Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


During cell division, it is critical to properly partition functional sets of organelles to each daughter cell. The partitioning of mitochondria shares some common features with that of other organelles, particularly in the use of interactions with cytoskeletal elements to facilitate delivery to the daughter cells. However, mitochondria have unique features - including their own genome and a maternal mode of germline transmission - that place additional demands on this process. Consequently, mechanisms have evolved to regulate mitochondrial segregation during cell division, oogenesis, fertilization and tissue development, as well as to ensure the integrity of these organelles and their DNA, including fusion-fission dynamics, organelle transport, mitophagy and genetic selection of functional genomes. Defects in these processes can lead to cell and tissue pathologies.

Free full text 


Logo of nihpaLink to Publisher's site
Nat Rev Mol Cell Biol. Author manuscript; available in PMC 2015 Apr 1.
Published in final edited form as:
PMCID: PMC4250044
NIHMSID: NIHMS644044
PMID: 25237825

Mitochondrial dynamics and inheritance during cell division, development and disease

Preface

During cell division, it is critical to properly partition functional sets of organelles to each daughter cell. The partitioning of mitochondria shares some common features with other organelles, particularly in their interactions with cytoskeletal elements to facilitate delivery to the daughter cells. However, mitochondria have unique features – including their own genome and a maternal mode of germline transmission – that place additional demands on this process. We discuss the mechanisms regulating mitochondrial segregation during cell division, oogenesis, fertilization and tissue development. The mechanisms that ensure the integrity of these organelles and their DNA include fusion-fission dynamics, organelle transport, mitophagy, and genetic selection of functional genomes. Defects in these processes can lead to cell and tissue pathologies.

Introduction

Organelles are a distinguishing feature of eukaryotic cells. During somatic cell proliferation, they must segregate properly to daughter cells, and during germline inheritance, a highly functional population of organelles must be transmitted to the offspring. One such organelle is the mitochondrion1, which is best known for its critical function in energy production via oxidative phosphorylation (OXPHOS). The OXPHOS pathway generates many more adenosine triphosphate (ATP) molecules per glucose molecule than the glycolysis pathway. Mitochondria also have important roles in other types of metabolism, in regulating intracellular calcium concentration and signalling in neurons, in assembly of iron-sulfur clusters that are important for oxidation-reduction reactions2, in apoptosis3, and in innate immunity4.

According to the endosymbiotic theory, mitochondria are descendants of ancient bacteria that entered into a symbiotic relationship with primitive host cells5. Mitochondria retain several characteristics of their putative bacterial ancestors: a double-membrane, a proteome similar to that of α-proteobacteria, and the ability to synthesize ATP via a proton gradient created across its inner membrane (Box 1). In addition to these prokaryotic characteristics, mitochondria also undergo membrane remodelling through cycles of fusion (two mitochondria joining to form a single mitochondrion) and division (or fission; a single mitochondrion dividing into two)6 (Box 2). The balance of fusion and fission controls mitochondrial structure, and depending on the cell type, the numerous separate mitochondria in the cell can shift to form a single, interconnected membranous structure.

Box 1

The mitochondrial genome and oxidative phosphorylation

Mitochondria contain a residual genome (mitochondrial DNA; mtDNA) that is critical for their function in oxidative phosphorylation (OXPHOS). In humans, the mtDNA is ~16.6 kilobases in length and contains 37 genes encoding 13 polypeptides, 2 ribosomal RNAs (rRNAs), and 22 transfer RNAs (tRNAs; see the figure, part a) The D-loop (yellow) is a regulatory sequence that controls mtDNA replication and transcription. The 13 polypeptides are distributed among the OXPHOS protein complexes I, III, IV, and V and are essential for OXPHOS (cellular respiration) activity (see the figure, part b). The specific mtDNA encoded subunits are listed under each respiratory chain complex. These complexes also have multiple subunits encoded by the nuclear genome. Complex II (gray) is entirely encoded by the nuclear genome. The 2 rRNAs and 22 tRNAs encoded by the mtDNA are essential for the mitochondrial translational machinery that is used to generate the 13 polypeptides, and because of this all 37 mtDNA genes are critical for OXPHOS. The remaining proteins (>1000) in the mitochondrial proteome are encoded by the nuclear genome, synthesized in the cytosol, and imported into the mitochondria.

Mitochondria are double-membrane organelles, consisting of an outer membrane surrounding an inner membrane of greater surface area. The space between the two membranes is termed the intermembrane space. Because of its greater surface area, the inner membrane folds back on itself, creating compartments termed cristae. The respiratory chain complexes are concentrated in the cristae membranes. Complexes I, III, and IV pump protons from the matrix (the compartment enclosed by the inner membrane) to the intermembrane space, generating an electro-chemical gradient across the inner membrane that is used to power ATP production by Complex V (ATP synthase). The ATP-ADP translocase (green) exports the ATP into the intermembrane space, where it freely diffuses out of the outer membrane into the cytosol. The mtDNA is packaged into nucleoid structures (not shown) located in the matrix.

An external file that holds a picture, illustration, etc.
Object name is nihms644044u1.jpg

Box 2

Mitochondrial fusion and fission in mammalian cells

Continual cycles of fusion and fission result in the intermixing of the mitochondrial population in the cell6. These opposing processes also determine mitochondrial morphology. Increased fusion or reduced fission promotes the formation of elongated mitochondrial networks, whereas increased fission or reduced fusion causes mitochondrial fragmentation (Figure 1A). The key factors in both fusion and fission are large GTP hydrolyzing enzymes of the dynamin superfamily. Mitochondrial fusion consists of two sequential steps: First the outer membrane (OM) undergoes fusion, followed by inner membrane (IM) fusion (see figure, part a). Outer membrane fusion is mediated by the mitofusin proteins Mfn1 and Mfn2, which are dynamin-related GTPases at the outer membrane. Inner membrane fusion is mediated by the dynamin-related protein Opa1. Opa1 exists in two forms: a long form integral in the inner membrane, and a short form (not shown) targeted to the intermembrane space (IMS). Whereas cells lacking both mitofusins have no outer membrane fusion, cells lacking Opa1 do undergo mitochondrial outer membrane fusion but cannot progress and undergo inner membrane fusion107, 108.

An external file that holds a picture, illustration, etc.
Object name is nihms644044f1.jpg
Cellular mechanisms involved in mitochondrial segregation, transport, and degradation

(A) The control of mitochondrial morphology by fusion and fission. Cells with increased fission or reduced fusion have small mitochondria (left), whereas cells with increased fusion or reduced fission have elongated mitochondrial tubules (right). Mitochondria were visualized in mouse embryonic fibroblasts by expression of mitochondrially targeted DsRed. Scale bar: 10 μm. (B) Cell cycle mitochondrial dynamics. Mitochondrial morphology is coordinated with the cell cycle and promotes equal segregation of mitochondria during cell division. At the G1 stage of the cell cycle, mitochondria have a variety of morphologies. During the G1-S transition, mitochondria fuse and elongate, presumably in preparation for the high metabolic demand associated with genome replication. In contrast, at the G2 and M phases, mitochondria undergo fission and form numerous individual organelles that are spatially distributed throughout the soma. This enables the equal distribution of mitochondria to each of the daughter cells during mitosis. Upon re-entry to the G1 phase, mitochondria regain some of their elongated structure. (C) Mitochondrial transport. Mitochondria can associate with cytoskeletal filaments and be transported along them by molecular motors. Mitochondria in mammalian cells are transported mostly on microtubule filaments by kinesin and dynein motors14. Kinesins transport mitochondria towards the plus end of microtubules, whereas dynein transport them towards the minus end (left). Milton acts as an adaptor to link each motor to Miro, which is on the mitochondrial surface. In the presence of high calcium concentrations (right), kinesin is released from the Milton/Miro complex and instead binds to syntaphilin (SNPH). SNPH inhibits the ATPase activity of kinesin and results in immobilization of mitochondria. (D) Mitophagy pathways in mammalian cells: Parkin (right) is recruited to dysfunctional mitochondria (purple) with reduced membrane potential (ΔΨ)48. Parkin is an E3 ubiquitin ligase and it polyubiquitinates a large number of mitochondrial outer membrane proteins, resulting in their degradation by the ubiquitin-proteasome system (UPS)50, 51, which is necessary for the subsequent removal of mitochondria from the cytosol50. The AAA ATPase p97 facilitates the degradation of outer membrane proteins by the 26S proteasome. In erythroid cells (left), removal of mitochondria requires the outer membrane protein Nix, which interacts with the LC3 homolog GABARAP-L1. In hypoxia-induced mitophagy (bottom), the PGAM5 phosphatase dephosphorylates the LC3 interacting region of FUNDC1. Dephosphorylated FUNDC1 then recruits LC3 to bring autophagosomes to the mitochondria. (E) Mitophagy in yeast: In budding yeast, mitophagy is induced during post-log growth in non-fermentable media. Such conditions induce the expression of ATG32, which acts as a receptor to recruit the autophagy machinery65, 66. ATG32 binds to ATG11, an autophagy adaptor that binds to ATG8 (the yeast orthologue of LC3), thereby linking the mitochondrion to autophagosomes. ATG11 also binds to Dnm167 (see Boxes 2 and 4), a central component of the mitochondrial fission machinery, to promote mitochondrial fission during mitophagy.

The opposing process of mitochondrial fission (see figure, part b) requires the recruitment of the dynamin-related protein Drp1 from the cytosol to the outer mitochondrial membrane. Assembly of Drp1 on the mitochondrial surface causes constriction of the mitochondria and eventual division of the organelle into two separate entities. Four Drp1 receptors exist in mammals and these are located on the outer mitochondrial membrane: Fis1, Mff, Mid49 and Mid51. The latter three play the predominant role in fission. Despite its importance in yeast, Fis1 appears to play a minor role in mitochondrial fission in mammals109, 110.

An external file that holds a picture, illustration, etc.
Object name is nihms644044u2.jpg

Mitochondria have two features that constrain their segregation. First, they have a genome (mitochondrial DNA or “mtDNA”) that encodes critical bioenergetic functions (Box 1). Each mitochondrion contains one or more mtDNA molecules that are organized into mtDNA-protein complexes termed nucleoids. In proliferating cells, the partitioning of mitochondria to daughter cells must also result in proper distribution of mtDNA. Second, mitochondria in mammals are transmitted to subsequent generations exclusively through the maternal lineage. Because of this uniparental inheritance pattern, mtDNA lacks the benefits of recombination arising from sexual reproduction. However, mechanisms have evolved to reduce the chance of transmitting pathogenic mtDNA mutations to offspring. In some cases, these mechanisms fail, and maternally inherited disorders termed mitochondrial encephalomyopathies arise when a significant load of mtDNA mutations are passed to the offspring. These diseases are characterized by reduced mitochondrial function, with clinical signs being most pronounced in tissues with high energy consumption.

