Lysosomes are key regulators of many fundamental cellular processes such as metabolism, autophagy, immune response, cell signalling and plasma membrane repair. These highly dynamic organelles are composed of various membrane and soluble proteins, which are essential for their proper functioning. The soluble proteins include numerous proteases, glycosidases and other hydrolases, along with activators, required for catabolism. The correct sorting of soluble lysosomal proteins is crucial to ensure the proper functioning of lysosomes and is achieved through the coordinated effort of many sorting receptors, resident ER and Golgi proteins, and several cytosolic components. Mutations in a number of proteins involved in sorting soluble proteins to lysosomes result in human disease. These can range from rare diseases such as lysosome storage disorders, to more prevalent ones, such as Alzheimer’s disease, Parkinson’s disease and others, including rare neurodegenerative diseases that affect children. In this review, we discuss the mechanisms that regulate the sorting of soluble proteins to lysosomes and highlight the effects of mutations in this pathway that cause human disease. More precisely, we will review the route taken by soluble lysosomal proteins from their translation into the ER, their maturation along the Golgi apparatus, and sorting at the trans-Golgi network. We will also highlight the effects of mutations in this pathway that cause human disease.

First discovered by Christian de Duve, lysosomes perform several key cellular functions, and their dysfunction has been implicated in several human pathologies (reviewed in [1,2]). One of the central functions of this organelle is to degrade cellular materials, originating from autophagy, phagocytosis or endocytosis. Substrates to be degraded include proteins, carbohydrates, nucleic acids as well as lipids, and is achieved by more than 70 soluble enzymes, along with enzyme activators and protective factors present in the lumen. Lysosomes are highly dynamic organelles in constant flux which fuse and reform following processes such as autophagosome/lysosome fusion [3]. The lysosomal membrane contains several types of membrane proteins that have a wide variety of functions including the protection of the membrane, the control of pH and ion homeostasis through membrane permeable channels and pumps. Furthermore, these membrane proteins also participate in the formation of membrane contact sites or the fusion of lysosomes with other organelles (reviewed in [4]). In this review, we provide an up-to-date view of the mechanisms by which soluble lysosomal proteins are sorted through various cellular compartments, starting at the endoplasmic reticulum (ER) all the way to their final destination in lysosomes. We will also emphasize the interdependence of anterograde and retrograde trafficking and their consequences on lysosomal activity. In the process, mutations in certain pathways will be discussed, when these changes result in human pathology.

Most soluble lysosomal proteins begin their journey as peptides that contain a 20–25 amino acid signal sequence that enables their translocation from the cytosol into the lumen of the ER during translation [5]. Some exceptions to this conventional route have been identified. For example, ceroid lipofuscinosis neuronal 5 (CLN5) and phospholipase D3 are initially translated as type II integral membrane proteins and are subsequently cleaved to produce mature soluble lysosomal proteins [6,7]. In the case of CLN5, it is cleaved by members of the signal peptide peptidase-like proteases (SPPL) family of intramembrane proteases, which are localized at the ER/Golgi interface [6]. On the other hand, cleavage of phospholipase D3 occurs in an acidic compartment, with a yet identified mechanism [7].

Newly synthesized proteins containing the 20–25 amino acid signal peptide undergo their first modifications: a signal peptide peptidase cleaves the signal peptide, while N-glycosylations are added to asparagine residues of Asn-X-Ser/Thr motifs by the oligosaccharyl transferase (OST) [8]. These soluble proteins move through the ER, are packaged into COPII vesicles and move on to the Golgi apparatus [9]. The efficiency and selectivity of protein sorting depends on specific interactions and cooperation between soluble cargos, cargo receptors and coat proteins. Transport out of the ER is mediated by the COPII coatomer complex [10]. This coat complex is recruited to the ER by the small GTPase Sar1 [11]. When activated and membrane bound, Sar1 acts as a docking factor for the Sec23 and Sec24 components of COPII to form a pre-budding complex at ER exit sites (ERES) (Figure 2). Sec24 is responsible for cargo binding and concentrating proteins into forming vesicles [12]. Next, the Sec13 and Sec31 subunits are recruited and function to deform the membrane to produce nascent vesicles, which are subsequently cleaved from the ER membrane [12] (Figure 2). Sec24 recognizes various sorting signals such as the phenylalanine-phenyalanine (Phe-Phe) motif present in the cytosolic C-terminal tail of the cargo receptor LMAN1/ERGIC53, which is responsible for the sorting of glycosylated proteins such as the lysosomal enzyme cathepsin C, and the well characterized ΦXΦXΦ (Φ is a hydrophobic amino acid, and X is any amino acid) ER export motif [13,14]. Although some ER cargo receptors have been identified, the mechanism regulating the sorting of soluble lysosomal proteins at the ER–Golgi interface remains only partially understood.

The mechanism as to how soluble lysosomal proteins move through the ER and are exported to the Golgi apparatus was recently demonstrated. Ceroid lipofuscinosis neuronal 6 (CLN6) and CLN8 are ubiquitously expressed and localized at the ER [15–19]. Mutations in either of the genes result in a neurodegenerative disease known as neuronal ceroid lipofuscinosis (NCL) [20,21]. Beyond CLN6 and CLN8, mutations in a total of 13 different genes (CLN1–CLN8 and CLN10–CLN14), are the cause of NCL. Disease onset can occur at any stage in life, but most often affects children. Common symptoms include cognitive regression, seizures, visual failure, and ataxia. At a cellular level, defective lysosomal function is a hallmark of these diseases, resulting in the accumulation of ceroid lipofuscin, among other molecules [22].

Using an elegant screening method based on interactions in live cell using bimolecular fluorescence complementation (BiFC), it was shown that CLN8 can interact with several soluble lysosomal proteins including other NCL proteins: PPT1 (CLN1), TPP1 (CLN2), cathepsin D (CLN10) and cathepsin F (CLN13) [23]. To test for a functional relationship, the localization of cathepsin D was compared in cortical and cerebellar sections from CLN8-deficient mice to wild-type mice, while the localization of TPP1 was compared in patient-derived CLN8-deficient fibroblasts with cells obtained from healthy donor. A significant decrease in the amount of cathepsin D was found in the LAMP1-positive compartment (endolysosomal compartment) of CLN8-deficient cells compared to wild-type cells, while the same was true for TPP1 in patient-derived cells compared to healthy cells [23]. Finally, a proteomic analysis of purified lysosomes found a significant decrease in a number of soluble lysosomal proteins, but little change in lysosomal integral membrane proteins in CLN8-deficient cells, supporting a role for this protein in trafficking soluble lysosomal proteins Table 1.

Table 1
Diseases associated with genes participating in the trafficking of soluble lysosomal proteins
Protein nameFunctionCellular localizationDiseaseReferences
AP-1
Heterotetrameric Adaptor Protein Complex 
Interacts with lysosomal sorting receptors and recruits clathrin coat to TGN membrane TGN MEDNIK syndrome
MEDNIK-like Syndrome
X-linked Intellectual Disability
Pettigrew Syndrome
Pustular Psoriasis 
[104–112
CLN3
Ceroid Lipofuscinosis Neuronal 3 
Rab7A/retromer interaction
Rab7A/sortilin interaction 
Endolysosomes Neuronal Ceroid Lipofuscinosis [147
CLN5
Ceroid Lipofuscinosis Neuronal 5 
Binds to CLN3 and and modulates its activity Endolysosomes Neuronal Ceroid Lipofuscinosis [148
CLN6
Ceroid Lipofuscinosis Neuronal 6 
Lysosomal soluble protein trafficking from ER- to-Golgi function together with CLN8 ER Neuronal Ceroid Lipofuscinosis [22,23
CLN8
Ceroid Lipofuscinosis Neuronal 8 
Lysosomal soluble protein trafficking from ER-to-Golgi, functions together with CLN6 ER Neuronal Ceroid Lipofuscinosis [22,23
LIMP2 Receptor protein for
β-Glucocerebrosidase 
Endolysosomes Parkinson's disease
Gaucher disease
Progressive Myoclonic epilepsy 
[81
N-acetylglucosiamne-1- phosphoptransferase Catalyzes Man-6-P modification on lysosomal enzymes as their recognition marker for their proper lysosomal transport Golgi Mucolipidosis II [44,45
Rab7A Small GTPase: acts as a molecular switch for early-late endosome maturation, recruitment and stabilization of retromer complex Endolysosomes Charcot–Marie–Tooth Disease [156
Retromer
Trimeric protein complex
VPS26/29/35 
Protein complex that interacts with cargo receptors and recycles them back to TGN Endolysosomes Parkinson’s disease
Alzheimer’s disease
Amyotrophic lateral sclerosis (ALS) Neuronal Ceroid Lipofuscinosis 
[147–155
Protein nameFunctionCellular localizationDiseaseReferences
AP-1
Heterotetrameric Adaptor Protein Complex 
Interacts with lysosomal sorting receptors and recruits clathrin coat to TGN membrane TGN MEDNIK syndrome
MEDNIK-like Syndrome
X-linked Intellectual Disability
Pettigrew Syndrome
Pustular Psoriasis 
[104–112
CLN3
Ceroid Lipofuscinosis Neuronal 3 
Rab7A/retromer interaction
Rab7A/sortilin interaction 
Endolysosomes Neuronal Ceroid Lipofuscinosis [147
CLN5
Ceroid Lipofuscinosis Neuronal 5 
Binds to CLN3 and and modulates its activity Endolysosomes Neuronal Ceroid Lipofuscinosis [148
CLN6
Ceroid Lipofuscinosis Neuronal 6 
Lysosomal soluble protein trafficking from ER- to-Golgi function together with CLN8 ER Neuronal Ceroid Lipofuscinosis [22,23
CLN8
Ceroid Lipofuscinosis Neuronal 8 
Lysosomal soluble protein trafficking from ER-to-Golgi, functions together with CLN6 ER Neuronal Ceroid Lipofuscinosis [22,23
LIMP2 Receptor protein for
β-Glucocerebrosidase 
Endolysosomes Parkinson's disease
Gaucher disease
Progressive Myoclonic epilepsy 
[81
N-acetylglucosiamne-1- phosphoptransferase Catalyzes Man-6-P modification on lysosomal enzymes as their recognition marker for their proper lysosomal transport Golgi Mucolipidosis II [44,45
Rab7A Small GTPase: acts as a molecular switch for early-late endosome maturation, recruitment and stabilization of retromer complex Endolysosomes Charcot–Marie–Tooth Disease [156
Retromer
Trimeric protein complex
VPS26/29/35 
Protein complex that interacts with cargo receptors and recycles them back to TGN Endolysosomes Parkinson’s disease
Alzheimer’s disease
Amyotrophic lateral sclerosis (ALS) Neuronal Ceroid Lipofuscinosis 
[147–155

CLN6 is the only other NCL related protein located in the ER. CLN6 interacts with a number of soluble lysosomal proteins, and its depletion results in the significant decrease of soluble lysosomal proteins localizing to lysosomes [24]. This decrease in localization also corresponds to diminished lysosomal protein activity. Soluble enzyme depletion in the lysosomal compartment was also observed in CLN6-deficient mice, suggesting that both CLN6 and CLN8 are involved in lysosomal soluble proteins trafficking [24]. Their role was recently identified as a two-step process.

First, CLN8 forms a homodimer that interacts with CLN6 on the membrane of the ERES to create a complex named EGRESS (ER-to-Golgi relaying of enzymes of the lysosomal system). This complex is responsible for the recruitment of soluble lysosomal proteins to promote their transport to the Golgi apparatus via the large luminal loop of CLN6 and the second luminal loop of CLN8 (Figure 2). Then, CLN8 interacts with the Sec24 subunit of COPII, which packages the receptor along with cargo into budding vesicles [23]. Sec24 recognizes the specific export signal VDWNF (ΦXΦXΦ) present at the cytosolic tail of CLN8 [23]. COPII coated vesicles, loaded with cargo are exported from the ER and transported to the cis-cisternae of the Golgi apparatus, while CLN6 remains in the ER [24] (Figures 1 and 2).

Schematic representation of sorting pathways taken by lysosomal cargo proteins

Figure 1
Schematic representation of sorting pathways taken by lysosomal cargo proteins
Figure 1
Schematic representation of sorting pathways taken by lysosomal cargo proteins
Close modal

CLN8 sorting at the ER

Figure 2
CLN8 sorting at the ER

Soluble cargo (red circles) are recognized by the luminal loop of CLN8 (orange) in the ER. Activated Sar1 is recruited to the membrane, recruits COPII, and packages CLN8 bound to cargo into trafficking vesicles.

