The coordinated trafficking and tethering of membrane cargo within cells relies on the function of distinct cytoskeletal motors that are targeted to specific subcellular compartments through interactions with protein adaptors and phospholipids. The unique actin motor myosin VI functions at distinct steps during clathrin-mediated endocytosis and the early endocytic pathway – both of which are involved in cargo trafficking and sorting – through interactions with Dab2, GIPC, Tom1 and LMTK2. This multifunctional ability of myosin VI can be attributed to its cargo-binding tail region that contains two protein–protein interaction interfaces, a ubiquitin-binding motif and a phospholipid binding domain. In addition, myosin VI has been shown to be a regulator of the autophagy pathway, because of its ability to link the endocytic and autophagic pathways through interactions with the ESCRT-0 protein Tom1 and the autophagy adaptor proteins T6BP, NDP52 and optineurin. This function has been attributed to facilitating autophagosome maturation and subsequent fusion with the lysosome. Therefore, in this Commentary, we discuss the relationship between myosin VI and the different myosin VI adaptor proteins, particularly with regards to the spatial and temporal regulation that is required for the sorting of cargo at the early endosome, and their impact on autophagy.

The actin cytoskeleton in endocytosis and autophagy

In vertebrate cells the actin cytoskeleton provides the tensile strength to maintain cell morphology and the necessary mechanical forces and contractility to facilitate cell migration, adhesion, nutrient sensing, wound healing and cytokinesis. Actin filaments also form the tracks for myosin motor proteins to tether and transport various membrane compartments and cargoes during these distinct cellular processes. Actin filaments (F-actin) are polar structures with a fast polymerizing barbed end and a slow polymerizing pointed end, which can be assembled and disassembled in a dynamic equilibrium with actin monomers (G-actin) (Pollard et al., 2000).

Actin filaments and myosin motors are not only important for the steady-state distribution of intracellular organelles but are also intimately involved in shaping and deforming membranes during scission and fusion events, and crucial for the sorting of cargo on endosomes and the bidirectional delivery between membrane compartments of the endocytic pathway (Anitei and Hoflack, 2011; Mooren et al., 2012). Importantly, after vesicle internalisation, cargo sorting and tubular and vesicular carrier formation on endosomes is dependent on spatially restricted actin polymerisation. The force that is required to bud a vesicle is variable for different donor membranes. This might explain why, for example, in yeast, actin and myosin motor proteins are essential for clathrin-mediated endocytosis, but why in mammalian cells actin and myosin VI are only required for endocytosis from the apical but not from the basolateral domain (Buss et al., 2001; Gottlieb et al., 1993; Jackman et al., 1994). In addition, the cortical F-actin network underneath the plasma membrane may provide tracks for the short-range retrograde movement and the transfer of cargo vesicles from the cell periphery to microtubules for their long-distance delivery by microtubule-based motors to areas within the cell.

One specialised and regulated membrane trafficking pathway that requires the actin cytoskeleton, and the coordinated recruitment of membrane and cargo from various subcellular compartments – including fusion with vesicles derived from the endocytic pathway – is autophagy (Axe et al., 2008; Hailey et al., 2010; Ravikumar et al., 2010; Yamamoto et al., 2012; Young et al., 2006). Autophagy, characterised by the formation of a double-membrane autophagosome, is a lysosomal degradation pathway that is required for the clearance of damaged organelles, pathogens and large cytosolic protein complexes, in addition to being a homeostatic process during times of starvation (Yang and Klionsky, 2010). Actin filaments are essential at a very early point during starvation-induced autophagosome biogenesis, at which they are potentially involved in facilitating cargo recognition and aggresome formation; later they also assemble around mature autophagosomes, where they mediate fusion with endocytic and lysosomal compartments by tethering the required components and bringing them into close contact (Aguilera et al., 2012; Lee et al., 2010). Importantly, before an autophagosome is competent to fuse with the lysosome and for its contents to be degraded, it requires influx from the endocytic pathway. It receives input from early and late endosomes as well as from multivesicular bodies (MVB), which are believed to deliver the tethering and adaptor molecules that are necessary for final fusion with the lysosome (Simonsen and Tooze, 2009). One essential component is the ESCRT machinery, which is important for cargo sorting, MVB formation, cytokinesis and viral budding (McCullough et al., 2013). Loss of ESCRT function by disruption of either the ESCRT-I protein TSG101 or the ESCRT-III protein CHMP2B, leads to an accumulation of autophagosomes that is most likely to be linked to a defect in lysosomal fusion (Filimonenko et al., 2007; Lee et al., 2007; Rusten et al., 2007). However, it remains to be tested whether the ESCRT machinery is required at multiple stages during the autophagy pathway, including fusion with endocytic and lysosomal compartments, as well as during the final closure of the double-membrane autophagosome (Rusten and Stenmark, 2009).

Motor proteins facilitate the spatiotemporal distribution of membrane compartments

The spatial organisation of membrane compartments and the temporal regulation of membrane trafficking between intracellular organelles is a coordinated process involving dynamic cytoskeletal remodelling coupled to the action of specific motor protein-adaptor complexes that tether or drive transport along the actin and microtubule cytoskeletons. Dynein and Kinesin motors provide respective minus and plus end directed long distance movement of cargo along microtubules, while the myosin superfamily can anchor organelles to actin filaments and provide short-range movement along the actin cytoskeleton (Peckham, 2011). In humans, the myosin superfamily comprises 39 known myosin genes that are divided into 12 classes, and which are made up of both the conventional and unconventional myosins, whose functions are specified by their unique interactions with various cargo proteins and phospholipids. The conventional myosins, which comprise the non-muscle myosin II family and the skeletal, cardiac and smooth muscle myosins, assemble to form bipolar filaments that mediate sliding and crosslinking of actin filaments to generate contractility and tension. By contrast, unconventional myosins do not form filaments but, rather, function as either monomeric or dimeric cargo transporters, regulators of actin organisation, adjustable tethers for organelles, and/or as load-dependent tension sensors (Bloemink and Geeves, 2011; Hartman et al., 2011). The precise cellular roles of the different classes of unconventional myosins are mainly due to their divergent cargo-binding tail regions that mediate distinct interactions for targeting to various subcellular locations.

In this Commentary, we will focus on myosin of class VI, which has the unique ability, unlike the rest of the myosin family, to translocate towards the pointed or minus end of actin filaments. We will first highlight the structural properties of myosin VI, and the role and function of its distinct cargo adaptor proteins, which target this myosin to specific subcellular compartments to mediate its function (Fig. 1). Next, we will discuss the function of multiple adaptor molecules involved in the recruitment of myosin VI to the early endocytic pathway, where this motor could be required for sorting events directing cargo either to be recycled back to the plasma membrane or delivered to the lysosome for degradation. Finally, we will discuss the role of myosin VI during autophagy and its potential impact on the sorting and delivery of endosomal membranes required for autophagosome maturation, a process that is mediated by the interaction of myosin VI with the adaptor protein target of Myb1 (Tom1) (Tumbarello et al., 2012).

Fig. 1.

Cellular functions of myosin VI are directed by its distinct cargo adaptors. Myosin VI exerts its cellular functions through its cargo adaptor proteins in a number of specialised cell types: e.g. polarised epithelial cells containing microvilli, hair cells of the inner ear with stereocilia, dividing cells undergoing cytokinesis, migratory cells and neuronal cells. Myosin VI functions in various membrane trafficking pathways, such as clathrin-dependent endocytosis at the apical domain of polarised epithelia, through interaction with Dab2, sorting at the early endosome, through interaction with Tom1, GIPC and LMTK2 and, in autophagy, through interactions with T6BP, NDP52, optineurin and Tom1. In addition, myosin VI has multiple functions in the secretory pathway. It maintains Golgi morphology, facilitates vesicle fusion at the plasma membrane and regulates basolateral sorting through interactions with optineurin. Myosin VI also functions as an actin tether to maintain stereocilia and microvilli organisation. Finally, myosin VI has specialised functions in regulating cytokinesis during mitosis, facilitating polarised delivery of EGFR during directed cell migration, and regulating neuronal synaptic transmission and dendritic spine morphology. Broken arrows indicate potential crosstalk between pathways. N, nucleus.

Fig. 1.

Cellular functions of myosin VI are directed by its distinct cargo adaptors. Myosin VI exerts its cellular functions through its cargo adaptor proteins in a number of specialised cell types: e.g. polarised epithelial cells containing microvilli, hair cells of the inner ear with stereocilia, dividing cells undergoing cytokinesis, migratory cells and neuronal cells. Myosin VI functions in various membrane trafficking pathways, such as clathrin-dependent endocytosis at the apical domain of polarised epithelia, through interaction with Dab2, sorting at the early endosome, through interaction with Tom1, GIPC and LMTK2 and, in autophagy, through interactions with T6BP, NDP52, optineurin and Tom1. In addition, myosin VI has multiple functions in the secretory pathway. It maintains Golgi morphology, facilitates vesicle fusion at the plasma membrane and regulates basolateral sorting through interactions with optineurin. Myosin VI also functions as an actin tether to maintain stereocilia and microvilli organisation. Finally, myosin VI has specialised functions in regulating cytokinesis during mitosis, facilitating polarised delivery of EGFR during directed cell migration, and regulating neuronal synaptic transmission and dendritic spine morphology. Broken arrows indicate potential crosstalk between pathways. N, nucleus.

