2000px-Animal_mitochondrion_diagram_en.svg.png

Nature Reviews Molecular Cell Biology (2019)

 

Abstract

Mitochondria are essential for the viability of eukaryotic cells as they perform crucial functions in bioenergetics, metabolism and signalling and have been associated with numerous diseases. Recent functional and proteomic studies have revealed the remarkable complexity of mitochondrial protein organization. Protein machineries with diverse functions such as protein translocation, respiration, metabolite transport, protein quality control and the control of membrane architecture interact with each other in dynamic networks. In this Review, we discuss the emerging role of the mitochondrial protein import machinery as a key organizer of these mitochondrial protein networks. The preprotein translocases that reside on the mitochondrial membranes not only function during organelle biogenesis to deliver newly synthesized proteins to their final mitochondrial destination but also cooperate with numerous other mitochondrial protein complexes that perform a wide range of functions. Moreover, these protein networks form membrane contact sites, for example, with the endoplasmic reticulum, that are key for integration of mitochondria with cellular function, and defects in protein import can lead to diseases.

Introduction

Mitochondria are a hallmark and the powerhouses of eukaryotic cells; they synthesize ATP via oxidative phosphorylation but are also deeply integrated into cellular metabolism and signalling pathways.

Mitochondria consist of two membranes and two aqueous compartments (Fig. 1). The surface area of the mitochondrial inner membrane is several-fold larger than that of the outer membrane; it therefore forms invaginations known as cristae, which contain the oxidative phosphorylation system, comprising the respiratory complexes I to IV and the F1F0-ATP synthase for ATP production. Only a small set of proteins are encoded by the mitochondrial genome, and these are typically hydrophobic membrane proteins that form core parts of the oxidative phosphorylation complexes of the mitochondrial inner membrane. Approximately 99% of mitochondrial proteins are encoded by nuclear genes and depend on specific targeting signals that direct them from the cytosol, where they are synthesized, to mitochondrial surface receptors and then into the proper mitochondrial subcompartments1,2.

41580_2018_92_Fig1_HTML.png

Mitochondria consist of four compartments: outer membrane (OM), intermembrane space (IMS), inner membrane (IM) and matrix. A large variety of functions have been assigned to mitochondrial proteins and protein complexes and are indicated in the figure: energy metabolism with respiration and synthesis of ATP; metabolism of amino acids, lipids and nucleotides; biosynthesis of iron–sulfur (Fe–S) clusters and cofactors; expression of the mitochondrial genome; quality control and degradation processes including mitophagy and apoptosis; signalling and redox processes; membrane architecture and dynamics; and the import and processing of precursor proteins that are synthesized on cytosolic ribosomes. AAA, ATP-dependent proteases of the inner membrane; E3, ubiquitin-protein ligase; ER, endoplasmic reticulum; mtDNA, mitochondrial DNA; TCA, tricarboxylic acid; Ub, ubiquitin.

Traditionally, research on mitochondria focused on bioenergetics, but studies in the past 15–20 years have revealed a greater than expected complexity and versatility of mitochondrial activities, integrating mitochondrial energetics with protein biogenesis, metabolic pathways, cellular signalling, stress responses and apoptosis. It is becoming increasingly evident that mitochondrial protein machineries, which have diverse functions, are physically and functionally connected rather than functioning as independent units. It is therefore important to understand the principles that govern the formation and maintenance of these complex networks.

In this Review, we first discuss recent proteomic studies that unveiled a large spectrum of mitochondrial protein types and functions. Quantitative proteomics revealed for the first time absolute numbers for the cellular abundance of all important mitochondrial machineries under respiratory and non-respiratory conditions. We then discuss how preprotein translocases are at the core of dynamic protein networks that link organelle biogenesis to energy metabolism, membrane morphology and dynamics. With this integrative view, we discuss how the mitochondrial protein import machinery is connected to mitochondrial stress responses, quality control mechanisms and diseases and discuss its role in the formation of membrane contact sites between mitochondria and other organelles, which are important for cell function.

Multifunctional mitochondria

Systematic analyses of the mitochondrial proteome have provided a comprehensive overview of the mitochondrial protein complement and of the large variety of functions performed by mitochondria (Fig. 1). Importantly, our understanding of mitochondrial activities has been shaped by the recent absolute quantification of the mitochondrial proteome.

Numerous and diverse functions of mitochondria

Textbooks typically describe mitochondria as comprising the respiratory complexes and the F1F0-ATP synthase in the inner membrane cristae and transporters and channels for metabolites and ions in both mitochondrial membranes and as the site of metabolic pathways, which are mainly localized to the matrix and inner membrane3 (Fig. 1). These include energy metabolism pathways, such as the tricarboxylic acid cycle, as well as amino acid, lipid and nucleotide metabolism pathways. Electrons derived from the oxidation of metabolites are fed into the respiratory chain, which generates an electrochemical gradient by pumping protons from the mitochondrial matrix into the intermembrane space. The resulting proton gradient is used to drive ATP synthesis by the F1F0-ATP synthase and to enable the import of precursor proteins and the transfer of some metabolites across the inner membrane.

Functions of mitochondria, which are essential for cell viability under all growth conditions, include the synthesis of iron–sulfur (Fe–S) clusters4 and mitochondrial protein import and maturation1,2. The mitochondrial matrix also contains a complete genetic system, which includes the mitochondrial genome, numerous factors that are crucial for the maintenance and regulation of the genome and mitochondrial ribosomes, which differ in size and composition from cytosolic ribosomes5. Proteins encoded by the mitochondrial genome are inserted into the inner membrane in a co-translational mechanism by coupling translating ribosomes to the oxidase assembly (OXA) insertase1,2,6.

Mitochondria form a dynamic network in most cell types that is continuously remodelled by fusion and fission of the organelles7,8. Several machineries have been identified that control mitochondrial membrane architecture and dynamics, which include factors that mediate fusion or fission, such as dynamin-related GTPases located at the outer and inner membranes, and membrane-shaping components. The mitochondrial contact site and cristae organizing system (MICOS) is crucial for maintaining the characteristic shape of inner membrane cristae9,10,11. MICOS and several other protein complexes form contact sites between the mitochondrial outer and inner membranes to promote the transfer of proteins, lipids and metabolites9,10,12,13.

Topics that are being intensively studied include quality control systems and the regulation of mitochondrial signalling (all of which have been recently reviewed14,15,16,17,18,19). Numerous cytosolic signalling cascades are connected to mitochondria under physiological and pathophysiological conditions. The metabolic activity of the organelle serves as a measure for mitochondrial fitness and quality14, and elaborate pathways for mitochondrial stress responses, selective degradation of damaged mitochondria by autophagy (mitophagy)15,16and programmed cell death (apoptosis) via mitochondria have been identified17. Mitochondria contain a set of internal proteases that are also involved in quality control and turnover of mitochondrial proteins18. In addition, mitochondria are a major site of cellular production of reactive oxygen species (ROS) and contain numerous redox pathways19.

Quantitative analysis of the mitochondrial proteome

The wide range of mitochondrial activities is challenging to study and has led to different views on how mitochondria are organized. We can envisage three stages as having led to our current understanding of mitochondrial functions. First, the original research on metabolism and ATP production established energetics and metabolism as hallmarks of mitochondria. Second, systematic proteomic studies on mitochondria, which began ~15 years ago, led to the identification of many new mitochondrial proteins. We currently estimate that mitochondria contain at least 1,000 (in yeast) to 1,500 (in humans) different proteins20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38. The functional classification of this high number of different proteins26,32 indicated that only up to 15% of the mitochondrial proteins are directly involved in energy metabolism, including the energy metabolizing pathways and all structural subunits of the oxidative phosphorylation system. It is estimated that 20–25% of the mitochondrial proteome maintains and regulates the mitochondrial genome, which encodes only ~1% of mitochondrial proteins. In addition, numerous mitochondrial components were found to be connected to a variety of cellular signalling pathways and membrane dynamics. Thus, mitochondria emerged as signalling platforms that are crucial in the regulation of cell function, extending their role well beyond that of cellular powerhouses. Lastly, a more refined understanding of the nature and functions of mitochondrial proteins in different conditions became possible with the recent systematic quantification of the majority of the mitochondrial proteome, which yielded the absolute copy numbers of mitochondrial proteins present in distinct cells26,39,40,41,42,43. Proteins involved in energy metabolism are by far the most abundant protein classes in respiring yeast mitochondria (Box 1). The ~15% of mitochondrial proteins with a direct role in energy metabolism and respiration mentioned above constitute more than half of the mitochondrial protein mass under respiratory conditions, reinforcing the original view of mitochondria as cellular powerhouses. Taking the various metabolic processes, oxidative phosphorylation and metabolite carriers and channels together, ~75% of the protein mass of respiring mitochondria is dedicated to metabolism and bioenergetics26.

These seemingly controversial views, of mitochondria as powerhouses versus mitochondria as organelles with a large number of different functions, must be combined to understand the cellular importance of mitochondria. On the basis of absolute protein mass —that is, from a quantitative perspective — metabolism and bioenergetics remain the major tasks of mitochondria (the powerhouses). However, other functions are of central importance for mitochondrial fitness, cellular growth and development. A striking example is the system for Fe–S cluster biosynthesis, which makes up <1% of the protein mass of respiring mitochondria but is essential for the viability of all eukaryotic cells4,26. Equally important are the machineries for mitochondrial membrane morphology and dynamics, which represent <1% of the mitochondrial protein mass26. These proteins have key roles in mitochondrial architecture, fusion and fission and are thus crucial for maintaining and remodelling the mitochondrial network under different growth conditions7,8. Mitochondria can thus be seen as super-powerhouses that, in addition to their predominant metabolic and energetic functions, are deeply integrated into cellular dynamics, signalling and biosynthetic pathways by performing a multitude of functions. We discuss below that these functions are not independent of one another, as the machineries and proteins involved are physically connected in large, dynamic networks.

Box 1 The mitochondrial proteome: from fermentation to respiration

Recent studies have revealed the absolute copy numbers of mitochondrial proteins per cell26,39,40,41,42,43 (see the figure). Proteins involved in energy metabolism are by far the most abundant classes of mitochondrial proteins26. The abundance of these proteins is strongly regulated by the growth conditions. When shifting yeast from fermentation to respiratory conditions, their protein levels are increased approximately threefold or more. The levels of proteins involved in signalling, redox processes, membrane dynamics and morphology are increased approximately twofold from fermentation to respiration, which is comparable to the increase in the total mitochondrial protein mass. Interestingly, proteins involved in two essential processes — protein import, maturation and turnover, and biosynthesis of iron–sulfur (Fe–S) clusters — are only mildly affected when cells are shifted from fermentation to respiration26. The overall abundance of proteins involved in mitochondrial gene expression and translation and in the metabolism of lipids, amino acids and nucleotides is only mildly affected by different growth conditions; however, differences can be observed for individual classes of protein. For example, proteins involved in lipid metabolism are considerably more expressed under respiratory conditions, whereas proteins involved in amino acid metabolism are more strongly expressed under fermentable conditions26.

41580_2018_92_Figa_HTML.png

Plasticity of the mitochondrial proteome

The mitochondrial content of a cell can vary considerably under different growth conditions and between different organisms and tissues21,24,26,27,28,44. Systematic analyses of yeast mitochondria revealed that the total mitochondrial protein mass constitutes ~9% of the cellular protein mass in fermentable conditions (when a lower activity of mitochondria is required) but is more than double in respiratory growth conditions, reaching ~20% of the cellular protein mass26. The changes from fermentation to respiration are quite different for mitochondrial proteins belonging to different functional classes (Box 1). There is more than a threefold increase in the absolute copy number per cell for proteins directly involved in energy metabolism and respiration. Proteins functioning in signalling, redox processes and membrane dynamics are increased by approximately twofold, like the overall increase in mitochondrial mass26.

Remarkably, the protein classes that are involved in protein biogenesis and folding and biosynthesis of Fe–S clusters — the two mitochondrial processes that are essential for cell viability in all cell types and growth conditions — are only moderately altered in their overall copy numbers when yeast cells transition from fermentation to respiration26. This indicates that cells are well equipped for these systems in non-respiratory conditions and require only a slight increase in their abundance during respiration. For example, the protein import machinery imports more than double the amount of proteins during respiration, and so it is evident that the machinery works at a considerably lower capacity during fermentation. We conclude that the protein import machinery and the system for biosynthesis of Fe–S clusters are essential housekeeping systems of mitochondria and are not or are only moderately regulated by their copy number. Indeed, studies on the translocase of the outer membrane (TOM) complex — which is the main protein entry gate, consisting of receptor proteins on the cytosolic face (Tom70, Tom20 and Tom22) and a pore-forming protein (Tom40) — revealed that at least four different cytosolic signalling systems regulate TOM complex activity by phosphorylation, leading to a sophisticated pattern of stimulatory and inhibitory effects depending on the kinase and TOM subunits involved45,46,47,48. Having a stable set of housekeeping systems that are regulated by reversible modification, such as phosphorylation, has the advantage of enabling rapid responses to changing conditions. The system would be rather slow in adapting to different requirements if it was dependent on increased gene expression, translation and import. Similarly, inhibition is faster when achieved by covalent modification rather than by reducing protein copy numbers. For example, during fermentable growth, when less translocation of metabolites in and out of mitochondria is needed, the activity of Tom70, which is required for the import of metabolite carriers, is inhibited by phosphorylation, leading to an immediate decrease in metabolite carrier import into mitochondria48. Thus, the largely stable protein copy numbers in housekeeping systems and their regulation by post-translational modification enable greater flexibility during mitochondrial biogenesis and remodelling of the mitochondrial network.

Assembly of functional protein networks

Studies of protein import from the cytosol into mitochondria were originally based on the assumption that all proteins were transported to their final destination along one central pathway. However, the characterization of precursor proteins carrying different targeting signals revealed that mitochondria use at least five major protein import pathways, each one directed by a different type of targeting signal. The complexity of the system is even higher, as preprotein translocases do not operate as isolated units but are connected to numerous mitochondrial protein complexes involved in seemingly unrelated functions.

