2 Å, and with bound nucleotides (ddCTP and dCTP) at 3.3 Å resolution [67,68] (Figure 2).

3.1. Fidelity of mtDNA Replication

The human pol γ maintains high base substitution fidelity through stringent nucleotide selectivity and intrinsic 3′→5′ exonucleolytic proofreading activity [69]. Pol γ incorporates nucleotides with high accuracy in non-iterated and short repetitive sequences, where a misinsertion occurs only once per ~500,000 nucleotides synthesized [69]. However, its fidelity decreases in homopolymeric tracts exceeding four nucleotides in length, resulting in an increased risk of frameshift errors. This suggests that such regions in mtDNA are particularly susceptible to replication errors by pol γ. The PolG2 accessory subunit, which enhances pol γ processivity, paradoxically reduces both frameshift and base substitution fidelity by facilitating extension of mismatched primer termini [69].
Multiple lines of evidence suggest that the primary driver of mitochondrial DNA mutagenesis is spontaneous errors of DNA replication [70,71], most likely due to the absence of mismatch repair in mammalian mitochondria [72]. Analysis of age-specific mtDNA sequences reveals mutation signatures more consistent with polymerase errors as opposed to oxidative lesion-driven mutations [73,74]. Furthermore, mtDNA appears to be resistant to mutagenesis following exposure to exogenous DNA-damaging agents [75,76].
Biochemical studies have demonstrated that substituting Asp198 and Glu200 with alanine in the ExoI motif of the exonuclease domain of pol γ abolishes 3′→5′ exonuclease activity [77]. This proofreading function normally enhances replication fidelity by at least 20-fold [69]. Expression of exonuclease-deficient pol γ in human cells leads to accumulation of mtDNA point mutations [78]. In pancreatic islet cells, this defect impairs glucose tolerance, induces diabetes, and increases β-cell apoptosis [79,80]. In mice, cardiac-specific overexpression of exonuclease-deficient pol γ results in a greater than 23-fold increase in mtDNA point mutations, detectable mtDNA deletions, elevated apoptosis, and the development of cardiomyopathy [81,82,83].
Previous studies have associated mtDNA mutation accumulation with aging. The strongest evidence linking mtDNA mutations to aging came from “mutator mice,” generated independently by two groups, which carried homozygous mutations disrupting the exonuclease function of pol γ [84,85]. These mice exhibited premature aging between 6 and 9 months of age, including hair graying and loss, hearing impairment, spinal curvature, cardiomegaly, reduced body weight, and decreased bone density. Mutation frequency in these exonuclease-deficient mice has been quantified and demonstrates significantly elevated mutation rates in their mitochondrial genome [86,87]. Notably, asymptomatic heterozygotes exhibited a mutation frequency far higher than aged wild-type mice, indicating that elevated mutagenesis alone is insufficient to produce aging phenotypes [86]. In addition to point mutations, homozygous mutator mice show a 90-fold increase in mtDNA deletions compared with wild-type or heterozygous controls [88]. While wild-type deletions typically occurred at direct repeats ≥6 nucleotides, deletions in exonuclease-deficient mice arose independently of repeats, implicating them in aging pathology. More recently, mutator mice have also shown nuclear genomic instability, likely resulting from nucleotide pool imbalances and compensatory sequestration of nucleotides within mitochondria [89].

3.2. POLG-Related Diseases

Mutations in POLG represent the most common cause of inherited mitochondrial disease [90,91]. To date, more than 300 pathogenic mutations have been identified in the POLG gene (https://tools.niehs.nih.gov/polg/, accessed on 25 September 2025). These pathogenic POLG variants are now known to cause a broad and overlapping spectrum of clinical phenotypes—many of which were defined clinically before their genetic basis was understood. Most individuals exhibit a subset, rather than the full set, of features associated with a given phenotype. The age of onset varies widely, ranging from infancy to late adulthood. The clinical spectrum includes the following (reviewed in [92,93]):
  • Alpers–Huttenlocher syndrome: A severe childhood-onset encephalopathy marked by intractable epilepsy and progressive liver failure.
  • Childhood myocerebrohepatopathy spectrum: Presents during the first months to three years of life with developmental delay or regression, lactic acidosis, and myopathy. Additional features may include liver failure, renal tubular acidosis, pancreatitis, cyclic vomiting, and sensorineural hearing loss.
  • Myoclonic epilepsy myopathy sensory ataxia (MEMSA): A group of disorders involving epilepsy, myopathy, and ataxia without ophthalmoplegia. This category includes what was previously described as spinocerebellar ataxia with epilepsy.
  • Ataxia–neuropathy spectrum: Encompasses mitochondrial recessive ataxia syndrome (MIRAS) and sensory ataxia with neuropathy, dysarthria, and ophthalmoplegia.
  • Autosomal recessive progressive external ophthalmoplegia (arPEO): Characterized by progressive weakness of the extraocular muscles, leading to ptosis and ophthalmoparesis. Although initially isolated to eye movement, many individuals later develop additional systemic symptoms of POLG-related disease.
  • Autosomal dominant progressive external ophthalmoplegia (adPEO): Typically involves generalized myopathy along with varying degrees of sensorineural hearing loss, axonal neuropathy, ataxia, depression, Parkinsonism, hypogonadism, and cataracts.
Early-onset diseases associated with POLG mutations such as Alpers–Huttenlocher syndrome cause depletion of mtDNA, whereas the adult-onset diseases such as ataxia and PEO are associated with multiple mtDNA deletions. Symptoms from depletion usually do not occur until >50% of the mtDNA is depleted while adult-onset symptoms from multiple DNA deletions are associated with deletions in >60% of the mtDNA genomes [94].
Most POLG mutations are linked to autosomal recessive progressive external ophthalmoplegia (arPEO), ataxia, or Alpers syndrome, and are frequently observed as compound heterozygous recessive variants. Most patients with recessive POLG mutations carry at least one of the common mutations: A467T, W748S, G848S, or the T251I-P587L allelic pair. Results from biochemical studies show that the A467T PolG disease variant exhibits only ~4% of wild-type polymerase activity, with minimal impact on exonuclease function [95]. Moreover, the A467T PolG variant fails to interact with its accessory subunit, p55, which is essential for highly processive DNA synthesis [95]. Despite its functional defects, A467T PolG is present in approximately 0.6% of the Belgian population [96].
The W748S mutation in POLG reduces polymerase catalytic activity and severely impairs DNA binding [97]. It is usually found in cis with the E1143G POLG polymorphism, which partially compensates for its deleterious effects and appears to modulate disease severity [97]. The G848S PolG variant, which is the third most common POLG mutation, is located in the thumb subdomain of the polymerase active site and displays less than 1% of normal polymerase activity due to defective DNA binding [98]. The T251I-P587L PolG substitutions, typically occurring in cis and present in up to 1% of the Italian population, are frequently found in PEO [99]. Individually, these mutations reduce polymerase activity by approximately 30%; however, when combined, they synergistically impair function to about 5% of normal, through decreased enzyme stability, reduced DNA binding affinity, and diminished catalytic efficiency [100].
Dominant POLG mutations are associated with adult-onset PEO and are almost exclusively located in the polymerase domain. Two key substitutions, R943H and Y955C PolG, directly alter residues that interact with incoming deoxynucleotide triphosphates (dNTPs) [101]. These variants retain less than 1% of wild-type polymerase activity and exhibit dramatically reduced processivity. The resulting stalling of DNA synthesis and low catalytic activity likely underlie the severe clinical phenotypes in heterozygotes [101]. Additionally, Y955C PolG increases nucleotide misinsertion errors by 10- to 100-fold in the absence of exonucleolytic proofreading [102].
A mouse model harboring the human A467T POLG mutation recapitulated the accessory subunit binding defect observed in humans [103]. In this model, the isolated PolG catalytic subunit becomes a substrate for the mitochondrial Lon protease (LonP1), leading to removal of the unstable A467T polymerase. Cross-linking mass spectrometry mapped this interaction to the accessory interacting determinant (AID) domain, which mediates PolG2 accessory subunit binding [104]. Thus, association with PolG2 not only enhances processivity, DNA binding, and salt tolerance but also protects the enzyme from proteolysis, suggesting an important regulatory role in polymerase activity as well as a site for therapeutic intervention (see Section 8).

4. Accessory Subunit, PolG2

The accessory subunit of human pol γ is encoded by the POLG2 gene on chromosome 17q23-24 and translates into a 485 amino-acid protein weighing 55 kDa [105]. PolG2 functions as a processivity factor for mtDNA replication catalyzed by PolG. As a processivity factor, PolG2 functions include increasing DNA binding of the holoenzyme, accelerating polymerization rate, and suppressing exonuclease activity [67]. The gene for the accessory subunit has been identified in humans as well as mice, Drosophila, Xenopus, and other eukaryotes. In contrast, yeast, fungi, and C. elegans lack an accessory subunit altogether, as their pol γ functions as a single catalytic subunit [62,106,107]. PolG2 is unique as a eukaryotic processivity factor because of its similarity in primary amino acid sequence and conservation of tertiary structure to prokaryotic aminoacyl-tRNA synthetases and, to a lesser extent, thioredoxin, the accessory subunit of T7 DNA polymerase [108,109,110]. Additionally, PolG2 has also been reported to serve as a primer recognition factor and clamp loader in mtDNA replication in Drosophila pol γ [108].
The PolG2 accessory subunit has been characterized as a homodimer in solution, comprising the proximal and distal subunits [67,109]. The p55 dimer forms a strong interface of approximately 4000 Å2 [111] and ultracentrifugation studies have confirmed its tight association, with a dissociation constant (KD) of less than 0.1 nM. Functional studies indicate that the proximal p55 subunit stabilizes DNA binding, whereas the distal subunit enhances nucleotide incorporation [67,111]. Each monomer is further divided into three functional domains: domain one lies downstream of the N-terminal mitochondrial targeting sequence (MTS) and functions in p55 dimerization and binding of PolG, domain two contains the four-helix bundle that functions in homodimerization of p55, and the C-terminal domain three is involved in binding PolG [62]. PolG2 has also recently been reported to possess dimeric and hexameric (trimer of PolG2 dimers (Figure 2)) DNA binding modes [112]. The dimeric PolG2 preferentially binds duplex DNA, while multimeric PolG2 binds to DNA crossings or forked DNA resembling DNA substrates found in the D-loop of mtDNA. These observations suggest that PolG2 DNA binding serves both PolG-dependent and -independent utility in mtDNA replication and maintenance [112,113]. Additionally, PolG2 also permits increased salt tolerance of the pol γ holoenzyme during DNA synthesis, protects the catalytic subunit from inhibition by N-ethylmaleimide, and from degradation by LonP1 [103,104,105].
The mitochondrial disorders associated with POLG2 mutations involve amino acid residues distributed across the three domains of the protein. Generally, disease mutations in POLG2 give rise to decreased interaction with the catalytic subunit, PolG, as well as instability of the homodimeric enzyme. The first pathogenic POLG2 mutation (c.1352G > A, p.G451E) was observed in a 60-year-old patient with late-onset adPEO, with multiple mtDNA deletions in the skeletal muscle reflected in mild weakness of facial and limb muscles, and ptosis [47]. In the PolG2 structure, this heterozygous mutation is located in a loop region in domain three that is not involved in the dimerization of the protein. Biochemical analysis of the G451E PolG2 variant revealed that this variant retained DNA binding capability but exhibits impaired interaction with the catalytic subunit, PolG, hence impairing processivity of the holoenzyme. The second heterozygous POLG2 disease mutation (c.1207_1208ins24) was identified in a patient with similar phenotypic manifestations as the first, including adPEO, late onset ptosis, and mild myopathy. This 24 bp insertion causes mis-splicing and skipping of exon 7 in cDNA analysis. PolG2 with exon 7 that has been spliced is predicted to cause disruption of the C-terminal region required for processivity [114]. This mutation was also reported with the dominant clinical symptom of camptocormia in a 68-year-old patient, further supporting the point that mitochondrial disorders may be similar but present with varied phenotypes [115].
In an analysis of 112 patients with mitochondrial diseases and no POLG mutations, eight variants were observed, seven of which were novel heterozygous PolG2 disease variants [116]. However, only the R369G PolG2 disease variant is linked to adPEO, similar to G415E PolG2. The R369G PolG2 variant exhibits less binding affinity for the catalytic subunit PolG; however, the G451E variant appears to be more defective than the R369G PolG2 homodimeric variant in DNA processivity assays [117]. Further analyses of heterozygous PolG2 disease mutants were performed due to the hypothesis that affected patients may harbor mixtures of variants and wild-type (WT) p55 molecules. Heterodimeric PolG2 disease variants were purified with a tandem affinity strategy, and biochemical analysis showed that the G451E PolG2 variant was a dominant negative that caused severe inhibition of DNA processivity with WT p140 in vitro [118]. The inhibition by WT/G451E p55 heterodimer is distinct from the inability of the G451E p55 homodimer to interact with WT p140, suggesting that a single mutant monomer in the heterodimer is dominant negative and disrupts the interaction with the catalytic subunit. The results from the WT/G451E p55 heterodimer were in contrast to the R369G p55 heterodimeric variant, which displayed wild-type-like activity. Additionally, in vivo studies with R369G and G451E PolG2 heterodimeric variants revealed that they failed to localize to mtDNA-containing nucleoids and that their expression is associated with diminished reserve respiratory capacity [118].
The first homozygous POLG2 disease mutation (c.544C > T, p.R182W) was identified in a 3-month-old patient with fulminant hepatic failure due to severe mtDNA depletion [119]. This pathogenic variant is classified within MDS, which is an autosomal recessive disorder characterized by significantly decreased mtDNA content that impairs energy production in multiple organ systems. The Arg182 residue is conserved across vertebrates and is located at the base of the four-helix bundle that is critical for dimer formation, suggesting that the mutation to tryptophan leads to loss of electrostatic interactions that may disrupt dimerization. HEK293 cells expressing R182W PolG2 showed significantly impaired respiratory capacity, while patient fibroblasts showed a reduction in mtDNA copy number. Interestingly, compared to WT PolG2, the R182W variant exhibits no difference in DNA binding affinity, p140 binding affinity, steady-state stimulation of p140 DNA polymerase activity, and stimulation of processive DNA synthesis by pol γ [120]. However, protein purification and results from size-exclusion chromatography–multi-angle light scattering (SEC-MALS) revealed the R182W PolG2 variant to be potentially unstable, potentially disrupting dimerization. This finding was confirmed by the significant reduction in thermostability of R182W PolG2 compared to wild-type PolG2, and this instability further impacted its ability to stimulate PolG in thermostability activity assays [120].
The second reported homozygous POLG2 disease mutant (c.694G > A, p.G232S) was documented in a patient with epileptic seizures without ophthalmoplegia. The p.G232S variant is also the second recessive POLG2 mutation reported, though it is a single-nucleotide polymorphism (SNP) with allele frequency ≤0.0002 [121]. The first ultrarare homozygous missense POLG2 mutation (p.Asp433Tyr) that resulted in MDS was identified in an adult patient with a childhood onset and complex neuro-ophthalmic phenotype [122]. Asp433 is located in the proximal monomer of PolG2, which is also close to the catalytic thumb subdomain of PolG. A network of hydrogen-bonding interactions was identified with this residue that enables further polar interactions with PolG in the holoenzyme. An Asp-to-Tyr mutation is predicted to disrupt the hydrogen bond interactions in the proximal monomer and to further abolish interactions with PolG. The A433T PolG2 variant exhibited reduced thermal stability, intermediate efficiency in stimulating processivity, and had no measurable change in exonuclease activity of the holoenzyme, which may be explained by the loss of interactions predicted with this variant [122]. Additional POLG2 mutations have since been reported; however, only a subset has undergone clinical evaluation, and extensive biochemical characterization still needs to be performed [123,124,125,126,127,128].
In addition to its role as a processivity factor, the accessory subunit has been proposed to stabilize the catalytic subunit, PolG. This is evidenced by both the severe decrease in PolG expression in Polg2 knockout cells and the ability of WT PolG2 to stabilize PolG from heat-induced inactivation [95,129,130]. The Polg2 gene has also been reported to be critical for development in Drosophila, as mutations that impair the function of this gene caused lethality in the early pupal stage, concomitant with mtDNA depletion [131]. Results from studies where heterozygous (Polg2+/−) and homozygous (Polg2−/−) Polg2 knockout mice were generated to investigate the role of the accessory subunit in mammalian systems show that Polg2+/− mice are haplosufficient and exhibit normal development after two years. However, homozygous Polg2 knockout mice exhibit embryonic lethality at 8–8.5 dpc, alongside loss of mtDNA and mtDNA gene products [65]. Furthermore, knockdown of Polg2 in porcine oocytes showed a decrease in mtDNA copy number, which led to a reduction in ATP synthesis, oocyte maturation rate, and maturation-promoting factor activity, resulting in an aging-type phenotype [132]. PolG2 has also been reported to be critical for nucleoid maintenance, and its role in the initiation of mitochondrial DNA replication has been proposed to occur through the recruitment of proteins to the D-loop [133]. Collectively, these results demonstrate the fundamental role of PolG2 in mammalian embryogenesis, mtDNA maintenance, replication, and initiation.