Structural and functional properties of myosin VI

Structurally, myosin VI, like all myosin motors, consists of an N-terminal highly conserved motor domain that can bind ATP and filamentous actin. The conformational changes in the motor domain generated by ATP hydrolysis are transmitted through the converter region into a large movement of the adjacent lever arm. Myosin VI contains a unique insert, the reverse gear, between the converter and the lever arm, which is stabilised by a calmodulin and swings the lever arm in the opposite direction to that of all the other myosins. A second calmodulin is bound through a canonical IQ motif in the neck region. The C-terminal tail contains a central single alpha helix (SAH) domain followed by a globular cargo-binding domain (CBD) (Peckham, 2011). Purified full-length myosin VI exists as a monomer (Lister et al., 2004), unless it is forced in vitro to dimerise artificially when including a leucine zipper motif or adding a myosin II coiled-coil motif.

Intriguingly, myosin VI might have the ability to function either as a monomer or as a processive dimeric motor, which may be regulated by ligand-induced dimerisation (Spudich et al., 2007; Yu et al., 2009); thus cargo-attachment might directly regulate its motor properties. Indeed, binding to the endocytic adaptor disabled-2 (Dab2) might induce myosin VI dimerisation, thereby generating a processive endocytic motor for the transport and clustering of transmembrane receptors in the area of a clathrin-coated pit or even for the short-range transport of internalised vesicles away from the cell surface. However, under certain conditions, for example, when it acts to regulate actin stabilisation during spermatid individualisation, myosin VI can fully function when it lacks the ability to dimerise. Under these circumstances it remains bound to actin for long periods, thus functioning as an actin tether (Noguchi et al., 2009; Noguchi et al., 2006). Furthermore, it has been suggested that, as load is increased on myosin VI, it switches from its transport function to an anchoring function (Altman et al., 2004), such as at the base of microvilli, stereocilia and on vesicles in the terminal synaptic bouton (Hertzano et al., 2008; Kisiel et al., 2011; Sakurai et al., 2011; Seiler et al., 2004). Thus, myosin VI most probably functions as both a weak processive motor and an actin tethering protein; however, further work in vivo is required to determine the conditions, the physiological regulation and the cellular mechanisms involved.

Myosin VI-interacting proteins

In the Snell's waltzer myosin VI knockout mouse – derived from a spontaneous deletion that resulted in a frameshift mutation and premature stop codon (Avraham et al., 1995) – and also in flies, the loss of functional myosin VI is associated with a wide spectrum of phenotypes, which suggest that myosin VI is a constituent of multiple protein complexes that function in a variety of different pathways (Fig. 1). In mammalian cells and also in Drosophila a substantial number of direct binding partners have been identified (Buss and Kendrick-Jones, 2011; Finan et al., 2011). However, additional transient or low-affinity myosin VI cargo adaptor molecules as well as larger protein complexes may be identified in the future, by using in situ crosslinking or spatially-restricted enzymatic tagging in living cells or tissues (Rhee et al., 2013). Although the ability of myosin VI to associate with multiple cargo adaptors under strict spatial and temporal control is very intriguing, it is not unique to this class of motor protein and highlights the mammalian paradigm that one motor can bind to a number of different cargoes. For example, cytoplasmic dynein – like myosin VI – is a unique motor due to its reverse directionality along microtubules, and transports a variety of cargoes – from endocytic and exocytic transport carriers, to organelles such as mitochondria and ribonucleoproteins (Allan, 2011). In yeast, the two classes of myosin V bind to a number of different receptors to drive movement of peroxisomes, mitochondria, ER, exocytic vesicles and the Golgi complex (Akhmanova and Hammer, 2010). The spatial and temporal regulation and coordination of motor-cargo attachment is likely to involve several mechanisms, including tissue-specific expression of motor protein isoforms and the stoichiometric concentration of adaptor proteins.

In the CBD of myosin VI, at least seven adaptor proteins compete to interact with two binding interfaces and/or subdomains that contain the amino acid motifs RRL or WWY – as suggested by the CBD susceptibility to tryptic digestion leading to two stable soluble peptides (Spudich et al., 2007; Yu et al., 2009) (Table 1). For example, the region containing the RRL motif interacts with nuclear dot protein 52 (NDP52), Traf6-binding protein (T6BP), optineurin and GAIP-interacting protein C-terminus (GIPC), and the region encompassing the WWY motif is required for binding to Tom1, Dab2 and lemur tyrosine kinase-2 (LMTK2) (Bunn et al., 1999; Chibalina et al., 2007; Morris et al., 2002; Morriswood et al., 2007; Sahlender et al., 2005; Spudich et al., 2007; Tumbarello et al., 2012). For other myosin VI adaptor molecules, such as SAP97 (Wu et al., 2002), otoferlin (Heidrych et al., 2009), phospholipase Cδ3 (Sakurai et al., 2011) and Dock7 (Majewski et al., 2012), the exact binding domain has not yet been determined. The two subdomains containing the RRL and WWY protein–protein interaction sites are linked by a phospholipid binding motif that interacts specifically and with high affinity to PtdIns(4,5)P2 (Spudich et al., 2007). The CBD also contains a conserved ubiquitin-binding motif (Penengo et al., 2006); however, ubiquitylated proteins that specifically bind to myosin VI have not yet been identified.

Table 1.
Characterised myosin VI cargo adaptors
graphic
graphic

Whereas myosin VI is widely expressed in most tissues, it exists in multiple tissue-specific splice isoforms with either a large insert (31 amino acids), a small insert (8 amino acids), both inserts, or no insert in the cargo-binding tail domain (Buss et al., 2001; Buss et al., 1998; Hasson and Mooseker, 1994). The tissue-specific expression of the four different myosin VI splice variants is most likely to be a further mechanism to control motor–cargo interactions and their subsequent subcellular localisation and function. In addition, the intracellular targeting of myosin VI might also depend on the stoichiometric expression level of its adaptors (Dance et al., 2004). However, the exact mechanism that regulates adaptor binding and cargo attachment is likely to be complex, involving a combination of different factors. Whereas the isoform with the large insert is preferentially expressed in polarised epithelia, the small and no-insert isoforms are found in non-polarised tissues (Buss et al., 2001). Interestingly, these alternatively spliced isoforms give myosin VI the ability to target to different intracellular locations through a yet undefined mechanism. Myosin VI that contains the large insert (LI) is recruited to clathrin-coated structures at the apical domain of polarised epithelial cells, where this motor is required for clathrin-dependent receptor-mediated endocytosis (Ameen and Apodaca, 2007; Buss et al., 2001). By contrast, myosin VI that lacks this insert (no insert isoform; NI) localises to uncoated vesicles near the cell periphery, where it is recruited by GIPC and the ESCRT-0 protein Tom1, and is suggested to either facilitate the movement of vesicles through the peripheral actin cytoskeleton towards the early endosome and/or to regulate a signalling platform for newly endocytosed receptors (Aschenbrenner et al., 2003; Dance et al., 2004; Naccache et al., 2006; Tumbarello et al., 2012). Myosin VI, however, is not only an endocytic motor but also functions in the exocytic pathway, in order to maintain Golgi morphology and facilitate the fusion of secretory vesicles at the plasma membrane (Bond et al., 2011; Sahlender et al., 2005; Warner et al., 2003). In addition, myosin VI loss-of-function affects the structural integrity of the stereocilia of the sensory hair cells of the inner ear (Self et al., 1999), the stability of cell–cell contacts in epithelia, efficient progression of cytokinesis (Arden et al., 2007) and directed cell migration (Chibalina et al., 2010; Geisbrecht and Montell, 2002). Further studies are required to show whether any of these complex cellular phenotypes that are caused by the loss-of-function of myosin VI are linked to trafficking defects in the endocytic or exocytic pathway.

Receptor internalisation

The tissue-specific expression of the different myosin VI isoforms enables this motor to function in at least two separate stages along the endocytic pathway because the isoforms differ in their intracellular targeting. In cell types that express myosin VI LI, several physiological phenotypes have been identified that demonstrate the role of myosin VI in apical endocytosis. For example, defects identified in the myosin VI knockout mouse show that myosin VI regulates endocytosis of the cystic fibrosis transmembrane conductance regulator from the apical domain of enterocytes. Also, in renal proximal tubule cells, myosin VI mediates the movement of two prominent sodium phosphate co-transporters, NaPi2a and NaPi2c, down the microvilli, and their subsequent removal from the brush-border membrane in response to parathyroid hormone stimulation (Ameen and Apodaca, 2007; Blaine et al., 2009; Lanzano et al., 2011). In polarised epithelial cells, myosin VI LI specifically associates – and can be coimmunoprecipitated – with the tumour suppressor Dab2, an endocytic adaptor protein that functions in cell signalling and the degradation of cell surface receptors (Inoue et al., 2002; Morris et al., 2002). Structurally, Dab2 contains a phosphotyrosine-binding region, which mediates its interaction with cell-surface receptors and phospholipids, multiple clathrin binding motifs as well as binding motifs for the clathrin adaptor AP-2; in addition it has a proline-rich region that interacts with the SH3 domain of Grb2 and a SYF motif that binds myosin VI (Mishra et al., 2002; Morris and Cooper, 2001; Xu et al., 1998). Thus, owing to the presence of binding motifs for AP-2, clathrin and myosin VI, Dab2 functions as an adaptor that links myosin VI LI to clathrin-mediated receptor endocytosis and degradation.

