Next Article in Journal
Molecular Alterations in Pancreatic Cancer: Transfer to the Clinic
Next Article in Special Issue
Post-Translational Modifications of Retroviral HIV-1 Gag Precursors: An Overview of Their Biological Role
Previous Article in Journal
Changes in Apoptotic Pathways in MOLM-13 Cell Lines after Induction of Resistance to Hypomethylating Agents
Previous Article in Special Issue
The Roles of the Ubiquitin–Proteasome System in the Endoplasmic Reticulum Stress Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

How Is the Fidelity of Proteins Ensured in Terms of Both Quality and Quantity at the Endoplasmic Reticulum? Mechanistic Insights into E3 Ubiquitin Ligases

1
Department of Biochemistry, College of Medicine, Chungnam National University, Daejeon 35015, Korea
2
Department of Medical Science, College of Medicine, Chungnam National University, Daejeon 35015, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(4), 2078; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22042078
Submission received: 29 January 2021 / Revised: 16 February 2021 / Accepted: 16 February 2021 / Published: 19 February 2021

Abstract

:
The endoplasmic reticulum (ER) is an interconnected organelle that plays fundamental roles in the biosynthesis, folding, stabilization, maturation, and trafficking of secretory and transmembrane proteins. It is the largest organelle and critically modulates nearly all aspects of life. Therefore, in the endoplasmic reticulum, an enormous investment of resources, including chaperones and protein folding facilitators, is dedicated to adequate protein maturation and delivery to final destinations. Unfortunately, the folding and assembly of proteins can be quite error-prone, which leads to the generation of misfolded proteins. Notably, protein homeostasis, referred to as proteostasis, is constantly exposed to danger by flows of misfolded proteins and subsequent protein aggregates. To maintain proteostasis, the ER triages and eliminates terminally misfolded proteins by delivering substrates to the ubiquitin–proteasome system (UPS) or to the lysosome, which is termed ER-associated degradation (ERAD) or ER-phagy, respectively. ERAD not only eliminates misfolded or unassembled proteins via protein quality control but also fine-tunes correctly folded proteins via protein quantity control. Intriguingly, the diversity and distinctive nature of E3 ubiquitin ligases determine efficiency, complexity, and specificity of ubiquitination during ERAD. ER-phagy utilizes the core autophagy machinery and eliminates ERAD-resistant misfolded proteins. Here, we conceptually outline not only ubiquitination machinery but also catalytic mechanisms of E3 ubiquitin ligases. Further, we discuss the mechanistic insights into E3 ubiquitin ligases involved in the two guardian pathways in the ER, ERAD and ER-phagy. Finally, we provide the molecular mechanisms by which ERAD and ER-phagy conduct not only protein quality control but also protein quantity control to ensure proteostasis and subsequent organismal homeostasis.

