Across eukaryotes, most genes required for mitochondrial function have been transferred to, or otherwise acquired by, the nucleus. Encoding genes in the nucleus has many advantages. So why do mitochondria retain any genes at all? Why does the set of mtDNA genes vary so much across different species? And how do species maintain functionality in the mtDNA genes they do retain? In this review, we will discuss some possible answers to these questions, attempting a broad perspective across eukaryotes. We hope to cover some interesting features which may be less familiar from the perspective of particular species, including the ubiquity of recombination outside bilaterian animals, encrypted chainmail-like mtDNA, single genes split over multiple mtDNA chromosomes, triparental inheritance, gene transfer by grafting, gain of mtDNA recombination factors, social networks of mitochondria, and the role of mtDNA dysfunction in feeding the world. We will discuss a unifying picture where organismal ecology and gene-specific features together influence whether organism X retains mtDNA gene Y, and where ecology and development together determine which strategies, importantly including recombination, are used to maintain the mtDNA genes that are retained.

Mitochondria in most eukaryotes contain mitochondrial DNA (mtDNA). MtDNA encodes a subset of genes required for mitochondrial functionality. The particular set of encoded genes, the genetic organisation, and the physical structure of mtDNA vary dramatically across eukaryotes (Figure 1) [1,2]. MtDNA is inherited via diverse mechanisms across species, few of which resemble the inheritance of nuclear DNA [3–5]. Furthermore, the cellular ploidy and arrangement of mtDNA vary not just across species, but between cells and tissues and over development and time within individuals [6,7]. Table 1, in the spirit of the comprehensive graphical summary in [2], illustrates some of this diversity.

Genetic diversity in mtDNA.

Figure 1.
Genetic diversity in mtDNA.

(A) Tiles show the number of samples in NCBI's Organelle Genome database with a given mtDNA length and gene count (darker colours denote more samples). Particular species of interest are labelled Xy, where X is the first letter of their genus and y the first letter of their species, with full names given in the box (for example, Hs is Homo sapiens). (B) Unique protein-coding mtDNA profiles, ordered by gene count, found in the NCBI Organelle Genome database. Each row is a unique profile (which may be observed in many individual species), each column is a gene, and dark pixels denote gene presence. Example profiles corresponding to completely random, random reductive, or completely stereotypical mtDNA evolution are shown on the right. The inset is a schematic of this article: retaining more or fewer genes may trade off local organelle control with genetic robustness, and species must maintain the genes they do retain against mutational hazard. Code to reproduce these figures is freely available at https://github.com/StochasticBiology/mt-gene-stats.

Figure 1.
Genetic diversity in mtDNA.

(A) Tiles show the number of samples in NCBI's Organelle Genome database with a given mtDNA length and gene count (darker colours denote more samples). Particular species of interest are labelled Xy, where X is the first letter of their genus and y the first letter of their species, with full names given in the box (for example, Hs is Homo sapiens). (B) Unique protein-coding mtDNA profiles, ordered by gene count, found in the NCBI Organelle Genome database. Each row is a unique profile (which may be observed in many individual species), each column is a gene, and dark pixels denote gene presence. Example profiles corresponding to completely random, random reductive, or completely stereotypical mtDNA evolution are shown on the right. The inset is a schematic of this article: retaining more or fewer genes may trade off local organelle control with genetic robustness, and species must maintain the genes they do retain against mutational hazard. Code to reproduce these figures is freely available at https://github.com/StochasticBiology/mt-gene-stats.

Close modal
Table 1.
Physical and structural diversity in mtDNA
FeatureExample valuesNotes
Presence/absence Simply absent from, for example, Encephalitozoan cuniculi and Giardia, Entamoeba, and Trichomonas (unicellular parasites)  
Structure Linear, branched, circular, multichromosomal  
Copies per cell Presumably >106 in Xenopus oocytes, as 107 mitochondria present
Single nucleoid in many Apicomplexans (unicellular parasites) 
[8
Inheritance Uniparental (maternal or paternal), biparental, doubly uniparental, uniparental with leakage, ‘triparental’ (from neither nuclear parent)  
Mutation rate 0.13 dS/mya Pelargonium exstipulatum; 2.53 × 10−5 dS/mya Ceratozamia hildae (flowering plants) These values are only from plants, as cross-taxa comparisons can be complicated [9
Protein-coding gene count ∼67 protein-coding genes (100 genes in total) Andalucia godoyi (jakobid protist)
Two protein-coding genes Chromella velia (coral endosymbiont) 
 
Length 11.3 Mb Silene conica (flowering plant)
6 kb Plasmodium falciparum (unicellular parasite) 
 
Chromosome count Single in many metazoans
Hundreds in Amoebidium parasiticum (unicellular parasite) 
 
Different genetic codes Vertebrate, yeast, protozoan, invertebrate, echinoderm, ascidian, alternative flatworm, chlorophycean, trematode, Scenedesmus obliquus, Thraustochytrium, Rhabdopleuridae See (https://www.ncbi.nlm.nih.gov/Taxonomy/Utils/wprintgc.cgi
Beyond above classification Trypanosoma brucei mtDNA is partitioned into interlocking, chainmail-like ‘mini’ and ‘maxi’ circles; minicircles encode guide RNA to ‘decrypt’ the content of the maxicircles  
FeatureExample valuesNotes
Presence/absence Simply absent from, for example, Encephalitozoan cuniculi and Giardia, Entamoeba, and Trichomonas (unicellular parasites)  
Structure Linear, branched, circular, multichromosomal  
Copies per cell Presumably >106 in Xenopus oocytes, as 107 mitochondria present
Single nucleoid in many Apicomplexans (unicellular parasites) 
[8
Inheritance Uniparental (maternal or paternal), biparental, doubly uniparental, uniparental with leakage, ‘triparental’ (from neither nuclear parent)  
Mutation rate 0.13 dS/mya Pelargonium exstipulatum; 2.53 × 10−5 dS/mya Ceratozamia hildae (flowering plants) These values are only from plants, as cross-taxa comparisons can be complicated [9
Protein-coding gene count ∼67 protein-coding genes (100 genes in total) Andalucia godoyi (jakobid protist)
Two protein-coding genes Chromella velia (coral endosymbiont) 
 
Length 11.3 Mb Silene conica (flowering plant)
6 kb Plasmodium falciparum (unicellular parasite) 
 
Chromosome count Single in many metazoans
Hundreds in Amoebidium parasiticum (unicellular parasite) 
 
Different genetic codes Vertebrate, yeast, protozoan, invertebrate, echinoderm, ascidian, alternative flatworm, chlorophycean, trematode, Scenedesmus obliquus, Thraustochytrium, Rhabdopleuridae See (https://www.ncbi.nlm.nih.gov/Taxonomy/Utils/wprintgc.cgi
Beyond above classification Trypanosoma brucei mtDNA is partitioned into interlocking, chainmail-like ‘mini’ and ‘maxi’ circles; minicircles encode guide RNA to ‘decrypt’ the content of the maxicircles  

A summary of several aspects of mtDNA diversity from the references in this article, particularly inspired by Smith and Keeling [2] but with other data sources cited throughout this article.

MtDNA has downsides as a site for information storage. Replicating frequently, with a low effective population size, in an environment surrounded by potential mutagens, and with less packaging than nuclear DNA, the risk of mutational damage is high [10–14]. In some organisms (including most animals) mtDNA recombination is limited, raising the possibility of genome erosion via Muller's ratchet — the ongoing buildup of deleterious mutations until function is lost [15,16]. Maintaining high-ploidy mtDNA is likely costly [17] and raises possible conflicts between nuclear- and mtDNA-encoded genes [18].

Given these challenges, an obvious question is — why do organisms encode any genes at all in mtDNA? And the necessary corollary to any answer — how do organisms maintain the function of their encoded mtDNA genes? This review will attempt to describe some of the diversity of mtDNA behaviour through the lens of these questions (Figure 1B inset), attempting to provide a plausible and general set of principles that shape mtDNA evolution and maintenance across eukaryotes.

We must first consider the history of mitochondria. It is generally accepted that they were originally independent organisms — the closest known modern approximation to the ‘proto-mitochondrion’ is an α-proteobacterium [1,19–21]. Through an endosymbiotic event, the proto-mitochondrion was absorbed by a host — thought to be similar to an Asgard archaeon [1,22–24] — beginning the symbiosis that would give rise to modern eukaryotes [25–29]. An excellent overview of the subsequent changes in metabolic, regulatory, and import profiles is given in [1]; we will focus on the genome. Studies have attempted to reconstruct the properties of the proto-mitochondrion [30–33], with some work suggesting that it was originally an energy parasite [34]. The consistent picture is that it originally possessed the full complement of genes that a free-living organism would require.

Following endosymbiosis, redundancy with the host genome led to rapid loss of many of these genes [35,36]. Other genes were transferred to the host cell nucleus [19,37–39]. Several advantages have been proposed for nuclear encoding of mitochondrial machinery [40], with several focussing on the mutational hazard experienced by genes encoded in mtDNA [12,41] — which will be discussed in the contexts of different taxa throughout this article. These advantages include avoidance of Muller's ratchet (the inevitable buildup of deleterious mutations) [15,42,43], protection from damaging chemicals [10], enhanced capacity to fix beneficial mutations [40,42], and an energetic advantage over maintaining multiple mtDNA copies [17]. The physical transfer of mtDNA to the nucleus (giving rise to so-called nuclear mitochondrial sequences or NUMTs) is not a rare event [44,45], occurring over generational timescales in humans [46] and readily in plants [47]. However, the transfer of mtDNA is not the same as the transfer of functional gene content, as differences in genetic code (Table 1), regulation, and more must be addressed for functionalisation of transferred content. Several specific mechanisms for transfer have been discussed in detail [37,48,49], with increased recent focus on the properties of the intermediate state where a gene is contained in both nuclear and mtDNA [50,51].

These losses reduced the gene content of mtDNA dramatically, so that the most gene-rich mtDNAs discovered in modern eukaryotes have only dozens of genes, with the highest protein-coding gene counts so far found in jakobid protists Andalucia godoyi and Reclinomonas americana [52,53]. Overwhelmingly, the collection of genes found in modern eukaryotes are a subset of those in these gene-rich protists (Figure 1B) [38,54,55]. Reconstruction suggests that the last common ancestor of modern eukaryotes had a gene complement slightly larger than these jakobids [55]. Rare examples of mtDNA containing genes not found in these protists do exist. For example, octocoral mtDNA has acquired the msh1 gene [56,57] — which we will meet again later — likely via virus-mediated horizontal gene transfer [58], and a restriction modification system has been acquired by the mitochondrion of a marine protist [59].

The physical structure of the mtDNA housing these genes is highly variable [2,60]. Many animal mtDNAs have a familiar circular structure, although mtDNA may form networks in human hearts [61], and mtDNA fragmentation is observed in lice [62] and cnidarians [63]. In contrast, plant and algal mitochondrial genomes are often split between many (often dozens of) different ‘subgenomic’ mtDNA molecules, each containing a subset of the full genome [64] and which may be linear or branched [65]. Linear mtDNA, including telomeres, is found across kingdoms [66,67]. Protist mtDNA structure exhibits substantial diversity [68], including branching and linear molecules, deviations from usual genetic codes [69], multiple chromosomes (sometimes with a single gene split across multiple mtDNA molecules and subsequently spliced together [70]), and the unusual ‘kinetoplast’ situation found in trypanosomes. Here, small ‘mini’ and large ‘maxi’ circles exist linked together in a ‘chainmail’ structure, with the minicircles encoding a guide RNA required to decode the mtDNA genome in the maxicircles [71].

Different eukaryotic kingdoms differ in both average number of mtDNA genes and the spread of gene count across different species (Figure 1B, Table 1, [38]). Focussing on the set of genes and not their ordering or arrangement (which does vary across species), animal mtDNA gene content is quite constant, with 13 protein-coding genes found across most animals. Exceptions to this complement include the aforementioned gain of msh1 in corals [57] and some instances of loss in taxa including nematodes [72]. The gene content of many fungi often similar, and in many cases quite constant [50], although rearrangements and structural complexity can be dramatic (cox1 in Agaricus bisporus contains 19 introns [73]). Plant mtDNA is generally more gene-rich and much more variable, with dozens of protein-coding genes and, often, substantial non-coding regions, which can range from 1% to >99% of the genome [74,75]. Across kingdoms, parasitism is often associated with reduced gene content [76]; in an extreme example, a cnidarian parasite retaining mitochondria but lacking mtDNA has been reported [77].

Among protists, gene profiles vary dramatically across different taxa [68]. Some unicellular parasites, with anaerobic lifestyles, have completely lost mtDNA [78–83]. Mitochondria that have undergone this — or even greater — reductive evolution are often referred to as mitochondrion-related organelles (MROs) including mitosomes and hydrogenosomes, depending on their particular metabolic properties. An anaerobic eukaryote without any organelle related to a mitochondrion has been reported [84]; reports of a dinoflagellate retaining aerobic mitochondria but lacking mtDNA [85] remain debated [86]. Other unicellular parasites, including many Apicomplexans, retain only three protein-coding genes cox1, cox3, cob; the related coral endosymbiont Chromera velia has additionally lost cob to retain only two protein-coding genes. On the other hand, the (also unicellular) jakobids above have the highest known mtDNA gene counts [52]. Different algae have markedly different profiles, with, for example, several dozen protein-coding genes retained by many red algae and some green algae retaining very few [87].

While not completely stereotypical, the genes retained across eukaryotic mtDNA are far from random [38,54] (Figure 1B). Several protein-coding genes, including cox1, cox3, cob, are retained in almost all species. Several specific nad and atp genes are also highly retained, while various rps and rpl genes are retained in a more limited and variable range of species. sdh genes, and a collection of others not encoding ETC subunits or ribosomal proteins, are retained by substantially fewer species [38,50,88]. Ribosomal RNA genes are consistently conserved (although often fragmented if ribosomal protein-coding genes are transferred from the organelle) [50]. Profiles of retained tRNA genes vary more substantially across taxa. A broad review is given in [89], which highlights some particular points of diversity. While many metazoans contain a complete, minimal set of tRNAs, other taxa vary substantially. Plant tRNA profiles are highly variable and rapidly evolving [90]; fungal profiles are also highly variable, with closely related species containing dramatically different sets. Trypanosome and alveolate mtDNA may completely lack tRNA genes, or contain a dramatically reduced set.

These observations turn our original question into two subquestions. First, what determines which genes are preferentially retained across species? And second, why does a particular species retain a given number of genes?

The question of why a given gene is more or less likely to be retained in mtDNA has been discussed for decades. We focus here on mtDNA diversity across extant species; many endosymbiont genes were lost early post-endosymbiosis (presumably due in part to redundancy with the host) [1,35], and are not present in modern eukaryotes. One classic hypothesis for protein-coding genes relates to the hydrophobicity of a gene product [91,92]. It was first hypothesised that hydrophobic products, produced outside the mitochondrion, would be hard to import through the mitochondrial membrane to their required position. More recent research has suggested that hydrophobic products may be prone to mistargeting to the endoplasmic reticulum [91].

Another classic hypothesis is ‘colocation for redox regulation’ or CoRR [93,94]. Here, retaining genes local to the mitochondrion allows the individual organelle a tighter degree of local control over its redox function. This tighter control potentially allows faster, and more efficient, responses to new challenges — a change in bioenergetic demand or the degradation of key proteins, for example. Nuclear encoding makes it harder to fulfil the specific requirements of a given mitochondrion, out of the hundreds in the cell [94].

Other hypotheses have also been proposed. The economics — in the sense of the ATP budget for expression and maintenance — of organelle encoding has been argued to favour retention under some conditions [17]. It has been suggested that organelle genes can act as redox sensors, reporting the bioenergetic performance of a cell over time and facilitating control [95]. Issues with nuclear transfer and expression, including potential cytosolic toxicity of products [96] and differences in genetic code [40,97] have also been proposed to explain retention.

In an attempt to examine support for these hypotheses from an unbiased perspective, our group has used large-scale organelle genome data (thousands of eukaryotic mtDNA sequences and dozens of full nuclear genomes) with structural data and Bayesian model selection to identify likely features predicting the retention profile of a given gene [38,54]. We found that a combination of the hydrophobicity of a gene product and the GC content of the gene itself (independently of the general low GC bias in mtDNA [98,99]) robustly predicted (in unseen data) both whether a given gene would be retained in mtDNA or transferred to the nucleus, along with a signal associated with the pKa of the gene product.1

1The GC content of the mtDNA genome as a whole will certainly have changed over evolutionary time, and these studies accounted for the diversity of GC content baselines across different taxa. But it is the between-gene differences in GC content that was found — consistently across taxa — to predict between-gene differences in retention in this work.

We also found that the ‘energetic centrality’ of a gene product — how physically central its position is in its containing complex — predicted mtDNA retention. Although correlations exist between these gene properties, their appearance together in the Bayesian model selection framework we used suggests that each provides independent power to predict retention. In contrast, features including molecular mass, energetic requirements for assembly, genetic code discrepancies and GC skew (G vs C usage) were not found to have any notable statistical support by this method. Although such an inference-based approach can only support and compare hypotheses statistically rather than directly test them experimentally (and can only consider the hypotheses with which it is presented), models based on these features predicted success of synthetic nuclear-mtDNA gene transfer experiments [88] (reviewed in [50]) and across other endosymbionts and organelles [100].