Cargo sorting in the endocytic pathway

Myosin VI NI is predominantly associated with a specific subpopulation of early endosomes underneath the plasma membrane that are characterised by the presence of the small GTPase Rab5 and its effector APPL1 (Chibalina et al., 2007; Tumbarello et al., 2012). These APPL1-positive early endosomes do not contain the early endosomal antigen 1 (EEA1) but, interestingly, recruit myosin VI together with its adaptor proteins GIPC and Tom1 (Chibalina et al., 2007; Tumbarello et al., 2012). During endosomal maturation, the synthesis of phosphatidylinositol 3-phosphate [PtdIns(3)P] and the recruitment of the PtdIns(3)P-binding protein EEA1 causes the loss of APPL1 from endosomes (Zoncu et al., 2009). APPL1 is a multifunctional adaptor protein containing a pleckstrin homology (PH) domain, a phosphotyrosine binding (PTB) domain and a leucine zipper motif. It binds through its PH domain to Rab5 and interacts through its very C-terminus with the PDZ-domain-containing protein GIPC (Miaczynska et al., 2004). In addition, it is targeted to lipid membranes through its BAR and PH domains, and associates through its PTB domain with the cytosolic domain of a diverse set of membrane receptors. The C-terminal PTB domain of APPL1 also mediates its direct interaction with signalling proteins, such as the serine/threonine kinase Akt and the phosphatidylinositol 3-kinase catalytic subunit p110α (PIK3CA). Thus, APPL1, by binding to several membrane receptors as well as signalling proteins, may regulate downstream signalling events on these specialised endosomes (Mitsuuchi et al., 1999). Thus, this distinct APPL1-positive early endocytic compartment may form a specialised class of signalling endosomes that act as a platform to regulate downstream signal transduction following receptor-mediated endocytosis; this, in turn, may then impact on various cellular functions, such as growth, migration, or homeostasis (Broussard et al., 2012; Cheng et al., 2012; Lin et al., 2006). However, little is known with regard to the mechanisms that determine which receptors are trafficked through this subset of endosomes, and/or the sorting processes that critically influence the lifetime of signalling events at this intracellular compartment.

The myosin VI adaptor protein GIPC, which was first identified as an interacting partner of the GTPase-activating protein RGS-GAIP for G-protein-coupled receptor subunit Gαi, can indirectly link myosin VI to APPL1 (De Vries et al., 1998; Varsano et al., 2006). GIPC acts as a scaffold to regulate receptor-mediated trafficking (Naccache et al., 2006; Varsano et al., 2006; Yi et al., 2007) and, after receptor internalisation, GIPC transiently associates with a pool of endocytic vesicles close to the plasma membrane (De Vries et al., 1998) before this compartment matures into an early endosome that is marked by EEA1. Thus, GIPC, through interaction with myosin VI, facilitates the retrograde movement of endosomes through the peripheral actin cytoskeleton to the early endosome (Aschenbrenner et al., 2003; Naccache et al., 2006; Spudich et al., 2007). This vesicle movement requires the motor activity of myosin VI, because expression of a myosin VI mutant that binds more tightly to actin (a rigor mutant) inhibits trafficking of peripheral vesicles (Aschenbrenner et al., 2004). Interestingly, both myosin VI and GIPC are recruited to the cleavage furrow in dividing cells and, like many endosomal proteins, are required during cytokinesis (Arden et al., 2007).

The second myosin VI adaptor protein that is present on APPL1-positive signalling endosomes is Tom1 and its close relative Tom1-like 2 (Tom1L2) (Fig. 2). The Drosophila homologue of Tom1/Tom1L2 (CG3529) was first identified as an interacting partner of myosin VI (Jaguar in Drosophila) by using a proteomics approach (Finan et al., 2011). We have confirmed this interaction in mammalian cells and demonstrated that Tom1 is an essential adaptor protein for intracellular targeting of myosin VI to early endosomes. In addition, both proteins function together to facilitate autophagosome maturation to enable autophagosome and lysosome fusion (Tumbarello et al., 2012). Tom1 was first identified as a target gene for the Myb oncogene (Burk et al., 1997) and, later characterised as part of the endosomal sorting complex required for transport (ESCRT type 0 family of proteins), which includes Stam1 and Stam2, and hepatocyte growth factor-regulated tyrosine kinase substrate (Hrs) (Seroussi et al., 1999). The ESCRT machinery functions as a multi-protein complex in the lysosomal degradation pathway. ESCRT-0 proteins mediate cargo recognition and sorting through interactions with ubiquitin moieties, and work in concert with the other ESCRT components to deform membranes and create intralumenal vesicles to form late endocytic multi-vesicular bodies (Henne et al., 2011). Tom1 is part of a smaller subfamily that includes the paralogues Tom1-like 1 (Tom1L1) and Tom1L2. Structurally, members of the Tom1 family consist of a conserved N-terminal Vps27–Hrs–STAM (VHS) domain, a central GAT domain that binds ubiquitin and a less-conserved C-terminal tail. Unlike the VHS domain in Hrs, Tom1 does not contain a FYVE membrane-interaction domain, but might interact directly with membranes through other potential regions within the VHS or indirectly through binding to its interacting partners endofin and Tollip (Misra et al., 2000; Seet et al., 2004; Yamakami et al., 2003). In addition, for cargo recognition, Tom1 contains two ubiquitin-binding motifs within its GAT domain and interacts directly with clathrin to mediate its recruitment to endosomes (Akutsu et al., 2005; Seet and Hong, 2005). Although all Tom1 family members are capable of interacting with the ESCRT-0 component Hrs, only Tom1L1 interacts with the ESCRT-I component TSG101. Moreover, only Tom1 can associate with endofin, suggesting that the family members have both redundant and unique functions (Puertollano, 2005; Seet et al., 2004; Yanagida-Ishizaki et al., 2008). Further divergence in the Tom1 family occurs within the C-terminal region, in which Tom1 and Tom1L2 are most similar, whereas Tom1L1 shows less similarity (Fig. 2). The C-terminus also contains a conserved IEXWL amino acid motif that is unique to Tom1 and Tom1L2, and mediates binding to myosin VI (F.B. and D.A.T., unpublished observations) (Fig. 2).

Fig. 2.

Tom1 family protein domains and their interacting partners. Tom1, Tom1L2 and Tom1L1, the organisation of their respective protein domains, and of the proteins that interact with them. VHS, Vps27–Hrs–STAM domain; GAT, GGA–Tom domain; Hrs, hepatocyte growth factor-regulated tyrosine kinase substrate; TSG101, tumour susceptibility gene101 protein (a subunit of the ESCRT-1 complex); Tyr-P, phosphotyrosine residue.

Fig. 2.

Tom1 family protein domains and their interacting partners. Tom1, Tom1L2 and Tom1L1, the organisation of their respective protein domains, and of the proteins that interact with them. VHS, Vps27–Hrs–STAM domain; GAT, GGA–Tom domain; Hrs, hepatocyte growth factor-regulated tyrosine kinase substrate; TSG101, tumour susceptibility gene101 protein (a subunit of the ESCRT-1 complex); Tyr-P, phosphotyrosine residue.

Although the cellular functions of Tom1 and Tom1L2 are not completely understood, Tom1 regulates signalling downstream of IL-1 and TNF-α receptors (Yamakami and Yokosawa, 2004) by linking them to the endosomal sorting machinery through its interaction with the membrane adaptor Toll-interaction protein (TOLLIP). The latter interacts with several Toll-like receptors, thereby mediating their lysosomal degradation (Brissoni et al., 2006). In addition, work from our lab has recently highlighted the importance of Tom1 and/or Tom1L2 in targeting myosin VI to early endosomes, and in potentially mediating cargo sorting in the endocytic pathway – which is required during autophagosome maturation (Tumbarello et al., 2012).

The third adaptor molecule associating with myosin VI in the early endocytic pathway is LMTK2, a transmembrane serine/threonine kinase and a binding partner for the protein phosphatase inhibitor 2 (PPP1R2) (Wang and Brautigan, 2002; Wang and Brautigan, 2006); it is phosphorylated and regulated by cyclin-dependent kinase-5 (Cdk5) (Kesavapany et al., 2003). LMTK2 directly binds to the WWY motif of myosin VI and regulates trafficking through the early endocytic pathway towards the endocytic recycling compartment (Chibalina et al., 2007). The open question remains whether myosin VI directs trafficking and targeting of LMTK2 or whether myosin VI activity is regulated through phosphorylation by LMTK2. Interestingly, both myosin VI and LMTK2 have been implicated in prostate cancer: myosin VI is highly overexpressed (Dunn et al., 2006) and genetic risk loci for LMTK2 have been identified (Eeles et al., 2008).

