Skip to main content

Induced pluripotent stem cells: ex vivo models for human diseases due to mitochondrial DNA mutations

Abstract

Mitochondria are essential organelles for cellular metabolism and physiology in eukaryotic cells. Human mitochondria have their own genome (mtDNA), which is maternally inherited with 37 genes, encoding 13 polypeptides for oxidative phosphorylation, and 22 tRNAs and 2 rRNAs for translation. mtDNA mutations are associated with a wide spectrum of degenerative and neuromuscular diseases. However, the pathophysiology of mitochondrial diseases, especially for threshold effect and tissue specificity, is not well understood and there is no effective treatment for these disorders. Especially, the lack of appropriate cell and animal disease models has been significant obstacles for deep elucidating the pathophysiology of maternally transmitted diseases and developing the effective therapy approach. The use of human induced pluripotent stem cells (iPSCs) derived from patients to obtain terminally differentiated specific lineages such as inner ear hair cells is a revolutionary approach to deeply understand pathogenic mechanisms and develop the therapeutic interventions of mitochondrial disorders. Here, we review the recent advances in patients-derived iPSCs as ex vivo models for mitochondrial diseases. Those patients-derived iPSCs have been differentiated into specific targeting cells such as retinal ganglion cells and eventually organoid for the disease modeling. These disease models have advanced our understanding of the pathophysiology of maternally inherited diseases and stepped toward therapeutic interventions for these diseases.

Background

Mitochondria are essential organelles for cellular metabolism and physiology in eukaryotic cells. The primary function of mitochondria acts as a powerhouse to generate the large quantities of ATP through oxidative phosphorylation (OXPHOS) [1]. Mitochondria also play the pivotal roles in regulating cellular calcium balance, adaptive thermogenesis, programmed cell death, and affecting inflammation and aging [2,3,4]. More than 1000 mitochondrial proteins encoded by nuclear genes are synthesized in the cytosol and transported to mitochondrial membranes or matrix [5], while 13 core components of OXPHOS complexes, encoded by mitochondrial DNA (mtDNA), are produced by mitochondrial translation machinery [6].

Human mtDNAs are referred as circular double-strand molecules with 16,569 base pairs, encoding 13 polypeptides for OXPHOS complexes, 2 ribosomal RNAs and 22 transfer RNAs required for the synthesis of polypeptides [7]. mtDNA replication is catalyzed by own replication machinery including DNA polymerase γ (Polγ), twinkle helicase and single stranded DNA binding protein (mtSSB) [8]. Unlike nuclear DNA, mtDNA is polyploid, which are multiple copies of mtDNA molecules within each cell. The numbers of mtDNA molecules vary greatly in different cell types and organs, from around 100 copies of mtDNA molecules in sperm to hundreds of thousands in oocytes [9,10,11]. During fertilization, mitochondria from sperm are selectively degraded in zygotes, thus mtDNA is inherited from mothers rather than fathers [12].

There is an inextricable connection between the polyploid mtDNA and genetic heterogeneity. Variations in mtDNA are present in either heteroplasmy (mixture of wild-type and mutated molecules) or homoplasmy (all mutated molecules) [13,14,15,16,17,18]. mtDNA mutations occur sporadically or arise due in large part to the high mutation rate of mtDNA and revealed the maternal inheritance [13,14,15]. The absence of abundant DNA repair system or protection of histones makes mtDNA more vulnerable to reactive oxygen species (ROS), generated in mitochondrial electron transport chain [1, 13]. mtDNA mutations also arise from defects in the biogenesis and mtDNA maintaince, reflecting the contribution of nuclear-encoded genes such as POLG to these processes, and in this case exhibit Mendelian inheritance [14, 17].

Mitochondrial diseases

Mitochondrial diseases are a group of genetic disorders that affect any organ at any age, caused by mutations in nuclear and mitochondrial genes [1,2,3]. Mitochondrial diseases are often multi-systemic disorders, affecting tissues or organ with high energy demands, such as the nervous system, heart, and skeletal muscles [13,14,15]. Mitochondrial diseases also manifest single organ and tissue, such as ear and eye [18]. In this review, we focus on the mitochondrial diseases arising from mtDNA alterations (Fig. 1). These mtDNA alterations included mtDNA rearrangements such as deletions, inversions or duplications, point mutations, or copy number depletion. mtDNA large scale deletions are present in heteroplasmic form to cause clinical abnormalities including Kearns–Sayre syndrome (KSS), Pearson syndrome (PS), or Chronic progressive external ophthalmoplegia (CPEO) [13]. The heteroplasmic point mutations affected more than one organ or tissues including mitochondrial encephalomyopathy lactic acidosis, and stroke-like episodes (MELAS) associated tRNALeu(UUR) 3243A > G [19], myoclonic epilepsy associated with ragged-red fibers (MERRF) linked tRNALys 8344A > G [20], Leigh syndrome-associated ATP6 8993T > G mutations [21]. By contrast, the homoplasmic or near homoplasmic mutations often manifest single organ or tissues, including Leber hereditary optic neuropathy (LHON)-associated ND4 11778G > A, ND1 3460G > A, and ND6 14484T > C [22], deafness-linked 12S rRNA 1555A > G and 1494C > T mutations [23, 24].

Fig. 1
figure 1

Mitochondrial genome and pathogenic mtDNA mutations. Human mitochondrial genome is shown as circular, double-stranded DNA molecule with annotations. The ribosomal RNA genes are shown in purple, while tRNA genes are shown in white and annotated with single letter abbreviations. The subunits of complex I (cyan), cytochrome b (green) of complex III, subunits of complex IV (yellow) and complex V (red) are denoted by the position along the mtDNA sequence, with the outer circle as the heavy chain, and inner circle as the light chain. The positions of pathogenic mtDNA mutations are marked by black arrows. KSS Kearns–Sayre syndrome, LHON Leber’s hereditary optic neuropathy, MELAS Mitochondrial Encephalomyopathy, Lactic Acidosis, and Stroke-like episodes, MERRF Myoclonic Epilepsy and Ragged Red Muscle Fibers

The heteroplasmic point mutations such as m.3243A > G mutation confer drastic biochemical defects and are by themselves enough to produce the clinical phenotype [1, 1315, 19]. Individuals carrying the heteroplasmic mtDNA mutation often present with more than one initial clinical manifestation. The clinical phenotype of m.3243A > G mutation is highly complex and variable, ranging from asymptomatic to lethal phenotypes, partly depending on the level and distribution of m.3243A > G heteroplasmy across cells and tissues [25]. The development of phenotypes arising from these mutations depends on the proportion of mutated mtDNA molecules in the cells, surpassed a threshold level to maintain full OXPHOS function in the vulnerability of tissue or organ [1, 15]. The threshold effects vary among cells, tissues or organs, and the organs most affected with relatively low threshold levels are often those with high energy demand, such as the brain, muscles and heart [26]. Moreover, the elevated proportion of mutated mtDNA moecules across generation leads to genetic anticipation. Heteroplasmy dynamics might be positively or negatively modulated by mitophagy and other cellular contexts, while the underlying mechanisms have remained to be elucidated [27].

Homoplasmic mtDNA mutations often caused mild biochemical defects and manifested only one tissue or organs. Only some of matrilineal relatives bearing the mtDNA mutation(s) developed a clinical phenotype such as deafness, while other mutation carriers did not develop a disease phenotype. These mtDNA mutation(s) is by itself insufficient to produce the clinical phenotypes and other genetic and environmental factors such as nuclear modifier gene and aminoglycosides contributed to the development of clinical phenotype. In particular, there were marked variations in the penetrance and gender bias occurring in the LHON pedigrees carrying the mtDNA mutation, reflecting the complex etiology of this disease [28,29,30]. Mutations in the X-linked modifier PRICKLE3 encoding a mitochondrial protein linked to the biogenesis of ATPase or autosomal recessive modifier YARS2 encoding mitochondrial tyrosyl-tRNA synthetase interact with m.11778G > A mutation to cause optic neuropathy [28, 29]. The administration of aminoglycosides can induce or worsen hearing loss in these subjects carrying the m.1555A > G or m.1494C > T mutation [23, 24]. In the absence of aminoglycosides, a modifier allele (c.28G > T, p.Ala10Sser) in TRMU encoding methylaminomethyl-2-thiouridylate-methyltransferase related to tRNA modification interacts with m.1555A > G mutation to cause deafness [31, 32]. Moreover, mitochondrial genetic modifiers such as m.3394T > C, m.4435A > G, and m.3866 T > C mutations act in synergy with the m.11778G > A mutations to increase penetrance and expressivity of LHON [33,34,35].