Why these features? The signal associated with hydrophobicity agrees with the hypothesis that difficulty in importing hydrophobic products — due to physical barriers and/or mistargeting — is a shaping factor. The energetic centrality of a product can intuitively — and explicitly [101,102] — be connected to its centrality in the assembly pathway of the complex. The control of complex assembly (in response to bioenergetic demand) in turn is a key determinant of redox regulation and therefore to CoRR [94].

GC content corresponds less readily to an established hypothesis. Following [103], we speculated that GC richness confers thermodynamic stability to a gene and therefore makes it more robust to the challenging environment of the mitochondrion. At a similarly speculative level, we proposed that ‘the synthesis of protein products enriched for higher-pKa amino acids may involve lower kinetic hurdles in the more alkaline pH of mitochondria…. favoring the retention of the corresponding genes’ [38]. Investigation of these hypotheses at a molecular level will be required to strengthen these arguments.

Our dual question was why a given species is more or less likely to retain mtDNA genes. For example, parasitic species are expected to atrophy their mtDNA (and their mitochondria) both due to their reduced requirements for intrinsic energy transduction and due to their often low-oxygen environments [39,79,104–106]. Self-pollinating plants often transfer more genes to the nucleus than other plants; selfing has been shown theoretically to accelerate the transfer process when it confers an advantage [107,108]. More general theory across taxa has also been proposed. The ‘mutational hazard hypothesis’ proposes that mtDNA gene retention is safer in taxa with lower mtDNA mutation rates (for example, plants) [12,41]. A recent ‘burst-upon-drift’ model has been proposed to jointly explain variability in retention profiles and how nuclear transfer becomes fixed [50].

We recently hypothesised that the CoRR argument could connect species-specific demands on redox regulation to retention profiles more generally [109]. We considered a cellular model for the expression and degradation of organelle-targeted gene products, expressed either from oDNA (where high mutation rate poses a challenge) or the nucleus (where mutation is lower). We assessed the possible ‘supply’ of these products in the face of a ‘demand’ for organelle machinery imposed by the environment, which could be low and stable or high and highly varying. We found that in environments imposing a high and variable demand, the advantage of rapid supply from oDNA encoding outweighed the disadvantage of mutational hazard; the opposite was true in stable, facile environments. This theory predicts semi-quantitatively that more oDNA encoding is advantageous in organisms subject to strong, variable environmental demands, while nuclear transfer is advantageous in stable, less demanding environments.

This is supported by a cross-taxa phylogenetic comparative investigation of mtDNA gene count and ecology [76]. Here, attempting to account for the difficulty of comparisons across the broad, sparse, uncertain datasets available, we found fewer genes retained in organelles exposed to limited demands (endoparasites, and plastids without photosynthetic demands) and more genes in those exposed to more varying environments (in sessile organisms, deserts, and tropical oceans).

It could never be claimed that these ideas give a complete answer to our first question. Indeed, it would be astonishing if a single, concise principle could explain all the diverse behaviour observed over billions of years of eukaryotic evolution. But the statistical treatments and connections to large-scale data above suggest that the proposed mechanisms do have some (not complete) explanatory power across a broad range of organisms. More genes are retained in mtDNA if species require tight local control of their redox machinery; properties of a gene including its product's hydrophobicity and centrality increase its propensity to be retained (Figure 1B inset). Overall, there would seem to be advantages to retaining genes in mtDNA in many cases. So…

Mutational hazard

It is worth beginning by expanding on some issues associated with encoding information in mtDNA. MtDNA is less packaged and protected than nuclear DNA, frequently replicates, and its physical environment contains mutagens including the reactive oxygen species resulting from mitochondrial activity [10]. The contributions of these features to the accumulation of mtDNA damage is debated [110], with some evidence that oxidative damage may not be the dominant source of mutation [111]. Oxidative damage may be more like to induce strand breaks and abasic sites, and the specific behaviour of the polymerase gamma that replicates and proofreads mtDNA (including avoiding misreading damaged bases) also shapes mutational profiles [112–114]. Across these specific mechanisms, mutational hazard is clearly an issue [11–13], and can be directly demonstrated [115]. The limited number of genomes per cell limits the effective population size, potentially amplifying the effects of Muller's ratchet [14] and imposing a ‘drift barrier’ to the maintenance of efficient repair machinery [116]. [50] highlight that mutation rate does not provide a direct selective advantage for gene transfer at the level of the organism; however, it can readily be demonstrated that transfer is nonetheless evolutionarily favoured in populations (Supplementary Information).

Observed mtDNA mutation rates vary dramatically across taxa [9,12], between males and females [117,118], and between genes [119] — although such rates are a combination of a basal damage process and repair capacity, which also vary dramatically. In many animals, mtDNA mutation rates are well known to be higher than nuclear mutation rates. However, in plants [120], fungi [12], and indeed some animals (corals and sponges) [121,122], mtDNA mutation rates may in fact be lower than those in the nucleus. In plants [123], and more speculatively in these other taxa, mtDNA recombination-mediated repair will allow the correction of mutations [124–126], albeit at the cost of structural rearrangements of the genome [120,127] constituting an important mode of evolution [128].

The consequences of this mutational pressure on mtDNA are not homogeneous. Biochemical asymmetry (favouring hydrolytic deamination of cytosine) has the effect of favouring C → T conversion in mtDNA [98,99]. The GC content of mtDNA influences the free energy of the DNA duplex, suggested to influence mutational susceptibility of mtDNA [103].

MtDNA mutations can be highly detrimental. Cells typically contain large (highly polyploid) populations of mtDNA molecules (Figure 2). The state where all these molecules have the same haplotype is termed ‘homoplasmic’; the converse, where at least two types exist, is ‘heteroplasmic’ [129–132]. Heteroplasmy, albeit on a small scale, is ubiquitous across many cell types and species [133–135]. In the case of two mtDNA types, the proportion of one (usually mutant) type is often referred to as the ‘heteroplasmy’ h of a sample, which could be a single cell, a tissue, or an organism2

2This terminology can be misleading, as if a mutant allele proportion exceeds 50% then heteroplasmy should arguably be redefined with respect to it as the major allele, but we will keep it for consistency with the literature.

(Figure 2B). A nonlinear threshold effect is often observed, where a cell can support a heteroplasmic fraction of a dysfunctional mutant, but if this mutant frequency is too high then the cell experiences negative consequences [136]. This threshold allows mtDNA mutations to persist in populations, occasionally manifesting at high enough levels to cause disease [132].

MtDNA-intrinsic processes shaping heteroplasmic mtDNA populations within cells.

Figure 2.
MtDNA-intrinsic processes shaping heteroplasmic mtDNA populations within cells.

(A) Coarse-grained schematic of some processes that influence mtDNA populations, (i) independent of and (ii) dependent on recombination. Dark and light circles denote a general “heteroplasmic” picture of different mtDNA types; the star denotes molecular damage. (iii) Illustrates how recombination between regions of the same mtDNA molecule can lead to genome fragmentation and stoichiometric complexity. (B) Evolution of heteroplasmic populations viewed as selection and segregation processes. Selection shifts mean heteroplasmy, favouring one mtDNA type over another (due to type-specific differences between rates in (A)). Segregation increases (cell-to-cell) heteroplasmy variance without shifting the mean.

Figure 2.
MtDNA-intrinsic processes shaping heteroplasmic mtDNA populations within cells.

(A) Coarse-grained schematic of some processes that influence mtDNA populations, (i) independent of and (ii) dependent on recombination. Dark and light circles denote a general “heteroplasmic” picture of different mtDNA types; the star denotes molecular damage. (iii) Illustrates how recombination between regions of the same mtDNA molecule can lead to genome fragmentation and stoichiometric complexity. (B) Evolution of heteroplasmic populations viewed as selection and segregation processes. Selection shifts mean heteroplasmy, favouring one mtDNA type over another (due to type-specific differences between rates in (A)). Segregation increases (cell-to-cell) heteroplasmy variance without shifting the mean.

Close modal

As well as driving mitochondrial evolution across eukaryotes, mtDNA mutations have important translational consequences. Devastating human diseases arise when deleterious mtDNA mutations are inherited at high heteroplasmy [131,132] and understanding the organism-scale evolution of mtDNA is important in clinical approaches to address these diseases [137]. In plants, dysfunction due to mtDNA variants, while damaging for the organisms, can counterintuitively have very positive consequences for humans. ‘Cytoplasmic male sterility’ (CMS), arising from mtDNA or mitonuclear properties (see below), allows the easy production of hybrid crops, which often have substantially higher yields than inbred lines [138–140]. Although hard to precisely quantify, CMS is involved in a substantial proportion, or majority, of the global production of many tabletop crop species [140,141]. In this sense, ‘pathologies’ arising from plant mtDNA issues genuinely help feed the world.

Intracellular competition and incompatibility between mtDNAs

An important parallel issue is the potential for competition between different mtDNA types within the same cell. There is some evidence that mtDNA heteroplasmy in and of itself is detrimental, even when no mtDNA types involved are deleterious [142–144].

Cell-to-cell distributions of heteroplasmy change over time in response to selection and segregation. Selection shifts the mean heteroplasmy over time; segregation increases the width of the cell-to-cell distribution (Figure 2B). Under various assumptions, the distribution of heteroplasmy has been shown [145] to correspond to population genetic solution in the absence [146] and presence [147] of selection. However, using this connection as suggested [145,148] to estimate selection and segregation rates from mtDNA measurements has several issues which recent statistical work has addressed [149]. Many other theoretical approaches have been used to explore the quantitative behaviour of heteroplasmy [150] including implementations of the Moran model [151] and Wright's models [152], classical models considering more specifics details of organelle genomes [153–156], and more detailed models including the roles of spatial structure and the microscopic processes involved [157–165].

Connected literature discusses selective differences between mtDNA types at this level as ‘segregation bias’ or ‘selfish proliferation’. Different mtDNA sequences may, for example, have different propensities for replication. A ‘replication–transcription switch’ has been proposed where favouring one process disfavours the other [166]. They may have different functional consequences for their host organelles and cells, so that selective pressures at those levels act to remove less functional types. A common picture is that an mtDNA type experiencing a replicative advantage is detrimental to cell, tissue, or organismal fitness. The different scales of selection in such cases can lead to proliferation (by replication) or removal (by removal of cells) of the selfish type [167–170]. Counterintuitively, physical properties of the system can lead to the proliferation of even deleterious mutations [159].

Mitonuclear incompatibility

Another issue arising from the cellular context of mtDNA variation is mitonuclear incompatibility [18,171]. Because mitochondria require products encoded both by the nucleus and the mtDNA, it is possible for negative effects to arise from a combination of the nuclear and mtDNA alleles. A striking recent example is a lethal incompatibility affecting Complex I in naturally occurring hybrids [172]. Such interactions may drive speciation [173–175] and have been implicated in ageing [176], the evolution of sex [177,178], and shaping environment–gene and gene–gene interactions [179].

In cases where mtDNA is inherited maternally, the ‘mother's curse’ effect can lead to the accumulation of mutations which are damaging to males but are neutral or beneficial for females [180]. Presumably, if mtDNA is inherited strictly paternally, the comparable accumulation of mutations damaging to females but neutral or beneficial to males may occur — akin to the ‘father's curse’ picture [181]. Mitonuclear interactions are a mechanism by which these curses can be resolved [182]. As mitochondrial functionality relies on cooperation between nuclear- and mtDNA-encoded mitochondrial genes, the presence of a damaging mtDNA variant may induce strong selection for a nuclear allele that compensates this damaging effect [183]. Such ‘restorative’ nuclear variants are observed, for example, in CMS in plants (where male fertility, compromised by mtDNA variants, is restored by a nuclear factor) [184].

Different cellular processes at the molecular, organelle, cellular, and organismal levels influence mtDNA evolution. Figure 2 gives a coarse-grained picture of some of the processes that shape cellular populations of mtDNA.

Intracellular repair and removal

At the level of an individual mtDNA molecule, damage-repair mechanisms can be used to correct lesions, for example via fixing double-strand breaks or templating corrections by gene conversion [9,124,125,128,185]. At the level of organelles, if an mtDNA mutation corresponds to an organelle phenotype that can be individually sensed, cellular machinery can attempt to preferentially remove the mutant within that single cell via ‘mitophagy’ [186,187]. This within-cell process is part of mitochondrial ‘quality control’ [188–190].

Intercellular removal

Between-cell selection can be used, removing whole cells if they contain an unacceptable proportion of the dysfunctional mutant. This scale of process is highly contingent on the broader context of a single cell. In a unicellular population, it simply corresponds to loss of less-fit individuals from the population. In a multicellular organism, it relies on the ability to remove cells, and is, therefore, more feasible in tissues with high rates of turnover than in quiescent tissues of static structure (for example, plant soma, animal brain, and muscle) [167,169].

In many organisms there is also a developmental axis to consider (Figure 3A). Depending on the germline structure of an organism, the timing and scale of selection can vary (for example, removing cells or embryos at different stages). For example, animal embryos containing (cells containing) a high mutant proportion may fail early developmental checkpoints and fail to develop further. The selection for mitochondrial quality, in the face of different mutational pressures, has been proposed to drive the evolution of a germline itself [192].

Segregation and developmental influences on mtDNA.

Figure 3.
Segregation and developmental influences on mtDNA.

(A) Illustration of mtDNA in the germline of (i) bilaterian animals (ii) plants. In (i), early developmental stages decrease mtDNA copy number per cell, subsampling the mtDNA population and imposing a physical ‘bottleneck’ that acts to accelerate drift due to other segregation processes. In (ii), a physical bottleneck is less pronounced or absent; segregation occurs due to other processes. (B) A mathematical model for segregation quantifies the heteroplasmy variance due to different processes [191]. All except gene conversion (arrowed) are amplified at low mtDNA copy number N; evidence suggests that animals employ turnover and partitioning (i, ii, iv–v) for segregation and plants make use of gene conversion (iii). Other pertinent parameters are fi (fragmented mitochondrial proportion, linking physical and genetic behaviour) and νi (mitophagy rate); a full description can be found in the original paper.

Figure 3.
Segregation and developmental influences on mtDNA.

(A) Illustration of mtDNA in the germline of (i) bilaterian animals (ii) plants. In (i), early developmental stages decrease mtDNA copy number per cell, subsampling the mtDNA population and imposing a physical ‘bottleneck’ that acts to accelerate drift due to other segregation processes. In (ii), a physical bottleneck is less pronounced or absent; segregation occurs due to other processes. (B) A mathematical model for segregation quantifies the heteroplasmy variance due to different processes [191]. All except gene conversion (arrowed) are amplified at low mtDNA copy number N; evidence suggests that animals employ turnover and partitioning (i, ii, iv–v) for segregation and plants make use of gene conversion (iii). Other pertinent parameters are fi (fragmented mitochondrial proportion, linking physical and genetic behaviour) and νi (mitophagy rate); a full description can be found in the original paper.

Close modal

It is worth taking a second to disambiguate the various meanings that ‘selection’ can have in this context. Given the centrality of mtDNA to bioenergetics and eukaryotic life, it is almost self-evident that some mutations will be selected against (negative selection). Pathogenic human mtDNA mutations [132] and sterility-causing mutations in plants [193] are intuitive examples. However, it can be hard to identify at what level selection acts — intracellular, intercellular, and/or at the level of the organism itself [169,194,195]. In bulk samples, distinguishing intracellular and intercellular selection is challenging, as the same bulk dynamics would be observed for either level of selection. The dominant level may depend on circumstance: work in mice has suggested organelle-level selection [196], while recent single-cell work has found more support for intercellular selection in some circumstances [194]. Powerful theoretical work has demonstrated the capacity of selection at these different levels to maintain mtDNA through germline development [197]. Another subtle (and debated) question is the extent to which positive selection has shaped natural mtDNA populations. Can mtDNA diversity be explained by non-adaptive processes, including neutral ratchets [198], or must selection be invoked?

Segregation

Any selection on or above the between-cell scale relies on there being diversity in heteroplasmy between cells. This ‘heteroplasmy variance’ (often written V(h)) is what intercellular or organismal selection can act upon to purify a population. The generation of V(h) is often referred to as ‘segregation’ or (particularly in the plant kingdom) ‘sorting out’. It can be achieved through various mechanisms (Figure 3) [191], and can occur in parallel with selection acting to change the heteroplasmy mean (Figure 2B) [160,197]. These include several process in Figure 2, including the random replication and degradation of mtDNA [157,161,199,200], the replication of a random subset of mtDNA molecules in a cell [201], random partitioning of mtDNA molecules at cell divisions [161,202–205], and gene conversion [191,206,207]. MtDNA sequence features partly determine segregation behaviour [208,209]. The physical distribution of mtDNA molecules in the mitochondrial population, which may be reticulated, fragmented, or a combination, shapes the segregation contribution of each of these processes [157,191,204,210] — the physical behaviour of mitochondria shapes the genetic segregation of mtDNA.