5. The Human Mitochondrial Single-Stranded DNA-Binding Protein, mtSSB

Single-stranded DNA-binding proteins (SSBs) are among the most abundant and evolutionarily conserved cellular proteins found in all domains of life. A major characteristic of this group of proteins is their high nonspecific affinity for ssDNA [134,135,136]. SSB proteins play essential roles in various cellular processes, including DNA replication, repair, and recombination. Single-stranded DNA molecules are known to form secondary structures and can reanneal to themselves, which in both cases may interfere with these aforementioned cellular processes [135,136,137]. In addition, ssDNA is vulnerable to being cleaved by intracellular nucleases. Therefore, SSB proteins are required to protect and stabilize ssDNA from nuclease degradation and to keep ssDNA in a functionally active state. Several enzymes and protein factors can interact with ssDNA during different stages of DNA replication, repair, and recombination [137]. SSB proteins are known to interact both functionally and physically with genome maintenance proteins, by directing them to their respective sites of action or by stimulating their biochemical functions [137,138,139,140,141]. SSB is present in both the eukaryotic mitochondria and in the nucleus; however, the human mitochondrial SSB (mtSSB) is structurally and evolutionarily distinct from its nuclear counterpart, replication protein A (RP-A) [137].
MtSSB is essential for mtDNA replication, initiation, and elongation, as it coats the displaced parental H-strand during mtDNA replication. Furthermore, during mtDNA replication, mtSSB prevents primer synthesis by POLRMT on the displaced H-strand and controls the initiation of L-strand mtDNA synthesis from its designated origin of replication, OL [24,142,143,144]. Unlike POLRMT, Twinkle, and pol γ which are evolutionarily related to their bacteriophage T7 counterparts, mtSSB does not share bacteriophage ancestry but instead exhibits structural and biochemical similarities to Escherichia coli SSB [145]. The gene encoding human mtSSB, SSBP1, was first cloned in 1993 [146]. The cDNA of SSBP1 is translated as a 148 amino-acid polypeptide, in which the first 16 amino acids make up the MTS required for import of this protein into the mitochondria [146]. Human mtSSB binds ssDNA as a tetramer, composed of four identical 16 kDa subunits, each containing an oligosaccharide/oligonucleotide binding fold (OB-fold) [147,148]. Similar to E. coli SSB, human mtSSB can bind ssDNA in two distinct modes, termed (SSB)30 and (SSB)60, where the number represents how many nucleotides are occluded per tetramer [136,149,150,151,152]. In a previous study by our group, atomic force microscopy of the human mtSSB with ssDNA demonstrated a single wrap of ssDNA around the mtSSB tetramer at physiological salt conditions [153]. Despite having only 32% amino acid identity, the overall tetramer folding of the human mtSSB resembles the tetrameric structure of E. coli SSB. However, E. coli SSB contains an additional 60-amino-acid acidic C-terminal tail required for protein–protein interactions, which is lacking in the human mtSSB [148,154,155].
Despite the lack of evidence for physical protein–protein interactions between mtSSB and other replisome proteins, mtSSB has been reported to functionally interact with pol γ and the Twinkle helicase at the replication fork [143,156]. Moreover, mtSSB facilitates the binding of pol γ to the primer-template DNA and enhances the processivity of pol γ during mtDNA initiation and elongation, thereby having an overall stimulatory effect on mtDNA synthesis [143]. In addition, mtSSB has been reported to stimulate the unwinding activity of the mitochondrial Twinkle helicase at the replication fork. Since ssDNA can re-anneal to itself and hinder mtDNA replication, mtSSB prevents ssDNA reannealing through binding of ssDNA, thereby stimulating the unwinding activity of Twinkle [144,156]. The functional interaction between human mtSSB and the mitochondrial Twinkle helicase appears to be highly specific, since the structurally similar E. coli SSB could not stimulate the unwinding activity of the human mitochondrial Twinkle helicase [157]. Functional interactions of mtSSB with pol γ and Twinkle have also been demonstrated through in vitro studies, in which, in the absence of mtSSB, pol γ and Twinkle can synthesize ssDNA products only up to ~2 kb. The addition of mtSSB to this reaction enhances the processivity of the minimal replication machinery, resulting in DNA products longer than ~16 kb in vitro [158]. Although biochemical analyses provide evidence that mtSSB functionally interacts with pol γ and Twinkle, direct physical protein–protein interactions are currently elusive [158].
While SSB proteins lack direct catalytic activity, they are essential for DNA metabolism [137]. In addition to humans, mtSSB proteins are present in higher eukaryotes, including yeast, mice, Xenopus laevis, and Drosophila [159]. In yeast, deletion of the RIM1 gene, encoding mtSSB, results in complete loss of mtDNA and yields strains that are unable to grow on non-fermentable carbon sources [48]. Similarly, RNAi knockdown of Drosophila mtSSB significantly reduces its expression in Schneider cells, resulting in growth defects and severe mtDNA depletion [160]. Knockdown of mtSSB in HeLa cells results in a moderate decrease in mtDNA synthesis, a gradual reduction in mtDNA copy number, and a significant reduction in the synthesis of 7S DNA [161]. Knockdown of mtSSB in osteosarcoma cells has also been shown to significantly affect the synthesis of 7S DNA, reinforcing the role of mtSSB as both an mtDNA maintenance factor and a key player in mtDNA replication [161].
Since 2018, over 10 mutations have been identified in the SSBP1 gene that have been associated with mitochondrial diseases. Patients with SSBP1 mutations typically suffer from optic atrophy and have also been reported to experience neurological dysfunction mainly affecting the visual system (retinopathy, foveopathy, ptosis, ophthalmoplegia) and auditory system (sensorineural hearing loss) [46,162,163,164,165]. With the exception of the I132V disease-associated mutation in SSBP1, all other reported mutations are heterozygous in patients [159,164]. These mutations are mostly missense mutations, with the lone exception of c.3G > A, which abolishes the start codon of SSBP1, resulting in decreased protein translation. The c.3G > A mutation in SSBP1 observed in a patient also co-segregated with a point mutation in the mtDNA (m.1555A > G), which resulted in decreased levels of 7S DNA in fibroblasts, mtDNA depletion, and multiple mtDNA deletions [162]. In 2019, a single de novo SSBP1 mutation (E27K) was observed in a 14-year-old patient who had mitochondrial disease manifestations across the full Pearson, Kearns–Sayre, and Leigh syndromes spectrum. In addition, the patient presented with infantile anemia, bone marrow and growth failure, ptosis, ophthalmoplegia, metabolic strokes, and chronic kidney disease [46]. Results from mtDNA genome sequencing using next-generation sequencing revealed a single large-scale mtDNA deletion (SLSMD) of 5440 base pairs (m.8629_14068del5440). While E27K mtSSB remains a stable tetramer in solution and exhibits similar binding affinity for ssDNA as the wild-type protein, this mutation resulted in a slight decrease in thermal stability [46]. Furthermore, molecular modeling of E27K mtSSB comprising different combinations of mutant subunits predicted changes in the surface-accessible charges, strength of inter-subunit interactions, and overall protein dynamics [46]. Altogether, it was proposed that these observed alterations could interfere with mtDNA replication, explaining the resulting SLSMD observed in the patient [46,159].
Like several other disease-associated residues, Glu27 is evolutionarily conserved across higher eukaryotes [46,159]. Within the homotetrameric structure of mtSSB, Glu27 is located at the monomer–monomer interface at the edge of a positively charged surface patch containing Arg38 and Arg107, which are both residues with mitochondrial disease alleles (R38Q and R107Q). In the structure, Glu27 participates in an inter-subunit interaction with Arg38 and Arg107, which has been proposed to be structurally significant for tetramer formation [46,159,165]. This inter-subunit interaction is facilitated by the guanidinium nitrogens of the Arg38 and Arg107 side chains, which interact with the side-chain carboxylate group of Glu27 in the neighboring subunit [46,159,165]. Even though the first X-ray crystallography structure of apo mtSSB was determined in 1997, a ssDNA-bound mtSSB structure was elusive for more than 20 years [148]. In 2024, our group was the first to determine a high-resolution X-ray crystallography structure of mtSSB bound to a dT20 ssDNA (Figure 2) [147]. This ssDNA-bound mtSSB structure revealed that three residues (Arg38, Gly40, and Gln62) with mitochondrial disease alleles (R38Q, G40V, and N62D) reside within 3.5Å of the ssDNA chain. These three residues were shown to participate in interactions with either the ssDNA backbone or the thymine nucleotides [147].
The autosomal dominant mutations, R38Q and R107Q, in SSBP1, were first identified in five unrelated families, all of whom exhibited optic atrophy. In addition to optic atrophy, nearly half of the affected individuals also presented with foveopathy [164,165]. Biochemical analyses using patient fibroblasts revealed that the R38Q and R107Q mtSSB variants impair tetramer stability, perturb mtDNA replication, and lead to mtDNA depletion. These results also revealed reduced levels of TFAM and PolG in patient fibroblasts [164,165]. Autosomal dominant mutations in the nuclear-encoded replication genes impair mtDNA replication by inducing stalled replication forks [166,167]. Therefore, frequent pausing of mtDNA replication would explain the depletion of mtDNA observed in patient fibroblasts [165]. In addition to mtDNA depletion, patients carrying the R107Q SSBP1 mutation exhibited a significant reduction in Complex I and IV, resulting in severe respiratory deficiency [164]. Several other mtSSB disease variants (G40V, N62D, R107Q, E111Q, and I132V mtSSB) have been evaluated based on their ability to stimulate DNA synthesis. Although all these disease variants were able to support full-length DNA synthesis, they were consistently less efficient than the wild-type protein [164]. In the same study, zebrafish models were employed to assess the relevance of ssbp1 in relation to the optic atrophy phenotype observed in patients with SSBP1 mutations. Loss of ssbp1 in zebrafish affected the development of the optic nerve and resulted in signs of nephropathy. While all dominant mtSSB variants were ineffective in rescuing the optic nerve phenotype, the recessive I132V mtSSB disease variant behaved as a hypomorph [164]. Finally, results from another study evaluating the ability of mtSSB disease variants to support replication initiation and elongation revealed that the G40V, N62D, R107Q, and E111Q mtSSB disease variants cannot support replication initiation in vitro. However, these variants did not impair replication elongation [168].

6. The Mitochondrial DNA Helicase, Twinkle

Helicases are a class of molecular motor proteins that employ the energy derived from nucleotide hydrolysis to drive the unwinding of double-stranded DNA/RNA [169]. They play diverse roles in DNA replication, repair, recombination, and protein synthesis. Helicases are classified into six families (SF 1–6) based on conserved sequence and structural motifs [170]. Although multiple helicases have been identified in mitochondria, Twinkle is the only helicase in mtDNA replication, where it functions in unwinding duplex DNA and during pol γ–catalyzed mtDNA synthesis [171,172]. Twinkle was first reported in 2001; it belongs to superfamily 4 (SF4) DNA helicases and shares 46% amino acid sequence similarity and 15% sequence identity with the bacteriophage T7 gene protein 4 (gp4) primase–helicase fusion protein [173,174]. In mammalian cells, Twinkle is encoded by the TWNK gene (also commonly referred to as C10orf2 or PEO1), and is translated into a 684-amino-acid protein with a molecular weight of approximately 77 kDa [144]. Twinkle helicase is highly conserved in many eukaryotes including Drosophila and mice but lacks orthologs in yeast [175,176].
The replicative Twinkle helicase is structurally organized into an N-terminal region (NTR), which includes the primase-like domain, connected by a flexible linker region to the C-terminal helicase domain. Although Twinkle shares significant homology with the bacteriophage T7 primase–helicase (gp4) through its C-terminal helicase domain (CTD), the NTR and primase-like domain lack critical conserved cysteine residues required for formation of the zinc finger binding domain (ZBD), which is essential for primase activity [173]. The different domains have unique functions: the N-terminal domain contributes to ssDNA binding, helicase activity, and processivity; the linker region is critical for multimerization; and the C-terminal helicase domain functions in oligomer stability [171,177]. Disease-causing mutations in the TWNK gene have been mapped mostly to the linker region and C-terminal domain [178]. Results from previous studies have reported the ability of Twinkle to interact with a variety of DNA substrates including circular and linear ssDNA, but it shows a greater affinity for linear dsDNA [144,179,180]. However, for the initiation of unwinding, Twinkle requires an open fork-like dsDNA comprising a single-stranded 5′ DNA loading site and a short 3′ tail [144].
Previous structural and biochemical studies have predicted that Twinkle exists in multiple oligomeric states and undergoes oligomeric shifts [156]. These oligomeric transitions have been observed to be regulated by salt concentrations and the presence of Mg2+ and dNTPs [181,182]. As a ring-structured 5′→3′ DNA helicase, Twinkle binds dNTPs at the interface and DNA in the central channel, and the mechanism by which Twinkle loads onto DNA has been predicted to occur through oligomeric transitions typical for gp4 helicases [144,183]. Recent cryo-EM structures of the Twinkle W315L disease variant revealed two oligomeric states, heptamer and octamer (Figure 2) [184]. Additionally, results from SEC-MALS revealed that the W315L Twinkle heptamer was observed upon DNA binding, which supports the ejection theory of a monomer subunit from the ring-shaped octamer [184]. Twinkle’s unwinding activity has also been shown to facilitate the degradation of linear mtDNA by endonucleases specific for ssDNA, including the 3′→5′ exonuclease activity of PolG [185]. In addition to its helicase activity, Twinkle also functions in strand exchange, reannealing, and branch migration, which are proposed to be important for DNA recombination [186,187,188,189]. Given that Twinkle functionally interacts with mtSSB and pol γ, it has been proposed that even subtle changes in Twinkle’s activity may disrupt the coordination with mtSSB and pol γ at the replication fork [143,156,158,190,191,192].
The mitochondrial DNA helicase, Twinkle, colocalizes with mtDNA in mitochondrial nucleoids [13,173,193]. The importance of Twinkle in mtDNA maintenance has been demonstrated through multiple lines of experimental evidence. In particular, studies using in vivo mouse models have provided evidence that Twinkle is critical for embryonic development, as homozygous knockout of Twinkle results in embryonic lethality at 8.5 dpc and co-segregates with mtDNA depletion, impaired mtDNA expression, and respiratory chain deficiency [178]. In contrast, mice harboring a conditional tissue-specific knockout of Twinkle in their heart and skeletal muscle remain viable [178]. Conversely, systemic overexpression of Twinkle in a transgenic mouse results in a significant increase in mtDNA copy number; however, these animals exhibited enlarged nucleoid structures, impaired mtDNA replication, and disrupted transcription [20,194]. In Drosophila, depletion of Twinkle in Schneider cells results in slow growth, decreased viability, and a fivefold reduction in mtDNA copy number [195]. Collectively, these results confirm that Twinkle cannot be substituted by other mitochondrial helicases and underscore the importance of Twinkle as a key player in mtDNA replication and maintenance [171].
The TWNK gene was initially identified through sequencing analysis in a group of patients suffering from PEO, which co-segregated with multiple mtDNA deletions and 11 different mutations across 12 families [173]. Typically, an autosomal dominant mode of inheritance is revealed by the combined existence of disease and mutant allele, in which one mutant allele is sufficient to develop adult-onset PEO. In contrast to their heterozygous relatives, homozygous individuals exhibit a more severe phenotype with earlier onset of PEO [173]. In addition to PEO, the T457I, Y508C, and A318T Twinkle disease variants are associated with neurological disorders, hepatocerebral MDS, and infantile-onset spinocerebellar ataxia (IOSCA), which manifest in the initial stages of life [196,197]. Individuals homozygous for a TWNK mutation causing a dual T457I substitution presented with hepatocerebral MDS and died by the age of three years; however, their heterozygous parents and grandparents remain unaffected [198]. A similar result was also observed for the Y508C Twinkle variant in which homozygous individuals developed IOSCA, whereas heterozygotes were unaffected [199]. Furthermore, biochemical analyses of recombinant Y508C Twinkle revealed that this variant has similar helicase activity as the wild-type; however, a modest decrease in ssDNA binding affinity has been reported [197,200]. Finally, the A318T Twinkle variant has only been observed as a heterozygous mutation with Y508C. Individuals harboring both the A318T and Y508C TWNK mutations presented with a severe early-onset encephalopathy and mtDNA depletion in the liver. Taken together, these findings suggest that one T457I-, Y508C-, or A318T-substituted copy of TWNK is not sufficient to induce disease [196,201]. These findings also emphasize that one wild-type TWNK allele can rescue any biochemical defects caused by these mutant alleles [200].
Many of the TWNK mutations found in adPEO patients are located within or in close proximity to the linker region [202]. The cryo-EM structure shows that these mutations cluster at subunit interfaces, and substitution of these residues disrupts this subunit–subunit interface [184]. The importance of the linker region was also demonstrated through in vitro biochemical analyses of several Twinkle disease variants. Unlike the wild-type enzyme, I367T and R374Q Twinkle disease variants eluted as monomers in gel filtration assays and displayed no ATPase and helicase activity [174]. Furthermore, the I367T, S369P, R374Q, and L381P Twinkle disease variants were unable to perform rolling-circle DNA synthesis. This inability to facilitate DNA synthesis was attributed to impaired ATPase and helicase activities as well as the failure of certain variants to form hexamers in solution [174]. In contrast, a previous study from our group demonstrated that 20 Twinkle disease variants have similar helicase activity as the wild-type enzyme under optimized in vitro conditions [200]. Despite preserved helicase activity, several of these 20 Twinkle disease variants displayed reduced DNA binding affinity, diminished nucleotide hydrolysis rates, and decreased thermal stability compared to the wild-type enzyme [200]. Given the moderate biochemical defects observed in vitro, we proposed that these results are consistent with the delayed onset of adPEO due to TWNK mutations [200]. It is noteworthy that the current literature contains significant discrepancies regarding different physical and biochemical properties of wild-type and mutant Twinkle. These discrepancies may be explained by the specific methods and conditions utilized for biochemical evaluation [166,174,177,200].
In addition to in vitro biochemical analyses, the use of in vivo models has been useful in elucidating disease outcomes associated with Twinkle variants. Transgenic mice overexpressing either the dominant A360T Twinkle mutation or a 13-amino-acid duplication in the linker region were found to accumulate multiple mtDNA deletions [203]. The transgenic mouse containing the duplication is also referred to as the deletor mouse in the literature and exhibits respiratory dysfunction and mitochondrial myopathy starting at one year of age [203]. The phenotypes observed with the deletor mouse closely resembled those seen in PEO patients [203,204]. A significant hallmark of PEO disease is the accumulation of multiple large-scale mtDNA deletions in postmitotic tissues of the patients, with the level of mutant mtDNA reaching up to 60% in the brain region and 40% in the skeletal muscle [205,206]. Furthermore, overexpression of several adPEO Twinkle variants in HEK293 cells resulted in the accumulation of replication intermediates. Expression of these adPEO Twinkle variants also resulted in decreased mtDNA copy number, enlarged nucleoids, and was proposed to induce replication stalling [166,167]. Beyond PEO, defects in Twinkle have also been linked to neurodegenerative diseases [207]. Analysis of neurons from the deletor mouse revealed cytochrome c oxidase deficiency, which was attributed to mtDNA depletion and the accumulation of mtDNA deletions in specific neuronal populations [203]. Collectively, these findings underscore the critical role of Twinkle in mtDNA replication and maintenance.