Autophagy receptors directly bind to myosin VI

In addition to cargo adaptor proteins, such as Dab2, GIPC, Tom1, Tom1L2 and LMTK2, that link myosin VI to distinct functions along the endocytic pathway, we identified three myosin-VI-interacting proteins – NDP52, T6BP and optineurin – that have recently been characterised as cargo-selective receptors in autophagy (Fig. 3). The targeting of ubiquitylated cargo for autophagy-dependent degradation requires specific adaptors that bind cargo through an ubiquitin-binding domain (UBD) and recruit LC3-associated autophagic membranes through an LC3-interaction region (LIR), thus facilitating autophagosome formation and degradation of its contents. Two of these autophagy receptors, NDP52 and optineurin, have an important function in innate immunity for the autophagy-dependent clearance of ubiquitin-coated Salmonella (Thurston et al., 2009; Wild et al., 2011). The third autophagy receptor that specifically binds to myosin VI is the NDP52-related protein T6BP, which also localises to autophagosomes and is required for autophagosome biogenesis (Tumbarello et al., 2012). Structurally, T6BP and NDP52 are most similar and consist of an N-terminal SKICH domain, followed by a non-canonical LIR, a coiled-coil domain and C-terminal zinc finger domains that mediate ubiquitin-binding (Morriswood et al., 2007; von Muhlinen et al., 2012). Optineurin also consists of multiple coiled-coil regions, a leucine zipper domain and a C-terminal zinc finger domain (Fig. 3). Optineurin contains multiple protein–protein interaction motifs, and binding to ubiquitin involves – like for its close homologue NF-κB essential modulator (NEMO) – a canonical motif within the central coiled-coiled domain together with its C-terminal zinc finger (Cordier et al., 2009; Laplantine et al., 2009). Recently, it has been shown that phosphorylation of optineurin at serine 177, adjacent to its LIR, regulates its interaction with LC3 and its ability to degrade ubiqutin-coated Salmonella (Wild et al., 2011). Recognition and clearance of ubiqutin-coated Salmonella by NDP52 is mediated through its interaction with the specific LC3C isoform (von Muhlinen et al., 2012). Interestingly, each of these three autophagy receptors is targeted sufficiently to autophagosomes under basal conditions and under conditions that induce selective autophagy (Tumbarello et al., 2012). Although this suggests that these myosin VI-associated autophagy adaptors have some redundant overlapping functions, it is also possible that they have distinct selective roles that are mediated by binding to specific LC3-isoforms or by displaying different specificity towards different types of ubiquitin chains, giving further support to a non-redundant function for the clearance of distinct substrates in the autophagy pathway (von Muhlinen et al., 2012; Wild et al., 2011).

Fig. 3.

Myosin VI autophagy adaptors, their protein domains and binding partners. Domain structure and interacting partners of T6BP, NDP52 and optineurin. LIR, LC3-interaction motif; SKICH, SKIP–carboxyl homology domain; CC, coiled-coil domain; ZF, zinc finger domain; GAL-8, galectin-8; HTT, huntingtin; CYLD, ubiquitin carboxyl-terminal hydrolase; TBK1, TANK-binding kinase 1; NAP1, nucleosome assembly protein 1; ITCH, E3 ubiquitin protein ligase; RIPK1, receptor-interacting serine/threonine-protein kinase 1; TBC1D17, TBC1 domain family member 17 (a Rab GTPase-activating protein).

Fig. 3.

Myosin VI autophagy adaptors, their protein domains and binding partners. Domain structure and interacting partners of T6BP, NDP52 and optineurin. LIR, LC3-interaction motif; SKICH, SKIP–carboxyl homology domain; CC, coiled-coil domain; ZF, zinc finger domain; GAL-8, galectin-8; HTT, huntingtin; CYLD, ubiquitin carboxyl-terminal hydrolase; TBK1, TANK-binding kinase 1; NAP1, nucleosome assembly protein 1; ITCH, E3 ubiquitin protein ligase; RIPK1, receptor-interacting serine/threonine-protein kinase 1; TBC1D17, TBC1 domain family member 17 (a Rab GTPase-activating protein).

We previously identified each of these autophagy adaptors – NDP52, optineurin and T6BP – as direct myosin VI-binding partners, and showed that myosin VI binding is mediated through the RRL motif in its C-terminal cargo-binding tail region (Morriswood et al., 2007; Sahlender et al., 2005). T6BP and NDP52 associate with myosin VI through their respective zinc finger domains, and optineurin through a central region that overlaps with the coiled-coiled- and ubiquitin-binding domains (Fig. 3). Functionally, these adaptors colocalise with myosin VI on autophagosomes, and depletion of the three adaptors inhibits autophagosome biogenesis (Tumbarello et al., 2012).

Myosin VI facilitates the delivery of endosomal membranes to autophagosomes

Myosin VI facilitates the delivery of endocytic membranes that contain the ESCRT-0 component Tom1 to autophagosomes, thereby mediating autophagosome maturation and its fusion with the lysosome (Tumbarello et al., 2012) (Fig. 4). How the ESCRT machinery works in concert with myosin VI in order to facilitate endocytic cargo delivery, vesicle fusion events and/or membrane deformation, is complicated by the fact that the exact temporal role of the ESCRT machinery during autophagosome maturation has not yet been established (Rusten and Stenmark, 2009).

Fig. 4.

Myosin VI functions in early endosomal sorting and autophagy maturation. (A) Ubiquitylated cargo, such as damaged organelles, aggregated proteins and pathogens, are recognised by autophagy adaptors. The latter include T6BP, NDP52 and optineurin, and recruit autophagic membranes through interactions with LC3, thus sequestering cargo within an autophagic vacuole. In addition, these autophagy adaptors localise to the outer limiting membrane of the autophagosome through interactions with LC3 (enlarged area). (B) Myosin VI localises to a Rab5- and APPL1-positive early endosomal compartment in the actin-rich cell-periphery just below the plasma membrane through interactions with its cargo-binding adaptors Tom1 and GIPC. (C,D) The recruitment of myosin VI by Tom1 to an early endosomal compartment results in the sorting and delivery of membrane associated cargo (yet to be determined) to the autophagosome (AP). Myosin VI is recruited to autophagosomes by binding to T6BP, NDP52 or optineurin present on the outer limiting membrane. Here, myosin VI can act either as (C) a tether that links endosomes to the actin filament network surrounding the autophagosomes or (D) a processive dimeric motor that moves endosomes on F-actin. By binding simultaneously to Tom1 and the autophagy adaptors, myosin VI facilitates the fusion of endosomes with autophagosomes as well as autophagosome maturation. Subsequently, the ‘mature’ autophagosome fuses with the lysosome (Lys) and its contents can be degraded. MVB, multivesicular bodies.

Fig. 4.

Myosin VI functions in early endosomal sorting and autophagy maturation. (A) Ubiquitylated cargo, such as damaged organelles, aggregated proteins and pathogens, are recognised by autophagy adaptors. The latter include T6BP, NDP52 and optineurin, and recruit autophagic membranes through interactions with LC3, thus sequestering cargo within an autophagic vacuole. In addition, these autophagy adaptors localise to the outer limiting membrane of the autophagosome through interactions with LC3 (enlarged area). (B) Myosin VI localises to a Rab5- and APPL1-positive early endosomal compartment in the actin-rich cell-periphery just below the plasma membrane through interactions with its cargo-binding adaptors Tom1 and GIPC. (C,D) The recruitment of myosin VI by Tom1 to an early endosomal compartment results in the sorting and delivery of membrane associated cargo (yet to be determined) to the autophagosome (AP). Myosin VI is recruited to autophagosomes by binding to T6BP, NDP52 or optineurin present on the outer limiting membrane. Here, myosin VI can act either as (C) a tether that links endosomes to the actin filament network surrounding the autophagosomes or (D) a processive dimeric motor that moves endosomes on F-actin. By binding simultaneously to Tom1 and the autophagy adaptors, myosin VI facilitates the fusion of endosomes with autophagosomes as well as autophagosome maturation. Subsequently, the ‘mature’ autophagosome fuses with the lysosome (Lys) and its contents can be degraded. MVB, multivesicular bodies.

It is possible that docking of Tom1-positive endosomes to autophagosomes indirectly involves myosin VI through its association with the autophagy adaptors T6BP, NDP52 and optineurin. These autophagy receptors might have a dual role by acting not only as adaptors during autophagosome biogenesis to facilitate cargo recognition and recruitment of autophagocytic membranes, but also by decorating the cytosol-facing outer membrane of the autophagosome, where they act as myosin VI receptors, thereby facilitating the delivery and docking of the endocytic membrane, or even fusion events required for its maturation. The ubiquitin-binding region in T6BP, NDP52 and optineurin overlaps with their myosin VI-interaction domain, suggesting that ubiquitylation and myosin VI binding is mutually exclusive, thereby, supporting our model that the autophagy adaptors function as multiple ligand receptors on both the inner and outer autophagocytic membranes (Fig. 4).