Segregation of deleterious mutations allows selection to remove entities (for example, individual cells, embryos, or organisms) in which a relatively high mutant load has been concentrated, leaving the remaining entities with lower mutant loads. This process can mitigate against Muller's ratchet because it allows descendant entities to inherit lower mutant loads than their ancestor. For example, average heteroplasmy amongst (surviving) offspring can be lower than in their mother — because high-heteroplasmy offspring did not survive. But segregation can also facilitate adaptation of beneficial mutations [211]. This is because fixing a new mtDNA type necessarily involves a heteroplasmic intermediate state (before all mitochondria in a cell harbour the new mitotype), and heteroplasmy can be detrimental even if neither mitotype is deleterious [142–144].

Inheritance and exchange

The inheritance patterns of mtDNA in a given species contribute to its ability to maintain function and reduce genomic conflicts [5,181,212]. Strictly maternal inheritance avoids generating heteroplasmy by mixing parental mtDNA contributions, and hence limits the negative consequences of mixed mtDNA [142–144,213]. But in some circumstances an alternative may be desirable. If some paternal contribution is allowed, and recombination supported [4,5,214], heterozygosity can be maintained in a population and more rapid adaptation to changing environments may be supported [215]. Purely paternal inheritance, rarely observed (though more common in plastids [5]), has been suggested to support strong selection through a severe bottleneck [181,216].

Some species may support horizontal gene transfer of mtDNA on various scales, from the transfer of individual mitochondria (and hence mtDNA) between cells, to large-scale exchange of mtDNA content between individuals. Introgression — where mitochondrial content from another organism not involved in the nuclear reproductive process — has been naturally observed in algae [217], and is a key component of human therapies targeting the inheritance of mtDNA disease [137,218,219]. Grafting plants, an essential aspect of agriculture, can lead to introgression [47,220]. At the cellular level, transfer of mitochondria (and therefore mtDNA) between cells via tunnelling nanotubes has received substantial recent attention [221,222]. From a mathematical perspective, such cellular introgression can help stabilise evolving mtDNA populations [161,223] and has experimentally been found to rescue deleterious phenotypes [224,225].

Taken together, there are clearly a collection of different strategies that organisms can in principle employ to balance the priorities of maintaining existing mtDNA integrity and allowing adaptation to new conditions. We will now discuss how these possible strategies are employed by different eukaryotic species, and attempt to crystallise some principles underlying this diversity. Due to the vast amount of research on these topics, especially in vertebrates, we cannot hope to connect to every relevant study. Our goal is not (indeed, cannot be) to exhaustively survey all studied mtDNA behaviour, but rather to provide a combined general picture and specific examples of diversity across kingdoms. We hope to provide a summary picture and also (see Discussion) propose a mechanism whereby this summary can by expanded over time outside the confines of a single article.

Animals

MtDNA mutation rates vary across animals [226], with vertebrates often having mtDNA mutation rates 20× higher than nuclear rates, and other lineages (for example, corals) having very low rates [121]. Recombination in the mtDNA of many animals is usually thought to be limited, with evidence against rapid mtDNA recombination occurring in mice [227]. Evidence has been reported for recombination in mussels [228] and carp [229], and recent work in Drosophila has shown that recombination can repair double-strand breaks in mtDNA [230]. In human cell lines, mtDNA damage has been reported as being removed through degradation rather than repair mechanisms [231,232]. The existence of mitochondrial quality control through mitophagy in animals has been more established, and reviewed extensively (for example, [188,189]).

At the cellular level, animal mtDNA exhibits selection both in germline and somatic tissues. Favouring of one mtDNA type over another in somatic animal tissues over the lifespan of one organism has been observed over many model systems and many mtDNA pairings [169]. Mouse lines constructed to be heteroplasmic have been a common study model here [233], and all mouse tissue-specific patterns of selective advantage and disadvantage observed to date can be grouped on an overall ‘atlas’ of tissue profiles [169]. That work proposed an overarching explanation in terms of different degrees of ‘selfishness’ — different propensities to replicate rather than transcribe useful machinery — across different sequences. Selfish replication of mtDNA has also been the focus of study in other animal systems including nematodes [234] and flies [168] (as well as in other kingdoms, described in later sections). Different mtDNA haplotypes have been shown to have different respiratory behaviours in mice [235] and humans [236]. Nuclear factors shaping heteroplasmy in different mouse tissues have been reported [196,237,238] along with a role for mitochondrial fission–fusion balance [239]. Bodies of work have also explored the multi-level selection shaping mtDNA populations in, for example, nematodes [167,240]. In humans, tissue-specific selection is also observed [241], including for disease-causing variants [242]. Nuclear factors shaping such heteroplasmy evolution have been identified [243,244]. Many open questions remain, however, including the reasons why pathogenic variants experience clear negative selection in some tissues (for example, blood for the 3243 mutation in humans [242]) and not others, and the molecular mechanisms of mtDNA selection [245,246] remain incompletely understood.

Germline selection for mtDNA in animals has also been demonstrated, including in mice [247–250], flies [251,252], and humans [253]. Several mechanisms have been identified, involving nuclear factors [143] and mitophagy with mitochondrial fragmentation [251,252]. Selection through germline development, particularly at the intracellular level, can help purify mtDNA and avoid mutations proliferating during the establishment of the high-ploidy oocytes found in mammals [197]. Correspondingly, population-level evidence for mtDNA selection has been observed in humans [254,255]. Selective pressures acting at this broader scale have been proposed to involve gene expression profiles [256], transcriptional pressures shaping gene ordering [257] and environmental cues, for example, of temperature and altitude in humans [254,255,258], altitude in birds [259], and temperature and metabolism in fish [260,261].

Many animals exploit a developmental mechanism variously called the ‘germline bottleneck’ or ‘mitochondrial bottleneck’ to segregate mtDNA [130,262,263]. This mechanism typically couples a developmental reduction in mtDNA copy number per cell with random processes that segregate heteroplasmy between cells (Figure 3) [150,160]. In such animals, mtDNA copy number in oocytes is often high (for example, ∼2 × 105 in mice [201,202,264]). During the first several cell divisions after fertilisation, this copy number per cell plummets to perhaps hundreds or thousands (the exact number is debated [202]) before being reamplified in the germ cells of the next generation. In parallel, random replication [200,201] and partitioning [202,203] generates cell-to-cell variability in heteroplasmy between developing germ cells, and hence between offspring [160,247]. This process, with different rates and numbers, occurs across bilaterians [150,265] including insects [266,267], humans [131,262,268], fish [269], and cattle, where it was originally observed [270,271]. Ongoing random replication of mtDNA continues this segregation throughout lifetimes [247,272]. Segregation also occurs in somatic tissue over time [240,269,273,274].

Several animals do not sequester a germline in the same way as vertebrates, including soft corals and sponges. Elegant theory work has connected this to the particular mutation pressures faced by these taxa (suggested to be relatively high background mutation rates and lower copying error rates), with the converse (low background mutation, high copying error) suggested to favour an early sequestered germline [192]. The absence of extreme mtDNA ploidy in these taxa and their modular growth plans have also been theoretically connected to their mtDNA maintenance [197]. Some members of these taxa, as mentioned above, have unusually acquired msh1 in their mtDNA. Theory work has suggested that these two features may be connected, and that msh1-supported mtDNA recombination may assist segregation in the absence of a vertebrate-like germline bottleneck [191]. In some of these organisms, mitochondria are fragmented and highly motile, recalling structure and dynamics in plants (see next section) — for example, freshwater sponges [275].

MtDNA inheritance in animals is predominantly maternal. This is the case observed in humans; most claims against this rule [276] are controversial [277], and recent observations have indeed shown a lack of intact mtDNA in human sperm [278]. The extent of paternal leakage varies across animals; substantial leakage is observed, for example, in bees [279]. An exception to the maternal rule is the doubly uniparental inheritance observed in some bivalves [280–282]. The benefits and costs of the consequential paternal contribution to mtDNA in some individuals is the target of ongoing study [181,283].

Plants

Mutation rates in plant mtDNA, while typically lower than nuclear mutation rates [12], vary dramatically across species [284] and are in part predicted by (somatic) genome copy number [9], in a relationship suggested to be linked to the availability of templates for repair. Plant mtDNA readily recombines [125,285–287]. This supports both homologous recombination-mediated damage-repair mechanisms [125,286,288–290] and gene conversion for templated repair [185] and segregation [207,291,292]. The relative plasticity of plant mtDNA has led to it being (rather unkindly) dubbed ‘the dumping ground’; a large amount of non-coding content, including material derived from the nucleus, plastid, and viral genomes is found in plant mtDNA [193,293,294]. The specific connection between recombination-driven mtDNA repair and genome evolution has been highlighted in [128,289,295].

As a consequence of this plasticity, the physical structure of plant mtDNA is both more complex and more variable than in animals [287,296,297]. The mtDNA genome is often spread over a collection of subgenomic mtDNA molecules [298, 299], and individual plant mitochondria typically contain less than a full genome [64]. Famous examples in the Silene genus involve the mtDNA genome partitioned into dozens of chromosomes, some of which contain no functional content [74,300]. These subgenomic molecules interact through recombination in a dynamic population [127,301,302], and individual mitochondria share mtDNA and its products through exchange on dynamic ‘social networks’ in the cell [141,298,299,303–305]. When msh1, responsible for organelle DNA maintenance, is perturbed, the dynamics of this social exchange are altered to support more mtDNA sharing [306]. Although less understood than in animals [307], quality control through mitophagy is established in plants [308–311] and likely serves to shape cellular mtDNA populations.

At the population level, the extent of selection on plant mtDNA has (like animals) been subject to debate [312]. MtDNA features clearly give rise to phenotypes that are detrimental to natural plants, including CMS. CMS involves the loss of male fertility which has been linked to mitonuclear interactions and both point mutations and structural rearrangements in mtDNA [139, 193, 313]. While detrimental to natural plants, CMS is of great use in agriculture, where sterile males support high-yielding hybrid production [138,140,141].

Non-chromosomal striping (NCS) is another example of selection linked to tissue-level differences in mitochondrial heteroplasmy. NCS is linked to deletions in mtDNA that impact the electron transport chain and has a more widespread impact on growth and development, including plant stature and yield in maize [314]. Tissue-level differences in heteroplasmy, possibly due to selective amplification of mtDNA fragments, have also been observed in tobacco [315] and rice [316]. Reduced nonsynonymous mutation in functional regions of genome has been reported in Ginkgo and rice [317] and even the selective neutrality of synonymous substitutions is debated, with some recent studies suggesting a role for selection [318].

Although known for over a century and foundational to organelle genetics [319], segregation in plants has classically been challenging to quantify, because the levels of heteroplasmy observed in naturally occurring plants was typically very low. Despite this, segregation has been reported in different taxa including carrot, olives, and Silene [320–322]. The existence and nature of a germline in plants is debated [323], and it does not seem to be the case that plants sequester an animal-like germline. Theory has explored the consequences of this for segregation mechanisms [191], finding that the increase in V(h) through gene conversion can proceed independently of cellular mtDNA copy number, and may, therefore, be a robust strategy in the absence of a physical mtDNA bottleneck.

To increase the quantitative understanding of plant segregation, recent work in Arabidopsis used an msh1 mutant, in which de novo mtDNA (and cpDNA) mutations were readily generated [123]. Some heteroplasmic plants containing an admixture of these mutations and wildtype mtDNA were then back-crossed to the wildtype msh1, leading to plants with substantial heteroplasmy with either wildtype nuclear DNA or the msh1 mutation. Heteroplasmy was tracked in these plants through development and between generations. Segregation was extremely rapid (an effective bottleneck size of ∼4) in the wildtype and seven times slower in the msh1 mutant, pointing to a role for gene conversion in this rapid generation of V(h) [291,292]. Rapid segregation of plant mtDNA is likely to support ‘substoichiometric shifting’ (SSS), a process whereby an mtDNA type that is initially rare comes to dominate a sample [324–326].

Indirect evidence for the role of gene conversion in other plant species comes from a bioinformatic survey showing high expression of organelle recombination machinery in the shoot apical meristem (which will be responsible for producing sex cells) in barley, Medicago, rice, and potato [191]. In the shoot apical meristem (responsible for the aboveground germline), plant mitochondria physically meet in a network [327,328], which could support recombination more readily than the fragmented arrangement in other cell types [191]. In Zostera, powerful modelling work has combined individual and population-wide pictures to explore the roles of segregation and selection in shaping mtDNA [329,330].

Plants have long been observed to display a variety of mitochondrial inheritance strategies [331,332]. [5] provide an excellent review illustrating several of these, including maternal inheritance (common); maternal with paternal leakage (e.g. alfalfa [333]); paternal inheritance (e.g. cucumber [334]) and biparental inheritance (e.g. zonal geranium [335]).

Fungi

Fungal mtDNA also has the capacity for recombination [191,336,337]. Evidence seems mixed on whether recombination occurs readily over organismal (as opposed to evolutionary) timescales, with some studies observing extensive recombination [338,339] and some with little observed [340]. Of course, the observation of recombination will depend on many features including species and the extent of heteroplasmy (as in plants, above).

In addition to random drift [341], various selective pressures have been shown to shape fungal mtDNA. A common example of ‘selfish’ mtDNA behaviour in yeast is the ‘petite’ mutant, harbouring a large-scale deletion that appears to confer a replicative advantage [342–344]. This mutant has been extensively studied, with over 100 nuclear factors shaping its evolutionary dynamics at the cellular level [345]. Recent single-molecule work has characterised the dynamics of generation and proliferation of this mutant, and its link to recombination hotspots in the mtDNA genome [346].

The proliferation of different mtDNA types in fungi in response to different environmental pressures has been observed across species, including for fungicide treatments [347,348], salinity [349], and host species [350] and mtDNA type has been shown to confer temperature tolerance [351]. The action of multi-level selection, within- and between cells, has been characterised in budding yeast [352], with roles for mitochondrial fission and mitophagy identified in shaping heteroplasmic populations [353].

In unicellular organisms, the behaviour of mtDNA at cell divisions determines (largely) mtDNA segregation and (completely) the inheritance of mtDNA [354–356]. The physical process of mtDNA segregation at cell divisions in unicellular fungi has been studied in depth [204], with evidence that yeast controls the partitioning of mtDNA at divisions more tightly than binomial partitioning. Yeast mtDNA inheritance is biparental [3], but selective inheritance of particular mtDNA types has long been observed [344]. In hybrid situations a colony can come to favour one paternal type through preferential (and environmentally determined) retention [357]. Other fungi, including the multicellular Neurospora crassa, exhibit uniparental inheritance and segregation of artificial heteroplasmy over time [358]. Across the kingdom, a range of inheritance and segregation behaviours are observed [336,337]

Protists

Presence of recombination machinery varies across protists [191], but many species have highly fragmented mtDNA genomes that might suggest recombination-mediated coupled [2,68]. Minicircles, almost corresponding to individual mtDNA genes, have been recently reported in red algae [359]. The euglenozoan Diplonema papillatum has multiple small mtDNA fragments smaller than the size of individual genes, which must be spliced together from these fragments [70]. Recent work dramatically increasing the sampling of protist mtDNA has revealed genome plasticity reminiscent of the plant kingdom in stramenopiles [68].

In several protists, a single mitochondrion with a single mtDNA nucleoid exists per cell [360]. The physical segregation machinery has been characterised in the unusual case of trypanosomes [361]. In multicellular protist species, segregation is not to our knowledge well explored. Multicellular algae can have relatively complex developmental plans, somewhat reminiscent of plants, that could conceivably harbour comparable segregation processes [362]. In an interesting parallel to the case of green plants above, ultrastructural analysis has found mitochondria in a brown alga to be generally fragmented except in female gametophytes (perhaps analogous to the reticulated mitochondria in the plant shoot apical meristem) [363].

Instances of external pressures shaping protist mtDNA are as diverse as the species in this section. Heteroplasmy profiles in Fucus have been observed to depend on geography [364]. Selective pressures acting on trypanosome mtDNA have been suggested to include intrinsic factors like translational efficiency and transcript cost [365], and it has been found that mtDNA is essential for the parasite's transmission stage [366]. An interesting branch of research has drawn parallels between mitochondrial disease in Dictyostelium and other taxa, finding that heteroplasmic mtDNA gene disruption has systemic effects on organism physiology [367,368].

Inheritance patterns in protists are as diverse as the species involved. In some slime moulds, mtDNA inheritance has been reported as uniparental [369]. In various marine algae, maternal, paternal, and heteroplasmic mtDNA inheritance has been observed (reviewed in [370]) — including maternal, paternal, and biparental modes within one Porphyra (Rhodophyta) species [371]. An unusual mechanism of triparental inheritance — where mtDNA is inherited from a cell that is neither of the (biparental) nuclear parents — has been observed in Dictyostelium [372] (recalling the artificial introduction of mtDNA from a third-party donor in mitochondrial replacement therapies [137,218,219]).

A synthesis of observations and theories

Having surveyed at least some of the diversity of mtDNA content and behaviour across eukaryotes, are we better placed to answer our original questions? We can at least attempt to synthesise some of the observations we have noted (Figure 4).