Mitochondrial diseases present a tissue specificity characterized by the fact that even if a mtDNA mutation is present in all tissues, only some will be affected and induce a pathology. In fact, one mtDNA mutation is able to cause different clinical presentation, or same disease can result from different mtDNA mutations. Clinical presentations due to mtDNA mutations share neurological clinical features such as seizure, epilepsy, or myopathy, and non-neurological features such as cardiomyopathy, diabetes mellitus, or respiratory failure [18]. Tissue-specific effects of heteroplasmy mtDNA mutations reflect the differing bioenergetic requirements of different tissue [36]. The heterogeneity in the phenotypic expression of m.3243A > G mutation ranges from mild phenotype such as deafness to severe symptoms such as stroke-like episodes cardiomyopathy [19, 37]. The tissue specific defects of homoplasmic mtDNA mutations such as m.11778G > A or m.1555A > G mutation are due to the contribution of nuclear modifier genes or the use of aminoglycoside [28, 29, 31, 32, 38, 39]. However, the lack of appropriate cell and animal disease models has been significant obstacles for deep elucidating the pathophysiology of maternally transmitted diseases, especially for tissue-specific effects [19, 40].

iPSCs

iPSCs can be induced to differentiate into nearly all cell types through guided stepwise differentiation protocols to recapitulate the embryonic development [41]. The induced pluripotency triggered great scientific enthusiasm and excitement for etiology research and cellular therapy applications [42]. In particular, the iPSCs can be generated from somatic cells by introducing four “Yamanaka factors”, Oct3/4, Sox2, c-Myc, and Klf4 [43, 44]. These iPSCs derived from human patient cells have provided the excellent disease models to investigate the etiology in affecting organs or tissues, subtly bypassing the impracticality to conduct biopsy on eyes or inner ears. Since the original discovery, the follow-up studies reported plentiful of reprogramming strategies including viral or non-viral system delivering distinct combinations of exogenous transcription factors, RNAs, and even small-molecule compounds [45].

Towards appropriate patient derived iPSCs, the source of somatic cells, transcription factors, reprogramming enhancers, and the delivery strategies should all be taken into consideration. Fibroblasts are most used source of cells for reprogramming for the well-established skin biopsy and culturing techniques [44, 46]. Peripheral blood mononuclear cells (PBMCs) and cells derived from urine, easily available in medical examinations, were also reported to be reprogrammed to iPSCs [47, 48], as well as CD34+ hematopoietic stem cells [49], T-lymphocytes [50], and keratinocytes [51]. Once the cell type was chosen, the set of reprogramming factors should be compatible. Variations of the combinations of reprogramming factors to optimize the efficiency for generating iPSCs have been essential issues to better understand induced pluripotency [52]. Epigenetics changes occurred during the reprogramming process. Hence small-molecule chemicals affecting epigenetics act as enhancers for reprogramming, for example, vitamin C inducing the demethylation of DNA [53], valproic acid and sodium butyrate inhibiting histone deacetylase [54, 55], and 5-azacytidine (AZA) inhibiting DNA methyltransferase [56]. Moreover, the choice of delivery strategy is the selection of vehicle to introduce the reprogramming factors to designated cells, affecting the efficiency and safety. Integrative strategies including retrovirus, lentivirus, and transposon were used in the early studies, yet raising the concerns of carcinogenicity [43, 44, 46, 57, 58]. Most recently, nonintegrative strategies including adenovirus, sendai virus and episomal vectors have been developed [59,60,61]. Furthermore, direct delivery of proteins, mRNAs, self-replicative RNAs, miRNAs, or small-molecule compounds have been used sufficiently for reprogramming [62,63,64,65,66].

The use of human-iPSC (hiPSC) derived from patients to obtain terminally differentiated specific targeting cells such as neurons and cardiomyocytes is a revolutionary approach to understanding pathogenic mechanisms and to develop effective therapeutic strategies for human diseases [67]. In particular, the iPSCs technology has offered great opportunity to shed new light on the threshold effect, incomplete penetrance, and tissue specificity of mitochondrial diseases. It should be taken into consideration in modelling mitochondrial diseases whether the status of mtDNA mutations is altered after reprogramming from patient cells to iPSCs, or after the differentiation to target cells [68]. The long-term monitor of mtDNA should be included to avoid de novo mtDNA mutations when culturing or manipulating the iPSCs.

iPSC models for mitochondrial diseases

A number of patient-derived iPSCs models for mitochondrial diseases have been established (Fig. 2). In this review, we summarized these patient-specific iPSCs models for mitochondrial disorders due to mtDNA large deletions, point mutations in protein-coding genes, tRNA and 12S rRNA genes (Table 1). These patients-derived iPSCs have been differentiated into specific targeting cells such as retinal ganglion cells (two-dimensional differentiation) and eventually organoid (three-dimensional differentiation) for disease modeling. These models have advanced our deep understanding the pathogenic mechanism, especially in tissue specific effects, and stepped toward therapeutic interventions for these diseases.

Fig. 2
figure 2

iPSC models for human diseases due to mtDNA alterations. In modelling mtDNA diseases, patient derived cells are firstly reprogrammed to iPSCs. With mitochondrial base editing and mito-TALEN, the manipulating of mtDNA in iPSCs is achievable. The patient derived iPSCs and genetically corrected iPSCs are differentiated to distinct types of targeting cells to investigate the pathophysiology and to develop the therapeutic intervention approaches for these diseases

Table 1 iPSC models of mtDNA mutation induced diseases

Kearns–Sayre syndrome (KSS) and Pearson syndrome (PS)

Several patient-derived iPSCs for KSS and PS with mtDNA deletions have been established and characterized. KSS and PS are the multisystem mitochondrial diseases caused by large-scale mitochondrial DNA deletions, especially the common 4,977-bp deletion [69,70,71,72,73,74]. KSS is a rare neuromuscular disorder, characterized by progressive external ophthalmoplegia, pigmentary retinopathy, with additional cardiac conduction defects, or ataxia [69, 70, 74], while PS is a multisystem mitochondrial disease with bone marrow failure [71,72,73]. iPSC derived from PS and KSS patients carrying the heteroplasmy mtDNA deletions have been established. The heteroplasmy level of mutant mtDNA in these iPSC lines ranged from high levels of mutation to normal mitochondrial DNA during cell expansion of iPSCs generated from patients. PS-iPSCs carrying a high burden of deleted mtDNA exhibited defects in growth, mitochondrial ultrastructure, oxidative phosphorylation and hematopoietic phenotype when differentiated in vitro, compared to isogenic iPS cells without deleted mtDNA [75, 76]. Furthermore, hiPSC derived from KSS patients with 7.3 kb mtDNA deletion were generated [77, 78]. Interestingly, some iPSCs from these KSS or PS patients showed normal differentiation toward to target tissue or organ without any mtDNA deletions over passages [75,76,77,78]. These open the new avenues to create isogenic mutation free iPSC with absent or very low level of expression of mtDNA deletion for future cell replacement therapies.

MELAS

The hiPSCs models for MELAS syndrome, caused by the heteroplasmic mtDNA mutation(s), covered disease modelling and therapeutic applications [79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94]. MELAS syndrome is one of the most frequent maternally inherited mitochondrial disorders, with broad manifestations including stroke-like episodes, dementia, epilepsy, lactic acidemia, myopathy, recurrent headaches, hearing loss, diabetes, and short stature [19, 94, 95]. The tRNALeu(UUR) 3243A > G mutation is the most prevalent MELAS-associated mtDNA mutation [94, 95], while other MELAS-associated mtDNA mutations included the tRNALeu(UUR) 3271T > C, tRNATrp 5541C > T and ND5 13513G > A mutations [94, 95]. In particular, the m.3243A > G mutation caused pleiotropic defects in tRNALeu(UUR) metabolism, ultimately impairing OXPHOS and mitochondrial functions [96, 97]. During the programming, the iPSC lines underwent an mtDNA bottleneck with bimodal segregation and carried the mutation at variable heteroplasmy, ranging from nearly homoplasmy to mutation-low or mutation-free [79,80,81,82]. Those iPSCs-derived specific types of cells such as neuron, cardiomyocytes and retinal pigment epithelium (RFE) cells have made possible the study of not only common pathophysiological alterations but also cell-specific defects in the disease [79,80,81,82,83]. MELAS iPSC-derived neurons harboring high m.3243A > G mutation load showed defective OXPHOS function, increasing ROS production, aberrant mitophagy and impaired neuronal network functionality and synchronicity [79, 81]. MELAS iPSC bearing nearly homoplasmic m.5541C > T mutation exhibited defective neuronal differentiation but without apparent impairment of neural progenitors or skeletal muscle cells [82]. Furthermore, MELAS iPSC-derived neurons with high m.3243A > G levels displayed abnormal electrophysiological properties: reduced mean firing rate, reduced network burst rates and increased random spike events outside of network bursts, while neurons with intermediate heteroplasmy levels (~ 30%) were relatively normal, indicative of the threshold effect for mtDNA mutations [81]. Other MELAS iPSC-derived cell types included the endothelial cells displaying pro-atherogenic properties and RPE cells exhibiting dysfunctional phagocytosis [84,85,86]. In particular, these RPE cells revealed abnormal mitochondria and melanosomes, marked functional defects in phagocytosis of photoreceptor outer segments, and dysregulated autophagic activity [84, 85, 89]. Moreover, motor neurons bearing the m.3243A > G mutation exhibited defective mitochondrial respiration and developmental deficit, evidenced by elevating Notch signaling that maintained the stem cell identity in NPCs and inhibited neurite outgrowth in neurons [90, 98, 99].