1. Introduction

Post-translational modification (PTM) by ubiquitin (ubiquitination) is a very powerful and finely tuned process that is spatiotemporally involved in nearly all aspects of biological and physiological functions for the maintenance of organismal homeostasis. Comprehensive proteomics studies have identified tens of thousands of ubiquitination sites on thousands of proteins, indicating that the vast majority of cellular proteins are ubiquitinated during their lifetime. Among the enzymes catalyzing ubiquitination—including E1 ubiquitin-activating enzymes, E2 ubiquitin-conjugating enzymes, and E3 ubiquitin ligases—E3 ubiquitin ligases govern the efficiency, complexity, and specificity of ubiquitination, since E3 ubiquitin ligases participate in the specific recognition of signals present on substrate proteins [1,2,3,4,5,6,7]. Importantly, ubiquitination leads to alterations in activity, interaction, intracellular trafficking, stability, and subcellular localization of substrate proteins through its reversible nature, rapid kinetics and versatility of outcomes, thereby orchestrating a plethora of biological processes involving stress responses, immune modulation, signaling transduction, control of cell cycle, division, and proliferation [5,8,9,10,11,12,13,14,15]. Therefore, aberrant function or dysregulation of ubiquitination could be at the roots of the development of many severe diseases such as cancer, immune disorders, neurodegeneration and susceptibility to infections.
In the endoplasmic reticulum (ER), folding of proteins is inherently error-prone. Even with an elaborate network to ensure adequate protein folding and assembly, misfolding and subsequent aggregates of proteins can occur quite frequently. To maintain protein homeostasis, referred to as proteostasis in the ER, eukaryotes have evolved multiple quality control mechanisms involving ER-associated degradation (ERAD), ER-phagy, Golgi quality control, and plasma membrane quality control [16]. While most of the misfolded proteins in the ER are efficiently eliminated via ERAD, ERAD can be constrained by aggregation propensity, the severity of mutation, the nature of folding lesions and kinetics, and the thermodynamics of folding, which leads to the development of another mechanism, ER-phagy. Interestingly, some misfolded proteins are subject to more than one quality control, suggesting the cooperation of the quality control mechanisms to efficiently eliminate the misfolded proteins. Preference among ERAD and ER-phagy could rely on the versatile characteristics of substrate proteins, including topology, misfolding positions and modifications of substrate proteins [17].
In this review, we first conceptually outline ubiquitination machinery and catalytic mechanisms of E3 ubiquitin ligases. We then highlight the mechanistic insights into the E3 ubiquitin ligases involved in the protein quality and quantity control processes at the ER. We finally discuss both protein quality control and quantity control at the ER, which are mainly mediated by ERAD and ER-phagy.

2. Ubiquitination

2.1. A Conceptual Overview of Ubiquitination

Ubiquitin is a small, compact and highly conserved globular protein, with the exception of its unrestrained and flexible C-terminal tail. Ubiquitin harbors a unique functional isoleucine–hydrophobic patch that supports essential noncovalent interactions during ubiquitination and signal recognition [18]. To achieve abundant cellular concentrations of ubiquitin, four different genes, UBB, UBC, RPS27, and UBA52, encode ubiquitin in mammals. Genes UBB and UBC, as polyubiquitin cassettes, encode linear fusions of three and nine ubiquitin molecules, respectively, while RPS27A and UBA52 encode ubiquitin as in-frame fusion to small and large ribosomal proteins, respectively, from which deubiquitinating enzymes (DUBs) cleave off the monomeric ubiquitin [19,20,21].