Knowledge graph-style synthesis of mtDNA influences.

Figure 4.
Knowledge graph-style synthesis of mtDNA influences.

An outline of the (non-exhaustive) set of influences on coarse-grained mtDNA structure that we have discussed. Nodes are concepts; edges denote links between concepts, labelled including with C, causes; F, favours; S, supports; I, includes. (Left) external factors affecting the poise of recombination and multiscale selection processes acting on mtDNA. (Right) the consequences of these processes for mtDNA behaviour. Code to reproduce this figure is freely available at https://github.com/StochasticBiology/mt-gene-stats.

Figure 4.
Knowledge graph-style synthesis of mtDNA influences.

An outline of the (non-exhaustive) set of influences on coarse-grained mtDNA structure that we have discussed. Nodes are concepts; edges denote links between concepts, labelled including with C, causes; F, favours; S, supports; I, includes. (Left) external factors affecting the poise of recombination and multiscale selection processes acting on mtDNA. (Right) the consequences of these processes for mtDNA behaviour. Code to reproduce this figure is freely available at https://github.com/StochasticBiology/mt-gene-stats.

Close modal

The first clear observation is that the textbook picture of an isolated mammalian mitochondrion with a non-recombining, 16 kb circular mtDNA encoding 13 proteins is unrepresentative of eukaryotes. Gene retention, physical structure, inheritance, and mutational hazard varies hugely across species. Given the similarities in process and machinery to bacterial recombination, mtDNA recombination is likely ancestral (discussed, for example, in [9]) and plays varied roles across kingdoms in repair and segregation of damage. Structural, genetic, and stoichiometric complexity result.

A path through the knowledge graph in Figure 4 can be used to summarise some of the principles in this article. A combination of the physical features of individual genes [38,88] and the challenges faced by mitochondria in an individual species together (and non-exclusively) influence mtDNA gene retention profiles (Figure 1B inset). Strong, dynamic environmental changes favour gene retention for CoRR [76,93,109]. Maintaining mtDNA heterozygosity to adapt to changing environments may also influence which inheritance patterns are favoured [211,215], and the necessity of dealing with differing mutational pressures and maintaining mtDNA may influence germline timing and properties [192,197].

The requirements for repairing consequent mtDNA damage then influence to what extent to mtDNA recombination may be usefully employed by a species. An organism's developmental germline profile also seems to affect whether recombination is used to segregate damage [191] or an animal-like bottleneck strategy of high ploidy is used [192,197]. As mtDNA molecules must physically meet to recombine, the physical dynamics of mitochondria also shape the genetic activity of recombination [191,304,306]. Multiscale mtDNA removal, at the organelle, cellular, or organismal levels, also contributes to damage control and function maintenance. The recombination benefits of templated repair and segregation via gene conversion are balanced by the structural variance induced by recombination, which can lead to genome fragmentation, junk inclusion, and the appearance of selfish elements [2,287].

Across eukaryotes — across organelles?

Many of the arguments outlined above do not particularly require the organelle of interest to be a mitochondrion. We found that the same features of hydrophobicity, GC content, and energetic centrality predict cpDNA gene retention as well as mtDNA retention — and, strikingly, this prediction is quantitative in the sense that a model trained on mtDNA retention profiles predicts cpDNA retention profiles [38]. The theory developed suggesting that strong and dynamic environmental demands favour organelle gene retention also applies to cpDNA [109], and we observed consistencies among environmental features statistically linked with gene retention profiles in both organelles [76]. Indeed, a weak but robust correlation between mtDNA and cpDNA gene counts is detectable in the subset of species for which records are available for both [373]. Symmetry particularly in sets of genes encoding ribosomal proteins in mtDNA and cpDNA has been observed [102]. CpDNA heteroplasmy appears to sorted rapidly and with similar drivers to mtDNA in plants [291,374]. However, the link is perhaps better founded on the left hand side of Figure 4 than the right hand side. The physical embedding of mtDNA and cpDNA can be very different. In plants, mitochondria contain less than a full genome copy [64] and continually meet to exchange contents. Chloroplasts contain many genome copies and are not known to exchange cpDNA [127], so the physical and ‘social’ dynamics described above are likely not comparable.

Beyond chloroplasts, hydrophobicity is also linked to the gene profiles of other endosymbionts [14], including the photosynthetic endosymbiont acquired more recently in Paulinella algae [375,376], the nitroplast [100], numerous endosymbiotic bacteria in insects [14], and other symbiotic bacteria [38]. It is tempting to speculate — though not without caution [2] — that these principles may constitute universal modulators of endosymbiont-organelle genome evolution.

An ongoing synthesis?

Any attempt to describe phenomena across all eukaryotes will necessarily be incomplete. We would like to do two things that are perhaps somewhat unusual. First, we offer our sincere apologies to the authors of studies which are aligned with the topic of this review which we have missed a connection with. In no cases was this deliberate and the corresponding author would (always!) appreciate suggestions of aligned literature. Second, we propose a public document where comments on the manuscript, suggestions of related content, and other aligned messages can be posted. This document can be found here https://tinyurl.com/mtdna-review, and readers should be able to post comments freely and anonymously. We will synthesise content and comments on the Github repository associated with this paper https://github.com/StochasticBiology/mt-gene-stats.

The authors declare that there are no competing interests associated with the manuscript.

Iain G. Johnston: Conceptualization, Resources, Data curation, Software, Supervision, Funding acquisition, Investigation, Visualization, Methodology, Writing — original draft, Project administration, Writing — review and editing. Shibani Veeraragavan: Conceptualization, Resources, Investigation, Writing — review and editing. Maria Johansen: Resources, Investigation, Writing — review and editing.

The authors are grateful to members of the Stochastic Biology Group for useful discussions. This project has received funding from the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (Grant agreement no. 805046 (EvoConBiO) to I.G.J.).