With mitochondria functioning in diverse cellular processes and their dysfunction implicated in several diseases, an understanding of mitochondrial dynamics, segregation and germline inheritance is of great interest. In this Review we first outline the molecular mechanisms that ensure proper partitioning of mitochondria during cell division: the fusion/fission cycle, mitochondrial interactions with cytoskeletal elements and the endoplasmic reticulum (ER), and mitophagy. In the second part of the Review we discuss the specific regulation of mtDNA inheritance and its uniparental transmission through the maternal lineage, including the evidence for genetic bottlenecks during oogenesis and embryonic development, and the mechanisms preventing transmission of paternal mitochondria to offspring. Throughout, our main focus is on mammalian mitochondria, but studies performed in yeast, worms and flies are discussed in cases where they have provided substantial insight to our understanding of those issues.

Mitochondrial dynamics and segregation

In discussing mitochondrial inheritance, “proper partitioning” of the mitochondria implies that mitochondria and mtDNA are distributed equally to daughter cells (in the case of symmetric cell division), and that the inherited mitochondria are highly functional. In mammalian cells that divide symmetrically, segregation could, in principle, be a largely passive process, as long as mechanisms exist to ensure a homogenous and evenly dispersed spatial distribution of organelles throughout the cell soma (Figure 1A). As discussed below, faithful segregation is facilitated by the induction of mitochondrial fission during the G2 and mitosis phases of the cell cycle (Fig. 1B) and by the interaction of mitochondria with the cytoskeleton and endoplasmic reticulum (ER) elements that are spread throughout the cell soma. Active mechanisms that deliver mitochondria to daughter cells are not known in mammalian cells; however, this may simply reflect a lack of suitable model systems to study this process. In particular, little is known about mitochondrial segregation in mammalian cells that divide asymmetrically. In budding yeast, which undergo asymmetric cell division, mitochondrial segregation is driven by mechanisms that actively deliver a subset of mitochondria in the daughter cell, while retaining another subset in the mother cell (Box 3).

Box 3

Active segregation of mitochondria in yeast

The budding yeast Saccharomyces cerevisiae is an attractive model system to study the mechanisms controlling mitochondrial cytoplasmic distribution and segregation. In this organism, cell division is a highly polarized process whereby a daughter bud forms at a distinct site on the mother cell surface. Mitochondria and other organelles are rapidly transported to the emerging bud, which eventually divides off the daughter cell. In fact, the amount of mitochondria increases in direct relation to the steadily increasing volume of the bud, until a specific ratio of mitochondrial to bud volume is met111. This suggests that the mitochondrial content in the bud is carefully monitored to ensure that a sufficient amount of mitochondria is distributed to the daughter cell, even if it compromises the mitochondrial content of the aging mother cell111. Yeast genetic screens tracking the delivery of mitochondria during budding have identified factors that are important to mitochondrial transport to the growing bud112.

The actin cytoskeleton actively transports mitochondria into the bud (see figure). During bud emergence, formin proteins (not shown) are deposited at the future bud site and direct polymerization of actin filaments, linking the mother cell to the bud113. These mother-to-bud actin filaments serve as ‘tracks’ to deliver mitochondria and other organelles to the daughter cell114116. Because Arp2/3 mutants have mitochondrial motility defects, it has been proposed that Arp2/3 nucleates a burst of actin polymerization on the mitochondrial surface that drives transport along the mother-to-bud actin filaments117, 118. This model has similarities to the transport mechanism used by some mobile intracellular bacteria119. However, subsequent studies have supported an alternative model in which Myo2 is a motor protein with a direct role in transporting mitochondria as cargo into the bud120122 (see figure). Myo2 is utilized for transport of most membrane-bound organelles in budding yeast123128. The Myo2 receptor on mitochondria remains to be defined, although candidates include Mmr1 and Ypt11 that physically interact with Myo2; indeed, Δmmr1 Δypt11 double mutants fail to transport mitochondria into the bud121, 129, 130. Myosin molecules may also play a role in organelle delivery during other asymmetric cell divisions; for instance, Myo6 has been implicated in active mitochondrial delivery during spermatogenesis in C.elegans131.

The delivery of mitochondria to the daughter bud is balanced by retention of mitochondria in the mother cell. Recent genetic and biochemical studies indicate that the cortical protein Num1 serves as a plasma membrane anchor that interacts with mitochondria and retains them in the mother cell132, 133 (see figure).

An external file that holds a picture, illustration, etc.
Object name is nihms644044u3.jpg

Microtubule-based mitochondrial motility

Mitochondrial distribution within the cell soma depends on mitochondrial trafficking along the cytoskeleton. In mammals, microtubule filaments, created by the polymerization of α- and β-tubulin subunits, form a cytoskeletal platform for the distribution and transport of mitochondria The colocalization of mitochondria and microtubules has been well documented, and treatment of cells with microtubule depolymerizing agents results in misalignment and redistribution of mitochondria7, 8. Depending on the cellular context, actin and intermediate filaments can also regulate mitochondrial distribution. For example, the actin cytoskeleton plays a prominent role in organizing the presynaptic terminal and dendritic spines of neurons, and actin filaments help to recruit mitochondria to these regions9.

Motor proteins transport mitochondria along microtubules: Kinesins transport mitochondria in the plus-end (anterograde) direction, and dyneins transport mitochondria in the minus-end (retrograde) direction. Numerous kinesin family members have been implicated in mitochondrial anterograde transport. For instance, deletion of KIF5B, a Kinesin-1 family member, in mammalian placental cells leads to a dramatic redistribution of mitochondria - rather than being evenly distributed throughout the cell, they accumulate into a single perinuclear cluster10. This observation suggests that kinesin-based motility in the microtubule plus-end direction is required for proper mitochondrial distribution throughout the cell body. Other studies have provided evidence for the involvement of kinesins in mitochondrial anterograde motility in mammalian neurons1113. In contrast to yeast (Box 3), deficiencies in mitochondrial motor proteins have not yet been causally linked with mitochondrial segregation defects in mammalian cells.

Additional components of the kinesin-based transport system have been identified in genetic screens of Drosophila melanogaster neuronal function14. The C-terminal cargo-binding domain of kinesin-1 interacts with an adaptor protein called Milton, which in turn binds to Miro, an integral membrane protein localized to the mitochondrial outer membrane. In fly neurons, anterograde mitochondrial transport is severely impaired in the absence of either Milton or Miro15, 16. Interestingly, Miro has calcium-binding EF hand domains that are involved in stalling mitochondria in response to local elevation of calcium concentration at active synapses within axons17. In mouse axons, an additional important factor that regulates mitochondrial localization is local stalling of mitochondria - at any given time, most mitochondria in axons are immobile. This pausing results from anchoring of axonal mitochondria to the microtubule cytoskeleton by syntaphilin (SNPH)18. When calcium levels are high, Miro releases Kinesin-1, which is then bound and inhibited by SNPH19. This close interplay between SNPH and the Milton-Miro transport complex regulates the saltatory movement of axonal mitochondria. Moreover, data from proliferating mammalian cells suggest that the mitochondrial kinesin complex plays an important role in regulating the mitochondrial distribution throughout the cell soma. In mammalian fibroblasts, co-overexpression of Milton and KIF5B, a member of the Kinesin-1 family, results in clustering of mitochondria at the cell periphery20, and KIF5B deletion results in mitochondrial accumulation around the nucleus. Thus, the Miro–Milton–KIF5B complex can regulate mitochondrial distribution through the promotion of anterograde motility (Figure 1C). Recent data suggests that Miro and Milton can also regulate retrograde motility through their association with dynein motor proteins21 (Figure 1C).

Because cells lacking Miro or Milton function show dramatic changes in mitochondrial distribution, it is possible that these molecules play a role in mitochondrial partitioning during cell division. However, this point remains to be experimentally addressed, because most studies have utilized non-dividing neurons. In contrast to mammals, mitochondrial motility and distribution in budding yeast is dependent on the actin cytoskeleton, and mutants in these processes show clear segregation defects (Box 3).

Fusion and fission ensure mitochondrial homogenization

During mitochondrial fusion, two mitochondria merge both their outer and inner membranes, resulting in a single, larger mitochondrion. Mitochondrial fission is the opposite process, whereby a mitochondrion divides into two organelles. Fusion of mitochondria is mediated by three dynamin-related GTPases: the Mitofusins 1 and 2 (Mfn1 and Mfn2) on the mitochondrial outer membrane and Optic atrophy 1 (Opa1) on the inner membrane, whereas fission is mediated by the Dynamin-related protein 1 (Drp1)6, which is also a large GTPase (Box 2). Complete deficiencies in any of these proteins is embryonic lethal in mice, and mutations in some of these proteins are associated with neuromuscular disease in humans6, 22. Studies in cell culture and tissues suggest that fusion and fission are critical for maintaining the health of the mitochondrial population2326. For example, when mitochondrial fusion is blocked, Purkinje neurons in the cerebellum degenerate due to defects in electron transport and mtDNA maintenance27.

How could fusion and fission maintain the proper function of mitochondria? Cells lacking the mitofusins or Opa1 show a striking heterogeneity between mitochondria in protein content, mtDNA nucleoid content and membrane potential23, 2729. As a result, mutant cells show defects in respiratory chain function28 and accumulation of mtDNA deletions23. Thus, cycles of fusion and fission promote mixing of mitochondrial content and homogenization of mitochondrial protein and DNA in the cell. Whereas functional mitochondria continually mix their contents, quality control mechanisms exist to prevent grossly dysfunctional mitochondria from fusing with healthy mitochondria30. For example, mitochondria that lose their inner membrane potential undergo proteolytic inactivation of Opa131, resulting in loss of inner membrane fusion that prevents their re-incorporation into the cellular mitochondrial population network30. In addition, many forms of mitochondrial dysfunction result in activation of mitochondrial fission or inactivation of fusion31, 32. These dynamic changes promote the conversion of dysfunctional mitochondria into smaller physical units that can be targeted for degradation via autophagy (see below). Conversely, mitochondria that elongate due to inhibition of the fission activity of Drp1 are protected from autophagy, presumably due to their enlarged size33, 34.

Mitochondrial fusion and fission dynamics are regulated during the cell cycle35. Enhanced fusion is associated with the G1-to-S phase transition, when the mitochondria tend to appear as large, interconnected networks (Figure 1B). Mitochondrial fission is activated during mitosis35, 36 through the phosphorylation of Drp1, resulting in individual mitochondrial units. Because they are associated with cytoskeletal elements and with the ER (Box 4), these individual mitochondria are distributed evenly throughout the cell soma, which ensures that they are passively partitioned in equal numbers to daughter cells during cytokinesis. In cells lacking Drp1, the mitochondria exist constitutively as elongated networks due to lack of mitochondrial fission. Nevertheless, mitochondria are segregated to daughter cells, presumably because the cytokinesis machinery is sufficient to forcibly cleave mitochondria25, 26. However, the partitioning of mitochondria to daughter cells in Drp1 mutant cells appears less uniform, perhaps because the mitochondria remain in large networks during mitosis25, although this observation remains to be quantified systematically. It is unclear whether this segregation defect has any physiological consequence in the daughter cells.