Five major import pathways of precursor proteins into mitochondria

The presequence pathway is the best-characterized pathway, responsible for the import of ~60% of all mitochondrial proteins49. The precursor proteins carry amino-terminal targeting signals, termed presequences, which form positively charged amphipathic α-helices. These presequences are typically recognized by the TOM receptors Tom20 and Tom22 (refs50,51) on the mitochondrial surface and are transported through the main protein translocation channel of the outer membrane Tom40 (refs1,2,52,53) (Fig. 2). Upon passage across the outer membrane, the preproteins are engaged by the presequence translocase of the inner membrane (TIM23), which directs their transfer across the inner membrane54,55,56,57. The membrane potential (Δψ) across the inner membrane (negative on the inside) activates the TIM23 channel and drives the positively charged presequences towards the matrix58,59,60,61. The presequence translocase-associated motor (PAM) contains the mitochondrial heat shock protein 70 (mtHsp70) as the central ATP-driven chaperone, which binds the precursor proteins to promote their unidirectional movement62. Together with five co-chaperones, mtHsp70 translocates the polypeptide chain into the matrix63,64,65,66,67,68, where the presequences are removed by the mitochondrial processing peptidase (MPP)1,2,69. The matrix contains additional processing enzymes involved in quality control functions, as mitochondrial proteins can be degraded by the N-end rule pathwaydepending on whether they contain stabilizing or destabilizing amino acid residues at their amino termini49,70. The intermediate cleaving peptidase of 55 kDa (Icp55) and the octapeptidyl peptidase (Oct1) remove destabilizing amino acid residues from the amino termini of imported proteins after cleavage by MPP, thus generating amino termini that are less prone to degradation by matrix proteases49,71,72. With the help of mtHsp70 and other chaperones such as the Hsp60–Hsp10 chaperonin complex73,74, the proteins are then folded into their active form.

41580_2018_92_Fig2_HTML.png

Five major pathways of mitochondrial protein import have been identified. The protein import machineries have been well conserved from fungi (shown in this figure) to mammals (shown in Box 4). First, the presequence pathway transports presequence-carrying cleavable preproteins through the translocase of the outer membrane (TOM) and the presequence translocase of the inner membrane (TIM23) with the presequence translocase-associated motor (PAM). The membrane potential (Δψ) across the inner membrane (IM) activates the TIM23 channel and drives translocation of the positively charged presequences into the matrix. The presequences are removed by the mitochondrial processing peptidase (MPP), and additional proteolytic processing can occur by intermediate cleaving peptidase of 55 kDa (Icp55) or octapeptidyl peptidase (Oct1). IM proteins are either laterally released from the TIM23 complex or are transported via the matrix and inserted into the IM by the oxidase assembly protein 1 (Oxa1) insertase. IM proteins synthesized on mitochondrial ribosomes are also inserted by Oxa1. Second, cysteine-rich proteins destined for the intermembrane space (IMS) are imported through the TOM complex and are recognized by the mitochondrial IMS import and assembly protein (Mia40), which functions as an oxidoreductase to insert disulfide bonds into the imported proteins. The sulfhydryl oxidase Erv1 forms a disulfide relay with Mia40, transferring disulfides from Erv1 to Mia40 to imported proteins. Third, the precursors of non-cleavable IM proteins such as the carrier proteins are imported by the TOM complex, followed by transfer to the small TIM chaperones in the IMS and insertion into the IM by the TIM22 carrier translocase. Fourth, the precursors of outer membrane (OM) β-barrel proteins use the TOM complex and small TIM chaperones and are inserted into the OM by the sorting and assembly machinery (SAM). Fifth, many OM proteins with α-helical transmembrane segments are inserted into the membrane by the mitochondrial import (MIM) complex. α-Helical OM proteins typically do not use the Tom40 channel, but Tom70 can be involved in their recognition. The inset shows the absolute copy numbers of characteristic translocase components in a respiring yeast cell and, for comparison, the abundance of respiratory complexes, metabolite channels and carriers of the mitochondrial membranes26. The porin isoform porin 1 (Por1) of the OM is one of the most abundant mitochondrial proteins, whereas the isoform Por2 is one of the least abundant proteins. mtHsp70, mitochondrial heat shock protein 70.

 

Presequence-carrying precursors that become integrated into the inner mitochondrial membrane follow two distinct routes (Fig. 2). Those that have a hydrophobic sorting signal behind the matrix targeting signal become arrested in the TIM23 complex and are then released by the lateral gatekeeper Mgr2 into the inner membrane (stop transfer pathway)75,76. Other inner membrane proteins are first transported into the matrix and are incorporated into the inner membrane by the OXA insertase, which is also used by mitochondrially encoded proteins (conservative sorting)6,77,78,79,80.

Most other protein import pathways also use the TOM channel for preprotein translocation across the outer membrane81,82 (Fig. 2), but the dependence on the three TOM receptors Tom20, Tom22 and Tom70 and the mode of delivery from the cytosol to the TOM complex can differ83,84,85. The carrier pathway is dedicated to the import of hydrophobic multi-spanning inner membrane proteins. These proteins do not have amino-terminal presequences; they have internal targeting signals that contain hydrophobic elements but have not been fully characterized. Cytosolic chaperones of the Hsp90 and Hsp70 classes deliver these inner membrane precursor proteins to the receptor Tom70 (ref.86). After their release from the cytosolic chaperones, the precursors pass through the Tom40 channel in a loop formation and enter the intermembrane space87,88. Here, they are bound by small TIM chaperones, which prevent their aggregation89,90,91,92 and guide them to the carrier translocase of the inner membrane (TIM22) complex. The TIM22 complex operates in a Δψ-dependent manner to insert these multi-spanning proteins in the inner membrane93,94,95,96,97 (Fig. 2).

Many proteins of the mitochondrial intermembrane space contain characteristic cysteine motifs that become oxidized to form stabilizing disulfide bonds in the mature proteins. The mitochondrial intermembrane space import and assembly (MIA) system, which mediates the import and oxidative folding of intermembrane space proteins, consists of two main components: the oxidoreductase Mia40 (refs98,99) and the sulfhydryl oxidase Erv1 (ref.100) (Fig. 2). Upon passage through the Tom40 channel, Mia40 recognizes the precursor proteins that contain an intermembrane space sorting signal, typically consisting of a hydrophobic element flanked by a cysteine residue101,102,103. Mia40 forms transient disulfide bonds with these precursors and then transfers the disulfide bonds to them, which leads to intramolecular disulfide bond formation via oxidation and stabilization of the proteins104,105. At each transfer of a disulfide bond to a protein, cysteines of Mia40 become reduced and are re-oxidized by Erv1. In this disulfide relay, disulfide bonds are thus transferred from Erv1 to Mia40 to imported proteins.

The mitochondrial outer membrane contains different classes of membrane protein: single-spanning and multi-spanning proteins with α-helical transmembrane segments and β-barrel proteins. The precursors of β-barrel proteins are initially translocated by the TOM complex106 to the intermembrane space and interact with small TIM chaperones like the carrier precursors92 (Fig. 2). Insertion of β-barrel precursors into the outer membrane is mediated by the sorting and assembly machinery (SAM)107,108,109 in a step-wise process that involves translocation into the SAM channel (formed by the Sam50 subunit) and lateral release into the lipid phase of the membrane110. The carboxy-terminal β-strand of these proteins functions as a β-signal that directs insertion via SAM111. α-Helical outer membrane proteins usually follow distinct import routes that do not involve the Tom40 channel. The sorting signal is typically contained within the α-helical transmembrane segments and flanking positively charged amino acid residues. Single-spanning proteins with an amino-terminal membrane anchor (signal-anchored proteins) and multi-spanning outer membrane proteins can use the mitochondrial import (MIM) channel for membrane insertion, assisted by Tom70 at least in the case of multi-spanning proteins112,113,114,115,116,117. In the case of single-spanning proteins with a carboxy-terminal membrane anchor (tail-anchored proteins) and some multi-spanning proteins, the lipid composition of the outer membrane seems to be important for membrane insertion, yet the exact molecular mechanism is unknown118,119,120,121. Several possibilities have been proposed, including protein-independent direct insertion into the phospholipid membrane, MIM complex-assisted insertion or the involvement of an unknown proteinaceous insertase of the outer membrane.

Abundance and versatility of import machineries

The absolute copy numbers of a variety of mitochondrial import components have been identified26 (Fig. 2). The TOM complex is the most abundant translocase, consistent with its role in feeding precursors into at least four distinct downstream translocase systems. The TIM23 complex is also abundant, as expected given the major role of the presequence import pathway. Interestingly, mtHsp70 is approximately ten times more abundant than Tim23 and other motor subunits such as Tim44. mtHsp70 plays a dual role: a small fraction of mtHsp70 molecules act in the TIM23-associated PAM to drive preprotein import, whereas the majority of mtHsp70 is dedicated to protein folding in the mitochondrial matrix2. Tim22 and Sam50 are present in quite low amounts. However, their major substrates are of high abundance, which underscores the importance and activity of these translocases. Inner membrane metabolite carriers, the substrates of the TIM22 complex, are highly abundant, and of note, one of the most abundant mitochondrial proteins, the outer membrane β-barrel metabolite channel porin 1 (Por1), is a substrate of the SAM complex. Thus, the TIM22 and SAM complexes are crucial for the biogenesis of mitochondrial metabolite carriers and channels that mediate the export of ATP and link mitochondrial and cellular metabolism.

The TOM complexes can form different types of dynamic supercomplex: a TOM–SAM supercomplex for efficient transfer of β-barrel precursors and a two-membrane-spanning TOM–TIM23–preprotein supercomplex122,123,124,125,126,127. TOM also interacts with the small TIM chaperones, and in mammals, TIM29 of the TIM22 complex was found to associate with TOM82,128,129. The differential abundance of the translocases (Fig. 2) indicates that TOM complexes are sufficiently abundant to form the different supercomplexes. Whether separate pools of TOM complexes exist for different import pathways or whether the TOM complexes are freely interchangeable in one large dynamic pool is currently a subject of debate. The translocation of several intermembrane space precursor proteins across the outer membrane depends on Tom40 but does not require the TOM receptor domains130,131. Competition experiments suggest that intermembrane space precursors and presequence-carrying precursors do not use the same TOM complexes, which is in support of the view that there may be distinct pools of the TOM complex130. We speculate that in addition to the full-size TOM complex, which comprises the Tom40 channels, all three receptors and three small Tom proteins82, mitochondria might also contain simpler forms of TOM complex. These simpler TOM complexes may just contain the Tom40 channel and possibly some of the small Tom subunits and may be dedicated, for example, to the import of intermembrane space precursors that are recognized by Mia40 and do not depend on classical TOM receptors. The differential phosphorylation of TOM complexes by cytosolic signalling cascades also contributes to the heterogeneity of TOM complexes and suggests that cellular signalling pathways control the activity of distinct import routes45,46,47,48.

In metazoans, there are two forms of the presequence translocase, which are differentially distributed across tissues: one form contains the stably expressed housekeeping subunit TIM17B, found in skeletal muscle, and the other form contains the stress-regulated subunit TIM17A, found in the brain132. Under stress conditions, the levels of TIM17A decrease as a result of reduced synthesis and increased degradation by the ATP-dependent AAA protease of the inner membrane that is exposed to the intermembrane space (iAAA protease) (Fig. 3). This decrease in TIM17A levels promotes a mitochondrial unfolded protein response (UPRmt)133. Thus, whereas the overall abundance of the mitochondrial protein import machinery in rapidly growing unicellular organisms like yeast is quite stable under different metabolic conditions, the abundance of tissue-specific isoforms varies in metazoans, suggesting that additional regulatory mechanisms operate in multicellular organisms132,133,134.

41580_2018_92_Fig3_HTML.png

Supercomplexes of the mitochondrial respiratory chain are integrated into functional networks with the presequence translocase of the inner membrane (TIM23) (see Box 2) and the ATP-dependent AAA proteases of the inner membrane (IM). The ATP-dependent AAA proteases degrade not only several IM proteins but also selected proteins of the matrix, intermembrane space (IMS) and outer membrane (OM), functioning as a quality control system of mitochondria. Several respiratory chain–AAA linker proteins, AAA–substrate adaptor proteins and assembly factors for respiratory supercomplexes were identified in fungi. The coenzyme Q (CoQ) biosynthetic complex on the matrix side of the IM provides CoQ for the respiratory chain and further enzymes. The precursors of the CoQ complex are imported by the translocase of the outer membrane (TOM) and TIM machineries. Proteolytic processing in the matrix can involve two steps, as it does for the precursor of Coq5. The mitochondrial respiratory chain is a main source for the generation of reactive oxygen species (ROS), which can exert harmful effects but also function in signalling. Targeting of the cytosolic translation machinery by ROS leads to decreased protein synthesis, providing a link between the status of the respiratory chain and protein biogenesis. ΔμH+, electrochemical proton gradient; CI, complex I; CII, complex II; CIII, complex III; CIV, complex IV; iAAA, IMS-exposed AAA; mAAA, matrix-exposed AAA; MPP, mitochondrial processing peptidase; Oct1, octapeptidyl peptidase; PAM, presequence translocase-associated motor.

 

Elements of different import pathways can be combined to create new pathways. For example, a large domain of the single-spanning outer membrane protein Om45 is exposed to the intermembrane space. The Om45 precursor is first transported by the presequence pathway through the TOM complex and interacts with the TIM23 complex but then escapes into the intermembrane space to be inserted into the outer membrane by a topologically opposite action of the MIM complex135,136. Cleavable carboxy-terminal targeting signals are another example of the versatility of the mitochondrial import machinery, as they are imported via the presequence pathway and removed by matrix or intermembrane space peptidases, followed by differential sorting to intramitochondrial destinations137,138,139. Thus, although import pathways are classified into five major pathways, import mechanisms are much more diverse and versatile, and we expect that the systematic analysis of the large number of substrates will uncover new import routes and possibly also new translocases.

Respiratory chain interactions link bioenergetics, biogenesis and quality control

The respiratory chain complexes of the mitochondrial inner membrane are the core of a large protein network that connects bioenergetics to mitochondrial biogenesis, regulation and turnover processes (Fig. 3). Respiratory complexes I (NADH:ubiquinone oxidoreductase), III (cytochrome c reductase) and IV (cytochrome c oxidase) assemble into large I–III–IV supercomplexes, also termed respirasomes140,141,142,143,144. The assembly into such supercomplexes is now generally accepted, but there are different views about their functions141,143. The supercomplexes may influence the assembly and stability of respiratory complexes, regulate the activity of the complexes and/or reduce the formation of ROS. Various factors involved in the formation of respiratory supercomplexes have been reported (Fig. 3), but whether they mainly function in the assembly of individual respiratory complexes or in the formation of supercomplexes remains to be elucidated141,143,145,146,147,148,149.