The hiPSC model for MELAS can be used for developing therapeutic interventions. The levels of heteroplasmic m.3243A > G and m.13513G > A mutations in iPSCs were efficiently reduced by mito-TALEN approach to reduce mutant mtDNA loads and restore oxidative phosphorylation function in mitochondrial diseases [91, 92]. Moreover, a small-molecule compound tryptolinamide (TLAM) activated mitochondrial respiration in fibroblasts from patient-derived iPSCs carrying the m.3243A > G mutation and rescued the defect in neuronal differentiation of iPSCs with high ratio of mutation [93].

MERRF

MERRF (myoclonic epilepsy with ragged red fibers) is a rare syndromic mitochondrial disorder, mainly caused by the heteroplasmic tRNALys 8344A > G mutation [20, 100]. The primary clinical features in this disorder are myoclonus, dementia, muscular weakness, ataxia and ragged-red fibers in the muscle, and other secondary clinical manifestations included peripheral neuropathy, renal dysfunction or cardiomyopathy or hearing impairment [100]. These iPSCs derived from MERRF patients carrying the m8344A > G mutation were differentiated into neuron like cells, which allow the study of these diseases in one of the most affected cellular types [101,102,103]. In particular, neural progenitor cells (NPCs) derived from iPSCs from MERRF patients bearing the m.8344A > G mutation exhibited the pathophysiological features previously observed in other disease models, such as an impaired mitochondrial structure and function, increased the level of ROS production, altered antioxidant enzyme expression (101). Furthermore, induced neurons derived from iPSc reprogrammed from skin fibroblasts of MERRF patients harboring the m.8344A > G mutation showed pathophysiological features including impaired mitochondrial morphology and function, reduced membrane potentials, elevated ROS levels, autophagy flux disruption and increased mitophagy [103]. Moreover, inner ear hair cells derived from iPSCs carrying the m.8344A > G mutation displayed significantly elevated ROS levels, more single cilia with a shorter length or fewer stereociliary bundle-like protrusions, and altered transcriptome [102].

Leigh syndrome

Leigh Syndrome (LS) is a rare, inherited neurometabolic disorder that mainly affects the central nervous system, caused by mutations in mitochondrial DNA or nuclear genes encoding mitochondrial proteins [104, 105]. The heterogeneous neurological symptoms include hypotonia, developmental delay, ataxia, seizures, dystonia, polyneuropathy, nystagmus, and optic atrophy [104]. The LS-causative mtDNA mutations included hetroplasmic ATP6 8993 T > G and 9185T > C and ND5 13513G > A mutation [104,105,106]. LS iPSC-derived patients carrying the m.8993T > G, m.9185T > C or m.13513G > A mutation were differentiated into neural progenitors, neurons, and brain organoids [107,108,109,110,111,112]. Neural cells differentiated from these LS iPSCs bearing the m.8993T > A or 9185T > C mutation revealed deficiencies in bioenergetics, mitochondrial morphologies, calcium homeostasis and increased susceptibility to glutamate toxicity [81, 107,108,109]. Neurons derived from patients carrying the heteroplasmic m.13513G > A mutation exhibited defects in OXPHOS function and calcium signaling [110]. Furthermore, brain organoids carrying the m.8993T > G mutation displayed developed abnormally, showing reduced size and altered cortical architecture [111]. Motor neurons from iPSCs bearing the m.9154C > T mutation were compatible with motor neuron identity and function to determine the threshold effect of heteroplasmic ATP6 mutation in reprogramming, Notch hyperactivation and motor neuron metabolism [112].

LHON

The LHON disease model generated by iPSCs derived from patients to obtain terminally differentiated RGCs allowed us to elucidate the tissue-specific effects, and develop effective therapeutic strategies for RGC disorders [113,114,115,116,117]. LHON is the most common maternally inherited eye disease characterized by selective loss of RGCs and their axons, which leads to rapidly progressive bilateral vision loss in young adults [118, 119]. The homoplasmic or nearly homoplasmic ND4 11778G > A, ND1 3460G > A, and ND6 14484T > C mutations, which affects the essential component of complex I (NADH: ubiquinone oxidoreductase), accounted for the majority of LHON cases worldwide [118,119,120,121,122,123]. The primary defects in the LHON-linked mtDNA mutation were the complex I deficiencies, thereby causing the respiratory deficiency, diminishing ATP synthesis, an increasing generation of reactive oxygen species and impaired apoptosis and mitophagy [29, 33, 35, 121, 122]. The incomplete penetrance and gender bias in patients presenting with optic neuropathy suggested nuclear modifier genes such as X-linked modifier necessary for the phenotypic manifestation of LHON-associated mtDNA mutations [28, 29, 116, 118, 119, 124]. Several LHON disease models were generated by iPSCs derived from patient’s cells bearing the m.11778G > A, m.14484T > C or m.3460G > A mutation [113,114,115,116,117, 125, 126]. These patients-derived iPSCs from LHON families included asymptomatic and symptomatic subjects carrying the mtDNA mutation(s) and control subjects. Those iPSCs were differentiated into neural progenitor cells and subsequently induced RGC-like cells using a stepwise differentiation procedure. These RGC-like cells derived from revealed more drastic reductions in oxygen consumption rates, levels of mitochondrial ATP and increasing productions of reactive oxygen species than those in other cell models such as cybrids [115, 117, 121, 122, 125, 126]. These mitochondrial dysfunctions increased susceptibility to apoptotic process for RGC degenerations [115, 125, 126]. These RGC-like cells derived from symptomatic individual harboring both mtDNA and nuclear modifier PRICKLE3 or YARS2 mutations exhibited greater defects in neuronal differentiation, morphology including reduced area of soma, numbers of neurites and shortened length of axons, electrophysiological properties than those in RGC-like cells derived from asymptomatic subjects bearing mtDNA mutation(s) [125, 126]. The nuclear modifier allele such as YARS2 c.572G > T mutation in iPSC lines from a syndromic individual can be corrected by CRISPR/Cas9. Strikingly, the genetic correction of YARS2 c.572G > T mutation led to morphologic and functional recovery of patient-derived RGC-like cells [126]. These findings provide new insights into pathophysiology of LHON arising from RGC-specific mitochondrial dysfunctions and a step toward therapeutic interventions for this disease.

Maternally inherited deafness

mtDNA mutations are associated with the maternal transmission of syndromic deafness (hearing loss with other medical problems such as diabetes) and nonsyndromic deafness (hearing loss is the only obvious medical problem). The 12S rRNA 1555A > G and 1494C > T mutations have been associated with both aminoglycoside-induced and nonsyndromic deafness in many families worldwide [23, 24, 39]. The syndromic deafness-associated mutations included the MELAS-associated tRNALeu(UUR) 3243A > G mutation and MERRF-associated tRNALys 8344A > G mutation and maternally inherited diabetes and deafness (MIDD)-associated tRNAGlu 14692A > G mutation [19, 20, 127]. The nonsyndromic deafness-associated mtDNA mutations included tRNASer(UCN) 7511T > C, tRNAHis 12201T > C, tRNAAsp 7551A > G, tRNAIle 4295A > G andtRNACys 5783C > T mutations [128,129,130,131,132]. In particular, the m.1555A > G mutation is a primary factor underlying the development of hearing loss and TRMU allele (c.28G > T, p.Ala10Sser) encoding tRNA thiouridylase interact with m.1555A > G mutation to cause hearing loss [31, 32]. Hsu et al. generated an iPSC line from the peripheral blood mononuclear cells derived from a hearing-impaired patient carrying the m.1555A > G mutation [133]. Chen et al. generated iPSCs from lymphoblastoid cell lines derived from members of an Arab–Israeli family (asymptomatic individual carrying only m.1555A > G mutation, symptomatic individual bearing both m.1555A > G and TRMU c.28G > T mutations, and control subject) [134]. The c.28G > T mutation in iPSC lines from a hearing-impaired subject harboring both m.1555A > G and TRMU c.28G > T mutations was corrected by CRISPR/Cas9 [134]. Four iPSC lines were differentiated into otic epithelial progenitors (OEPs) and subsequent inner ear hair cell-like cells (HCs). The iPSCs bearing the m.1555A > G mutation displayed mild defects in differentiation and resultant HC-like cells displayed mild defects in morphology and electrophysiological properties. Those HC-like cells harboring m.1555A > G and TRMU c.28G > T mutations exihibited greater defects in the development, morphology and functions than those in cells bearing only m.1555A > G mutation. Genetic correction of TRMU c.28G > T mutation resulted in morphologic and functional recovery of patient derived HC-like cells [134]. These findings provided new insights into pathophysiology of deafness arising from HC-specific mitochondrial deficiencies and a step toward therapeutic interventions for this disease.