Box 4

Mitochondria-ER connections

In electron micrographs, contact sites between the ER and mitochondria are often found, with suggestions of a proteinaceous material tethering the two organelles together. Several proteins, including Mfn237, have been implicated in formation of these contacts. These close contacts are important for transmission of calcium signals from the ER to the mitochondria134 and for phospholipid biosynthesis135, 136. In addition, ER tubules appear to initiate mitochondrial fission in both yeast and mammals by wrapping around and constricting the organelle38. This ER-mitochondria contact is an early event observed in many, but not all, fission events. It occurs prior to recruitment of the fission factor Dnm1 (in yeast; known as Drp1 in mammals)38, which further constricts the mitochondrial tubule to mediate scission. The actin cytoskeleton has also been shown to be important for mitochondrial fission39.

Components of the mitochondria-ER tether, termed the ERMES (ER-mitochondria encounter structure) complex, have been identified in budding yeast. It consists of the mitochondrial outer membrane proteins Mdm10 and Mdm34, the integral ER membrane protein Mmm1, and the peripheral mitochondrial outer membrane protein Mdm12. Because of its role in mediating ER-mitochondrial contacts, ERMES may have a significant role in regulating yeast mitochondrial fission. Recent results further suggest that ERMES may have a role in linking mitochondrial fission to segregation of mtDNA nucleoids140,. Components of the ERMES complex are located adjacent to replicating mtDNA nucleoids and may facilitate the equal distribution of mtDNA nucleoids between daughter mitochondria during the fission event140 (see Figure). Perhaps due to these roles in mitochondrial fission, the ERMES components Mdm10, Mdm12 and Mdm34 were originally isolated in yeast screens for genes required for mitochondrial transport to the daughter bud112, 142, 143 (see Box 3). No orthologs of the ERMES complex have been found in mammals.

An external file that holds a picture, illustration, etc.
Object name is nihms644044u4.jpg

Thus, the combined effects of fusion and fission promote a homogenous and healthy population by mixing mitochondrial protein and DNA contents throughout the mitochondrial population of the cell, whereas enhanced fission at the onset of mitosis facilitates equal passive segregation mitochondria to daughter cells.

Mitochondrial clearance by mitophagy

Mitophagy is the sequestration of mitochondria by autophagosomes followed by their degradation in lysosomes. Thought to be the primary mechanism for degradation of mitochondria, mitophagy controls the composition of the mitochondrial population in the cell by culling subpopulations of mitochondria that are dysfunctional. Although mitochondria can be non-selectively removed as part of a bulk autophagy response, mitophagy can be selective for damaged mitochondria under certain cellular contexts. For instance, mitochondria that are experimentally depolarized attract the autophagy machinery and undergo mitophagy40, 41. In yeast, nitrogen-starvation leads to bulk cellular autophagy that includes mitophagy. However, if cells starved for nitrogen are kept on a carbon source requiring mitochondrial metabolism, mitophagy (but not cellular autophagy in general) is reduced, indicating that mitophagy can be differentially regulated from bulk autophagy42. By selectively targeting dysfunctional mitochondria, mitophagy could serve to maintain the quality of the mitochondrial population. However, how efficiently mitophagy is in achieving this remains unclear, as there are many pathological states, such as mitochondrial encephalomyopathies, in which highly dysfunctional mitochondria or pathogenic mtDNA mutations persist in cells and tissues43.

The Pink1-Parkin Mitophagy pathway

The mitophagy pathway most studied involves Pink1 and Parkin; mutations in their encoding genes are involved in familial cases of Parkinson’s disease44. As first shown in fly mutants, Parkin acts downstream of Pink1 to maintain mitochondrial function4547. In mammalian cell culture models, these two proteins promote mitophagy of compromised mitochondria. In response to depolarization of the mitochondrial inner membrane, Parkin, which is an E3 ubiquitin ligase, translocates from the cytosol onto the mitochondrial outer membrane in a Pink1-dependent manner48, 49, where it causes widespread polyubiquitination of mitochondrial outer membrane proteins50, 51. This polyubiquitination promotes degradation of mitochondrial outer membrane proteins by the 26S proteasome50, 52. Protein degradation is also dependent on the AAA ATPase p9753, which may have a role in extracting proteins from the mitochondrial outer membrane in order to make them more accessible to the proteasome54. Degradation of mitochondrial outer membrane proteins is necessary for the mitophagic response50 (Figure 1D). Overexpression of Parkin in cytoplasmic hybrid (cybrid) cells which are heteroplasmic for a mutation in the mitochondrial COX1 gene reduces the level of the pathogenic mutation, suggesting that the Pink1-Parkin pathway can introduce bias in the maintenance of mtDNA genotypes55. The functions of Pink1 and Parkin in mitophagy have mostly been studied by overexpression in mammalian cell lines. However, their involvement in mitophagy is supported by proteomic analyses indicating that flies mutated in these genes have slower turnover of mitochondrial proteins56.

Other mitophagy pathways

During erythrocyte maturation, intracellular organelles such as mitochondria are completely degraded. It has been suggested that the removal of damaged mitochondria is important in preventing downstream pathways that lead to erythrocyte cell death57. The mitochondria are eliminated by sequestration into autophagosomes, but the process may involve both canonical and non-canonical autophagy pathways. At least some mitochondrial degradation persists in the absence of core components of the autophagic machinery, such as the ATG558 or ATG759 proteins. The BH3 (Bcl2 homology domain 3)-containing protein Nix, located on the mitochondrial outer membrane, has been shown to be important for elimination of mitochondria from erythrocytes (Figure 1D). In Nix knockout mice, about a third to a half of the circulating erythrocytes show aberrant persistence of mitochondria, even though other cellular components, such as ribosomes, are degraded60, 61. During maturation, erythroid mitochondria normally show loss of membrane potential, which likely acts as a signal to promote mitophagy. Mitochondrial membrane depolarization fails to occur in Nix-deficient erythroid cells, and artificial dissipation of the membrane potential can restore mitophagy61. Nix contains an N-terminal LIR (LC3-interacting region) motif that binds to LC3 (Microtubule-associated proteins 1A/1B light chain 3)-like proteins, especially GABARAP-L162, which are ubiquitin-related proteins that are located on autophagosomes and are important for their maturation. The LIR motif is essential for the ability of Nix to mediate mitophagy62. These observations suggest that Nix is a selective mitophagy receptor that physically connects the autophagy machinery to the mitochondrial surface in erythroid cells.

In cultured mammalian cells, another mitophagy receptor is the outer membrane protein FUNDC1, which regulates autophagic degradation of mitochondria in response to hypoxia63 (Figure 1D). Overexpression of FUNDC1 reduces total mitochondrial mass in cells, and FUNDC1 depletion conserves mitochondrial mass during hypoxia63. FUNDC1 has a LIR motif that is important for recruitment of LC3 and for mitochondrial degradation. Under hypoxic conditions, the mitochondrial phosphatase PGAM5 (phosphoglycerate mutase family member 5) dephosphorylates the LIR motif, which increases its physical association with LC3 and consequently promotes mitophagy64.

In yeast, genetic screens have identified the outer membrane protein ATG32 as a mitophagy receptor65, 66 (Figure 1E). ATG32 contains an LIR motif that binds ATG8, the yeast LC3 ortholog that is located on the surface of autophagosomes. In addition, it binds ATG11, an adaptor for several selective forms of autophagy. ATG11 has been shown to interact with Dnm167, the yeast ortholog of Drp1 that is involved in mitochondrial fission. Based on these observations, it has been suggested that once ATG11 has been recruited by ATG32 to mitochondria, Dnm1 is recruited to these marked mitochondria to facilitate fission67, and LC3 is recruited to target them to autophagosomes. The process of fission may be important for generating smaller mitochondrial fragments that can be segregated into autophagosomes. Mitochondrial fission has also been shown to facilitate Parkin-mediated mitophagy53. Nevertheless, a study in yeast indicates that Dnm1 and other components of the mitochondrial fission machinery are dispensable for mitophagy68.

Taken together, these results suggest that mitochondrial degradation depends on specific mitophagy receptors present on the mitochondrial outer membrane that function to target mitochondria to the autophagic machinery. It is also possible that in some cases ubiquitination serves to mark mitochondria for autophagy. In addition, mitophagy can also involve outer membrane protein degradation by the ubiquitin-proteasome system and the AAA ATPase p97, and reduction of mitochondrial size by mitochondrial fission50, 53. By culling damaged mitochondria, particularly mitochondria with defective mtDNA, mitophagy has the potential to regulate the functionality of the mitochondrial populations inherited by daughter cells; however more work is necessary to understand the physiological impact of this process.

mtDNA germline dynamics and inheritance

A typical human cell contains hundreds of copies of mtDNA and until a few years ago it was assumed that individuals exhibit homoplasmy for a single mtDNA genotype. This assumption has certain consequences, for example when analysing mtDNA genotypes in forensic cases and for tracing human lineages during evolution. Significant heteroplasmy was known to occur in patients with mitochondrial encephalomyopathies (see below), but these cases were thought to be the exception rather than the rule. With the advent of deep sequencing, however, it has become clear that low-level heteroplasmy in humans is normally present in most tissues69, 70. The source of this heteroplasmy is two-fold. A portion is maternally inherited, as evident by the observation that some of the low frequency alleles found in an individual can be found also in the mother. Another portion, not present in the mother, is presumed to arise from de novo mutations that occur during tissue development69, 70.

Because the mtDNA encodes for respiratory chain proteins and for the tRNAs and rRNAs that are necessary for their translation (Box 1), mutations in the mitochondrial genome result in OXPHOS defects that cause a large collection of diseases termed encephalomyopathies. These diseases have a surprisingly broad range of clinical phenotypes, but tissues that are highly metabolically active, such as skeletal muscle, heart muscle, nerve, and brain, tend to be most commonly affected43. Encephalomyopathies can be familial or sporadic. When familial they have a maternal inheritance pattern, because mtDNA is passed from the mother to her offspring. The most common familial encephalomyopathy is Leber’s hereditary optic neuropathy (LHON), which is caused by Complex I deficiency due to mutations predominantly in the ND1, ND4, or ND6 genes of the mitochondrial genome (Box 1). LHON typically causes bilateral deterioration of central vision due to loss of retinal ganglion cells and atrophy of the optic nerve71, 72. In sporadic encephalomyopathies, respiratory chain defects originate from a de novo mutation. For example, in chronic progressive external ophthalmoplegia (CPEO), mtDNA genomes containing a partial deletion are found in muscle tissue only, implying that the mutation arose in precursor cells of this lineage43. CPEO causes weakness in the eye-associated muscles, resulting in inability to raise the eyelids (ptosis) and difficulty in moving the eyeballs (ophthalmoplegia).