Figure 2
CLN8 sorting at the ER

Soluble cargo (red circles) are recognized by the luminal loop of CLN8 (orange) in the ER. Activated Sar1 is recruited to the membrane, recruits COPII, and packages CLN8 bound to cargo into trafficking vesicles.

Close modal

The Golgi apparatus is the cellular centre for the modification, sorting and trafficking of cargo proteins and lipids to their final destinations. Newly synthesized proteins and lipids enter the cis-cisternae of the Golgi apparatus from the ER. While they progress through the Golgi, escaped ER resident proteins, as well as empty receptors are retrieved back to the ER, enabling the latter to take part in other cycles of trafficking. Once cargo proteins, such as soluble lysosomal proteins, and lipids reach the trans-cisternae and the trans-Golgi Network (TGN), which is the tubular structure of the most trans site of the Golgi complex, they leave for their final destinations [25–28]. How proteins are transported through the Golgi cisternae is still an active area of research. Several models have been proposed including (1) vesicular transport, (2) diffusion, (3) kiss and run, and (4) cisternae maturation.

(1) The vesicular transport model, as the name suggests, includes the formation of vesicles, in which cargos are packed and transported from one cisternae to another for further modifications by the enzymes within that cisternae and eventually exit from the Golgi. (2) In the diffusion model, small soluble cargo, such as albumin, transverse through the Golgi via interconnected Golgi stacks by simple diffusion [29]. However, this model excludes the transport of large cargo and probably underlies the co-existence of more than one model for intra-Golgi transport for different cargos. (3) In the symmetric kiss and run model, two Golgi compartments can fuse together through a transient narrow neck, which then undergoes fission to separate into two compartments. In the asymmetric kiss and run model, cargo molecules accumulate in a domain within the cisternae. The domain goes through fusion with the ‘accepting’ cisterna and subsequently undergoes fission with its originating cisterna [30]. The kiss and run model provides an advantage for cargo to move bidirectionally through the fusion neck without a separate mechanism. (4) Lastly, the Cisternal maturation model predicts that each cisternae moves and matures to distal cisternae. The key difference is that while the cargo stays in the cisternae, it is the Golgi resident proteins that recycle back to the previous cisternae. Each model poses its own advantages and disadvantages which are reviewed in detail here [31].

At the cis-Golgi, empty receptors, mis-sorted or mis-folded proteins, along with escaped ER resident proteins are retrieved back to the ER. This retrograde transport is enabled by a network of vesicles trafficking from the cis-Golgi to the ER, a process mediated by the COPI complex (Figure 3). COPI is located in the cytosol and composed of seven subunits : α, β′, ε, β, δ, γ and ζ. α, β′ and ε subunits form a cage-like B-subcomplex whereas β, δ, γ and ζ subunits form an adaptor-like F-subcomplex. The formation of vesicles starts with the activation of the small GTPase, ADP-ribosylation Factor 1 (Arf1) by guanine nucleotide exchange factors (GEFs) such as GBF1 at the cis-Golgi [32]. Arf proteins undergo N-terminal myristoylation, which is an important step in anchoring Arf GTPases to membranes [33].

Retrieval of CLN8 from the Golgi

Figure 3
Retrieval of CLN8 from the Golgi

Once CLN8 (orange) delivers cargo (red circles) to the Golgi apparatus, Arf1 recruits COPI to retrieve CLN8 back to the ER

Figure 3
Retrieval of CLN8 from the Golgi

Once CLN8 (orange) delivers cargo (red circles) to the Golgi apparatus, Arf1 recruits COPI to retrieve CLN8 back to the ER

Close modal

Activated, GTP-loaded Arf1 localizes to membranes of the cis-Golgi, interacting with the γζ and βδ subunits, leading to the recruitment of COPI [32,34]. Studies have shown that other Arf GTPases could also be involved in COPI recruitment [35]. ArfGEFs and Arf GTPase-activating proteins (ArfGAPs) such as ArfGAP1, 2 and 3, which return Arf1 to an inactive GDP bound form, were also shown to play important roles in promoting COPI recruitment, stability and vesicle budding (reviewed by [36]). Integral membrane proteins to be retrieved are then packaged into the nascent vesicle by interacting with COPI via specific sorting signals, while soluble proteins located in the lumen must bind to transmembrane receptors. Several retrieval signals and their receptors have been characterized (reviewed in [37]).

One of the most studied cargo receptors for ER-to-Golgi trafficking is the KDEL receptor (KDELR). This seven transmembrane domain containing protein localizes mainly to the cis-cisternae of the Golgi and is predicted to behave like a G-protein coupled receptor (GPCR) [38]. KDELR recognizes the KDEL (Lys-Asp-Glu-Leu) motif found on ER resident proteins and retrieves them back to the ER in a pH-dependent manner [39,40]. It was recently proposed that three amino acids of the signal binding pocket of KDELR (Y158, E127 and H12) could play an essential role in KDELR activation and cargo locking. Using quantum mechanical modelling, it is predicted that in the more acidic environment of lumen of the Golgi, H12 undergoes protonation, which would strengthen the hydrogen bond between Y158 and E127, resulting in a stronger interaction of KDEL-tagged proteins with the KDELR. Cargo binding then induces a conformational change of the KDELR and its activation, required for the recruitment of COPI complex [40,41]. KDELR activation also initiates a signaling cascade inside the Golgi complex that results in the phosphorylation of Src and Src Family Kinases (SFK). This signalling cascade regulates Golgi-to-ER trafficking and Intra-Golgi transport [38], such that SFK inactivation by inhibitors resulted in the accumulation of vesicular stomatitis virus G protein (VSVG) in the Golgi and prevented it from reaching the TGN [42]. While KDELR induced SFK activation regulates intra-Golgi trafficking, a PKA signaling cascade plays an important role in regulating Golgi-to-ER retrograde transport, as well as promoting the expression of genes involved in vesicular trafficking through CREB1 activation [38,42,43]. A recent study suggests that activation by KDELR signaling inside the Golgi also results in lysosomal repositioning, bringing lysosomes close to the perinuclear space, and sustains Golgi secretion [44].

Enzymes present in the different cisternae of the Golgi apparatus catalyze important post-translational modifications including phosphorylation, glycosylation, proteolytic cleavage and sulfation [45]. Some modifications are required for the proper sorting of cargos to their final destination, as well as their function. Indeed, in 1972, Hickman & Neufeld showed that soluble lysosomal enzymes required certain modifications to be sorted to lysosomes using skin-derived fibroblasts from I-cell disease patients [46]. This disease, also known as mucolipidosis II, is a rare lysosomal storage disorder caused by mutations in N-acetylglucosamine-1-phosphotransferase [45,46]. N-acetylglucosaminyl-1-phosphotransferase (GlcNac-1-phosphotransferase) and N-acetylglucosamine-1-phosphodiester α-N-acetylglucosaminidase act sequentially to synthesize mannose 6-phosphate (M6P) on many lysosomal soluble N-glycosylated enzymes. The latter enzyme is also known as the uncovering enzyme that exposes the M6P. N-acetylglucosamine-1-phosphodiester α-N-acetylglucosaminidase mainly localizes to the TGN, while a small amount can be found at the plasma membrane, suggesting the enzymes cycles between the TGN and the plasma membrane [45,46]. As such, tagging proteins with M6P separates proteins from a general secretion pathway, and targets them for sorting towards the lysosome.

The trans-Golgi Network (TGN), is the trans-most distal cisternae site of the Golgi apparatus and has a tubular network shape. TGN structure and size vary from one cell type to another, and incoming and outgoing trafficking to the TGN dynamically regulates its size and morphology [47–50]. In the classical point of view, the TGN is depicted as the ultimate sorting hub for cargo proteins and lipids, where they are sorted to their final destinations [47,51]. This view has been challenged with results from polarized Madin–Darby canine kidney (MDCK) cells implicated in biosynthetic cargo trafficking. Studies showed that, after cargos exit from the TGN, they visit endosomal compartments inside the cell before their final destinations, for example, common recycling endosomes, apical and basal sorting endosomes [51,52].

Whether sorting happens ultimately in the TGN or in post-TGN endosomal compartments, sorting of cargo proteins requires an elaborate collaboration between receptor, adaptor and coat proteins, which will be discussed below.

The mannose 6-phosphate receptors

Newly tagged proteins with M6P are recognized by two different lysosomal sorting receptors, namely: the 46 kDa cation-dependent (CD), and the 300 kDa cation-independent (CI) Mannose 6-phosphate receptors (CD-MPR and CI-MPR, respectively). These receptors are type I transmembrane domain proteins that co-operate in the delivery of lysosomal proteins but also have non-overlapping functions in the targeting of a subset of lysosomal enzymes [53,54] (Figure 1). The MPRs are found in vertebrates including zebrafish (Danio rerio), but it is not well defined in invertebrates or other organisms [55]. Beyond transport of lysosomal proteins, CI-MPR, also known as insulin-like growth factor II receptor (IGFRII), also plays a significant role in the clearance of IGFII hormone through endocytosis [56,57].

The binding of M6P tagged proteins to the receptors is a pH-dependent process: the relatively neutral pH of the TGN (pH 6,5) results in conformational changes of the receptors, enabling their interaction with M6P-tagged cargos. Once delivered in the more acidic environment of the lumen of the endosomal compartment (pH≤5), cargos are dissociated from the receptor [58,59]

In addition to M6P, a study combining NMR spectroscopy and molecular modelling showed that CI-MPR can also bind Man-P-GlcNAc residues via domain 5 of its 5th mannose 6-phosphate receptor homology (MRH) domain. Man-P-GlcNAc containing proteins are cargos that eluded the GlcNAc hydrolysis catalyzed by the uncovering enzyme, exposing the M6P residue [60]. This suggests that cargos which didn’t undergo complete maturation in the Golgi apparatus could still be sorted at the TGN. This could also explain why mice lacking the uncovering enzyme still showed some enzyme sorting to the lysosome [61].

Sortilin

It has been observed, that despite defects in their MPR trafficking pathway, some cell lines still showed a significant level of lysosomal activity, suggesting that a MPR-independent sorting of lysosomal enzymes exists [62,63]. Sortilin, a type I transmembrane protein, has been identified due to its high homology to yeast Vps10p [64]. Vps10p functions in the sorting of soluble lysosomal proteins between the TGN and endosomes [65]. Sortilin is a cargo receptor that cycles between the TGN and endosomes, and sorts various proteins such as GM2AP (GM2 activator protein), acid sphingomyelinase and prosaposin, as well as cathepsins D and H to lysosomes [66–68] (Figure 1). Sortilin also localizes to the plasma membrane and regulates the endocytosis of progranulin, lipoprotein lipase, apolipoprotein A-V and EGFR [69–72]. Beyond its role in intracellular trafficking, sortilin was shown to be associated with extracellular vesicles regulating vascular calcification [73].

The cytosolic tail of sortilin is the site of various forms of post-translational modifications (PTMs) regulating its function. Phosphorylation can occur at serine 793 via the p21-activated kinase family [74]. This modification appears to modulate the interaction between sortilin and Adaptor Protein 1 (AP-1), as constitutively active phosphomimetic constructs of sortilin were unable to bind AP-1 in a yeast 2-hybrid system [74]. A second phosphorylation site in the cytosolic tail of sortilin has also been identified. Serine 825 is phosphorylated, and phosphorylation at this site modulates GGA1 binding [75]. Beyond phosphorylation, palmitoylation and ubiquitination are also PTMs found in the cytosolic tail of sortilin. Palmitoylation is required for the efficient endosome-to-TGN retrieval of sortilin, while ubiquitination labels the protein for lysosomal degradation [76,77]. Interestingly, these two PTMs interplay with one another as non-palmitoylated sortilin is ubiquitinated, and degraded in lysosomes [77].

LIMP-2

LIMP-2 is a type III lysosomal integral protein, which binds β-glucocerebrosidase (β-GC) and traffics it to lysosomes in a pH-dependant manner [78]. Interestingly, the interaction of LIMP-2 with β-GC is suggested to occur in the ER, unlike cargo binding to sortilin and MPRs, which take place in the TGN [78]. Sorting of LIMP-2 to the lysosomal compartment requires its interaction with AP-1 via its cytosolic tail through both dileucine and tyrosine signals [79,80]. More recent studies have shown that LIMP-2 could be more than a lysosomal protein sorting receptor. In fact, LIMP-2 is also involved in the transport of cholesterol and phospholipids such as phosphatidylserine [81,82].