Ubiquitination is a multistep process achieved by E1 ubiquitin-activating enzymes, E2 ubiquitin-conjugating enzymes and E3 ubiquitin ligase sequentially activating, conjugating and ligating ubiquitin to substrate proteins (Figure 1) [8,22]. E1 ubiquitin-activating enzymes consist of ubiquitin-like modifier activating enzyme 1 (UBA1) and UBA6. E1 activating enzymes possess a multidomain architecture, including the adenylation domain, which binds ubiquitin and ATP-Mg2+, the catalytic cysteine domain and the ubiquitin-fold domain (UFD), which supports interactions with E2 conjugating enzymes [23,24,25,26,27]. Additionally, a long flexible linker connects the active adenylation domain with the catalytic cysteine domain [28]. Upon binding to both ubiquitin and ATP-Mg2+, the E1 ubiquitin-activating enzyme mediates the formation of a high-energy ubiquitin–AMP complex via ubiquitin C-terminal acyl adenylation and then a catalytic cysteine on the E1 activating enzyme replaces the AMP group in the ubiquitin–AMP complex, thereby leading to the generation of an activated ubiquitin via a covalent thioester bond [29].
The E1 ubiquitin complex transfers ubiquitin to the E2 ubiquitin-conjugating enzyme through a transthioesterification reaction [30]. There are approximately 40 E2 ubiquitin-conjugating enzymes in human genome, of which 35 are dedicated to ubiquitination. E2 conjugating enzymes possess a core ubiquitin conjugation fold (UBC), along with N-terminal and C-terminal extensions [31,32]. In accordance, E2 conjugating enzymes are subdivided into class II, which possess N-terminal extensions, class III, which possess C-terminal extensions, class I, which possess neither extensions, and class IV, which possess both N-terminal and C-terminal extensions [32].
E3 ubiquitin ligases finally associate with both the E2 ubiquitin complex and a substrate protein to catalyze the covalent transfer of ubiquitin, thereby leading to the formation of an amide bond of ubiquitin, most commonly with a lysine residue or the N-terminus of the substrate [33]. Interestingly, ubiquitination of non-lysine residues has been reported [34,35]. Ubiquitin can be targeted to thiol or hydroxyl side chains in cysteine, serine, and threonine residues on the substrate proteins [35,36,37,38]. Moreover, ubiquitin can also be conjugated to a serine residue on the substrate protein in a phosphoribosylation-dependent manner [39,40].
DUBs reverse ubiquitination and switch off the ubiquitin signal, which deals with the matched complexity of ubiquitination. In humans, more than 100 DUBs have been identified. DUBs consist of six structurally distinct families of different cysteine and metalloproteases. The five families of cysteine proteases are the ubiquitin-specific proteases (USPs) with 54 members, the ubiquitin C-terminal hydrolases (UCHs) with four members, the ovarian tumor proteases (OTUs) with 16 members, the Josephin family with four members and the motif interacting with ubiquitin (MIU)-containing novel DUB family (MINDYs) with four members [14,41,42,43]. The sixth family of DUBs is Zn-dependent JAB1/MPN/MOV34 metalloproteases (JAMMs) with 16 members [42]. The biochemistry and the roles of DUBs in regulating diverse cellular processes are well-appreciated in recent reviews and thus we will not describe DUBs in further detail here.