Understanding mtDNA segregation and inheritance has biomedical implications for understanding how mutations in mtDNA can cause the wide variability and diversity of tissue defects found in mitochondrial encephalomyopathies. Such diseases also often have a progressive and variable clinical outcome. Patients inherit a load of pathogenic mtDNA from their mothers, but typically not at high enough levels to cause an immediate OXPHOS deficiency in cells. As an individual ages, genetic drift during cell proliferation results in variability in the levels of pathogenic mtDNA between cells. Only cells or tissues that acquire very high levels of pathogenic mtDNA will exhibit an OXPHOS deficiency. The threshold for this bioenergetic breakdown is quite high: most mtDNA mutations need to accumulate to greater than 60–90% of total mtDNA in the cell before OXPHOS activity is compromised7375. In both mouse76 and human cells77, the complementation of mtDNA mutations was proposed to depend on mixing of genomes by mitochondrial fusion. In a mouse genetically engineered to have high mtDNA mutation rates, reduction in mitochondrial fusion indeed exacerbated the effects of mtDNA mutations23.

mtDNA undergoes a genetic bottleneck during oogenesis

Analysis of mtDNA transmission in animal pedigrees has provided evidence that heteroplasmic mtDNA genotypes can rapidly segregate in the offspring. This phenomenon has been particularly well-documented in Holstein cows78, where offspring from a heteroplasmic mother can have mtDNA haplotype ratios that differ significantly from each other or from the mother. In extreme cases, a mother with low levels of heteroplasmy can give rise to offspring that are homoplasmic for the rare haplotype78. These observations support the occurrence during oogenesis of a severe mtDNA genetic “bottleneck”, such that only a small fraction of mtDNA molecules are ultimately represented in the mature egg (Figure 2). Experimental and theoretical analyses suggest that this bottleneck in mice consists of a couple hundred segregating molecules of mtDNA79. A similar situation occurs in humans and has implications on our understanding of familial mitochondrial encephalomyopathies, as clinical studies indicate that a heteroplasmic mother can have offspring whose pathogenic mtDNA content varies substantially from her own80, 81.

An external file that holds a picture, illustration, etc.
Object name is nihms644044f2.jpg
mtDNA segregation during maternal transmission and early embryogenesis

Two major mechanisms affect mtDNA genotypes during oocyte development — purifying selection (A) and the mtDNA genetic bottleneck (B). This is followed by a second bottleneck in early embryogenesis (C). By eliminating oocytes containing deleterious mtDNA mutations, purifying selection reduces the possibility of severe mtDNA disease and favors the preservation of functional mtDNA. Little is known about the exact timing of this purifying selection, but one study suggests that it occurs during oogenesis prior to the genetic bottleneck84. The mtDNA genetic bottleneck may occur due to reduction of mtDNA copy number or due to amplification of a subset of mtDNA molecules. Regardless of the mechanism, the result of the bottleneck is to decrease the number of haplotypes within an egg. This phenomenon also increases genotypic variance between mature eggs, a feature that facilitates the rapid segregation of mtDNA variants between generations. During very early embryonic development, mtDNA replication is not active, and the mtDNA content per cell is diluted due to an increase in cell number. It has been proposed that this reduction in mtDNA content per cell followed by the resumption of mtDNA replication results in a second bottleneck (C) during early embryonic development91. The graph of mtDNA copy number is meant to illustrate trends during development; direct quantification is available for only selected stages.

Although the occurrence of an mtDNA bottleneck during oogenesis is widely accepted, its molecular mechanism is unclear as recent studies have provided conflicting data about it. During oocyte maturation, a dramatic increase occurs in mitochondrial content and mtDNA copy number, culminating in mature (“metaphase II”) oocytes containing >100,000 copies of mtDNA, presumably in preparation for the metabolic demands of fertilization and implantation. The mtDNA bottleneck is though to occur prior to this period of rapid mtDNA accumulation. Several mechanisms could potentially lead to rapid segregation of mtDNA haplotypes. The bottleneck may be numerical, that is, the level of mtDNA might be transiently constricted earlier in oogenesis, thereby limiting the diversity of mtDNA molecules that will be present in the mature oocyte. In this case, rapid segregation of haplotypes between oocytes is achieved by the reduction of the number of mtDNA molecules present in each oocyte and by random genetic drift. Quantitative measurements of mtDNA numbers during oogenesis provide evidence for82 and against83, 84 the severe developmental contraction of mtDNA content predicted by this model. Alternatively, selective replication and amplification of only a subpopulation of mtDNA molecules84 may take place. The timing of mtDNA replication occurs independently of the cell cycle85. As a result, it is possible for a few mtDNA genomes in the developing oocyte to be preferentially replicated at the expense of others, leading to an apparent bottleneck. In this case, it is unclear whether this selective replication is stochastic or biased towards particular mtDNA haplotypes.

Mitochondrial quality control in the female germline

Genetic bottlenecks and genetic drift would result in oocytes with randomly skewed mtDNA genomes compared to the mtDNA population of primordial germ cells. However, two lines of research involving mice with pathogenic mtDNA mutations indicate the presence of a purifying selection process that monitors the quality and functionality of mtDNA genomes (Figure 2). In the first set of studies, female mice that are heteroplasmic for a severe mtDNA mutation, such as a large deletion86 or a frameshift in the ND6 gene87, show selective reduction of the pathogenic mitochondrial genomes during oogenesis. Offspring of such females have a lower mutation burden, and this tendency to reduce the mutational load is enhanced with maternal age. In contrast, offspring born to mice heteroplasmic for non-pathogenic polymorphisms show mutation burdens that are similar to the maternal burden. This suggests that oocytes containing severe mtDNA mutations are selectively eliminated from the germline in an age-dependent manner. Further evidence for the existence of an mtDNA quality control filter comes from the second set of studies, which analysed mice containing high levels of mtDNA mutations generated by a knock-in of an error-prone mtDNA polymerase. Female mice harboring high levels of mtDNA mutations were mated to a wildtype male, and the spectrum of mtDNA mutations in the offspring was compared to that in the mother. The offspring showed a strong suppression of mutations that change the coding sequence of protein-coding genes88, a clear signature of genetic selection. Collectively, these studies point towards a process of “purifying selection” that helps maintain the functionality of mtDNA in the female germline and therefore in the offspring. The presence of this filter likely explains why, in humans, most of the mutations causing familial encephalomyopathies are relatively mild, and most encephalomyopathies caused by large deletions are sporadic rather than familial43. mtDNA genomes with large deletions do arise de novo in somatic tissues with age89, but these mutations are not passed through the germline.

Little is known about the molecular mechanism of this quality control filter. Even severe mtDNA mutations, such as deletions, do not cause OXPHOS defects until they accumulate beyond high thresholds7375. In the studies described above using heteroplasmic mice, purifying selection is observed at mutation loads far lower than these thresholds86, 87, at least when the thresholds are assessed in commonly used cell models. Therefore, if the filter operates at the cellular level, it is surprisingly sensitive and likely involves a mechanism assessing a feature other than total cellular OXPHOS capacity. Because some mtDNA mutations increase the levels of reactive oxygen species (ROS), ROS sensing is one possible mechanism for this quality control filter. Another possibility is that the filter is operating on the level of individual organelles, so that dysfunctional organelles containing pathogenic genomes might be segregated and eliminated, for example by mitophagy.

A second genetic bottleneck during early embryogenesis

Between the zygote and the blastocyst stages90 there is no mtDNA replication, which results in an mtDNA copy number reduction during early embryogenesis due to dilution as cells rapidly divide. Later in development, mtDNA replication provides an increase in copy number, depending on tissue type. Experiments in rhesus monkeys suggest the existence of an additional genetic bottleneck during this early embryonic period prior to the resumption of mtDNA replication91 (Figure 2). When heteroplasmic oocytes are fertilized, the resulting 8-cell embryo shows a broad range of heteroplasmy in its individual blastomeres. However, later in development, fetuses arising from heteroplasmic oocytes show a tendency toward homoplasmy91. These observations suggest that somatic cells undergo an mtDNA bottleneck sometime after the formation of the blastocyst. In contrast, female germline cells in the foetus retain a broad range of heteroplasmy, suggesting that mtDNA inheritance in somatic and germline lineages are differentially regulated during embryogenesis. Another experiment with heteroplasmic mice has revealed that some somatic tissues can rapidly change the composition of their mtDNA92: In mice heteroplasmic for wildtype Balb/c and NZB mtDNA haplotypes, blood and spleen invariably show progressive accumulation of the Balb/c haplotype, whereas the kidney and liver show accumulation of the NZB haplotype. The basis for this is unknown, but the two haplotypes do not result in differences in respiratory activity of the mitochondria93. These observations suggest that the genetic bottlenecks in somatic cells may include a considerable degree of complexity and tissue specificity.

Depletion of paternal mtDNA

A dramatic example of bias in mitochondrial segregation is the exclusively maternal inheritance of mitochondria in mammals. This has been recently further validated in humans by deep sequencing techniques, which enable the detection of even minor maternal genotypes, but failed to detect any signatures of paternal mtDNA in the tissues of offspring69.

Although uniparental inheritance of mtDNA is widespread in eukaryotes, the reason for its evolutionary conservation remains unclear. Most naturally occurring polymorphisms in mtDNA have negligible effects on mitochondrial function, and it is not obvious why it would be detrimental to have an equal mixture of two different, wildtype mtDNA genotypes (though as mentioned above, humans normally do have low levels of heteroplasmy). A recent mouse study has provided striking evidence that substantial germline-inherited heteroplasmy can result in metabolic and behavioural aberrations97. Cytoplasts from the NZB mouse strain were fused with female embryonic stem cells from the 129S6 mouse strain to generate heteroplasmic stem cells, which were used to produce lines of heteroplasmic mice. When passed through the female germ line, a marked tendency towards the reduction of the NZB mtDNA genotype was observed, indicating that a bias in transmission between generations can occur even with two wildtype mtDNA genotypes. The NZB and 129S6 mtDNA genotypes are presumably equivalent in function, because mice homoplasmic for either show indistinguishable levels of activity, fertility, or behaviour. However when NZB-129S6 heteroplasmic animals were compared to their homoplasmic NZB or 129S6 littermates, they exhibited a number of metabolic and behavioural changes, including reduced physical activity and lower food intake. In addition, they showed aberrations in tests that measure anxiety and spatial learning. This mouse model therefore provides evidence that heteroplasmy can have metabolic and behavioural effects and indicates that even apparently equivalent wild type mitochondrial genomes can be distinguished during germline inheritance97. Uniparental inheritance of mtDNA may therefore be advantageous by promoting homoplasmy and avoiding such phenotypes in the offspring. The molecular mechanisms underlying these heteroplasmic defects are unknown, but may be related to subtle differences in bioenergetic properties or ROS generation between the two mtDNA haplotypes.

An early hypothesis suggested that uniparental inheritance of mitochondria may be a passive consequence of the asymmetry in male and female gamete size; for example, in mammals, the male gamete is much smaller and contains many fewer mitochondria compared to the oocyte. However, some organisms, such as the unicellular green alga Chlamydomonas reinhardtii, have equal-sized gametes but nevertheless show uniparental inheritance of mtDNA98. As discussed below, recent studies indicate that active mechanisms that function pre- and post-fertilization have evolved to ensure uniparental inheritance of mtDNA. There are indications that these mechanisms can occasionally go awry or be bypassed. For example, there is a single clinical case reported of a patient with exercise intolerance who inherited mtDNA haplotypes from both his mother and father, with the paternal haplotype accumulating to high levels in skeletal muscles99. The patient’s myopathy resulted from an apparent de novo mutation that arose in the paternal mtDNA.