Additionally, the overexpression of LIMP-2 in COS cells results in enlarged early and late endosomes, which is suppressed by dominant-negative expression of Rab5b; a small GTPase that functions as a molecular switch for early to late endosome progression [83]. Lastly, LIMP-2 has been implicated in other neurological diseases such as Parkinson’s disease, Gaucher disease and progressive myoclonic epilepsy (reviewed extensively here [84]), demonstrating its importantance, but partially understood role in lysosomal function.

Sorting at the TGN: the role of the Clathrin adaptors

Transport of soluble lysosomal proteins from the TGN to the endolysosomal compartment is mediated by clathrin coated vesicles. Their formation, and the specific selection of cargos to sort, is regulated by the cooperation of cargos, cargo receptors, clathrin adaptor proteins, accessory proteins and clathrin, at the interface of the cytosol and the lumen of the trans-Golgi Network. There are two major types of clathrin adaptors: The multimeric adaptor proteins, and the Golgi-localized, γ-ear containing, ADP-ribosylation factor-binding (GGA) proteins.

Multimeric adaptor protein

Adaptor protein-1 (AP-1) is a heterotetrameric protein complex composed of two large subunits (β1 and γ): a medium subunit (µ1) and a small subunit (σ1). The complex is arranged into two sections: a core composed of the N-terminal portions of β1 and γ, and the entire µ1 and σ1 subunits, while the C-terminal portions of β1 and γ extend out away from the core [9]. AP-1 can interact with the cytosolic tail of CD-MPR, CI-MPR and sortilin through specific motifs (Figure 4). Two such motifs are contained in the tail of each receptor, the tyrosine motif (YXXØ, where X is any amino acids and Ø is a bulky hydrophobic amino acid) and the dileucine motif ([D/E]XXXL[L/I]) [85,86]. Binding of the tyrosine motifs to AP-1 is mediated by the µ1 subunit [85,87], while dileucine motifs binding requires σ1 and the core portion of γ [86,88].

Sorting of the lysosomal sorting receptors at the TGN - AP-1

Figure 4
Sorting of the lysosomal sorting receptors at the TGN - AP-1

AP-1 is recruited to the membrane of the TGN by the small GTPase Arf1, which also open the conformation of AP-1.

This enables the binding of AP-1 to the sorting receptor (green), which is loaded with soluble lysosomal cargo (red circles).

Figure 4
Sorting of the lysosomal sorting receptors at the TGN - AP-1

AP-1 is recruited to the membrane of the TGN by the small GTPase Arf1, which also open the conformation of AP-1.

This enables the binding of AP-1 to the sorting receptor (green), which is loaded with soluble lysosomal cargo (red circles).

Close modal

Beyond these classical sorting signals, atypical sorting motifs have also been identified for proper lysosomal trafficking. For example CLN3, a multi-spanning membrane protein requires the extended dileucine motif [EEEX(8)LI] in its second (large) cytoplasmic loop domain and the [MX(9)G] motif, where methionine and glycine, separated by nine amino acids in its C-terminal tail, for its proper lysosomal localization [89]. AP-1 has been shown to bind the atypical sorting motifs in CLN3 for its trafficking; however, there are contradictory results whether AP complexes binds these motifs or not [90,91].

The interaction between the sorting receptor and AP-1 occurs on the membranes of the TGN. The spatiotemporal recruitment of AP-1 to membranes requires phosphatidylinositol 4-phosphate (PI4P) and the small GTPase Arf1 [92–94]. Small GTPases such as Arf1 cycle between an inactive cytosolic form which is bound to GDP, and a membrane bound active form [95]. Guanine nucleotide exchange factors (GEFs) activate small GTPases, while GTPase activating proteins (GAPs) shut off the signal, returning small GTPases to the cytosol [96]. Activation of Arf1 for AP-1 recruitment is modulated by BIG2 [97,98], while the GAP implicated in this pathway has not been fully determined. Beyond membrane recruitment, Arf1 activates AP-1 by changing its conformation leading to an increased binding to the cytosolic tail of the sorting receptors [99,100].

Beyond binding to proteins sorted toward the lysosomal compartment, AP-1 has also been implicated in endosome-to-TGN sorting of a wide range of cargos including CI-MPR and CD-MPR [101,102], and TGN-to-PM sorting. AP-1 also plays an important role in basolateral sorting in polarized cells [103] and is essential for embryonic development in mice. Knockout of the γ1 or µ1a subunits of AP-1 in mice results in embryonic lethality [104,105]. Mutations in AP-1 subunits cause a range of human disorders, which are classified as coatopathies [106]. For example, X-linked intellectual disability and Pettigrew syndrome are associated with mutations in the σ1B gene [107,108]. On the other hand, mutations in σ1A gene cause MEDNIK syndrome (mental disability, enteropathy, deafness, neuropathy, ichthyosis and keratoderma), and loss-of-function mutation in the β1 gene results in MEDNIK-like syndrome [109–111]. Mutations in the σ1C subunit of the AP-1 complex lead to pustular psoriasis, an autoinflammatory skin condition [112,113].

MEDNIK syndrome is characterized by several symptoms including profound intellectual impairment, deafness, cerebral motor disorder, severe intestinal impairments and scaly/thickened skin. Mutations in the AP1S1 gene encoding the σ1A subunit of the AP-1 protein complex have been identified as the cause of MEDNIK syndrome [114,115]. The most common mutation results in the introduction of a premature stop codon at the beginning of exon 4, translating to a truncated protein of 19 amino acids [114].

The Golgi-localized, γ-ear containing, ADP-ribosylation factor-binding proteins

The GGA (GGA1-3) proteins are monomeric adaptors of 65–80 kDa, which were originally discovered as Arf interacting proteins [116–118]. The GGAs contain four domains; VHS (Vps27p/Hrs/STAM), GAT (GGA and TOM1), hinge and GAE (gamma-adaptin homology domain). As the name implies, the C-terminus of the GGA proteins show high sequence homology with the γ subunit of AP-1 complex [119]. In mammalian cells, the GGA proteins localize mainly to the TGN, and similarly to AP-1, bind to sortilin and the mannose 6-phosphate receptors through the interaction of their dileucine signals with the VHS domain [120–122] (Figure 5). The GAT domain is required for the TGN localization of the GGAs, and is also required for their interaction with Arf1 [116,123]. The hinge domain contains clathrin binding sites, and has been shown to interact with this protein [123,124], while the GAE domain binds several accessory proteins [125,126]. The hinge domain of GGA and the γ ear domain of AP-1 were shown to interact, which suggests that GGA can function along with AP-1 to sequester sorting receptor into AP-1 positive vesicles [127]. In HeLa cells, where all three GGAs were deleted using CRISPR/Cas9 (GGAKO), CI-MPR localized mainly to the Golgi, in contrast with wild-type cells where CI-MPR showed both Golgi and peripheral puncta staining. Moreover, in the triple GGAKO cells, cathepsin D localization to endolysosomes was impaired, however not completely abolished, suggesting that AP-1 could maintain its trafficking [128].

Sorting of the lysosomal sortig receptors at the TGN - GGAs

Figure 5
Sorting of the lysosomal sortig receptors at the TGN - GGAs

GGA proteins are recruited to the TGN membranes by Arf1, which enables their interaction with cargo (red dots) bound sorting receptor (green).

Figure 5
Sorting of the lysosomal sortig receptors at the TGN - GGAs

GGA proteins are recruited to the TGN membranes by Arf1, which enables their interaction with cargo (red dots) bound sorting receptor (green).

Close modal

GTP-loaded Arf1 is required to recruit GGAs to the TGN [116,123]. However, the mechanism regulating the Arf1 cycle for GGA recruitment is not well understood. The GEF GBF1 has been implicated in the recruitment of GGAs, as its depletion using RNAi resulted in decreased GGA3 recruitment, whereas expressing catalytically dead GBF1 led to the mis-sorting of the soluble lysosomal protein prosaposin [129]. Depletion of the GAP ArfGAP3 caused a decreased membrane recruitment of the GGAs, which resulted in the mis-sorting of the soluble lysosomal enzyme cathepsin D. However, CI-MPR was relocalized to endosomes in ArfGAP3 deleted cells, suggesting a blockage in endosome-to-TGN trafficking [130].

It is not fully understood how GGAs function in relation to AP-1 at the TGN. Some work has suggested that GGAs hand off cargo to AP-1, in order to sort cargo at the TGN. More work will be required to fully understand the role of these proteins in lysosomal sorting.

Cargo receptor retrieval to the TGN

Once the receptor/cargo complex that was packaged into clathrin-coated vesicles and fuses with the late endosome, the more acidic pH in this organelle causes a conformational change in the receptor, which enables cargo to be released. Meanwhile, the receptor is sorted out of the endosome and retrieved to the TGN. The receptor escapes lysosomal degradation and is instead used for subsequent rounds of sorting. Originally identified in yeast, retromer plays a key role in retrograde trafficking and receptor recycling. Retromer is a cytosolic protein complex that is recruited to endosomal membranes and can interact with the cytosolic tail of the receptors (Figure 6). This trimeric complex is composed of vacuolar protein sorting (Vps)-26, Vps29 and Vps35 [131,132]. In retromer-depleted cells, the retrieval of the receptors is less efficient, resulting in their lysosomal degradation (Figure 1). This has a significant negative impact on the sorting of soluble lysosomal cargos, resulting in lysosomal dysfunction [131,132].

Retrieval of the lysosomal sorting receptors from the endosome

Figure 6
Retrieval of the lysosomal sorting receptors from the endosome

In order to retrieve the lysosomal sorting receptors (green) from the endosome, Rab7A is recruited to membranes and enables the binding of retromer to the receptor CLN3 (blue) and CLN5 coordinate this process.

Figure 6
Retrieval of the lysosomal sorting receptors from the endosome

In order to retrieve the lysosomal sorting receptors (green) from the endosome, Rab7A is recruited to membranes and enables the binding of retromer to the receptor CLN3 (blue) and CLN5 coordinate this process.

Close modal

The spatiotemporal recruitment of retromer is regulated by a small GTPase, Rab7A (Figure 6). In its GTP active form, Rab7A is recruited to membranes and stabilizes the recruitment of retromer [133,134]. In Rab7A-depleted cells, retromer recruitment is significantly reduced, but the receptors are not degraded, as Rab7A is also required to enable lysosomal degradation. Instead, the receptors accumulate at the late endosome [133].

As a small GTPase, the recruitment of Rab7A to membranes is regulated by its GTP loading status [135]. In an inactive form, Rab7A is bound to GDP and is localized to the cytosol. The GEF MON1/CCZ1 loads Rab7A with GTP, which enables its membrane recruitment [136,137]. Two GAPs, TBC1D5 or TBC1D15, hydrolyze GTP back to GDP, resulting in Rab7A shutoff [134,138,139]. Beyond this mechanism, Rab7A can be regulated specifically in its function of retromer recruitment. First, it is known that several post-translational modifications on Rab7A can regulate its function. Rab7A has been shown to be phosphorylated [140–143], palmitoylated [144] and ubiquitinated [145,146]. Of these, palmitoylation is required for the interaction between Rab7A and retromer, although nonpalmitoylatable Rab7A is still membrane localized and can mediate the degradation of internalized cell surface proteins [144].

Beyond PTMs, other mechanisms have been shown to regulate the retrieval of the lysosomal sorting receptors. For instance CLN3, an integral membrane protein localized to endolysosomal membranes acts as a scaffold to favor the Rab7A/retromer and retromer/sortilin interactions as CLN3 can interact with all three of these proteins [147]. In CLN3-depleted HeLa cells, these interactions are significantly weakened, resulting in the receptor degradation in lysosomes and mis-processing of cathepsin D. Additionally, CLN5, which is a soluble lysosomal protein that binds CLN3, is also required for retromer recruitment, as it modulates CLN3 activity [148]. In CLN5-depleted HeLa cells, the Rab7A/retromer and retromer/sortilin interactions are also disrupted, as are the CLN3 interactions with Rab7A, retromer and sortilin. This also leads to receptor degradation in lysosomes and defective lysosomal function. In this case, CLN5 appears to regulate Rab7A PTMs, as its palmitoylation level is significantly reduced in CLN5-deleted HeLa cells [148]. Although the mechanism of Rab7A palmitoylation is not known, one could speculate that CLN5 modulates an interaction of the palmitoylation machinery with CLN3, promoting Rab7A palmitoylation.