7. Other Proteins Involved in mtDNA Maintenance

Several proteins are involved in maintaining mtDNA genomic stability and mitochondrial function. Beyond the core replication machinery discussed in this review, defects in proteins involved in mtDNA transcription, replication, nucleotide pool maintenance, and DNA repair pathways of the mitochondrial genome can all influence mtDNA integrity (Table 1). Additionally, disease alleles in proteins involved in mitochondrial dynamics and mitochondrial quality control pathways may influence mtDNA integrity, as these pathways are involved in removing damaged mitochondrial genomes.
In mitochondria, RNase H1 is crucial to mtDNA replication [208] and has been implicated in both primer removal during mtDNA replication [209,210] and primer maturation [211]. Additionally, RNase H1 is critical for the segregation of daughter mtDNA molecules following replication [212], R-loop resolution [213], regulating transcription of 7S RNA [214], and pre-rRNA processing [215]. Disease-associated alleles in RNASEH1 have been identified and affected patients display classic mitochondrial disease phenotypes, including mitochondrial myopathy, PEO, cerebellar ataxia, muscle weakness, dysphagia, mtDNA depletion, and the accumulation of mtDNA deletions [216,217,218,219,220].
DNA replication helicase/nuclease 2 (DNA2) localizes to both the mitochondria and the nucleus. DNA2 is essential to mtDNA replication and the long-patch base excision repair pathway in the mitochondria [221,222,223,224]. Numerous pathogenic variants in DNA2 have been identified with clinical outcomes such as mtDNA instability, mitochondrial myopathy [225,226,227,228], congenital ptosis, and hypotonia [229]. Other disease outcomes include microcephalic primordial dwarfism [230,231], mtDNA depletion syndrome associated with hearing loss and muscle weakness [232], and mesial temporal lobe epilepsy [233]. Recently, a patient carrying a DNA2 mutation presented with adult-onset mitochondrial encephalomyopathy, mtDNA depletion, and mtDNA deletions [234].
Mitochondrial genome maintenance exonuclease 1 (MGME1) is a mitochondrial deoxyribonuclease that can cleave ssDNA in both 5′→3′ and 3′→5′ directions, dsDNA flaps, and RNA-DNA flaps [235,236]. MGME1 is crucial to mtDNA maintenance, the regulation of 7S DNA, and is involved in the degradation of linear mtDNA [185]. Loss-of-function mutations in MGME1 are associated with mtDNA depletion and mtDNA rearrangements [235,237,238]. Clinical manifestations linked to MGME1 mutations include PEO, skeletal muscle weakness, respiratory distress, and neurological symptoms [235,238].
Top3α localizes to both the nucleus and the mitochondria. The mitochondrial isoform of Top3α is essential for genome segregation during mtDNA replication [32] and for regulating topology of mtDNA at the replication fork [239]. Pathogenic variants in TOP3A have been associated with Bloom syndrome-like disorders [240,241] as well as primary mitochondrial disease characterized by adult-onset PEO [32,33,242,243].
TFAM, which facilitates mtDNA compaction, initiates transcription, plays a role in the regulation of mtDNA replication, and participates in DNA repair processes, has very few disease alleles. Although there are numerous documented allelic variants of the TFAM gene, only two pathogenic variants have been identified [244]. A c.533C > T (p.Pro178Leu) mutation was associated with mtDNA depletion and neonatal liver failure [245]. The P178L TFAM variant has been reported to have a deficiency in transcription initiation [246]. A c.694C > T (p.Arg232Cys) mutation has also been reported and is associated with primary ovarian insufficiency in females, abnormal sex hormone levels in males, neurological symptoms, and mtDNA depletion [247]. Among the non-disease-causing allelic variants of TFAM that have been reported, evidence suggests that some variants are implicated in the risk of Alzheimer’s disease and Parkinson’s disease [248,249], reduced survival rates in patients with head and neck cancers [250], and are co-occurring with colorectal cell line and tumor cancers [251].
The mitochondrial RNA polymerase initiates mtDNA replication via the synthesis of RNA primers for pol γ to begin replication. Recently, 14 pathogenic variants of POLRMT associated with mitochondrial disease have been reported, with phenotypes including adult-onset PEO, motor delays, and neurological symptoms [252,253]. Although all patient mutations in POLRMT were associated with reduced in vitro primer synthesis, patient fibroblasts harboring the POLRMT mutations did not exhibit mtDNA depletion or deletions. Notably, these fibroblasts exhibited a significant loss of OXPHOS mRNA transcripts [252].
While we have briefly described several disease-associated variants in proteins that are crucial to maintaining mitochondrial genomic integrity and supporting mtDNA replication, this list is not exhaustive. Additional genes not covered in this review include those involved in DNA repair processes of the mitochondria, nucleotide pool maintenance, mitophagy, mitochondrial quality control mechanisms, and those for which disease variants cause mtDNA depletion without a known mechanism (Table 1).

8. Therapies for Mitochondrial Diseases

Current therapy for PolG-related disorders is primarily symptomatic and supportive, with a focus on managing symptoms and preventing progression. Seizure management needs to avoid the use of valproate, which can cause fatal liver failure in patients with POLG mutations. Mitochondrial-cocktail supplements including coenzyme A, riboflavin, L-carnitine, creatine, arginine, and other cofactors, as well as antioxidants, are frequently used to improve mitochondrial function and enhance energy metabolism in patients. Below, we summarize recent research-supported therapies that may offer promising clinical benefits.
Nucleotide therapy: MtDNA replication is dependent on dNTP pools in the mitochondria, and one of the main sources of dNTPs is the salvage pathway involving phosphorylation of pyrimidines by thymidine kinase 2 (TK2) and purines by deoxyguanosine kinase (DGUOK). Mutations in the genes TK2 and DGUOK result in early childhood mitochondrial DNA depletion (reviewed in [51]). Over the past decade, nucleotide therapy has proven to be highly effective against TK2 deficiencies as a result of TK2 gene mutations [254,255]. Nucleotide therapy application for PolG defects has also been considered. Given that mitochondrial deoxynucleotide pools are usually limiting for mtDNA replication, increasing dNTP availability has shown promise in fibroblasts from patients harboring either POLG or TWNK mutations in ethidium bromide-forced depletion and recovery [256,257]. This approach has been extended to iPSC-derived neural stem cells with POLG mutations and has been shown to improve mtDNA depletion after ethidium bromide treatment [258]. Furthermore, in a 6-month, open-label trial, supplementation with deoxycytidine and deoxythymidine in children with PolG-related disorders demonstrated improved energy levels, enhanced EEG profiles, and favorable changes in multiple biomarkers of mitochondrial dysfunction [259,260].
Metformin and nicotinamide riboside: Mitophagy is a key process for the removal of damaged mitochondria that can result from a buildup of mtDNA mutations derived from PolG defects. Investigation into modulating mitophagy in POLG-mutant astrocytes has identified that combined treatment with nicotinamide riboside (NR) and metformin restored mitophagy and mitochondrial function [261]. These findings indicate that impaired mitophagy contributes to mitochondrial dysfunction in POLG mutations and highlight NR and metformin as potential therapies for PolG-related mitochondrial diseases.
Targeted drug screening: Recently, a small effector molecule, PZL-A, was identified in a high-throughput screen of 270,000 compounds based on its ability to stimulate pol γ activity and mitigate the phenotypes associated with impaired pol γ function [262]. The PZL-A molecule binds allosterically at the interface between the catalytic PolG and accessory PolG2 subunits. PZL-A effectively restores near wild-type activity to the most common mutant forms of PolG in vitro, reviving mtDNA synthesis. Furthermore, in cells from pediatric patients with lethal diseases associated with POLG mutations, PZL-A enhances mitochondrial DNA replication, boosts the biogenesis of oxidative phosphorylation complexes, and improves cellular respiration. Structural studies support these findings, including several high-resolution cryo-EM structures demonstrating that both wild-type and mutant PolG complexes bound to PZL-A, illuminating how the molecule interacts with and activates the enzyme [262]. This compound represents a promising therapeutic candidate for conditions caused by POLG mutations.

9. Concluding Remarks

One of the most important functions of the mitochondria is the production of cellular ATP via OXPHOS, which is subsequently used for various metabolic processes. Cellular ATP is generated through a series of electron transfer reactions requiring genes expressed by both the nuclear and mitochondrial genomes. Defects in this energy production pathway cause a broad spectrum of diseases, collectively known as mitochondrial diseases. Mutations in the nuclear-encoded mtDNA replication machinery continue to be one of the main causes of mitochondrial diseases. These mutations impair the rate and efficiency of the replisome proteins during progression of the replication fork, mtDNA synthesis, protein–DNA affinity, as well as protein and complex stability. Collectively, alterations in the biochemical functions of the replisome proteins may result in stalled replication forks, uncoupling of the replication machinery, DNA strand breaks, and exposure of ssDNA [263]. Consequently, stalled replication forks and DNA strand breaks can contribute to an increase in the number of point mutations, the formation of large-scale mtDNA deletions, and mtDNA depletion. Therefore, understanding the molecular mechanisms pertaining to mtDNA replication is critical to understanding the pathogenesis of mitochondrial diseases. Mechanistic knowledge of replisome proteins, combined with the use of animal models and biochemical characterization in the investigation of disease variants, provides a foundation for developing effective therapeutic interventions for mitochondrial diseases.

Author Contributions

S.S., conceptualization, writing—original draft, review and editing, and figure preparation; C.H.A., conceptualization, writing—original draft, review and editing, and figure preparation; D.E.K., conceptualization, writing—original draft, review and editing, and figure preparation; W.C.C., conceptualization, writing—original draft, review, supervision, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Intramural Research Program of the National Institutes of Health (NIH), National Institute of Environmental Health Sciences (Z01 ES065078 and Z01 ES065080 to W.C.C.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The contributions of the NIH author(s) are considered works of the United States Government. The findings and conclusions presented in this paper are those of the author(s) and do not necessarily reflect the views of the NIH or the U.S. Department of Health and Human Services.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

adPEOautosomal dominant progressive external ophthalmoplegia
AIDaccessory interacting determinant
arPEOautosomal recessive progressive external ophthalmoplegia
ATPadenosine 5′-triphosphate
bpbase pairs
CTDC-terminal helicase domain
DGUOKdeoxyguanosine kinase
DNA2DNA replication helicase/nuclease 2
dNTPsdeoxynucleotide triphosphates
dsDNAdouble-stranded DNA
ETCelectron transport chain
gp4T7 gene protein 4 primase–helicase
H-strandheavy strand
IOSCAinfantile-onset spinocerebellar ataxia
L-strandlight strand
LonP1Lon protease
MDSmitochondrial DNA depletion syndrome
MEMSAmyoclonic epilepsy myopathy sensory ataxia
MGME1mitochondrial genome maintenance exonuclease 1
MIRASmitochondrial recessive ataxia syndrome
mtDNAmitochondrial DNA
MTSmitochondrial targeting sequence
mtSSBmitochondrial single-stranded DNA binding protein
NCRnoncoding region
NRnicotinamide riboside
NTRN-terminal region
OXPHOSoxidative phosphorylation
p140catalytic subunit PolG
p55accessory subunit PolG2
PEOprogressive external ophthalmoplegia
Pol γDNA polymerase γ
POLRMTmitochondrial RNA polymerase
ROSreactive oxygen species
RP-Areplication protein A
SEC-MALSsize-exclusion chromatography–multi-angle light scattering
SF4superfamily 4
SLSMDsingle large-scale mtDNA deletion
SNPsingle nucleotide polymorph
SSBssingle-stranded DNA-binding proteins
ssDNAsingle-stranded DNA
TCAtricarboxylic acid
TFAMtranscription factor A, mitochondrial
TK2thymidine kinase 2
TOP3Atopoisomerase 3α
WTwild-type
ZBDzinc finger binding domain