Not only are preprotein translocases in crosstalk with each other, they also form physical contacts with other mitochondrial machineries, including machineries involved in mitochondrial energy metabolism. The TIM23 complex forms a hub in the sorting of preproteins at the inner membrane cooperating with import complexes TOM and PAM and forms supercomplexes with the respiratory complexes III and IV as well as with the ADP/ATP carrier (Box 2; Fig. 3). These interactions of the TIM23 complex support protein import under energy-limiting conditions150,151,152 and can also promote the assembly of respiratory complexes153,154,155 (Box 2). The respiratory complexes as well as the ADP/ATP carrier are several-fold more abundant than the TIM23 complex26 (Fig. 2), and thus, only a fraction of them are engaged in the interaction with TIM23. Respiratory complexes are preferentially located in cristae membranes, yet a smaller fraction is found in the inner boundary membrane, which is adjacent to the outer membrane156, and can thus interact with the TIM23 complexes. TIM23 and respiratory complexes seem to form dynamic, non-permanent supercomplexes150,151.

A link between the mitochondrial respiratory chain and the machinery for protein synthesis was found by analysing the effects of mitochondrially generated ROS. The respiratory chain is a major source of ROS, and stress conditions and dysfunction of the respiratory chain can lead to increased ROS production that causes oxidative damage in proteins, DNA and membranes. As ROS have signalling functions, mitochondrially produced ROS can signal the functional state of mitochondria19,143,157. ROS were found to target redox-sensitive cysteine residues (redox switches) of the cytosolic translation apparatus, including the ribosome and translation factors157. When ROS production is increased, translation efficiency is decreased in a reversible manner (Fig. 3). The respiratory chain thus participates in controlling cytosolic protein synthesis to decrease the protein load under mitochondrial stress conditions157.

Coenzyme Q (CoQ), also termed ubiquinone, is a central molecule for the function of the respiratory chain158,159. CoQ is a redox-active lipid that mediates electron transfer from respiratory complexes I and II to complex III (Fig. 3) and functions as a cofactor of many enzymes. Several components of CoQ biosynthesis were recently identified using systematic mass spectrometry profiling at the proteomic, lipidomic and metabolomic levels26,71,158,159. The CoQ biosynthetic complex, which is located at the matrix side of the mitochondrial inner membrane, contains numerous enzymes involved in CoQ biosynthesis. All protein subunits of this dynamic CoQ biosynthetic complex are encoded by nuclear genes and are imported by the TOM and TIM machineries158,159. For example, the precursor of the methyltransferase Coq5 is processed twice. A first cleavage by MPP generates an unstable intermediate and a second cleavage by Oct1 generates the stable mature Coq5 enzyme (Fig. 3). Disturbance of processing by Oct1 leads to CoQ deficiency and respiratory defects71. Thus, import and specific processing of Coq precursors by the mitochondrial protein import machinery are functionally linked to the CoQ biosynthetic complex. A recent study showed that the machinery for assembly of Fe–S clusters is associated with respiratory chain supercomplexes160, underscoring a close connection between respiratory functions and cofactor biogenesis.

The mitochondrial inner membrane carries two large ATP-dependent protease complexes, the iAAA protease and the matrix-exposed mAAA protease18,161,162,163. These proteases are main elements of a quality control system for protein processing and turnover in mitochondria. AAA proteases cleave or degrade different inner membrane proteins such as some subunits of respiratory complexes and preprotein translocases and are also involved in the quality control and turnover of selected matrix, intermembrane space and outer membrane proteins (Fig. 3). In fungi, the adaptor proteins Mgr1 and Mgr3 associate with the iAAA protease and promote substrate recognition by the protease161,163. A recent proteomic study of yeast mitochondria identified the respiratory chain-interacting proteins Rci37 and Rci50 and demonstrated that they also interacted with the mAAA protease and the iAAA protease26, respectively, revealing specific connections between respiratory complexes III and IV and the inner membrane quality control system.

The functional network of the mitochondrial respiratory chain thus includes respiratory supercomplexes and machineries for protein biogenesis, cofactor biosynthesis and mitochondrial quality control.

Box 2 Mitochondrial presequence translocase and respiratory chain assembly

Interaction network of the presequence translocase

The translocase of the inner membrane (TIM23) complex is a central junction in the presequence import pathway and interacts with several partner complexes in a dynamic manner: with the translocase of the outer membrane (TOM) complex during preprotein transfer from the outer membrane (OM) to the inner membrane (IM), forming a TOM–TIM–preprotein supercomplex; with the ATP-driven presequence translocase-associated motor (PAM); with respiratory chain complexes III (CIII) and IV (CIV), which generate an electrochemical proton gradient (ΔμH+) driving preprotein insertion150,151; and with the ADP/ATP carrier, which also supports preprotein translocation152,234 (see the figure, part a; interactions depicted by double-headed arrows). The translocase subunit Tim21 functions as a dynamic coupling factor that interacts with TOM and the respiratory supercomplex CIII–CIV in an alternating manner. In fully active, respiring yeast mitochondria, the activity of the presequence translocase drives preprotein import efficiently. However, when the respiratory activity is decreased, coupling of the translocase to machineries involved in bioenergetics is beneficial to maintain the energy-dependent action of the translocase150,151,152. Preprotein translocases in the immediate vicinity of proton pumping respiratory complexes likely experience an increased proton motive force (localized proton gradients), and thus, under energy-limiting conditions, preprotein insertion into the IM is still possible150,151, ensuring that the biogenesis of respiratory complexes continues even in conditions of limited food supply.

Coupling of presequence translocase to respiratory chain assembly

The characterization of respiratory chain biogenesis in human mitochondria revealed a further level of cooperation between the protein import machinery and the respiratory chain, which is mediated by the mitochondrial translation regulation assembly intermediate of cytochrome c oxidase (MITRAC)153,154,155. MITRAC comprises several assembly intermediate complexes of the respiratory chain and plays a dual role (see the figure, part b). It links the presequence translocase to respiratory chain assembly intermediates via the TIM21-mediated transfer of imported proteins from TIM23 to MITRAC154. Moreover, MITRAC cooperates with the machineries for mitochondrial protein synthesis and insertion (oxidase assembly (OXA) insertion machinery) by adapting the efficiency of mitochondrial translation to the import of nuclear-encoded partner proteins (translational plasticity). MITRAC assembly factors bind to partially synthesized membrane-inserted proteins, such as the highly hydrophobic COX1 protein, and delay their synthesis until the appropriate partner proteins have been imported to ensure a proper balance of nuclear and mitochondrially encoded subunits155. The functions of human TIM14 (also known as DNAJC19 or PAM18) and ROMO1 (also known as MGR2) in the TIM23–PAM machinery have not been defined so far (indicated by dashed borders).

41580_2018_92_Figb_HTML.png

 

Mitochondrial membrane architecture and membrane contact sites

Contact sites between the mitochondrial outer and inner membranes and between the outer membrane and the endoplasmic reticulum (ER) are crucial elements of a large network of membrane contact sites that functions in protein and lipid biogenesis, mitochondrial membrane architecture and dynamics, metabolite and ion transport and mitochondrial inheritance (Fig. 4). Preprotein translocases form central building blocks of this ER–mitochondria organizing network (ERMIONE)164. The TOM and SAM complexes of the outer membrane interact with the large MICOS complex of the inner membrane9,10,11,13,165,166,167,168. MICOS is enriched at crista junctions169,170, and its largest subunit, Mic60, plays an important role in the formation of outer–inner membrane contact sites. In addition, Mic60 transiently interacts with the receptor and oxidoreductase Mia40 of the intermembrane space assembly machinery9. MICOS thus helps to position the downstream machineries MIA and SAM close to the import channel TOM and promotes the efficient import of cysteine-rich precursors into the intermembrane space and of β-barrel precursors into the outer membrane9,165.

41580_2018_92_Fig4_HTML.png

The mitochondrial contact site and cristae organizing system (MICOS) of the inner membrane (IM) and the protein translocases translocase of the outer membrane (TOM) and sorting and assembly machinery (SAM) of the outer membrane (OM) form the core of a large endoplasmic reticulum (ER)–mitochondria organizing network (ERMIONE) that includes multiple dynamic interactions: with the ER–mitochondria encounter structure (ERMES); with further ER–mitochondria contact sites that involve the receptor Tom70 and inositol trisphosphate (InsP3) receptors or the lipid transfer protein Lam6, as well as with vacuole–mitochondria contact sites (including Tom40 and the bridging protein Vps39); with the kinase PTEN-induced putative kinase 1 (PINK1) and the metabolite channel porin; with the mitochondrial intermembrane space (IMS) protein import and assembly system (Mia40); with respiratory chain complexes, the F1F0-ATP synthase and the fusion protein optic atrophy 1 (OPA1) of the IM; and with mitochondrial DNA (mtDNA) nucleoids (with the mtDNA packaging factor, termed mitochondrial transcription factor A (TFAM))262 of the matrix. Most components shown have been functionally conserved from yeast to humans; proteins that have been characterized in fungi only are indicated by a dashed border. In sum, ERMIONE forms a membrane-spanning system for the coordination of protein and lipid biogenesis, energetics, inheritance and quality control of mitochondria. CI, complex I; CII, complex II; CIII, complex III; CIV, complex IV.

 

The MICOS–SAM–TOM core of ERMIONE interacts with other mitochondrial machineries, which results in a large and complex network of functional interactions (Fig. 4). Most ERMIONE-interacting partners have been identified biochemically (via direct binding) or genetically (identifying synthetic growth defects).

The mechanisms that regulate the interactions in this complex network and their functional importance are the subject of intensive research and are being gradually revealed. In the inner membrane, Mic10, a core component of MICOS171,172, and its partner protein Mic27 are in dynamic contact with the dimeric F1F0-ATP synthase that shapes cristae rims, leading to a crosstalk between the two major membrane-shaping machineries of the inner membrane, MICOS and F1F0-ATP synthase144,173,174. Assembly of the Mic10-containing subcomplex of MICOS is linked to respiratory complexes and the mitochondria-specific dimeric phospholipid cardiolipin175,176. In addition, MICOS is connected to the machineries for mitochondrial fusion, including the inner membrane fusion protein optic atrophy 1 (termed OPA1 in mammals and Mgm1 in yeast)144,177,178. Studies of mutants have uncovered a functional link between MICOS and nucleoid aggregation and inheritance of mitochondrial DNA (mtDNA)179,180, but the underlying molecular mechanisms require further analysis. At the outer membrane, the SAM complex not only directly interacts with a fraction of TOM complexes in TOM–SAM supercomplexes125,127 but also exchanges Mdm10 with the ER–mitochondria encounter structure(ERMES) that links the mitochondrial outer membrane to the ER (Fig. 4). The outer membrane β-barrel protein Mdm10 is a subunit of both SAM, where it functions in TOM biogenesis, and ERMES, where it contributes to lipid transfer and maintenance of mitochondrial morphology181,182,183,184. The shuttling of Mdm10 between SAM and ERMES is regulated by the small protein Tom7. Tom7 has a dual localization: it is mainly located in the TOM complex but also functions outside the TOM complex as a regulatory factor that promotes Mdm10 transfer to ERMES185,186,187,188. Non-assembled Tom7 retards TOM assembly by shifting Mdm10 from the SAM form to the ERMES form, which constitutes a regulatory mechanism that is active when an excess of non-assembled TOM subunits accumulate in mitochondria. The major outer membrane metabolite channel porin, also termed voltage-dependent anion channel (VDAC), interacts with MICOS as well as the TOM complex10,189, linking metabolite transport to ERMIONE. Tom70, one of the receptors of the TOM complex, is also found outside the TOM complex and has crucial roles in forming ER–mitochondria contact sites (Fig. 4). Tom70 and its isoform Tom71 interact with Lam6 (also known as Ltc1), a lipid transfer protein that is anchored to membranes through lipid-binding sites and regulates contact sites between mitochondria, ER and other organelles190,191. Mammalian TOM70 also interacts with inositol trisphosphate (inositol-1,4,5-trisphosphate) receptors of the ER to promote Ca2+ transfer from the ER to mitochondria192. Moreover, Tom40 participates in the formation of vacuole–mitochondria contact sites, known as vacuole and mitochondria patch (vCLAMP), involving the vacuolar GTPase Ypt7 and the bridging protein Vps39 (also known as Vam6)190,193. Furthermore, vesicles can be released from the mitochondrial outer membrane to direct selected cargo to lysosomal degradation194,195,196 and Tom70 and Tom71 were found to be required for the formation of at least some mitochondria-derived vesicles197, linking the transport machinery to quality control and degradation systems. The AAA-type ATPase Msp1 promotes the extraction of mistargeted proteins from the outer membrane198,199. Upon accumulation of non-imported precursor proteins, Msp1 is recruited to Tom70 via the peripheral membrane protein Cis1 (also known as Atg31), leading to removal of non-imported proteins and their degradation by the proteasome200. Lastly, as discussed below, TOM and MICOS are involved in the accumulation of PTEN-induced putative kinase 1 (PINK1) at the outer membrane of dysfunctional or damaged mitochondria, which promotes their removal by mitophagy15,16,201,202.

Thus, the mitochondrial membranes contain at least two large protein networks, both containing TOM complexes: the TOM–TIM23–respiratory chain–AAA network, which couples protein import to bioenergetics and quality control mechanisms, and the MICOS–SAM–TOM–ER network (ERMIONE), which links protein biogenesis to membrane contact sites and membrane morphology. Whereas MICOS is enriched at crista junctions, TOM–TIM23–preprotein supercomplexes are preferentially found ~30–60 nm away from crista junctions124. It remains to be determined whether these two large networks function independently of each other or whether they exchange components as a means of coordinating protein biogenesis, energetics, membrane morphology and quality control. Although substantial future work will be required to fully understand how ERMIONE functions at the molecular level, the identification of these networks clearly demonstrates that mitochondrial machineries do not function as stand-alone units but are intimately linked to each other.