Much like mutations in CLN6 or CLN8, mutations in either CLN3 or CLN5 in humans also cause NCL. No cure has yet been found for these diseases which dramatically reduce the lifespan of patients, with death often occurring before the third decade of life. Beyond the NCLs, retromer mutations have been identified in Parkinson's disease [149–152], while defective retromer has also been associated with Alzheimer’s disease [153,154] and amyotrophic lateral sclerosis (ALS) [155]. Finally, mutations in Rab7A have been shown to cause Charcot–Marie–Tooth disease (CMT), a peripheral nerve disease which induces loss of muscles and sensory defects [156].

Retromer is not the only complex regulating the retrieval of proteins from endolysosomes to the TGN. Indeed, it has been shown that a combination of sorting nexin (SNX) proteins can interact directly with CI-MPR for its recycling either to the TGN or to the plasma membrane [157]. This sorting complex has been termed Endosomal SNX–BAR sorting complex for promoting exit-1 (ESCPE-1) (Figure 1), and is composed of SNX1 (or SNX2) and SNX5 (or SNX6). While SNX5 and SNX6 bind a specific motif in the cytosolic tail of CI-MPR, SNX1 and SNX2 interact with PI3P at the endosomal membrane and initiate the formation of tubulo-vesicular transport carriers, via their C-terminal BAR (Bin/Amphiphysin/Rvs) domain [157].

Two distinct sorting mechanisms function at endosomes to retrieve receptors. Why the cell needs these two mechanisms is unclear. Perhaps the redundancy ensures proper retrieval. What is emerging, is that these two pathways package cargo receptors into distinct vesicles that reach the TGN and tether through different mechanisms [158,159].

Soluble lysosomal protein delivery to their final destination

As the soluble lysosomal proteins are progressively dissociated from their receptors and the receptors are retrieved back to the TGN, endosomes continue their routes towards lysosomes. Endosomal pH slowly decreases to reach lysosomal pH 4.5, resulting in final cleavage and activation of lysosomal hydrolases such as cathepsin D [160,161]. When in close proximity, late endosome and lysosome membranes tether and can undergo heterotypic fusion, resulting in material exchange between the two compartments. The mechanisms behind endosome/lysosome fusion is still a matter of debate, although two models have been proposed: the ‘kiss and run’ and the ‘fusion-fission’ models. The first is based on transient fusions and fissions events between the two compartments, enabling content mixing [162,163]. On the contrary, the ‘fusion-fission’ model suggests that late endosome and lysosome completely fuse to form a hybrid organelle called endolysosome. Later on, lysosomes would reform, packing lysosomal proteins with it and serving as an enzyme storage compartment, until the next fusion with another organelle [164,165]. Evidence also suggests that both mechanisms could occur concurrently in mammal cells [166].

Fusion events between late-endosomes and lysosomes are a three-step process: membrane tethering, followed by SNARE complex assembly, which results in the twio membranes fusing. It is mediated and regulated by several proteins. In mammalian cells, the homotypic fusion and vacuole protein sorting (HOPS) complex is one of the main actors involved. This tethering complex has a seahorse-like structure and is composed of six subunits : VPS11, VPS16, VPS18, VPS33, VPS39 and VPS41. HOPS bridges membranes from different compartments and keep them in close proximity. In yeast, the Rab7A homolog Ypt7 was showed to be involved in the recruitment of VPS39 and VPS41 to membranes [167,168]. In mammalian cells, the recruitments of those subunits to the endosomal membranes would not be directly mediated by Rab7A, but through its effectors RILP [169], and Pleckstrin homology domain-containing family M member 1 (PLEKHM1) [170]. However, Rab7A could play an indirect role in endosome fusion, through the formation of a complex with RILP and the cholesterol sensor ORP1L. Rab7A–RILP–ORP1L complex would regulate the coordination of endosome movement and fusion, and the switch from the transport machinery to the fusion machinery [169,171]. The activated small GTPase Arl8 is also responsible for the recruitment of VPS41 to the lysosomal membrane [172,173]. In addition, Arl8 and Rab7 were also shown to collaborate to recruit PLEKHM1 onto lysosome and late endosome membranes respectively [174]. PLEKHM1 is an essential adaptor protein which promotes the recruitment and stabilization of HOPS subunits, and the tethering of endosome and autophagosome membranes with lysosomes [170]. HOPS complex assembles and enables the recruitment of SNARE proteins [175]; for example, a study conducted in yeast found that VPS33 is the subunit where the interaction with SNARE proteins occurs [176,177].

In mammalian cells, trans-SNARE complexes consist in the association of one R-SNARE (VAMP7 or VAMP8) in association with three Q-SNAREs (Syntaxin-7, Syntaxin-8 and VTI1b). VAMP8 would be involved in homotypic fusion of endosomes whereas VAMP7 would be required for heterotypic fusion of endosomes with lysosomes [178]. SNARE association acts as a zipper that brings membranes close enough to each other to result in membrane curvature change and spontaneous proteolipid fusion pore formation [179]. Lumen content of both organelle then mix, leading to soluble enzyme release in endolysosomes.

Sorting of soluble lysosomal proteins is a fundamental cellular process required to maintain lysosomal activity. It involves the close cooperation of a wide range of proteins from various compartments, which act together to ensure the proper localization and maturation of soluble lysosomal proteins. The variety of protein adaptors, cargo receptors and recognized sorting motifs enable the sorting numerous proteins with high specificity. For a number of years, defects in this process have been known to cause lysosomal storage disorders [1], but it is only more recently that they have been associated to more common human diseases, especially age-related neurodegenerative diseases [153–155,180]. Those studies have especially highlighted the underestimated importance of the retrograde trafficking of receptors and misrouted proteins, in the process and the regulation of lysosomal protein sorting. Although some of those pathways have been investigated for more than 40 years, mechanisms regulating the sorting of lysosomal protein have not been fully elucidated. Revisiting models and identifying new effector functions, using novel methods such as protein-protein interactions studies in live cells, CRISPR/Cas9 knockouts generation, and unbiased ‘omics’ approaches, appear to be essential for a better understanding of the sorting of lysosomal proteins. Increasing our comprehension of these pathways could hence lead to the identification of novel therapeutic targets for a number of diseases.

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

Work in the S.L. lab is supported by grants from CIHR [grant numbers ENG-155186 and PJT-173419]; NSERC [grant number RGPIN-2020-06549]; CFI [grant number 35258] and ForeBatten Foundation.

Original draft: I.M., J.M., S.L. Revisions: I.M., J.M., S.L.