If myosin VI simultaneously binds to Tom1 on endosomal membranes through its tail domain that contains the WWY-motif as well as to autophagy receptors through its second cargo-binding site, the RRL-motif, then myosin VI could bring both of its cargo binding subdomains into close proximity to prepare for SNARE-mediated fusion of endosomes with autophagosomes. Indeed, we have shown previously (Spudich et al., 2007) that the CBD of monomeric myosin VI consists of two subdomains that can be separated: one contains the RRL motif that is bound by optineurin, NDP52 and T6BP; the other contains the WWY motif that is bound by Tom1. In vitro studies have further shown that binding to optineurin and Dab2 can induce dimerisation of myosin VI (Phichith et al., 2009). Therefore, it will be important to test whether monomeric and dimeric myosin VI molecules can bind two different adaptor proteins (cargo) because, if this were the case, it could lead to tethering or positioning and, ultimately, crosslinking of different subcellular compartments. A tethering function might not be regarded as the typical activity for a motor protein because, characteristically, it drives transport of cargo along a cytoskeletal track. However, in vitro motility measurements have shown that myosin VI is a slow motor that moves with a maximum speed of ∼30–60 nm/second. In the cell, it might move even slower. Therefore, consistent with its known motor properties, myosin VI is likely to be involved in short-distance transport, such as positioning and tethering of endosomes around autophagosomes or other vesicular organelles to enhance fusion. In fact, in the final stages of the secretory pathway, myosin VI is required, through interaction with optineurin, for the fusion events that take place between secretory vesicles and the plasma membrane (Bond et al., 2011).

It is clear that myosin VI is a multifunctional motor protein that operates during multiple steps of the endocytic pathway. Myosin VI dysfunction has been predominantly associated with hearing loss due to the degeneration of stereocilia in the inner ear. However, the consequences of loss-of-function of many of its adaptors (Table 1) suggests that a number of other pathologies are associated with myosin VI dysfunction, which results in defects along the endocytic and autophagy-dependent pathways. These defects would be likely to lead to a disruption in protein quality control, cell homeostasis and innate immunity. Furthermore, because myosin VI is overexpressed in ovarian and prostate cancers (Dunn et al., 2006; Yoshida et al., 2004), the identification of associated adaptors and functional consequences of this overexpression on cancer progression is of great importance. Further characterisation of the myosin VI knockout mice will also yield further insights into the physiological effects of myosin VI loss-of-function. It has already been established that the physiological loss-of-function of myosin VI is associated with astrogliosis in the brain, hypertrophic cardiomyopathy, defects in endocytosis in neuronal and epithelial tissues, disruption in the integrity of the intestinal brush border, and various metabolic dysfunctions (Ameen and Apodaca, 2007; Hegan et al., 2012; Mohiddin et al., 2004; Osterweil et al., 2005) (see also ‘myo6’ at the Mouse Resources Portal of the Wellcome Trust Sanger Institute, http://www.sanger.ac.uk/mouseportal/search?query=myo6). Further work is required, however, before we can understand the myosin VI-dependent mechanisms that underlay these pathologies.

Importantly, in order to understand the mechanisms behind the above stated pathologies, we need to apprehend the complex nature of the spatiotemporal regulation of myosin VI with respect to its functions within the cell. This will entail understanding how the CBD of myosin VI interacts with different cargo adaptors and whether there is a further layer of regulation, such as phosphorylation, ubiquitylation or cargo-induced structural modulations. In addition, more cargo adaptors are likely to be identified when new techniques for probing these interactions are developed, as illustrated recently by Finan et al. (Finan et al., 2011). Regardless of whether these new or previously identified cargo-binding adaptors are tissue-specific or regulatory in function, they make the mechanism of how myosin VI functions at distinct phases of membrane trafficking more complex. New and innovative approaches are, therefore, needed to delineate the molecular mechanisms of the interaction between cargo adaptors and myosin VI.

Funding

This work is financially supported by the Wellcome Trust and the Medical Research Council (F.B. and J.K.-J.). The CIMR is in receipt of a strategic award from the Wellcome Trust (100410).