Protein import and pathophysiology

The efficiency of protein import into mitochondria is a sensitive indicator of the energetic state and the fitness of mitochondria. Various disorders of mitochondrial respiration and metabolism lead to reduction in the inner membrane potential203,204. As the membrane potential is crucial for protein translocation into and across the inner membrane, the import of preproteins is diminished1,2. Defects of protein homeostasis in the mitochondrial matrix by the accumulation of misfolded proteins also lead to a reduced protein import, likely by disturbing the mtHsp70 import motor205. The impaired activity of the mitochondrial protein import machinery under stress conditions or in mitochondrial diseases is a direct indicator of impaired mitochondrial functions and can induce stress responses or lead to the removal of damaged mitochondria by mitophagy (Box 3).

A mild disturbance of mitochondrial protein import can trigger the activation of the UPRmt as a result of the failure to import the transcription factor ATFS-1 (also known as ATF5) into the mitochondria, leading to its transport into the nucleus and a transcriptional stress response206,207 to rescue partially damaged mitochondria (Box 3). The stress-induced decrease in TIM17A levels also leads to decreased mitochondrial protein import and promotes the induction of a UPRmt133. In addition, accumulation of mitochondrial precursor proteins in the cytosol leads to an attenuation of cytosolic protein synthesis and activation of the proteasome to clear the mistargeted proteins from the cytosol208,209,210,211,212. This process is known as unfolded protein response activated by mistargeted mitochondrial proteins (UPRam) or mitochondrial precursor over-accumulation stress (mPOS).

Upon severe damage of mitochondrial protein import, the kinase PINK1 is not imported, processed and degraded but associates with the TOM complex as full-length protein, initiating a cascade that leads to removal of damaged mitochondria by mitophagy15,16,201,202 (Box 3). As mutations in PINK1 are linked to Parkinson disease, it has been proposed that insufficient mitophagy may be one of the causes underlying the development of the disease201. Recently, PINK1 and MIC60 of the MICOS complex were found to interact transiently, suggesting a crosstalk between PINK1 accumulation and inner membrane cristae remodelling213,214 (Fig. 4).

A recent study in yeast has suggested that the mitochondrial protein import machinery removes misfolded proteins from the cytosol and transports them to degradation inside mitochondria215. This process, which was named mitochondria as guardian in cytosol (MAGIC), requires further investigation to define its relevance for cellular protein homeostasis (proteostasis) and its relation to stress responses that are initiated by a decreased mitochondrial protein import efficiency such as UPRmt, UPRam and the PINK1–parkin pathway.

Proteolytic processing of precursor proteins also plays a role in mitochondrial quality control. In one mechanism, the removal of destabilizing amino-terminal amino acid residues by the processing enzyme Icp55 or Oct1 stabilizes imported proteins against proteolytic degradation49,71,72 (Fig. 2). In another mechanism, imported proteins are differentially processed, yielding two or more isoforms with distinct amino termini. For example, the inner membrane fusion protein OPA1 is first processed by MPP, generating a long isoform, and further processing by inner membrane proteases such as AAA proteases and OMA1 in mammals generates short isoforms216,217,218. The balance between long and short isoforms, which is important for membrane fusion and fission, is modulated by stress and the energetic state of the inner membrane (that is, mitochondrial activity)216,217,218.

The impaired processing of preproteins has been linked to mitochondrial dysfunctions in Alzheimer disease219. The matrix peptidasome degrades presequences and other peptides such as Alzheimer-linked amyloid-β peptides. Upon accumulation of amyloid-β peptides in mitochondria, the degradation of presequences is slowed down competitively, leading to an inhibition of processing peptidases. As a consequence, proteins imported into mitochondria are retained in precursor or intermediate forms that cannot fold properly and are prone to rapid degradation. The accumulation of amyloid-β peptides thus causes numerous changes in mitochondrial protein composition, providing possible explanations for a wide variety of mitochondrial alterations observed in Alzheimer disease.

Studies in recent years have provided increasing evidence for the involvement of mitochondrial protein import and processing in the pathogenesis of human diseases. At present, however, there are different views on whether mitochondrial dysfunctions are directly or indirectly involved in the development of major neurodegenerative diseases such as Parkinson disease and Alzheimer disease220. In Box 4, we provide an overview of more rare diseases and disorders that have been linked to specific components of the mitochondrial machineries for protein import and maturation, suggesting an involvement in disease pathogenesis. The diseases mostly affect the nervous system and other tissues with a high energy demand such as heart, muscles and kidney. On a mechanistic level, defects in preprotein targeting, the presequence pathway, processing and folding, the MIA pathway and the carrier pathway have been observed (Box 4).

The elaborate networks between preprotein translocases and other mitochondrial machineries have mostly been studied under physiological conditions. We expect that these networks will play an important role in understanding the mechanistic basis of mitochondrial stress responses and pathogenesis of diseases, exemplified by the role of TOM and MICOS in the accumulation of PINK1 at the outer membrane and the subsequent removal of damaged mitochondria by mitophagy213,221,222,223.

Box 3 Quality control pathways associated with the protein import machinery

Mitochondrial unfolded protein response (UPRmt)

The stress-activated transcription factor ATFS-1 contains mitochondrial and nuclear localization signals. The factor is imported into healthy mitochondria and degraded. When mitochondrial import is impaired, the transcription factor accumulates in the cytosol, is translocated into the nucleus and induces expression of chaperones, proteases and further factors to promote recovery of impaired mitochondria206,207.

Unfolded protein response activated by mistargeted mitochondrial proteins (UPRam)

Upon disturbance of mitochondrial protein import, precursor proteins accumulating in the cytosol trigger a stress response, the UPRam (also known as mitochondrial precursor over-accumulation stress (mPOS)), that reduces the efficiency of cytosolic protein synthesis and increases the activity of the proteasome, thus reducing the accumulation of mistargeted proteins in the cytosol208,209.

The PINK1–parkin pathway

The mitochondrial kinase PTEN-induced putative kinase 1 (PINK1) has been identified in familial cases of Parkinson disease. In healthy mitochondria, PINK1 is imported by the presequence pathway and processed by mitochondrial processing peptidase (MPP) and the presenilin-associated rhomboid-like protease PARL, followed by release into the cytosol and degradation by the proteasome. When protein import or processing by the presequence pathway is disturbed, unprocessed PINK1 accumulates at the translocase of the outer membrane (TOM) complex221,222,223 (Fig. 4), where it phosphorylates ubiquitin and the E3 ubiquitin ligase parkin, triggering the removal of damaged mitochondria by mitophagy.

Mitochondria as guardian in cytosol (MAGIC)

Some aggregation-prone or misfolded cytosolic proteins may be imported into mitochondria and degraded215, suggesting a role of mitochondria in cytosolic proteostasis.

 

Box 4 Disorders and diseases associated with distinct steps of human mitochondrial protein import and maturation

Mitochondrial targeting

Mutations in mitochondrial targeting signals can impair import of individual proteins, causing pyruvate dehydrogenase E1α deficiency235 or mitochondrial aspartyl-tRNA synthetase import defect linked to leukoencephalopathy with brainstem and spinal cord involvement and lactate elevation (LBSL)236L-Alanine:glyoxylate aminotransferase resides in peroxisomes in humans; however, mutations can generate a mitochondrial targeting signal, leading to mistargeting to mitochondria and primary hyperoxaluria type 1 (refs237,238). Similarly, in a form of renal Fanconi syndrome, a mutation generates a mitochondrial targeting signal in a peroxisomal protein involved in fatty acid oxidation, causing its mistargeting to mitochondria and disturbance of mitochondrial energy production in the proximal tubule239.

Mitochondrial intermembrane space import

A mutation in the ERV1 gene, which encodes the disulfide relay component GFER (also known as ALR), causes impairment of FAD cofactor binding, resulting in myopathy with cataract and combined respiratory chain deficiency240,241.

Mitochondrial presequence import pathway

Mutations in TIMM50, the gene encoding the presequence translocase receptor translocase of the inner membrane 50 (TIM50), lead to an impaired import by the presequence pathway, reduced levels of oxidative phosphorylation components, increased reactive oxygen species (ROS) production, severe epileptic encephalopathy, 3-methylglutaconic aciduria and lactic acidosis242,243. Mutations in the DNAJC19 gene, encoding the human mitochondrial import inner membrane (IM) translocase subunit TIM14 (also known as DNAJC19 or PAM18), cause dilated cardiomyopathy with ataxia, anaemia and testicular dysgenesis244,245. As TIM14 is mainly associated with prohibitin complexes affecting cardiolipin metabolism, cardiolipin alteration is likely involved in disease pathogenesis246. DNAJC15 (also known as methylation-controlled J protein (MCJ)), another human TIM14 homologue247, has been linked to tumorigenesis. TIM14 and DNAJC15 have been connected to distinct presequence translocase forms132, yet their exact relevance for protein import needs further analysis (indicated by a dashed border). A mutation in the MAGMAS gene, encoding the human TIM16 J-like co-chaperone (also known as PAM16) of the presequence translocase-associated import motor is linked to a severe spondylodysplastic dysplasia248.

Precursor protein processing

Mutations in the genes encoding the mitochondrial processing peptidase (MPP) subunits α (PMPCA) and β (PMPCB) cause defects in preprotein processing, which are linked to cerebellar ataxia249or early childhood neurodegeneration with cerebellar atrophy250. Mutations in the MIPEP gene, encoding the human octapeptidyl peptidase, cause left ventricular non-compaction cardiomyopathy with hypotonia and developmental delay251. Mutations in XPNPEP3, the gene encoding the human intermediate cleaving peptidase, are linked to nephronophthisis-like cystic kidney disease252,253.

Chaperones for precursor protein folding

Mutations in HSPD1, encoding the chaperonin HSP60, lead to neurodegenerative disorders, spastic paraplegia and mitochondrial chaperonin-60 disease254,255. A mutation in the HSPE1 gene, encoding the co-chaperonin HSP10, is associated with infantile spasms and developmental delay256.

Membrane protein transfer chaperone

Mutations in the DDP1 gene (also known as TIMM8A), encoding a small TIM chaperone (subunit TIM8A), cause deafness dystonia syndrome, which is also termed Mohr–Tranebjaerg syndrome257,258.

Carrier translocase

Mutations in the acyl glycerol kinase AGK gene lead to cataracts, cardiomyopathy and skeletal myopathy (Sengers syndrome)259. AGK plays a dual role as a lipid kinase and as subunit of the human TIM22 carrier translocase, linking lipid metabolism and protein import to Sengers syndrome260,261.

ΔΨ, membrane potential; IMS, inner membrane space; MIA, mitochondrial intermembrane space import and assembly; mtHSP70, mitochondrial heat shock protein 70 kDa; OM, outer membrane; TOM, translocase of the outer membrane.

41580_2018_92_Figc_HTML.png

 

Conclusions and perspectives

We have discussed that mitochondrial preprotein translocases, respiratory complexes, metabolite transporters, proteases, morphology complexes and membrane contact sites do not function as independent machineries but are physically and functionally connected in large dynamic networks. The protein translocases represent an essential housekeeping system of mitochondria. Not only are the translocases responsible for importing ~1,000–1,500 different proteins; they also form stable building blocks of the mitochondrial protein networks.

The rapid progress in identifying connections between machineries of different functions11,26,144,158,159,224,225 indicates that we have not reached saturation in the analysis of mitochondrial protein networks. In addition to the experimentally established connections described in this Review, interesting further network candidates include scaffold protein complexes that locally organize the protein–lipid composition of the inner membrane, such as the prohibitin ring complexes and stomatin-like protein 2, which associates with protease complexes and regulates the processing of PINK1 and OPA1 (refs11,226); lipid biosynthesis and remodelling enzymes; and cytosolic machineries that are involved in transferring preproteins, lipids or metabolites to mitochondria. Whereas several contact sites between mitochondria and other cell organelles have been identified recently, we have only a limited understanding of the interplay between cytosolic proteins or protein complexes and the mitochondrial outer membrane. This includes the potential involvement of specialized pools of cytosolic ribosomes in protein delivery to mitochondria227, the role of cytosolic chaperones, co-chaperones and potential targeting factors in cytosol–mitochondria crosstalk86,228,229,230, the rerouting of mitochondrial preproteins from the surface of the ER to mitochondria231 and the emerging evidence that numerous mitochondrial proteins possess a dual function and localization26,232.

Important questions concern the dynamics, regulation and turnover of the protein networks. It is likely that partner complexes in networks are turned over at different rates. Examples are the stress-regulated degradation of the TIM17A isoform of metazoan presequence translocases133 and the selective degradation of the outer membrane proteins Tom22 and porin-associated Om45 by the iAAA protease163(Fig. 3), whereas the other subunits of the complexes are turned over by different proteolytic machineries. The differential control of the networks by mitochondrial proteolytic systems, the cytosolic ubiquitin–proteasome system and mitophagy, as well as the role of lipids in establishing and maintaining the networks, will become central topics of research.

The large number of distinct functions observed in mitochondrial protein networks may give the initial impression that collaborations of protein machineries have developed in a random manner. The mechanistic studies performed so far, however, indicate that the interactions are highly specialized and specifically regulated, such as those between presequence translocase and respiratory supercomplexes and those between MICOS, TOM, SAM and the ER-mitochondria contact sites. To date, the studies have been mainly performed in yeast and partially in human mitochondria, which both belong to the same supergroup of eukaryotes, opisthokonts, including fungal and metazoan kingdoms. As the characterization of the mitochondrial protein import machinery in different supergroups yielded remarkable insight into core machineries and the high variability of transport complexes233, a systematic analysis of mitochondrial protein networks in the five eukaryotic supergroups will represent a rich source for defining core principles and variable parts of mitochondrial organization.


References

  1. 1.

    Neupert, W. A perspective on transport of proteins into mitochondria: a myriad of open questions. J. Mol. Biol. 427, 1135–1158 (2015).

  2. 2.

    Wiedemann, N. & Pfanner, N. Mitochondrial machineries for protein import and assembly. Annu. Rev. Biochem. 86, 685–714 (2017).

  3. 3.

    van der Bliek, A. M., Sedensky, M. M. & Morgan, P. G. Cell biology of the mitochondrion. Genetics 207, 843–871 (2017).

  4. 4.

    Lill, R. Function and biogenesis of iron–sulphur proteins. Nature460, 831–838 (2009).

  5. 5.

    Ott, M., Amunts, A. & Brown, A. Organization and regulation of mitochondrial protein synthesis. Annu. Rev. Biochem. 85, 77–101 (2016).

  6. 6.