β-GC

β-glucocerebrosidase

CMT

Charcot–Marie–Tooth disease

GAE

gamma-adaptin homology domain

GAP

GTPase activating protein

GEF

guanine nucleotide exchange factor

HOPS

homotypic fusion and vacuole protein sorting

PLEKHM1

Pleckstrin homology domain-containing family M member 1

1.
Sharma
J.
,
di Ronza
A.
,
Lotfi
P.
and
Sardiello
M.
(
2018
)
Lysosomes and brain health
.
Annu. Rev. Neurosci.
41
,
255
276
[PubMed]
2.
Ballabio
A.
and
Bonifacino
J.S.
(
2020
)
Lysosomes as dynamic regulators of cell and organismal homeostasis
.
Nat. Rev. Mol. Cell Biol.
21
,
101
118
[PubMed]
3.
Saffi
G.T.
and
Botelho
R.J.
(
2019
)
Lysosome fission: planning for an exit
.
Trends Cell Biol.
29
,
635
646
[PubMed]
4.
Schwake
M.
,
Schroder
B.
and
Saftig
P.
(
2013
)
Lysosomal membrane proteins and their central role in physiology
.
Traffic
14
,
739
748
[PubMed]
5.
Walter
P.
,
Gilmore
R.
and
Blobel
G.
(
1984
)
Protein translocation across the endoplasmic reticulum
.
Cell
38
,
5
8
[PubMed]
6.
Jules
F.
,
Sauvageau
E.
,
Dumaresq-Doiron
K.
,
Mazzaferri
J.
,
Haug-Kroper
M.
,
Fluhrer
R.
et al.
(
2017
)
CLN5 is cleaved by members of the SPP/SPPL family to produce a mature soluble protein
.
Exp. Cell Res.
357
,
40
50
[PubMed]
7.
Gonzalez
A.C.
,
Schweizer
M.
,
Jagdmann
S.
,
Bernreuther
C.
,
Reinheckel
T.
,
Saftig
P.
et al.
(
2018
)
Unconventional trafficking of mammalian phospholipase D3 to lysosomes
.
Cell Rep.
22
,
1040
1053
[PubMed]
8.
Mohorko
E.
,
Glockshuber
R.
and
Aebi
M.
(
2011
)
Oligosaccharyltransferase: the central enzyme of N-linked protein glycosylation
.
J. Inherit. Metab. Dis.
34
,
869
878
[PubMed]
9.
Braulke
T.
and
Bonifacino
J.S.
(
2009
)
Sorting of lysosomal proteins
.
Biochim. Biophys. Acta
1793
,
605
614
[PubMed]
10.
Stephens
D.J.
(
2003
)
De novo formation, fusion and fission of mammalian COPII-coated endoplasmic reticulum exit sites
.
EMBO Rep.
4
,
210
217
[PubMed]
11.
Kuge
O.
,
Dascher
C.
,
Orci
L.
,
Rowe
T.
,
Amherdt
M.
,
Plutner
H.
et al.
(
1994
)
Sar1 promotes vesicle budding from the endoplasmic reticulum but not Golgi compartments
.
J. Cell Biol.
125
,
51
65
[PubMed]
12.
Stagg
S.M.
,
LaPointe
P.
,
Razvi
A.
,
Gurkan
C.
,
Potter
C.S.
,
Carragher
B.
et al.
(
2008
)
Structural basis for cargo regulation of COPII coat assembly
.
Cell
134
,
474
484
[PubMed]
13.
Vollenweider
F.
,
Kappeler
F.
,
Itin
C.
and
Hauri
H.P.
(
1998
)
Mistargeting of the lectin ERGIC-53 to the endoplasmic reticulum of HeLa cells impairs the secretion of a lysosomal enzyme
.
J. Cell Biol.
142
,
377
389
[PubMed]
14.
Otsu
W.
,
Kurooka
T.
,
Otsuka
Y.
,
Sato
K.
and
Inaba
M.
(
2013
)
A new class of endoplasmic reticulum export signal PhiXPhiXPhi for transmembrane proteins and its selective interaction with Sec24C
.
J. Biol. Chem.
288
,
18521
18532
[PubMed]
15.
Lonka
L.
,
Kyttala
A.
,
Ranta
S.
,
Jalanko
A.
and
Lehesjoki
A.E.
(
2000
)
The neuronal ceroid lipofuscinosis CLN8 membrane protein is a resident of the endoplasmic reticulum
.
Hum. Mol. Genet.
9
,
1691
1697
[PubMed]
16.
Heine
C.
,
Koch
B.
,
Storch
S.
,
Kohlschutter
A.
,
Palmer
D.N.
and
Braulke
T.
(
2004
)
Defective endoplasmic reticulum-resident membrane protein CLN6 affects lysosomal degradation of endocytosed arylsulfatase A
.
J. Biol. Chem.
279
,
22347
22352
[PubMed]
17.
Lonka
L.
,
Salonen
T.
,
Siintola
E.
,
Kopra
O.
,
Lehesjoki
A.E.
and
Jalanko
A.
(
2004
)
Localization of wild-type and mutant neuronal ceroid lipofuscinosis CLN8 proteins in non-neuronal and neuronal cells
.
J. Neurosci. Res.
76
,
862
871
[PubMed]
18.
Mole
S.E.
,
Michaux
G.
,
Codlin
S.
,
Wheeler
R.B.
,
Sharp
J.D.
and
Cutler
D.F.
(
2004
)
CLN6, which is associated with a lysosomal storage disease, is an endoplasmic reticulum protein
.
Exp. Cell Res.
298
,
399
406
[PubMed]
19.
Heine
C.
,
Quitsch
A.
,
Storch
S.
,
Martin
Y.
,
Lonka
L.
,
Lehesjoki
A.E.
et al.
(
2007
)
Topology and endoplasmic reticulum retention signals of the lysosomal storage disease-related membrane protein CLN6
.
Mol. Membr. Biol.
24
,
74
87
[PubMed]
20.
Weimer
J.M.
,
Kriscenski-Perry
E.
,
Elshatory
Y.
and
Pearce
D.A.
(
2002
)
The neuronal ceroid lipofuscinoses: mutations in different proteins result in similar disease
.
Neuromolecular Med.
1
,
111
124
[PubMed]
21.
Mole
S.E.
and
Cotman
S.L.
(
2015
)
Genetics of the neuronal ceroid lipofuscinoses (Batten disease)
.
Biochim. Biophys. Acta
1852
,
2237
2241
[PubMed]
22.
Anderson
G.W.
,
Goebel
H.H.
and
Simonati
A.
(
2013
)
Human pathology in NCL
.
Biochim. Biophys. Acta
1832
,
1807
26
23.
di Ronza
A.
,
Bajaj
L.
,
Sharma
J.
,
Sanagasetti
D.
,
Lotfi
P.
,
Adamski
C.J.
et al.
(
2018
)
CLN8 is an endoplasmic reticulum cargo receptor that regulates lysosome biogenesis
.
Nat. Cell Biol.
20
,
1370
1377
[PubMed]
24.
Bajaj
L.
,
Sharma
J.
,
di Ronza
A.
,
Zhang
P.
,
Eblimit
A.
,
Pal
R.
et al.
(
2020
)
A CLN6-CLN8 complex recruits lysosomal enzymes at the ER for Golgi transfer
.
J. Clin. Invest.
130
,
4118
4132
[PubMed]
25.
Ladinsky
M.S.
,
Mastronarde
D.N.
,
McIntosh
J.R.
,
Howell
K.E.
and
Staehelin
L.A.
(
1999
)
Golgi structure in three dimensions: functional insights from the normal rat kidney cell
.
J. Cell Biol.
144
,
1135
1149
[PubMed]
26.
Klumperman
J.
(
2011
)
Architecture of the mammalian Golgi
.
Cold Spring Harbor Perspect. Biol.
3
,
a005181
27.
Pantazopoulou
A.
and
Glick
B.S.
(
2019
)
A kinetic view of membrane traffic pathways can transcend the classical view of Golgi compartments
.
Front. Cell Development. Biol.
7
,
153
28.
David
Y.
,
Castro
I.G.
and
Schuldiner
M.
(
2021
)
The fast and the furious: Golgi contact sites
.
Contact
4
,
25152564211034424
29.
Beznoussenko
G.V.
,
Parashuraman
S.
,
Rizzo
R.
,
Polishchuk
R.
,
Martella
O.
,
Di Giandomenico
D.
et al.
(
2014
)
Transport of soluble proteins through the Golgi occurs by diffusion via continuities across cisternae
.
Elife
3
,
[PubMed]
30.
Mironov
A.A.
and
Beznoussenko
G.V.
(
2012
)
The kiss-and-run model of intra-Golgi transport
.
Int. J. Mol. Sci.
13
,
6800
6819
[PubMed]
31.
Mironov
A.A.
and
Beznoussenko
G.V.
(
2019
)
Models of intracellular transport: pros and cons
.
Front. Cell Development. Biol.
7
,
146
32.
Yu
X.
,
Breitman
M.
and
Goldberg
J.
(
2012
)
A structure-based mechanism for Arf1-dependent recruitment of coatomer to membranes
.
Cell
148
,
530
542
[PubMed]
33.
Liu
Y.
,
Kahn
R.A.
and
Prestegard
J.H.
(
2009
)
Structure and membrane interaction of myristoylated ARF1
.
Structure
17
,
79
87
[PubMed]
34.
Serafini
T.
,
Orci
L.
,
Amherdt
M.
,
Brunner
M.
,
Kahn
R.A.
and
Rothman
J.E.
(
1991
)
ADP-ribosylation factor is a subunit of the coat of Golgi-derived COP-coated vesicles: a novel role for a GTP-binding protein
.
Cell
67
,
239
253
[PubMed]
35.
Popoff
V.
,
Langer
J.D.
,
Reckmann
I.
,
Hellwig
A.
,
Kahn
R.A.
,
Brugger
B.
et al.
(
2011
)
Several ADP-ribosylation factor (Arf) isoforms support COPI vesicle formation
.
J. Biol. Chem.
286
,
35634
35642
[PubMed]
36.
Arakel
E.C.
and
Schwappach
B.
(
2018
)
Formation of COPI-coated vesicles at a glance
.
J. Cell Sci.
131
,
37.
Gomez-Navarro
N.
and
Miller
E.
(
2016
)
Protein sorting at the ER-Golgi interface
.
J. Cell Biol.
215
,
769
778
[PubMed]
38.
Pulvirenti
T.
,
Giannotta
M.
,
Capestrano
M.
,
Capitani
M.
,
Pisanu
A.
,
Polishchuk
R.S.
et al.
(
2008
)
A traffic-activated Golgi-based signalling circuit coordinates the secretory pathway
.
Nat. Cell Biol.
10
,
912
922
[PubMed]
39.
Johannes
L.
,
Tenza
D.
,
Antony
C.
and
Goud
B.
(
1997
)
Retrograde transport of KDEL-bearing B-fragment of Shiga toxin
.
J. Biol. Chem.
272
,
19554
19561
[PubMed]
40.
Brauer
P.
,
Parker
J.L.
,
Gerondopoulos
A.
,
Zimmermann
I.
,
Seeger
M.A.
,
Barr
F.A.
et al.
(
2019
)
Structural basis for pH-dependent retrieval of ER proteins from the Golgi by the KDEL receptor
.
Science
363
,
1103
1107
[PubMed]
41.
Wu
Z.
,
Newstead
S.
and
Biggin
P.C.
(
2020
)
The KDEL trafficking receptor exploits pH to tune the strength of an unusual short hydrogen bond
.
Sci. Rep.
10
,
16903
[PubMed]
42.
Bard
F.
,
Mazelin
L.
,
Pechoux-Longin
C.
,
Malhotra
V.
and
Jurdic
P.
(
2003
)
Src regulates Golgi structure and KDEL receptor-dependent retrograde transport to the endoplasmic reticulum
.
J. Biol. Chem.
278
,
46601
46606
[PubMed]
43.
Cancino
J.
,
Capalbo
A.
,
Di Campli
A.
,
Giannotta
M.
,
Rizzo
R.
,
Jung
J.E.
et al.
(
2014
)
Control systems of membrane transport at the interface between the endoplasmic reticulum and the Golgi
.
Dev. Cell.
30
,
280
294
[PubMed]
44.
Tapia
D.
,
Jimenez
T.
,
Zamora
C.
,
Espinoza
J.
,
Rizzo
R.
,
Gonzalez-Cardenas
A.
et al.
(
2019
)
KDEL receptor regulates secretion by lysosome relocation- and autophagy-dependent modulation of lipid-droplet turnover
.
Nat. Commun.
10
,
735
[PubMed]
45.
Potelle
S.
,
Klein
A.
and
Foulquier
F.
(
2015
)
Golgi post-translational modifications and associated diseases
.
J. Inherit. Metab. Dis.
38
,
741
751
[PubMed]
46.
Hickman
S.
and
Neufeld
E.F.
(
1972
)
A hypothesis for I-cell disease: defective hydrolases that do not enter lysosomes
.
Biochem. Biophys. Res. Commun.
49
,
992
999
[PubMed]
47.
Griffiths
G.
and
Simons
K.
(
1986
)
The trans Golgi network: sorting at the exit site of the Golgi complex
.
Science
234
,
438
443
[PubMed]
48.
Clermont
Y.
,
Rambourg
A.
and
Hermo
L.
(
1995
)
Trans‐Golgi network (TGN) of different cell types: Three‐dimensional structural characteristics and variability
.
Anat. Rec.
242
,
289
301
[PubMed]
49.
Trucco
A.
,
Polishchuk
R.S.
,
Martella
O.
,
Di Pentima
A.
,
Fusella
A.
,
Di Giandomenico
D.
et al.
(
2004
)
Secretory traffic triggers the formation of tubular continuities across Golgi sub-compartments
.
Nat. Cell Biol.
6
,
1071
1081
[PubMed]
50.
De Matteis
M.A.
and
Luini
A.
(
2008
)
Exiting the Golgi complex