Aguilera
M. O.
,
Berón
W.
,
Colombo
M. I.
(
2012
).
The actin cytoskeleton participates in the early events of autophagosome formation upon starvation induced autophagy.
Autophagy
8
,
1590
1603
.
Akhmanova
A.
,
Hammer
J. A.
 III
(
2010
).
Linking molecular motors to membrane cargo.
Curr. Opin. Cell Biol.
22
,
479
487
.
Akutsu
M.
,
Kawasaki
M.
,
Katoh
Y.
,
Shiba
T.
,
Yamaguchi
Y.
,
Kato
R.
,
Kato
K.
,
Nakayama
K.
,
Wakatsuki
S.
(
2005
).
Structural basis for recognition of ubiquitinated cargo by Tom1-GAT domain.
FEBS Lett.
579
,
5385
5391
.
Allan
V. J.
(
2011
).
Cytoplasmic dynein.
Biochem. Soc. Trans.
39
,
1169
1178
.
Altman
D.
,
Sweeney
H. L.
,
Spudich
J. A.
(
2004
).
The mechanism of myosin VI translocation and its load-induced anchoring.
Cell
116
,
737
749
.
Ameen
N.
,
Apodaca
G.
(
2007
).
Defective CFTR apical endocytosis and enterocyte brush border in myosin VI-deficient mice.
Traffic
8
,
998
1006
.
Anitei
M.
,
Hoflack
B.
(
2011
).
Bridging membrane and cytoskeleton dynamics in the secretory and endocytic pathways.
Nat. Cell Biol.
14
,
11
19
.
Arden
S. D.
,
Puri
C.
,
Au
J. S.
,
Kendrick-Jones
J.
,
Buss
F.
(
2007
).
Myosin VI is required for targeted membrane transport during cytokinesis.
Mol. Biol. Cell
18
,
4750
4761
.
Aschenbrenner
L.
,
Lee
T.
,
Hasson
T.
(
2003
).
Myo6 facilitates the translocation of endocytic vesicles from cell peripheries.
Mol. Biol. Cell
14
,
2728
2743
.
Aschenbrenner
L.
,
Naccache
S. N.
,
Hasson
T.
(
2004
).
Uncoated endocytic vesicles require the unconventional myosin, Myo6, for rapid transport through actin barriers.
Mol. Biol. Cell
15
,
2253
2263
.
Avraham
K. B.
,
Hasson
T.
,
Steel
K. P.
,
Kingsley
D. M.
,
Russell
L. B.
,
Mooseker
M. S.
,
Copeland
N. G.
,
Jenkins
N. A.
(
1995
).
The mouse Snell's waltzer deafness gene encodes an unconventional myosin required for structural integrity of inner ear hair cells.
Nat. Genet.
11
,
369
375
.
Axe
E. L.
,
Walker
S. A.
,
Manifava
M.
,
Chandra
P.
,
Roderick
H. L.
,
Habermann
A.
,
Griffiths
G.
,
Ktistakis
N. T.
(
2008
).
Autophagosome formation from membrane compartments enriched in phosphatidylinositol 3-phosphate and dynamically connected to the endoplasmic reticulum.
J. Cell Biol.
182
,
685
701
.
Blaine
J.
,
Okamura
K.
,
Giral
H.
,
Breusegem
S.
,
Caldas
Y.
,
Millard
A.
,
Barry
N.
,
Levi
M.
(
2009
).
PTH-induced internalization of apical membrane NaPi2a: role of actin and myosin VI.
Am. J. Physiol. Cell Physiol.
297
,
C1339
C1346
.
Bloemink
M. J.
,
Geeves
M. A.
(
2011
).
Shaking the myosin family tree: biochemical kinetics defines four types of myosin motor.
Semin. Cell Dev. Biol.
22
,
961
967
.
Bond
L. M.
,
Peden
A. A.
,
Kendrick-Jones
J.
,
Sellers
J. R.
,
Buss
F.
(
2011
).
Myosin VI and its binding partner optineurin are involved in secretory vesicle fusion at the plasma membrane.
Mol. Biol. Cell
22
,
54
65
.
Brissoni
B.
,
Agostini
L.
,
Kropf
M.
,
Martinon
F.
,
Swoboda
V.
,
Lippens
S.
,
Everett
H.
,
Aebi
N.
,
Janssens
S.
,
Meylan
E.
 et al. (
2006
).
Intracellular trafficking of interleukin-1 receptor I requires Tollip.
Curr. Biol.
16
,
2265
2270
.
Broussard
J. A.
,
Lin
W. H.
,
Majumdar
D.
,
Anderson
B.
,
Eason
B.
,
Brown
C. M.
,
Webb
D. J.
(
2012
).
The endosomal adaptor protein APPL1 impairs the turnover of leading edge adhesions to regulate cell migration.
Mol. Biol. Cell
23
,
1486
1499
.
Bunn
R. C.
,
Jensen
M. A.
,
Reed
B. C.
(
1999
).
Protein interactions with the glucose transporter binding protein GLUT1CBP that provide a link between GLUT1 and the cytoskeleton.
Mol. Biol. Cell
10
,
819
832
.
Burk
O.
,
Worpenberg
S.
,
Haenig
B.
,
Klempnauer
K. H.
(
1997
).
tom-1, a novel v-Myb target gene expressed in AMV- and E26-transformed myelomonocytic cells.
EMBO J.
16
,
1371
1380
.
Buss
F.
,
Kendrick-Jones
J.
(
2011
).
Multifunctional myosin VI has a multitude of cargoes.
Proc. Natl. Acad. Sci. USA
108
,
5927
5928
.
Buss
F.
,
Kendrick-Jones
J.
,
Lionne
C.
,
Knight
A. E.
,
Côté
G. P.
,
Paul Luzio
J.
(
1998
).
The localization of myosin VI at the golgi complex and leading edge of fibroblasts and its phosphorylation and recruitment into membrane ruffles of A431 cells after growth factor stimulation.
J. Cell Biol.
143
,
1535
1545
.
Buss
F.
,
Arden
S. D.
,
Lindsay
M.
,
Luzio
J. P.
,
Kendrick-Jones
J.
(
2001
).
Myosin VI isoform localized to clathrin-coated vesicles with a role in clathrin-mediated endocytosis.
EMBO J.
20
,
3676
3684
.
Cheng
K. K.
,
Lam
K. S.
,
Wu
D.
,
Wang
Y.
,
Sweeney
G.
,
Hoo
R. L.
,
Zhang
J.
,
Xu
A.
(
2012
).
APPL1 potentiates insulin secretion in pancreatic β cells by enhancing protein kinase Akt-dependent expression of SNARE proteins in mice.
Proc. Natl. Acad. Sci. USA
109
,
8919
8924
.
Chibalina
M. V.
,
Seaman
M. N.
,
Miller
C. C.
,
Kendrick-Jones
J.
,
Buss
F.
(
2007
).
Myosin VI and its interacting protein LMTK2 regulate tubule formation and transport to the endocytic recycling compartment.
J. Cell Sci.
120
,
4278
4288
.
Chibalina
M. V.
,
Poliakov
A.
,
Kendrick-Jones
J.
,
Buss
F.
(
2010
).
Myosin VI and optineurin are required for polarized EGFR delivery and directed migration.
Traffic
11
,
1290
1303
.
Cordier
F.
,
Grubisha
O.
,
Traincard
F.
,
Véron
M.
,
Delepierre
M.
,
Agou
F.
(
2009
).
The zinc finger of NEMO is a functional ubiquitin-binding domain.
J. Biol. Chem.
284
,
2902
2907
.
Dance
A. L.
,
Miller
M.
,
Seragaki
S.
,
Aryal
P.
,
White
B.
,
Aschenbrenner
L.
,
Hasson
T.
(
2004
).
Regulation of myosin-VI targeting to endocytic compartments.
Traffic
5
,
798
813
.
De Vries
L.
,
Lou
X.
,
Zhao
G.
,
Zheng
B.
,
Farquhar
M. G.
(
1998
).
GIPC, a PDZ domain containing protein, interacts specifically with the C terminus of RGS-GAIP.
Proc. Natl. Acad. Sci. USA
95
,
12340
12345
.
Dunn
T. A.
,
Chen
S.
,
Faith
D. A.
,
Hicks
J. L.
,
Platz
E. A.
,
Chen
Y.
,
Ewing
C. M.
,
Sauvageot
J.
,
Isaacs
W. B.
,
De Marzo
A. M.
 et al. (
2006
).
A novel role of myosin VI in human prostate cancer.
Am. J. Pathol.
169
,
1843
1854
.
Eeles
R. A.
,
Kote-Jarai
Z.
,
Giles
G. G.
,
Olama
A. A.
,
Guy
M.
,
Jugurnauth
S. K.
,
Mulholland
S.
,
Leongamornlert
D. A.
,
Edwards
S. M.
,
Morrison
J.
, et al. 
UK Genetic Prostate Cancer Study Collaborators; British Association of Urological Surgeons' Section of Oncology; UK ProtecT Study Collaborators
(
2008
).
Multiple newly identified loci associated with prostate cancer susceptibility.
Nat. Genet.
40
,
316
321
.
Filimonenko
M.
,
Stuffers
S.
,
Raiborg
C.
,
Yamamoto
A.
,
Malerød
L.
,
Fisher
E. M.
,
Isaacs
A.
,
Brech
A.
,
Stenmark
H.
,
Simonsen
A.
(
2007
).
Functional multivesicular bodies are required for autophagic clearance of protein aggregates associated with neurodegenerative disease.
J. Cell Biol.
179
,
485
500
.
Finan
D.
,
Hartman
M. A.
,
Spudich
J. A.
(
2011
).
Proteomics approach to study the functions of Drosophila myosin VI through identification of multiple cargo-binding proteins.
Proc. Natl. Acad. Sci. USA
108
,
5566
5571
.
Geisbrecht
E. R.
,
Montell
D. J.
(
2002
).
Myosin VI is required for E-cadherin-mediated border cell migration.
Nat. Cell Biol.
4
,
616
620
.
Gottlieb
T. A.
,
Ivanov
I. E.
,
Adesnik
M.
,
Sabatini
D. D.
(
1993
).
Actin microfilaments play a critical role in endocytosis at the apical but not the basolateral surface of polarized epithelial cells.
J. Cell Biol.
120
,
695
710
.
Hailey
D. W.
,
Rambold
A. S.
,
Satpute-Krishnan
P.
,
Mitra
K.
,
Sougrat
R.
,
Kim
P. K.
,
Lippincott-Schwartz
J.
(
2010
).