Diversity of mtDNA variants (non-pathogenic) in the iPSCs models

The genetic and functional integrity of iPSCs is an important consideration towards high-quality therapeutic application. Aside from the pathogenic mtDNA mutations, mtDNA variations accumulated later in life affect mitochondrial functions. Perales-Clemente et al. showed that one possible cause of intra-person iPSC variability was the impaired mitochondrial respiration detected in iPSC-derived cardiomyocytes, due to expanded mtDNA mutations non-related with human diseases [135]. Kang et al. examined the accumulation of mtDNA mutations in iPSCs derived from young and elderly individuals [136].These mtDNA mutations randomly arose within cultured cells and resulted in respiratory deficiency. The frequency of mutations in iPSCs increased with age [136]. In support of these observations, by deep-sequencing the iPSCs, increasing numbers of variants in mtDNA were detected during time in culture [68, 137, 138]. Wei et al. reported a rather high mutation rate at 8.62 × 10–5/base pair of iPSC mtDNAs [68]. These variants indicate appearance of chromosomal abnormalities [137]. The mutation rates of mtDNA are inherently higher than that of nuclear DNA. One of the possible hypotheses is that mtDNA is extensively exposed to high levels of ROS, and the fidelity of mitochondrial DNA polymerase gamma is inadequate in mtDNA replication. Specific mtDNA variants lead to faster replication causing heteroplasmy shifts over time. During cytokinesis, asymmetric partitioning of mtDNA variants is also hypothesized to affect the segregation of variants [26]. Furthermore, de novo mutations occurring in the mtDNA of cultured iPSCs were reported to encode neoantigens that provoke specific immune responses, suggesting that iPSCs and their derivatives are not inherently immunologically available for autologous transplantation [138]. Recently, the Schrepfer group showed that hypoimmunogenic iPSCs with inactivated major histocompatibility complex (MHC) class I and II genes, and over-expression of CD47 lost their immunogenicity [139]. Derivatives of these hypoimmunogenic iPSCs evaded immune rejection in vivo, providing a more successful and more economically viable strategy for regenerative medicine [139]. These results highlight the necessity of monitoring mtDNA mutations in iPSCs and immunologically properties to ensure the effectiveness in disease studies and clinical applications.

iPSCs for therapeutic interventions for mitochondrial diseases

These iPSC patient-derived disease models can be used for developing therapeutic intervention strategy for mitochondrial diseases. Mitochondria targeting restriction endonucleases (mitoREs), TALENs (mitoTALENs), and ZFNs (mtZFNs) have been used to eliminate mutated mtDNA molecules in the iPSCs or organoid bearing heteroplasmic mtDNA mutations including the m.3243A > G mutation (Fig. 3) [91, 92, 140]. The base editing for mtDNA in iPSCs may enable to generate mitochondrial disease model and develop potential therapeutic intervention. In fact, the conventional CRISPR/Cas systems relying on small guide RNAs (sgRNAs) are incompatible with mitochondria, because sgRNAs are hardly imported into mitochondria [140]. The Liu group engineered inactive split-DddA halves until brought together on target DNA in mitochondria by adjacently bound programmable DNA-binding proteins (TALE) [141]. The employment of this RNA-free DddA-derived cytosine base editors (DdCBEs) achieved in catalysing C•G-to-T•A conversions in human mtDNA. Moreover, the Kim group engineered deoxyadenosine deaminase derived from the bacterial TadA protein with TALE to catalyze A-to-G conversions [142], extending the application range for mtDNA base editing. The combination of iPSC technology with these gene editing technology will provides the powerful tools to generate disease models and develop the effectively therapeutic approaches for these diseases.

Fig. 3
figure 3

Therapeutic approaches combining iPSCs with mtDNA gene editing technology. For iPSCs derived from patients carrying heteroplasmic mtDNA mutation, mitochondrial restriction endonucleases (mtREs), mitochondrial-targeted transcription activator-like effector nucleases (mtTALENs), and zinc-finger nucleases (mtZFN) are utilized to eliminate the mutant mtDNA molecules. For iPSCs derived from patients bearing homoplasmic mtDNA mutation, mtDNA base editing technology is preferred to correct the mutations. After gene editing, the corrected iPSCs differentiate to target cells with remodeling cellular functions. The black circles indicate mtDNA. The red dots on black circles denote mtDNA mutations. The green color of mitochondria indicates normal condition, and the pink color of mitochondria show damaged function. Other colors of cells do not indicate any information

Conclusions

Despite the discovery of maternally inherited diseases caused by mtDNA mutations 35 years ago, there are many major challenges in elucidating the pathophysiology and developing the treatments for these disorders. To overcome these difficulties, many groups have turned to hiPSCs derived from patients carrying the mtDNA mutation(s) to model mitochondrial diseases. Here, we review the recent advances in patients-derived iPSCs as mitochondrial disease models. Those patients-derived iPSCs have been differentiated into specific cells (2-dimension differentiation) such as retinal ganglion cells, neuronal cells, muscle, cardiac cells and organoids (3-dimension differentiation). The iPSC-based modeling, in combining with mtDNA base editing technology, lead to deep elucidation of pathologic mechanism and opening new revenue to interventional strategies for these disorders.

Availability of data and materials

Not applicable.

Abbreviations

mtDNA:

Mitochondrial DNA

iPSCs:

Induced pluripotent stem cells

OXPHOS:

Oxidative phosphorylation

ROS:

Reactive oxygen species

POLG:

DNA polymerase gamma

KSS:

Kearns–Sayre syndrome

PS:

Pearson syndrome

CPEO:

Chronic progressive external ophthalmoplegia

MELAS:

Mitochondrial encephalomyopathy lactic acidosis, and stroke-like episodes

MERRF:

Myoclonic epilepsy associated with ragged-red fibers

LHON:

Leber hereditary optic neuropathy

ATP6:

Mitochondrially encoded atp synthase membrane subunit 6

ND4:

Mitochondrially encoded nadh:ubiquinone oxidoreductase core subunit 4

ND1:

Mitochondrially encoded nadh:ubiquinone oxidoreductase core subunit 1

ND6:

Mitochondrially encoded nadh:ubiquinone oxidoreductase core subunit 6

PRICKLE3:

Prickle planar cell polarity protein 3

YARS2:

Mitochondrial tyrosyl-tRNA synthetase

TRMU:

Methylaminomethyl-2-thiouridylate-methyltransferase

Oct3/4:

Octamer-binding transcription factor 3/4

Sox2:

SRY-box transcription factor 2

c-Myc:

MYC proto-oncogene, BHLH transcription factor

Klf4:

Kruppel-like factor 4

PBMCs:

Peripheral blood mononuclear cells

AZA:

5-Azacytidine

NPCs:

Neural progenitor cells

ND5:

Mitochondrially encoded nadh:ubiquinone oxidoreductase core subunit 5

CO3:

Mitochondrially encoded cytochrome c oxidase III

CYB:

Mitochondrially encoded cytochrome b

RPE:

Retinal pigment epithelial

DAPT:

(N-[N-(3, 5-difluorophenacetyl)-l-alanyl]-s-phenylglycinet-butyl ester)

TALEN:

Transcription activator-like effector nuclease

TALM:

Tryptolinamide

SCNT:

Somatic cell nuclear transfer

RGC:

Retinal ganglion cells

KIF5A:

Kinesin family member 5A

MIDD:

Maternally inherited diabetes and deafness

OEP:

Otic epithelial progenitor

HC:

Hair cell

MHC:

Major histocompatibility complex

DdCBE:

DddA-derived cytosine base editor

References

  1. Wallace DC. Mitochondrial genetic medicine. Nat Genet. 2018;50:1642–9.

    CAS  PubMed  Google Scholar 

  2. Nunnari J, Suomalainen A. Mitochondria: in sickness and in health. Cell. 2012;148(6):1145–59.

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Chan DC. Mitochondria: dynamic organelles in disease, aging, and development. Cell. 2006;125(7):1241–52.

    CAS  PubMed  Google Scholar 

  4. Picard M, Shirihai OS. Mitochondrial signal transduction. Cell Metab. 2022;34(11):1620–53.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Hallberg BM, Larsson NG. Making proteins in the powerhouse. Cell Metab. 2014;20(2):226–40.

    PubMed  Google Scholar 

  7. Anderson S, Bankier AT, Barrell BG, de Bruijn MH, Coulson AR, Drouin J, et al. Sequence and organization of the human mitochondrial genome. Nature. 1981;290(5806):457–65.

    CAS  PubMed  Google Scholar 

  8. Falkenberg M. Mitochondrial DNA replication in mammalian cells: overview of the pathway. Essays Biochem. 2018;62(3):287–96.

    PubMed  PubMed Central  Google Scholar 

  9. Chen X, Prosser R, Simonetti S, Sadlock J, Jagiello G, Schon EA. Rearranged mitochondrial genomes are present in human oocytes. Am J Hum Genet. 1995;57(2):239–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  10. D’Erchia AM, Atlante A, Gadaleta G, Pavesi G, Chiara M, De Virgilio C, et al. Tissue-specific mtDNA abundance from exome data and its correlation with mitochondrial transcription, mass and respiratory activity. Mitochondrion. 2015;20:13–21.

    CAS  PubMed  Google Scholar 

  11. Popova D, Bhide P, D’Antonio F, Basnet P, Acharya G. Sperm mitochondrial DNA copy numbers in normal and abnormal semen analysis: a systematic review and meta-analysis. BJOG. 2022;129(9):1434–46.