Pre-fertilization depletion of paternal mtDNA

In Drosophila melanogaster, mtDNA are eliminated from spermatozoa as they develop from the “onion stage” spermatid to fully elongated spermatids that are subsequently individualized (Figure 3A). Spermatid maturation takes place within a syncytium; sperm individualization is a process necessary for generating individual gametes, whereby each spermatid is packaged into a separate plasma membrane, separating it from the syncytium. During the final stage of spermatid elongation, the mtDNA is progressively removed from the nuclear end towards the tail100. By the time spermatids are fully elongated, few mtDNA nucleoids persist. As a result, mature sperm in Drosophila melanogaster normally have little if any mtDNA by the time of fertilization.

An external file that holds a picture, illustration, etc.
Object name is nihms644044f3.jpg
Mechanisms ensuring the elimination of paternal mtDNA

(A) Pre-fertilization mechanisms in flies. In Drosophila melanogaster, two mechanisms remove mtDNA from sperm. Left: mtDNA is normally degraded during sperm elongation by the mitochondrial endonuclease EndoG. Right: In EndoG mutant flies, the mtDNA persists beyond the elongation stage but is ultimately removed by a second mechanism (extrusion) during the individualization stage, in which mtDNA and cellular debris are sequestered into a waste compartment that is extruded from the sperm body. (B) Post-fertilization mechanism in the nematode Caenorhabditis elegans. Shortly after fertilization, paternal mitochondria (bright green) co-localize with autophagy markers (not shown) and are eliminated by mitophagy by the 16–64 cell stages. Worms deficient for autophagy show persistence of paternal mitochondria long beyond this stage.

This degradation process is dependent on EndoG, one of five nucleases with mitochondrial targeting signals100. In EndoG mutants, mtDNA nucleoids aberrantly persist in the tail regions of fully elongated and individualized spermatids. Nevertheless, eggs fertilized by EndoG mutant males do not contain detectable paternal mtDNA. Analysis of EndoG mutants reveals a redundant process in later spermatid development that further eliminates mtDNA from the sperm. When sperm undergo individualization in EndoG mutant flies, the residual mtDNA in the spermatid tail is transported, along with other cellular components, into a waste compartment that is culled from the spermatid (Figure 3A). It will be interesting to determine whether similar pre-fertilization mechanisms occur in mammals. Indeed in mice, there is a 10-fold decrease in mtDNA levels as sperm mature101, but the molecular basis of this is unknown.

Post-fertilization depletion of paternal mtDNA

In bovine and primate zygotes, the paternal mitochondria can be selectively stained using ubiquitin antibodies, and ubiquitin has been proposed to be a mark that targets the paternal organelles for proteasomal destruction102. According to this hypothesis mitochondrial ubiquitination is performed in the sperm and no active mechanism in the egg would be needed to mark the paternal mitochondria for post-fertilization destruction102.

Genetic studies in Caenorhabditis elegans clearly indicate a role for mitophagy in removal of paternal mitochondria after fertilization103105. Shortly after fertilization, autophagosomes are recruited to paternal mitochondria and consequently they disappear by the 16-cell or the 64-cell stages103, 104 (Figure 3B). Mutants in the autophagy pathway show persistence of paternal mitochondria well beyond these stages103, 104. Similar results were obtained in embryos treated with inhibitors of lysosomal function105. In contrast to mammals, the sperm mitochondria in nematode embryos are not polyubiquitinated103, 104. Nevertheless, nematodes with mutations in proteasome ubiquitin receptors are impaired in the degradation of their paternal mitochondria105.

Fertilized mouse oocytes show co-localization of the autophagic markers LC3, GABARAP, and p62 with the paternal mitochondria103, suggesting that autophagy is a conserved mechanism for paternal mtDNA depletion in mammals as well. However, this view has been challenged by the recent observation that the association of LC3 and p62 with paternal mitochondria is transient and occurs well before the loss of mitochondria106. As a result, it has been suggested that elimination of mtDNA in mammalian sperm is a passive process, largely accomplished prior to fertilization by reduction of mtDNA in sperm106. In addition, mitochondrial partitioning in the dividing blastomeres is uneven, and this may remove the remaining paternal mtDNA from most tissues of the embryo106. More work is needed in to resolve the mechanism of mammalian uniparental mtDNA inheritance.

Concluding remarks

In mammalian cells, the partitioning of mitochondria to daughter cells during cell division appears to involve passive mechanisms. Mitochondria undergo fragmentation during mitosis and are well distributed throughout the cell soma due to their interactions with the ER and cytoskeleton. As a result, cytokinesis can result in the partitioning of roughly equal amounts of mitochondria to daughter cells. The role of active transport processes, such as the Milton-Miro-Kinesin complex, in mitochondrial partitioning during cell division has not been examined. Active partitioning mechanisms have been delineated in budding yeast; it seems likely that proliferating mammalian cells would utilize analogous mechanisms, but thus far there are no data to support this. Because cells do not survive without mitochondria, the availability of temperature-sensitive mutants in S. cerevisiae has been instrumental in identifying mitochondrial transport mechanisms to the bud. To make further progress in our understanding of mammalian mitochondrial partitioning, new assays and model systems will need to be developed to analyse mitochondrial distribution, heterogeneity and segregation in mammalian cells, and to study asymmetric cell division.

During maternal transmission of mitochondria, mtDNA genetic bottleneck and purifying selection during oogenesis have major influences on the physiology of offspring and potentially on the overall evolution of species. The molecular mechanisms underlying these complex phenomena are poorly understood, and it is unknown whether the mechanisms regulating mitochondrial partitioning in cell division are relevant to understanding mitochondrial transmission between generations. A future challenge will be to synthesize these parallel lines of research. Mechanisms of mitochondrial quality control are another area for future exploration. Several pathways of mitophagy have been identified so far, and it will be important to understand the extent to which these pathways contribute to maintenance of mitochondrial function in health and disease.

Online Summary

  • Mitochondria are organelles with key roles in cellular metabolism. They have unique cellular dynamics to ensure their proper distribution to dividing cells and high fidelity of inheritance of their genome in a maternal mode of transmission.

  • In mammals, mitochondrial segregation during cell division appears to be a primarily a passive process. Mitochondrial fusion, fission, transport, contacts with the ER, and mitophagy all play a role in maintaining a homogeneous population that is well spatially distributed in the cell soma and thus can be partitioned equally to daughter cells.

  • mtDNA inheritance from one generation to another is strongly influenced by mtDNA bottlenecks and genetic selection that occur during oogenesis and early embryonic development. Quality control mechanisms are likely present to minimize the accumulation of pathogenic mutations, which lead to a class of diseases termed mitochondrial encephalomyopathies.

  • The depletion of paternal mitochondria during fertilization is nearly universal in metozoans, though its timing and mechanisms vary substantially between species. Proteasome-dependent degradation, mtDNA degradation and mitophagy have been implicated in this process in different organisms.

Acknowledgments

Work in laboratory of D.C.C. is supported by NIH grants GM062967 and GM110039, the Muscular Dystrophy Association, and the Howard Hughes Medical Institute. P.M. is supported by a Baxter Postdoctoral Fellowship.

Glossary

mtDNA nucleoidA punctate structure within mitochondria containing one or more copies of the mtDNA genome in complex with proteins
Oxidative phosphorylationA biochemical pathway within mitochondria that generates ATP through the oxidation of nutrients
MyosinA family of ATP-dependent molecular motors that transports cargo along actin filaments
KinesinA family of ATP-dependent molecular motors that transports cargo along microtubule filaments, usually towards the plus-end
DyneinA family of ATP-dependent molecular motors that transports cargo along microtubule filaments towards the minus-end
Endosymbiotic theoryA theory postulating that genome-containing organelles, such as mitochondria and chloroplasts, are derived from prokaryotes that have undergone endosymbiosis with the ancestral eukaryotic cells
HeteroplasmyA state in which more than one haplotype of mtDNA exists in a cell or organism
HomoplasmyA state in which a single haplotype of mtDNA exists in a cell or organism
dynamin-related GTPasesa family of large GTP hydrolyzing enzymes related to “classical” dynamin, which uses GTP hydrolysis to mechanically remodel membranes
microtubule directionality (plus/minus)Describes the polarity of microtubule filaments, which are polymerized from asymmetric dimers of α-tublin and β-tubulin
formin proteinsa group of proteins that regulate the actin cytoskeleton and cell signaling
cytoplasmic hybrid (cybrid) cella hybrid cell created by the fusion of a cell with an enucleated second cell, thereby combining the nucleus and mitochondria from two distinct cells
genetic bottlenecka drastic reduction in the genetic diversity of a population
genetic drifta change in the allele or haplotype frequencies within a population due to stochastic forces
purifying selectiona mode of natural selection in which detrimental genetic variants are selected against
macroautophagyThe most commonly studied autophagic process, whereby cellular components are incorporated into autophagosomes and ultimately degraded or recycled through the lysosome. Macroautophagy is commonly referred to as simply “autophagy”
autophagosomesa double-membrane structure that engulfs cellular components for delivery to the lysosome
lysosomesorganelles that degrade cellular components using acid hydrolases
motor proteinsmolecular motors that use ATP hydrolysis to power movement of cargo along substrate surfaces

Biographies

• 

Prashant Mishra is a postdoctoral fellow in the laboratory of David Chan at the California Institute of Technology (Pasadena, CA, USA). After finishing his bachelor’s degree in Biochemical Sciences at Harvard University, he completed his M.D./Ph.D degrees at the University of Texas Southwestern Medical Center. His Ph.D. studies were carried out in the laboratory of Rama Ranganathan. His current research focuses on regulation of mitochondrial fusion.

• 

David C. Chan is Professor of Biology in the Division of Biology and Biological Engineering at the California Institute of Technology (Pasadena, CA, USA). After M.D./Ph.D. studies at Harvard Medical School, he completed postdoctoral work at the Whitehead Institute at the Massachusetts Institute of Technology. His laboratory works on the biochemistry and physiological functions of mitochondrial dynamics.