CMS

cytoplasmic male sterility

MROs

mitochondrion-related organelles

mtDNA

mitochondrial DNA

NCS

non-chromosomal striping

SSS

substoichiometric shifting

1
Roger
,
A.J.
,
Muñoz-Gómez
,
S.A.
and
Kamikawa
,
R.
(
2017
)
The origin and diversification of mitochondria
.
Curr. Biol.
27
,
R1177
R1192
2
Smith
,
D.R.
and
Keeling
,
P.J.
(
2015
)
Mitochondrial and plastid genome architecture: reoccurring themes, but significant differences at the extremes
.
Proc. Natl Acad. Sci. U.S.A.
112
,
10177
10184
3
Birky
,
C.W.
(
2001
)
The inheritance of genes in mitochondria and chloroplasts: laws, mechanisms, and models
.
Annu. Rev. Genet.
35
,
125
148
4
Camus
,
M.F.
,
Alexander-Lawrie
,
B.
,
Sharbrough
,
J.
and
Hurst
,
G.D.
(
2022
)
Inheritance through the cytoplasm
.
Heredity
129
,
31
43
5
Greiner
,
S.
,
Sobanski
,
J.
and
Bock
,
R.
(
2015
)
Why are most organelle genomes transmitted maternally?
Bioessays
37
,
80
94
6
Bendich
,
A.J.
(
1987
)
Why do chloroplasts and mitochondria contain so many copies of their genome?
Bioessays
6
,
279
282
7
Cole
,
L.W.
(
2016
)
The evolution of per-cell organelle number
.
Front. Cell Dev. Biol.
4
,
85
8
Marinos
,
E.
(
1985
)
The number of mitochondria in Xenopus laevis ovulated oocytes
.
Cell Differ.
16
,
139
143
9
Zwonitzer
,
K.D.
,
Tressel
,
L.G.
,
Wu
,
Z.
,
Kan
,
S.
,
Broz
,
A.K.
,
Mower
,
J.P.
et al (
2024
)
Genome copy number predicts extreme evolutionary rate variation in plant mitochondrial DNA
.
Proc. Natl Acad. Sci. U.S.A.
121
,
e2317240121
10
Allen
,
J.F.
and
Raven
,
J.A.
(
1996
)
Free-radical-induced mutation vs redox regulation: costs and benefits of genes in organelles
.
J. Mol. Evol.
42
,
482
492
11
Lynch
,
M.
(
1997
)
Mutation accumulation in nuclear, organelle, and prokaryotic transfer RNA genes
.
Mol. Biol. Evol.
14
,
914
925
12
Lynch
,
M.
,
Koskella
,
B.
and
Schaack
,
S.
(
2006
)
Mutation pressure and the evolution of organelle genomic architecture
.
Science
311
,
1727
1730
13
Lynch
,
M.
and
Blanchard
,
J.L.
(
1998
) Deleterious mutation accumulation in organelle genomes. In
Mutation and Evolution
(
Woodruff
,
R.C.
,
Thompson
,
J.N.
, eds), vol.
7
, pp.
29
39
,
Springer
,
Dordrecht, Netherlands
14
McCutcheon
,
J.P.
and
Moran
,
N.A.
(
2012
)
Extreme genome reduction in symbiotic bacteria
.
Nat. Rev. Microbiol.
10
,
13
26
15
Muller
,
H.J.
(
1964
)
The relation of recombination to mutational advance
.
Mutat. Res.
1
,
2
9
16
Radzvilavicius
,
A.L.
,
Kokko
,
H.
and
Christie
,
J.R.
(
2017
)
Mitigating mitochondrial genome erosion without recombination
.
Genetics
207
,
1079
1088
17
Kelly
,
S.
(
2021
)
The economics of organellar gene loss and endosymbiotic gene transfer
.
Genome Biol.
22
,
345
18
Hill
,
G.E.
,
Havird
,
J.C.
,
Sloan
,
D.B.
,
Burton
,
R.S.
,
Greening
,
C.
and
Dowling
,
D.K.
(
2019
)
Assessing the fitness consequences of mitonuclear interactions in natural populations
.
Biol. Rev.
94
,
1089
1104
19
Gray
,
M.W.
(
2012
)
Mitochondrial evolution
.
Cold Spring Harb. Perspect. Biol.
4
,
a011403
20
Yang
,
D.
,
Oyaizu
,
Y.
,
Oyaizu
,
H.
,
Olsen
,
G.J.
and
Woese
,
C.R.
(
1985
)
Mitochondrial origins
.
Proc. Natl Acad. Sci. U.S.A.
82
,
4443
4447
21
Wang
,
Z.
and
Wu
,
M.
(
2015
)
An integrated phylogenomic approach toward pinpointing the origin of mitochondria
.
Sci. Rep.
5
,
7949
22
Eme
,
L.
,
Spang
,
A.
,
Lombard
,
J.
,
Stairs
,
C.W.
and
Ettema
,
T.J.
(
2017
)
Archaea and the origin of eukaryotes
.
Nat. Rev. Microbiol.
15
,
711
723
23
Spang
,
A.
,
Saw
,
J.H.
,
Jørgensen
,
S.L.
,
Zaremba-Niedzwiedzka
,
K.
,
Martijn
,
J.
,
Lind
,
A.E.
et al (
2015
)
Complex archaea that bridge the gap between prokaryotes and eukaryotes
.
Nature
521
,
173
179
24
Zaremba-Niedzwiedzka
,
K.
,
Caceres
,
E.F.
,
Saw
,
J.H.
,
Bäckström
,
D.
,
Juzokaite
,
L.
,
Vancaester
,
E.
et al (
2017
)
Asgard archaea illuminate the origin of eukaryotic cellular complexity
.
Nature
541
,
353
358
25
Embley
,
T.M.
and
Martin
,
W.
(
2006
)
Eukaryotic evolution, changes and challenges
.
Nature
440
,
623
630
26
Goksøyr
,
J.
(
1967
)
Evolution of eucaryotic cells
.
Nature
214
,
1161
27
Martin
,
W.F.
,
Garg
,
S.
and
Zimorski
,
V.
(
2015
)
Endosymbiotic theories for eukaryote origin
.
Philos. Trans. R. Soc. Lond. B Biol. Sci.
370
,
20140330
28
Sagan
,
L.
(
1967
)
On the origin of mitosing cells
.
J. Theor. Biol.
14
,
225-IN6
29
Zimorski
,
V.
,
Ku
,
C.
,
Martin
,
W.F.
and
Gould
,
S.B.
(
2014
)
Endosymbiotic theory for organelle origins
.
Curr. Opin. Microbiol.
22
,
38
48
30
Gabaldón
,
T.
and
Huynen
,
M.A.
(
2007
)
From endosymbiont to host-controlled organelle: the hijacking of mitochondrial protein synthesis and metabolism
.
PLoS Comput. Biol.
3
,
e219
31
Geiger
,
O.
,
Sanchez-Flores
,
A.
,
Padilla-Gomez
,
J.
and
Degli Esposti
,
M.
(
2023
)
Multiple approaches of cellular metabolism define the bacterial ancestry of mitochondria
.
Sci. Adv.
9
,
eadh0066
32
Gabaldón
,
T.
and
Huynen
,
M.A.
(
2003
)
Reconstruction of the proto-mitochondrial metabolism
.
Science
301
,
609
33
Thiergart
,
T.
,
Landan
,
G.
,
Schenk
,
M.
,
Dagan
,
T.
and
Martin
,
W.F.
(
2012
)
An evolutionary network of genes present in the eukaryote common ancestor polls genomes on eukaryotic and mitochondrial origin
.
Genome Biol. Evol.
4
,
466
485
34
Wang
,
Z.
and
Wu
,
M.
(
2014
)
Phylogenomic reconstruction indicates mitochondrial ancestor was an energy parasite
.
PLoS ONE
9
,
e110685
35
Janouškovec
,
J.
,
Tikhonenkov
,
D.V.
,
Burki
,
F.
,
Howe
,
A.T.
,
Rohwer
,
F.L.
,
Mylnikov
,
A.P.
et al (
2017
)
A New lineage of eukaryotes illuminates early mitochondrial genome reduction
.
Curr. Biol.
27
,
3717
3724.e5
36
Speijer
,
D.
,
Hammond
,
M.
and
Lukeš
,
J.
(
2020
)
Comparing early eukaryotic integration of mitochondria and chloroplasts in the light of internal ROS challenges: timing is of the essence
.
mBio
11
,
10.1128/mbio.00955-20
37
Doolittle
,
W.F.
(
1998
)
You are what you eat: a gene transfer ratchet could account for bacterial genes in eukaryotic nuclear genomes
.
Trends Genet.
14
,
307
311
38
Giannakis
,
K.
,
Arrowsmith
,
S.J.
,
Richards
,
L.
,
Gasparini
,
S.
,
Chustecki
,
J.M.
,
Røyrvik
,
E.C.
et al (
2022
)
Evolutionary inference across eukaryotes identifies universal features shaping organelle gene retention
.
Cell Syst.
13
,
874
884.e5
39
Timmis
,
J.N.
,
Ayliffe
,
M.A.
,
Huang
,
C.Y.
and
Martin
,
W.
(
2004
)
Endosymbiotic gene transfer: organelle genomes forge eukaryotic chromosomes
.
Nat. Rev. Genet.
5
,
123
135
40
Adams
,
K.L.
and
Palmer
,
J.D.
(
2003
)
Evolution of mitochondrial gene content: gene loss and transfer to the nucleus
.
Mol. Phylogenet. Evol.
29
,
380
395
41
Smith
,
D.R.
(
2016
)
The mutational hazard hypothesis of organelle genome evolution: 10 years on
.
Mol. Ecol.
25
,
3769
3775
42
Blanchard
,
J.L.
and
Lynch
,
M.
(
2000
)
Organellar genes: why do they end up in the nucleus?
Trends Genet.
16
,
315
320
43
Saccone
,
C.
,
Gissi
,
C.
,
Lanave
,
C.
,
Larizza
,
A.
,
Pesole
,
G.
and
Reyes
,
A.
(
2000
)
Evolution of the mitochondrial genetic system: an overview
.
Gene
261
,
153
159
44
Hazkani-Covo
,
E.
and
Martin
,
W.F.
(
2017
)
Quantifying the number of independent organelle DNA insertions in genome evolution and human health
.
Genome Biol. Evol.
9
,
1190
1203
45
Richly
,
E.
and
Leister
,
D.
(
2004
)
NUMTs in sequenced eukaryotic genomes
.
Mol. Biol. Evol.
21
,
1081
1084
46
Wei
,
W.
,
Schon
,
K.R.
,
Elgar
,
G.
,
Orioli
,
A.
,
Tanguy
,
M.
,
Giess
,
A.
et al (
2022
)
Nuclear-embedded mitochondrial DNA sequences in 66,083 human genomes
.
Nature
611
,
105
114
47
Bock
,
R.
(
2017
)
Witnessing genome evolution: experimental reconstruction of endosymbiotic and horizontal gene transfer
.
Annu. Rev. Genet.
51
,
1
22
48
Hazkani-Covo
,
E.
,
Zeller
,
R.M.
and
Martin
,
W.
(
2010
)
Molecular poltergeists: mitochondrial DNA copies (numts) in sequenced nuclear genomes
.
PLoS Genet.
6
,
e1000834
49
Berg
,
O.G.
and
Kurland
,
C.G.
(
2000
)
Why mitochondrial genes are most often found in nuclei
.
Mol. Biol. Evol.
17
,
951
961
50
Butenko
,
A.
,
Lukeš
,
J.
,
Speijer
,
D.
and
Wideman
,
J.G.
(
2024
)
Mitochondrial genomes revisited: why do different lineages retain different genes?
BMC Biol.
22
,
15
51
Brennicke
,
A.
,
Grohmann
,
L.
,
Hiesel
,
R.
,
Knoop
,
V.
and
Schuster
,
W.
(
1993
)
The mitochondrial genome on its way to the nucleus: different stages of gene transfer in higher plants
.
FEBS Lett.
325
,
140
145
52
Burger
,
G.
,
Gray
,
M.W.
,
Forget
,
L.
and
Lang
,
B.F.
(
2013
)
Strikingly bacteria-like and gene-rich mitochondrial genomes throughout jakobid protists
.
Genome Biol. Evol.
5
,
418
438
53
Lang
,
B.F.
,
Burger
,
G.
,
O'Kelly
,
C.J.
,
Cedergren
,
R.
,
Golding
,
G.B.
,
Lemieux
,
C.
et al (
1997
)
An ancestral mitochondrial DNA resembling a eubacterial genome in miniature
.
Nature
387
,
493
497
54
Johnston
,
I.G.
and
Williams
,
B.P.
(
2016
)
Evolutionary inference across eukaryotes identifies specific pressures favoring mitochondrial gene retention
.
Cell Syst.
2
,
101
111
55
Kannan
,
S.
,
Rogozin
,
I.B.
and
Koonin
,
E.V.
(
2014
)
MitoCOGs: clusters of orthologous genes from mitochondria and implications for the evolution of eukaryotes
.
BMC Evol. Biol.
14
,
237
56
Muthye
,
V.
and
Lavrov
,
D.V.
(
2021
)
Multiple losses of MSH1, gain of mtMutS, and other changes in the MutS family of DNA repair proteins in animals
.
Genome Biol. Evol.
13
,
evab191
57
Pont-Kingdon
,
G.
,
Okada
,
N.A.
,
Macfarlane
,
J.L.
,
Beagley
,
C.T.
,
Watkins-Sims
,
C.D.
,
Cavalier-Smith
,
T.
et al (
1998
)
Mitochondrial DNA of the coral Sarcophyton glaucum contains a gene for a homologue of bacterial muts: a possible case of gene transfer from the nucleus to the mitochondrion
.
J. Mol. Evol.
46
,
419
431
58
Bilewitch
,
J.P.
and
Degnan
,
S.M.
(
2011
)
A unique horizontal gene transfer event has provided the octocoral mitochondrial genome with an active mismatch repair gene that has potential for an unusual self-contained function
.
BMC Evol. Biol.
11
,
228
59
Milner
,
D.S.
,
Wideman
,
J.G.
,
Stairs
,
C.W.
,
Dunn
,
C.D.
and
Richards
,
T.A.
(
2021
)
A functional bacteria-derived restriction modification system in the mitochondrion of a heterotrophic protist
.
PLoS Biol.
19
,
e3001126
60
Burger
,
G.
,
Gray
,
M.W.
and
Lang
,
B.F.
(
2003
)
Mitochondrial genomes: anything goes
.
Trends Genet.
19
,
709
716
61
Pohjoismäki
,
J.L.
,
Goffart
,
S.
,
Tyynismaa
,
H.
,
Willcox
,
S.
,
Ide
,
T.
,
Kang
,
D.
et al (
2009
)
Human heart mitochondrial DNA is organized in complex catenated networks containing abundant four-way junctions and replication forks
.
J. Biol. Chem.
284
,
21446
21457
62
Shao
,
R.
,
Zhu
,
X.-Q.
,
Barker
,
S.C.
and
Herd
,
K.
(
2012
)
Evolution of extensively fragmented mitochondrial genomes in the lice of humans
.
Genome Biol. Evol.
4
,
1088
1101
63
Smith
,
D.R.
,
Kayal
,
E.
,
Yanagihara
,
A.A.
,
Collins
,
A.G.
,
Pirro
,
S.
and
Keeling
,
P.J.
(
2012
)
First complete mitochondrial genome sequence from a box jellyfish reveals a highly fragmented linear architecture and insights into telomere evolution
.
Genome Biol. Evol.
4
,
52
58
64
Preuten
,
T.
,
Cincu
,
E.
,
Fuchs
,
J.
,
Zoschke
,
R.
,
Liere
,
K.
and
Börner
,
T.
(
2010
)
Fewer genes than organelles: extremely low and variable gene copy numbers in mitochondria of somatic plant cells
.
Plant J.
64
,
948
959
65
Bendich
,
A.J.
(
2007
)
The size and form of chromosomes are constant in the nucleus, but highly variable in bacteria, mitochondria and chloroplasts
.
Bioessays
29
,
474
483
66
Smith
,
D.R.
and
Keeling
,
P.J.
(
2013
)
Gene conversion shapes linear mitochondrial genome architecture
.
Genome Biol. Evol.
5
,
905
912
67
Nosek
,
J.
and
Tomáška
,
L.
(
2003
)
Mitochondrial genome diversity: evolution of the molecular architecture and replication strategy
.
Curr. Genet.
44
,
73
84
68
Wideman
,
J.G.
,
Monier
,
A.
,
Rodríguez-Martínez
,
R.
,
Leonard
,
G.
,
Cook
,
E.
,
Poirier
,
C.
et al (
2020
)
Unexpected mitochondrial genome diversity revealed by targeted single-cell genomics of heterotrophic flagellated protists
.
Nat. Microbiol.
5
,
154
165
69
Smith
,
D.R.
and
Keeling
,
P.J.
(
2016
)
Protists and the wild, wild west of gene expression: new frontiers, lawlessness, and misfits
.
Annu. Rev. Microbiol.
70
,
161
178
70
Vlcek
,
C.
,
Marande
,
W.
,
Teijeiro
,
S.
,
Lukeš
,
J.
and
Burger
,
G.
(
2011
)
Systematically fragmented genes in a multipartite mitochondrial genome
.
Nucleic Acids Res.
39
,
979
988
71
Shapiro
,
T.A.
and
Englund
,
P.T.
(
1995
)
The structure and replication of kinetoplast DNA
.
Annu. Rev. Microbiol.
49
,
117
143
72
Clark
,
K.A.
,
Howe
,
D.K.
,
Gafner
,
K.
,
Kusuma
,
D.
,
Ping
,
S.
,
Estes
,
S.
et al
2012
)
Selfish little circles: transmission bias and evolution of large deletion-bearing mitochondrial DNA in Caenorhabditis briggsae nematodes
.
PLoS ONE
7
,
e41433
.
73
Férandon
,
C.
,
Xu
,
J.
and
Barroso
,
G.
(
2013
)
The 135 kbp mitochondrial genome of Agaricus bisporus is the largest known eukaryotic reservoir of group I introns and plasmid-related sequences
.
Fungal Genet. Biol.
55
,
85
91
74
Sloan
,
D.B.
,
Alverson
,
A.J.
,
Chuckalovcak
,
J.P.
,
Wu
,
M.
,
McCauley
,
D.E.
,
Palmer
,
J.D.
et al (
2012
)
Rapid evolution of enormous, multichromosomal genomes in flowering plant mitochondria with exceptionally high mutation rates
.
PLoS Biol.
10
,
e1001241
75
Mower
,
J.P.
(
2020
)
Variation in protein gene and intron content among land plant mitogenomes
.
Mitochondrion
53
,
203
213
76
Giannakis
,
K.
,
Richards
,
L.
and
Johnston
,
I.G.
(
2024
)
Ecological predictors of organelle genome evolution: phylogenetic correlations with taxonomically broad, sparse, unsystematized data
.
Syst. Biol.
syae009
73
,
419
433
77
Yahalomi
,
D.
,
Atkinson
,
S.D.
,
Neuhof
,
M.
,
Chang
,
E.S.
,
Philippe
,
H.
,
Cartwright
,
P.
et al (
2020
)
A cnidarian parasite of salmon (Myxozoa: Henneguya) lacks a mitochondrial genome
.
Proc. Natl Acad. Sci. U.S.A.
117
,
5358
5363
78
de Paula
,
W.B.M.
,
Allen
,
J.F.
and
van der Giezen
,
M.
(
2012
) Mitochondria, hydrogenosomes and mitosomes in relation to the corr hypothesis for genome function and evolution. In