    CAS  PubMed  Google Scholar 

  12. May-Panloup P, Boguenet M, Hachem HE, Bouet PE, Reynier P. Embryo and its mitochondria. Antioxidants (Basel). 2021;10(2):139.

    CAS  PubMed  Google Scholar 

  13. Tuppen HA, Blakely EL, Turnbull DM, Taylor RW. Mitochondrial DNA mutations and human disease. Biochim Biophys Acta. 2010;1797(2):113–28.

    CAS  PubMed  Google Scholar 

  14. Viscomi C, Zeviani M. MtDNA-maintenance defects: syndromes and genes. J Inherit Metab Dis. 2017;40(4):587–99.

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Schapira AH. Mitochondrial diseases. Lancet. 2012;379(9828):1825–34.

    CAS  PubMed  Google Scholar 

  16. He Y, Wu J, Dressman DC, Iacobuzio-Donahue C, Markowitz SD, Velculescu VE, et al. Heteroplasmic mitochondrial DNA mutations in normal and tumour cells. Nature. 2010;464(7288):610–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Gorman GS, Schaefer AM, Ng Y, Gomez N, Blakely EL, Alston CL, et al. Prevalence of nuclear and mitochondrial DNA mutations related to adult mitochondrial disease. Ann Neurol. 2015;77(5):753–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Gorman GS, Chinnery PF, DiMauro S, Hirano M, Koga Y, McFarland R, et al. Mitochondrial diseases. Nat Rev Dis Primers. 2016;2:16080.

    PubMed  Google Scholar 

  19. Goto Y, Nonaka I, Horai S. A mutation in the tRNALeu(UUR) gene associated with the MELAS subgroup of mitochondrial encephalomyopathies. Nature. 1990;348(6302):651–3.

    CAS  PubMed  Google Scholar 

  20. Shoffner JM, Lott MT, Lezza AM, Seibel P, Ballinger SW, Wallace DC. Myoclonic epilepsy and ragged-red fiber disease (MERRF) is associated with a mitochondrial DNA tRNALys mutation. Cell. 1990;61(6):931–7.

    CAS  PubMed  Google Scholar 

  21. Santorelli FM, Shanske S, Macaya A, DeVivo DC, DiMauro S. The mutation at nt 8993 of mitochondrial DNA is a common cause of Leigh’s syndrome. Ann Neurol. 1993;34(6):827–34.

    CAS  PubMed  Google Scholar 

  22. Zeviani M, Carelli V. Mitochondrial retinopathies. Int J Mol Sci. 2021;23(1):210.

    PubMed  PubMed Central  Google Scholar 

  23. Prezant TR, Agapian JV, Bohlman MC, Bu X, Oztas S, Qiu WQ, et al. Mitochondrial ribosomal RNA mutation associated with both antibiotic-induced and non-syndromic deafness. Nat Genet. 1993;4(3):289–94.

    CAS  PubMed  Google Scholar 

  24. Zhao H, Li R, Wang Q, Yan Q, Deng JH, Han D, et al. Maternally inherited aminoglycoside-induced and nonsyndromic deafness is associated with the novel C1494T mutation in the mitochondrial 12S rRNA gene in a large Chinese family. Am J Hum Genet. 2004;74(1):139–52.

    CAS  PubMed  Google Scholar 

  25. Sproule DM, Kaufmann P. Mitochondrial encephalopathy, lactic acidosis, and strokelike episodes: basic concepts, clinical phenotype, and therapeutic management of MELAS syndrome. Ann Ny Acad Sci. 2008;1142:133–58.

    CAS  PubMed  Google Scholar 

  26. Chiaratti MR, Chinnery PF. Modulating mitochondrial DNA mutations: factors shaping heteroplasmy in the germ line and somatic cells. Pharmacol Res. 2022;185: 106466.

    CAS  PubMed  Google Scholar 

  27. Nissanka N, Moraes CT. Mitochondrial DNA heteroplasmy in disease and targeted nuclease-based therapeutic approaches. EMBO Rep. 2020;21(3): e49612.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Yu J, Liang X, Ji Y, Ai C, Liu J, Zhu L, et al. PRICKLE3 linked to ATPase biogenesis manifested Leber’s hereditary optic neuropathy. J Clin Invest. 2020;130(9):4935–46.

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Jiang P, Jin X, Peng Y, Wang M, Liu H, Liu X, et al. The exome sequencing identified the mutation in YARS2 encoding the mitochondrial tyrosyl-tRNA synthetase as a nuclear modifier for the phenotypic manifestation of Leber’s hereditary optic neuropathy-associated mitochondrial DNA mutation. Hum Mol Genet. 2016;25(3):584–96.

    CAS  PubMed  Google Scholar 

  30. Caporali L, Maresca A, Capristo M, Del Dotto V, Tagliavini F, Valentino ML, et al. Incomplete penetrance in mitochondrial optic neuropathies. Mitochondrion. 2017;36:130–7.

    CAS  PubMed  Google Scholar 

  31. Guan MX, Yan Q, Li X, Bykhovskaya Y, Gallo-Teran J, Hajek P, et al. Mutation in TRMU related to transfer RNA modification modulates the phenotypic expression of the deafness-associated mitochondrial 12S ribosomal RNA mutations. Am J Hum Genet. 2006;79(2):291–302.

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Meng F, Cang X, Peng Y, Li R, Zhang Z, Li F, et al. Biochemical evidence for a nuclear modifier allele (A10S) in TRMU (Methylaminomethyl-2-thiouridylate-methyltransferase) related to mitochondrial tRNA modification in the phenotypic manifestation of deafness-associated 12S rRNA Mutation. J Biol Chem. 2017;292(7):2881–92.

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Ji Y, Zhang J, Yu J, Wang Y, Lu Y, Liang M, et al. Contribution of mitochondrial ND1 3394T>C mutation to the phenotypic manifestation of Leber’s hereditary optic neuropathy. Hum Mol Genet. 2019;28(9):1515–29.

    CAS  PubMed  Google Scholar 

  34. Qu J, Li R, Zhou X, Tong Y, Lu F, Qian Y, et al. The novel A4435G mutation in the mitochondrial tRNAMet may modulate the phenotypic expression of the LHON-associated ND4 G11778A mutation. Invest Ophthalmol Vis Sci. 2006;47(2):475–83.

    PubMed  Google Scholar 

  35. Ji Y, Zhang J, Lu Y, Yi Q, Chen M, Xie S, et al. Complex I mutations synergize to worsen the phenotypic expression of Leber’s hereditary optic neuropathy. J Biol Chem. 2020;295(38):13224–38.

    CAS  PubMed  PubMed Central  Google Scholar 

  36. He Q, He X, Xiao Y, Zhao Q, Ye Z, Cui L, et al. Tissue-specific expression atlas of murine mitochondrial tRNAs. J Biol Chem. 2021;297(2): 100960.

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Nesbitt V, Pitceathly RD, Turnbull DM, Taylor RW, Sweeney MG, Mudanohwo EE, et al. The UK MRC Mitochondrial disease patient cohort study: clinical phenotypes associated with the m.3243A>G mutation—implications for diagnosis and management. J Neurol Neurosurg Psychiatry. 2013;84(8):936–8.

    PubMed  Google Scholar 

  38. Carelli V, Ross-Cisneros FN, Sadun AA. Mitochondrial dysfunction as a cause of optic neuropathies. Prog Retin Eye Res. 2004;23(1):53–89.

    CAS  PubMed  Google Scholar 

  39. Guan MX. Mitochondrial 12S rRNA mutations associated with aminoglycoside ototoxicity. Mitochondrion. 2011;11(2):237–45.

    CAS  PubMed  Google Scholar 

  40. Lightowlers RN, Taylor RW, Turnbull DM. Mutations causing mitochondrial disease: what is new and what challenges remain? Science. 2015;349(6255):1494–9.

    CAS  PubMed  Google Scholar 

  41. Yilmaz A, Benvenisty N. Defining human pluripotency. Cell Stem Cell. 2019;25(1):9–22.

    CAS  PubMed  Google Scholar 

  42. Yamanaka S. Pluripotent stem cell-based cell therapy-promise and challenges. Cell Stem Cell. 2020;27(4):523–31.

    CAS  PubMed  Google Scholar 

  43. Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 2006;126(4):663–76.

    CAS  PubMed  Google Scholar 

  44. Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell. 2007;131(5):861–72.

    CAS  PubMed  Google Scholar 

  45. Bailly A, Milhavet O, Lemaitre JM. RNA-based strategies for cell reprogramming toward pluripotency. Pharmaceutics. 2022;14(2):317.

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Yu J, Vodyanik MA, Smuga-Otto K, Antosiewicz-Bourget J, Frane JL, Tian S, et al. Induced pluripotent stem cell lines derived from human somatic cells. Science. 2007;318(5858):1917–20.

    CAS  PubMed  Google Scholar 

  47. Dowey SN, Huang X, Chou BK, Ye Z, Cheng L. Generation of integration-free human induced pluripotent stem cells from postnatal blood mononuclear cells by plasmid vector expression. Nat Protoc. 2012;7(11):2013–21.