.
Nat. Rev. Mol. Cell Biol.
9
,
273
284
[PubMed]
51.
Rodriguez-Boulan
E.
and
Müsch
A.
(
2005
)
Protein sorting in the Golgi complex: shifting paradigms
.
Biochim. Biophys. Acta Mol. Cell Res.
1744
,
455
464
52.
Rodriguez-Boulan
E.
,
Kreitzer
G.
and
Müsch
A.
(
2005
)
Organization of vesicular trafficking in epithelia
.
Nat. Rev. Mol. Cell Biol.
6
,
233
247
[PubMed]
53.
Koster
A.
,
Saftig
P.
,
Matzner
U.
,
von Figura
K.
,
Peters
C.
and
Pohlmann
R.
(
1993
)
Targeted disruption of the M(r) 46,000 mannose 6-phosphate receptor gene in mice results in misrouting of lysosomal proteins
.
EMBO J.
12
,
5219
5223
[PubMed]
54.
Ludwig
T.
,
Ovitt
C.E.
,
Bauer
U.
,
Hollinshead
M.
,
Remmler
J.
,
Lobel
P.
et al.
(
1993
)
Targeted disruption of the mouse cation-dependent mannose 6-phosphate receptor results in partial missorting of multiple lysosomal enzymes
.
EMBO J.
12
,
5225
5235
[PubMed]
55.
Nolan
C.M.
,
McCarthy
K.
,
Eivers
E.
,
Jirtle
R.L.
and
Byrnes
L.
(
2006
)
Mannose 6-phosphate receptors in an ancient vertebrate, zebrafish
.
Dev. Genes Evol.
216
,
144
151
[PubMed]
56.
Sohar
I.
,
Sleat
D.
,
Liu
C.-G.
,
Ludwig
T.
and
Lobel
P.
(
1998
)
Mouse mutants lacking the cation-independent mannose 6-phosphate/insulin-like growth factor II receptor are impaired in lysosomal enzyme transport: comparison of cation-independent and cation-dependent mannose 6-phosphate receptor-deficient mice
.
Biochem. J.
330
,
903
908
[PubMed]
57.
Bajaj
L.
,
Lotfi
P.
,
Pal
R.
,
Ronza
A.d.
,
Sharma
J.
and
Sardiello
M.
(
2019
)
Lysosome biogenesis in health and disease
.
J. Neurochem.
148
,
573
589
[PubMed]
58.
Olson
L.J.
,
Hindsgaul
O.
,
Dahms
N.M.
and
Kim
J.J.
(
2008
)
Structural insights into the mechanism of pH-dependent ligand binding and release by the cation-dependent mannose 6-phosphate receptor
.
J. Biol. Chem.
283
,
10124
10134
[PubMed]
59.
Olson
L.J.
,
Misra
S.K.
,
Ishihara
M.
,
Battaile
K.P.
,
Grant
O.C.
,
Sood
A.
et al.
(
2020
)
Allosteric regulation of lysosomal enzyme recognition by the cation-independent mannose 6-phosphate receptor
.
Commun. Biol.
3
,
498
[PubMed]
60.
Olson
L.J.
,
Peterson
F.C.
,
Castonguay
A.
,
Bohnsack
R.N.
,
Kudo
M.
,
Gotschall
R.R.
et al.
(
2010
)
Structural basis for recognition of phosphodiester-containing lysosomal enzymes by the cation-independent mannose 6-phosphate receptor
.
Proc. Natl. Acad. Sci. U.S.A.
107
,
12493
12498
[PubMed]
61.
Boonen
M.
,
Vogel
P.
,
Platt
K.A.
,
Dahms
N.
and
Kornfeld
S.
(
2009
)
Mice lacking mannose 6-phosphate uncovering enzyme activity have a milder phenotype than mice deficient for N-acetylglucosamine-1-phosphotransferase activity
.
Mol. Biol. Cell.
20
,
4381
4389
[PubMed]
62.
Gabel
C.A.
,
Goldberg
D.E.
and
Kornfeld
S.
(
1983
)
Identification and characterization of cells deficient in the mannose 6-phosphate receptor: evidence for an alternate pathway for lysosomal enzyme targeting
.
Proc. Natl. Acad. Sci.
80
,
775
779
63.
Glickman
J.N.
and
Kornfeld
S.
(
1993
)
Mannose 6-phosphate-independent targeting of lysosomal enzymes in I-cell disease B lymphoblasts
.
J. Cell Biol.
123
,
99
108
[PubMed]
64.
Petersen
C.M.
,
Nielsen
M.S.
,
Nykjær
A.
,
Jacobsen
L.
,
Tommerup
N.
,
Rasmussen
H.H.
et al.
(
1997
)
Molecular identification of a novel candidate sorting receptor purified from human brain by receptor-associated protein affinity chromatography
.
J. Biol. Chem.
272
,
3599
3605
[PubMed]
65.
Cooper
A.A.
and
Stevens
T.H.
(
1996
)
Vps10p cycles between the late-Golgi and prevacuolar compartments in its function as the sorting receptor for multiple yeast vacuolar hydrolases
.
J. Cell Biol.
133
,
529
541
[PubMed]
66.
Lefrancois
S.
,
Zeng
J.
,
Hassan
A.J.
,
Canuel
M.
and
Morales
C.R.
(
2003
)
The lysosomal trafficking of sphingolipid activator proteins (SAPs) is mediated by sortilin
.
EMBO J.
22
,
6430
6437
[PubMed]
67.
Canuel
M.
,
Lefrancois
S.
,
Zeng
J.
and
Morales
C.R.
(
2008
)
AP-1 and retromer play opposite roles in the trafficking of sortilin between the Golgi apparatus and the lysosomes
.
Biochem. Biophys. Res. Commun.
366
,
724
730
[PubMed]
68.
Canuel
M.
,
Bhattacharyya
N.
,
Balbis
A.
,
Yuan
L.
and
Morales
C.R.
(
2009
)
Sortilin and prosaposin localize to detergent-resistant membrane microdomains
.
Exp. Cell Res.
315
,
240
247
[PubMed]
69.
Nielsen
M.S.
,
Jacobsen
C.
,
Olivecrona
G.
,
Gliemann
J.
and
Petersen
C.M.
(
1999
)
Sortilin/neurotensin receptor-3 binds and mediates degradation of lipoprotein lipase
.
J. Biol. Chem.
274
,
8832
8836
[PubMed]
70.
Nilsson
S.K.
,
Christensen
S.
,
Raarup
M.K.
,
Ryan
R.O.
,
Nielsen
M.S.
and
Olivecrona
G.
(
2008
)
Endocytosis of apolipoprotein AV by members of the low density lipoprotein receptor and the VPS10p domain receptor families
.
J. Biol. Chem.
283
,
25920
25927
[PubMed]
71.
Hu
F.
,
Padukkavidana
T.
,
Vægter
C.B.
,
Brady
O.A.
,
Zheng
Y.
,
Mackenzie
I.R.
et al.
(
2010
)
Sortilin-mediated endocytosis determines levels of the frontotemporal dementia protein, progranulin
.
Neuron
68
,
654
667
[PubMed]
72.
Al-Akhrass
H.
,
Naves
T.
,
Vincent
F.
,
Magnaudeix
A.
,
Durand
K.
,
Bertin
F.
et al.
(
2017
)
Sortilin limits EGFR signaling by promoting its internalization in lung cancer
.
Nat. Commun.
8
,
1
15
[PubMed]
73.
Goettsch
C.
,
Hutcheson
J.D.
,
Aikawa
M.
,
Iwata
H.
,
Pham
T.
,
Nykjaer
A.
et al.
(
2016
)
Sortilin mediates vascular calcification via its recruitment into extracellular vesicles
.
J. Clin. Invest.
126
,
1323
1336
[PubMed]
74.
Pallesen
L.T.
,
Gustafsen
C.
,
Cramer
J.F.
,
Petersen
S.V.
,
Thirup
S.S.
,
Madsen
P.
et al.
(
2020
)
PAK Kinases target sortilin and modulate its sorting
.
Mol. Cell. Biol.
40
,
e00411
19
[PubMed]
75.
Cramer
J.F.
,
Gustafsen
C.
,
Behrens
M.A.
,
Oliveira
C.L.
,
Pedersen
J.S.
,
Madsen
P.
et al.
(
2010
)
GGA autoinhibition revisited
.
Traffic
11
,
259
273
[PubMed]
76.
McCormick
P.J.
,
Dumaresq-Doiron
K.
,
Pluviose
A.S.
,
Pichette
V.
,
Tosato
G.
and
Lefrancois
S.
(
2008
)
Palmitoylation controls recycling in lysosomal sorting and trafficking
.
Traffic
9
,
1984
1997
[PubMed]
77.
Dumaresq-Doiron
K.
,
Jules
F.
and
Lefrancois
S.
(
2013
)
Sortilin turnover is mediated by ubiquitination
.
Biochem. Biophys. Res. Commun.
433
,
90
95
[PubMed]
78.
Reczek
D.
,
Schwake
M.
,
Schroder
J.
,
Hughes
H.
,
Blanz
J.
,
Jin
X.
et al.
(
2007
)
LIMP-2 is a receptor for lysosomal mannose-6-phosphate-independent targeting of beta-glucocerebrosidase
.
Cell
131
,
770
783
[PubMed]
79.
Vega
M.A.
,
Rodriguez
F.
,
Segui
B.
,
Cales
C.
,
Alcalde
J.
and
Sandoval
I.V.
(
1991
)
Targeting of lysosomal integral membrane protein LIMP II. The tyrosine-lacking carboxyl cytoplasmic tail of LIMP II is sufficient for direct targeting to lysosomes
.
J. Biol. Chem.
266
,
16269
16272
[PubMed]
80.
Ogata
S.
and
Fukuda
M.
(
1994
)
Lysosomal targeting of Limp II membrane glycoprotein requires a novel Leu-Ile motif at a particular position in its cytoplasmic tail
.
J. Biol. Chem.
269
,
5210
5217
[PubMed]
81.
Conrad
K.S.
,
Cheng
T.W.
,
Ysselstein
D.
,
Heybrock
S.
,
Hoth
L.R.
,
Chrunyk
B.A.
et al.
(
2017
)
Lysosomal integral membrane protein-2 as a phospholipid receptor revealed by biophysical and cellular studies
.
Nat. Commun.
8
,
1908
[PubMed]
82.
Heybrock
S.
,
Kanerva
K.
,
Meng
Y.
,
Ing
C.
,
Liang
A.
,
Xiong
Z.J.
et al.
(
2019
)
Lysosomal integral membrane protein-2 (LIMP-2/SCARB2) is involved in lysosomal cholesterol export
.
Nat. Commun.
10
,
3521
[PubMed]
83.
Kuronita
T.
,
Eskelinen
E.L.
,
Fujita
H.
,
Saftig
P.
,
Himeno
M.
and
Tanaka
Y.
(
2002
)
A role for the lysosomal membrane protein LGP85 in the biogenesis and maintenance of endosomal and lysosomal morphology
.
J. Cell Sci.
115
,
4117
4131
[PubMed]
84.
Gonzalez
A.
,
Valeiras
M.
,
Sidransky
E.
and
Tayebi
N.
(
2014
)
Lysosomal integral membrane protein-2: a new player in lysosome-related pathology
.
Mol. Genet. Metab.
111
,
84
91
[PubMed]
85.
Ohno
H.
,
Aguilar
R.C.
,
Yeh
D.
,
Taura
D.
,
Saito
T.
and
Bonifacino
J.S.
(
1998
)
The medium subunits of adaptor complexes recognize distinct but overlapping sets of tyrosine-based sorting signals
.
J. Biol. Chem.
273
,
25915
25921
[PubMed]
86.
Doray
B.
,
Lee
I.
,
Knisely
J.
,
Bu
G.
and
Kornfeld
S.
(
2007
)
The gamma/sigma1 and alpha/sigma2 hemicomplexes of clathrin adaptors AP-1 and AP-2 harbor the dileucine recognition site
.
Mol. Biol. Cell.
18
,
1887
1896
[PubMed]
87.
Ohno
H.
,
Stewart
J.
,
Fournier
M.C.
,
Bosshart
H.
,
Rhee
I.
,
Miyatake
S.
et al.
(
1995
)
Interaction of tyrosine-based sorting signals with clathrin-associated proteins
.
Science
269
,
1872
1875
[PubMed]
88.
Janvier
K.
,
Kato
Y.
,
Boehm
M.
,
Rose
J.R.
,
Martina
J.A.
,
Kim
B.Y.
et al.
(
2003
)
Recognition of dileucine-based sorting signals from HIV-1 Nef and LIMP-II by the AP-1 gamma-sigma1 and AP-3 delta-sigma3 hemicomplexes
.
J. Cell Biol.
163
,
1281
1290
[PubMed]
89.
Kyttälä
A.
,
Ihrke
G.
,
Vesa
J.
,
Schell
M.J.
and
Luzio
J.P.
(
2004
)
Two motifs target Batten disease protein CLN3 to lysosomes in transfected nonneuronal and neuronal cells
.
Mol. Biol. Cell.
15
,
1313
1323
[PubMed]
90.
Storch
S.
,
Pohl
S.
and
Braulke
T.
(
2004
)
A dileucine motif and a cluster of acidic amino acids in the second cytoplasmic domain of the batten disease-related CLN3 protein are required for efficient lysosomal targeting
.
J. Biol. Chem.
279
,
53625
53634
[PubMed]
91.
Kyttälä
A.
,
Yliannala
K.
,
Schu
P.
,
Jalanko
A.
and
Luzio
J.P.
(
2005
)
AP-1 and AP-3 facilitate lysosomal targeting of Batten disease protein CLN3 via its dileucine motif
.
J. Biol. Chem.
280
,
10277
10283
[PubMed]
92.
Zhu
Y.
,
Traub
L.M.
and
Kornfeld
S.
(
1998
)
ADP-ribosylation factor 1 transiently activates high-affinity adaptor protein complex AP-1 binding sites on Golgi membranes