References

  1. 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. Sequence and organization of the human mitochondrial genome. Nature 1981, 290, 457–465. [Google Scholar] [CrossRef] [PubMed]
  2. Casuso, R.A.; Huertas, J.R. Mitochondrial Functionality in Inflammatory Pathology-Modulatory Role of Physical Activity. Life 2021, 11, 61. [Google Scholar] [CrossRef] [PubMed]
  3. Harris, D.A.; Das, A.M. Control of mitochondrial ATP synthesis in the heart. Biochem. J. 1991, 280 Pt 3, 561–573. [Google Scholar] [CrossRef] [PubMed]
  4. Spinelli, J.B.; Haigis, M.C. The multifaceted contributions of mitochondria to cellular metabolism. Nat. Cell Biol. 2018, 20, 745–754. [Google Scholar] [CrossRef]
  5. Picard, M.; Shirihai, O.S. Mitochondrial signal transduction. Cell Metab. 2022, 34, 1620–1653. [Google Scholar] [CrossRef]
  6. Hatefi, Y. The mitochondrial electron transport and oxidative phosphorylation system. Annu. Rev. Biochem. 1985, 54, 1015–1069. [Google Scholar] [CrossRef]
  7. Berg, O.G.; Kurland, C.G. Why Mitochondrial Genes are Most Often Found in Nuclei. Mol. Biol. Evol. 2000, 17, 951–961. [Google Scholar] [CrossRef]
  8. Larsson, N.G.; Wang, J.; Wilhelmsson, H.; Oldfors, A.; Rustin, P.; Lewandoski, M.; Barsh, G.S.; Clayton, D.A. Mitochondrial transcription factor A is necessary for mtDNA maintenance and embryogenesis in mice. Nat. Genet. 1998, 18, 231–236. [Google Scholar] [CrossRef]
  9. Wiese, M.; Bannister, A.J. Two genomes, one cell: Mitochondrial-nuclear coordination via epigenetic pathways. Mol. Metab. 2020, 38, 100942. [Google Scholar] [CrossRef]
  10. Wallace, D.C. Mitochondrial diseases in man and mouse. Science 1999, 283, 1482–1488. [Google Scholar] [CrossRef]
  11. Tuppen, H.A.; Blakely, E.L.; Turnbull, D.M.; Taylor, R.W. Mitochondrial DNA mutations and human disease. Biochim. Biophys. Acta 2010, 1797, 113–128. [Google Scholar] [CrossRef]
  12. Wallace, D.C. Mitochondria and cancer. Nat. Rev. Cancer 2012, 12, 685–698. [Google Scholar] [CrossRef]
  13. Lee, S.R.; Han, J. Mitochondrial Nucleoid: Shield and Switch of the Mitochondrial Genome. Oxidative Med. Cell. Longev. 2017, 2017, 8060949. [Google Scholar] [CrossRef]
  14. Iborra, F.J.; Kimura, H.; Cook, P.R. The functional organization of mitochondrial genomes in human cells. BMC Biol. 2004, 2, 9. [Google Scholar] [CrossRef] [PubMed]
  15. Vinograd, J.; Morris, J.; Davidson, N.; Dove, W.F., Jr. The bouyant behavior of viral and bacterial DNA in alkaline CsCl. Proc. Natl. Acad. Sci. USA 1963, 49, 12–17. [Google Scholar] [CrossRef] [PubMed]
  16. Wells, R.D.; Larson, J.E. Buoyant density studies on natural and synthetic deoxyribonucleic acids in neutral and alkaline solutions. J. Biol. Chem. 1972, 247, 3405–3409. [Google Scholar] [CrossRef]
  17. Tan, B.G.; Mutti, C.D.; Shi, Y.; Xie, X.; Zhu, X.; Silva-Pinheiro, P.; Menger, K.E.; Díaz-Maldonado, H.; Wei, W.; Nicholls, T.J.; et al. The human mitochondrial genome contains a second light strand promoter. Mol. Cell 2022, 82, 3646–3660.e9. [Google Scholar] [CrossRef]
  18. Zhu, X.; Xie, X.; Das, H.; Tan, B.G.; Shi, Y.; Al-Behadili, A.; Peter, B.; Motori, E.; Valenzuela, S.; Posse, V.; et al. Non-coding 7S RNA inhibits transcription via mitochondrial RNA polymerase dimerization. Cell 2022, 185, 2309–2323.e24. [Google Scholar] [CrossRef]
  19. Magnusson, J.; Orth, M.; Lestienne, P.; Taanman, J.W. Replication of mitochondrial DNA occurs throughout the mitochondria of cultured human cells. Exp. Cell Res. 2003, 289, 133–142. [Google Scholar] [CrossRef]
  20. Ylikallio, E.; Tyynismaa, H.; Tsutsui, H.; Ide, T.; Suomalainen, A. High mitochondrial DNA copy number has detrimental effects in mice. Hum. Mol. Genet. 2010, 19, 2695–2705. [Google Scholar] [CrossRef]
  21. Falkenberg, M. Mitochondrial DNA replication in mammalian cells: Overview of the pathway. Essays Biochem. 2018, 62, 287–296. [Google Scholar] [CrossRef]
  22. Robberson, D.L.; Kasamatsu, H.; Vinograd, J. Replication of mitochondrial DNA. Circular replicative intermediates in mouse L cells. Proc. Natl. Acad. Sci. USA 1972, 69, 737–741. [Google Scholar] [CrossRef]
  23. Fusté, J.M.; Wanrooij, S.; Jemt, E.; Granycome, C.E.; Cluett, T.J.; Shi, Y.; Atanassova, N.; Holt, I.J.; Gustafsson, C.M.; Falkenberg, M. Mitochondrial RNA polymerase is needed for activation of the origin of light-strand DNA replication. Mol. Cell 2010, 37, 67–78. [Google Scholar] [CrossRef]
  24. Miralles Fusté, J.; Shi, Y.; Wanrooij, S.; Zhu, X.; Jemt, E.; Persson, Ö.; Sabouri, N.; Gustafsson, C.M.; Falkenberg, M. In vivo occupancy of mitochondrial single-stranded DNA binding protein supports the strand displacement mode of DNA replication. PLoS Genet. 2014, 10, e1004832. [Google Scholar] [CrossRef] [PubMed]
  25. Tapper, D.P.; Clayton, D.A. Mechanism of replication of human mitochondrial DNA. Localization of the 5′ ends of nascent daughter strands. J. Biol. Chem. 1981, 256, 5109–5115. [Google Scholar] [CrossRef]
  26. Phillips, A.F.; Millet, A.R.; Tigano, M.; Dubois, S.M.; Crimmins, H.; Babin, L.; Charpentier, M.; Piganeau, M.; Brunet, E.; Sfeir, A. Single-Molecule Analysis of mtDNA Replication Uncovers the Basis of the Common Deletion. Mol. Cell 2017, 65, 527–538.e6. [Google Scholar] [CrossRef] [PubMed]
  27. Clayton, D.A. Replication and transcription of vertebrate mitochondrial DNA. Annu. Rev. Cell Biol. 1991, 7, 453–478. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, Y.; Chen, Z.; Wang, Z.H.; Delaney, K.M.; Tang, J.; Pirooznia, M.; Lee, D.Y.; Tunc, I.; Li, Y.; Xu, H. The PPR domain of mitochondrial RNA polymerase is an exoribonuclease required for mtDNA replication in Drosophila melanogaster. Nat. Cell Biol. 2022, 24, 757–765. [Google Scholar] [CrossRef]
  29. Wanrooij, S.; Fusté, J.M.; Farge, G.; Shi, Y.; Gustafsson, C.M.; Falkenberg, M. Human mitochondrial RNA polymerase primes lagging-strand DNA synthesis in vitro. Proc. Natl. Acad. Sci. USA 2008, 105, 11122–11127. [Google Scholar] [CrossRef]
  30. Tan, B.G.; Gustafsson, C.M.; Falkenberg, M. Mechanisms and regulation of human mitochondrial transcription. Nat. Rev. Mol. Cell Biol. 2024, 25, 119–132. [Google Scholar] [CrossRef]
  31. Wanrooij, S.; Miralles Fusté, J.; Stewart, J.B.; Wanrooij, P.H.; Samuelsson, T.; Larsson, N.G.; Gustafsson, C.M.; Falkenberg, M. In vivo mutagenesis reveals that OriL is essential for mitochondrial DNA replication. EMBO Rep. 2012, 13, 1130–1137. [Google Scholar] [CrossRef]
  32. Nicholls, T.J.; Nadalutti, C.A.; Motori, E.; Sommerville, E.W.; Gorman, G.S.; Basu, S.; Hoberg, E.; Turnbull, D.M.; Chinnery, P.F.; Larsson, N.-G.; et al. Topoisomerase 3α Is Required for Decatenation and Segregation of Human mtDNA. Mol. Cell 2018, 69, 9–23.e6. [Google Scholar] [CrossRef]
  33. Erdinc, D.; Rodríguez-Luis, A.; Fassad, M.R.; Mackenzie, S.; Watson, C.M.; Valenzuela, S.; Xie, X.; Menger, K.E.; Sergeant, K.; Craig, K.; et al. Pathological variants in TOP3A cause distinct disorders of mitochondrial and nuclear genome stability. EMBO Mol. Med. 2023, 15, e16775. [Google Scholar] [CrossRef]
  34. Miller, F.J.; Rosenfeldt, F.L.; Zhang, C.; Linnane, A.W.; Nagley, P. Precise determination of mitochondrial DNA copy number in human skeletal and cardiac muscle by a PCR-based assay: Lack of change of copy number with age. Nucleic Acids Res. 2003, 31, e61. [Google Scholar] [CrossRef]
  35. Kotrys, A.V.; Durham, T.J.; Guo, X.A.; Vantaku, V.R.; Parangi, S.; Mootha, V.K. Single-cell analysis reveals context-dependent, cell-level selection of mtDNA. Nature 2024, 629, 458–466. [Google Scholar] [CrossRef] [PubMed]
  36. Greaves, L.C.; Nooteboom, M.; Elson, J.L.; Tuppen, H.A.; Taylor, G.A.; Commane, D.M.; Arasaradnam, R.P.; Khrapko, K.; Taylor, R.W.; Kirkwood, T.B.; et al. Clonal expansion of early to mid-life mitochondrial DNA point mutations drives mitochondrial dysfunction during human ageing. PLoS Genet. 2014, 10, e1004620. [Google Scholar] [CrossRef]
  37. Picard, M.; Zhang, J.; Hancock, S.; Derbeneva, O.; Golhar, R.; Golik, P.; O’Hearn, S.; Levy, S.; Potluri, P.; Lvova, M.; et al. Progressive increase in mtDNA 3243A>G heteroplasmy causes abrupt transcriptional reprogramming. Proc. Natl. Acad. Sci. USA 2014, 111, E4033–E4042. [Google Scholar] [CrossRef] [PubMed]
  38. Stewart, J.B.; Chinnery, P.F. The dynamics of mitochondrial DNA heteroplasmy: Implications for human health and disease. Nat. Rev. Genet. 2015, 16, 530–542. [Google Scholar] [CrossRef]
  39. Pickett, S.J.; Taylor, R.W.; McFarland, R. Fit for purpose: Selecting the best mitochondrial DNA for the job. Cell Metab. 2024, 36, 1436–1438. [Google Scholar] [CrossRef] [PubMed]
  40. Gupta, R.; Kanai, M.; Durham, T.J.; Tsuo, K.; McCoy, J.G.; Kotrys, A.V.; Zhou, W.; Chinnery, P.F.; Karczewski, K.J.; Calvo, S.E.; et al. Nuclear genetic control of mtDNA copy number and heteroplasmy in humans. Nature 2023, 620, 839–848, Erratum in Nature 2024, 630, E10. [Google Scholar] [CrossRef]
  41. Medeiros, T.C.; Thomas, R.L.; Ghillebert, R.; Graef, M. Autophagy balances mtDNA synthesis and degradation by DNA polymerase POLG during starvation. J. Cell Biol. 2018, 217, 1601–1611. [Google Scholar] [CrossRef]
  42. Copeland, W.C. Defects in mitochondrial DNA replication and human disease. Crit. Rev. Biochem. Mol. Biol. 2012, 47, 64–74. [Google Scholar] [CrossRef]
  43. Taylor, S.D.; Zhang, H.; Eaton, J.S.; Rodeheffer, M.S.; Lebedeva, M.A.; O’Rourke, T.W.; Siede, W.; Shadel, G.S. The conserved Mec1/Rad53 nuclear checkpoint pathway regulates mitochondrial DNA copy number in Saccharomyces cerevisiae. Mol. Biol. Cell 2005, 16, 3010–3018. [Google Scholar] [CrossRef] [PubMed]
  44. Young, M.J.; Copeland, W.C. Human mitochondrial DNA replication machinery and disease. Curr. Opin. Genet. Dev. 2016, 38, 52–62. [Google Scholar] [CrossRef] [PubMed]
  45. El-Hattab, A.W.; Scaglia, F. Mitochondrial DNA depletion syndromes: Review and updates of genetic basis, manifestations, and therapeutic options. Neurotherapeutics 2013, 10, 186–198. [Google Scholar] [CrossRef] [PubMed]
  46. Gustafson, M.A.; McCormick, E.M.; Perera, L.; Longley, M.J.; Bai, R.; Kong, J.; Dulik, M.; Shen, L.; Goldstein, A.C.; McCormack, S.E.; et al. Mitochondrial single-stranded DNA binding protein novel de novo SSBP1 mutation in a child with single large-scale mtDNA deletion (SLSMD) clinically manifesting as Pearson, Kearns-Sayre, and Leigh syndromes. PLoS ONE 2019, 14, e0221829. [Google Scholar] [CrossRef]
  47. Longley, M.J.; Clark, S.; Yu Wai Man, C.; Hudson, G.; Durham, S.E.; Taylor, R.W.; Nightingale, S.; Turnbull, D.M.; Copeland, W.C.; Chinnery, P.F. Mutant POLG2 disrupts DNA polymerase gamma subunits and causes progressive external ophthalmoplegia. Am. J. Hum. Genet. 2006, 78, 1026–1034. [Google Scholar] [CrossRef]
  48. Van Dyck, E.; Foury, F.; Stillman, B.; Brill, S.J. A single-stranded DNA binding protein required for mitochondrial DNA replication in S. cerevisiae is homologous to E. coli SSB. EMBO J. 1992, 11, 3421–3430. [Google Scholar] [CrossRef]
  49. Van Goethem, G.; Dermaut, B.; Löfgren, A.; Martin, J.J.; Van Broeckhoven, C. Mutation of POLG is associated with progressive external ophthalmoplegia characterized by mtDNA deletions. Nat. Genet. 2001, 28, 211–212. [Google Scholar] [CrossRef]
  50. Khan, N.A.; Govindaraj, P.; Meena, A.K.; Thangaraj, K. Mitochondrial disorders: Challenges in diagnosis & treatment. Indian J. Med. Res. 2015, 141, 13–26. [Google Scholar] [CrossRef]
  51. Copeland, W.C. Inherited mitochondrial diseases of DNA replication. Annu. Rev. Med. 2008, 59, 131–146. [Google Scholar] [CrossRef]
  52. Wortmann, S.B.; Zweers-van Essen, H.; Rodenburg, R.J.; van den Heuvel, L.P.; de Vries, M.C.; Rasmussen-Conrad, E.; Smeitink, J.A.; Morava, E. Mitochondrial energy production correlates with the age-related BMI. Pediatr. Res. 2009, 65, 103–108. [Google Scholar] [CrossRef] [PubMed]
  53. Morava, E.; Rodenburg, R.; van Essen, H.Z.; De Vries, M.; Smeitink, J. Dietary intervention and oxidative phosphorylation capacity. J. Inherit. Metab. Dis. 2006, 29, 589. [Google Scholar] [CrossRef] [PubMed]
  54. Aldossary, A.M.; Tawfik, E.A.; Alomary, M.N.; Alsudir, S.A.; Alfahad, A.J.; Alshehri, A.A.; Almughem, F.A.; Mohammed, R.Y.; Alzaydi, M.M. Recent advances in mitochondrial diseases: From molecular insights to therapeutic perspectives. Saudi Pharm. J. 2022, 30, 1065–1078. [Google Scholar] [CrossRef]
  55. Gorman, G.S.; Chinnery, P.F.; DiMauro, S.; Hirano, M.; Koga, Y.; McFarland, R.; Suomalainen, A.; Thorburn, D.R.; Zeviani, M.; Turnbull, D.M. Mitochondrial diseases. Nat. Rev. Dis. Primers 2016, 2, 16080. [Google Scholar] [CrossRef]
  56. Gorman, G.S.; Schaefer, A.M.; Ng, Y.; Gomez, N.; Blakely, E.L.; Alston, C.L.; Feeney, C.; Horvath, R.; Yu-Wai-Man, P.; Chinnery, P.F.; et al. Prevalence of nuclear and mitochondrial DNA mutations related to adult mitochondrial disease. Ann. Neurol. 2015, 77, 753–759. [Google Scholar] [CrossRef] [PubMed]
  57. Schon, K.R.; Horvath, R.; Wei, W.; Calabrese, C.; Tucci, A.; Ibañez, K.; Ratnaike, T.; Pitceathly, R.D.S.; Bugiardini, E.; Quinlivan, R.; et al. Use of whole genome sequencing to determine genetic basis of suspected mitochondrial disorders: Cohort study. BMJ 2021, 375, e066288. [Google Scholar] [CrossRef]
  58. Finsterer, J.; Scorza, F.A. Renal manifestations of primary mitochondrial disorders. Biomed. Rep. 2017, 6, 487–494. [Google Scholar] [CrossRef]
  59. Muraresku, C.C.; McCormick, E.M.; Falk, M.J. Mitochondrial Disease: Advances in clinical diagnosis, management, therapeutic development, and preventative strategies. Curr. Genet. Med. Rep. 2018, 6, 62–72. [Google Scholar] [CrossRef]
  60. Graziewicz, M.A.; Longley, M.J.; Copeland, W.C. DNA polymerase gamma in mitochondrial DNA replication and repair. Chem. Rev. 2006, 106, 383–405. [Google Scholar] [CrossRef]
  61. Ropp, P.A.; Copeland, W.C. Cloning and characterization of the human mitochondrial DNA polymerase, DNA polymerase gamma. Genomics 1996, 36, 449–458. [Google Scholar] [CrossRef]
  62. Young, M.J.; Copeland, W.C. Mitochondrial Disorders Associated with the Mitochondrial DNA Polymerase g: A Focus on Intersubunit Interactions. In Mitochondrial Disorders Caused by Nuclear Genes; Wong, L.-J.C., Ed.; Springer: New York, NY, USA, 2013; pp. 49–72. [Google Scholar]
  63. Wang, Y.; Kaguni, L.S. Baculovirus expression reconstitutes Drosophila mitochondrial DNA polymerase. J. Biol. Chem. 1999, 274, 28972–28977. [Google Scholar] [CrossRef] [PubMed]
  64. Hance, N.; Ekstrand, M.I.; Trifunovic, A. Mitochondrial DNA polymerase gamma is essential for mammalian embryogenesis. Hum. Mol. Genet. 2005, 14, 1775–1783. [Google Scholar] [CrossRef] [PubMed]
  65. Humble, M.M.; Young, M.J.; Foley, J.F.; Pandiri, A.R.; Travlos, G.S.; Copeland, W.C. Polg2 is essential for mammalian embryogenesis and is required for mtDNA maintenance. Hum. Mol. Genet. 2013, 22, 1017–1025. [Google Scholar] [CrossRef] [PubMed]
  66. Longley, M.J.; Prasad, R.; Srivastava, D.K.; Wilson, S.H.; Copeland, W.C. Identification of 5′-deoxyribose phosphate lyase activity in human DNA polymerase gamma and its role in mitochondrial base excision repair in vitro. Proc. Natl. Acad. Sci. USA 1998, 95, 12244–12248. [Google Scholar] [CrossRef]