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Zhou T, Benda C, Dunzinger S, Huang Y, Ho JC, Yang J, et al. Generation of human induced pluripotent stem cells from urine samples. Nat Protoc. 2012;7(12):2080–9.

    CAS  PubMed  Google Scholar 

  49. Ye Z, Zhan H, Mali P, Dowey S, Williams DM, Jang YY, et al. Human-induced pluripotent stem cells from blood cells of healthy donors and patients with acquired blood disorders. Blood. 2009;114(27):5473–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Staerk J, Dawlaty MM, Gao Q, Maetzel D, Hanna J, Sommer CA, et al. Reprogramming of human peripheral blood cells to induced pluripotent stem cells. Cell Stem Cell. 2010;7(1):20–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Aasen T, Izpisua Belmonte JC. Isolation and cultivation of human keratinocytes from skin or plucked hair for the generation of induced pluripotent stem cells. Nat Protoc. 2010;5(2):371–82.

    CAS  PubMed  Google Scholar 

  52. Brouwer M, Zhou H, Nadif KN. Choices for induction of pluripotency: recent developments in human induced pluripotent stem cell reprogramming strategies. Stem Cell Rev Rep. 2016;12(1):54–72.

    CAS  PubMed  Google Scholar 

  53. Esteban MA, Wang T, Qin B, Yang J, Qin D, Cai J, et al. Vitamin C enhances the generation of mouse and human induced pluripotent stem cells. Cell Stem Cell. 2010;6(1):71–9.

    CAS  PubMed  Google Scholar 

  54. Huangfu D, Osafune K, Maehr R, Guo W, Eijkelenboom A, Chen S, et al. Induction of pluripotent stem cells from primary human fibroblasts with only Oct4 and Sox2. Nat Biotechnol. 2008;26(11):1269–75.

    CAS  PubMed  Google Scholar 

  55. Mali P, Chou BK, Yen J, Ye Z, Zou J, Dowey S, et al. Butyrate greatly enhances derivation of human induced pluripotent stem cells by promoting epigenetic remodeling and the expression of pluripotency-associated genes. Stem Cells. 2010;28(4):713–20.

    CAS  PubMed  Google Scholar 

  56. Mikkelsen TS, Hanna J, Zhang XL, Ku MC, Wernig M, Schorderet P, et al. Dissecting direct reprogramming through integrative genomic analysis. Nature. 2008;454(7200):49-U1.

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Kaji K, Norrby K, Paca A, Mileikovsky M, Mohseni P, Woltjen K. Virus-free induction of pluripotency and subsequent excision of reprogramming factors. Nature. 2009;458(7239):771–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Woltjen K, Michael IP, Mohseni P, Desai R, Mileikovsky M, Hamalainen R, et al. piggyBac transposition reprograms fibroblasts to induced pluripotent stem cells. Nature. 2009;458(7239):766–70.

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Zhou WB, Freed CR. Adenoviral gene delivery can reprogram human fibroblasts to induced pluripotent stem cells. Stem Cells. 2009;27(11):2667–74.

    CAS  PubMed  Google Scholar 

  60. Ye L, Muench MO, Fusaki N, Beyer AI, Wang J, Qi Z, et al. Blood cell-derived induced pluripotent stem cells free of reprogramming factors generated by Sendai viral vectors. Stem Cells Transl Med. 2013;2(8):558–66.

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Yu J, Hu K, Smuga-Otto K, Tian S, Stewart R, Slukvin II, et al. Human induced pluripotent stem cells free of vector and transgene sequences. Science. 2009;324(5928):797–801.

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Zhou H, Wu S, Joo JY, Zhu S, Han DW, Lin T, et al. Generation of induced pluripotent stem cells using recombinant proteins. Cell Stem Cell. 2009;4(5):381–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Warren L, Manos PD, Ahfeldt T, Loh YH, Li H, Lau F, et al. Highly efficient reprogramming to pluripotency and directed differentiation of human cells with synthetic modified mRNA. Cell Stem Cell. 2010;7(5):618–30.

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Yoshioka N, Gros E, Li HR, Kumar S, Deacon DC, Maron C, et al. Efficient generation of human iPSCs by a synthetic self-replicative RNA. Cell Stem Cell. 2013;13(2):246–54.

    CAS  PubMed  Google Scholar 

  65. Anokye-Danso F, Trivedi CM, Juhr D, Gupta M, Cui Z, Tian Y, et al. Highly efficient miRNA-mediated reprogramming of mouse and human somatic cells to pluripotency. Cell Stem Cell. 2011;8(4):376–88.

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Hou P, Li Y, Zhang X, Liu C, Guan J, Li H, et al. Pluripotent stem cells induced from mouse somatic cells by small-molecule compounds. Science. 2013;341(6146):651–4.

    CAS  PubMed  Google Scholar 

  67. Shi Y, Inoue H, Wu JC, Yamanaka S. Induced pluripotent stem cell technology: a decade of progress. Nat Rev Drug Discov. 2017;16(2):115–30.

    CAS  PubMed  Google Scholar 

  68. Wei W, Gaffney DJ, Chinnery PF. Cell reprogramming shapes the mitochondrial DNA landscape. Nat Commun. 2021;12(1):5241.

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Maceluch JA, Niedziela M. The clinical diagnosis and molecular genetics of kearns-sayre syndrome: a complex mitochondrial encephalomyopathy. Pediatr Endocrinol Rev. 2006;4(2):117–37.

    PubMed  Google Scholar 

  70. Zeviani M, Moraes CT, DiMauro S, Nakase H, Bonilla E, Schon EA, et al. Deletions of mitochondrial DNA in Kearns-Sayre syndrome. Neurology. 1988;38(9):1339–46.

    CAS  PubMed  Google Scholar 

  71. Rotig A, Colonna M, Bonnefont JP, Blanche S, Fischer A, Saudubray JM, et al. Mitochondrial DNA deletion in Pearson’s marrow/pancreas syndrome. Lancet. 1989;1(8643):902–3.

    CAS  PubMed  Google Scholar 

  72. Rotig A, Cormier V, Blanche S, Bonnefont JP, Ledeist F, Romero N, et al. Pearson’s marrow-pancreas syndrome. A multisystem mitochondrial disorder in infancy. J Clin Invest. 1990;86(5):1601–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Yoshimi A, Ishikawa K, Niemeyer C, Grunert SC. Pearson syndrome: a multisystem mitochondrial disease with bone marrow failure. Orphanet J Rare Dis. 2022;17(1):379.

    PubMed  PubMed Central  Google Scholar 

  74. Moraes CT, DiMauro S, Zeviani M, Lombes A, Shanske S, Miranda AF, et al. Mitochondrial DNA deletions in progressive external ophthalmoplegia and Kearns-Sayre syndrome. New Engl J Med. 1989;320(20):1293–9.

    CAS  PubMed  Google Scholar 

  75. Cherry AB, Gagne KE, McLoughlin EM, Baccei A, Gorman B, Hartung O, et al. Induced pluripotent stem cells with a mitochondrial DNA deletion. Stem Cells. 2013;31(7):1287–97.

    CAS  PubMed  Google Scholar 

  76. Hernández-Ainsa C, López-Gallardo E, García-Jiménez MC, Climent-Alcalá FJ, Rodríguez-Vigil C, García Fernández de Villalta M, et al. Development and characterization of cell models harbouring mtDNA deletions for in vitro study of Pearson syndrome. Dis Model Mech. 2022;15(3):dmm049083.

    PubMed  PubMed Central  Google Scholar 

  77. Lester Sequiera G, Srivastava A, Alagarsamy KN, Rockman-Greenberg C, Dhingra S. Generation and evaluation of isogenic iPSC as a source of cell replacement therapies in patients with Kearns Sayre Syndrome. Cells. 2021;10(3):568.

    PubMed  PubMed Central  Google Scholar 

  78. Srivastava A, Lester Sequiera G, Narayan Alagarsamy K, Rockman-Greenberg C, Dhingra S. Generation of human induced pluripotent stem cell (hiPSC) line UOMi005-A from PBMCs of a patient with Kearns-Sayre syndrome. Stem Cell Res. 2021;53: 102283.

    CAS  PubMed  Google Scholar 

  79. Hamalainen RH, Manninen T, Koivumaki H, Kislin M, Otonkoski T, Suomalainen A. Tissue- and cell-type-specific manifestations of heteroplasmic mtDNA 3243A>G mutation in human induced pluripotent stem cell-derived disease model. Proc Natl Acad Sci USA. 2013;110(38):E3622-30.

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Folmes CD, Martinez-Fernandez A, Perales-Clemente E, Li X, McDonald A, Oglesbee D, et al. Disease-causing mitochondrial heteroplasmy segregated within induced pluripotent stem cell clones derived from a patient with MELAS. Stem Cells. 2013;31(7):1298–308.

    CAS  PubMed  Google Scholar 

  81. Klein Gunnewiek TM, Van Hugte EJH, Frega M, Guardia GS, Foreman K, Panneman D, et al. m.3243A>G-induced mitochondrial dysfunction impairs human neuronal development and reduces neuronal network activity and synchronicity. Cell Rep. 2020;31(3):107538.