.
Mol. Biol. Cell.
9
,
1323
1337
[PubMed]
93.
Zhu
Y.
,
Drake
M.T.
and
Kornfeld
S.
(
1999
)
ADP-ribosylation factor 1 dependent clathrin-coat assembly on synthetic liposomes
.
Proc. Natl. Acad. Sci. U.S.A.
96
,
5013
5018
[PubMed]
94.
Wang
J.
,
Sun
H.Q.
,
Macia
E.
,
Kirchhausen
T.
,
Watson
H.
,
Bonifacino
J.S.
et al.
(
2007
)
PI4P promotes the recruitment of the GGA adaptor proteins to the trans-Golgi network and regulates their recognition of the ubiquitin sorting signal
.
Mol. Biol. Cell.
18
,
2646
2655
[PubMed]
95.
Donaldson
J.G.
and
Jackson
C.L.
(
2011
)
ARF family G proteins and their regulators: roles in membrane transport, development and disease
.
Nat. Rev. Mol. Cell Biol.
12
,
362
375
[PubMed]
96.
Barr
F.
and
Lambright
D.G.
(
2010
)
Rab GEFs and GAPs
.
Curr. Opin. Cell Biol.
22
,
461
470
[PubMed]
97.
Shinotsuka
C.
,
Waguri
S.
,
Wakasugi
M.
,
Uchiyama
Y.
and
Nakayama
K.
(
2002
)
Dominant-negative mutant of BIG2, an ARF-guanine nucleotide exchange factor, specifically affects membrane trafficking from the trans-Golgi network through inhibiting membrane association of AP-1 and GGA coat proteins
.
Biochem. Biophys. Res. Commun.
294
,
254
260
[PubMed]
98.
Shinotsuka
C.
,
Yoshida
Y.
,
Kawamoto
K.
,
Takatsu
H.
and
Nakayama
K.
(
2002
)
Overexpression of an ADP-ribosylation factor-guanine nucleotide exchange factor, BIG2, uncouples brefeldin A-induced adaptor protein-1 coat dissociation and membrane tubulation
.
J. Biol. Chem.
277
,
9468
9473
[PubMed]
99.
Ren
X.
,
Farias
G.G.
,
Canagarajah
B.J.
,
Bonifacino
J.S.
and
Hurley
J.H.
(
2013
)
Structural basis for recruitment and activation of the AP-1 clathrin adaptor complex by Arf1
.
Cell
152
,
755
767
[PubMed]
100.
Sauvageau
E.
,
McCormick
P.J.
and
Lefrancois
S.
(
2017
)
In vivo monitoring of the recruitment and activation of AP-1 by Arf1
.
Sci. Rep.
7
,
7148
[PubMed]
101.
Meyer
C.
,
Zizioli
D.
,
Lausmann
S.
,
Eskelinen
E.L.
,
Hamann
J.
,
Saftig
P.
et al.
(
2000
)
mu1A-adaptin-deficient mice: lethality, loss of AP-1 binding and rerouting of mannose 6-phosphate receptors
.
EMBO J.
19
,
2193
2203
[PubMed]
102.
Folsch
H.
,
Pypaert
M.
,
Schu
P.
and
Mellman
I.
(
2001
)
Distribution and function of AP-1 clathrin adaptor complexes in polarized epithelial cells
.
J. Cell Biol.
152
,
595
606
[PubMed]
103.
Bonifacino
J.S.
(
2014
)
Adaptor proteins involved in polarized sorting
.
J. Cell Biol.
204
,
7
17
[PubMed]
104.
Zizioli
D.
,
Meyer
C.
,
Guhde
G.
,
Saftig
P.
,
von Figura
K.
and
Schu
P.
(
1999
)
Early embryonic death of mice deficient in γ-adaptin
.
J. Biol. Chem.
274
,
5385
5390
[PubMed]
105.
Meyer
C.
,
Zizioli
D.
,
Lausmann
S.
,
Eskelinen
E.L.
,
Hamann
J.
,
Saftig
P.
et al.
(
2000
)
μ1A‐adaptin‐deficient mice: lethality, loss of AP‐1 binding and rerouting of mannose 6‐phosphate receptors
.
EMBO J.
19
,
2193
2203
[PubMed]
106.
Dell'Angelica
E.C.
and
Bonifacino
J.S.
(
2019
)
Coatopathies: genetic disorders of protein coats
.
Annu. Rev. Cell Dev. Biol.
35
,
131
168
[PubMed]
107.
Tarpey
P.S.
,
Stevens
C.
,
Teague
J.
,
Edkins
S.
,
O'Meara
S.
,
Avis
T.
et al.
(
2006
)
Mutations in the gene encoding the Sigma 2 subunit of the adaptor protein 1 complex, AP1S2, cause X-linked mental retardation
.
Am. J. Hum. Genet.
79
,
1119
1124
108.
Cacciagli
P.
,
Desvignes
J.-P.
,
Girard
N.
,
Delepine
M.
,
Zelenika
D.
,
Lathrop
M.
et al.
(
2014
)
AP1S2 is mutated in X-linked Dandy–Walker malformation with intellectual disability, basal ganglia disease and seizures (Pettigrew syndrome)
.
Eur. J. Hum. Genet.
22
,
363
368
[PubMed]
109.
Montpetit
A.
,
Côté
S.
,
Brustein
E.
,
Drouin
C.A.
,
Lapointe
L.
,
Boudreau
M.
et al.
(
2008
)
Disruption of AP1S1, causing a novel neurocutaneous syndrome, perturbs development of the skin and spinal cord
.
PLos Genet.
4
,
e1000296
[PubMed]
110.
Martinelli
D.
,
Travaglini
L.
,
Drouin
C.A.
,
Ceballos-Picot
I.
,
Rizza
T.
,
Bertini
E.
et al.
(
2013
)
MEDNIK syndrome: a novel defect of copper metabolism treatable by zinc acetate therapy
.
Brain
136
,
872
881
[PubMed]
111.
Alsaif
H.S.
,
Al-Owain
M.
,
Barrios-Llerena
M.E.
,
Gosadi
G.
,
Binamer
Y.
,
Devadason
D.
et al.
(
2019
)
Homozygous loss-of-function mutations in AP1B1, encoding beta-1 subunit of adaptor-related protein complex 1, cause MEDNIK-like syndrome
.
Am. J. Human Genetics
105
,
1016
1022
112.
Mahil
S.K.
,
Twelves
S.
,
Farkas
K.
,
Setta-Kaffetzi
N.
,
Burden
A.D.
,
Gach
J.E.
et al.
(
2016
)
AP1S3 mutations cause skin autoinflammation by disrupting keratinocyte autophagy and up-regulating IL-36 production
.
J. Invest. Dermatol.
136
,
2251
2259
[PubMed]
113.
Sanger
A.
,
Hirst
J.
,
Davies
A.K.
and
Robinson
M.S.
(
2019
)
Adaptor protein complexes and disease at a glance
.
J. Cell Sci.
132
,
jcs222992
[PubMed]
114.
Montpetit
A.
,
Cote
S.
,
Brustein
E.
,
Drouin
C.A.
,
Lapointe
L.
,
Boudreau
M.
et al.
(
2008
)
Disruption of AP1S1, causing a novel neurocutaneous syndrome, perturbs development of the skin and spinal cord
.
PLos Genet.
4
,
e1000296
[PubMed]
115.
Martinelli
D.
and
Dionisi-Vici
C.
(
2014
)
AP1S1 defect causing MEDNIK syndrome: a new adaptinopathy associated with defective copper metabolism
.
Ann. N. Y. Acad. Sci.
1314
,
55
63
[PubMed]
116.
Dell'Angelica
E.C.
,
Puertollano
R.
,
Mullins
C.
,
Aguilar
R.C.
,
Vargas
J.D.
,
Hartnell
L.M.
et al.
(
2000
)
GGAs: a family of ADP ribosylation factor-binding proteins related to adaptors and associated with the Golgi complex
.
J. Cell Biol.
149
,
81
94
[PubMed]
117.
Hirst
J.
,
Lui
W.W.
,
Bright
N.A.
,
Totty
N.
,
Seaman
M.N.
and
Robinson
M.S.
(
2000
)
A family of proteins with gamma-adaptin and VHS domains that facilitate trafficking between the trans-Golgi network and the vacuole/lysosome
.
J. Cell Biol.
149
,
67
80
[PubMed]
118.
Boman
A.L.
(
2001
)
GGA proteins: new players in the sorting game
.
J. Cell Sci.
114
,
3413
3418
[PubMed]
119.
Boman
A.L.
,
Zhang
C.-j.
,
Zhu
X.
and
Kahn
R.A.
(
2000
)
A family of ADP-ribosylation factor effectors that can alter membrane transport through the trans-Golgi
.
Mol. Biol. Cell.
11
,
1241
1255
[PubMed]
120.
Puertollano
R.
,
Aguilar
R.C.
,
Gorshkova
I.
,
Crouch
R.J.
and
Bonifacino
J.S.
(
2001
)
Sorting of mannose 6-phosphate receptors mediated by the GGAs
.
Science
292
,
1712
1716
[PubMed]
121.
Takatsu
H.
,
Katoh
Y.
,
Shiba
Y.
and
Nakayama
K.
(
2001
)
Golgi-localizing, γ-adaptin ear homology domain, ADP-ribosylation factor-binding (GGA) proteins interact with acidic dileucine sequences within the cytoplasmic domains of sorting receptors through their Vps27p/Hrs/STAM (VHS) domains
.
J. Biol. Chem.
276
,
28541
28545
[PubMed]
122.
Zhu
Y.
,
Doray
B.
,
Poussu
A.
,
Lehto
V.-P.
and
Kornfeld
S.
(
2001
)
Binding of GGA2 to the lysosomal enzyme sorting motif of the mannose 6-phosphate receptor
.
Science
292
,
1716
1718
[PubMed]
123.
Puertollano
R.
,
Randazzo
P.A.
,
Presley
J.F.
,
Hartnell
L.M.
and
Bonifacino
J.S.
(
2001
)
The GGAs promote ARF-dependent recruitment of clathrin to the TGN
.
Cell
105
,
93
102
[PubMed]
124.
Zhu
Y.
,
Doray
B.
,
Poussu
A.
,
Lehto
V.P.
and
Kornfeld
S.
(
2001
)
Binding of GGA2 to the lysosomal enzyme sorting motif of the mannose 6-phosphate receptor
.
Science
292
,
1716
1718
[PubMed]
125.
Lui
W.W.
,
Collins
B.M.
,
Hirst
J.
,
Motley
A.
,
Millar
C.
,
Schu
P.
et al.
(
2003
)
Binding partners for the COOH-terminal appendage domains of the GGAs and gamma-adaptin
.
Mol. Biol. Cell.
14
,
2385
2398
[PubMed]
126.
Mattera
R.
,
Arighi
C.N.
,
Lodge
R.
,
Zerial
M.
and
Bonifacino
J.S.
(
2003
)
Divalent interaction of the GGAs with the Rabaptin-5-Rabex-5 complex
.
EMBO J.
22
,
78
88
[PubMed]
127.
Doray
B.
,
Ghosh
P.
,
Griffith
J.
,
Geuze
H.J.
and
Kornfeld
S.
(
2002
)
Cooperation of GGAs and AP-1 in packaging MPRs at the trans-Golgi network
.
Science
297
,
1700
1703
[PubMed]
128.
Doray
B.
,
Liu
L.
,
Lee
W.S.
,
Jennings
B.C.
and
Kornfeld
S.
(
2021
)
Inactivation of the three GGA genes in HeLa cells partially compromises lysosomal enzyme sorting
.
FEBS Open Bio.
11
,
367
374
[PubMed]
129.
Lefrancois
S.
and
McCormick
P.J.
(
2007
)
The Arf GEF GBF1 is required for GGA recruitment to Golgi membranes
.
Traffic
8
,
1440
1451
[PubMed]
130.
Shiba
Y.
,
Kametaka
S.
,
Waguri
S.
,
Presley
J.F.
and
Randazzo
P.A.
(
2013
)
ArfGAP3 regulates the transport of cation-independent mannose 6-phosphate receptor in the post-Golgi compartment
.
Curr. Biol.
23
,
1945
1951
[PubMed]
131.
Arighi
C.N.
,
Hartnell
L.M.
,
Aguilar
R.C.
,
Haft
C.R.
and
Bonifacino
J.S.
(
2004
)
Role of the mammalian retromer in sorting of the cation-independent mannose 6-phosphate receptor
.
J. Cell Biol.
165
,
123
133
[PubMed]
132.
Seaman
M.N.
(
2004
)
Cargo-selective endosomal sorting for retrieval to the Golgi requires retromer
.
J. Cell Biol.
165
,
111
122
[PubMed]
133.
Rojas
R.
,
van Vlijmen
T.
,
Mardones
G.A.
,
Prabhu
Y.
,
Rojas
A.L.
,
Mohammed
S.
et al.
(
2008
)
Regulation of retromer recruitment to endosomes by sequential action of Rab5 and Rab7
.
J. Cell Biol.
183
,
513
526
[PubMed]
134.
Seaman
M.N.
,
Harbour
M.E.
,
Tattersall
D.
,
Read
E.
and
Bright
N.
(
2009
)
Membrane recruitment of the cargo-selective retromer subcomplex is catalysed by the small GTPase Rab7 and inhibited by the Rab-GAP TBC1D5
.
J. Cell Sci.
122
,
2371
2382
[PubMed]
135.
Hutagalung
A.H.
and
Novick
P.J.
(
2011
)
Role of Rab GTPases in membrane traffic and cell physiology
.
Physiol. Rev.
91
,
119
149
[PubMed]
136.
Kinchen
J.M.
and
Ravichandran
K.S.
(
2010
)
Identification of two evolutionarily conserved genes regulating processing of engulfed apoptotic cells
.
Nature
464
,
778
782
[PubMed]
137.
Yan
B.R.
,
Li
T.
,
Coyaud
E.
,
Laurent
E.M.N.
,
St-Germain
J.
,
Zhou
Y.
et al.
(
2022
)