  67. Lee, Y.S.; Kennedy, W.D.; Yin, Y.W. Structural insight into processive human mitochondrial DNA synthesis and disease-related polymerase mutations. Cell 2009, 139, 312–324. [Google Scholar] [CrossRef]
  68. Szymanski, M.R.; Kuznetsov, V.B.; Shumate, C.; Meng, Q.; Lee, Y.S.; Patel, G.; Patel, S.; Yin, Y.W. Structural basis for processivity and antiviral drug toxicity in human mitochondrial DNA replicase. EMBO J. 2015, 34, 1959–1970. [Google Scholar] [CrossRef]
  69. Longley, M.J.; Nguyen, D.; Kunkel, T.A.; Copeland, W.C. The fidelity of human DNA polymerase gamma with and without exonucleolytic proofreading and the p55 accessory subunit. J. Biol. Chem. 2001, 276, 38555–38562. [Google Scholar] [CrossRef]
  70. Zheng, W.; Khrapko, K.; Coller, H.A.; Thilly, W.G.; Copeland, W.C. Origins of human mitochondrial point mutations as DNA polymerase gamma-mediated errors. Mutat. Res. 2006, 599, 11–20. [Google Scholar] [CrossRef]
  71. Ju, Y.S.; Alexandrov, L.B.; Gerstung, M.; Martincorena, I.; Nik-Zainal, S.; Ramakrishna, M.; Davies, H.R.; Papaemmanuil, E.; Gundem, G.; Shlien, A.; et al. Origins and functional consequences of somatic mitochondrial DNA mutations in human cancer. eLife 2014, 3, e02935. [Google Scholar] [CrossRef]
  72. King, D.E.; Copeland, W.C. DNA Repair Pathways in the Mitochondria. DNA Repair 2025, 146, 103814. [Google Scholar] [CrossRef] [PubMed]
  73. Kennedy, S.R.; Salk, J.J.; Schmitt, M.W.; Loeb, L.A. Ultra-sensitive sequencing reveals an age-related increase in somatic mitochondrial mutations that are inconsistent with oxidative damage. PLoS Genet. 2013, 9, e1003794. [Google Scholar] [CrossRef] [PubMed]
  74. Itsara, L.S.; Kennedy, S.R.; Fox, E.J.; Yu, S.; Hewitt, J.J.; Sanchez-Contreras, M.; Cardozo-Pelaez, F.; Pallanck, L.J. Oxidative stress is not a major contributor to somatic mitochondrial DNA mutations. PLoS Genet. 2014, 10, e1003974. [Google Scholar] [CrossRef] [PubMed]
  75. Valente, W.J.; Ericson, N.G.; Long, A.S.; White, P.A.; Marchetti, F.; Bielas, J.H. Mitochondrial DNA exhibits resistance to induced point and deletion mutations. Nucleic Acids Res. 2016, 44, 8513–8524. [Google Scholar] [CrossRef]
  76. Leuthner, T.C.; Meyer, J.N. Mitochondrial DNA Mutagenesis: Feature of and Biomarker for Environmental Exposures and Aging. Curr. Environ. Health Rep. 2021, 8, 294–308. [Google Scholar] [CrossRef]
  77. Longley, M.J.; Ropp, P.A.; Lim, S.E.; Copeland, W.C. Characterization of the native and recombinant catalytic subunit of human DNA polymerase gamma: Identification of residues critical for exonuclease activity and dideoxynucleotide sensitivity. Biochemistry 1998, 37, 10529–10539. [Google Scholar] [CrossRef]
  78. Spelbrink, J.N.; Toivonen, J.M.; Hakkaart, G.A.; Kurkela, J.M.; Cooper, H.M.; Lehtinen, S.K.; Lecrenier, N.; Back, J.W.; Speijer, D.; Foury, F.; et al. In vivo functional analysis of the human mitochondrial DNA polymerase POLG expressed in cultured human cells. J. Biol. Chem. 2000, 275, 24818–24828. [Google Scholar] [CrossRef]
  79. Bensch, K.G.; Degraaf, W.; Hansen, P.A.; Zassenhaus, H.P.; Corbett, J.A. A transgenic model to study the pathogenesis of somatic mtDNA mutation accumulation in beta-cells. Diabetes Obes. Metab. 2007, 9 (Suppl. S2), 74–80. [Google Scholar] [CrossRef]
  80. Bensch, K.G.; Mott, J.L.; Chang, S.W.; Hansen, P.A.; Moxley, M.A.; Chambers, K.T.; de Graaf, W.; Zassenhaus, H.P.; Corbett, J.A. Selective mtDNA mutation accumulation results in beta-cell apoptosis and diabetes development. Am. J. Physiol. 2009, 296, E672–E680. [Google Scholar] [CrossRef]
  81. Zhang, D.; Mott, J.L.; Chang, S.W.; Denniger, G.; Feng, Z.; Zassenhaus, H.P. Construction of transgenic mice with tissue-specific acceleration of mitochondrial DNA mutagenesis. Genomics 2000, 69, 151–161. [Google Scholar] [CrossRef]
  82. Zhang, D.; Mott, J.L.; Farrar, P.; Ryerse, J.S.; Chang, S.W.; Stevens, M.; Denniger, G.; Zassenhaus, H.P. Mitochondrial DNA mutations activate the mitochondrial apoptotic pathway and cause dilated cardiomyopathy. Cardiovasc. Res. 2003, 57, 147–157. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, D.; Mott, J.L.; Chang, S.W.; Stevens, M.; Mikolajczak, P.; Zassenhaus, H.P. Mitochondrial DNA mutations activate programmed cell survival in the mouse heart. Am. J. Physiol. Heart Circ. Physiol. 2005, 288, H2476–H2483. [Google Scholar] [CrossRef] [PubMed]
  84. Trifunovic, A.; Wredenberg, A.; Falkenberg, M.; Spelbrink, J.N.; Rovio, A.T.; Bruder, C.E.; Bohlooly, Y.M.; Gidlof, S.; Oldfors, A.; Wibom, R.; et al. Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature 2004, 429, 417–423. [Google Scholar] [CrossRef]
  85. Kujoth, G.C.; Hiona, A.; Pugh, T.D.; Someya, S.; Panzer, K.; Wohlgemuth, S.E.; Hofer, T.; Seo, A.Y.; Sullivan, R.; Jobling, W.A.; et al. Mitochondrial DNA mutations, oxidative stress, and apoptosis in mammalian aging. Science 2005, 309, 481–484. [Google Scholar] [CrossRef]
  86. Vermulst, M.; Bielas, J.H.; Kujoth, G.C.; Ladiges, W.C.; Rabinovitch, P.S.; Prolla, T.A.; Loeb, L.A. Mitochondrial point mutations do not limit the natural lifespan of mice. Nat. Genet. 2007, 39, 540–543. [Google Scholar] [CrossRef]
  87. Williams, S.L.; Huang, J.; Edwards, Y.J.; Ulloa, R.H.; Dillon, L.M.; Prolla, T.A.; Vance, J.M.; Moraes, C.T.; Zuchner, S. The mtDNA mutation spectrum of the progeroid Polg mutator mouse includes abundant control region multimers. Cell Metab. 2010, 12, 675–682. [Google Scholar] [CrossRef]
  88. Vermulst, M.; Bielas, J.H.; Loeb, L.A. Quantification of random mutations in the mitochondrial genome. Methods 2008, 46, 263–268. [Google Scholar] [CrossRef]
  89. Hämäläinen, R.H.; Landoni, J.C.; Ahlqvist, K.J.; Goffart, S.; Ryytty, S.; Rahman, M.O.; Brilhante, V.; Icay, K.; Hautaniemi, S.; Wang, L.; et al. Defects in mtDNA replication challenge nuclear genome stability through nucleotide depletion and provide a unifying mechanism for mouse progerias. Nat. Metab. 2019, 1, 958–965, Erratum in Nat. Metab. 2020, 2, 793. [Google Scholar] [CrossRef]
  90. Gorman, E.; Dai, H.; Feng, Y.; Craigen, W.J.; Chen, D.C.Y.; Xia, F.; Meng, L.; Liu, P.; Rigobello, R.; Neogi, A.; et al. Experiences from dual genome next-generation sequencing panel testing for mitochondrial disorders: A comprehensive molecular diagnosis. Front. Genet. 2025, 16, 1488956. [Google Scholar] [CrossRef]
  91. Barca, E.; Long, Y.; Cooley, V.; Schoenaker, R.; Emmanuele, V.; DiMauro, S.; Cohen, B.H.; Karaa, A.; Vladutiu, G.D.; Haas, R.; et al. Mitochondrial diseases in North America: An analysis of the NAMDC Registry. Neurol. Genet. 2020, 6, e402. [Google Scholar] [CrossRef]
  92. Rahman, S.; Copeland, W.C. POLG-related disorders and their neurological manifestations. Nat. Rev. Neurol. 2019, 15, 40–52. [Google Scholar] [CrossRef]
  93. Cohen, B.H.; Chinnery, P.F.; Copeland, W.C. POLG-Related Disorders. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Mirzaa, G., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 2024. [Google Scholar]
  94. Lujan, S.A.; Longley, M.J.; Humble, M.H.; Lavender, C.A.; Burkholder, A.; Blakely, E.L.; Alston, C.L.; Gorman, G.S.; Turnbull, D.M.; McFarland, R.; et al. Ultrasensitive deletion detection links mitochondrial DNA replication, disease, and aging. Genome Biol. 2020, 21, 248. [Google Scholar] [CrossRef]
  95. Chan, S.S.; Longley, M.J.; Copeland, W.C. The common A467T mutation in the human mitochondrial DNA polymerase (POLG) compromises catalytic efficiency and interaction with the accessory subunit. J. Biol. Chem. 2005, 280, 31341–31346. [Google Scholar] [CrossRef]
  96. Van Goethem, G.; Martin, J.J.; Dermaut, B.; Lofgren, A.; Wibail, A.; Ververken, D.; Tack, P.; Dehaene, I.; Van Zandijcke, M.; Moonen, M.; et al. Recessive POLG mutations presenting with sensory and ataxic neuropathy in compound heterozygote patients with progressive external ophthalmoplegia. Neuromuscul. Disord. 2003, 13, 133–142. [Google Scholar] [CrossRef]
  97. Chan, S.S.; Longley, M.J.; Copeland, W.C. Modulation of the W748S mutation in DNA polymerase gamma by the E1143G polymorphismin mitochondrial disorders. Hum. Mol. Genet. 2006, 15, 3473–3483. [Google Scholar] [CrossRef]
  98. Kasiviswanathan, R.; Longley, M.J.; Chan, S.S.; Copeland, W.C. Disease mutations in the human mitochondrial DNA polymerase thumb subdomain impart severe defects in mitochondrial DNA replication. J. Biol. Chem. 2009, 284, 19501–19510. [Google Scholar] [CrossRef] [PubMed]
  99. Ferrari, G.; Lamantea, E.; Donati, A.; Filosto, M.; Briem, E.; Carrara, F.; Parini, R.; Simonati, A.; Santer, R.; Zeviani, M. Infantile hepatocerebral syndromes associated with mutations in the mitochondrial DNA polymerase-gammaA. Brain 2005, 128, 723–731. [Google Scholar] [CrossRef] [PubMed]
  100. DeBalsi, K.L.; Longley, M.J.; Hoff, K.E.; Copeland, W.C. Synergistic Effects of the in cis T251I and P587L Mitochondrial DNA Polymerase gamma Disease Mutations. J. Biol. Chem. 2017, 292, 4198–4209. [Google Scholar] [CrossRef] [PubMed]
  101. Graziewicz, M.A.; Longley, M.J.; Bienstock, R.J.; Zeviani, M.; Copeland, W.C. Structure-function defects of human mitochondrial DNA polymerase in autosomal dominant progressive external ophthalmoplegia. Nat. Struct. Mol. Biol. 2004, 11, 770–776. [Google Scholar] [CrossRef]
  102. Ponamarev, M.V.; Longley, M.J.; Nguyen, D.; Kunkel, T.A.; Copeland, W.C. Active site mutation in DNA polymerase gamma associated with progressive external ophthalmoplegia causes error-prone DNA synthesis. J. Biol. Chem. 2002, 277, 15225–15228. [Google Scholar] [CrossRef]
  103. Silva-Pinheiro, P.; Pardo-Hernández, C.; Reyes, A.; Tilokani, L.; Mishra, A.; Cerutti, R.; Li, S.; Rozsivalova, D.H.; Valenzuela, S.; Dogan, S.A.; et al. DNA polymerase gamma mutations that impair holoenzyme stability cause catalytic subunit depletion. Nucleic Acids Res. 2021, 49, 5230–5248, Erratum in Nucleic Acids Res. 2021, 49, 10803. [Google Scholar] [CrossRef] [PubMed]
  104. Riccio, A.A.; Brannon, A.J.; Krahn, J.M.; Bouvette, J.; Williams, J.G.; Borgnia, M.J.; Copeland, W.C. Coordinated DNA polymerization by Polgamma and the region of LonP1 regulated proteolysis. Nucleic Acids Res. 2024, 52, 7863–7875. [Google Scholar] [CrossRef] [PubMed]
  105. Lim, S.E.; Longley, M.J.; Copeland, W.C. The Mitochondrial p55 Accessory Subunit of Human DNA Polymerase γ Enhances DNA Binding, Promotes Processive DNA Synthesis, and Confers N-Ethylmaleimide Resistance. J. Biol. Chem. 1999, 274, 38197–38203. [Google Scholar] [CrossRef]
  106. Olson, M.W.; Wang, Y.; Elder, R.H.; Kaguni, L.S. Subunit structure of mitochondrial DNA polymerase from Drosophila embryos. Physical and immunological studies. J. Biol. Chem. 1995, 270, 28932–28937. [Google Scholar] [CrossRef]
  107. Wang, Y.; Farr, C.L.; Kaguni, L.S. Accessory subunit of mitochondrial DNA polymerase from Drosophila embryos. Cloning, molecular analysis, and association in the native enzyme. J. Biol. Chem. 1997, 272, 13640–13646. [Google Scholar] [CrossRef]
  108. Fan, L.; Sanschagrin, P.C.; Kaguni, L.S.; Kuhn, L.A. The accessory subunit of mtDNA polymerase shares structural homology with aminoacyl-tRNA synthetases: Implications for a dual role as a primer recognition factor and processivity clamp. Proc. Natl. Acad. Sci. USA 1999, 96, 9527–9532. [Google Scholar] [CrossRef]
  109. Carrodeguas, J.A.; Theis, K.; Bogenhagen, D.F.; Kisker, C. Crystal structure and deletion analysis show that the accessory subunit of mammalian DNA polymerase gamma, Pol gamma B, functions as a homodimer. Mol. Cell 2001, 7, 43–54. [Google Scholar] [CrossRef]
  110. Carrodeguas, J.A.; Kobayashi, R.; Lim, S.E.; Copeland, W.C.; Bogenhagen, D.F. The accessory subunit of Xenopus laevis mitochondrial DNA polymerase gamma increases processivity of the catalytic subunit of human DNA polymerase gamma and is related to class II aminoacyl-tRNA synthetases. Mol. Cell Biol. 1999, 19, 4039–4046. [Google Scholar] [CrossRef]
  111. Lee, Y.S.; Lee, S.; Demeler, B.; Molineux, I.J.; Johnson, K.A.; Yin, Y.W. Each monomer of the dimeric accessory protein for human mitochondrial DNA polymerase has a distinct role in conferring processivity. J. Biol. Chem. 2010, 285, 1490–1499. [Google Scholar] [CrossRef]
  112. Wojtaszek, J.L.; Hoff, K.E.; Longley, M.J.; Kaur, P.; Andres, S.N.; Wang, H.; Williams, R.S.; Copeland, W.C. Structure-specific roles for PolG2–DNA complexes in maintenance and replication of mitochondrial DNA. Nucleic Acids Res. 2023, 51, 9716–9732. [Google Scholar] [CrossRef]
  113. Farge, G.; Pham, X.H.; Holmlund, T.; Khorostov, I.; Falkenberg, M. The accessory subunit B of DNA polymerase gamma is required for mitochondrial replisome function. Nucleic Acids Res. 2007, 35, 902–911. [Google Scholar] [CrossRef] [PubMed]
  114. Walter, M.C.; Czermin, B.; Muller-Ziermann, S.; Bulst, S.; Stewart, J.D.; Hudson, G.; Schneiderat, P.; Abicht, A.; Holinski-Feder, E.; Lochmuller, H.; et al. Late-onset ptosis and myopathy in a patient with a heterozygous insertion in POLG2. J. Neurol. 2010, 257, 1517–1523. [Google Scholar] [CrossRef] [PubMed]
  115. Lehmann Urban, D.; Motlagh Scholle, L.; Alt, K.; Ludolph, A.C.; Rosenbohm, A. Camptocormia as a Novel Phenotype in a Heterozygous POLG2 Mutation. Diagnostics 2020, 10, 68. [Google Scholar] [CrossRef]
  116. Young, M.J.; Longley, M.J.; Li, F.Y.; Kasiviswanathan, R.; Wong, L.J.; Copeland, W.C. Biochemical analysis of human POLG2 variants associated with mitochondrial disease. Hum. Mol. Genet. 2011, 20, 3052–3066. [Google Scholar] [CrossRef]
  117. Craig, K.; Young, M.J.; Blakely, E.L.; Longley, M.J.; Turnbull, D.M.; Copeland, W.C.; Taylor, R.W. A p.R369G POLG2 mutation associated with adPEO and multiple mtDNA deletions causes decreased affinity between polymerase γ subunits. Mitochondrion 2012, 12, 313–319. [Google Scholar] [CrossRef]
  118. Young, M.J.; Humble, M.M.; DeBalsi, K.L.; Sun, K.Y.; Copeland, W.C. POLG2 disease variants: Analyses reveal a dominant negative heterodimer, altered mitochondrial localization and impaired respiratory capacity. Hum. Mol. Genet. 2015, 24, 5184–5197. [Google Scholar] [CrossRef]
  119. Varma, H.; Faust, P.L.; Iglesias, A.D.; Lagana, S.M.; Wou, K.; Hirano, M.; DiMauro, S.; Mansukani, M.M.; Hoff, K.E.; Nagy, P.L.; et al. Whole exome sequencing identifies a homozygous POLG2 missense variant in an infant with fulminant hepatic failure and mitochondrial DNA depletion. Eur. J. Med. Genet. 2016, 59, 540–545. [Google Scholar] [CrossRef]
  120. Hoff, K.E.; DeBalsi, K.L.; Sanchez-Quintero, M.J.; Longley, M.J.; Hirano, M.; Naini, A.B.; Copeland, W.C. Characterization of the human homozygous R182W POLG2 mutation in mitochondrial DNA depletion syndrome. PLoS ONE 2018, 13, e0203198. [Google Scholar] [CrossRef]
  121. Lee, S.J.; Kanwal, S.; Yoo, D.H.; Park, H.R.; Choi, B.O.; Chung, K.W. A POLG2 Homozygous Mutation in an Autosomal Recessive Epilepsy Family Without Ophthalmoplegia. J. Clin. Neurol. 2019, 15, 418–420. [Google Scholar] [CrossRef]
  122. Dosekova, P.; Dubiel, A.; Karlowicz, A.; Zietkiewicz, S.; Rydzanicz, M.; Habalova, V.; Pienkowski, V.M.; Skirkova, M.; Han, V.; Mosejova, A.; et al. Whole exome sequencing identifies a homozygous POLG2 missense variant in an adult patient presenting with optic atrophy, movement disorders, premature ovarian failure and mitochondrial DNA depletion. Eur. J. Med. Genet. 2020, 63, 103821. [Google Scholar] [CrossRef]