Mitochondria supply membranes for autophagosome biogenesis during starvation.
Cell
141
,
656
667
.
Hartman
M. A.
,
Finan
D.
,
Sivaramakrishnan
S.
,
Spudich
J. A.
(
2011
).
Principles of unconventional myosin function and targeting.
Annu. Rev. Cell Dev. Biol.
27
,
133
155
.
Hasson
T.
,
Mooseker
M. S.
(
1994
).
Porcine myosin-VI: characterization of a new mammalian unconventional myosin.
J. Cell Biol.
127
,
425
440
.
Hegan
P. S.
,
Giral
H.
,
Levi
M.
,
Mooseker
M. S.
(
2012
).
Myosin VI is required for maintenance of brush border structure, composition, and membrane trafficking functions in the intestinal epithelial cell.
Cytoskeleton (Hoboken)
69
,
235
251
.
Heidrych
P.
,
Zimmermann
U.
,
Kuhn
S.
,
Franz
C.
,
Engel
J.
,
Duncker
S. V.
,
Hirt
B.
,
Pusch
C. M.
,
Ruth
P.
,
Pfister
M.
 et al. (
2009
).
Otoferlin interacts with myosin VI: implications for maintenance of the basolateral synaptic structure of the inner hair cell.
Hum. Mol. Genet.
18
,
2779
2790
.
Henne
W. M.
,
Buchkovich
N. J.
,
Emr
S. D.
(
2011
).
The ESCRT pathway.
Dev. Cell
21
,
77
91
.
Hertzano
R.
,
Shalit
E.
,
Rzadzinska
A. K.
,
Dror
A. A.
,
Song
L.
,
Ron
U.
,
Tan
J. T.
,
Shitrit
A. S.
,
Fuchs
H.
,
Hasson
T.
 et al. (
2008
).
A Myo6 mutation destroys coordination between the myosin heads, revealing new functions of myosin VI in the stereocilia of mammalian inner ear hair cells.
PLoS Genet.
4
,
e1000207
.
Inoue
A.
,
Sato
O.
,
Homma
K.
,
Ikebe
M.
(
2002
).
DOC-2/DAB2 is the binding partner of myosin VI.
Biochem. Biophys. Res. Commun.
292
,
300
307
.
Jackman
M. R.
,
Shurety
W.
,
Ellis
J. A.
,
Luzio
J. P.
(
1994
).
Inhibition of apical but not basolateral endocytosis of ricin and folate in Caco-2 cells by cytochalasin D. J. Cell Sci.
107
,
2547
2556
.
Kesavapany
S.
,
Lau
K. F.
,
Ackerley
S.
,
Banner
S. J.
,
Shemilt
S. J.
,
Cooper
J. D.
,
Leigh
P. N.
,
Shaw
C. E.
,
McLoughlin
D. M.
,
Miller
C. C.
(
2003
).
Identification of a novel, membrane-associated neuronal kinase, cyclin-dependent kinase 5/p35-regulated kinase.
J. Neurosci.
23
,
4975
4983
.
Kisiel
M.
,
Majumdar
D.
,
Campbell
S.
,
Stewart
B. A.
(
2011
).
Myosin VI contributes to synaptic transmission and development at the Drosophila neuromuscular junction.
BMC Neurosci.
12
,
65
.
Lanzano
L.
,
Lei
T.
,
Okamura
K.
,
Giral
H.
,
Caldas
Y.
,
Masihzadeh
O.
,
Gratton
E.
,
Levi
M.
,
Blaine
J.
(
2011
).
Differential modulation of the molecular dynamics of the type IIa and IIc sodium phosphate cotransporters by parathyroid hormone.
Am. J. Physiol. Cell Physiol.
301
,
C850
C861
.
Laplantine
E.
,
Fontan
E.
,
Chiaravalli
J.
,
Lopez
T.
,
Lakisic
G.
,
Véron
M.
,
Agou
F.
,
Israël
A.
(
2009
).
NEMO specifically recognizes K63-linked poly-ubiquitin chains through a new bipartite ubiquitin-binding domain.
EMBO J.
28
,
2885
2895
.
Lee
J. A.
,
Beigneux
A.
,
Ahmad
S. T.
,
Young
S. G.
,
Gao
F. B.
(
2007
).
ESCRT-III dysfunction causes autophagosome accumulation and neurodegeneration.
Curr. Biol.
17
,
1561
1567
.
Lee
J. Y.
,
Koga
H.
,
Kawaguchi
Y.
,
Tang
W.
,
Wong
E.
,
Gao
Y. S.
,
Pandey
U. B.
,
Kaushik
S.
,
Tresse
E.
,
Lu
J.
 et al. (
2010
).
HDAC6 controls autophagosome maturation essential for ubiquitin-selective quality-control autophagy.
EMBO J.
29
,
969
980
.
Lin
D. C.
,
Quevedo
C.
,
Brewer
N. E.
,
Bell
A.
,
Testa
J. R.
,
Grimes
M. L.
,
Miller
F. D.
,
Kaplan
D. R.
(
2006
).
APPL1 associates with TrkA and GIPC1 and is required for nerve growth factor-mediated signal transduction.
Mol. Cell. Biol.
26
,
8928
8941
.
Lister
I.
,
Schmitz
S.
,
Walker
M.
,
Trinick
J.
,
Buss
F.
,
Veigel
C.
,
Kendrick-Jones
J.
(
2004
).
A monomeric myosin VI with a large working stroke.
EMBO J.
23
,
1729
1738
.
Majewski
L.
,
Sobczak
M.
,
Havrylov
S.
,
Jóźwiak
J.
,
Rędowicz
M. J.
(
2012
).
Dock7: a GEF for Rho-family GTPases and a novel myosin VI-binding partner in neuronal PC12 cells.
Biochem. Cell Biol.
90
,
565
574
.
McCullough
J.
,
Colf
L. A.
,
Sundquist
W. I.
(
2013
).
Membrane fission reactions of the mammalian ESCRT pathway.
Annu. Rev. Biochem.
[Epub ahead of print] doi:10.1146/annurev-biochem-072909-101058
Miaczynska
M.
,
Christoforidis
S.
,
Giner
A.
,
Shevchenko
A.
,
Uttenweiler-Joseph
S.
,
Habermann
B.
,
Wilm
M.
,
Parton
R. G.
,
Zerial
M.
(
2004
).
APPL proteins link Rab5 to nuclear signal transduction via an endosomal compartment.
Cell
116
,
445
456
.
Mishra
S. K.
,
Keyel
P. A.
,
Hawryluk
M. J.
,
Agostinelli
N. R.
,
Watkins
S. C.
,
Traub
L. M.
(
2002
).
Disabled-2 exhibits the properties of a cargo-selective endocytic clathrin adaptor.
EMBO J.
21
,
4915
4926
.
Misra
S.
,
Beach
B. M.
,
Hurley
J. H.
(
2000
).
Structure of the VHS domain of human Tom1 (target of myb 1): insights into interactions with proteins and membranes.
Biochemistry
39
,
11282
11290
.
Mitsuuchi
Y.
,
Johnson
S. W.
,
Sonoda
G.
,
Tanno
S.
,
Golemis
E. A.
,
Testa
J. R.
(
1999
).
Identification of a chromosome 3p14.3-21.1 gene, APPL, encoding an adaptor molecule that interacts with the oncoprotein-serine/threonine kinase AKT2.
Oncogene
18
,
4891
4898
.
Mohiddin
S. A.
,
Ahmed
Z. M.
,
Griffith
A. J.
,
Tripodi
D.
,
Friedman
T. B.
,
Fananapazir
L.
,
Morell
R. J.
(
2004
).
Novel association of hypertrophic cardiomyopathy, sensorineural deafness, and a mutation in unconventional myosin VI (MYO6).
J. Med. Genet.
41
,
309
314
.
Mooren
O. L.
,
Galletta
B. J.
,
Cooper
J. A.
(
2012
).
Roles for actin assembly in endocytosis.
Annu. Rev. Biochem.
81
,
661
686
.
Morris
S. M.
,
Cooper
J. A.
(
2001
).
Disabled-2 colocalizes with the LDLR in clathrin-coated pits and interacts with AP-2.
Traffic
2
,
111
123
.
Morris
S. M.
,
Arden
S. D.
,
Roberts
R. C.
,
Kendrick-Jones
J.
,
Cooper
J. A.
,
Luzio
J. P.
,
Buss
F.
(
2002
).
Myosin VI binds to and localises with Dab2, potentially linking receptor-mediated endocytosis and the actin cytoskeleton.
Traffic
3
,
331
341
.
Morriswood
B.
,
Ryzhakov
G.
,
Puri
C.
,
Arden
S. D.
,
Roberts
R.
,
Dendrou
C.
,
Kendrick-Jones
J.
,
Buss
F.
(
2007
).
T6BP and NDP52 are myosin VI binding partners with potential roles in cytokine signalling and cell adhesion.
J. Cell Sci.
120
,
2574
2585
.
Naccache
S. N.
,
Hasson
T.
,
Horowitz
A.
(
2006
).
Binding of internalized receptors to the PDZ domain of GIPC/synectin recruits myosin VI to endocytic vesicles.
Proc. Natl. Acad. Sci. USA
103
,
12735
12740
.
Noguchi
T.
,
Lenartowska
M.
,
Miller
K. G.
(
2006
).
Myosin VI stabilizes an actin network during Drosophila spermatid individualization.
Mol. Biol. Cell
17
,
2559
2571
.
Noguchi
T.
,
Frank
D. J.
,
Isaji
M.
,
Miller
K. G.
(
2009
).
Coiled-coil-mediated dimerization is not required for myosin VI to stabilize actin during spermatid individualization in Drosophila melanogaster.
Mol. Biol. Cell
20
,
358
367
.
Osterweil
E.
,
Wells
D. G.
,
Mooseker
M. S.
(
2005
).
A role for myosin VI in postsynaptic structure and glutamate receptor endocytosis.
J. Cell Biol.
168
,
329
338
.
Peckham
M.
(
2011
).
Coiled coils and SAH domains in cytoskeletal molecular motors.
Biochem. Soc. Trans.
39
,
1142
1148
.
Penengo
L.
,
Mapelli
M.
,
Murachelli
A. G.
,
Confalonieri
S.
,
Magri
L.
,
Musacchio
A.
,
Di Fiore
P. P.
,
Polo
S.
,
Schneider
T. R.
(
2006
).
Crystal structure of the ubiquitin binding domains of rabex-5 reveals two modes of interaction with ubiquitin.
Cell
124
,
1183
1195
.
Phichith
D.
,
Travaglia
M.
,
Yang
Z.
,
Liu
X.
,
Zong
A. B.
,
Safer
D.
,
Sweeney
H. L.
(
2009
).
Cargo binding induces dimerization of myosin VI.