2.2. Understanding the Catalytic Mechanisms of E3 Ubiquitin Ligases

Despite the mechanism of catalysis being conserved in all eukaryotes, the network of ubiquitin is more sophisticated in mammals due to the large number of E2 ubiquitin-conjugating and E3 ubiquitin ligase enzymes, compared to only two E1 ubiquitin-activating enzymes. There are approximately 40 E2 conjugating enzymes and 700 E3 ligases in humans that function in numerous combinations to ubiquitinate virtually every protein in the cell [44]. E3 ubiquitin ligases determine the efficiency, complexity, and specificity of ubiquitination, since E3 ligases participate in the specific recognition of signals present on the substrate proteins and E2 ubiquitin-conjugating enzymes work only with a limited set of E3 ligases [1,2,3,4,5,6,7].
E3 ligases are divided into three main families on the basis of the structure of the ubiquitin-charged E2 conjugating enzyme-binding domain and the ubiquitin transfer mechanism (Figure 1) [5,9,33]. Really interesting new gene (RING) E3 ligases are characterized by their RING or U-box catalytic domain and catalyze a direct transfer of ubiquitin from the ubiquitin-charged E2 conjugating enzyme to the substrate proteins. With approximately 700 predicted E3 ligases, the RING E3 ligases constitute the largest family of E3 ligases. RING domains are structurally characterized by the presence of two zinc ions essential for RING domain folding, which are coordinated by cysteine and histidine residues into a cross-braced configuration [45]. Although U-box E3 ligases are also classified as RING E3 ligases, the U-box E3 ligases do not have zinc ions [46]. RING and U-box domains interact with the ubiquitin-charged E2 conjugating enzyme in a closed and active conformation and optimize the geometry of the E2-ubiquitin thioester bond and its surrounding residues to favor the nucleophilic attack by the lysine residue on the substrate, which facilitates the transfer of ubiquitin with minimal binding of the E3 ligases to ubiquitin itself [5,33,47,48,49]. RING E3 ligases are a highly diverse group functioning as monomers, homodimers, or heterodimers. Casitas B-lineage Lymphoma (CBL), C-C chemokine receptor type 4 (CCR4)-NOT transcription complex, subunit 4 (CNOT4), and RING finger protein (RNF38) are active as monomers [50,51,52], whereas cellular inhibitor of apoptosis protein 2 (cIAP2), TNF receptor associated factor 6 (TRAF6), and RNF4 are only active as homodimers [53,54,55]. Additionally, breast cancer gene 1 (BRCA1)-associated RING domain protein 1 (BARD1), B cell-specific Moloney murine leukemia virus integration site 1 (BMI1), and mouse double minute X (MDMX) become functional upon heterodimerization with a RING domain-containing partner, BRCA1, RNF2 and MDM2, respectively [56,57,58,59,60]. Further, the tripartite motif (TRIM) family of RING E3 ligases assembles homodimers, heterodimers, and oligomers via the RING domains and coiled coil regions [61]. Some RING E3 ligases, such as cullin-RING ligases (CRLs) and anaphase-promoting complex/cyclosome (APC/C), are composed of multiple subunits. CRLs, a highly diverse class of E3 ubiquitin ligases, are composed of a RING E3 (RING-box protein 1 (RBX1) or RBX2), a cullin protein (CUL1, CUL2, CUL3, CUL4A/4B, CUL5, or CUL7) as a scaffold, and a substrate receptor [62,63]. APC/C consists of a 14-subunit core complex, including a RING E3 APC11, a cullin-like subunit APC2, and a co-activator protein, cell division cycle protein 20 (CDC20) or Cdh1, for recruitment of a substrate protein [64].
With approximately 30 members in the human genome, homologous to the E6-associated protein (E6AP) carboxyl terminus (HECT)-type E3 ligases contain not only a C-terminal HECT domain that forms a thioester intermediate with ubiquitin but also various N-terminal domains that govern substrate specificity and cellular localization, both of which adopt a characteristic bilobal structure with a short hinge that enables the lobes to rotate and supports ubiquitin transfer [65]. On the basis of their N-terminal protein–protein interaction domains, HECT E3 ligases are subdivided into three groups: the neural precursor cell expressed developmentally down-regulated protein 4 (NEDD4) family, characterized by the presence of WW domains, crucial for the interaction with PY motifs in substrates; the HERC family, characterized by the presence of a HECT domain and one or more regulator of chromosome condensation 1 (RCC1)-like domains; and HECTs with other protein–protein interaction domains [66]. HECT-type E3 ligases mediate ubiquitin transfer to the substrate proteins via a two-step reaction, in which ubiquitin is transferred to an active-site cysteine on the E3 ligase to form a thioester bond with ubiquitin and then is conjugated to the substrate protein.
RING-in-between-RINGs (IBR)-RING (RBR)-type E3 ligases, as a unique family of RING–HECT hybrid E3 ligases, are characterized by a homologous sequence, consisting of two predicted RING fingers, RING1 and RING2, and a central IBR zinc-binding domain [67]. Additionally, RBR E3 ligases also possess unique N- and C-terminal flanking domains, which regulate the catalytic activities of the E3 ligases [68,69,70,71]. The human genome encodes approximately 14 RBR-type E3 ligases, including HOIL-1L interacting protein (HOIP), PRRKIN (the product of the PARK2 gene mutated in Parkinson’s disease), the central E3 subunit of the linear ubiquitin chain assembly complex (LUBAC), two RING fingers and DRIL 1 (TRAID1), and human homologue of Ariadne (HHARI), all of which are highly conserved from yeast to humans [72,73,74,75,76]. RBR E3 ligases use an E2-binding RING domain and a second RING2 domain that contains an active cysteine residue for the formation of an E3~ubiquitin intermediate, from which ubiquitin is eventually transferred to the substrate proteins [77].
In spite of a plethora of structurally unrelated proteins, their ubiquitination is highly selective. Substrate specificity and type of polyubiquitin chains are conferred by E2 conjugating enzymes in the process of RING- and RBR-type E3 ligases-mediated ubiquitination, whereas they are governed by E3 ligases in the process of HECT-type E3 ligases-mediated ubiquitination, suggesting the importance of E2 conjugating enzymes, E3 ligases, and combinations of these two enzymes [30,66,78].