    CAS  PubMed  Google Scholar 

  82. Ma H, Folmes CD, Wu J, Morey R, Mora-Castilla S, Ocampo A, et al. Metabolic rescue in pluripotent cells from patients with mtDNA disease. Nature. 2015;524(7564):234–8.

    CAS  PubMed  Google Scholar 

  83. Hatakeyama H, Katayama A, Komaki H, Nishino I, Goto Y. Molecular pathomechanisms and cell-type-specific disease phenotypes of MELAS caused by mutant mitochondrial tRNATrp. Acta Neuropathol Commun. 2015;3:52.

    PubMed  PubMed Central  Google Scholar 

  84. Chichagova V, Hallam D, Collin J, Buskin A, Saretzki G, Armstrong L, et al. Human iPSC disease modelling reveals functional and structural defects in retinal pigment epithelial cells harbouring the m.3243A>G mitochondrial DNA mutation. Sci Rep. 2017;7(1):12320.

    PubMed  PubMed Central  Google Scholar 

  85. Bhattacharya S, Yin J, Huo W, Chaum E. Modeling of mitochondrial bioenergetics and autophagy impairment in MELAS-mutant iPSC-derived retinal pigment epithelial cells. Stem Cell Res Ther. 2022;13(1):260.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Pek NMQ, Phua QH, Ho BX, Pang JKS, Hor JH, An O, et al. Mitochondrial 3243A > G mutation confers pro-atherogenic and pro-inflammatory properties in MELAS iPS derived endothelial cells. Cell Death Dis. 2019;10(11):802.

    PubMed  PubMed Central  Google Scholar 

  87. Ryytty S, Modi SR, Naumenko N, Shakirzyanova A, Rahman MO, Vaara M, et al. Varied responses to a high m.3243A>G mutation load and respiratory chain dysfunction in patient-derived cardiomyocytes. Cells. 2022;11(16):2593.

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Lin DS, Huang YW, Ho CS, Hung PL, Hsu MH, Wang TJ, et al. Oxidative insults and mitochondrial DNA mutation promote enhanced autophagy and mitophagy compromising cell viability in pluripotent cell model of mitochondrial disease. Cells. 2019;8(1):65.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Yokota M, Hatakeyama H, Ono Y, Kanazawa M, Goto YI. Mitochondrial respiratory dysfunction disturbs neuronal and cardiac lineage commitment of human iPSCs. Cell Death Dis. 2017;8(1): e2551.

    PubMed  PubMed Central  Google Scholar 

  90. Winanto N, Khong ZJ, Soh BS, Fan Y, Ng SY. Organoid cultures of MELAS neural cells reveal hyperactive Notch signaling that impacts neurodevelopment. Cell Death Dis. 2020;11(3):182.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Yang Y, Wu H, Kang X, Liang Y, Lan T, Li T, et al. Targeted elimination of mutant mitochondrial DNA in MELAS-iPSCs by mitoTALENs. Protein Cell. 2018;9(3):283–97.

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Yahata N, Matsumoto Y, Omi M, Yamamoto N, Hata R. TALEN-mediated shift of mitochondrial DNA heteroplasmy in MELAS-iPSCs with m.13513G>A mutation. Sci Rep. 2017;7(1):15557.

    PubMed  PubMed Central  Google Scholar 

  93. Kobayashi H, Hatakeyama H, Nishimura H, Yokota M, Suzuki S, Tomabechi Y, et al. Chemical reversal of abnormalities in cells carrying mitochondrial DNA mutations. Nat Chem Biol. 2021;17(3):335–43.

    CAS  PubMed  Google Scholar 

  94. Ruiz-Pesini E, Lott MT, Procaccio V, Poole JC, Brandon MC, Mishmar D, et al. An enhanced MITOMAP with a global mtDNA mutational phylogeny. Nucleic Acids Res. 2007;35(Database issue):D823-8.

    CAS  PubMed  Google Scholar 

  95. El-Hattab AW, Adesina AM, Jones J, Scaglia F. MELAS syndrome: clinical manifestations, pathogenesis, and treatment options. Mol Genet Metab. 2015;116(1–2):4–12.

    CAS  PubMed  Google Scholar 

  96. Yoneda M, Chomyn A, Martinuzzi A, Hurko O, Attardi G. Marked replicative advantage of human mtDNA carrying a point mutation that causes the MELAS encephalomyopathy. Proc Natl Acad Sci USA. 1992;89(23):11164–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Li R, Guan MX. Human mitochondrial leucyl-tRNA synthetase corrects mitochondrial dysfunctions due to the tRNALeu(UUR) A3243G mutation, associated with mitochondrial encephalomyopathy, lactic acidosis, and stroke-like symptoms and diabetes. Mol Cell Biol. 2010;30(9):2147–54.

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Tan GC, Mazzoni EO, Wichterle H. Iterative role of Notch signaling in spinal motor neuron diversification. Cell Rep. 2016;16(4):907–16.

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Berezovska O, McLean P, Knowles R, Frosh M, Lu FM, Lux SE, et al. Notch1 inhibits neurite outgrowth in postmitotic primary neurons. Neuroscience. 1999;93(2):433–9.

    CAS  PubMed  Google Scholar 

  100. Finsterer J, Zarrouk-Mahjoub S. Management of epilepsy in MERRF syndrome. Seizure. 2017;50:166–70.

    PubMed  Google Scholar 

  101. Chou SJ, Tseng WL, Chen CT, Lai YF, Chien CS, Chang YL, et al. Impaired ROS scavenging system in human induced pluripotent stem cells generated from patients with MERRF syndrome. Sci Rep. 2016;6:23661.

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Chen YC, Tsai CL, Wei YH, Wu YT, Hsu WT, Lin HC, et al. ATOH1/RFX1/RFX3 transcription factors facilitate the differentiation and characterisation of inner ear hair cell-like cells from patient-specific induced pluripotent stem cells harbouring A8344G mutation of mitochondrial DNA. Cell Death Dis. 2018;9(4):437.

    PubMed  PubMed Central  Google Scholar 

  103. Villanueva-Paz M, Povea-Cabello S, Villalon-Garcia I, Suarez-Rivero JM, Alvarez-Cordoba M, de la Mata M, et al. Pathophysiological characterization of MERRF patient-specific induced neurons generated by direct reprogramming. Biochim Biophys Acta Mol Cell Res. 2019;1866(5):861–81.

    CAS  PubMed  Google Scholar 

  104. Rahman S. Leigh syndrome. Handbook Clin Neurol. 2023;194:43–63.

    Google Scholar 

  105. Lake NJ, Compton AG, Rahman S, Thorburn DR. Leigh syndrome: one disorder, more than 75 monogenic causes. Ann Neurol. 2016;79(2):190–203.

    PubMed  Google Scholar 

  106. Tatuch Y, Christodoulou J, Feigenbaum A, Clarke JTR, Wherret J, Smith C, et al. Heteroplasmic mtDNA mutation (T-G) at 8993 can cause leigh disease when the percentage of abnormal mtDNA is high. Am J Hum Genet. 1992;50(4):852–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Lorenz C, Lesimple P, Bukowiecki R, Zink A, Inak G, Mlody B, et al. Human iPSC-derived neural progenitors are an effective drugdiscovery model for neurological mtDNA disorders. Cell Stem Cell. 2017;20(5):659-74.e9.

    CAS  PubMed  Google Scholar 

  108. Zheng X, Boyer L, Jin M, Kim Y, Fan W, Bardy C, et al. Alleviation of neuronal energy deficiency by mTOR inhibition as a treatment for mitochondria-related neurodegeneration. Elife. 2016;5:e13378.

    PubMed  PubMed Central  Google Scholar 

  109. Meshrkey F, Cabrera Ayuso A, Rao RR, Iyer S. Quantitative analysis of mitochondrial morphologies in human induced pluripotent stem cells for Leigh syndrome. Stem Cell Res. 2021;57: 102572.

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Galera-Monge T, Zurita-Diaz F, Canals I, Hansen MG, Rufian-Vazquez L, Ehinger JK, et al. Mitochondrial dysfunction and calcium dysregulation in Leigh syndrome induced pluripotent stem cell derived neurons. Int J Mol Sci. 2020;21(9):3191.

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Romero-Morales AI, Robertson GL, Rastogi A, Rasmussen ML, Temuri H, McElroy GS, et al. Human iPSC-derived cerebral organoids model features of Leigh syndrome and reveal abnormal corticogenesis. Development. 2022;149(20):dev199914.

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Kenvin S, Torregrosa-Munumer R, Reidelbach M, Pennonen J, Turkia JJ, Rannila E, et al. Threshold of heteroplasmic truncating MT-ATP6 mutation in reprogramming, Notch hyperactivation and motor neuron metabolism. Hum Mol Genet. 2022;31(6):958–74.

    CAS  PubMed  Google Scholar 

  113. Wong RCB, Lim SY, Hung SSC, Jackson S, Khan S, Van Bergen NJ, et al. Mitochondrial replacement in an iPSC model of Leber’s hereditary optic neuropathy. Aging. 2017;9(4):1341–50.