Organelle Genetics: Evolution of Organelle Genomes and Gene Expression
(
Bullerwell
,
C.E.
, ed.), pp.
105
119
,
Springer
,
Berlin, Heidelberg
79
Hjort
,
K.
,
Goldberg
,
A.V.
,
Tsaousis
,
A.D.
,
Hirt
,
R.P.
and
Embley
,
T.M.
(
2010
)
Diversity and reductive evolution of mitochondria among microbial eukaryotes
.
Philos. Trans. R. Soc. Lond. B Biol. Sci.
365
,
713
727
80
Maciszewski
,
K.
and
Karnkowska
,
A.
(
2019
)
Should I stay or should I go? Retention and loss of components in vestigial endosymbiotic organelles
.
Curr. Opin. Genet. Dev.
58–59
,
33
39
81
Makiuchi
,
T.
and
Nozaki
,
T.
(
2014
)
Highly divergent mitochondrion-related organelles in anaerobic parasitic protozoa
.
Biochimie
100
,
3
17
82
Stairs
,
C.W.
,
Leger
,
M.M.
and
Roger
,
A.J.
(
2015
)
Diversity and origins of anaerobic metabolism in mitochondria and related organelles
.
Phil. Trans. R. Soc. B
370
,
20140326
83
Müller
,
M.
,
Mentel
,
M.
,
van Hellemond
,
J.J.
,
Henze
,
K.
,
Woehle
,
C.
,
Gould
,
S.B.
et al (
2012
)
Biochemistry and evolution of anaerobic energy metabolism in eukaryotes
.
Microbiol. Mol. Biol. Rev.
76
,
444
495
84
Karnkowska
,
A.
,
Vacek
,
V.
,
Zubáčová
,
Z.
,
Treitli
,
S.C.
,
Petrželková
,
R.
,
Eme
,
L.
et al (
2016
)
A eukaryote without a mitochondrial organelle
.
Curr. Biol.
26
,
1274
1284
85
John
,
U.
,
Lu
,
Y.
,
Wohlrab
,
S.
,
Groth
,
M.
,
Janouškovec
,
J.
,
Kohli
,
G.S.
et al (
2019
)
An aerobic eukaryotic parasite with functional mitochondria that likely lacks a mitochondrial genome
.
Sci. Adv.
5
,
eaav1110
86
Kayal
,
E.
and
Smith
,
D.R.
(
2021
)
Is the dinoflagellate amoebophrya really missing an mtDNA?
Mol. Biol. Evol.
38
,
2493
2496
87
Lee
,
R.W.
and
Hua
,
J
. (
2018
) Mitochondrial genomes of green, red and glaucophyte algae. In
Molecular Life Sciences: An Encyclopedic Reference
(
Wells
,
R.D.
,
Bond
,
J.S.
,
Klinman
,
J.
,
Masters
,
B.S.S.
, eds), pp.
762
769
,
Springer
,
New York, NY
88
Johnston
,
I.G.
and
Williams
,
B.P.
(
2016
)
Evolutionary inference across eukaryotes identifies specific pressures favoring mitochondrial gene retention
.
Cell Syst.
2
,
101
111
89
Salinas-Giegé
,
T.
,
Giegé
,
R.
and
Giegé
,
P.
(
2015
)
tRNA biology in mitochondria
.
Int. J. Mol. Sci.
16
,
4518
4559
90
Warren
,
J.M.
,
Broz
,
A.K.
,
Martinez-Hottovy
,
A.
,
Elowsky
,
C.
,
Christensen
,
A.C.
and
Sloan
,
D.B.
(
2023
)
Rewiring of aminoacyl-tRNA synthetase localization and interactions in plants with extensive mitochondrial tRNA gene loss
.
Mol. Biol. Evol.
40
,
msad163
91
Björkholm
,
P.
,
Harish
,
A.
,
Hagström
,
E.
,
Ernst
,
A.M.
and
Andersson
,
S.G.
(
2015
)
Mitochondrial genomes are retained by selective constraints on protein targeting
.
Proc. Natl Acad. Sci. U.S.A.
112
,
10154
10161
92
von Heijne
,
G.
(
1986
)
Why mitochondria need a genome
.
FEBS Lett.
198
,
1
4
93
Allen
,
J.F.
(
2015
)
Why chloroplasts and mitochondria retain their own genomes and genetic systems: colocation for redox regulation of gene expression
.
Proc. Natl Acad. Sci. U.S.A.
112
,
10231
10238
94
Allen
,
J.F.
and
Martin
,
W.F.
(
2016
)
Why have organelles retained genomes?
Cell Syst.
2
,
70
72
95
Wright
,
A.F.
,
Murphy
,
M.P.
and
Turnbull
,
D.M.
(
2009
)
Do organellar genomes function as long-term redox damage sensors?
Trends Genet.
25
,
253
261
96
Martin
,
W.
and
Schnarrenberger
,
C.
(
1997
)
The evolution of the Calvin cycle from prokaryotic to eukaryotic chromosomes: a case study of functional redundancy in ancient pathways through endosymbiosis
.
Curr. Genet.
32
,
1
18
97
De Grey A
,
D.N.J.
(
2005
)
Forces maintaining organellar genomes: is any as strong as genetic code disparity or hydrophobicity?
Bioessays
27
,
436
446
98
Reyes
,
A.
,
Gissi
,
C.
,
Pesole
,
G.
and
Saccone
,
C.
(
1998
)
Asymmetrical directional mutation pressure in the mitochondrial genome of mammals
.
Mol. Biol. Evol.
15
,
957
966
99
Smith
,
D.R.
(
2012
)
Updating our view of organelle genome nucleotide landscape
.
Front. Gene
3
,
175
100
Johnston
,
I.G.
(
2024
)
The nitroplast and its relatives support a universal model of features predicting gene retention in endosymbiont and organelle genomes
.
Genome Biol. Evol.
16
,
evae132
101
Levy
,
E.D.
,
Erba
,
E.B.
,
Robinson
,
C.V.
and
Teichmann
,
S.A.
(
2008
)
Assembly reflects evolution of protein complexes
.
Nature
453
,
1262
1265
102
Maier
,
U.-G.
,
Zauner
,
S.
,
Woehle
,
C.
,
Bolte
,
K.
,
Hempel
,
F.
,
Allen
,
J.F.
et al (
2013
)
Massively convergent evolution for ribosomal protein gene content in plastid and mitochondrial genomes
.
Genome Biol. Evol.
5
,
2318
2329
103
Samuels
,
D.C.
(
2005
)
Life span is related to the free energy of mitochondrial DNA
.
Mech. Ageing Dev.
126
,
1123
1129
104
Mathur
,
V.
,
Wakeman
,
K.C.
and
Keeling
,
P.J.
(
2021
)
Parallel functional reduction in the mitochondria of apicomplexan parasites
.
Curr. Biol.
31
,
2920
2928
105
Sanchez-Puerta
,
M.V.
,
Ceriotti
,
L.F.
,
Gatica-Soria
,
L.M.
,
Roulet
,
M.E.
,
Garcia
,
L.E.
and
Sato
,
H.A.
(
2023
)
Beyond parasitic convergence: unravelling the evolution of the organellar genomes in holoparasites
.
Ann. Bot.
132
,
909
928
106
Santos
,
H.J.
,
Makiuchi
,
T.
and
Nozaki
,
T.
(
2018
)
Reinventing an organelle: the reduced mitochondrion in parasitic protists
.
Trends Parasitol.
34
,
1038
1055
107
Brandvain
,
Y.
,
Barker
,
M.S.
and
Wade
,
M.J.
(
2007
)
Gene co-inheritance and gene transfer
.
Science
315
,
1685
108
Brandvain
,
Y.
and
Wade
,
M.J.
(
2009
)
The functional transfer of genes from the mitochondria to the nucleus: the effects of selection, mutation, population size and rate of self-fertilization
.
Genetics
182
,
1129
1139
109
García-Pascual
,
B.
,
Nordbotten
,
J.M.
and
Johnston
,
I.G.
(
2022
)
Cellular and environmental dynamics influence species-specific extents of organelle gene retention
.
Proc. R. Soc. B: Biol. Sci.
290
,
20222140
110
Itsara
,
L.S.
,
Kennedy
,
S.R.
,
Fox
,
E.J.
,
Yu
,
S.
,
Hewitt
,
J.J.
,
Sanchez-Contreras
,
M.
et al (
2014
)
Oxidative stress is not a major contributor to somatic mitochondrial DNA mutations
.
PLoS Genet.
10
,
e1003974
111
Kennedy
,
S.R.
,
Salk
,
J.J.
,
Schmitt
,
M.W.
and
Loeb
,
L.A.
(
2013
)
Ultra-sensitive sequencing reveals an age-related increase in somatic mitochondrial mutations that are inconsistent with oxidative damage
.
PLoS Genet.
9
,
e1003794
112
Cadet
,
J.
,
Delatour
,
T.
,
Douki
,
T.
,
Gasparutto
,
D.
,
Pouget
,
J.-P.
,
Ravanat
,
J.-L.
et al (
1999
)
Hydroxyl radicals and DNA base damage
.
Mutat. Res.
424
,
9
21
113
Graziewicz
,
M.A.
,
Bienstock
,
R.J.
and
Copeland
,
W.C.
(
2007
)
The DNA polymerase γ Y955C disease variant associated with PEO and parkinsonism mediates the incorporation and translesion synthesis opposite 7,8-dihydro-8-oxo-2′-deoxyguanosine
.
Hum. Mol. Genet.
16
,
2729
2739
114
Pinz
,
K.G.
,
Shibutani
,
S.
and
Bogenhagen
,
D.F.
(
1995
)
Action of mitochondrial DNA polymerase γ at sites of base loss or oxidative damage
.
J. Biol. Chem.
270
,
9202
9206
115
Lynch
,
M.
(
1996
)
Mutation accumulation in transfer RNAs: molecular evidence for Muller's ratchet in mitochondrial genomes
.
Mol. Biol. Evol.
13
,
209
220
116
Lynch
,
M.
(
2010
)
Evolution of the mutation rate
.
Trends Genet.
26
,
345
352
117
Whittle
,
C.-A.
and
Johnston
,
M.O.
(
2002
)
Male-driven evolution of mitochondrial and chloroplastidial DNA sequences in plants
.
Mol. Biol. Evol.
19
,
938
949
118
Xu
,
S.
,
Van Tran
,
K.
,
Neupane
,
S.
,
Snyman
,
M.
,
Huynh
,
T.V.
and
Sung
,
W.
(
2017
)
Single-sperm sequencing reveals the accelerated mitochondrial mutation rate in male Daphnia pulex (Crustacea. Cladocera)
.
Proc. R. Soc. B: Biol. Sci.
284
,
20171548
119
Zhu
,
A.
,
Guo
,
W.
,
Jain
,
K.
and
Mower
,
J.P.
(
2014
)
Unprecedented heterogeneity in the synonymous substitution rate within a plant genome
.
Mol. Biol. Evol.
31
,
1228
1236
120
Palmer
,
J.D.
and
Herbon
,
L.A.
(
1988
)
Plant mitochondrial DNA evolved rapidly in structure, but slowly in sequence
.
J. Mol. Evol.
28
,
87
97
121
Hellberg
,
M.E.
(
2006
)
No variation and low synonymous substitution rates in coral mtDNA despite high nuclear variation
.
BMC Evol. Biol.
6
,
24
122
Huang
,
D.
,
Meier
,
R.
,
Todd
,
P.A.
and
Chou
,
L.M.
(
2008
)
Slow mitochondrial COI sequence evolution at the base of the metazoan tree and its implications for DNA barcoding
.
J. Mol. Evol.
66
,
167
174
123
Wu
,
Z.
,
Waneka
,
G.
,
Broz
,
A.K.
,
King
,
C.R.
and
Sloan
,
D.B.
(
2020
)
MSH1 is required for maintenance of the low mutation rates in plant mitochondrial and plastid genomes
.
Proc. Natl Acad. Sci. U.S.A.
117
,
16448
16455
124
Chen
,
X.J.
(
2013
)
Mechanism of homologous recombination and implications for aging-related deletions in mitochondrial DNA
.
Microbiol. Mol. Biol. Rev.
77
,
476
496
125
Gualberto
,
J.M.
,
Mileshina
,
D.
,
Wallet
,
C.
,
Niazi
,
A.K.
,
Weber-Lotfi
,
F.
and
Dietrich
,
A.
(
2014
)
The plant mitochondrial genome: dynamics and maintenance
.
Biochimie
100
,
107
120
126
Oldenburg
,
D.J.
and
Bendich
,
A.J.
(
2015
)
DNA maintenance in plastids and mitochondria of plants
.
Front. Plant. Sci.
6
,
883
127
Johnston
,
I.G.
(
2019
)
Tension and resolution: dynamic, evolving populations of organelle genomes within plant cells
.
Mol. Plant.
12
,
764
783
128
Christensen
,
A.C.
(
2017
)
Mitochondrial DNA repair and genome evolution
.
Annu. Plant Rev.
50
,
11
31
129
Johnston
,
I.G.
and
Burgstaller
,
J.P.
(
2019
)
Evolving mtDNA populations within cells
.
Biochem. Soc. Trans.
47
,
1367
1382
130
Stewart
,
J.B.
and
Chinnery
,
P.F.
(
2015
)
The dynamics of mitochondrial DNA heteroplasmy: implications for human health and disease
.
Nat. Rev. Genet.
16
,
530
542
131
Van den Ameele
,
J.
,
Li
,
A.Y.
,
Ma
,
H.
and
Chinnery
,
P.F.
(
2020
)
Mitochondrial heteroplasmy beyond the oocyte bottleneck
.
Semin. Cell Dev. Biol.
97
,
156
166
132
Wallace
,
D.C.
and
Chalkia
,
D.
(
2013
)
Mitochondrial DNA genetics and the heteroplasmy conundrum in evolution and disease
.
Cold Spring Harb. Perspect. Biol.
5
,
a021220
133
Guo
,
Y.
,
Li
,
C.-I.
,
Sheng
,
Q.
,
Winther
,
J.F.
,
Cai
,
Q.
,
Boice
,
J.D.
et al (
2013
)
Very low-level heteroplasmy mtDNA variations are inherited in humans
.
J. Genet. Genomics
40
,
607
615
134
Payne
,
B.A.
,
Wilson
,
I.J.
,
Yu-Wai-Man
,
P.
,
Coxhead
,
J.
,
Deehan
,
D.
,
Horvath
,
R.
et al (
2013
)
Universal heteroplasmy of human mitochondrial DNA
.
Hum. Mol. Genet.
22
,
384
390
135
Rensch
,
T.
,
Villar
,
D.
,
Horvath
,
J.
,
Odom
,
D.T.
and
Flicek
,
P.
(
2016
)
Mitochondrial heteroplasmy in vertebrates using ChIP-sequencing data
.
Genome Biol.
17
,
139
136
Rossignol
,
R.
,
Faustin
,
B.
,
Rocher
,
C.
,
Malgat
,
M.
,
Mazat
,
J.-P.
and
Letellier
,
T.
(
2003
)
Mitochondrial threshold effects
.
Biochem. J.
370
,
751
762
137
Burgstaller
,
J.P.
,
Johnston
,
I.G.
and
Poulton
,
J.
(
2015
)
Mitochondrial DNA disease and developmental implications for reproductive strategies
.
Mol. Hum. Reprod.
21
,
11
22
138
Bohra
,
A.
,
Jha
,
U.C.
,
Adhimoolam
,
P.
,
Bisht
,
D.
and
Singh
,
N.P.
(
2016
)
Cytoplasmic male sterility (CMS) in hybrid breeding in field crops
.
Plant Cell Rep.
35
,
967
993
139
Chen
,
L.
and
Liu
,
Y.-G.
(
2014
)
Male sterility and fertility restoration in crops
.
Annu. Rev. Plant Biol.
65
,
579
606
140
Havey
,
M.J.
(
2004
) The use of cytoplasmic male sterility for hybrid seed production. In
Molecular Biology and Biotechnology of Plant Organelles: Chloroplasts and Mitochondria
(
Daniell
,
H.
,
Chase
,
C
., eds), pp.
623
634
,
Springer
,
Dordrecht, Netherlands
141
Chustecki
,
J.M.
and
Johnston
,
I.G.
(
2024
)
Collective mitochondrial dynamics resolve conflicting cellular tensions: from plants to general principles
.
Semin. Cell Dev. Biol.
156
,
253
265
142
Lane
,
N.
(
2012
)
The problem with mixing mitochondria
.
Cell
151
,
246
248
143
Latorre-Pellicer
,
A.
,
Lechuga-Vieco
,
A.V.
,
Johnston
,
I.G.
,
Hämäläinen
,
R.H.
,
Pellico
,
J.
,
Justo-Mendez
,
R.
et al (
2019
)
Regulation of mother-to-offspring transmission of mtDNA heteroplasmy
.
Cell Metab.
30
,
1120
1130
144
Sharpley
,
M.S.
,
Marciniak
,
C.
,
Eckel-Mahan
,
K.
,
McManus
,
M.
,
Crimi
,
M.
,
Waymire
,
K.
et al (
2012
)
Heteroplasmy of mouse mtDNA is genetically unstable and results in altered behavior and cognition
.
Cell
151
,
333
343
145
Wonnapinij
,
P.
,
Chinnery
,
P.F.
and
Samuels
,
D.C.
(
2008
)
The distribution of mitochondrial DNA heteroplasmy due to random genetic drift
.
Am. J. Hum. Genet.
83
,
582
593
146
Kimura
,
M.
(
1955
)
Solution of a process of random genetic drift with a continuous model
.
Proc. Natl Acad. Sci. U.S.A.
41
,
144
150
147
Kimura
,
M.
(
1955
)
Stochastic processes and distribution of gene frequencies under natural selection
.
Cold Spring Harb. Symp. Quant. Biol.
20
,
33
148
Wonnapinij
,
P.
,
Chinnery
,
P.F.
and
Samuels
,
D.C.
(
2010
)
Previous estimates of mitochondrial DNA mutation level variance did not account for sampling error: comparing the mtDNA genetic bottleneck in mice and humans
.
Am. J. Hum. Genet.
86
,
540
550
149
Giannakis
,
K.
,
Broz
,
A.K.
,
Sloan
,
D.B.
and
Johnston
,
I.G.
(
2023
)
Avoiding misleading estimates using mtDNA heteroplasmy statistics to study bottleneck size and selection
.
G3
13
,
jkad068
150
Johnston
,
I.G.
(
2019
)
Varied mechanisms and models for the varying mitochondrial bottleneck
.
Front. Cell Dev. Biol.
7
,
294
151
Moran
,
P.A.P.
(
1958
)
Random processes in genetics
.
Math. Proc. Camb. Philos. Soc.
54
,
60
71
152
Wright
,
S.
(
1942
)
Statistical genetics and evolution
.
Bull. Am. Math. Soc.
48
,
223
153
Takahata
,
N.
(
1984
)
A model of extranuclear genomes and the substitution rate under within-generation selection
.
Gene. Res.
44
,
109
115
154
Bergstrom
,
C.T.
and
Pritchard
,
J.
(
1998
)
Germline Bottlenecks and the Evolutionary Maintenance of Mitochondrial Genomes
.
Genetics
149
,
2135
2146
155
Roze
,
D.
,
Rousset
,
F.
and
Michalakis
,
Y.
(
2005
)
Germline Bottlenecks, Biparental Inheritance and Selection on Mitochondrial Variants: A Two-Level Selection Model
.
Genetics
170
,
1385
1399
.
156
Walsh
,
J.B.
(
1992
)
Intracellular selection, conversion bias, and the expected substitution rate of organelle genes
.
Genetics
130
,
939
946
157
Aryaman
,
J.
,