    Hell, K., Neupert, W. & Stuart, R. A. Oxa1p acts as a general membrane insertion machinery for proteins encoded by mitochondrial DNA. EMBO J. 20, 1281–1288 (2001).

  7. 7.

    Labbé, K., Murley, A. & Nunnari, J. Determinants and functions of mitochondrial behavior. Annu. Rev. Cell Dev. Biol. 30, 357–391 (2014).

  8. 8.

    Westermann, B. Mitochondrial fusion and fission in cell life and death. Nat. Rev. Mol. Cell. Biol. 11, 872–884 (2010).

  9. 9.

    von der Malsburg, K. et al. Dual role of mitofilin in mitochondrial membrane organization and protein biogenesis. Dev. Cell 21, 694–707 (2011).

  10. 10.

    Harner, M. et al. The mitochondrial contact site complex, a determinant of mitochondrial architecture. EMBO J. 30, 4356–4370 (2011).

  11. 11.

    Hoppins, S. et al. A mitochondrial-focused genetic interaction map reveals a scaffold-like complex required for inner membrane organization in mitochondria. J. Cell Biol. 195, 323–340 (2011). References 9–11 report the identification of the MICOS, a multisubunit complex that links outer and inner membranes and is crucial for the maintenance of crista junctions.

  12. 12.

    Aaltonen, M. J. et al. MICOS and phospholipid transfer by Ups2–Mdm35 organize membrane lipid synthesis in mitochondria. J. Cell Biol. 213, 525–534 (2016).

  13. 13.

    Ott, C. et al. Sam50 functions in mitochondrial intermembrane space bridging and biogenesis of respiratory complexes. Mol. Cell. Biol. 32, 1173–1188 (2012).

  14. 14.

    Shpilka, T. & Haynes, C. M. The mitochondrial UPR: mechanisms, physiological functions and implications in ageing. Nat. Rev. Mol. Cell. Biol. 19, 109–120 (2018).

  15. 15.

    Pickles, S., Vigié, P. & Youle, R. J. Mitophagy and quality control mechanisms in mitochondrial maintenance. Curr. Biol. 28, R170–R185 (2018).

  16. 16.

    Harper, J. W., Ordureau, A. & Heo, J.-M. Building and decoding ubiquitin chains for mitophagy. Nat. Rev. Mol. Cell. Biol. 19, 93–108 (2018).

  17. 17.

    Cosentino, K. & Garcia-Saez, A. J. Bax and Bak pores: are we closing the circle? Trends Cell Biol. 27, 266–275 (2017).

  18. 18.

    Rugarli, E. I. & Langer, T. Mitochondrial quality control: a matter of life and death for neurons. EMBO J. 31, 1336–1349 (2012).

  19. 19.

    Sena, L. A. & Chandel, N. S. Physiological roles of mitochondrial reactive oxygen species. Mol. Cell 48, 158–167 (2012).

  20. 20.

    Calvo, S. E., Clauser, K. R. & Mootha, V. K. MitoCarta2.0: an updated inventory of mammalian mitochondrial proteins. Nucleic Acids Res. 44, D1251–D1257 (2016).

  21. 21.

    Forner, F., Foster, L. J., Campanaro, S., Valle, G. & Mann, M. Quantitative proteomic comparison of rat mitochondria from muscle, heart, and liver. Mol. Cell. Proteom. 5, 608–619 (2006).

  22. 22.

    Gaucher, S. P. et al. Expanded coverage of the human heart mitochondrial proteome using multidimensional liquid chromatography coupled with tandem mass spectrometry. J. Proteome Res. 3, 495–505 (2004).

  23. 23.

    Hung, V. et al. Proteomic mapping of the human mitochondrial intermembrane space in live cells via ratiometric APEX tagging. Mol. Cell 55, 332–341 (2014).

  24. 24.

    Lefort, N. et al. Proteome profile of functional mitochondria from human skeletal muscle using one-dimensional gel electrophoresis and HPLC-ESI-MS/MS. J. Proteom. 72, 1046–1060 (2009).

  25. 25.

    McDonald, T. et al. Expanding the subproteome of the inner mitochondria using protein separation technologies: one- and two-dimensional liquid chromatography and two-dimensional gel electrophoresis. Mol. Cell. Proteom. 5, 2392–2411 (2006).

  26. 26.

    Morgenstern, M. et al. Definition of a high-confidence mitochondrial proteome at quantitative scale. Cell Rep. 19, 2836–2852 (2017). This is a systematic quantitative analysis of the proteome of yeast mitochondria, revealing the absolute copy numbers of most mitochondrial protein machineries under fermentable and respiratory growth conditions.

  27. 27.

    Ohlmeier, S., Kastaniotis, A. J., Hiltunen, J. K. & Bergmann, U. The yeast mitochondrial proteome, a study of fermentative and respiratory growth. J. Biol. Chem. 279, 3956–3979 (2004).

  28. 28.

    Pagliarini, D. J. et al. A mitochondrial protein compendium elucidates complex I disease biology. Cell 134, 112–123 (2008).

  29. 29.

    Prokisch, H. et al. Integrative analysis of the mitochondrial proteome in yeast. PLOS Biol. 2, e160 (2004).

  30. 30.

    Reinders, J., Zahedi, R. P., Pfanner, N., Meisinger, C. & Sickmann, A. Toward the complete yeast mitochondrial proteome: multidimensional separation techniques for mitochondrial proteomics. J. Proteome Res. 5, 1543–1554 (2006).

  31. 31.

    Rhee, H.-W. et al. Proteomic mapping of mitochondria in living cells via spatially restricted enzymatic tagging. Science 339, 1328–1331 (2013).

  32. 32.

    Sickmann, A. et al. The proteome of Saccharomyces cerevisiaemitochondria. Proc. Natl Acad. Sci. USA 100, 13207–13212 (2003).

  33. 33.

    Smith, A. C. & Robinson, A. J. MitoMinerv3.1, an update on the mitochondrial proteomics database. Nucleic Acids Res. 44, D1258–D1261 (2016).

  34. 34.

    Taylor, S. W. et al. Characterization of the human heart mitochondrial proteome. Nat. Biotechnol. 21, 281–286 (2003).

  35. 35.

    Vögtle, F. N. et al. Landscape of submitochondrial protein distribution. Nat. Commun. 8, 290 (2017).

  36. 36.

    Vögtle, F. N. et al. Intermembrane space proteome of yeast mitochondria. Mol. Cell. Proteom. 11, 1840–1852 (2012).

  37. 37.

    Zahedi, R. P. et al. Proteomic analysis of the yeast mitochondrial outer membrane reveals accumulation of a subclass of preproteins. Mol. Biol. Cell 17, 1436–1450 (2006).

  38. 38.

    Zhang, J. et al. Systematic characterization of the murine mitochondrial proteome using functionally validated cardiac mitochondria. Proteomics 8, 1564–1575 (2008).

  39. 39.

    de Godoy, L. M. F. et al. Comprehensive mass-spectrometry-based proteome quantification of haploid versus diploid yeast. Nature 455, 1251–1254 (2008).

  40. 40.

    Schwanhäusser, B. et al. Global quantification of mammalian gene expression control. Nature 473, 337–342 (2011).

  41. 41.

    Beck, M. et al. The quantitative proteome of a human cell line. Mol. Systems Biol. 7, 549 (2011).

  42. 42.

    Kulak, N. A., Pichler, G., Paron, I., Nagaraj, N. & Mann, M. Minimal, encapsulated proteomic-sample processing applied to copy-number estimation in eukaryotic cells. Nat. Methods 11, 319–324 (2014).

  43. 43.

    Wiśniewski, J. R., Hein, M. Y., Cox, J. & Mann, M. A ‘proteomic ruler’ for protein copy number and concentration estimation without spike-in standards. Mol. Cell. Proteom. 13, 3497–3506 (2014).

  44. 44.

    Paulo, J. A. et al. Quantitative mass spectrometry-based multiplexing compares the abundance of 5000 S. cerevisiaeproteins across 10 carbon sources. J. Proteom. 148, 85–93 (2016).

  45. 45.

    Harbauer, A. B. et al. Mitochondria: cell cycle-dependent regulation of mitochondrial preprotein translocase. Science 346, 1109–1113 (2014).

  46. 46.

    Gerbeth, C. et al. Glucose-induced regulation of protein import receptor Tom22 by cytosolic and mitochondria-bound kinases. Cell Metab. 18, 578–587 (2013).

  47. 47.

    Rao, S. et al. Biogenesis of the preprotein translocase of the outer mitochondrial membrane: protein kinase A phosphorylates the precursor of Tom40 and impairs its import. Mol. Biol. Cell 23, 1618–1627 (2012).

  48. 48.

    Schmidt, O. et al. Regulation of mitochondrial protein import by cytosolic kinases. Cell 144, 227–239 (2011). This study reports that biogenesis and activity of the TOM complex are regulated by cytosolic kinases, revealing the main protein import site of mitochondria as a major target for cytosolic signalling pathways.

  49. 49.

    Vögtle, F. N. et al. Global analysis of the mitochondrial N-proteome identifies a processing peptidase critical for protein stability. Cell 139, 428–439 (2009).

  50. 50.

    Abe, Y. et al. Structural basis of presequence recognition by the mitochondrial protein import receptor Tom20. Cell 100, 551–560 (2000).

  51. 51.

    van Wilpe, S. et al. Tom22 is a multifunctional organizer of the mitochondrial preprotein translocase. Nature 401, 485–489 (1999).

  52. 52.

    Kuszak, A. J. et al. Evidence of distinct channel conformations and substrate binding affinities for the mitochondrial outer membrane protein translocase pore Tom40. J. Biol. Chem. 290, 26204–26217 (2015).

  53. 53.

    Melin, J. et al. Presequence recognition by the Tom40 channel contributes to precursor translocation into the mitochondrial matrix. Mol. Cell. Biol. 34, 3473–3485 (2014).

  54. 54.

    Lohret, T. A., Jensen, R. E. & Kinnally, K. W. Tim23, a protein import component of the mitochondrial inner membrane, is required for normal activity of the multiple conductance channel, MCC. J. Cell Biol. 137, 377–386 (1997).

  55. 55.

    Bauer, M. F., Sirrenberg, C., Neupert, W. & Brunner, M. Role of Tim23 as voltage sensor and presequence receptor in protein import into mitochondria. Cell 87, 33–41 (1996).

  56. 56.

    Dekker, P. J. et al. Identification of MIM23, a putative component of the protein import machinery of the mitochondrial inner membrane. FEBS Lett. 330, 66–70 (1993).

  57. 57.

    Demishtein-Zohary, K., Marom, M., Neupert, W., Mokranjac, D. & Azem, A. GxxxG motifs hold the TIM23 complex together. FEBS J.282, 2178–2186 (2015).

  58. 58.

    Truscott, K. N. et al. A presequence- and voltage-sensitive channel of the mitochondrial preprotein translocase formed by Tim23. Nat. Struct. Biol. 8, 1074–1082 (2001).

  59. 59.

    Malhotra, K. et al. Cardiolipin mediates membrane and channel interactions of the mitochondrial TIM23 protein import complex receptor Tim50. Sci. Adv. 3, e1700532 (2017).

  60. 60.

    Denkert, N. et al. Cation selectivity of the presequence translocase channel Tim23 is crucial for efficient protein import. eLife 6, e28324 (2017).

  61. 61.

    Ramesh, A. et al. A disulfide bond in the TIM23 complex is crucial for voltage gating and mitochondrial protein import. J. Cell Biol.214, 417–431 (2016).

  62. 62.

    Kang, P. J. et al. Requirement for hsp70 in the mitochondrial matrix for translocation and folding of precursor proteins. Nature348, 137–143 (1990).

  63. 63.

    Demishtein-Zohary, K. et al. Role of Tim17 in coupling the import motor to the translocation channel of the mitochondrial presequence translocase. eLife 6, e22696 (2017).

  64. 64.

    Ting, S.-Y., Yan, N. L., Schilke, B. A. & Craig, E. A. Dual interaction of scaffold protein Tim44 of mitochondrial import motor with channel-forming translocase subunit Tim23. eL ife 6, e23609 (2017).

  65. 65.

    Banerjee, R., Gladkova, C., Mapa, K., Witte, G. & Mokranjac, D. Protein translocation channel of mitochondrial inner membrane and matrix-exposed import motor communicate via two-domain coupling protein. eLife 4, e11897 (2015).

  66. 66.

    Sikor, M., Mapa, K., von Voithenberg, L. V., Mokranjac, D. & Lamb, D. C. Real-time observation of the conformational dynamics of mitochondrial Hsp70 by spFRET. EMBO J. 32, 1639–1649 (2013).

  67. 67.

    Schendzielorz, A. B. et al. Two distinct membrane potential-dependent steps drive mitochondrial matrix protein translocation. J. Cell Biol. 216, 83–92 (2017).

  68. 68.

    Schulz, C. & Rehling, P. Remodelling of the active presequence translocase drives motor-dependent mitochondrial protein translocation. Nat. Commun. 5, 4349 (2014).

  69. 69.

    Fukasawa, Y. et al. MitoFates: improved prediction of mitochondrial targeting sequences and their cleavage sites. Mol. Cell. Proteom. 14, 1113–1126 (2015).

  70. 70.

    Varshavsky, A. The N-end rule pathway and regulation by proteolysis. Protein Sci. 20, 1298–1345 (2011).

  71. 71.

    Veling, M. T. et al. Multi-omic mitoprotease profiling defines a role for Oct1p in coenzyme Q production. Mol. Cell 68, 970–977 (2017).

  72. 72.

    Vögtle, F. N. et al. Mitochondrial protein turnover: role of the precursor intermediate peptidase Oct1 in protein stabilization. Mol. Biol. Cell 22, 2135–2143 (2011).

  73. 73.

    Cheng, M. Y. et al. Mitochondrial heat-shock protein hsp60 is essential for assembly of proteins imported into yeast mitochondria. Nature 337, 620–625 (1989).

  74. 74.

    Ostermann, J., Horwich, A. L., Neupert, W. & Hartl, F. U. Protein folding in mitochondria requires complex formation with hsp60 and ATP hydrolysis. Nature 341, 125–130 (1989).

  75. 75.

    Ieva, R. et al. Mgr2 functions as lateral gatekeeper for preprotein sorting in the mitochondrial inner membrane. Mol. Cell 56, 641–652 (2014). This paper identifies a lateral gatekeeper protein at the TIM23 complex that controls the proper release of preproteins with hydrophobic sorting signals into the inner membrane.