Proc. Natl. Acad. Sci. USA
106
,
17320
17324
.
Pollard
T. D.
,
Blanchoin
L.
,
Mullins
R. D.
(
2000
).
Molecular mechanisms controlling actin filament dynamics in nonmuscle cells.
Annu. Rev. Biophys. Biomol. Struct.
29
,
545
576
.
Puertollano
R.
(
2005
).
Interactions of TOM1L1 with the multivesicular body sorting machinery.
J. Biol. Chem.
280
,
9258
9264
.
Ravikumar
B.
,
Moreau
K.
,
Jahreiss
L.
,
Puri
C.
,
Rubinsztein
D. C.
(
2010
).
Plasma membrane contributes to the formation of pre-autophagosomal structures.
Nat. Cell Biol.
12
,
747
757
.
Rhee
H. W.
,
Zou
P.
,
Udeshi
N. D.
,
Martell
J. D.
,
Mootha
V. K.
,
Carr
S. A.
,
Ting
A. Y.
(
2013
).
Proteomic mapping of mitochondria in living cells via spatially restricted enzymatic tagging.
Science
339
,
1328
1331
.
Rusten
T. E.
,
Stenmark
H.
(
2009
).
How do ESCRT proteins control autophagy?
J. Cell Sci.
122
,
2179
2183
.
Rusten
T. E.
,
Vaccari
T.
,
Lindmo
K.
,
Rodahl
L. M.
,
Nezis
I. P.
,
Sem-Jacobsen
C.
,
Wendler
F.
,
Vincent
J. P.
,
Brech
A.
,
Bilder
D.
 et al. (
2007
).
ESCRTs and Fab1 regulate distinct steps of autophagy.
Curr. Biol.
17
,
1817
1825
.
Sahlender
D. A.
,
Roberts
R. C.
,
Arden
S. D.
,
Spudich
G.
,
Taylor
M. J.
,
Luzio
J. P.
,
Kendrick-Jones
J.
,
Buss
F.
(
2005
).
Optineurin links myosin VI to the Golgi complex and is involved in Golgi organization and exocytosis.
J. Cell Biol.
169
,
285
295
.
Sakurai
K.
,
Hirata
M.
,
Yamaguchi
H.
,
Nakamura
Y.
,
Fukami
K.
(
2011
).
Phospholipase Cδ3 is a novel binding partner of myosin VI and functions as anchoring of myosin VI on plasma membrane.
Adv. Enzyme Regul.
51
,
171
181
.
Seet
L. F.
,
Hong
W.
(
2005
).
Endofin recruits clathrin to early endosomes via TOM1.
J. Cell Sci.
118
,
575
587
.
Seet
L. F.
,
Liu
N.
,
Hanson
B. J.
,
Hong
W.
(
2004
).
Endofin recruits TOM1 to endosomes.
J. Biol. Chem.
279
,
4670
4679
.
Seiler
C.
,
Ben-David
O.
,
Sidi
S.
,
Hendrich
O.
,
Rusch
A.
,
Burnside
B.
,
Avraham
K. B.
,
Nicolson
T.
(
2004
).
Myosin VI is required for structural integrity of the apical surface of sensory hair cells in zebrafish.
Dev. Biol.
272
,
328
338
.
Self
T.
,
Sobe
T.
,
Copeland
N. G.
,
Jenkins
N. A.
,
Avraham
K. B.
,
Steel
K. P.
(
1999
).
Role of myosin VI in the differentiation of cochlear hair cells.
Dev. Biol.
214
,
331
341
.
Seroussi
E.
,
Kedra
D.
,
Kost-Alimova
M.
,
Sandberg-Nordqvist
A. C.
,
Fransson
I.
,
Jacobs
J. F.
,
Fu
Y.
,
Pan
H. Q.
,
Roe
B. A.
,
Imreh
S.
 et al. (
1999
).
TOM1 genes map to human chromosome 22q13.1 and mouse chromosome 8C1 and encode proteins similar to the endosomal proteins HGS and STAM.
Genomics
57
,
380
388
.
Simonsen
A.
,
Tooze
S. A.
(
2009
).
Coordination of membrane events during autophagy by multiple class III PI3-kinase complexes.
J. Cell Biol.
186
,
773
782
.
Spudich
G.
,
Chibalina
M. V.
,
Au
J. S.
,
Arden
S. D.
,
Buss
F.
,
Kendrick-Jones
J.
(
2007
).
Myosin VI targeting to clathrin-coated structures and dimerization is mediated by binding to Disabled-2 and PtdIns(4,5)P2.
Nat. Cell Biol.
9
,
176
183
.
Thurston
T. L.
,
Ryzhakov
G.
,
Bloor
S.
,
von Muhlinen
N.
,
Randow
F.
(
2009
).
The TBK1 adaptor and autophagy receptor NDP52 restricts the proliferation of ubiquitin-coated bacteria.
Nat. Immunol.
10
,
1215
1221
.
Tumbarello
D. A.
,
Waxse
B. J.
,
Arden
S. D.
,
Bright
N. A.
,
Kendrick-Jones
J.
,
Buss
F.
(
2012
).
Autophagy receptors link myosin VI to autophagosomes to mediate Tom1-dependent autophagosome maturation and fusion with the lysosome.
Nat. Cell Biol.
14
,
1024
1035
.
Varsano
T.
,
Dong
M. Q.
,
Niesman
I.
,
Gacula
H.
,
Lou
X.
,
Ma
T.
,
Testa
J. R.
,
Yates
J. R.
 3rd
,
Farquhar
M. G.
(
2006
).
GIPC is recruited by APPL to peripheral TrkA endosomes and regulates TrkA trafficking and signaling.
Mol. Cell. Biol.
26
,
8942
8952
.
von Muhlinen
N.
,
Akutsu
M.
,
Ravenhill
B. J.
,
Foeglein
A.
,
Bloor
S.
,
Rutherford
T. J.
,
Freund
S. M.
,
Komander
D.
,
Randow
F.
(
2012
).
LC3C, bound selectively by a noncanonical LIR motif in NDP52, is required for antibacterial autophagy.
Mol. Cell
48
,
329
342
.
Wang
H.
,
Brautigan
D. L.
(
2002
).
A novel transmembrane Ser/Thr kinase complexes with protein phosphatase-1 and inhibitor-2.
J. Biol. Chem.
277
,
49605
49612
.
Wang
H.
,
Brautigan
D. L.
(
2006
).
Peptide microarray analysis of substrate specificity of the transmembrane Ser/Thr kinase KPI-2 reveals reactivity with cystic fibrosis transmembrane conductance regulator and phosphorylase.
Mol. Cell. Proteomics
5
,
2124
2130
.
Warner
C. L.
,
Stewart
A.
,
Luzio
J. P.
,
Steel
K. P.
,
Libby
R. T.
,
Kendrick-Jones
J.
,
Buss
F.
(
2003
).
Loss of myosin VI reduces secretion and the size of the Golgi in fibroblasts from Snell's waltzer mice.
EMBO J.
22
,
569
579
.
Wild
P.
,
Farhan
H.
,
McEwan
D. G.
,
Wagner
S.
,
Rogov
V. V.
,
Brady
N. R.
,
Richter
B.
,
Korac
J.
,
Waidmann
O.
,
Choudhary
C.
 et al. (
2011
).
Phosphorylation of the autophagy receptor optineurin restricts Salmonella growth.
Science
333
,
228
233
.
Wu
H.
,
Nash
J. E.
,
Zamorano
P.
,
Garner
C. C.
(
2002
).
Interaction of SAP97 with minus-end-directed actin motor myosin VI. Implications for AMPA receptor trafficking.
J. Biol. Chem.
277
,
30928
30934
.
Xu
X. X.
,
Yi
T.
,
Tang
B.
,
Lambeth
J. D.
(
1998
).
Disabled-2 (Dab2) is an SH3 domain-binding partner of Grb2.
Oncogene
16
,
1561
1569
.
Yamakami
M.
,
Yokosawa
H.
(
2004
).
Tom1 (target of Myb 1) is a novel negative regulator of interleukin-1- and tumor necrosis factor-induced signaling pathways.
Biol. Pharm. Bull.
27
,
564
566
.
Yamakami
M.
,
Yoshimori
T.
,
Yokosawa
H.
(
2003
).
Tom1, a VHS domain-containing protein, interacts with tollip, ubiquitin, and clathrin.
J. Biol. Chem.
278
,
52865
52872
.
Yamamoto
H.
,
Kakuta
S.
,
Watanabe
T. M.
,
Kitamura
A.
,
Sekito
T.
,
Kondo-Kakuta
C.
,
Ichikawa
R.
,
Kinjo
M.
,
Ohsumi
Y.
(
2012
).
Atg9 vesicles are an important membrane source during early steps of autophagosome formation.
J. Cell Biol.
198
,
219
233
.
Yanagida-Ishizaki
Y.
,
Takei
T.
,
Ishizaki
R.
,
Imakagura
H.
,
Takahashi
S.
,
Shin
H. W.
,
Katoh
Y.
,
Nakayama
K.
(
2008
).
Recruitment of Tom1L1/Srcasm to endosomes and the midbody by Tsg101.
Cell Struct. Funct.
33
,
91
100
.
Yang
Z.
,
Klionsky
D. J.
(
2010
).
Mammalian autophagy: core molecular machinery and signaling regulation.
Curr. Opin. Cell Biol.
22
,
124
131
.
Yi
Z.
,
Petralia
R. S.
,
Fu
Z.
,
Swanwick
C. C.
,
Wang
Y. X.
,
Prybylowski
K.
,
Sans
N.
,
Vicini
S.
,
Wenthold
R. J.
(
2007
).
The role of the PDZ protein GIPC in regulating NMDA receptor trafficking.
J. Neurosci.
27
,
11663
11675
.
Yoshida
H.
,
Cheng
W.
,
Hung
J.
,
Montell
D.
,
Geisbrecht
E.
,
Rosen
D.
,
Liu
J.
,
Naora
H.
(
2004
).
Lessons from border cell migration in the Drosophila ovary: A role for myosin VI in dissemination of human ovarian cancer.
Proc. Natl. Acad. Sci. USA
101
,
8144
8149
.
Young
A. R.
,
Chan
E. Y.
,
Hu
X. W.
,
Köchl
R.
,
Crawshaw
S. G.
,
High
S.
,
Hailey
D. W.
,
Lippincott-Schwartz
J.
,
Tooze
S. A.
(
2006
).
Starvation and ULK1-dependent cycling of mammalian Atg9 between the TGN and endosomes.
J. Cell Sci.
119
,
3888
3900
.
Yu
C.
,
Feng
W.
,
Wei
Z.
,
Miyanoiri
Y.
,
Wen
W.
,
Zhao
Y.
,
Zhang
M.
(
2009
).
Myosin VI undergoes cargo-mediated dimerization.
Cell
138
,
537
548
.
Zoncu
R.
,
Perera
R. M.
,
Balkin
D. M.
,
Pirruccello
M.
,
Toomre
D.
,
De Camilli
P.
(
2009
).
A phosphoinositide switch controls the maturation and signaling properties of APPL endosomes.
Cell
136
,
1110
1121
.