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Wu YR, Wang AG, Chen YT, Yarmishyn AA, Buddhakosai W, Yang TC, et al. Bioactivity and gene expression profiles of hiPSC-generated retinal ganglion cells in MT-ND4 mutated Leber’s hereditary optic neuropathy. Exp Cell Res. 2018;363(2):299–309.

    CAS  PubMed  Google Scholar 

  115. Yang TC, Yarmishyn AA, Yang YP, Lu PC, Chou SJ, Wang ML, et al. Mitochondrial transport mediates survival of retinal ganglion cells in affected LHON patients. Hum Mol Genet. 2020;29(9):1454–64.

    CAS  PubMed  Google Scholar 

  116. Peron C, Mauceri R, Cabassi T, Segnali A, Maresca A, Iannielli A, et al. Generation of a human iPSC line, FINCBi001-A, carrying a homoplasmic m.G3460A mutation in MT-ND1 associated with Leber’s Hereditary optic Neuropathy (LHON). Stem Cell Res. 2020;48:101939.

    CAS  PubMed  Google Scholar 

  117. Danese A, Patergnani S, Maresca A, Peron C, Raimondi A, Caporali L, et al. Pathological mitophagy disrupts mitochondrial homeostasis in Leber’s hereditary optic neuropathy. Cell Rep. 2022;40(3): 111124.

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Wallace DC, Lott MT. Leber hereditary optic neuropathy: exemplar of an mtDNA disease. Handb Exp Pharmacol. 2017;240:339–76.

    CAS  PubMed  Google Scholar 

  119. Carelli V, La Morgia C, Yu-Wai-Man P. Mitochondrial optic neuropathies. Handb Clin Neurol. 2023;194:23–42.

    PubMed  Google Scholar 

  120. Jiang P, Liang M, Zhang J, Gao Y, He Z, Yu H, et al. Prevalence of mitochondrial ND4 mutations in 1281 Han Chinese subjects with Leber’s hereditary optic neuropathy. Invest Ophthalmol Vis Sci. 2015;56(8):4778–88.

    CAS  PubMed  Google Scholar 

  121. Zhang J, Ji Y, Chen J, Xu M, Wang G, Ci X, et al. Assocation Between Leber’s hereditary optic neuropathy and MT-ND1 3460G>A mutation-induced alterations in mitochondrial function, apoptosis, and mitophagy. Invest Ophthalmol Vis Sci. 2021;62(9):38.

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Liang M, Ji Y, Zhang L, Wang X, Hu C, Zhang J, et al. Leber’s hereditary optic neuropathy-associated ND6 14484T > C mutation caused pleiotropic effects on the complex I, RNA homeostasis, apoptosis and mitophagy. Hum Mol Genet. 2022;31(19):3299–312.

    CAS  PubMed  Google Scholar 

  123. Yen MY, Wang AG, Wei YH. Leber’s hereditary optic neuropathy: a multifactorial disease. Prog Retin Eye Res. 2006;25(4):381–96.

    PubMed  Google Scholar 

  124. Chen C, Chen Y, Guan MX. A peep into mitochondrial disorder: multifaceted from mitochondrial DNA mutations to nuclear gene modulation. Protein Cell. 2015;6(12):862–70.

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Nie Z, Wang C, Chen J, Ji Y, Zhang H, Zhao F, et al. Abnormal morphology and function in retinal ganglion cells derived from patients-specific iPSCs generated from individuals with Leber’s hereditary optic neuropathy. Hum Mol Genet. 2023;32(2):231–43.

    CAS  PubMed  Google Scholar 

  126. Chen JR, Chen C, Chen J, Ji Y, Lian Y, Zhang J, et al. Nuclear modifier YARS2 allele correction restored retinal ganglion cells-specific deficiencies in Leber’s hereditary optic neuropathy. Hum Mol Genet. 2023;32(9):1539–51.

    CAS  PubMed  Google Scholar 

  127. Wang M, Liu H, Zheng J, Chen B, Zhou M, Fan W, et al. A deafness- and diabetes-associated tRNA mutation causes deficient pseudouridinylation at position 55 in tRNAGlu and mitochondrial dysfunction. J Biol Chem. 2016;291(40):21029–41.

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Li X, Fischel-Ghodsian N, Schwartz F, Yan Q, Friedman RA, Guan MX. Biochemical characterization of the mitochondrial tRNASer(UCN) T7511C mutation associated with nonsyndromic deafness. Nucleic Acids Res. 2004;32(3):867–77.

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Gong S, Peng Y, Jiang P, Wang M, Fan M, Wang X, et al. A deafness-associated tRNAHis mutation alters the mitochondrial function, ROS production and membrane potential. Nucleic Acids Res. 2014;42(12):8039–48.

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Wang M, Peng Y, Zheng J, Zheng B, Jin X, Liu H, et al. A deafness-associated tRNAAsp mutation alters the m1G37 modification, aminoacylation and stability of tRNAAsp and mitochondrial function. Nucleic Acids Res. 2016;44(22):10974–85.

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Meng F, Zhou M, Xiao Y, Mao X, Zheng J, Lin J, et al. A deafness-associated tRNA mutation caused pleiotropic effects on the m1G37 modification, processing, stability and aminoacylation of tRNAIle and mitochondrial translation. Nucleic Acids Res. 2021;49(2):1075–93.

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Meng F, Jia Z, Zheng J, Ji Y, Wang J, Xiao Y, et al. A deafness-associated mitochondrial DNA mutation caused pleiotropic effects on DNA replication and tRNA metabolism. Nucleic Acids Res. 2022;50(16):9453–69.

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Hsu YH, Wu YT, Huang CY, Ho MC, Cheng YC, et al. Generation of an induced pluripotent stem cell line from a 39-year-old female patient with severe-to-profound non-syndromic sensorineural hearing loss and a A1555G mutation in the mitochondrial MTRNR1 gene. Stem Cell Res. 2017;25:245–9.

    CAS  PubMed  Google Scholar 

  134. Chen C, Guan MX. Genetic correction of TRMU allele restored the mitochondrial dysfunction-induced deficiencies in iPSCs-derived hair cells of hearing-impaired patients. Hum Mol Genet. 2022;31(18):3068–82.

    CAS  PubMed  Google Scholar 

  135. Perales-Clemente E, Cook AN, Evans JM, Roellinger S, Secreto F, Emmanuele V, et al. Natural underlying mtDNA heteroplasmy as a potential source of intra-person hiPSC variability. EMBO J. 2016;35(18):1979–90.

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Kang E, Wang X, Tippner-Hedges R, Ma H, Folmes CD, Gutierrez NM, et al. Age-related accumulation of somatic mitochondrial DNA mutations in adult-derived human iPSCs. Cell Stem Cell. 2016;18(5):625–36.

    CAS  PubMed  Google Scholar 

  137. Zambelli F, Mertens J, Dziedzicka D, Sterckx J, Markouli C, Keller A, et al. Random mutagenesis, clonal events, and embryonic or somatic origin determine the mtDNA variant type and load in human pluripotent stem cells. Stem Cell Reports. 2018;11(1):102–14.

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Deuse T, Hu X, Agbor-Enoh S, Koch M, Spitzer MH, Gravina A, et al. De novo mutations in mitochondrial DNA of iPSCs produce immunogenic neoepitopes in mice and humans. Nat Biotechnol. 2019;37(10):1137–44.

    CAS  PubMed  Google Scholar 

  139. Deuse T, Hu X, Gravina A, Wang D, Tediashvili G, De C, Thayer WO, et al. Hypoimmunogenic derivatives of induced pluripotent stem cells evade immune rejection in fully immunocompetent allogeneic recipients. Nat Biotechnol. 2019;37(3):252–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Falabella M, Minczuk M, Hanna MG, Viscomi C, Pitceathly RDS. Gene therapy for primary mitochondrial diseases: experimental advances and clinical challenges. Nat Rev Neurol. 2022;18(11):689–98.

    CAS  PubMed  Google Scholar 

  141. Mok BY, de Moraes MH, Zeng J, Bosch DE, Kotrys AV, Raguram A, et al. A bacterial cytidine deaminase toxin enables CRISPR-free mitochondrial base editing. Nature. 2020;583(7817):631–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Cho SI, Lee S, Mok YG, Lim K, Lee J, Lee JM, et al. Targeted A-to-G base editing in human mitochondrial DNA with programmable deaminases. Cell. 2022;185(10):1764-76.e12.

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to Ms. Jiji Sun for technical support.

Funding

This work was supported by the grants 2021YFC2700900 from National Key Research and Development Program of China to M.X.G and 31970557 and 81330024 from the National Natural Science Foundation of China to M.X.G.

Author information

Authors and Affiliations

Authors

Contributions

CC prepared the initial draft of the manuscript. MXG made the final version of the manuscript. All authors reviewed the manuscript.

Corresponding author

Correspondence to Guan Min-Xin.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

All authors have no proprietary or commercial interest in any of materials discussed in this article.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, C., Guan, MX. Induced pluripotent stem cells: ex vivo models for human diseases due to mitochondrial DNA mutations. J Biomed Sci 30, 82 (2023). https://doi.org/10.1186/s12929-023-00967-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12929-023-00967-7

Keywords