  76. 76.

    Schendzielorz, A. B. et al. Motor recruitment to the TIM23 channel’s lateral gate restricts polypeptide release into the inner membrane. Nat. Commun. 9, 4028 (2018).

  77. 77.

    Stiller, S. B. et al. Mitochondrial OXA translocase plays a major role in biogenesis of inner-membrane proteins. Cell Metab. 23, 901–908 (2016).

  78. 78.

    Park, K., Botelho, S. C., Hong, J., Osterberg, M. & Kim, H. Dissecting stop transfer versus conservative sorting pathways for mitochondrial inner membrane proteins in vivo. J. Biol. Chem.288, 1521–1532 (2013).

  79. 79.

    Bohnert, M. et al. Cooperation of stop-transfer and conservative sorting mechanisms in mitochondrial protein transport. Curr. Biol. 20, 1227–1232 (2010).

  80. 80.

    Herrmann, J. M., Neupert, W. & Stuart, R. A. Insertion into the mitochondrial inner membrane of a polytopic protein, the nuclear-encoded Oxa1p. EMBO J. 16, 2217–2226 (1997).

  81. 81.

    Bausewein, T. et al. Cryo-EM structure of the TOM core complex from Neurospora crassaCell 170, 693–700 (2017). This paper provides a cryoelectron microscopy structure of the TOM core complex of the mitochondrial outer membrane. Two translocation pores formed by Tom40 β-barrels are connected by two copies of the central receptor Tom22.

  82. 82.

    Shiota, T. et al. Molecular architecture of the active mitochondrial protein gate. Science 349, 1544–1548 (2015). This study maps the TOM complex architecture by site-specific crosslinking in the native membrane, revealing different translocation pathways for hydrophilic and hydrophobic precursor proteins through the Tom40 channel.

  83. 83.

    Backes, S. et al. Tom70 enhances mitochondrial preprotein import efficiency by binding to internal targeting sequences. J. Cell Biol.217, 1369–1382 (2018). This study reports a role of Tom70 in the import of presequence-containing preproteins that contain additional internal targeting signals. While Tom20 and Tom22 bind the amino-terminal presequences, Tom70 binds the internal signals to prevent aggregation of the preproteins.

  84. 84.

    Melin, J. et al. A presequence-binding groove in Tom70 supports import of Mdl1 into mitochondria. Biochim. Biophys. Acta 1853, 1850–1859 (2015).

  85. 85.

    Yamamoto, H. et al. Roles of Tom70 in Import of presequence-containing mitochondrial proteins. J. Biol. Chem. 284, 31635–31646 (2009).

  86. 86.

    Young, J. C., Hoogenraad, N. J. & Hartl, F. U. Molecular chaperones Hsp90 and Hsp70 deliver preproteins to the mitochondrial import receptor Tom70. Cell 112, 41–50 (2003).

  87. 87.

    Wiedemann, N., Pfanner, N. & Ryan, M. T. The three modules of ADP/ATP carrier cooperate in receptor recruitment and translocation into mitochondria. EMBO J. 20, 951–960 (2001).

  88. 88.

    Curran, S. P., Leuenberger, D., Schmidt, E. & Koehler, C. M. The role of the Tim8p-Tim13p complex in a conserved import pathway for mitochondrial polytopic inner membrane proteins. J. Cell Biol. 158, 1017–1027 (2002).

  89. 89.

    Vial, S. et al. Assembly of Tim9 and Tim10 into a functional chaperone. J. Biol. Chem. 277, 36100–36108 (2002).

  90. 90.

    Koehler, C. M. et al. Tim9p, an essential partner subunit of Tim10p for the import of mitochondrial carrier proteins. EMBO J. 17, 6477–6486 (1998).

  91. 91.

    Sirrenberg, C. et al. Carrier protein import into mitochondria mediated by the intermembrane proteins Tim10/Mrs11 and Tim12/Mrs5. Nature 391, 912–915 (1998).

  92. 92.

    Weinhäupl, K. et al. Structural basis of membrane protein chaperoning through the mitochondrial intermembrane space. Cell 175, 1365–1379 (2018).

  93. 93.

    Okamoto, H., Miyagawa, A., Shiota, T., Tamura, Y. & Endo, T. Intramolecular disulfide bond of Tim22 protein maintains integrity of the TIM22 complex in the mitochondrial inner membrane. J. Biol. Chem. 289, 4827–4838 (2014).

  94. 94.

    Wrobel, L., Trojanowska, A., Sztolsztener, M. E. & Chacinska, A. Mitochondrial protein import: Mia40 facilitates Tim22 translocation into the inner membrane of mitochondria. Mol. Biol. Cell 24, 543–554 (2013).

  95. 95.

    Rehling, P. et al. Protein insertion into the mitochondrial inner membrane by a twin-pore translocase. Science 299, 1747–1751 (2003).

  96. 96.

    Kerscher, O., Holder, J., Srinivasan, M., Leung, R. S. & Jensen, R. E. The Tim54p-Tim22p complex mediates insertion of proteins into the mitochondrial inner membrane. J. Cell Biol. 139, 1663–1675 (1997).

  97. 97.

    Sirrenberg, C., Bauer, M. F., Guiard, B., Neupert, W. & Brunner, M. Import of carrier proteins into the mitochondrial inner membrane mediated by Tim22. Nature 384, 582–585 (1996).

  98. 98.

    Koch, J. R. & Schmid, F. X. Mia40 combines thiol oxidase and disulfide isomerase activity to efficiently catalyze oxidative folding in mitochondria. J. Mol. Biol. 426, 4087–4098 (2014).

  99. 99.

    Chacinska, A. et al. Essential role of Mia40 in import and assembly of mitochondrial intermembrane space proteins. EMBO J. 23, 3735–3746 (2004).

  100. 100.

    Mesecke, N. et al. A disulfide relay system in the intermembrane space of mitochondria that mediates protein import. Cell 121, 1059–1069 (2005).

  101. 101.

    Milenkovic, D. et al. Identification of the signal directing Tim9 and Tim10 into the intermembrane space of mitochondria. Mol. Biol. Cell 20, 2530–2539 (2009).

  102. 102.

    Sideris, D. P. et al. A novel intermembrane space-targeting signal docks cysteines onto Mia40 during mitochondrial oxidative folding. J. Cell Biol. 187, 1007–1022 (2009).

  103. 103.

    Peleh, V., Cordat, E. & Herrmann, J. M. Mia40 is a trans-site receptor that drives protein import into the mitochondrial intermembrane space by hydrophobic substrate binding. eLife 5, e16177 (2016).

  104. 104.

    Neal, S. E. et al. Mia40 protein serves as an electron sink in the Mia40-Erv1 import pathway. J. Biol. Chem. 290, 20804–20814 (2015).

  105. 105.

    Kojer, K., Peleh, V., Calabrese, G., Herrmann, J. M. & Riemer, J. Kinetic control by limiting glutaredoxin amounts enables thiol oxidation in the reducing mitochondrial intermembrane space. Mol. Biol. Cell 26, 195–204 (2015).

  106. 106.

    Jores, T. et al. Characterization of the targeting signal in mitochondrial β-barrel proteins. Nat. Commun. 7, 12036 (2016).

  107. 107.

    Klein, A. et al. Characterization of the insertase for β-barrel proteins of the outer mitochondrial membrane. J. Cell Biol. 199, 599–611 (2012).

  108. 108.

    Wiedemann, N. et al. Machinery for protein sorting and assembly in the mitochondrial outer membrane. Nature 424, 565–571 (2003).

  109. 109.

    Paschen, S. A. et al. Evolutionary conservation of biogenesis of beta-barrel membrane proteins. Nature 426, 862–866 (2003).

  110. 110.

    Höhr, A. I. C. et al. Membrane protein insertion through a mitochondrial β-barrel gate. Science 359, eaah6834 (2018). This study maps the membrane insertion pathway of β-barrel precursors through the SAM channel, signal-induced opening of the lateral gate of Sam50 and release of the folded β-barrel protein into the outer membrane.

  111. 111.

    Kutik, S. et al. Dissecting membrane insertion of mitochondrial β-barrel proteins. Cell 132, 1011–1024 (2008).

  112. 112.

    Krüger, V. et al. Identification of new channels by systematic analysis of the mitochondrial outer membrane. J. Cell Biol. 216, 3485–3495 (2017).

  113. 113.

    Dimmer, K. S. et al. A crucial role for Mim2 in the biogenesis of mitochondrial outer membrane proteins. J. Cell Sci. 125, 3464–3473 (2012).

  114. 114.

    Papic, D., Krumpe, K., Dukanovic, J., Dimmer, K. S. & Rapaport, D. Multispan mitochondrial outer membrane protein Ugo1 follows a unique Mim1-dependent import pathway. J. Cell Biol. 194, 397–405 (2011).

  115. 115.

    Hulett, J. M. et al. The transmembrane segment of Tom20 is recognized by Mim1 for docking to the mitochondrial TOM complex. J. Mol. Biol. 376, 694–704 (2008).

  116. 116.

    Popov-Čeleketić, J., Waizenegger, T. & Rapaport, D. Mim1 functions in an oligomeric form to facilitate the integration of Tom20 into the mitochondrial outer membrane. J. Mol. Biol. 376, 671–680 (2008).

  117. 117.

    Becker, T. et al. The mitochondrial import protein Mim1 promotes biogenesis of multispanning outer membrane proteins. J. Cell Biol.194, 387–395 (2011).

  118. 118.

    Dukanovic, J. & Rapaport, D. Multiple pathways in the integration of proteins into the mitochondrial outer membrane. Biochim. Biophys. Acta 1808, 971–980 (2011).

  119. 119.

    Keskin, A., Akdoğan, E. & Dunn, C. D. Evidence for amino acid snorkeling from a high-resolution, in vivo analysis of Fis1 tail-anchor insertion at the mitochondrial outer membrane. Genetics205, 691–705 (2017).

  120. 120.

    Vögtle, F. N. et al. The fusogenic lipid phosphatidic acid promotes the biogenesis of mitochondrial outer membrane protein Ugo1. J. Cell Biol. 210, 951–960 (2015).

  121. 121.

    Sauerwald, J. et al. Genome-wide screens in Saccharomyces cerevisiae highlight a role for cardiolipin in biogenesis of mitochondrial outer membrane multispan proteins. Mol. Cell. Biol.35, 3200–3211 (2015).

  122. 122.

    Albrecht, R. et al. The Tim21 binding domain connects the preprotein translocases of both mitochondrial membranes. EMBO Rep. 7, 1233–1238 (2006).

  123. 123.

    Chacinska, A. et al. Distinct forms of mitochondrial TOM-TIM supercomplexes define signal-dependent states of preprotein sorting. Mol. Cell. Biol. 30, 307–318 (2010).

  124. 124.

    Gold, V. A. M. et al. Visualizing active membrane protein complexes by electron cryotomography. Nat. Commun. 5, 4129 (2014).

  125. 125.

    Qiu, J. et al. Coupling of mitochondrial import and export translocases by receptor-mediated supercomplex formation. Cell154, 596–608 (2013). This study identifies a TOM–SAM supercomplex that facilitates the efficient transfer of β-barrel precursors from the TOM import channel to the SAM membrane insertion sites of the mitochondrial outer membrane.

  126. 126.

    Waegemann, K., Popov-Čeleketić, D., Neupert, W., Azem, A. & Mokranjac, D. Cooperation of TOM and TIM23 complexes during translocation of proteins into mitochondria. J. Mol. Biol. 427, 1075–1084 (2015).

  127. 127.

    Wenz, L.-S. et al. Sam37 is crucial for formation of the mitochondrial TOM-SAM supercomplex, thereby promoting β-barrel biogenesis. J. Cell Biol. 210, 1047–1054 (2015).

  128. 128.

    Callegari, S. et al. TIM29 is a subunit of the human carrier translocase required for protein transport. FEBS Lett. 590, 4147–4158 (2016).

  129. 129.

    Kang, Y. et al. Tim29 is a novel subunit of the human TIM22 translocase and is involved in complex assembly and stability. eLife 5, e17463 (2016).

  130. 130.

    Gornicka, A. et al. A discrete pathway for the transfer of intermembrane space proteins across the outer membrane of mitochondria. Mol. Biol. Cell 25, 3999–4009 (2014).

  131. 131.

    Wiedemann, N. et al. Biogenesis of yeast mitochondrial cytochrome c: a unique relationship to the TOM machinery. J. Mol. Biol. 327, 465–474 (2003).

  132. 132.

    Sinha, D., Srivastava, S., Krishna, L. & D’Silva, P. Unraveling the intricate organization of mammalian mitochondrial presequence translocases: existence of multiple translocases for maintenance of mitochondrial function. Mol. Cell. Biol. 34, 1757–1775 (2014).

  133. 133.

    Rainbolt, T. K., Atanassova, N., Genereux, J. C. & Wiseman, R. L. Stress-regulated translational attenuation adapts mitochondrial protein import through Tim17A degradation. Cell Metab. 18, 908–919 (2013).

  134. 134.

    Opalińska, M., Parys, K., Murcha, M. W. & Jańska, H. The plant i-AAA protease controls the turnover of an essential mitochondrial protein import component. J. Cell Sci. 131, jcs200733 (2018).

  135. 135.

    Wenz, L.-S. et al. The presequence pathway is involved in protein sorting to the mitochondrial outer membrane. EMBO Rep. 15, 678–685 (2014).

  136. 136.

    Song, J., Tamura, Y., Yoshihisa, T. & Endo, T. A novel import route for an N-anchor mitochondrial outer membrane protein aided by the TIM23 complex. EMBO Rep. 15, 670–677 (2014).

  137. 137.

    Lee, C. M., Sedman, J., Neupert, W. & Stuart, R. A. The DNA helicase, Hmi1p, is transported into mitochondria by a C-terminal cleavable targeting signal. J. Biol. Chem. 274, 20937–20942 (1999).

  138. 138.

    Ieva, R. et al. Mitochondrial inner membrane protease promotes assembly of presequence translocase by removing a carboxy-terminal targeting sequence. Nat. Commun. 4, 2853 (2013).

  139. 139.

    Sinzel, M. et al. Mcp3 is a novel mitochondrial outer membrane protein that follows a unique IMP-dependent biogenesis pathway. EMBO Rep. 17, 965–981 (2016).