  123. Van Maldergem, L.; Besse, A.; De Paepe, B.; Blakely, E.L.; Appadurai, V.; Humble, M.M.; Piard, J.; Craig, K.; He, L.; Hella, P.; et al. POLG2 deficiency causes adult-onset syndromic sensory neuropathy, ataxia and parkinsonism. Ann. Clin. Transl. Neurol. 2017, 4, 4–14. [Google Scholar] [CrossRef] [PubMed]
  124. Valencia, C.A.; Wang, X.; Wang, J.; Peters, A.; Simmons, J.R.; Moran, M.C.; Mathur, A.; Husami, A.; Qian, Y.; Sheridan, R.; et al. Deep Sequencing Reveals Novel Genetic Variants in Children with Acute Liver Failure and Tissue Evidence of Impaired Energy Metabolism. PLoS ONE 2016, 11, e0156738. [Google Scholar] [CrossRef] [PubMed]
  125. Hou, Y.; Zhao, X.; Xie, Z.; Yu, M.; Lv, H.; Zhang, W.; Yuan, Y.; Wang, Z. Novel and recurrent nuclear gene variations in a cohort of Chinese progressive external ophthalmoplegia patients with multiple mtDNA deletions. Mol. Genet. Genom. Med. 2022, 10, e1921. [Google Scholar] [CrossRef] [PubMed]
  126. Borsche, M.; Dulovic-Mahlow, M.; Baumann, H.; Tunc, S.; Lüth, T.; Schaake, S.; Özcakir, S.; Westenberger, A.; Münchau, A.; Knappe, E.; et al. POLG2-Linked Mitochondrial Disease: Functional Insights from New Mutation Carriers and Review of the Literature. Cerebellum 2024, 23, 479–488. [Google Scholar] [CrossRef]
  127. Siegert, S.; Mindler, G.T.; Brücke, C.; Kranzl, A.; Patsch, J.; Ritter, M.; Janecke, A.R.; Vodopiutz, J. Expanding the Phenotype of the FAM149B1-Related Ciliopathy and Identification of Three Neurogenetic Disorders in a Single Family. Genes 2021, 12, 1648. [Google Scholar] [CrossRef]
  128. Kim, M.; Kim, A.R.; Kim, J.S.; Park, J.; Youn, J.; Ahn, J.H.; Mun, J.K.; Lee, C.; Kim, N.S.; Kim, N.K.D.; et al. Clarification of undiagnosed ataxia using whole-exome sequencing with clinical implications. Park. Relat. Disord. 2020, 80, 58–64. [Google Scholar] [CrossRef]
  129. Do, Y.; Matsuda, S.; Inatomi, T.; Nakada, K.; Yasukawa, T.; Kang, D. The accessory subunit of human DNA polymerase gamma is required for mitochondrial DNA maintenance and is able to stabilize the catalytic subunit. Mitochondrion 2020, 53, 133–139. [Google Scholar] [CrossRef]
  130. Johnson, A.A.; Tsai, Y.-c.; Graves, S.W.; Johnson, K.A. Human Mitochondrial DNA Polymerase Holoenzyme: Reconstitution and Characterization. Biochemistry 2000, 39, 1702–1708. [Google Scholar] [CrossRef]
  131. Iyengar, B.; Luo, N.; Farr, C.L.; Kaguni, L.S.; Campos, A.R. The accessory subunit of DNA polymerase gamma is essential for mitochondrial DNA maintenance and development in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 2002, 99, 4483–4488. [Google Scholar] [CrossRef]
  132. Lee, S.K.; Zhao, M.H.; Zheng, Z.; Kwon, J.W.; Liang, S.; Kim, S.H.; Kim, N.H.; Cui, X.S. Polymerase subunit gamma 2 affects porcine oocyte maturation and subsequent embryonic development. Theriogenology 2015, 83, 121–130. [Google Scholar] [CrossRef]
  133. Di Re, M.; Sembongi, H.; He, J.; Reyes, A.; Yasukawa, T.; Martinsson, P.; Bailey, L.J.; Goffart, S.; Boyd-Kirkup, J.D.; Wong, T.S.; et al. The accessory subunit of mitochondrial DNA polymerase gamma determines the DNA content of mitochondrial nucleoids in human cultured cells. Nucleic Acids Res. 2009, 37, 5701–5713. [Google Scholar] [CrossRef]
  134. Sigal, N.; Delius, H.; Kornberg, T.; Gefter, M.L.; Alberts, B. A DNA-unwinding protein isolated from Escherichia coli: Its interaction with DNA and with DNA polymerases. Proc. Natl. Acad. Sci. USA 1972, 69, 3537–3541. [Google Scholar] [CrossRef]
  135. Wold, M.S. Replication protein A: A heterotrimeric, single-stranded DNA-binding protein required for eukaryotic DNA metabolism. Annu. Rev. Biochem. 1997, 66, 61–92. [Google Scholar] [CrossRef]
  136. Lohman, T.M.; Ferrari, M.E. Escherichia coli single-stranded DNA-binding protein: Multiple DNA-binding modes and cooperativities. Annu. Rev. Biochem. 1994, 63, 527–570. [Google Scholar] [CrossRef]
  137. Pestryakov, P.E.; Lavrik, O.I. Mechanisms of single-stranded DNA-binding protein functioning in cellular DNA metabolism. Biochemistry 2008, 73, 1388–1404. [Google Scholar] [CrossRef] [PubMed]
  138. Perales, C.; Cava, F.; Meijer, W.J.; Berenguer, J. Enhancement of DNA, cDNA synthesis and fidelity at high temperatures by a dimeric single-stranded DNA-binding protein. Nucleic Acids Res. 2003, 31, 6473–6480. [Google Scholar] [CrossRef] [PubMed]
  139. Chase, J.W.; Williams, K.R. Single-stranded DNA binding proteins required for DNA replication. Annu. Rev. Biochem. 1986, 55, 103–136. [Google Scholar] [CrossRef]
  140. Zou, Y.; Liu, Y.; Wu, X.; Shell, S.M. Functions of human replication protein A (RPA): From DNA replication to DNA damage and stress responses. J. Cell Physiol. 2006, 208, 267–273. [Google Scholar] [CrossRef]
  141. Shereda, R.D.; Kozlov, A.G.; Lohman, T.M.; Cox, M.M.; Keck, J.L. SSB as an organizer/mobilizer of genome maintenance complexes. Crit. Rev. Biochem. Mol. Biol. 2008, 43, 289–318. [Google Scholar] [CrossRef]
  142. Clayton, D.A. Replication of animal mitochondrial DNA. Cell 1982, 28, 693–705. [Google Scholar] [CrossRef]
  143. Farr, C.L.; Wang, Y.; Kaguni, L.S. Functional interactions of mitochondrial DNA polymerase and single-stranded DNA-binding protein. Template-primer DNA binding and initiation and elongation of DNA strand synthesis. J. Biol. Chem. 1999, 274, 14779–14785. [Google Scholar] [CrossRef] [PubMed]
  144. Korhonen, J.A.; Gaspari, M.; Falkenberg, M. TWINKLE Has 5′→3′ DNA helicase activity and is specifically stimulated by mitochondrial single-stranded DNA-binding protein. J. Biol. Chem. 2003, 278, 48627–48632. [Google Scholar] [CrossRef] [PubMed]
  145. Shutt, T.E.; Gray, M.W. Bacteriophage origins of mitochondrial replication and transcription proteins. Trends Genet. 2006, 22, 90–95. [Google Scholar] [CrossRef] [PubMed]
  146. Tiranti, V.; Rocchi, M.; DiDonato, S.; Zeviani, M. Cloning of human and rat cDNAs encoding the mitochondrial single-stranded DNA-binding protein (SSB). Gene 1993, 126, 219–225. [Google Scholar] [CrossRef]
  147. Riccio, A.A.; Bouvette, J.; Pedersen, L.C.; Somai, S.; Dutcher, R.C.; Borgnia, M.J.; Copeland, W.C. Structures of the mitochondrial single-stranded DNA binding protein with DNA and DNA polymerase γ. Nucleic Acids Res. 2024, 52, 10329–10340. [Google Scholar] [CrossRef]
  148. Yang, C.; Curth, U.; Urbanke, C.; Kang, C. Crystal structure of human mitochondrial single-stranded DNA binding protein at 2.4 A resolution. Nat. Struct. Biol. 1997, 4, 153–157. [Google Scholar] [CrossRef]
  149. Qian, Y.; Johnson, K.A. The human mitochondrial single-stranded DNA-binding protein displays distinct kinetics and thermodynamics of DNA binding and exchange. J. Biol. Chem. 2017, 292, 13068–13084. [Google Scholar] [CrossRef]
  150. Chrysogelos, S.; Griffith, J. Escherichia coli single-strand binding protein organizes single-stranded DNA in nucleosome-like units. Proc. Natl. Acad. Sci. USA 1982, 79, 5803–5807. [Google Scholar] [CrossRef]
  151. Lohman, T.M.; Bujalowski, W. Negative cooperativity within individual tetramers of Escherichia coli single strand binding protein is responsible for the transition between the (SSB)35 and (SSB)56 DNA binding modes. Biochemistry 1988, 27, 2260–2265. [Google Scholar] [CrossRef]
  152. Lohman, T.M.; Bujalowski, W.; Overman, L.B.; Wei, T.F. Interactions of the E. coli single strand binding (SSB) protein with ss nucleic acids. Binding mode transitions and equilibrium binding studies. Biochem. Pharmacol. 1988, 37, 1781–1782. [Google Scholar] [CrossRef]
  153. Kaur, P.; Longley, M.J.; Pan, H.; Wang, H.; Copeland, W.C. Single-molecule DREEM imaging reveals DNA wrapping around human mitochondrial single-stranded DNA binding protein. Nucleic Acids Res. 2018, 46, 11287–11302. [Google Scholar] [CrossRef] [PubMed]
  154. Curth, U.; Genschel, J.; Urbanke, C.; Greipel, J. In Vitro and in Vivo Function of the C-Terminus of Escherichia coli Single-Stranded DNA Binding Protein. Nucleic Acids Res. 1996, 24, 2706–2711. [Google Scholar] [CrossRef] [PubMed]
  155. Williams, K.R.; Spicer, E.K.; LoPresti, M.B.; Guggenheimer, R.A.; Chase, J.W. Limited proteolysis studies on the Escherichia coli single-stranded DNA binding protein. Evidence for a functionally homologous domain in both the Escherichia coli and T4 DNA binding proteins. J. Biol. Chem. 1983, 258, 3346–3355. [Google Scholar] [CrossRef]
  156. Kaur, P.; Longley, M.J.; Pan, H.; Wang, W.; Countryman, P.; Wang, H.; Copeland, W.C. Single-molecule level structural dynamics of DNA unwinding by human mitochondrial Twinkle helicase. J. Biol. Chem. 2020, 295, 5564–5576. [Google Scholar] [CrossRef]
  157. Ciesielski, G.L.; Bermek, O.; Rosado-Ruiz, F.A.; Hovde, S.L.; Neitzke, O.J.; Griffith, J.D.; Kaguni, L.S. Mitochondrial Single-stranded DNA-binding Proteins Stimulate the Activity of DNA Polymerase γ by Organization of the Template DNA. J. Biol. Chem. 2015, 290, 28697–28707. [Google Scholar] [CrossRef]
  158. Korhonen, J.A.; Pham, X.H.; Pellegrini, M.; Falkenberg, M. Reconstitution of a minimal mtDNA replisome in vitro. EMBO J. 2004, 23, 2423–2429. [Google Scholar] [CrossRef]
  159. Gustafson, M.A.; Perera, L.; Shi, M.; Copeland, W.C. Mechanisms of SSBP1 variants in mitochondrial disease: Molecular dynamics simulations reveal stable tetramers with altered DNA binding surfaces. DNA Repair 2021, 107, 103212. [Google Scholar] [CrossRef]
  160. Maier, D.; Farr, C.L.; Poeck, B.; Alahari, A.; Vogel, M.; Fischer, S.; Kaguni, L.S.; Schneuwly, S. Mitochondrial single-stranded DNA-binding protein is required for mitochondrial DNA replication and development in Drosophila melanogaster. Mol. Biol. Cell 2001, 12, 821–830. [Google Scholar] [CrossRef]
  161. Ruhanen, H.; Borrie, S.; Szabadkai, G.; Tyynismaa, H.; Jones, A.W.; Kang, D.; Taanman, J.W.; Yasukawa, T. Mitochondrial single-stranded DNA binding protein is required for maintenance of mitochondrial DNA and 7S DNA but is not required for mitochondrial nucleoid organisation. Biochim. Biophys. Acta 2010, 1803, 931–939. [Google Scholar] [CrossRef]
  162. Kullar, P.J.; Gomez-Duran, A.; Gammage, P.A.; Garone, C.; Minczuk, M.; Golder, Z.; Wilson, J.; Montoya, J.; Häkli, S.; Kärppä, M.; et al. Heterozygous SSBP1 start loss mutation co-segregates with hearing loss and the m.1555A>G mtDNA variant in a large multigenerational family. Brain 2018, 141, 55–62. [Google Scholar] [CrossRef]
  163. Jurkute, N.; Leu, C.; Pogoda, H.M.; Arno, G.; Robson, A.G.; Nürnberg, G.; Altmüller, J.; Thiele, H.; Motameny, S.; Toliat, M.R.; et al. SSBP1 mutations in dominant optic atrophy with variable retinal degeneration. Ann. Neurol. 2019, 86, 368–383. [Google Scholar] [CrossRef]
  164. Del Dotto, V.; Ullah, F.; Di Meo, I.; Magini, P.; Gusic, M.; Maresca, A.; Caporali, L.; Palombo, F.; Tagliavini, F.; Baugh, E.H.; et al. SSBP1 mutations cause mtDNA depletion underlying a complex optic atrophy disorder. J. Clin. Investig. 2020, 130, 108–125. [Google Scholar] [CrossRef]
  165. Piro-Mégy, C.; Sarzi, E.; Tarrés-Solé, A.; Péquignot, M.; Hensen, F.; Quilès, M.; Manes, G.; Chakraborty, A.; Sénéchal, A.; Bocquet, B.; et al. Dominant mutations in mtDNA maintenance gene SSBP1 cause optic atrophy and foveopathy. J. Clin. Investig. 2020, 130, 143–156. [Google Scholar] [CrossRef]
  166. Goffart, S.; Cooper, H.M.; Tyynismaa, H.; Wanrooij, S.; Suomalainen, A.; Spelbrink, J.N. Twinkle mutations associated with autosomal dominant progressive external ophthalmoplegia lead to impaired helicase function and in vivo mtDNA replication stalling. Hum. Mol. Genet. 2009, 18, 328–340. [Google Scholar] [CrossRef]
  167. Wanrooij, S.; Goffart, S.; Pohjoismäki, J.L.; Yasukawa, T.; Spelbrink, J.N. Expression of catalytic mutants of the mtDNA helicase Twinkle and polymerase POLG causes distinct replication stalling phenotypes. Nucleic Acids Res. 2007, 35, 3238–3251. [Google Scholar] [CrossRef] [PubMed]
  168. Jiang, M.; Xie, X.; Zhu, X.; Jiang, S.; Milenkovic, D.; Misic, J.; Shi, Y.; Tandukar, N.; Li, X.; Atanassov, I.; et al. The mitochondrial single-stranded DNA binding protein is essential for initiation of mtDNA replication. Sci. Adv. 2021, 7, eabf8631. [Google Scholar] [CrossRef] [PubMed]
  169. Gorbalenya, A.E.; Koonin, E.V. Helicases: Amino acid sequence comparisons and structure-function relationships. Curr. Opin. Struct. Biol. 1993, 3, 419–429. [Google Scholar] [CrossRef]
  170. Fairman-Williams, M.E.; Guenther, U.-P.; Jankowsky, E. SF1 and SF2 helicases: Family matters. Curr. Opin. Struct. Biol. 2010, 20, 313–324. [Google Scholar] [CrossRef]
  171. Peter, B.; Falkenberg, M. TWINKLE and Other Human Mitochondrial DNA Helicases: Structure, Function and Disease. Genes 2020, 11, 408. [Google Scholar] [CrossRef]
  172. Singleton, M.R.; Dillingham, M.S.; Wigley, D.B. Structure and Mechanism of Helicases and Nucleic Acid Translocases. Annu. Rev. Biochem. 2007, 76, 23–50. [Google Scholar] [CrossRef]
  173. Spelbrink, J.N.; Li, F.-Y.; Tiranti, V.; Nikali, K.; Yuan, Q.-P.; Tariq, M.; Wanrooij, S.; Garrido, N.; Comi, G.; Morandi, L.; et al. Human mitochondrial DNA deletions associated with mutations in the gene encoding Twinkle, a phage T7 gene 4-like protein localized in mitochondria. Nat. Genet. 2001, 28, 223–231, Erratum in Nat. Genet. 2001, 29, 100. [Google Scholar] [CrossRef]
  174. Korhonen, J.A.; Pande, V.; Holmlund, T.; Farge, G.; Pham, X.H.; Nilsson, L.; Falkenberg, M. Structure–Function Defects of the TWINKLE Linker Region in Progressive External Ophthalmoplegia. J. Mol. Biol. 2008, 377, 691–705. [Google Scholar] [CrossRef]
  175. Westermann, B. Mitochondrial inheritance in yeast. Biochim. Biophys. Acta 2014, 1837, 1039–1046. [Google Scholar] [CrossRef] [PubMed]
  176. Ding, L.; Liu, Y. Borrowing nuclear DNA helicases to protect mitochondrial DNA. Int. J. Mol. Sci. 2015, 16, 10870–10887. [Google Scholar] [CrossRef] [PubMed]
  177. Holmlund, T.; Farge, G.; Pande, V.; Korhonen, J.; Nilsson, L.; Falkenberg, M. Structure-function defects of the twinkle amino-terminal region in progressive external ophthalmoplegia. Biochim. Biophys. Acta 2009, 1792, 132–139. [Google Scholar] [CrossRef] [PubMed]
  178. Milenkovic, D.; Matic, S.; Kühl, I.; Ruzzenente, B.; Freyer, C.; Jemt, E.; Park, C.B.; Falkenberg, M.; Larsson, N.G. TWINKLE is an essential mitochondrial helicase required for synthesis of nascent D-loop strands and complete mtDNA replication. Hum. Mol. Genet. 2013, 22, 1983–1993. [Google Scholar] [CrossRef]
  179. Farge, G.; Holmlund, T.; Khvorostova, J.; Rofougaran, R.; Hofer, A.; Falkenberg, M. The N-terminal domain of TWINKLE contributes to single-stranded DNA binding and DNA helicase activities. Nucleic Acids Res. 2008, 36, 393–403. [Google Scholar] [CrossRef]
  180. Jemt, E.; Farge, G.; Bäckström, S.; Holmlund, T.; Gustafsson, C.M.; Falkenberg, M. The mitochondrial DNA helicase TWINKLE can assemble on a closed circular template and support initiation of DNA synthesis. Nucleic Acids Res. 2011, 39, 9238–9249. [Google Scholar] [CrossRef]
  181. Ziebarth, T.D.; Farr, C.L.; Kaguni, L.S. Modular Architecture of the Hexameric Human Mitochondrial DNA Helicase. J. Mol. Biol. 2007, 367, 1382–1391. [Google Scholar] [CrossRef]
  182. Ziebarth, T.D.; Gonzalez-Soltero, R.; Makowska-Grzyska, M.M.; Núñez-Ramírez, R.; Carazo, J.M.; Kaguni, L.S. Dynamic effects of cofactors and DNA on the oligomeric state of human mitochondrial DNA helicase. J. Biol. Chem. 2010, 285, 14639–14647. [Google Scholar] [CrossRef]