References

1. Scheffler IE. Mitochondria. John Wiley & Sons, Inc; Hoboken, NJ: 2009. [Google Scholar]
2. Stehling O, Wilbrecht C, Lill R. Mitochondrial iron-sulfur protein biogenesis and human disease. Biochimie. 2014;100C:61–77. [Abstract] [Google Scholar]
3. Tait SW, Green DR. Mitochondria and cell death: outer membrane permeabilization and beyond. Nat Rev Mol Cell Biol. 2010;11:621–32. [Abstract] [Google Scholar]
4. Cloonan SM, Choi AM. Mitochondria: sensors and mediators of innate immune receptor signaling. Curr Opin Microbiol. 2013;16:327–38. [Europe PMC free article] [Abstract] [Google Scholar]
5. Gray MW, Burger G, Lang BF. Mitochondrial evolution. Science. 1999;283:1476–81. [Abstract] [Google Scholar]
6. Chan DC. Fusion and fission: interlinked processes critical for mitochondrial health. Annu Rev Genet. 2012;46:265–87. [Abstract] [Google Scholar]
7. Ball EH, Singer SJ. Mitochondria are associated with microtubules and not with intermediate filaments in cultured fibroblasts. Proc Natl Acad Sci U S A. 1982;79:123–6. [Europe PMC free article] [Abstract] [Google Scholar]
8. Heggeness MH, Simon M, Singer SJ. Association of mitochondria with microtubules in cultured cells. Proc Natl Acad Sci U S A. 1978;75:3863–6. [Europe PMC free article] [Abstract] [Google Scholar]
9. Sheng ZH, Cai Q. Mitochondrial transport in neurons: impact on synaptic homeostasis and neurodegeneration. Nat Rev Neurosci. 2012;13:77–93. [Europe PMC free article] [Abstract] [Google Scholar]
10. Tanaka Y, et al. Targeted disruption of mouse conventional kinesin heavy chain, kif5B, results in abnormal perinuclear clustering of mitochondria. Cell. 1998;93:1147–58. [Abstract] [Google Scholar]
11. Tanaka K, Sugiura Y, Ichishita R, Mihara K, Oka T. KLP6: a newly identified kinesin that regulates the morphology and transport of mitochondria in neuronal cells. J Cell Sci. 2011;124:2457–65. [Abstract] [Google Scholar]
12. Hirokawa N, Noda Y. Intracellular transport and kinesin superfamily proteins, KIFs: structure, function, and dynamics. Physiol Rev. 2008;88:1089–118. [Abstract] [Google Scholar]
13. Nangaku M, et al. KIF1B, a novel microtubule plus end-directed monomeric motor protein for transport of mitochondria. Cell. 1994;79:1209–20. Shows that the kinesin family member KIF1B mediates the transport of mitochondria in the cytoplasm in vitro and in vivo. [Abstract] [Google Scholar]
14. Saxton WM, Hollenbeck PJ. The axonal transport of mitochondria. J Cell Sci. 2012;125:2095–104. [Europe PMC free article] [Abstract] [Google Scholar]
15. Stowers RS, Megeath LJ, Gorska-Andrzejak J, Meinertzhagen IA, Schwarz TL. Axonal transport of mitochondria to synapses depends on milton, a novel Drosophila protein. Neuron. 2002;36:1063–77. [Abstract] [Google Scholar]
16. Guo X, et al. The GTPase dMiro is required for axonal transport of mitochondria to Drosophila synapses. Neuron. 2005;47:379–93. [Abstract] [Google Scholar]
17. Wang X, Schwarz TL. The mechanism of Ca2+-dependent regulation of kinesin-mediated mitochondrial motility. Cell. 2009;136:163–74. Shows that the Milton-kinesin complex mediates calcium-dependent motility of mitochondria in neurons. [Europe PMC free article] [Abstract] [Google Scholar]
18. Kang JS, et al. Docking of axonal mitochondria by syntaphilin controls their mobility and affects short-term facilitation. Cell. 2008;132:137–48. [Europe PMC free article] [Abstract] [Google Scholar]
19. Chen Y, Sheng ZH. Kinesin-1-syntaphilin coupling mediates activity-dependent regulation of axonal mitochondrial transport. J Cell Biol. 2013;202:351–64. [Europe PMC free article] [Abstract] [Google Scholar]
20. Glater EE, Megeath LJ, Stowers RS, Schwarz TL. Axonal transport of mitochondria requires milton to recruit kinesin heavy chain and is light chain independent. J Cell Biol. 2006;173:545–57. [Europe PMC free article] [Abstract] [Google Scholar]
21. van Spronsen M, et al. TRAK/Milton motor-adaptor proteins steer mitochondrial trafficking to axons and dendrites. Neuron. 2013;77:485–502. [Abstract] [Google Scholar]
22. Youle RJ, van der Bliek AM. Mitochondrial fission, fusion, and stress. Science. 2012;337:1062–5. [Europe PMC free article] [Abstract] [Google Scholar]
23. Chen H, et al. Mitochondrial fusion is required for mtDNA stability in skeletal muscle and tolerance of mtDNA mutations. Cell. 2010;141:280–9. Shows that mitochondrial fusion is important for the tolerance of mtDNA mutations. [Europe PMC free article] [Abstract] [Google Scholar]
24. Chen H, et al. Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development. J Cell Biol. 2003;160:189–200. [Europe PMC free article] [Abstract] [Google Scholar]
25. Ishihara N, et al. Mitochondrial fission factor Drp1 is essential for embryonic development and synapse formation in mice. Nat Cell Biol. 2009;11:958–66. [Abstract] [Google Scholar]
26. Wakabayashi J, et al. The dynamin-related GTPase Drp1 is required for embryonic and brain development in mice. J Cell Biol. 2009;186:805–16. [Europe PMC free article] [Abstract] [Google Scholar]
27. Chen H, McCaffery JM, Chan DC. Mitochondrial fusion protects against neurodegeneration in the cerebellum. Cell. 2007;130:548–62. [Abstract] [Google Scholar]
28. Chen H, Chomyn A, Chan DC. Disruption of fusion results in mitochondrial heterogeneity and dysfunction. J Biol Chem. 2005;280:26185–92. Shows that the deletion of the mitofusin 1 and 2 genes leads to heterogeneity in mitochondrial membrane potential and defects in oxidative phosphorylation. [Abstract] [Google Scholar]
29. Weaver D, et al. Distribution and apoptotic function of outer membrane proteins depend on mitochondrial fusion. Mol Cell. 2014;54:870–8. [Europe PMC free article] [Abstract] [Google Scholar]
30. Twig G, et al. Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J. 2008;27:433–46. [Europe PMC free article] [Abstract] [Google Scholar]
31. Ishihara N, Fujita Y, Oka T, Mihara K. Regulation of mitochondrial morphology through proteolytic cleavage of OPA1. EMBO J. 2006;25:2966–77. [Europe PMC free article] [Abstract] [Google Scholar]
32. Frank S, et al. The role of dynamin-related protein 1, a mediator of mitochondrial fission, in apoptosis. Dev Cell. 2001;1:515–25. [Abstract] [Google Scholar]
33. Gomes LC, Di Benedetto G, Scorrano L. During autophagy mitochondria elongate, are spared from degradation and sustain cell viability. Nat Cell Biol. 2011;13:589–98. [Europe PMC free article] [Abstract] [Google Scholar]
34. Rambold AS, Kostelecky B, Elia N, Lippincott-Schwartz J. Tubular network formation protects mitochondria from autophagosomal degradation during nutrient starvation. Proc Natl Acad Sci U S A. 2011;108:10190–5. [Europe PMC free article] [Abstract] [Google Scholar]
35. Mitra K, Wunder C, Roysam B, Lin G, Lippincott-Schwartz J. A hyperfused mitochondrial state achieved at G1-S regulates cyclin E buildup and entry into S phase. Proc Natl Acad Sci U S A. 2009;106:11960–5. [Europe PMC free article] [Abstract] [Google Scholar]
36. Taguchi N, Ishihara N, Jofuku A, Oka T, Mihara K. Mitotic phosphorylation of dynamin-related GTPase Drp1 participates in mitochondrial fission. J Biol Chem. 2007;282:11521–9. [Abstract] [Google Scholar]
37. de Brito OM, Scorrano L. Mitofusin 2 tethers endoplasmic reticulum to mitochondria. Nature. 2008;456:605–10. [Abstract] [Google Scholar]
38. Friedman JR, et al. ER tubules mark sites of mitochondrial division. Science. 2011;334:358–62. Shows that endoplasmic reticulum tubules are involved in the constriction of mitochondria at fission sites. [Europe PMC free article] [Abstract] [Google Scholar]
39. Korobova F, Ramabhadran V, Higgs HN. An actin-dependent step in mitochondrial fission mediated by the ER-associated formin INF2. Science. 2013;339:464–7. [Europe PMC free article] [Abstract] [Google Scholar]
40. Kim I, Rodriguez-Enriquez S, Lemasters JJ. Selective degradation of mitochondria by mitophagy. Arch Biochem Biophys. 2007;462:245–53. [Europe PMC free article] [Abstract] [Google Scholar]
41. Rodriguez-Enriquez S, He L, Lemasters JJ. Role of mitochondrial permeability transition pores in mitochondrial autophagy. Int J Biochem Cell Biol. 2004;36:2463–72. [Abstract] [Google Scholar]
42. Kanki T, Klionsky DJ. Mitophagy in yeast occurs through a selective mechanism. J Biol Chem. 2008;283:32386–93. [Europe PMC free article] [Abstract] [Google Scholar]
43. Schon EA, DiMauro S, Hirano M. Human mitochondrial DNA: roles of inherited and somatic mutations. Nat Rev Genet. 2012;13:878–90. [Europe PMC free article] [Abstract] [Google Scholar]
44. Houlden H, Singleton AB. The genetics and neuropathology of Parkinson’s disease. Acta Neuropathol. 2012;124:325–38. [Europe PMC free article] [Abstract] [Google Scholar]
45. Clark IE, et al. Drosophila pink1 is required for mitochondrial function and interacts genetically with parkin. Nature. 2006;441:1162–6. [Abstract] [Google Scholar]
46. Park J, et al. Mitochondrial dysfunction in Drosophila PINK1 mutants is complemented by parkin. Nature. 2006;441:1157–61. [Abstract] [Google Scholar]
47. Yang Y, et al. Mitochondrial pathology and muscle and dopaminergic neuron degeneration caused by inactivation of Drosophila Pink1 is rescued by Parkin. Proc Natl Acad Sci U S A. 2006;103:10793–8. [Europe PMC free article] [Abstract] [Google Scholar]
48. Narendra D, Tanaka A, Suen DF, Youle RJ. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol. 2008;183:795–803. Shows that Parkin translocates to mitochondria with depolarized membranes and promotes their removal by mitophagy. [Europe PMC free article] [Abstract] [Google Scholar]
49. Narendra DP, et al. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol. 2010;8:e1000298. [Europe PMC free article] [Abstract] [Google Scholar]
50. Chan NC, et al. Broad activation of the ubiquitin-proteasome system by Parkin is critical for mitophagy. Hum Mol Genet. 2011;20:1726–37. [Europe PMC free article] [Abstract] [Google Scholar]
51. Sarraf SA, et al. Landscape of the PARKIN-dependent ubiquitylome in response to mitochondrial depolarization. Nature. 2013;496:372–6. [Europe PMC free article] [Abstract] [Google Scholar]