C5orf51 is a component of the MON1-CCZ1 complex and controls RAB7A localization and stability during mitophagy
.
Autophagy
18
,
829
840
138.
Peralta
E.R.
,
Martin
B.C.
and
Edinger
A.L.
(
2010
)
Differential effects of TBC1D15 and mammalian Vps39 on Rab7 activation state, lysosomal morphology, and growth factor dependence
.
J. Biol. Chem.
285
,
16814
16821
[PubMed]
139.
Seaman
M.N.J.
,
Mukadam
A.S.
and
Breusegem
S.Y.
(
2018
)
Inhibition of TBC1D5 activates Rab7a and can enhance the function of the retromer cargo-selective complex
.
J. Cell Sci.
131
,
[PubMed]
140.
Heo
J.M.
,
Ordureau
A.
,
Swarup
S.
,
Paulo
J.A.
,
Shen
K.
,
Sabatini
D.M.
et al.
(
2018
)
RAB7A phosphorylation by TBK1 promotes mitophagy via the PINK-PARKIN pathway
.
Sci. Adv.
4
,
eaav0443
[PubMed]
141.
Hanafusa
H.
,
Yagi
T.
,
Ikeda
H.
,
Hisamoto
N.
,
Nishioka
T.
,
Kaibuchi
K.
et al.
(
2019
)
LRRK1 phosphorylation of Rab7 at S72 links trafficking of EGFR-containing endosomes to its effector RILP
.
J. Cell Sci.
132
,
[PubMed]
142.
Babur
O.
,
Melrose
A.R.
,
Cunliffe
J.M.
,
Klimek
J.
,
Pang
J.
,
Sepp
A.I.
et al.
(
2020
)
Phosphoproteomic quantitation and causal analysis reveal pathways in GPVI/ITAM-mediated platelet activation programs
.
Blood
136
,
2346
2358
[PubMed]
143.
Ritter
J.L.
,
Zhu
Z.
,
Thai
T.C.
,
Mahadevan
N.R.
,
Mertins
P.
,
Knelson
E.H.
et al.
(
2020
)
Phosphorylation of RAB7 by TBK1/IKKepsilon regulates innate immune signaling in triple-negative breast cancer
.
Cancer Res.
80
,
44
56
[PubMed]
144.
Modica
G.
,
Skorobogata
O.
,
Sauvageau
E.
,
Vissa
A.
,
Yip
C.M.
,
Kim
P.K.
et al.
(
2017
)
Rab7 palmitoylation is required for efficient endosome-to-TGN trafficking
.
J. Cell Sci.
130
,
2579
2590
[PubMed]
145.
Song
P.
,
Trajkovic
K.
,
Tsunemi
T.
and
Krainc
D.
(
2016
)
Parkin modulates endosomal organization and function of the endo-lysosomal pathway
.
J. Neurosci.
36
,
2425
2437
[PubMed]
146.
Sapmaz
A.
,
Berlin
I.
,
Bos
E.
,
Wijdeven
R.H.
,
Janssen
H.
,
Konietzny
R.
et al.
(
2019
)
USP32 regulates late endosomal transport and recycling through deubiquitylation of Rab7
.
Nat. Commun.
10
,
1454
[PubMed]
147.
Yasa
S.
,
Modica
G.
,
Sauvageau
E.
,
Kaleem
A.
,
Hermey
G.
and
Lefrancois
S.
(
2020
)
CLN3 regulates endosomal function by modulating Rab7A-effector interactions
.
J. Cell Sci.
133
,
[PubMed]
148.
Yasa
S.
,
Sauvageau
E.
,
Modica
G.
and
Lefrancois
S.
(
2021
)
CLN5 and CLN3 function as a complex to regulate endolysosome function
.
Biochem. J.
478
,
2339
2357
[PubMed]
149.
Follett
J.
,
Norwood
S.J.
,
Hamilton
N.A.
,
Mohan
M.
,
Kovtun
O.
,
Tay
S.
et al.
(
2014
)
The Vps35 D620N mutation linked to Parkinson's disease disrupts the cargo sorting function of retromer
.
Traffic
15
,
230
244
[PubMed]
150.
McGough
I.J.
,
Steinberg
F.
,
Jia
D.
,
Barbuti
P.A.
,
McMillan
K.J.
,
Heesom
K.J.
et al.
(
2014
)
Retromer binding to FAM21 and the WASH complex is perturbed by the Parkinson disease-linked VPS35(D620N) mutation
.
Curr. Biol.
24
,
1670
1676
[PubMed]
151.
Zavodszky
E.
,
Seaman
M.N.
,
Moreau
K.
,
Jimenez-Sanchez
M.
,
Breusegem
S.Y.
,
Harbour
M.E.
et al.
(
2014
)
Mutation in VPS35 associated with Parkinson's disease impairs WASH complex association and inhibits autophagy
.
Nat. Commun.
5
,
3828
[PubMed]
152.
Follett
J.
,
Bugarcic
A.
,
Yang
Z.
,
Ariotti
N.
,
Norwood
S.J.
,
Collins
B.M.
et al.
(
2016
)
Parkinson disease-linked Vps35 R524W mutation impairs the endosomal association of retromer and induces alpha-synuclein aggregation
.
J. Biol. Chem.
291
,
18283
18298
[PubMed]
153.
Small
S.A.
,
Kent
K.
,
Pierce
A.
,
Leung
C.
,
Kang
M.S.
,
Okada
H.
et al.
(
2005
)
Model-guided microarray implicates the retromer complex in Alzheimer’s disease
.
Ann. Neurol.
58
,
909
919
[PubMed]
154.
Muhammad
A.
,
Flores
I.
,
Zhang
H.
,
Yu
R.
,
Staniszewski
A.
,
Planel
E.
et al.
(
2008
)
Retromer deficiency observed in Alzheimer’s disease causes hippocampal dysfunction, neurodegeneration, and Abeta accumulation
.
Proc. Natl. Acad. Sci. U.S.A.
105
,
7327
7332
[PubMed]
155.
Muzio
L.
,
Sirtori
R.
,
Gornati
D.
,
Eleuteri
S.
,
Fossaghi
A.
,
Brancaccio
D.
et al.
(
2020
)
Retromer stabilization results in neuroprotection in a model of amyotrophic lateral sclerosis
.
Nat. Commun.
11
,
3848
[PubMed]
156.
Romano
R.
,
Rivellini
C.
,
De Luca
M.
,
Tonlorenzi
R.
,
Beli
R.
,
Manganelli
F.
et al.
(
2021
)
Alteration of the late endocytic pathway in Charcot-Marie-Tooth type 2B disease
.
Cell. Mol. Life Sci.:CMLS
78
,
351
372
[PubMed]
157.
Simonetti
B.
,
Paul
B.
,
Chaudhari
K.
,
Weeratunga
S.
,
Steinberg
F.
,
Gorla
M.
et al.
(
2019
)
Molecular identification of a BAR domain-containing coat complex for endosomal recycling of transmembrane proteins
.
Nat. Cell Biol.
21
,
1219
1233
[PubMed]
158.
Cui
Y.
,
Carosi
J.M.
,
Yang
Z.
,
Ariotti
N.
,
Kerr
M.C.
,
Parton
R.G.
et al.
(
2019
)
Retromer has a selective function in cargo sorting via endosome transport carriers
.
J. Cell Biol.
218
,
615
631
[PubMed]
159.
Tu
Y.
and
Seaman
M.N.J.
(
2021
)
Navigating the controversies of retromer-mediated endosomal protein sorting
.
Front. Cell Development. Biol.
9
,
658741
160.
Gieselmann
V.
,
Hasilik
A.
and
von Figura
K.
(
1985
)
Processing of human cathepsin D in lysosomes in vitro
.
J. Biol. Chem.
260
,
3215
3220
[PubMed]
161.
Laurent-Matha
V.
,
Derocq
D.
,
Prebois
C.
,
Katunuma
N.
and
Liaudet-Coopman
E.
(
2006
)
Processing of human cathepsin D is independent of its catalytic function and auto-activation: involvement of cathepsins L and B
.
J. Biochem. (Tokyo)
139
,
363
371
162.
Desjardins
M.
(
1995
)
Biogenesis of phagolysosomes: the ‘kiss and run’ hypothesis
.
Trends Cell Biol.
5
,
183
186
[PubMed]
163.
Bright
N.A.
,
Reaves
B.J.
,
Mullock
B.M.
and
Luzio
J.P.
(
1997
)
Dense core lysosomes can fuse with late endosomes and are re-formed from the resultant hybrid organelles
.
J. Cell Sci.
110
,
2027
2040
[PubMed]
164.
Mullock
B.M.
,
Bright
N.A.
,
Fearon
C.W.
,
Gray
S.R.
and
Luzio
J.P.
(
1998
)
Fusion of lysosomes with late endosomes produces a hybrid organelle of intermediate density and is NSF dependent
.
J. Cell Biol.
140
,
591
601
[PubMed]
165.
Luzio
J.P.
,
Hackmann
Y.
,
Dieckmann
N.M.
and
Griffiths
G.M.
(
2014
)
The biogenesis of lysosomes and lysosome-related organelles
.
Cold Spring Harbor Perspect. Biol.
6
,
a016840
166.
Bright
N.A.
,
Gratian
M.J.
and
Luzio
J.P.
(
2005
)
Endocytic delivery to lysosomes mediated by concurrent fusion and kissing events in living cells
.
Curr. Biol.
15
,
360
365
[PubMed]
167.
Cabrera
M.
,
Ostrowicz
C.W.
,
Mari
M.
,
LaGrassa
T.J.
,
Reggiori
F.
and
Ungermann
C.
(
2009
)
Vps41 phosphorylation and the Rab Ypt7 control the targeting of the HOPS complex to endosome-vacuole fusion sites
.
Mol. Biol. Cell.
20
,
1937
1948
[PubMed]
168.
Brocker
C.
,
Kuhlee
A.
,
Gatsogiannis
C.
,
Balderhaar
H.J.
,
Honscher
C.
,
Engelbrecht-Vandre
S.
et al.
(
2012
)
Molecular architecture of the multisubunit homotypic fusion and vacuole protein sorting (HOPS) tethering complex
.
Proc. Natl. Acad. Sci. U.S.A.
109
,
1991
1996
[PubMed]
169.
van der Kant
R.
,
Fish
A.
,
Janssen
L.
,
Janssen
H.
,
Krom
S.
,
Ho
N.
et al.
(
2013
)
Late endosomal transport and tethering are coupled processes controlled by RILP and the cholesterol sensor ORP1L
.
J. Cell Sci.
126
,
3462
3474
[PubMed]
170.
McEwan
D.G.
,
Popovic
D.
,
Gubas
A.
,
Terawaki
S.
,
Suzuki
H.
,
Stadel
D.
et al.
(
2015
)
PLEKHM1 regulates autophagosome-lysosome fusion through HOPS complex and LC3/GABARAP proteins
.
Mol. Cell.
57
,
39
54
[PubMed]
171.
Wijdeven
R.H.
,
Janssen
H.
,
Nahidiazar
L.
,
Janssen
L.
,
Jalink
K.
,
Berlin
I.
et al.
(
2016
)
Cholesterol and ORP1L-mediated ER contact sites control autophagosome transport and fusion with the endocytic pathway
.
Nat. Commun.
7
,
11808
[PubMed]
172.
Garg
S.
,
Sharma
M.
,
Ung
C.
,
Tuli
A.
,
Barral
D.C.
,
Hava
D.L.
et al.
(
2011
)
Lysosomal trafficking, antigen presentation, and microbial killing are controlled by the Arf-like GTPase Arl8b
.
Immunity
35
,
182
193
[PubMed]
173.
Khatter
D.
,
Raina
V.B.
,
Dwivedi
D.
,
Sindhwani
A.
,
Bahl
S.
and
Sharma
M.
(
2015
)
The small GTPase Arl8b regulates assembly of the mammalian HOPS complex on lysosomes
.
J. Cell Sci.
128
,
1746
1761
[PubMed]
174.
Marwaha
R.
,
Arya
S.B.
,
Jagga
D.
,
Kaur
H.
,
Tuli
A.
and
Sharma
M.
(
2017
)
The Rab7 effector PLEKHM1 binds Arl8b to promote cargo traffic to lysosomes
.
J. Cell Biol.
216
,
1051
1070
[PubMed]
175.
Zick
M.
and
Wickner
W.
(
2013
)
The tethering complex HOPS catalyzes assembly of the soluble SNARE Vam7 into fusogenic trans-SNARE complexes
.
Mol. Biol. Cell.
24
,
3746
3753
[PubMed]
176.
Baker
R.W.
,
Jeffrey
P.D.
,
Zick
M.
,
Phillips
B.P.
,
Wickner
W.T.
and
Hughson
F.M.
(
2015
)
A direct role for the Sec1/Munc18-family protein Vps33 as a template for SNARE assembly
.
Science
349
,
1111
1114
[PubMed]
177.
Song
H.
,
Orr
A.S.
,
Lee
M.
,
Harner
M.E.
and
Wickner
W.T.
(
2020
)
HOPS recognizes each SNARE, assembling ternary trans-complexes for rapid fusion upon engagement with the 4th SNARE
.
Elife
9
,
178.
Pryor
P.R.
,
Mullock
B.M.
,
Bright
N.A.
,
Lindsay
M.R.
,
Gray
S.R.
,
Richardson
S.C.
et al.
(
2004
)
Combinatorial SNARE complexes with VAMP7 or VAMP8 define different late endocytic fusion events
.
EMBO Rep.
5
,
590
595
[PubMed]
179.
Sharma
S.
and
Lindau
M.
(
2018
)
Molecular mechanism of fusion pore formation driven by the neuronal SNARE complex
.
Proc. Natl. Acad. Sci. U.S.A.
115
,
12751
12756
[PubMed]
180.
Follett
J.
,
Bugarcic
A.
,
Collins
B.M.
and
Teasdale
R.D.
(
2018
)
Retromer’s role in endosomal trafficking and impaired function in neurodegenerative diseases
.
Curr. Protein Pept. Sci.
18
,
687
701
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).