  183. Fernández-Millán, P.; Lázaro, M.; Cansız-Arda, Ş.; Gerhold, J.M.; Rajala, N.; Schmitz, C.A.; Silva-Espiña, C.; Gil, D.; Bernadó, P.; Valle, M.; et al. The hexameric structure of the human mitochondrial replicative helicase Twinkle. Nucleic Acids Res. 2015, 43, 4284–4295. [Google Scholar] [CrossRef] [PubMed]
  184. Riccio, A.A.; Bouvette, J.; Perera, L.; Longley, M.J.; Krahn, J.M.; Williams, J.G.; Dutcher, R.; Borgnia, M.J.; Copeland, W.C. Structural insight and characterization of human Twinkle helicase in mitochondrial disease. Proc. Natl. Acad. Sci. USA 2022, 119, e2207459119. [Google Scholar] [CrossRef] [PubMed]
  185. Peeva, V.; Blei, D.; Trombly, G.; Corsi, S.; Szukszto, M.J.; Rebelo-Guiomar, P.; Gammage, P.A.; Kudin, A.P.; Becker, C.; Altmüller, J.; et al. Linear mitochondrial DNA is rapidly degraded by components of the replication machinery. Nat. Commun. 2018, 9, 1727. [Google Scholar] [CrossRef] [PubMed]
  186. Sen, D.; Nandakumar, D.; Tang, G.Q.; Patel, S.S. Human mitochondrial DNA helicase TWINKLE is both an unwinding and annealing helicase. J. Biol. Chem. 2012, 287, 14545–14556. [Google Scholar] [CrossRef]
  187. Khan, I.; Crouch, J.D.; Bharti, S.K.; Sommers, J.A.; Carney, S.M.; Yakubovskaya, E.; Garcia-Diaz, M.; Trakselis, M.A.; Brosh, R.M., Jr. Biochemical Characterization of the Human Mitochondrial Replicative Twinkle Helicase: SUBSTRATE SPECIFICITY, DNA BRANCH MIGRATION, AND ABILITY TO OVERCOME BLOCKADES TO DNA UNWINDING. J. Biol. Chem. 2016, 291, 14324–14339. [Google Scholar] [CrossRef]
  188. Sen, D.; Patel, G.; Patel, S.S. Homologous DNA strand exchange activity of the human mitochondrial DNA helicase TWINKLE. Nucleic Acids Res. 2016, 44, 4200–4210. [Google Scholar] [CrossRef]
  189. Singh, A.; Patel, G.; Patel, S.S. Twinkle-Catalyzed Toehold-Mediated DNA Strand Displacement Reaction. J. Am. Chem. Soc. 2023, 145, 24522–24534. [Google Scholar] [CrossRef]
  190. Oliveira, M.T.; Kaguni, L.S. Functional roles of the N- and C-terminal regions of the human mitochondrial single-stranded DNA-binding protein. PLoS ONE 2010, 5, e15379. [Google Scholar] [CrossRef]
  191. Oliveira, M.T.; Kaguni, L.S. Reduced stimulation of recombinant DNA polymerase γ and mitochondrial DNA (mtDNA) helicase by variants of mitochondrial single-stranded DNA-binding protein (mtSSB) correlates with defects in mtDNA replication in animal cells. J. Biol. Chem. 2011, 286, 40649–40658. [Google Scholar] [CrossRef]
  192. Wang, H.Y.; Lee, Y.Y.; Chien, P.J.; Tsai, W.A.; Sun, P.J.; Wang, L.T.; Wu, C.C.; Fan, H.F. Single-molecule fluorescence reveals the DNA unwinding mechanism of mitochondrial helicase TWINKLE and its interplay with single-stranded DNA-binding proteins. Nucleic Acids Res. 2025, 53, gkaf803. [Google Scholar] [CrossRef]
  193. Garrido, N.; Griparic, L.; Jokitalo, E.; Wartiovaara, J.; van der Bliek, A.M.; Spelbrink, J.N. Composition and dynamics of human mitochondrial nucleoids. Mol. Biol. Cell 2003, 14, 1583–1596. [Google Scholar] [CrossRef]
  194. Tyynismaa, H.; Sembongi, H.; Bokori-Brown, M.; Granycome, C.; Ashley, N.; Poulton, J.; Jalanko, A.; Spelbrink, J.N.; Holt, I.J.; Suomalainen, A. Twinkle helicase is essential for mtDNA maintenance and regulates mtDNA copy number. Hum. Mol. Genet. 2004, 13, 3219–3227. [Google Scholar] [CrossRef] [PubMed]
  195. Matsushima, Y.; Kaguni, L.S. Differential phenotypes of active site and human autosomal dominant progressive external ophthalmoplegia mutations in Drosophila mitochondrial DNA helicase expressed in Schneider cells. J. Biol. Chem. 2007, 282, 9436–9444. [Google Scholar] [CrossRef] [PubMed]
  196. Lönnqvist, T.; Paetau, A.; Valanne, L.; Pihko, H. Recessive twinkle mutations cause severe epileptic encephalopathy. Brain 2009, 132, 1553–1562. [Google Scholar] [CrossRef]
  197. Hakonen, A.H.; Goffart, S.; Marjavaara, S.; Paetau, A.; Cooper, H.; Mattila, K.; Lampinen, M.; Sajantila, A.; Lönnqvist, T.; Spelbrink, J.N.; et al. Infantile-onset spinocerebellar ataxia and mitochondrial recessive ataxia syndrome are associated with neuronal complex I defect and mtDNA depletion. Hum. Mol. Genet. 2008, 17, 3822–3835. [Google Scholar] [CrossRef]
  198. Sarzi, E.; Goffart, S.; Serre, V.; Chrétien, D.; Slama, A.; Munnich, A.; Spelbrink, J.N.; Rötig, A. Twinkle helicase (PEO1) gene mutation causes mitochondrial DNA depletion. Ann. Neurol. 2007, 62, 579–587. [Google Scholar] [CrossRef]
  199. Nikali, K.; Suomalainen, A.; Saharinen, J.; Kuokkanen, M.; Spelbrink, J.N.; Lönnqvist, T.; Peltonen, L. Infantile onset spinocerebellar ataxia is caused by recessive mutations in mitochondrial proteins Twinkle and Twinky. Hum. Mol. Genet. 2005, 14, 2981–2990. [Google Scholar] [CrossRef]
  200. Longley, M.J.; Humble, M.M.; Sharief, F.S.; Copeland, W.C. Disease variants of the human mitochondrial DNA helicase encoded by C10orf2 differentially alter protein stability, nucleotide hydrolysis, and helicase activity. J. Biol. Chem. 2010, 285, 29690–29702. [Google Scholar] [CrossRef]
  201. Hakonen, A.H.; Isohanni, P.; Paetau, A.; Herva, R.; Suomalainen, A.; Lönnqvist, T. Recessive Twinkle mutations in early onset encephalopathy with mtDNA depletion. Brain 2007, 130, 3032–3040. [Google Scholar] [CrossRef]
  202. Sanchez-Martinez, A.; Calleja, M.; Peralta, S.; Matsushima, Y.; Hernandez-Sierra, R.; Whitworth, A.J.; Kaguni, L.S.; Garesse, R. Modeling pathogenic mutations of human twinkle in Drosophila suggests an apoptosis role in response to mitochondrial defects. PLoS ONE 2012, 7, e43954. [Google Scholar] [CrossRef]
  203. Tyynismaa, H.; Mjosund, K.P.; Wanrooij, S.; Lappalainen, I.; Ylikallio, E.; Jalanko, A.; Spelbrink, J.N.; Paetau, A.; Suomalainen, A. Mutant mitochondrial helicase Twinkle causes multiple mtDNA deletions and a late-onset mitochondrial disease in mice. Proc. Natl. Acad. Sci. USA 2005, 102, 17687–17692. [Google Scholar] [CrossRef] [PubMed]
  204. Tyynismaa, H.; Carroll, C.J.; Raimundo, N.; Ahola-Erkkilä, S.; Wenz, T.; Ruhanen, H.; Guse, K.; Hemminki, A.; Peltola-Mjøsund, K.E.; Tulkki, V.; et al. Mitochondrial myopathy induces a starvation-like response. Hum. Mol. Genet. 2010, 19, 3948–3958. [Google Scholar] [CrossRef] [PubMed]
  205. Zeviani, M.; Servidei, S.; Gellera, C.; Bertini, E.; DiMauro, S.; DiDonato, S. An autosomal dominant disorder with multiple deletions of mitochondrial DNA starting at the D-loop region. Nature 1989, 339, 309–311. [Google Scholar] [CrossRef] [PubMed]
  206. Suomalainen, A.; Majander, A.; Haltia, M.; Somer, H.; Lönnqvist, J.; Savontaus, M.L.; Peltonen, L. Multiple deletions of mitochondrial DNA in several tissues of a patient with severe retarded depression and familial progressive external ophthalmoplegia. J. Clin. Investig. 1992, 90, 61–66. [Google Scholar] [CrossRef]
  207. Baloh, R.H.; Salavaggione, E.; Milbrandt, J.; Pestronk, A. Familial parkinsonism and ophthalmoplegia from a mutation in the mitochondrial DNA helicase twinkle. Arch. Neurol. 2007, 64, 998–1000. [Google Scholar] [CrossRef]
  208. Cerritelli, S.M.; Frolova, E.G.; Feng, C.; Grinberg, A.; Love, P.E.; Crouch, R.J. Failure to Produce Mitochondrial DNA Results in Embryonic Lethality in Rnaseh1 Null Mice. Mol. Cell 2003, 11, 807–815. [Google Scholar] [CrossRef]
  209. Holmes, J.B.; Akman, G.; Wood, S.R.; Sakhuja, K.; Cerritelli, S.M.; Moss, C.; Bowmaker, M.R.; Jacobs, H.T.; Crouch, R.J.; Holt, I.J. Primer retention owing to the absence of RNase H1 is catastrophic for mitochondrial DNA replication. Proc. Natl. Acad. Sci. USA 2015, 112, 9334–9339. [Google Scholar] [CrossRef]
  210. Al-Behadili, A.; Uhler, J.P.; Berglund, A.-K.; Peter, B.; Doimo, M.; Reyes, A.; Wanrooij, S.; Zeviani, M.; Falkenberg, M. A two-nuclease pathway involving RNase H1 is required for primer removal at human mitochondrial OriL. Nucleic Acids Res. 2018, 46, 9471–9483. [Google Scholar] [CrossRef]
  211. Posse, V.; Al-Behadili, A.; Uhler, J.P.; Clausen, A.R.; Reyes, A.; Zeviani, M.; Falkenberg, M.; Gustafsson, C.M. RNase H1 directs origin-specific initiation of DNA replication in human mitochondria. PLoS Genet. 2019, 15, e1007781. [Google Scholar] [CrossRef]
  212. Akman, G.; Desai, R.; Bailey, L.J.; Yasukawa, T.; Dalla Rosa, I.; Durigon, R.; Holmes, J.B.; Moss, C.F.; Mennuni, M.; Houlden, H.; et al. Pathological ribonuclease H1 causes R-loop depletion and aberrant DNA segregation in mitochondria. Proc. Natl. Acad. Sci. USA 2016, 113, E4276–E4285. [Google Scholar] [CrossRef]
  213. Lima, W.F.; Murray, H.M.; Damle, S.S.; Hart, C.E.; Hung, G.; De Hoyos, C.L.; Liang, X.-H.; Crooke, S.T. Viable RNaseH1 knockout mice show RNaseH1 is essential for R loop processing, mitochondrial and liver function. Nucleic Acids Res. 2016, 44, 5299–5312. [Google Scholar] [CrossRef]
  214. Reyes, A.; Rusecka, J.; Tońska, K.; Zeviani, M. RNase H1 Regulates Mitochondrial Transcription and Translation via the Degradation of 7S RNA. Front. Genet. 2020, 10, 1393. [Google Scholar] [CrossRef]
  215. Wu, H.; Sun, H.; Liang, X.; Lima, W.F.; Crooke, S.T. Human RNase H1 Is Associated with Protein P32 and Is Involved in Mitochondrial Pre-rRNA Processing. PLoS ONE 2013, 8, e71006. [Google Scholar] [CrossRef]
  216. Misic, J.; Milenkovic, D.; Al-Behadili, A.; Xie, X.; Jiang, M.; Jiang, S.; Filograna, R.; Koolmeister, C.; Siira, S.J.; Jenninger, L.; et al. Mammalian RNase H1 directs RNA primer formation for mtDNA replication initiation and is also necessary for mtDNA replication completion. Nucleic Acids Res. 2022, 50, 8749–8766. [Google Scholar] [CrossRef]
  217. Carreño-Gago, L.; Blázquez-Bermejo, C.; Díaz-Manera, J.; Cámara, Y.; Gallardo, E.; Martí, R.; Torres-Torronteras, J.; García-Arumí, E. Identification and Characterization of New RNASEH1 Mutations Associated With PEO Syndrome and Multiple Mitochondrial DNA Deletions. Front. Genet. 2019, 10, 576. [Google Scholar] [CrossRef] [PubMed]
  218. Sachdev, A.; Fratter, C.; McMullan, T.F.W. Novel mutation in the RNASEH1 gene in a chronic progressive external ophthalmoplegia patient. Can. J. Ophthalmol. 2018, 53, e203–e205. [Google Scholar] [CrossRef] [PubMed]
  219. Reyes, A.; Melchionda, L.; Nasca, A.; Carrara, F.; Lamantea, E.; Zanolini, A.; Lamperti, C.; Fang, M.; Zhang, J.; Ronchi, D.; et al. RNASEH1 Mutations Impair mtDNA Replication and Cause Adult-Onset Mitochondrial Encephalomyopathy. Am. J. Hum. Genet. 2015, 97, 186–193. [Google Scholar] [CrossRef]
  220. Bugiardini, E.; Poole, O.V.; Manole, A.; Pittman, A.M.; Horga, A.; Hargreaves, I.; Woodward, C.E.; Sweeney, M.G.; Holton, J.L.; Taanman, J.-W.; et al. Clinicopathologic and molecular spectrum of RNASEH1-related mitochondrial disease. Neurol. Genet. 2017, 3, e149. [Google Scholar] [CrossRef]
  221. Duxin, J.P.; Dao, B.; Martinsson, P.; Rajala, N.; Guittat, L.; Campbell, J.L.; Spelbrink, J.N.; Stewart, S.A. Human Dna2 is a nuclear and mitochondrial DNA maintenance protein. Mol. Cell Biol. 2009, 29, 4274–4282. [Google Scholar] [CrossRef]
  222. Zheng, L.; Zhou, M.; Guo, Z.; Lu, H.; Qian, L.; Dai, H.; Qiu, J.; Yakubovskaya, E.; Bogenhagen, D.F.; Demple, B.; et al. Human DNA2 Is a Mitochondrial Nuclease/Helicase for Efficient Processing of DNA Replication and Repair Intermediates. Mol. Cell 2008, 32, 325–336. [Google Scholar] [CrossRef]
  223. Zheng, L.; Meng, Y.; Campbell, J.L.; Shen, B. Multiple roles of DNA2 nuclease/helicase in DNA metabolism, genome stability and human diseases. Nucleic Acids Res. 2019, 48, 16–35. [Google Scholar] [CrossRef]
  224. Copeland, W.C.; Longley, M.J. Mitochondrial genome maintenance in health and disease. DNA Repair 2014, 19, 190–198. [Google Scholar] [CrossRef] [PubMed]
  225. Ronchi, D.; Di Fonzo, A.; Lin, W.; Bordoni, A.; Liu, C.; Fassone, E.; Pagliarani, S.; Rizzuti, M.; Zheng, L.; Filosto, M.; et al. Mutations in DNA2 Link Progressive Myopathy to Mitochondrial DNA Instability. Am. J. Hum. Genet. 2013, 92, 293–300. [Google Scholar] [CrossRef] [PubMed]
  226. Ronchi, D.; Liu, C.; Caporali, L.; Piga, D.; Li, H.; Tagliavini, F.; Valentino, M.L.; Ferrò, M.T.; Bini, P.; Zheng, L.; et al. Novel mutations in DNA2 associated with myopathy and mtDNA instability. Ann. Clin. Transl. Neurol. 2019, 6, 1893–1899. [Google Scholar] [CrossRef] [PubMed]
  227. Chae, J.H.; Vasta, V.; Cho, A.; Lim, B.C.; Zhang, Q.; Eun, S.H.; Hahn, S.H. Utility of next generation sequencing in genetic diagnosis of early onset neuromuscular disorders. J. Med. Genet. 2015, 52, 208–216. [Google Scholar] [CrossRef]
  228. González-del Angel, A.; Bisciglia, M.; Vargas-Cañas, S.; Fernandez-Valverde, F.; Kazakova, E.; Escobar, R.E.; Romero, N.B.; Jardel, C.; Rucheton, B.; Stojkovic, T.; et al. Novel Phenotypes and Cardiac Involvement Associated with DNA2 Genetic Variants. Front. Neurol. 2019, 10, 1049. [Google Scholar] [CrossRef]
  229. Phowthongkum, P.; Sun, A. Novel truncating variant in DNA2-related congenital onset myopathy and ptosis suggests genotype–phenotype correlation. Neuromuscul. Disord. 2017, 27, 616–618. [Google Scholar] [CrossRef]
  230. Tarnauskaitė, Ž.; Bicknell, L.S.; Marsh, J.A.; Murray, J.E.; Parry, D.A.; Logan, C.V.; Bober, M.B.; de Silva, D.C.; Duker, A.L.; Sillence, D.; et al. Biallelic variants in DNA2 cause microcephalic primordial dwarfism. Hum. Mutat. 2019, 40, 1063–1070. [Google Scholar] [CrossRef]
  231. Shaheen, R.; Faqeih, E.; Ansari, S.; Abdel-Salam, G.; Al-Hassnan, Z.N.; Al-Shidi, T.; Alomar, R.; Sogaty, S.; Alkuraya, F.S. Genomic analysis of primordial dwarfism reveals novel disease genes. Genome Res. 2014, 24, 291–299. [Google Scholar] [CrossRef]
  232. Sun, J.; Su, W.; Deng, J.; Qin, Y.; Wang, Z.; Liu, Y. DNA2 mutation causing multisystemic disorder with impaired mitochondrial DNA maintenance. J. Hum. Genet. 2022, 67, 691–699. [Google Scholar] [CrossRef]
  233. Liu, Y.; Yang, H.; Gan, S.; He, L.; Zeng, R.; Xiao, T.; Wu, L. A novel mutation of DNA2 regulates neuronal cell membrane potential and epileptogenesis. Cell Death Discov. 2024, 10, 259. [Google Scholar] [CrossRef] [PubMed]
  234. Datta, A.; Manousakis, G.; Chompoopong, P. Novel DNA2 Mutation Leading to Mitochondrial DNA Deletions with MELAS-like Phenotype (P2-11.031). Neurology 2025, 104, 3261. [Google Scholar] [CrossRef]
  235. Kornblum, C.; Nicholls, T.J.; Haack, T.B.; Schöler, S.; Peeva, V.; Danhauser, K.; Hallmann, K.; Zsurka, G.; Rorbach, J.; Iuso, A.; et al. Loss-of-function mutations in MGME1 impair mtDNA replication and cause multisystemic mitochondrial disease. Nat. Genet. 2013, 45, 214–219. [Google Scholar] [CrossRef] [PubMed]
  236. Szczesny, R.J.; Hejnowicz, M.S.; Steczkiewicz, K.; Muszewska, A.; Borowski, L.S.; Ginalski, K.; Dziembowski, A. Identification of a novel human mitochondrial endo-/exonuclease Ddk1/c20orf72 necessary for maintenance of proper 7S DNA levels. Nucleic Acids Res. 2013, 41, 3144–3161. [Google Scholar] [CrossRef]
  237. Nicholls, T.J.; Zsurka, G.; Peeva, V.; Schöler, S.; Szczesny, R.J.; Cysewski, D.; Reyes, A.; Kornblum, C.; Sciacco, M.; Moggio, M.; et al. Linear mtDNA fragments and unusual mtDNA rearrangements associated with pathological deficiency of MGME1 exonuclease. Hum. Mol. Genet. 2014, 23, 6147–6162. [Google Scholar] [CrossRef]