Bowles
,
C.
,
Jones
,
N.S.
and
Johnston
,
I.G.
(
2019
)
Mitochondrial network state scales mtDNA genetic dynamics
.
Genetics
212
,
1429
1443
158
Hoitzing
,
H.
,
Gammage
,
P.A.
,
Haute
,
L.V.
,
Minczuk
,
M.
,
Johnston
,
I.G.
and
Jones
,
N.S.
(
2019
)
Energetic costs of cellular and therapeutic control of stochastic mitochondrial DNA populations
.
PLoS Comput. Biol.
15
,
e1007023
159
Insalata
,
F.
,
Hoitzing
,
H.
,
Aryaman
,
J.
and
Jones
,
N.S.
(
2022
)
Stochastic survival of the densest and mitochondrial DNA clonal expansion in aging
.
Proc. Natl Acad. Sci. U.S.A.
119
,
e2122073119
160
Johnston
,
I.G.
,
Burgstaller
,
J.P.
,
Havlicek
,
V.
,
Kolbe
,
T.
,
Rülicke
,
T.
,
Brem
,
G.
et al (
2015
)
Stochastic modelling, Bayesian inference, and new in vivo measurements elucidate the debated mtDNA bottleneck mechanism
.
eLife
4
,
e07464
161
Johnston
,
I.G.
and
Jones
,
N.S.
(
2016
)
Evolution of cell-to-cell variability in stochastic, controlled, heteroplasmic mtDNA populations
.
Am. J. Hum. Genet.
99
,
1150
1162
162
Mouli
,
P.K.
,
Twig
,
G.
and
Shirihai
,
O.S.
(
2009
)
Frequency and selectivity of mitochondrial fusion are key to its quality maintenance function
.
Biophys. J.
96
,
3509
3518
163
Poovathingal
,
S.K.
,
Gruber
,
J.
,
Halliwell
,
B.
and
Gunawan
,
R.
(
2009
)
Stochastic drift in mitochondrial DNA point mutations: a novel perspective ex silico
.
PLoS Comput. Biol.
5
,
e1000572
164
Tam
,
Z.Y.
,
Gruber
,
J.
,
Halliwell
,
B.
and
Gunawan
,
R.
(
2013
)
Mathematical modeling of the role of mitochondrial fusion and fission in mitochondrial DNA maintenance
.
PLoS ONE
8
,
e76230
165
Tam
,
Z.Y.
,
Gruber
,
J.
,
Halliwell
,
B.
and
Gunawan
,
R.
(
2015
)
Context-dependent role of mitochondrial fusion-fission in clonal expansion of mtDNA mutations
.
PLoS Comput. Biol.
11
,
e1004183
166
Agaronyan
,
K.
,
Morozov
,
Y.I.
,
Anikin
,
M.
and
Temiakov
,
D.
(
2015
)
Replication-transcription switch in human mitochondria
.
Science
347
,
548
551
167
Gitschlag
,
B.L.
,
Tate
,
A.T.
and
Patel
,
M.R.
(
2020
)
Nutrient status shapes selfish mitochondrial genome dynamics across different levels of selection
.
eLife
9
,
e56686
168
Ma
,
H.
and
O'Farrell
,
P.H.
(
2016
)
Selfish drive can trump function when animal mitochondrial genomes compete
.
Nat. Genet.
48
,
798
802
169
Røyrvik
,
E.C.
and
Johnston
,
I.G.
(
2020
)
MtDNA sequence features associated with ‘selfish genomes’ predict tissue-specific segregation and reversion
.
Nucleic Acids Res.
48
,
8290
8301
170
Aanen
,
D.K.
,
Spelbrink
,
J.N.
and
Beekman
,
M.
(
2014
)
What cost mitochondria? The maintenance of functional mitochondrial DNA within and across generations
.
Philos. Trans. R. Soc. B: Biol. Sci.
369
,
20130438
171
Ma
,
H.
,
Gutierrez
,
N.M.
,
Morey
,
R.
,
Van Dyken
,
C.
,
Kang
,
E.
,
Hayama
,
T.
et al (
2016
)
Incompatibility between nuclear and mitochondrial genomes contributes to an interspecies reproductive barrier
.
Cell Metab.
24
,
283
294
172
Moran
,
B.M.
,
Payne
,
C.Y.
,
Powell
,
D.L.
,
Iverson
,
E.N.
,
Donny
,
A.E.
,
Banerjee
,
S.M.
et al (
2024
)
A lethal mitonuclear incompatibility in complex I of natural hybrids
.
Nature
626
,
119
173
Burton
,
R.S.
(
2022
)
The role of mitonuclear incompatibilities in allopatric speciation
.
Cell Mol. Life Sci.
79
,
103
174
Sloan
,
D.B.
,
Havird
,
J.C.
and
Sharbrough
,
J.
(
2017
)
The on-again, off-again relationship between mitochondrial genomes and species boundaries
.
Mol. Ecol.
26
,
2212
2236
175
Telschow
,
A.
,
Gadau
,
J.
and
Kobayashi
,
Y.
(
2019
)
Genetic incompatibilities between mitochondria and nuclear genes: effect on gene flow and speciation
.
Front. Genet.
10
,
421056
176
Lane
,
N.
(
2011
)
Mitonuclear match: optimizing fitness and fertility over generations drives ageing within generations
.
Bioessays
33
,
860
869
177
Hadjivasiliou
,
Z.
,
Pomiankowski
,
A.
,
Seymour
,
R.M.
and
Lane
,
N.
(
2012
)
Selection for mitonuclear co-adaptation could favour the evolution of two sexes
.
Proc. R. Soc. B
279
,
1865
1872
178
Radzvilavicius
,
A.L.
and
Blackstone
,
N.W.
(
2015
)
Conflict and cooperation in eukaryogenesis: implications for the timing of endosymbiosis and the evolution of sex
.
J. R. Soc. Interface
12
,
20150584
179
Rand
,
D.M.
and
Mossman
,
J.A.
(
2020
)
Mitonuclear conflict and cooperation govern the integration of genotypes, phenotypes and environments
.
Phil. Trans. R. Soc. B
375
,
20190188
180
Gemmell
,
N.J.
,
Metcalf
,
V.J.
and
Allendorf
,
F.W.
(
2004
)
Mother's curse: the effect of mtDNA on individual fitness and population viability
.
Trends Ecol. Evol.
19
,
238
244
181
Munasinghe
,
M.
and
Ågren
,
J.A.
(
2023
)
When and why are mitochondria paternally inherited?
Curr. Opin. Genet. Dev.
80
,
102053
182
Connallon
,
T.
,
Camus
,
M.F.
,
Morrow
,
E.H.
and
Dowling
,
D.K.
(
2018
)
Coadaptation of mitochondrial and nuclear genes, and the cost of mother's curse
.
Proc. R. Soc. B: Biol. Sci.
285
,
20172257
183
Beekman
,
M.
,
Dowling
,
D.K.
and
Aanen
,
D.K.
(
2014
)
The costs of being male: are there sex-specific effects of uniparental mitochondrial inheritance?
Philos. Trans. R. Soc. B: Biol. Sci.
369
,
20130440
184
Budar
,
F.
,
Touzet
,
P.
and
De Paepe
,
R.
(
2003
)
The nucleo-mitochondrial conflict in cytoplasmic male sterilities revisited
.
Genetica
117
,
3
16
185
Christensen
,
A.C.
(
2014
)
Genes and junk in plant mitochondria—repair mechanisms and selection
.
Genome Biol. Evol.
6
,
1448
1453
186
Onishi
,
M.
,
Yamano
,
K.
,
Sato
,
M.
,
Matsuda
,
N.
and
Okamoto
,
K.
(
2021
)
Molecular mechanisms and physiological functions of mitophagy
.
EMBO J.
40
,
e104705
187
Youle
,
R.J.
and
Narendra
,
D.P.
(
2011
)
Mechanisms of mitophagy
.
Nat. Rev. Mol. Cell Biol.
12
,
9
14
188
Ni
,
H.-M.
,
Williams
,
J.A.
and
Ding
,
W.-X.
(
2015
)
Mitochondrial dynamics and mitochondrial quality control
.
Redox Biol.
4
,
6
13
189
Sedlackova
,
L.
and
Korolchuk
,
V.I.
(
2019
)
Mitochondrial quality control as a key determinant of cell survival
.
Biochim. Biophys. Acta Mol. Cell Res.
1866
,
575
587
190
Twig
,
G.
,
Hyde
,
B.
and
Shirihai
,
O.S.
(
2008
)
Mitochondrial fusion, fission and autophagy as a quality control axis: the bioenergetic view
.
Biochim. Biophys. Acta
1777
,
1092
1097
191
Edwards
,
D.M.
,
Røyrvik
,
E.C.
,
Chustecki
,
J.M.
,
Giannakis
,
K.
,
Glastad
,
R.C.
,
Radzvilavicius
,
A.L.
et al (
2021
)
Avoiding organelle mutational meltdown across eukaryotes with or without a germline bottleneck
.
PLoS Biol.
19
,
e3001153
192
Radzvilavicius
,
A.L.
,
Hadjivasiliou
,
Z.
,
Pomiankowski
,
A.
and
Lane
,
N.
(
2016
)
Selection for mitochondrial quality drives evolution of the germline
.
PLoS Biol.
14
,
e2000410
193
Chen
,
Z.
,
Zhao
,
N.
,
Li
,
S.
,
Grover
,
C.E.
,
Nie
,
H.
,
Wendel
,
J.F.
et al (
2017
)
Plant mitochondrial genome evolution and cytoplasmic male sterility
.
Crit. Rev. Plant Sci.
36
,
55
69
194
Kotrys
,
A.V.
,
Durham
,
T.J.
,
Guo
,
X.A.
,
Vantaku
,
V.R.
,
Parangi
,
S.
and
Mootha
,
V.K.
(
2024
)
Single-cell analysis reveals context-dependent, cell-level selection of mtDNA
.
Nature
629
,
458
466
195
Yu
,
L.
,
Boström
,
C.
,
Franzenburg
,
S.
,
Bayer
,
T.
,
Dagan
,
T.
and
Reusch
,
T.B.H.
(
2020
)
Somatic genetic drift and multilevel selection in a clonal seagrass
.
Nat. Ecol. Evol.
4
,
952
962
196
Lechuga-Vieco
,
A.V.
,
Latorre-Pellicer
,
A.
,
Johnston
,
I.G.
,
Prota
,
G.
,
Gileadi
,
U.
,
Justo-Méndez
,
R.
et al (
2020
)
Cell identity and nucleo-mitochondrial genetic context modulate OXPHOS performance and determine somatic heteroplasmy dynamics
.
Sci. Adv.
6
,
eaba5345
197
Colnaghi
,
M.
,
Pomiankowski
,
A.
and
Lane
,
N.
(
2021
)
The need for high-quality oocyte mitochondria at extreme ploidy dictates mammalian germline development
.
eLife
10
,
e69344
198
Gray
,
M.W.
,
Lukeš
,
J.
,
Archibald
,
J.M.
,
Keeling
,
P.J.
and
Doolittle
,
W.F.
(
2010
)
Irremediable complexity?
Science
330
,
920
921
199
Capps
,
G.J.
,
Samuels
,
D.C.
and
Chinnery
,
P.F.
(
2003
)
A model of the nuclear control of mitochondrial DNA replication
.
J. Theor. Biol.
221
,
565
583
200
Cree
,
L.M.
,
Samuels
,
D.C.
,
de Sousa Lopes
,
S.C.
,
Rajasimha
,
H.K.
,
Wonnapinij
,
P.
,
Mann
,
J.R.
et al (
2008
)
A reduction of mitochondrial DNA molecules during embryogenesis explains the rapid segregation of genotypes
.
Nat. Genet.
40
,
249
254
201
Wai
,
T.
,
Teoli
,
D.
and
Shoubridge
,
E.A.
(
2008
)
The mitochondrial DNA genetic bottleneck results from replication of a subpopulation of genomes
.
Nat. Genet.
40
,
1484
1488
202
Cao
,
L.
,
Shitara
,
H.
,
Horii
,
T.
,
Nagao
,
Y.
,
Imai
,
H.
,
Abe
,
K.
et al (
2007
)
The mitochondrial bottleneck occurs without reduction of mtDNA content in female mouse germ cells
.
Nat. Genet.
39
,
386
390
203
Huh
,
D.
and
Paulsson
,
J.
(
2011
)
Non-genetic heterogeneity from stochastic partitioning at cell division
.
Nat. Genet.
43
,
95
100
204
Jajoo
,
R.
,
Jung
,
Y.
,
Huh
,
D.
,
Viana
,
M.P.
,
Rafelski
,
S.M.
,
Springer
,
M.
et al (
2016
)
Accurate concentration control of mitochondria and nucleoids
.
Science
351
,
169
172
205
Johnston
,
I.G.
and
Jones
,
N.S.
(
2015
)
Closed-form stochastic solutions for non-equilibrium dynamics and inheritance of cellular components over many cell divisions
.
Proc. R. Soc. A: Math. Phys. Eng. Sci.
471
,
20150050
206
Khakhlova
,
O.
and
Bock
,
R.
(
2006
)
Elimination of deleterious mutations in plastid genomes by gene conversion
.
Plant J.
46
,
85
94
207
Lonsdale
,
D.M.
,
Brears
,
T.
,
Hodge
,
T.P.
,
Melville
,
S.E.
,
Rottmann
,
W.H.
,
Leaver
,
C.J.
et al (
1997
)
The plant mitochondrial genome: homologous recombination as a mechanism for generating heterogeneity
.
Philos. Trans. R. Soc. Lond. B. Biol. Sci.
319
,
149
163
208
Otten
,
A.B.C.
,
Sallevelt
,
S.C.E.H.
,
Carling
,
P.J.
,
Dreesen
,
J.C.F.M.
,
Drüsedau
,
M.
,
Spierts
,
S.
et al (
2018
)
Mutation-specific effects in germline transmission of pathogenic mtDNA variants
.
Hum. Reprod.
33
,
1331
1341
209
Wilson
,
I.J.
,
Carling
,
P.J.
,
Alston
,
C.L.
,
Floros
,
V.I.
,
Pyle
,
A.
,
Hudson
,
G.
et al (
2016
)
Mitochondrial DNA sequence characteristics modulate the size of the genetic bottleneck
.
Hum. Mol. Genet.
25
,
1031
1041
210
Glastad
,
R.C.
and
Johnston
,
I.G.
(
2023
)
Mitochondrial network structure controls cell-to-cell mtDNA variability generated by cell divisions
.
PLoS Comput. Biol.
19
,
e1010953
211
Radzvilavicius
,
A.L.
and
Johnston
,
I.G.
(
2022
)
Organelle bottlenecks facilitate evolvability by traversing heteroplasmic fitness valleys
.
Front. Genet.
13
,
974472
212
Cosmides
,
L.M.
and
Tooby
,
J.
(
1981
)
Cytoplasmic inheritance and intragenomic conflict
.
J. Theor. Biol.
89
,
83
129
213
Christie
,
J.R.
,
Schaerf
,
T.M.
and
Beekman
,
M.
(
2015
)
Selection against heteroplasmy explains the evolution of uniparental inheritance of mitochondria
.
PLoS Genet.
11
,
e1005112
214
Birky
,
C.W.
(
1995
)
Uniparental inheritance of mitochondrial and chloroplast genes: mechanisms and evolution
.
Proc. Natl Acad. Sci. U.S.A.
92
,
11331
11338
215
Radzvilavicius
,
A.L.
and
Johnston
,
I.G.
(
2020
)
Paternal leakage of organelles can improve adaptation to changing environments
.
bioRxiv
216
Havey
,
M.J
. (
2017
) Organellar genomes of the cucurbits. In
Genetics and Genomics of Cucurbitaceae
(
Grumet
,
R.
,
Katzir
,
N.
,
Garcia-Mas
,
J.
, eds), pp.
241
252
,
Springer International Publishing
,
Cham
217
Neiva
,
J.
,
Pearson
,
G.A.
,
Valero
,
M.
and
Serrão
,
E.A.
(
2010
)
Surfing the wave on a borrowed board: range expansion and spread of introgressed organellar genomes in the seaweed Fucus ceranoides L
.
Mol. Ecol.
19
,
4812
4822
218
Craven
,
L.
,
Tuppen
,
H.A.
,
Greggains
,
G.D.
,
Harbottle
,
S.J.
,
Murphy
,
J.L.
,
Cree
,
L.M.
et al (
2010
)
Pronuclear transfer in human embryos to prevent transmission of mitochondrial DNA disease
.
Nature
465
,
82
85
219
Wolf
,
D.P.
,
Mitalipov
,
N.
and
Mitalipov
,
S.
(
2015
)
Mitochondrial replacement therapy in reproductive medicine
.
Trends Mol. Med.
21
,
68
76
220
Gurdon
,
C.
,
Svab
,
Z.
,
Feng
,
Y.
,
Kumar
,
D.
and
Maliga
,
P.
(
2016
)
Cell-to-cell movement of mitochondria in plants
.
Proc. Natl Acad. Sci. U.S.A.
113
,
3395
3400
221
Berridge
,
M.V.
,
McConnell
,
M.J.
,
Grasso
,
C.
,
Bajzikova
,
M.
,
Kovarova
,
J.
and
Neuzil
,
J.
(
2016
)
Horizontal transfer of mitochondria between mammalian cells: beyond co-culture approaches
.
Curr. Opin. Genet. Dev.
38
,
75
82
222
Sinha
,
P.
,
Islam
,
M.N.
,
Bhattacharya
,
S.
and
Bhattacharya
,
J.
(
2016
)
Intercellular mitochondrial transfer: bioenergetic crosstalk between cells
.
Curr. Opin. Genet. Dev.
38
,
97
101
223
Jayaprakash
,
A.D.
,
Benson
,
E.K.
,
Gone
,
S.
,
Liang
,
R.
,
Shim
,
J.
,
Lambertini
,
L.
et al (
2015
)
Stable heteroplasmy at the single-cell level is facilitated by intercellular exchange of mtDNA
.
Nucleic Acids Res.
43
,
2177
2187
224
Spees
,
J.L.
,
Olson
,
S.D.
,
Whitney
,
M.J.
and
Prockop
,
D.J.
(
2006
)
Mitochondrial transfer between cells can rescue aerobic respiration
.
Proc. Natl Acad. Sci. U.S.A.
103
,
1283
1288
225
Tan
,
A.S.
,
Baty
,
J.W.
,
Dong
,
L.-F.
,
Bezawork-Geleta
,
A.
,
Endaya
,
B.
,
Goodwin
,
J.
et al (
2015
)
Mitochondrial genome acquisition restores respiratory function and tumorigenic potential of cancer cells without mitochondrial DNA
.
Cell Metab.
21
,
81
94
226
Allio
,
R.
,
Donega
,
S.
,
Galtier
,
N.
and
Nabholz
,
B.
(
2017
)
Large variation in the ratio of mitochondrial to nuclear mutation rate across animals: implications for genetic diversity and the use of mitochondrial DNA as a molecular marker
.
Mol. Biol. Evol.
34
,
2762
2772
227
Hagström
,
E.
,
Freyer
,
C.
,
Battersby
,
B.J.
,
Stewart
,
J.B.
and
Larsson
,
N.-G.
(
2013
)
No recombination of mtDNA after heteroplasmy for 50 generations in the mouse maternal germline
.
Nucleic Acids Res.
42
,
1111
1116
228
Ladoukakis
,
E.D.
and
Zouros
,
E.
(
2001
)
Direct evidence for homologous recombination in mussel (Mytilus galloprovincialis) mitochondrial DNA
.
Mol. Biol. Evol.
18
,
1168
1175
229
Guo
,
X.
,
Liu
,
S.
and
Liu
,
Y.
(
2006
)
Evidence for recombination of mitochondrial DNA in triploid crucian carp
.
Genetics
172
,
1745
1749
230
Klucnika
,
A.