52. Yoshii SR, Kishi C, Ishihara N, Mizushima N. Parkin mediates proteasome-dependent protein degradation and rupture of the outer mitochondrial membrane. J Biol Chem. 2011;286:19630–40. [Europe PMC free article] [Abstract] [Google Scholar]
53. Tanaka A, et al. Proteasome and p97 mediate mitophagy and degradation of mitofusins induced by Parkin. J Cell Biol. 2010;191:1367–80. [Europe PMC free article] [Abstract] [Google Scholar]
54. Xu S, Peng G, Wang Y, Fang S, Karbowski M. The AAA-ATPase p97 is essential for outer mitochondrial membrane protein turnover. Mol Biol Cell. 2011;22:291–300. [Europe PMC free article] [Abstract] [Google Scholar]
55. Suen DF, Narendra DP, Tanaka A, Manfredi G, Youle RJ. Parkin overexpression selects against a deleterious mtDNA mutation in heteroplasmic cybrid cells. Proc Natl Acad Sci U S A. 2010;107:11835–40. [Europe PMC free article] [Abstract] [Google Scholar]
56. Vincow ES, et al. The PINK1-Parkin pathway promotes both mitophagy and selective respiratory chain turnover in vivo. Proc Natl Acad Sci U S A. 2013;110:6400–5. [Europe PMC free article] [Abstract] [Google Scholar]
57. Mortensen M, et al. Loss of autophagy in erythroid cells leads to defective removal of mitochondria and severe anemia in vivo. Proc Natl Acad Sci U S A. 2010;107:832–7. [Europe PMC free article] [Abstract] [Google Scholar]
58. Matsui M, Yamamoto A, Kuma A, Ohsumi Y, Mizushima N. Organelle degradation during the lens and erythroid differentiation is independent of autophagy. Biochem Biophys Res Commun. 2006;339:485–9. [Abstract] [Google Scholar]
59. Zhang J, et al. Mitochondrial clearance is regulated by Atg7-dependent and -independent mechanisms during reticulocyte maturation. Blood. 2009;114:157–64. [Europe PMC free article] [Abstract] [Google Scholar]
60. Schweers RL, et al. NIX is required for programmed mitochondrial clearance during reticulocyte maturation. Proc Natl Acad Sci U S A. 2007;104:19500–5. [Europe PMC free article] [Abstract] [Google Scholar]
61. Sandoval H, et al. Essential role for Nix in autophagic maturation of erythroid cells. Nature. 2008;454:232–5. [Europe PMC free article] [Abstract] [Google Scholar]
62. Novak I, et al. Nix is a selective autophagy receptor for mitochondrial clearance. EMBO Rep. 2010;11:45–51. [Europe PMC free article] [Abstract] [Google Scholar]
63. Liu L, et al. Mitochondrial outer-membrane protein FUNDC1 mediates hypoxia-induced mitophagy in mammalian cells. Nat Cell Biol. 2012;14:177–85. [Abstract] [Google Scholar]
64. Chen G, et al. A Regulatory Signaling Loop Comprising the PGAM5 Phosphatase and CK2 Controls Receptor-Mediated Mitophagy. Mol Cell. 2014;54:362–77. [Abstract] [Google Scholar]
65. Okamoto K, Kondo-Okamoto N, Ohsumi Y. Mitochondria-anchored receptor Atg32 mediates degradation of mitochondria via selective autophagy. Dev Cell. 2009;17:87–97. [Abstract] [Google Scholar]
66. Kanki T, Wang K, Cao Y, Baba M, Klionsky DJ. Atg32 is a mitochondrial protein that confers selectivity during mitophagy. Dev Cell. 2009;17:98–109. [Europe PMC free article] [Abstract] [Google Scholar]
67. Mao K, Wang K, Liu X, Klionsky DJ. The scaffold protein Atg11 recruits fission machinery to drive selective mitochondria degradation by autophagy. Dev Cell. 2013;26:9–18. [Europe PMC free article] [Abstract] [Google Scholar]
68. Mendl N, et al. Mitophagy in yeast is independent of mitochondrial fission and requires the stress response gene WHI2. J Cell Sci. 2011;124:1339–50. [Abstract] [Google Scholar]
69. He Y, et al. Heteroplasmic mitochondrial DNA mutations in normal and tumour cells. Nature. 2010;464:610–4. [Europe PMC free article] [Abstract] [Google Scholar]
70. Payne BA, et al. Universal heteroplasmy of human mitochondrial DNA. Hum Mol Genet. 2013;22:384–90. [Europe PMC free article] [Abstract] [Google Scholar]
71. Yu-Wai-Man P, Griffiths PG, Chinnery PF. Mitochondrial optic neuropathies - disease mechanisms and therapeutic strategies. Prog Retin Eye Res. 2011;30:81–114. [Europe PMC free article] [Abstract] [Google Scholar]
72. Carelli V, et al. Retinal ganglion cell neurodegeneration in mitochondrial inherited disorders. Biochim Biophys Acta. 2009;1787:518–528. [Abstract] [Google Scholar]
73. Chomyn A, et al. MELAS mutation in mtDNA binding site for transcription termination factor causes defects in protein synthesis and in respiration but no change in levels of upstream and downstream mature transcripts. Proc Natl Acad Sci U S A. 1992;89:4221–5. [Europe PMC free article] [Abstract] [Google Scholar]
74. Hayashi J, et al. Introduction of disease-related mitochondrial DNA deletions into HeLa cells lacking mitochondrial DNA results in mitochondrial dysfunction. Proc Natl Acad Sci U S A. 1991;88:10614–8. [Europe PMC free article] [Abstract] [Google Scholar]
75. Rossignol R, et al. Mitochondrial threshold effects. Biochem J. 2003;370:751–62. [Europe PMC free article] [Abstract] [Google Scholar]
76. Nakada K, et al. Inter-mitochondrial complementation: Mitochondria-specific system preventing mice from expression of disease phenotypes by mutant mtDNA. Nat Med. 2001;7:934–40. [Abstract] [Google Scholar]
77. Ono T, Isobe K, Nakada K, Hayashi JI. Human cells are protected from mitochondrial dysfunction by complementation of DNA products in fused mitochondria. Nat Genet. 2001;28:272–5. [Abstract] [Google Scholar]
78. Koehler CM, et al. Replacement of bovine mitochondrial DNA by a sequence variant within one generation. Genetics. 1991;129:247–55. [Europe PMC free article] [Abstract] [Google Scholar]
79. Jenuth JP, Peterson AC, Fu K, Shoubridge EA. Random genetic drift in the female germline explains the rapid segregation of mammalian mitochondrial DNA. Nat Genet. 1996;14:146–51. [Abstract] [Google Scholar]
80. Ghosh SS, Fahy E, Bodis-Wollner I, Sherman J, Howell N. Longitudinal study of a heteroplasmic 3460 Leber hereditary optic neuropathy family by multiplexed primer-extension analysis and nucleotide sequencing. Am J Hum Genet. 1996;58:325–34. [Europe PMC free article] [Abstract] [Google Scholar]
81. Larsson NG, et al. Segregation and manifestations of the mtDNA tRNA(Lys) A-->G(8344) mutation of myoclonus epilepsy and ragged-red fibers (MERRF) syndrome. Am J Hum Genet. 1992;51:1201–12. [Europe PMC free article] [Abstract] [Google Scholar]
82. Cree LM, et al. A reduction of mitochondrial DNA molecules during embryogenesis explains the rapid segregation of genotypes. Nat Genet. 2008;40:249–54. [Abstract] [Google Scholar]
83. Cao L, et al. New evidence confirms that the mitochondrial bottleneck is generated without reduction of mitochondrial DNA content in early primordial germ cells of mice. PLoS Genet. 2009;5:e1000756. [Europe PMC free article] [Abstract] [Google Scholar]
84. Wai T, Teoli D, Shoubridge EA. The mitochondrial DNA genetic bottleneck results from replication of a subpopulation of genomes. Nat Genet. 2008;40:1484–8. [Abstract] [Google Scholar]
85. Bogenhagen D, Clayton DA. Mouse L cell mitochondrial DNA molecules are selected randomly for replication throughout the cell cycle. Cell. 1977;11:719–27. [Abstract] [Google Scholar]
86. Sato A, et al. Deletion-mutant mtDNA increases in somatic tissues but decreases in female germ cells with age. Genetics. 2007;177:2031–7. [Europe PMC free article] [Abstract] [Google Scholar]
87. Fan W, et al. A mouse model of mitochondrial disease reveals germline selection against severe mtDNA mutations. Science. 2008;319:958–62. [Europe PMC free article] [Abstract] [Google Scholar]
88. Stewart JB, Freyer C, Elson JL, Larsson NG. Purifying selection of mtDNA and its implications for understanding evolution and mitochondrial disease. Nat Rev Genet. 2008;9:657–62. Together with reference 87, provides evidence for the existance of a purifying selection mechanism for mtDNA mutations in the mouse female germline. [Abstract] [Google Scholar]
89. Chomyn A, Attardi G. MtDNA mutations in aging and apoptosis. Biochem Biophys Res Commun. 2003;304:519–29. [Abstract] [Google Scholar]
90. Piko L, Taylor KD. Amounts of mitochondrial DNA and abundance of some mitochondrial gene transcripts in early mouse embryos. Dev Biol. 1987;123:364–74. [Abstract] [Google Scholar]
91. Lee HS, et al. Rapid mitochondrial DNA segregation in primate preimplantation embryos precedes somatic and germline bottleneck. Cell Rep. 2012;1:506–15. [Europe PMC free article] [Abstract] [Google Scholar]
92. Jenuth JP, Peterson AC, Shoubridge EA. Tissue-specific selection for different mtDNA genotypes in heteroplasmic mice. Nat Genet. 1997;16:93–5. [Abstract] [Google Scholar]
93. Battersby BJ, Shoubridge EA. Selection of a mtDNA sequence variant in hepatocytes of heteroplasmic mice is not due to differences in respiratory chain function or efficiency of replication. Hum Mol Genet. 2001;10:2469–79. [Abstract] [Google Scholar]
94. Bredenoord A, et al. Preimplantation genetic diagnosis for mitochondrial DNA disorders: ethical guidance for clinical practice. Eur J Hum Genet. 2009;17:1550–9. [Europe PMC free article] [Abstract] [Google Scholar]
95. Tachibana M, et al. Mitochondrial gene replacement in primate offspring and embryonic stem cells. Nature. 2009;461:367–72. [Europe PMC free article] [Abstract] [Google Scholar]
96. Craven L, et al. Pronuclear transfer in human embryos to prevent transmission of mitochondrial DNA disease. Nature. 2010;465:82–5. [Europe PMC free article] [Abstract] [Google Scholar]
97. Sharpley MS, et al. Heteroplasmy of mouse mtDNA is genetically unstable and results in altered behavior and cognition. Cell. 2012;151:333–43. [Europe PMC free article] [Abstract] [Google Scholar]
98. Boynton JE, Harris EH, Burkhart BD, Lamerson PM, Gillham NW. Transmission of mitochondrial and chloroplast genomes in crosses of Chlamydomonas. Proc Natl Acad Sci U S A. 1987;84:2391–5. [Europe PMC free article] [Abstract] [Google Scholar]
99. Schwartz M, Vissing J. Paternal inheritance of mitochondrial DNA. N Engl J Med. 2002;347:576–80. [Abstract] [Google Scholar]
100. DeLuca SZ, O’Farrell PH. Barriers to male transmission of mitochondrial DNA in sperm development. Dev Cell. 2012;22:660–8. [Europe PMC free article] [Abstract] [Google Scholar]