  238. Hebbar, M.; Girisha, K.M.; Srivastava, A.; Bielas, S.; Shukla, A. Homozygous c.359del variant in MGME1 is associated with early onset cerebellar ataxia. Eur. J. Med. Genet. 2017, 60, 533–535. [Google Scholar] [CrossRef]
  239. Hangas, A.; Kekäläinen, N.J.; Potter, A.; Michell, C.; Aho, K.J.; Rutanen, C.; Spelbrink, J.N.; Pohjoismäki, J.L.; Goffart, S. Top3α is the replicative topoisomerase in mitochondrial DNA replication. Nucleic Acids Res. 2022, 50, 8733–8748. [Google Scholar] [CrossRef]
  240. Jiang, W.; Jia, N.; Guo, C.; Wen, J.; Wu, L.; Ogi, T.; Zhang, H. Predominant cellular mitochondrial dysfunction in the TOP3A gene-caused Bloom syndrome-like disorder. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2021, 1867, 166106. [Google Scholar] [CrossRef]
  241. Martin, C.A.; Sarlós, K.; Logan, C.V.; Thakur, R.S.; Parry, D.A.; Bizard, A.H.; Leitch, A.; Cleal, L.; Ali, N.S.; Al-Owain, M.A.; et al. Mutations in TOP3A Cause a Bloom Syndrome-like Disorder. Am. J. Hum. Genet. 2018, 103, 456. [Google Scholar] [CrossRef]
  242. Primiano, G.; Torraco, A.; Verrigni, D.; Sabino, A.; Bertini, E.; Carrozzo, R.; Silvestri, G.; Servidei, S. Novel TOP3A Variant Associated With Mitochondrial Disease. Neurol. Genet. 2022, 8, e200007. [Google Scholar] [CrossRef]
  243. Llauradó, A.; Rovira-Moreno, E.; Codina-Solà, M.; Martínez-Saez, E.; Salvadó, M.; Sanchez-Tejerina, D.; Sotoca, J.; López-Diego, V.; Restrepo-Vera, J.L.; Garcia-Arumi, E.; et al. Chronic progressive external ophthalmoplegia plus syndrome due to homozygous missense variant in TOP3A gene. Clin. Genet. 2023, 103, 492–494. [Google Scholar] [CrossRef]
  244. Kozhukhar, N.; Alexeyev, M.F. 35 Years of TFAM Research: Old Protein, New Puzzles. Biology 2023, 12, 823. [Google Scholar] [CrossRef]
  245. Stiles, A.R.; Simon, M.T.; Stover, A.; Eftekharian, S.; Khanlou, N.; Wang, H.L.; Magaki, S.; Lee, H.; Partynski, K.; Dorrani, N.; et al. Mutations in TFAM, encoding mitochondrial transcription factor A, cause neonatal liver failure associated with mtDNA depletion. Mol. Genet. Metab. 2016, 119, 91–99. [Google Scholar] [CrossRef]
  246. Mehmedović, M.; Martucci, M.; Spåhr, H.; Ishak, L.; Mishra, A.; Sanchez-Sandoval, M.E.; Pardo-Hernández, C.; Peter, B.; van den Wildenberg, S.M.; Falkenberg, M.; et al. Disease causing mutation (P178L) in mitochondrial transcription factor A results in impaired mitochondrial transcription initiation. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2022, 1868, 166467. [Google Scholar] [CrossRef] [PubMed]
  247. Ullah, F.; Rauf, W.; Khan, K.; Khan, S.; Bell, K.M.; de Oliveira, V.C.; Tariq, M.; Bakhshalizadeh, S.; Touraine, P.; Katsanis, N.; et al. A recessive variant in TFAM causes mtDNA depletion associated with primary ovarian insufficiency, seizures, intellectual disability and hearing loss. Hum. Genet. 2021, 140, 1733–1751. [Google Scholar] [CrossRef]
  248. Belin, A.C.; Bjork, B.F.; Westerlund, M.; Galter, D.; Sydow, O.; Lind, C.; Pernold, K.; Rosvall, L.; Hakansson, A.; Winblad, B.; et al. Association study of two genetic variants in mitochondrial transcription factor A (TFAM) in Alzheimer’s and Parkinson’s disease. Neurosci. Lett. 2007, 420, 257–262. [Google Scholar] [CrossRef] [PubMed]
  249. Gaweda-Walerych, K.; Safranow, K.; Maruszak, A.; Bialecka, M.; Klodowska-Duda, G.; Czyzewski, K.; Slawek, J.; Rudzinska, M.; Styczynska, M.; Opala, G.; et al. Mitochondrial transcription factor A variants and the risk of Parkinson’s disease. Neurosci. Lett. 2010, 469, 24–29. [Google Scholar] [CrossRef]
  250. Golubickaite, I.; Ugenskiene, R.; Bartnykaite, A.; Poskiene, L.; Vegiene, A.; Padervinskis, E.; Rudzianskas, V.; Juozaityte, E. Mitochondria-Related TFAM and POLG Gene Variants and Associations with Tumor Characteristics and Patient Survival in Head and Neck Cancer. Genes 2023, 14, 434. [Google Scholar] [CrossRef] [PubMed]
  251. Guo, J.; Zheng, L.; Liu, W.; Wang, X.; Wang, Z.; Wang, Z.; French, A.J.; Kang, D.; Chen, L.; Thibodeau, S.N.; et al. Frequent truncating mutation of TFAM induces mitochondrial DNA depletion and apoptotic resistance in microsatellite-unstable colorectal cancer. Cancer Res. 2011, 71, 2978–2987. [Google Scholar] [CrossRef]
  252. Oláhová, M.; Peter, B.; Szilagyi, Z.; Diaz-Maldonado, H.; Singh, M.; Sommerville, E.W.; Blakely, E.L.; Collier, J.J.; Hoberg, E.; Stránecký, V.; et al. POLRMT mutations impair mitochondrial transcription causing neurological disease. Nat. Commun. 2021, 12, 1135. [Google Scholar] [CrossRef]
  253. Fassad, M.R.; Valenzuela, S.; Oláhová, M.; Collier, J.J.; Knowles, C.V.Y.; Mavraki, E.; Elbracht, M.; Güzel, N.; Herberhold, T.; Kurth, I.; et al. Expanding the Genetic and Phenotypic Spectrum of POLRMT-Related Mitochondrial Disease. Clin. Genet. 2025, 1–9. [Google Scholar] [CrossRef] [PubMed]
  254. Garone, C.; Garcia-Diaz, B.; Emmanuele, V.; Lopez, L.C.; Tadesse, S.; Akman, H.O.; Tanji, K.; Quinzii, C.M.; Hirano, M. Deoxypyrimidine monophosphate bypass therapy for thymidine kinase 2 deficiency. EMBO Mol. Med. 2014, 6, 1016–1027. [Google Scholar] [CrossRef]
  255. Lopez-Gomez, C.; Levy, R.J.; Sanchez-Quintero, M.J.; Juanola-Falgarona, M.; Barca, E.; Garcia-Diaz, B.; Tadesse, S.; Garone, C.; Hirano, M. Deoxycytidine and Deoxythymidine Treatment for Thymidine Kinase 2 Deficiency. Ann. Neurol. 2017, 81, 641–652. [Google Scholar] [CrossRef] [PubMed]
  256. Blazquez-Bermejo, C.; Carreno-Gago, L.; Molina-Granada, D.; Aguirre, J.; Ramon, J.; Torres-Torronteras, J.; Cabrera-Perez, R.; Martin, M.A.; Dominguez-Gonzalez, C.; de la Cruz, X.; et al. Increased dNTP pools rescue mtDNA depletion in human POLG-deficient fibroblasts. FASEB J. 2019, 33, 7168–7179. [Google Scholar] [CrossRef]
  257. Dombi, E.; Marinaki, T.; Spingardi, P.; Millar, V.; Hadjichristou, N.; Carver, J.; Johnston, I.G.; Fratter, C.; Poulton, J. Nucleoside supplements as treatments for mitochondrial DNA depletion syndrome. Front. Cell Dev. Biol. 2024, 12, 1260496. [Google Scholar] [CrossRef] [PubMed]
  258. Kristiansen, C.K.; Furriol, J.; Chen, A.; Sullivan, G.J.; Bindoff, L.A.; Liang, K.X. Deoxyribonucleoside treatment rescues EtBr-induced mtDNA depletion in iPSC-derived neural stem cells with POLG mutations. FASEB J. 2023, 37, e23139. [Google Scholar] [CrossRef]
  259. Berrahmoune, S.; Dassi, C.; Pekeles, H.; Cheung, A.C.T.; Gagnon, T.; Waters, P.J.; Buhas, D.; Myers, K.A. Investigating the safety and efficacy of deoxycytidine/deoxythymidine in mitochondrial DNA depletion disorders: Phase 2 open-label trial. J. Neurol. 2025, 272, 307. [Google Scholar] [CrossRef]
  260. Pekeles, H.; Berrahmoune, S.; Dassi, C.; Cheung, A.C.T.; Gagnon, T.; Waters, P.J.; Eberhard, R.; Buhas, D.; Myers, K.A. Safety and efficacy of deoxycytidine/deoxythymidine combination therapy in POLG-related disorders: 6-month interim results of an open-label, single arm, phase 2 trial. eClinicalMedicine 2024, 74, 102740. [Google Scholar] [CrossRef]
  261. Chen, A.; Kristiansen, C.K.; Hong, Y.; Kianian, A.; Fang, E.F.; Sullivan, G.J.; Wang, J.; Li, X.; Bindoff, L.A.; Liang, K.X. Nicotinamide Riboside and Metformin Ameliorate Mitophagy Defect in Induced Pluripotent Stem Cell-Derived Astrocytes with POLG Mutations. Front. Cell Dev. Biol. 2021, 9, 737304. [Google Scholar] [CrossRef]
  262. Valenzuela, S.; Zhu, X.; Macao, B.; Stamgren, M.; Geukens, C.; Charifson, P.S.; Kern, G.; Hoberg, E.; Jenninger, L.; Gruszczyk, A.V.; et al. Small molecules restore mutant mitochondrial DNA polymerase activity. Nature 2025, 642, 501–507. [Google Scholar] [CrossRef]
  263. Gustafson, M.A.; Sullivan, E.D.; Copeland, W.C. Consequences of compromised mitochondrial genome integrity. DNA Repair 2020, 93, 102916. [Google Scholar] [CrossRef]
Figure 1. Human mitochondrial DNA. (A) Schematic of the human mitochondrial genome. Human mtDNA is circular and 16,569 base pairs in length. Human mtDNA encodes 13 mRNAs, all of which are subunits of electron transport chain complexes, as well as 2 rRNAs, and 22 tRNAs. The noncoding region of the genome contains the heavy-strand origin of replication (OH), the heavy-strand promoter (HSP), and two light-strand promoters (LSP1 and LSP2). The light-strand origin of replication (OL) is located approximately two-thirds away from OH, near COX1 and ND2. Gene names annotated on the outside of the diagram are on the heavy strand, and gene names annotated on the inside of the diagram are on the light strand. (B) Overview of the strand displacement model of mtDNA replication. Replication is initiated first at OH, where the heavy strand becomes displaced and is stabilized by mtSSB. POLRMT synthesizes a short RNA primer, and Twinkle helicase unwinds the DNA as pol γ synthesizes a new heavy strand of DNA. Once the replication machinery reaches OL, a stem-loop structure is formed and POLRMT can initiate primer synthesis that allows replication of a new light strand to proceed in the opposite direction. Replication of mtDNA results in two identical daughter molecules of mtDNA.
Figure 1. Human mitochondrial DNA. (A) Schematic of the human mitochondrial genome. Human mtDNA is circular and 16,569 base pairs in length. Human mtDNA encodes 13 mRNAs, all of which are subunits of electron transport chain complexes, as well as 2 rRNAs, and 22 tRNAs. The noncoding region of the genome contains the heavy-strand origin of replication (OH), the heavy-strand promoter (HSP), and two light-strand promoters (LSP1 and LSP2). The light-strand origin of replication (OL) is located approximately two-thirds away from OH, near COX1 and ND2. Gene names annotated on the outside of the diagram are on the heavy strand, and gene names annotated on the inside of the diagram are on the light strand. (B) Overview of the strand displacement model of mtDNA replication. Replication is initiated first at OH, where the heavy strand becomes displaced and is stabilized by mtSSB. POLRMT synthesizes a short RNA primer, and Twinkle helicase unwinds the DNA as pol γ synthesizes a new heavy strand of DNA. Once the replication machinery reaches OL, a stem-loop structure is formed and POLRMT can initiate primer synthesis that allows replication of a new light strand to proceed in the opposite direction. Replication of mtDNA results in two identical daughter molecules of mtDNA.
Ijms 26 10275 g001
Figure 2. Structures of the minimal mitochondrial DNA replication machinery and the disease outcomes associated with mutations. The replicative polymerase, pol γ, is comprised of PolG, the catalytic subunit, and PolG2, the accessory subunit. The catalytic subunit alone is represented in PDB: 4ZTU bound to DNA. The accessory subunit bound to DNA as a trimer of dimers is represented in PDB: 8F6B. The holoenzyme in complex with DNA is shown in PDB: 4ZTU. Mitochondrial single-stranded binding protein bound to DNA is shown in PDB: 8UZT. The Twinkle helicase is represented in PDB: 7T8C.
Figure 2. Structures of the minimal mitochondrial DNA replication machinery and the disease outcomes associated with mutations. The replicative polymerase, pol γ, is comprised of PolG, the catalytic subunit, and PolG2, the accessory subunit. The catalytic subunit alone is represented in PDB: 4ZTU bound to DNA. The accessory subunit bound to DNA as a trimer of dimers is represented in PDB: 8F6B. The holoenzyme in complex with DNA is shown in PDB: 4ZTU. Mitochondrial single-stranded binding protein bound to DNA is shown in PDB: 8UZT. The Twinkle helicase is represented in PDB: 7T8C.
Ijms 26 10275 g002
Table 1. Mitochondrial disease genes associated with mtDNA maintenance.
Table 1. Mitochondrial disease genes associated with mtDNA maintenance.
GeneDisorderLocusFunction
mtDNA replication and repair
APTXataxia9p21.1DNA repair
DNA2mtDNA deletions, PEO, epilepsy10q21.3Mito/nuclear helicase–nuclease
MGME1PEO, mtDNA depletion20p11.23Single-stranded DNA nuclease
POLGPEO, Alpers, ataxia, epilepsy,
mtDNA depletion
15q26.1Pol γ catalytic subunit
POLG2PEO, ataxia, mtDNA depletion17q23.3Pol γ accessory subunit
POLRMTPEO19p13.3RNA polymerase
RNASEH1PEO, encephalopathy, ataxia, mtDNA deletions2p25.3RNA/DNA hybrid endoribonuclease
SSBP1Optic atrophy, mtDNA depletion/deletions7q34Single-stranded DNA-binding protein
TFAMmtDNA depletion10q21.1DNA compaction, transcription factor
TOP3ABloom-syndrome-like disorders, PEO17p11.2Topoisomerase
TWNKPEO, ataxia, mtDNA depletion10q24.31Replicative helicase
nucleotide pool metabolism and maintenance
ABATmtDNA deletions, depletion16p13.24-Aminobutyrate aminotransferase
DGUOKmtDNA depletion, PEO2p13.1Deoxyguanosine kinase
RRM2BPEO, mtDNA deletions, depletion8q22.3p53-Ribonucleotide reductase subunit
SAMHD1mtDNA deletions20q11.23dNTP triphosphohydrolase
SLC25A4PEO4q34.1Adenine nucleotide translocator
SUCLA2mtDNA depletion13q14.2ATP-dep Succinate-CoA ligase
SUCLG1mtDNA depletion2p11.2GTP-dep Succinate-CoA ligase
TK2PEO, mtDNA depletion16q21Mitochondrial thymidine kinase
TYMPMNGIE, mtDNA deletions/depletion22q13.33Thymidine phosphorylase
mitochondrial homeostasis and dynamics
AFG3L2Spinocerebellar ataxia, mtDNA deletions18p11.21Mitochondrial metalloprotease
DNM1LEncephalopathy, neurological disorders,
epilepsy
12p11.21GTPase involved in mitochondrial fission
FBXL4mtDNA depletion, encephalopathy6q16.1–16.2Mitochondrial LLR F-Box protein
GDAP1CMT disease8q21.11Mitochondrial fission protein
GFERmtDNA deletions, myopathy16p13.3Protein import to inner membrane
MFFEncephalopathy, hypotonia, neurological
disorders
2q36.3Mitochondrial fission protein
MFN2CMT disease, dominant optic atrophy, mtDNA deletions1p36.22Mitochondrial fusion protein
MPV17mtDNA depletion, CMT disease2p23.3Unknown, inner membrane protein
NME3Neurodegeneration, hypotonia16p13.3Nucleoside diphosphate kinase, fusion
OPA1Dominant optic atrophy, mtDNA deletions, ataxia3q29Dynamin-related GTPase
SLC25A46Leigh syndrome, optic atrophy, ataxia,
CMT disease
5q22.1Mitochondrial fission protein
SPG7ataxia, spastic paraplegia16q24.3Mitochondrial metalloprotease
STAT2Immunodeficiency12q13.3Mitochondrial fission protein
TMEM65mtDNA depletion8q24.13Na+/Ca2+ exchange
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

Somai, S.; Aloh, C.H.; King, D.E.; Copeland, W.C. Mitochondrial DNA Replication and Disease: A Historical Perspective on Molecular Insights and Therapeutic Advances. Int. J. Mol. Sci. 2025, 26, 10275. https://doi.org/10.3390/ijms262110275

AMA Style

Somai S, Aloh CH, King DE, Copeland WC. Mitochondrial DNA Replication and Disease: A Historical Perspective on Molecular Insights and Therapeutic Advances. International Journal of Molecular Sciences. 2025; 26(21):10275. https://doi.org/10.3390/ijms262110275

Chicago/Turabian Style

Somai, Shruti, Chioma H. Aloh, Dillon E. King, and William C. Copeland. 2025. "Mitochondrial DNA Replication and Disease: A Historical Perspective on Molecular Insights and Therapeutic Advances" International Journal of Molecular Sciences 26, no. 21: 10275. https://doi.org/10.3390/ijms262110275

APA Style

Somai, S., Aloh, C. H., King, D. E., & Copeland, W. C. (2025). Mitochondrial DNA Replication and Disease: A Historical Perspective on Molecular Insights and Therapeutic Advances. International Journal of Molecular Sciences, 26(21), 10275. https://doi.org/10.3390/ijms262110275

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