  140. 140.

    Acin-Perez, R., Fernández-Silva, P., Peleato, M. L., Pérez-Martos, A. & Enríquez, J. A. Respiratory active mitochondrial supercomplexes. Mol. Cell 32, 529–539 (2008).

  141. 141.

    Melber, A. & Winge, D. R. Inner secrets of the respirasome. Cell167, 1450–1452 (2016).

  142. 142.

    Wu, M., Gu, J., Guo, R., Huang, Y. & Yang, M. Structure of mammalian respiratory supercomplex I1III2IV1Cell 167, 1598–1609 (2016). This study provides a cryoelectron microscopy structure of the 1.7 MDa respiratory supercomplex at near-atomic resolution, revealing the arrangement of complexes I, III and IV and the position of cofactors and phospholipids.

  143. 143.

    Milenkovic, D., Blaza, J. N., Larsson, N.-G. & Hirst, J. The enigma of the respiratory chain supercomplex. Cell Metab. 25, 765–776 (2017).

  144. 144.

    Schweppe, D. K. et al. Mitochondrial protein interactome elucidated by chemical cross-linking mass spectrometry. Proc. Natl Acad. Sci. USA 114, 1732–1737 (2017).

  145. 145.

    Chen, Y.-C. et al. Identification of a protein mediating respiratory supercomplex stability. Cell Metab. 15, 348–360 (2012).

  146. 146.

    Strogolova, V., Furness, A., Robb-McGrath, M., Garlich, J. & Stuart, R. A. Rcf1 and Rcf2, members of the hypoxia-induced gene 1 protein family, are critical components of the mitochondrial cytochrome bc 1-cytochrome c oxidase supercomplex. Mol. Cell. Biol. 32, 1363–1373 (2012).

  147. 147.

    Vukotic, M. et al. Rcf1 mediates cytochrome oxidase assembly and respirasome formation, revealing heterogeneity of the enzyme complex. Cell Metab. 15, 336–347 (2012).

  148. 148.

    Singhal, R. K. et al. Coi1 is a novel assembly factor of the yeast complex III-complex IV supercomplex. Mol. Biol. Cell 28, 2609–2622 (2017).

  149. 149.

    Dannenmaier, S. et al. Complete native stable isotope labeling by amino acids of Saccharomyces cerevisiae for global proteomic analysis. Anal. Chem. 90, 10501–10509 (2018).

  150. 150.

    van der Laan, M. et al. A role for Tim21 in membrane-potential-dependent preprotein sorting in mitochondria. Curr. Biol. 16, 2271–2276 (2006).

  151. 151.

    Wiedemann, N., van der Laan, M., Hutu, D. P., Rehling, P. & Pfanner, N. Sorting switch of mitochondrial presequence translocase involves coupling of motor module to respiratory chain. J. Cell Biol. 179, 1115–1122 (2007).

  152. 152.

    Mehnert, C. S. et al. The mitochondrial ADP/ATP carrier associates with the inner membrane presequence translocase in a stoichiometric manner. J. Biol. Chem. 289, 27352–27362 (2014).

  153. 153.

    Dennerlein, S. et al. MITRAC7 acts as a COX1-specific chaperone and reveals a checkpoint during cytochrome c oxidase assembly. Cell Rep. 12, 1644–1655 (2015).

  154. 154.

    Mick, D. U. et al. MITRAC links mitochondrial protein translocation to respiratory-chain assembly and translational regulation. Cell 151, 1528–1541 (2012).

  155. 155.

    Richter-Dennerlein, R. et al. Mitochondrial protein synthesis adapts to influx of nuclear-encoded protein. Cell 167, 471–483 (2016). This paper identifies a mechanism by which (MITRAC) assembly factors adjust the efficiency of mitochondrial synthesis of membrane-integrated respiratory chain subunits to the import of nuclear-encoded partner proteins, which is termed mitochondrial translational plasticity.

  156. 156.

    Stoldt, S. et al. Spatial orchestration of mitochondrial translation and OXPHOS complex assembly. Nat. Cell Biol. 20, 528–534 (2018).

  157. 157.

    Topf, U. et al. Quantitative proteomics identifies redox switches for global translation modulation by mitochondrially produced reactive oxygen species. Nat. Commun. 9, 324 (2018).

  158. 158.

    Floyd, B. J. et al. Mitochondrial protein interaction mapping identifies regulators of respiratory chain function. Mol. Cell 63, 621–632 (2016).

  159. 159.

    Stefely, J. A. et al. Mitochondrial protein functions elucidated by multi-omic mass spectrometry profiling. Nat. Biotechnol. 34, 1191–1197 (2016).

  160. 160.

    Böttinger, L. et al. Respiratory chain supercomplexes associate with the cysteine desulfurase complex of the iron-sulfur cluster assembly machinery. Mol. Biol. Cell 29, 776–785 (2018).

  161. 161.

    Gerdes, F., Tatsuta, T. & Langer, T. Mitochondrial AAA proteases—towards a molecular understanding of membrane-bound proteolytic machines. Biochim. Biophys. Acta 1823, 49–55 (2012).

  162. 162.

    Puchades, C. et al. Structure of the mitochondrial inner membrane AAA+protease YME1 gives insight into substrate processing. Science 358, eaao0464 (2017).

  163. 163.

    Wu, X., Li, L. & Jiang, H. Mitochondrial inner-membrane protease Yme1 degrades outer-membrane proteins Tom22 and Om45. J. Cell Biol. 217, 139–149 (2018).

  164. 164.

    van der Laan, M., Bohnert, M., Wiedemann, N. & Pfanner, N. Role of MINOS in mitochondrial membrane architecture and biogenesis. Trends Cell Biol. 22, 185–192 (2012).

  165. 165.

    Bohnert, M. et al. Role of mitochondrial inner membrane organizing system in protein biogenesis of the mitochondrial outer membrane. Mol. Biol. Cell 23, 3948–3956 (2012).

  166. 166.

    Zerbes, R. M. et al. Role of MINOS in mitochondrial membrane architecture: cristae morphology and outer membrane interactions differentially depend on mitofilin domains. J. Mol. Biol. 422, 183–191 (2012).

  167. 167.

    Körner, C. et al. The C-terminal domain of Fcj1 is required for formation of crista junctions and interacts with the TOB/SAM complex in mitochondria. Mol. Biol. Cell 23, 2143–2155 (2012).

  168. 168.

    Xie, J., Marusich, M. F., Souda, P., Whitelegge, J. & Capaldi, R. A. The mitochondrial inner membrane protein Mitofilin exists as a complex with SAM50, metaxins 1 and 2, coiled-coil-helix coiled-coil-helix domain-containing protein 3 and 6 and DnaJC11. FEBS Lett. 581, 3545–3549 (2007).

  169. 169.

    Rabl, R. et al. Formation of cristae and crista junctions in mitochondria depends on antagonism between Fcj1 and Su e/g. J. Cell Biol. 185, 1047–1063 (2009).

  170. 170.

    Jans, D. C. et al. STED super-resolution microscopy reveals an array of MINOS clusters along human mitochondria. Proc. Natl Acad. Sci. USA 110, 8936–8941 (2013).

  171. 171.

    Barbot, M. et al. Mic10 oligomerizes to bend mitochondrial inner membranes at cristae junctions. Cell Metab. 21, 756–763 (2015).

  172. 172.

    Bohnert, M. et al. Central role of Mic10 in the mitochondrial contact site and cristae organizing system. Cell Metab. 21, 747–755 (2015).

  173. 173.

    Rampelt, H. et al. Mic10, a core subunit of the mitochondrial contact site and cristae organizing system, interacts with the dimeric F1F0-ATP synthase. J. Mol. Biol. 429, 1162–1170 (2017).

  174. 174.

    Eydt, K., Davies, K. M., Behrendt, C., Wittig, I. & Reichert, A. S. Cristae architecture is determined by an interplay of the MICOS complex and the F1F0 ATP synthase via Mic27 and Mic10. Microb. Cell 4, 259–272 (2017).

  175. 175.

    Friedman, J. R., Mourier, A., Yamada, J., McCaffery, J. M. & Nunnari, J. MICOS coordinates with respiratory complexes and lipids to establish mitochondrial inner membrane architecture. Elife 4, e07739 (2015).

  176. 176.

    Rampelt, H. et al. Assembly of the mitochondrial cristae organizer Mic10 is regulated by Mic26-Mic27 antagonism and cardiolipin. J. Mol. Biol. 430, 1883–1890 (2018).

  177. 177.

    Barrera, M., Koob, S., Dikov, D., Vogel, F. & Reichert, A. S. OPA1 functionally interacts with MIC60 but is dispensable for crista junction formation. FEBS Lett. 590, 3309–3322 (2016).

  178. 178.

    Glytsou, C. et al. Optic atrophy 1 is epistatic to the core MICOS component MIC60 in mitochondrial cristae shape control. Cell Rep. 17, 3024–3034 (2016).

  179. 179.

    Itoh, K., Tamura, Y., Iijima, M. & Sesaki, H. Effects of Fcj1-Mos1 and mitochondrial division on aggregation of mitochondrial DNA nucleoids and organelle morphology. Mol. Biol. Cell 24, 1842–1851 (2013).

  180. 180.

    Li, H. et al. Mic60/Mitofilin determines MICOS assembly essential for mitochondrial dynamics and mtDNA nucleoid organization. Cell Death Differ. 23, 380–392 (2016).

  181. 181.

    Kornmann, B. et al. An ER-mitochondria tethering complex revealed by a synthetic biology screen. Science 325, 477–481 (2009). This paper identifies the ERMES. ERMES-mediated contact sites between ER and mitochondria are involved in lipid transfer and maintenance of mitochondrial morphology.

  182. 182.

    Meisinger, C. et al. The mitochondrial morphology protein Mdm10 functions in assembly of the preprotein translocase of the outer membrane. Dev. Cell 7, 61–71 (2004).

  183. 183.

    Yamano, K., Tanaka-Yamano, S. & Endo, T. Mdm10 as a dynamic constituent of the TOB/SAM complex directs coordinated assembly of Tom40. EMBO Rep. 11, 187–193 (2010).

  184. 184.

    Flinner, N. et al. Mdm10 is an ancient eukaryotic porin co-occurring with the ERMES complex. Biochim. Biophys. Acta 1833, 3314–3325 (2013).

  185. 185.

    Ellenrieder, L. et al. Separating mitochondrial protein assembly and endoplasmic reticulum tethering by selective coupling of Mdm10. Nat. Commun. 7, 13021 (2016).

  186. 186.

    Meisinger, C. et al. Mitochondrial protein sorting: differentiation of beta-barrel assembly by Tom7-mediated segregation of Mdm10. J. Biol. Chem. 281, 22819–22826 (2006).

  187. 187.

    Stroud, D. A. et al. Composition and topology of the endoplasmic reticulum-mitochondria encounter structure. J. Mol. Biol. 413, 743–750 (2011).

  188. 188.

    Yamano, K., Tanaka-Yamano, S. & Endo, T. Tom7 regulates Mdm10-mediated assembly of the mitochondrial import channel protein Tom40. J. Biol. Chem. 285, 41222–41231 (2010).

  189. 189.

    Müller, C. S. et al. Cryo-slicing blue native-mass spectrometry (csBN-MS), a novel technology for high resolution complexome profiling. Mol. Cell. Proteom. 15, 669–681 (2016).

  190. 190.

    Elbaz-Alon, Y. et al. Lam6 regulates the extent of contacts between organelles. Cell Rep. 12, 7–14 (2015).

  191. 191.

    Murley, A. et al. Ltc1 is an ER-localized sterol transporter and a component of ER-mitochondria and ER-vacuole contacts. J. Cell Biol. 209, 539–548 (2015). References 190 and 191 report that the lipid transfer protein Lam6 is located at and regulates contact sites between ER, mitochondria and further organelles. Lam6 interacts with the receptor Tom70 in ER–mitochondria contact sites.

  192. 192.

    Filadi, R. et al. TOM70 sustains cell bioenergetics by promoting IP3R3-mediated ER to mitochondria Ca2+ transfer. Curr. Biol. 28, 369–382 (2018). This paper shows that the receptor TOM70 of the mitochondrial outer membrane interacts with inositol trisphosphate receptors of the ER, supporting the formation of contact sites for Ca 2+ transfer to mitochondria.

  193. 193.

    González Montoro, A. et al. Vps39 interacts with Tom40 to establish one of two functionally distinct vacuole-mitochondria contact sites. Dev. Cell 45, 621–636 (2018).

  194. 194.

    McLelland, G.-L., Lee, S. A., McBride, H. M. & Fon, E. A. Syntaxin-17 delivers PINK1/parkin-dependent mitochondrial vesicles to the endolysosomal system. J. Cell Biol. 214, 275–291 (2016).

  195. 195.

    Soubannier, V., Rippstein, P., Kaufman, B. A., Shoubridge, E. A. & McBride, H. M. Reconstitution of mitochondria derived vesicle formation demonstrates selective enrichment of oxidized cargo. PLOS ONE 7, e52830 (2012).

  196. 196.

    Soubannier, V. et al. A vesicular transport pathway shuttles cargo from mitochondria to lysosomes. Curr. Biol. 22, 135–141 (2012). References 195 and 196 report that stress conditions can induce the formation of mitochondria-derived vesicles that transport selected cargo such as oxidized proteins to lysosomes as part of a mitochondrial quality control system.

  197. 197.

    Hughes, A. L., Hughes, C. E., Henderson, K. A., Yazvenko, N. & Gottschling, D. E. Selective sorting and destruction of mitochondrial membrane proteins in aged yeast. Elife 5, e13943 (2016).

  198. 198.

    Chen, Y.-C. et al. Msp1/ATAD1 maintains mitochondrial function by facilitating the degradation of mislocalized tail-anchored proteins. EMBO J. 33, 1548–1564 (2014).

  199. 199.

    Okreglak, V. & Walter, P. The conserved AAA-ATPase Msp1 confers organelle specificity to tail-anchored proteins. Proc. Natl Acad. Sci. USA 111, 8019–8024 (2014).

  200. 200.

    Weidberg, H. & Amon, A. MitoCPR — a surveillance pathway that protects mitochondria in response to protein import stress. Science 360, eaan4146 (2018). References 198, 199 and 200 report that the AAA-type ATPase Msp1 extracts mistargeted or non-imported proteins from the mitochondrial outer membrane for degradation by the proteasome in the cytosol.