,
Mu
,
P.
,
Jezek
,
J.
,
McCormack
,
M.
,
Di
,
Y.
,
Bradshaw
,
C.R.
et al (
2022
)
REC drives recombination to repair double-strand breaks in animal mtDNA
.
J. Cell Biol.
222
,
e202201137
231
Shokolenko
,
I.N.
,
Wilson
,
G.L.
and
Alexeyev
,
M.F.
(
2013
)
Persistent damage induces mitochondrial DNA degradation
.
DNA Repair
12
,
488
499
232
Shokolenko
,
I.
,
Venediktova
,
N.
,
Bochkareva
,
A.
,
Wilson
,
G.L.
and
Alexeyev
,
M.F.
(
2009
)
Oxidative stress induces degradation of mitochondrial DNA
.
Nucleic Acids Res.
37
,
2539
2548
233
Jenuth
,
J.P.
,
Peterson
,
A.C.
and
Shoubridge
,
E.A.
(
1997
)
Tissue-specific selection for different mtDNA genotypes in heteroplasmic mice
.
Nat. Genet.
16
,
93
95
234
Phillips
,
W.S.
,
Coleman-Hulbert
,
A.L.
,
Weiss
,
E.S.
,
Howe
,
D.K.
,
Ping
,
S.
,
Wernick
,
R.I.
et al (
2015
)
Selfish mitochondrial DNA proliferates and diversifies in small, but not large, experimental populations of Caenorhabditis briggsae
.
Genome Biol. Evol.
7
,
2023
2037
235
Moreno-Loshuertos
,
R.
,
Acín-Pérez
,
R.
,
Fernández-Silva
,
P.
,
Movilla
,
N.
,
Pérez-Martos
,
A.
,
de Cordoba
,
S.R.
et al (
2006
)
Differences in reactive oxygen species production explain the phenotypes associated with common mouse mitochondrial DNA variants
.
Nat. Genet.
38
,
1261
1268
236
Gómez-Durán
,
A.
,
Pacheu-Grau
,
D.
,
López-Gallardo
,
E.
,
Díez-Sánchez
,
C.
,
Montoya
,
J.
,
López-Pérez
,
M.J.
et al (
2010
)
Unmasking the causes of multifactorial disorders: OXPHOS differences between mitochondrial haplogroups
.
Hum. Mol. Genet.
19
,
3343
3353
237
Battersby
,
B.J.
,
Loredo-Osti
,
J.C.
and
Shoubridge
,
E.A.
(
2003
)
Nuclear genetic control of mitochondrial DNA segregation
.
Nat. Genet.
33
,
183
186
238
Jokinen
,
R.
,
Marttinen
,
P.
,
Sandell
,
H.K.
,
Manninen
,
T.
,
Teerenhovi
,
H.
,
Wai
,
T.
et al (
2010
)
Gimap3 regulates tissue-specific mitochondrial DNA segregation
.
PLoS Genet.
6
,
e1001161
239
Jokinen
,
R.
,
Marttinen
,
P.
,
Stewart
,
J.B.
,
Neil Dear
,
T.
and
Battersby
,
B.J.
(
2016
)
Tissue-specific modulation of mitochondrial DNA segregation by a defect in mitochondrial division
.
Hum. Mol. Genet.
25
,
706
714
240
Tsyba
,
N.
,
Feng
,
G.
,
Grub
,
L.K.
,
Held
,
J.P.
,
Strozak
,
A.M.
,
Burkewitz
,
K.
et al (
2023
)
Tissue-specific heteroplasmy segregation is accompanied by a sharp mtDNA decline in Caenorhabditis elegans soma
.
Iscience
26
,
106349
241
Li
,
M.
,
Schröder
,
R.
,
Ni
,
S.
,
Madea
,
B.
and
Stoneking
,
M.
(
2015
)
Extensive tissue-related and allele-related mtDNA heteroplasmy suggests positive selection for somatic mutations
.
Proc. Natl Acad. Sci. U.S.A.
112
,
2491
2496
242
Pyle
,
A.
,
Taylor
,
R.W.
,
Durham
,
S.E.
,
Deschauer
,
M.
,
Schaefer
,
A.M.
,
Samuels
,
D.C.
et al (
2007
)
Depletion of mitochondrial DNA in leucocytes harbouring the 3243A→ G mtDNA mutation
.
J. Med. Genet.
44
,
69
74
243
Gupta
,
R.
,
Kanai
,
M.
,
Durham
,
T.J.
,
Tsuo
,
K.
,
McCoy
,
J.G.
,
Kotrys
,
A.V.
et al (
2023
)
Nuclear genetic control of mtDNA copy number and heteroplasmy in humans
.
Nature
620
,
839
848
244
Chiaratti
,
M.R.
and
Chinnery
,
P.F.
(
2022
)
Modulating mitochondrial DNA mutations: factors shaping heteroplasmy in the germ line and somatic cells
.
Pharmacol. Res.
185
,
106466
245
Scott
,
L.
,
Karuppagounder
,
S.S.
,
Neifert
,
S.
,
Kang
,
B.G.
,
Wang
,
H.
,
Dawson
,
V.L.
et al (
2022
)
The absence of parkin does not promote dopamine or mitochondrial dysfunction in PolgAD257A/D257A mitochondrial mutator mice
.
J. Neurosci.
42
,
9263
9277
246
Pickrell
,
A.M.
,
Huang
,
C.-H.
,
Kennedy
,
S.R.
,
Ordureau
,
A.
,
Sideris
,
D.P.
,
Hoekstra
,
J.G.
et al (
2015
)
Endogenous parkin preserves dopaminergic substantia nigral neurons following mitochondrial DNA mutagenic stress
.
Neuron
87
,
371
381
247
Burgstaller
,
J.P.
,
Kolbe
,
T.
,
Havlicek
,
V.
,
Hembach
,
S.
,
Poulton
,
J.
,
Piálek
,
J.
et al (
2018
)
Large-scale genetic analysis reveals mammalian mtDNA heteroplasmy dynamics and variance increase through lifetimes and generations
.
Nat. Commun.
9
,
2488
248
Burr
,
S.P.
,
Pezet
,
M.
and
Chinnery
,
P.F.
(
2018
)
Mitochondrial DNA heteroplasmy and purifying selection in the mammalian female germ line
.
Dev. Growth Differ.
60
,
21
32
249
Fan
,
W.
,
Waymire
,
K.G.
,
Narula
,
N.
,
Li
,
P.
,
Rocher
,
C.
,
Coskun
,
P.E.
et al (
2008
)
A mouse model of mitochondrial disease reveals germline selection against severe mtDNA mutations
.
Science
319
,
958
962
250
Stewart
,
J.B.
,
Freyer
,
C.
,
Elson
,
J.L.
,
Wredenberg
,
A.
,
Cansu
,
Z.
,
Trifunovic
,
A.
et al (
2008
)
Strong purifying selection in transmission of mammalian mitochondrial DNA
.
PLoS Biol.
6
,
e10
251
Lieber
,
T.
,
Jeedigunta
,
S.P.
,
Palozzi
,
J.M.
,
Lehmann
,
R.
and
Hurd
,
T.R.
(
2019
)
Mitochondrial fragmentation drives selective removal of deleterious mtDNA in the germline
.
Nature
570
,
380
384
252
Palozzi
,
J.M.
,
Jeedigunta
,
S.P.
,
Minenkova
,
A.V.
,
Monteiro
,
V.L.
,
Thompson
,
Z.S.
,
Lieber
,
T.
et al (
2022
)
Mitochondrial DNA quality control in the female germline requires a unique programmed mitophagy
.
Cell Metab.
34
,
1809
1823
253
Wei
,
W.
,
Tuna
,
S.
,
Keogh
,
M.J.
,
Smith
,
K.R.
,
Aitman
,
T.J.
,
Beales
,
P.L.
et al (
2019
)
Germline selection shapes human mitochondrial DNA diversity
.
Science
364
,
eaau6520
254
Mishmar
,
D.
,
Ruiz-Pesini
,
E.
,
Golik
,
P.
,
Macaulay
,
V.
,
Clark
,
A.G.
,
Hosseini
,
S.
et al (
2003
)
Natural selection shaped regional mtDNA variation in humans
.
Proc. Natl Acad. Sci. U.S.A.
100
,
171
176
255
Ruiz-Pesini
,
E.
,
Mishmar
,
D.
,
Brandon
,
M.
,
Procaccio
,
V.
and
Wallace
,
D.C.
(
2004
)
Effects of purifying and adaptive selection on regional variation in human mtDNA
.
Science
303
,
223
226
256
Nabholz
,
B.
,
Ellegren
,
H.
and
Wolf
,
J.B.W.
(
2013
)
High levels of gene expression explain the strong evolutionary constraint of mitochondrial protein-coding genes
.
Mol. Biol. Evol.
30
,
272
284
257
Shtolz
,
N.
and
Mishmar
,
D.
(
2023
)
The metazoan landscape of mitochondrial DNA gene order and content is shaped by selection and affects mitochondrial transcription
.
Commun. Biol.
6
,
1
15
258
Luo
,
Y.
,
Yang
,
X.
and
Gao
,
Y.
(
2013
)
Mitochondrial DNA response to high altitude: a new perspective on high-altitude adaptation
.
Mitochondrial DNA
24
,
313
319
259
Graham
,
A.M.
,
Lavretsky
,
P.
,
Wilson
,
R.E.
and
McCracken
,
K.G.
(
2024
)
High-altitude adaptation is accompanied by strong signatures of purifying selection in the mitochondrial genomes of three Andean waterfowl
.
PLoS ONE
19
,
e0294842
260
Consuegra
,
S.
,
John
,
E.
,
Verspoor
,
E.
and
de Leaniz
,
C.G.
(
2015
)
Patterns of natural selection acting on the mitochondrial genome of a locally adapted fish species
.
Genet. Sel. Evol.
47
,
58
261
Cam
,
S.L.
,
Mortz
,
M.
and
Blier
,
P.
(
2024
)
Longevity and environmental temperature modulate mitochondrial DNA evolution in fishes
.
bioRxiv
2024.03.07.583929
262
Zhang
,
H.
,
Burr
,
S.P.
and
Chinnery
,
P.F.
(
2018
)
The mitochondrial DNA genetic bottleneck: inheritance and beyond
.
Essays Biochem.
62
,
225
234
263
Jokinen
,
R.
and
Battersby
,
B.J.
(
2013
)
Insight into mammalian mitochondrial DNA segregation
.
Ann. Med.
45
,
149
155
264
Jenuth
,
J.P.
,
Peterson
,
A.C.
,
Fu
,
K.
and
Shoubridge
,
E.A.
(
1996
)
Random genetic drift in the female germline explains the rapid segregation of mammalian mitochondrial DNA
.
Nat. Genet.
14
,
146
151
265
Wolff
,
J.N.
,
White
,
D.J.
,
Woodhams
,
M.
,
White
,
H.E.
and
Gemmell
,
N.J.
(
2011
)
The strength and timing of the mitochondrial bottleneck in salmon suggests a conserved mechanism in vertebrates
.
PLoS ONE
6
,
e20522
266
Rand
,
D.M.
and
Harrison
,
R.G.
(
1986
)
Mitochondrial DNA transmission genetics in crickets
.
Genetics
114
,
955
970
267
Solignac
,
M.
,
Génermont
,
J.
,
Monnerot
,
M.
and
Mounolou
,
J.-C.
(
1984
)
Genetics of mitochondria in Drosophila: mtDNA inheritance in heteroplasmic strains of D. mauritiana
.
Mol. Gen. Genet.
197
,
183
188
268
Li
,
M.
,
Rothwell
,
R.
,
Vermaat
,
M.
,
Wachsmuth
,
M.
,
Schröder
,
R.
,
Laros
,
J.F.
et al (
2016
)
Transmission of human mtDNA heteroplasmy in the genome of the Netherlands families: support for a variable-size bottleneck
.
Genome Res.
26
,
417
426
269
Otten
,
A.B.C.
,
Theunissen
,
T.E.J.
,
Derhaag
,
J.G.
,
Lambrichs
,
E.H.
,
Boesten
,
I.B.W.
,
Winandy
,
M.
et al (
2016
)
Differences in strength and timing of the mtDNA bottleneck between zebrafish germline and non-germline cells
.
Cell Rep.
16
,
622
630
270
Hauswirth
,
W.W.
and
Laipis
,
P.J.
(
1982
)
Mitochondrial DNA polymorphism in a maternal lineage of Holstein cows
.
Proc. Natl Acad. Sci. U.S.A.
79
,
4686
4690
271
Ashley
,
M.V.
,
Laipis
,
P.J.
and
Hauswirth
,
W.W.
(
1989
)
Rapid segregation of heteroplasmic bovine mitodiondria
.
Nucleic Acids Res.
17
,
7325
7331
272
Rebolledo-Jaramillo
,
B.
,
Su
,
M.S.-W.
,
Stoler
,
N.
,
McElhoe
,
J.A.
,
Dickins
,
B.
,
Blankenberg
,
D.
et al (
2014
)
Maternal age effect and severe germ-line bottleneck in the inheritance of human mitochondrial DNA
.
Proc. Natl Acad. Sci. U.S.A.
111
,
15474
15479
273
Barrett
,
A.
,
Arbeithuber
,
B.
,
Zaidi
,
A.
,
Wilton
,
P.
,
Paul
,
I.M.
,
Nielsen
,
R.
et al (
2020
)
Pronounced somatic bottleneck in mitochondrial DNA of human hair
.
Phil. Trans. R. Soc. B
375
,
20190175
274
Wilton
,
P.R.
,
Zaidi
,
A.
,
Makova
,
K.
and
Nielsen
,
R.
(
2018
)
A population phylogenetic view of mitochondrial heteroplasmy
.
Genetics
208
,
1261
1274
275
Wachtmann
,
D.
and
Stockem
,
W.
(
1992
)
Significance of the cytoskeleton for cytoplasmic organization and cell organelle dynamics in epithelial cells of fresh-water sponges
.
Protoplasma
169
,
107
119
276
Luo
,
S.
,
Valencia
,
C.A.
,
Zhang
,
J.
,
Lee
,
N.-C.
,
Slone
,
J.
,
Gui
,
B.
et al (
2018
)
Biparental inheritance of mitochondrial DNA in humans
.
Proc. Natl Acad. Sci. U.S.A.
115
,
13039
13044
277
Lutz-Bonengel
,
S.
and
Parson
,
W.
(
2019
)
No further evidence for paternal leakage of mitochondrial DNA in humans yet
.
Proc. Natl Acad. Sci. U.S.A.
116
,
1821
1822
278
Lee
,
W.
,
Zamudio-Ochoa
,
A.
,
Buchel
,
G.
,
Podlesniy
,
P.
,
Marti Gutierrez
,
N.
,
Puigròs
,
M.
et al (
2023
)
Molecular basis for maternal inheritance of human mitochondrial DNA
.
Nat. Genet.
55
,
1632
1639
279
Meusel
,
M.S.
and
Moritz
,
R.F.
(
1993
)
Transfer of paternal mitochondrial DNA during fertilization of honeybee (Apis mellifera L.) eggs
.
Curr. Genet.
24
,
539
543
280
Passamonti
,
M.
and
Ghiselli
,
F.
(
2009
)
Doubly uniparental inheritance: two mitochondrial genomes, one precious model for organelle DNA inheritance and evolution
.
DNA Cell Biol.
28
,
79
89
281
Zouros
,
E.
,
Oberhauser Ball
,
A.
,
Saavedra
,
C.
and
Freeman
,
K.R.
(
1994
)
An unusual type of mitochondrial DNA inheritance in the blue mussel Mytilus
.
Proc. Natl Acad. Sci. U.S.A.
91
,
7463
7467
282
Zouros
,
E.
,
Freeman
,
K.R.
,
Ball
,
A.O.
and
Pogson
,
G.H.
(
1992
)
Direct evidence for extensive paternal mitochondrial DNA inheritance in the marine mussel Mytilus
.
Nature
359
,
412
414
283
Breton
,
S.
,
Beaupré
,
H.D.
,
Stewart
,
D.T.
,
Hoeh
,
W.R.
and
Blier
,
P.U.
(
2007
)
The unusual system of doubly uniparental inheritance of mtDNA: isn't one enough?
Trends Genet.
23
,
465
474
284
Mower
,
J.P.
,
Touzet
,
P.
,
Gummow
,
J.S.
,
Delph
,
L.F.
and
Palmer
,
J.D.
(
2007
)
Extensive variation in synonymous substitution rates in mitochondrial genes of seed plants
.
BMC Evol. Biol.
7
,
135
285
Arrieta-Montiel
,
M.P.
and
Mackenzie
,
S.A.
(
2011
) Plant mitochondrial genomes and recombination. In
Plant Mitochondria
(
Kempken
,
F.
, ed.), pp.
65
82
,
Springer
,
New York, NY
286
Maréchal
,
A.
and
Brisson
,
N.
(
2010
)
Recombination and the maintenance of plant organelle genome stability
.
New Phytol.
186
,
299
317
287
Woloszynska
,
M.
(
2010
)
Heteroplasmy and stoichiometric complexity of plant mitochondrial genomes—though this be madness, yet there's method in't
.
J. Exp. Bot.
61
,
657
671
288
Wu
,
Z.
,
Waneka
,
G.
,
Broz
,
A.K.
,
King
,
C.R.
and
Sloan
,
D.B.
(
2020
)
MSH1 is required for maintenance of the low mutation rates in plant mitochondrial and plastid genomes
.
Proc. Natl Acad. Sci. U.S.A.
117
,
16448
16455
289
Davila
,
J.I.
,
Arrieta-Montiel
,
M.P.
,
Wamboldt
,
Y.
,
Cao
,
J.
,
Hagmann
,
J.
,
Shedge
,
V.
et al (
2011
)
Double-strand break repair processes drive evolution of the mitochondrial genome in Arabidopsis
.
BMC Biol.
9
,
64
290
Miller-Messmer
,
M.
,
Kühn
,
K.
,
Bichara
,
M.
,
Le Ret
,
M.
,
Imbault
,
P.
and
Gualberto
,
J.M.
(
2012
)
RecA-dependent DNA repair results in increased heteroplasmy of the Arabidopsis mitochondrial genome
.
Plant Physiol.
159
,
211
226
291
Broz
,
A.K.
,
Keene
,
A.
,
Fernandes Gyorfy
,
M.
,
Hodous
,
M.
,
Johnston
,
I.G.
and
Sloan
,
D.B.
(
2022
)
Sorting of mitochondrial and plastid heteroplasmy in Arabidopsis is extremely rapid and depends on MSH1 activity
.
Proc. Natl Acad. Sci. U.S.A.
119
,
e2206973119
292
Broz
,
A.K.
,
Sloan
,
D.B.
and
Johnston
,
I.G.
(
2024
)
Stochastic organelle genome segregation through Arabidopsis development and reproduction
.
New Phytol.
241
,
896
910
293
Kitazaki
,
K.
and
Kubo
,
T.
(
2010
)
Cost of having the largest mitochondrial genome: evolutionary mechanism of plant mitochondrial genome
.
J. Bot.
2010
,
620137
294
Sloan
,
D.B.
and
Wu
,
Z.
(
2014
)
History of plastid DNA insertions reveals weak deletion and at mutation biases in angiosperm mitochondrial genomes
.
Genome Biol. Evol.
6
,
3210
3221
295
Christensen
,
A.C.
(
2013
)
Plant mitochondrial genome evolution can be explained by DNA repair mechanisms
.
Genome Biol. Evol.
5
,
1079
1086
296
Wu
,
Z.-Q.
,
Liao
,
X.-Z.
,
Zhang
,
X.-N.
,
Tembrock
,
L.R.
and
Broz
,
A.
(
2022
)
Genomic architectural variation of plant mitochondria—a review of multichromosomal structuring
.
J. Syst. Evol.
60
,
160
168
297
Chevigny
,
N.
,
Schatz-Daas
,
D.
,
Lotfi
,
F.
and
Gualberto
,
J.M.
(
2020
)
DNA repair and the stability of the plant mitochondrial genome
.
Int. J. Mol. Sci.
21
,
328
298
Arimura
,
S.
(
2018
)
Fission and fusion of plant mitochondria, and genome maintenance
.
Plant Physiol.
176
,
152
161