2.3. E4 Ubiuqitin Ligases and Ubiquitin Chain Elongation

E4 ubiquitin ligases play roles in highly processive chain elongation but not in the initial steps of ubiquitination [79,80,81,82,83]. E4 ligases not only play complementary roles to E3 ligases but also facilitate the function of E3 ligases under certain circumstances, which extends the length of polyubiquitin chains [80,84]. Ubiquitination factor E4B (UBE4B) and its isoform UBE4A are a U-box-containing RING family of ubiquitin ligases [85]. UBE4B and UBE4A are involved in the degradation of substrates through the ubiquitin fusion degradation pathway (UFD) in a similar manner to the yeast orthologue UFD2 [79,86,87]. UBE4B possesses a conserved U-box catalytic domain of about 70 amino acids, which mediates the association of UBE4B with the ubiquitin-charged E2 enzyme to facilitate attachment of a polyubiquitin chain on a selected target of UBE4B, based on its E3 and E4 ligase activity [88,89,90,91]. Additionally, UBE2A interacts with ubiquitin moieties of preformed conjugates and facilitates the elongation of the polyubiquitin chain [79,86,87].
p300 and CREB-binding protein (CBP) possess E4 ligase activity in their N-terminal regions [92,93,94]. The N-termini of p300 and CBP are cysteine-histidine-rich regions, whereas no signatures for RING, HECT, U-box, or plant homeodomain (PHD) domains can be found.
The C-terminus of the Hsc70-interacting protein (CHIP) possesses not only one U-box domain but also three tandem tetratricopeptide repeat (TPR) motifs that bind to Hsc70 and Hsc90 [95]. Similarly to yeast UFD2, CHIP acts as an E4 ligase for ubiquitin chain elongation in cooperative manner with an additional E3 ligase. CHIP is involved in ubiquitination of the PARKIN-associated endothelin receptor-like (Pael) receptor in collaboration with PARKIN [96]. Even though PARKIN can ubiquitinate the Pael receptor in vitro, only a combination of PARKIN and CHIP can sufficiently accomplish its polyubiquitination.
Yin-Yang 1 (YY1) is an E4 ligase and facilitates p53 polyubiquitination by enhancing the interaction between p53 and MDM2 [97]. Further, YY1 interacts with p300, thereby facilitating p300-mediated polyubiquitination of p53 [98].

2.4. Ubiquitin Code and Its Biological Significance

Ubiquitination determines the fate of substrate proteins. Functions of ubiquitination are dictated by their distinct structural topologies that are recognized by specific ubiquitin adaptors. Monoubiquitination, modification by a single ubiquitin moiety, is the most abundant modification and controls protein recognition, complex formation, allosteric regulation, endocytosis and even proteasomal degradation [99,100]. Further, ubiquitin possesses eight ubiquitination sites, seven internal lysine (K) residues (including K6, K11, K27, K29, K33, K48, and K63), and a primary amine at the N-terminal methionine (M1), all of which depend on the ubiquitination site and the length of the chains and participate in distinct homotypic polyubiquitin chain formation and chain topologies. The architecture of polyubiquitin chains is determined by the lysine acceptor site to be conjugated, the number of modified sites and the length of the added ubiquitin molecules, and combinations of these parameters have been referred to as ubiquitin code [2]. Recent discoveries have revealed that heterotypic polyubiquitin chains can be generated by the formation of mixed ubiquitin chains or branched ubiquitin chains in which ubiquitin can be modified on multiple sites [44,101,102,103,104,105]. Intriguingly, the dynamic and complex ubiquitin architecture, which includes monoubiquitination, multiple monoubiquitination, and eight different modes of homotypic and numerous types of heterotypic and branched polyubiquitin linkages, generates sophisticated and versatile ubiquitin code, thereby governing the fate of substrate proteins and providing additional regulatory nodes for cellular homeostasis [2,9,44,106].
Although the functional significance of K48- and K63-linked polyubiquitination is largely known, the biological significance of K6-, K11-, K27-, K29-, K33-, and M1-linked polyubiquitination is still far from being fully understood. Four ubiquitin molecules linked via K48 are enough to target substrate proteins to proteasomal degradation and K63-linked polyubiquitin chains serve as signals for modulation of protein trafficking, DNA repair and signal transduction [107,108,109,110,111]. Moreover, protein aggregates or dysfunctional organelles are eliminated via the autophagy–lysosome pathway, which is mediated by K6- or K63-linked chains. Additionally, K11-linked polyubiquitin chains serve as a signal for proteasome to recognize substrate, facilitating ERAD [104,112]. However, K27-, K29-, and K33-linked polyubiquitinations have begun to emerge and little is known about the biochemical information [113,114].
Interestingly, heterotypic conjugates potentially lead to combinatorial control of signaling. K11/K48-branched chains function as proteasomal priority signals and target aggregation-prone and cytotoxic proteins to proteasomal degradation [112,115]. Additionally, capping of K63-linked conjugates with M1 linkages results in the formation of M1/K63 heterotypic ubiquitin chains, which facilitates NF-κB activation [116,117,118].