101. Hecht NB, Liem H, Kleene KC, Distel RJ, Ho SM. Maternal inheritance of the mouse mitochondrial genome is not mediated by a loss or gross alteration of the paternal mitochondrial DNA or by methylation of the oocyte mitochondrial DNA. Dev Biol. 1984;102:452–61. [Abstract] [Google Scholar]
102. Sutovsky P, et al. Ubiquitin tag for sperm mitochondria. Nature. 1999;402:371–2. [Abstract] [Google Scholar]
103. Al Rawi S, et al. Postfertilization autophagy of sperm organelles prevents paternal mitochondrial DNA transmission. Science. 2011;334:1144–7. [Abstract] [Google Scholar]
104. Sato M, Sato K. Degradation of paternal mitochondria by fertilization-triggered autophagy in C. elegans embryos. Science. 2011;334:1141–4. Together with reference 103, demonstrates a role for autophagy in the removal of paternal mitochondria in the early nematode embryo. [Abstract] [Google Scholar]
105. Zhou Q, Li H, Xue D. Elimination of paternal mitochondria through the lysosomal degradation pathway in C. elegans. Cell Res. 2011;21:1662–9. [Europe PMC free article] [Abstract] [Google Scholar]
106. Luo SM, et al. Unique insights into maternal mitochondrial inheritance in mice. Proc Natl Acad Sci U S A. 2013;110:13038–43. [Europe PMC free article] [Abstract] [Google Scholar]
107. Meeusen S, et al. Mitochondrial inner-membrane fusion and crista maintenance requires the dynamin-related GTPase Mgm1. Cell. 2006;127:383–95. [Abstract] [Google Scholar]
108. Song Z, Ghochani M, McCaffery JM, Frey TG, Chan DC. Mitofusins and OPA1 mediate sequential steps in mitochondrial membrane fusion. Mol Biol Cell. 2009;20:3525–32. [Europe PMC free article] [Abstract] [Google Scholar]
109. Loson OC, Song Z, Chen H, Chan DC. Fis1, Mff, MiD49, and MiD51 mediate Drp1 recruitment in mitochondrial fission. Mol Biol Cell. 2013;24:659–67. [Europe PMC free article] [Abstract] [Google Scholar]
110. Otera H, et al. Mff is an essential factor for mitochondrial recruitment of Drp1 during mitochondrial fission in mammalian cells. J Cell Biol. 2010;191:1141–58. [Europe PMC free article] [Abstract] [Google Scholar]
111. Rafelski SM, et al. Mitochondrial network size scaling in budding yeast. Science. 2012;338:822–4. [Europe PMC free article] [Abstract] [Google Scholar]
112. McConnell SJ, Stewart LC, Talin A, Yaffe MP. Temperature-sensitive yeast mutants defective in mitochondrial inheritance. J Cell Biol. 1990;111:967–76. [Europe PMC free article] [Abstract] [Google Scholar]
113. Evangelista M, Pruyne D, Amberg DC, Boone C, Bretscher A. Formins direct Arp2/3-independent actin filament assembly to polarize cell growth in yeast. Nat Cell Biol. 2002;4:260–9. [Abstract] [Google Scholar]
114. Fagarasanu A, Mast FD, Knoblach B, Rachubinski RA. Molecular mechanisms of organelle inheritance: lessons from peroxisomes in yeast. Nat Rev Mol Cell Biol. 2010;11:644–54. [Abstract] [Google Scholar]
115. Lazzarino DA, Boldogh I, Smith MG, Rosand J, Pon LA. Yeast mitochondria contain ATP-sensitive, reversible actin-binding activity. Mol Biol Cell. 1994;5:807–18. [Europe PMC free article] [Abstract] [Google Scholar]
116. Drubin DG, Jones HD, Wertman KF. Actin structure and function: roles in mitochondrial organization and morphogenesis in budding yeast and identification of the phalloidin-binding site. Mol Biol Cell. 1993;4:1277–94. [Europe PMC free article] [Abstract] [Google Scholar]
117. Boldogh IR, et al. Arp2/3 complex and actin dynamics are required for actin-based mitochondrial motility in yeast. Proc Natl Acad Sci U S A. 2001;98:3162–7. [Europe PMC free article] [Abstract] [Google Scholar]
118. Boldogh IR, Ramcharan SL, Yang HC, Pon LA. A type V myosin (Myo2p) and a Rab-like G-protein (Ypt11p) are required for retention of newly inherited mitochondria in yeast cells during cell division. Mol Biol Cell. 2004;15:3994–4002. [Europe PMC free article] [Abstract] [Google Scholar]
119. Lambrechts A, Gevaert K, Cossart P, Vandekerckhove J, Van Troys M. Listeria comet tails: the actin-based motility machinery at work. Trends Cell Biol. 2008;18:220–7. [Abstract] [Google Scholar]
120. Altmann K, Frank M, Neumann D, Jakobs S, Westermann B. The class V myosin motor protein, Myo2, plays a major role in mitochondrial motility in Saccharomyces cerevisiae. J Cell Biol. 2008;181:119–30. Identifies a myosin family member, Myo2, as a mitochondrial motor protein. [Europe PMC free article] [Abstract] [Google Scholar]
121. Chernyakov I, Santiago-Tirado F, Bretscher A. Active segregation of yeast mitochondria by Myo2 is essential and mediated by Mmr1 and Ypt11. Curr Biol. 2013;23:1818–24. [Europe PMC free article] [Abstract] [Google Scholar]
122. Fortsch J, Hummel E, Krist M, Westermann B. The myosin-related motor protein Myo2 is an essential mediator of bud-directed mitochondrial movement in yeast. J Cell Biol. 2011;194:473–88. [Europe PMC free article] [Abstract] [Google Scholar]
123. Ishikawa K, et al. Identification of an organelle-specific myosin V receptor. J Cell Biol. 2003;160:887–97. [Europe PMC free article] [Abstract] [Google Scholar]
124. Tang F, et al. Regulated degradation of a class V myosin receptor directs movement of the yeast vacuole. Nature. 2003;422:87–92. [Abstract] [Google Scholar]
125. Govindan B, Bowser R, Novick P. The role of Myo2, a yeast class V myosin, in vesicular transport. J Cell Biol. 1995;128:1055–68. [Europe PMC free article] [Abstract] [Google Scholar]
126. Pruyne DW, Schott DH, Bretscher A. Tropomyosin-containing actin cables direct the Myo2p-dependent polarized delivery of secretory vesicles in budding yeast. J Cell Biol. 1998;143:1931–45. [Abstract] [Google Scholar]
127. Rossanese OW, et al. A role for actin, Cdc1p, and Myo2p in the inheritance of late Golgi elements in Saccharomyces cerevisiae. J Cell Biol. 2001;153:47–62. [Europe PMC free article] [Abstract] [Google Scholar]
128. Hoepfner D, van den Berg M, Philippsen P, Tabak HF, Hettema EH. A role for Vps1p, actin, and the Myo2p motor in peroxisome abundance and inheritance in Saccharomyces cerevisiae. J Cell Biol. 2001;155:979–90. [Europe PMC free article] [Abstract] [Google Scholar]
129. Itoh T, Toh EA, Matsui Y. Mmr1p is a mitochondrial factor for Myo2p-dependent inheritance of mitochondria in the budding yeast. EMBO J. 2004;23:2520–30. [Europe PMC free article] [Abstract] [Google Scholar]
130. Itoh T, Watabe A, Toh EA, Matsui Y. Complex formation with Ypt11p, a rab-type small GTPase, is essential to facilitate the function of Myo2p, a class V myosin, in mitochondrial distribution in Saccharomyces cerevisiae. Mol Cell Biol. 2002;22:7744–57. [Europe PMC free article] [Abstract] [Google Scholar]
131. Kelleher JF, et al. Myosin VI is required for asymmetric segregation of cellular components during C. elegans spermatogenesis. Curr Biol. 2000;10:1489–96. [Abstract] [Google Scholar]
132. Klecker T, Scholz D, Fortsch J, Westermann B. The yeast cell cortical protein Num1 integrates mitochondrial dynamics into cellular architecture. J Cell Sci. 2013;126:2924–30. [Abstract] [Google Scholar]
133. Lackner LL, Ping H, Graef M, Murley A, Nunnari J. Endoplasmic reticulum-associated mitochondria-cortex tether functions in the distribution and inheritance of mitochondria. Proc Natl Acad Sci U S A. 2013;110:E458–67. [Europe PMC free article] [Abstract] [Google Scholar]
134. Rizzuto R, Brini M, Murgia M, Pozzan T. Microdomains with high Ca2+ close to IP3-sensitive channels that are sensed by neighboring mitochondria. Science. 1993;262:744–7. [Abstract] [Google Scholar]
135. Achleitner G, et al. Association between the endoplasmic reticulum and mitochondria of yeast facilitates interorganelle transport of phospholipids through membrane contact. Eur J Biochem. 1999;264:545–53. [Abstract] [Google Scholar]
136. Rusinol AE, Cui Z, Chen MH, Vance JE. A unique mitochondria-associated membrane fraction from rat liver has a high capacity for lipid synthesis and contains pre-Golgi secretory proteins including nascent lipoproteins. J Biol Chem. 1994;269:27494–502. [Abstract] [Google Scholar]
137. Swayne TC, et al. Role for cER and Mmr1p in anchorage of mitochondria at sites of polarized surface growth in budding yeast. Curr Biol. 2011;21:1994–9. [Europe PMC free article] [Abstract] [Google Scholar]
138. Kornmann B, et al. An ER-mitochondria tethering complex revealed by a synthetic biology screen. Science. 2009;325:477–81. Reports the use of a genetic synthetic biology screen to identify components of the ERMES complex in S.cerevisiae. [Europe PMC free article] [Abstract] [Google Scholar]
139. Bockler S, Westermann B. Mitochondrial ER contacts are crucial for mitophagy in yeast. Dev Cell. 2014;28:450–8. [Abstract] [Google Scholar]
140. Murley A, et al. ER-associated mitochondrial division links the distribution of mitochondria and mitochondrial DNA in yeast. Elife. 2013;2:e00422. Provides a model linking mitochondrial division with inheritance of mtDNA nucleoids in yeast. [Europe PMC free article] [Abstract] [Google Scholar]
141. Ban-Ishihara R, Ishihara T, Sasaki N, Mihara K, Ishihara N. Dynamics of nucleoid structure regulated by mitochondrial fission contributes to cristae reformation and release of cytochrome c. Proc Natl Acad Sci U S A. 2013;110:11863–8. [Europe PMC free article] [Abstract] [Google Scholar]
142. Berger KH, Sogo LF, Yaffe MP. Mdm12p, a component required for mitochondrial inheritance that is conserved between budding and fission yeast. J Cell Biol. 1997;136:545–53. [Europe PMC free article] [Abstract] [Google Scholar]
143. Sogo LF, Yaffe MP. Regulation of mitochondrial morphology and inheritance by Mdm10p, a protein of the mitochondrial outer membrane. J Cell Biol. 1994;126:1361–73. [Europe PMC free article] [Abstract] [Google Scholar]

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/2691911
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/2691911

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1038/nrm3877

Supporting
Mentioning
Contrasting
11
690
0

Article citations


Go to all (537) article citations

Data 


Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.

Funding 


Funders who supported this work.

Howard Hughes Medical Institute

    NIGMS NIH HHS (4)