  201. 201.

    Sekine, S. & Youle, R. J. PINK1 import regulation; a fine system to convey mitochondrial stress to the cytosol. BMC Biol. 16, 2 (2018).

  202. 202.

    Okamoto, K. Quality control: organellophagy: eliminating cellular building blocks via selective autophagy. J. Cell Biol. 205, 435–445 (2014).

  203. 203.

    Suomalainen, A. & Battersby, B. J. Mitochondrial diseases: the contribution of organelle stress responses to pathology. Nat. Rev. Mol. Cell. Biol. 19, 77–92 (2018).

  204. 204.

    Frazier, A. E., Thorburn, D. R. & Compton, A. G. Mitochondrial energy generation disorders: genes, mechanisms and clues to pathology. J. Biol. Chem. https://doi-org.autorpa.cmu.edu.tw/10.1074/jbc.R117.809194 (2017).

  205. 205.

    Melber, A. & Haynes, C. M. UPRmt regulation and output: a stress response mediated by mitochondrial-nuclear communication. Cell Res. 28, 281–295 (2018).

  206. 206.

    Nargund, A. M., Pellegrino, M. W., Fiorese, C. J., Baker, B. M. & Haynes, C. M. Mitochondrial import efficiency of ATFS-1 regulates mitochondrial UPR activation. Science 337, 587–590 (2012). Impaired mitochondrial protein import activates an unfolded protein response. The transcription factor ATFS-1 is normally imported into mitochondria and degraded. Disturbance of mitochondrial import leads to cytosolic accumulation of ATFS-1 and translocation into the nucleus, where it induces expression of chaperones and further rescue factors.

  207. 207.

    Fiorese, C. J. et al. The transcription factor ATF5 mediates a mammalian mitochondrial UPR. Curr. Biol. 26, 2037–2043 (2016).

  208. 208.

    Wrobel, L. et al. Mistargeted mitochondrial proteins activate a proteostatic response in the cytosol. Nature 524, 485–488 (2015).

  209. 209.

    Wang, X. & Chen, X. J. A cytosolic network suppressing mitochondria-mediated proteostatic stress and cell death. Nature524, 481–484 (2015). References 208 and 209 identify a stress response that is induced by accumulation of mitochondrial precursor proteins in the cytosol. The response leads to a decrease in cytosolic protein synthesis and increased proteasomal activity to reduce the accumulation of toxic mistargeted proteins.

  210. 210.

    Bragoszewski, P. et al. Retro-translocation of mitochondrial intermembrane space proteins. Proc. Natl Acad. Sci. USA 112, 7713–7718 (2015).

  211. 211.

    Bragoszewski, P., Gornicka, A., Sztolsztener, M. E. & Chacinska, A. The ubiquitin-proteasome system regulates mitochondrial intermembrane space proteins. Mol. Cell. Biol. 33, 2136–2148 (2013).

  212. 212.

    Kowalski, L. et al. Determinants of the cytosolic turnover of mitochondrial intermembrane space proteins. BMC Biol. 16, 66 (2018).

  213. 213.

    Akabane, S. et al. PKA regulates PINK1 stability and Parkin recruitment to damaged mitochondria through phosphorylation of MIC60. Mol. Cell 62, 371–384 (2016).

  214. 214.

    Tsai, P.-I. et al. PINK1 phosphorylates MIC60/mitofilin to control structural plasticity of mitochondrial crista junctions. Mol. Cell69, 744–756 (2018).

  215. 215.

    Ruan, L. et al. Cytosolic proteostasis through importing of misfolded proteins into mitochondria. Nature 543, 443–446 (2017).

  216. 216.

    Anand, R. et al. The i-AAA protease YME1L and OMA1 cleave OPA1 to balance mitochondrial fusion and fission. J. Cell Biol. 204, 919–929 (2014).

  217. 217.

    Ehses, S. et al. Regulation of OPA1 processing and mitochondrial fusion by m-AAA protease isoenzymes and OMA1. J. Cell Biol. 187, 1023–1036 (2009).

  218. 218.

    Song, Z., Chen, H., Fiket, M., Alexander, C. & Chan, D. C. OPA1 processing controls mitochondrial fusion and is regulated by mRNA splicing, membrane potential, and Yme1L. J. Cell Biol. 178, 749–755 (2007).

  219. 219.

    Mossmann, D. et al. Amyloid-β peptide induces mitochondrial dysfunction by inhibition of preprotein maturation. Cell Metab.20, 662–669 (2014).

  220. 220.

    Schapira, A. H. Mitochondrial diseases. Lancet 379, 1825–1834 (2012).

  221. 221.

    Okatsu, K., Kimura, M., Oka, T., Tanaka, K. & Matsuda, N. Unconventional PINK1 localization to the outer membrane of depolarized mitochondria drives Parkin recruitment. J. Cell Sci.128, 964–978 (2015).

  222. 222.

    Bertolin, G. et al. The TOMM machinery is a molecular switch in PINK1 and PARK2/PARKIN-dependent mitochondrial clearance. Autophagy 9, 1801–1817 (2013).

  223. 223.

    Lazarou, M., Jin, S. M., Kane, L. A. & Youle, R. J. Role of PINK1 binding to the TOM complex and alternate intracellular membranes in recruitment and activation of the E3 ligase Parkin. Dev. Cell 22, 320–333 (2012). This study shows that upon dissipation of the mitochondrial membrane potential, the kinase PINK1 accumulates at the mitochondrial outer membrane in a complex with the TOM.

  224. 224.

    Liu, F., Rijkers, D. T. S., Post, H. & Heck, A. J. R. Proteome-wide profiling of protein assemblies by cross-linking mass spectrometry. Nat. Methods 12, 1179–1184 (2015).

  225. 225.

    Huttlin, E. L. et al. Architecture of the human interactome defines protein communities and disease networks. Nature 545, 505–509 (2017).

  226. 226.

    Wai, T. et al. The membrane scaffold SLP2 anchors a proteolytic hub in mitochondria containing PARL and the i-AAA protease YME1L. EMBO Rep. 17, 1844–1856 (2016).

  227. 227.

    Segev, N. & Gerst, J. E. Specialized ribosomes and specific ribosomal protein paralogs control translation of mitochondrial proteins. J. Cell Biol. 217, 117–126 (2018).

  228. 228.

    Hoseini, H. et al. The cytosolic cochaperone Sti1 is relevant for mitochondrial biogenesis and morphology. FEBS J. 283, 3338–3352 (2016).

  229. 229.

    Jores, T. et al. Cytosolic Hsp70 and Hsp40 chaperones enable the biogenesis of mitochondrial β-barrel proteins. J. Cell Biol. 217, 3091–3108 (2018).

  230. 230.

    Opalinski, L. et al. Recruitment of cytosolic J-proteins by TOM receptors promotes mitochondrial protein biogenesis. Cell Rep.25, 2036–2043 (2018).

  231. 231.

    Hansen, K. G. et al. An ER surface retrieval pathway safeguards the import of mitochondrial membrane proteins in yeast. Science361, 1118–1122 (2018).

  232. 232.

    Ben-Menachem, R. & Pines, O. Detection of dual targeting and dual function of mitochondrial proteins in yeast. Methods Mol. Biol. 1567, 179–195 (2017).

  233. 233.

    Harsman, A. & Schneider, A. Mitochondrial protein import in trypanosomes: expect the unexpected. Traffic 18, 96–109 (2017).

  234. 234.

    Dienhart, M. K. & Stuart, R. A. The yeast Aac2 protein exists in physical association with the cytochrome bc1-COX supercomplex and the TIM23 machinery. Mol. Biol. Cell 19, 3934–3943 (2008).

  235. 235.

    Takakubo, F. et al. An amino acid substitution in the pyruvate dehydrogenase E1 alpha gene, affecting mitochondrial import of the precursor protein. Am. J. Hum. Genet. 57, 772–780 (1995).

  236. 236.

    Messmer, M. et al. A human pathology-related mutation prevents import of an aminoacyl-tRNA synthetase into mitochondria. Biochem. J. 433, 441–446 (2011).

  237. 237.

    Purdue, P. E., Allsop, J., Isaya, G., Rosenberg, L. E. & Danpure, C. J. Mistargeting of peroxisomal L-alanine:glyoxylate aminotransferase to mitochondria in primary hyperoxaluria patients depends upon activation of a cryptic mitochondrial targeting sequence by a point mutation. Proc. Natl Acad. Sci. USA88, 10900–10904 (1991).

  238. 238.

    Danpure, C. J., Cooper, P. J., Wise, P. J. & Jennings, P. R. An enzyme trafficking defect in two patients with primary hyperoxaluria type 1: peroxisomal alanine/glyoxylate aminotransferase rerouted to mitochondria. J. Cell Biol. 108, 1345–1352 (1989).

  239. 239.

    Klootwijk, E. D. et al. Mistargeting of peroxisomal EHHADH and inherited renal Fanconi’s syndrome. N. Engl. J. Med. 370, 129–138 (2014).

  240. 240.

    Di Fonzo, A. et al. The mitochondrial disulfide relay system protein GFER is mutated in autosomal-recessive myopathy with cataract and combined respiratory-chain deficiency. Am. J. Hum. Gen. 84, 594–604 (2009).

  241. 241.

    Ceh-Pavia, E., Ang, S. K., Spiller, M. P. & Lu, H. The disease-associated mutation of the mitochondrial thiol oxidase Erv1 impairs cofactor binding during its catalytic reaction. Biochem. J.464, 449–459 (2014).

  242. 242.

    Shahrour, M. A. et al. Mitochondrial epileptic encephalopathy, 3-methylglutaconic aciduria and variable complex V deficiency associated with TIMM50 mutations. Clin. Genet. 91, 690–696 (2017).

  243. 243.

    Reyes, A. et al. Mutations in TIMM50 compromise cell survival in OxPhos-dependent metabolic conditions. EMBO Mol. Med. 10, e8698 (2018).

  244. 244.

    Ojala, T. et al. New mutation of mitochondrial DNAJC19 causing dilated and noncompaction cardiomyopathy, anemia, ataxia, and male genital anomalies. Pediatr. Res. 72, 432–437 (2012).

  245. 245.

    Davey, K. M. et al. Mutation of DNAJC19, a human homologue of yeast inner mitochondrial membrane co-chaperones, causes DCMA syndrome, a novel autosomal recessive Barth syndrome-like condition. J. Med. Gen. 43, 385–393 (2006).

  246. 246.

    Richter-Dennerlein, R. et al. DNAJC19, a mitochondrial cochaperone associated with cardiomyopathy, forms a complex with prohibitins to regulate cardiolipin remodeling. Cell Metab.20, 158–171 (2014).

  247. 247.

    Schusdziarra, C., Blamowska, M., Azem, A. & Hell, K. Methylation-controlled J-protein MCJ acts in the import of proteins into human mitochondria. Hum. Mol. Genet. 22, 1348–1357 (2013).

  248. 248.

    Mehawej, C. et al. The impairment of MAGMAS function in human is responsible for a severe skeletal dysplasia. PLOS Genet. 10, e1004311 (2014).

  249. 249.

    Jobling, R. K. et al. PMPCA mutations cause abnormal mitochondrial protein processing in patients with non-progressive cerebellar ataxia. Brain 138, 1505–1517 (2015).

  250. 250.

    Vögtle, F. N. et al. Mutations in PMPCB encoding the catalytic subunit of the mitochondrial presequence protease cause neurodegeneration in early childhood. Am. J. Hum. Gen. 102, 557–573 (2018).

  251. 251.

    Eldomery, M. K. et al. MIPEP recessive variants cause a syndrome of left ventricular non-compaction, hypotonia, and infantile death. Genome Med. 8, 106 (2016).

  252. 252.

    Otto, E. A. et al. Mutation analysis of 18 nephronophthisis associated ciliopathy disease genes using a DNA pooling and next generation sequencing strategy. J. Med. Gen. 48, 105–116 (2011).

  253. 253.

    O’Toole, J. F. et al. Individuals with mutations in XPNPEP3, which encodes a mitochondrial protein, develop a nephronophthisis-like nephropathy. J. Clin. Invest. 120, 791–802 (2010).

  254. 254.

    Magen, D. et al. Mitochondrial Hsp60 chaperonopathy causes an autosomal-recessive neurodegenerative disorder linked to brain hypomyelination and leukodystrophy. Am. J. Hum. Gen. 83, 30–42 (2008).

  255. 255.

    Hansen, J. J. et al. Hereditary spastic paraplegia SPG13 is associated with a mutation in the gene encoding the mitochondrial chaperonin Hsp60. Am. J. Hum. Genet. 70, 1328–1332 (2002).

  256. 256.

    Bie, A. S. et al. Effects of a mutation in the HSPE1 gene encoding the mitochondrial co-chaperonin HSP10 and its potential association with a neurological and developmental disorder. Front. Mol. Biosci. 3, e874 (2016).

  257. 257.

    Koehler, C. M. et al. Human deafness dystonia syndrome is a mitochondrial disease. Proc. Natl Acad. Sci. USA 96, 2141–2146 (1999).

  258. 258.

    Roesch, K., Curran, S. P., Tranebjaerg, L. & Koehler, C. M. Human deafness dystonia syndrome is caused by a defect in assembly of the DDP1/TIMM8a-TIMM13 complex. Hum. Mol. Genet. 11, 477–486 (2002).

  259. 259.

    Mayr, J. A. et al. Lack of the mitochondrial protein acylglycerol kinase causes Sengers syndrome. Am. J. Hum. Gen. 90, 314–320 (2012).

  260. 260.

    Kang, Y. et al. Sengers syndrome-associated mitochondrial acylglycerol kinase is a subunit of the human TIM22 protein import complex. Mol. Cell 67, 457–470 (2017).

  261. 261.

    Vukotic, M. et al. Acylglycerol kinase mutated in sengers syndrome is a subunit of the TIM22 protein translocase in mitochondria. Mol. Cell 67, 471–483 (2017).

  262. 262.

    Kukat, C. et al. Cross-strand binding of TFAM to a single mtDNA molecule forms the mitochondrial nucleoid. Proc. Natl Acad. Sci. USA 112, 11288–11293 (2015).

arrow
arrow
    文章標籤
    Mitochondrion
    全站熱搜

    快樂小藥師 發表在 痞客邦 留言(0) 人氣()