3. “Protein Quality Control” and “Protein Quantity Control” at the ER: ERAD

Despite an elaborate network within the ER for co- and post-translational folding and maturation of polypeptides, the process of protein folding and maturation is inherently error-prone and therefore misfolded proteins must be eliminated [119,120,121,122,123]. ERAD was discovered nearly 30 years ago. ERAD is highly conserved protein quality control machinery of the ER used to eliminate misfolded or unassembled proteins via the cytosolic ubiquitin-proteasome system (UPS) (Figure 2) [8,122,124,125,126]. Additionally, growing evidence indicates that ERAD also targets and fine-tunes correctly folded proteins, involving metabolically controlled enzymes, plasma membrane transporters and transcription factors, suggesting the role of ERAD in protein quantity control [127]. ERAD is a sophisticated and multistep process composed of substrate recognition, energy-dependent retrotranslocation from the ER, ubiquitination, and proteasomal degradation of substrates [128,129,130,131]. Intriguingly, deletion of key components of ERAD results in embryonic lethality in mice, pinpointing the physiological significance of ERAD in development [132,133]. Accordingly, any failure to interrupt the degradation of a specific protein in a timely manner could result in malfunction of essential cellular processes, which leads to the development of various diseases, including cancer, neurodegeneration, and metabolic diseases.

3.1. Substrate Recognition

Substrate recognition, the commitment step for ERAD, is accomplished by molecular chaperones and chaperone-like lectins [134]. During folding of glycoprotein, high-mannose “core” glycan possessing Glc3Man9GlcNAc2 (Glc: glucose, Man: mannose, GlcNAc: N-acetylglucosamine) is first appended to consensus asparagine residues within canonical N-glycosylation sites (NxS/T), which plays a leading part in monitoring conformational maturation, directing correctly folded proteins to the ER exit and directing misfolded proteins to ERAD [135]. The glycan is subsequently trimmed by lectin-type chaperone, calnexin, or calreticulin. Calnexin or calreticulin interacts with Glc1Man9GlcNAc2 generated by cleavage of three terminal glucose residues from the core glycans, thereby monitoring the folding of glycoproteins. To specifically generate a glycan signal for ERAD, terminally misfolded proteins must escape from the calnexin or calreticulin, which is conducted by mannosidases—including ER mannosidase I (ERManI) [136,137], ER degradation-enhancing α-mannosidase like protein 1 (EDEM1), EDEM2, EDEM3 [138,139,140,141], or Golgi-resident mannosidase α class 1C member 1 (Man1C1) [142]—that progressively eliminate terminal mannose residues from core glycans and generate an exposed α-1,6-mannose, thereby leading to association with mannose-specific lectins for ERAD [143,144]. These misfolded glycoproteins are then captured, recognized by osteosarcoma 9 (OS-9) and XTP3-transactivated gene B precursor (XTP3-B), and subsequently recruited to the protein penetration channel, retrotranslocon, for ERAD [145,146,147]. Interestingly, non-glycosylated proteins can be degraded via ERAD. Non-glycosylated proteins bind to non-lectin chaperone binding immunoglobulin protein (BiP) and are subsequently targeted to ERAD [148,149]. Additionally, EDEM1 or protein disulfide isomerase (PDI) is involved in the degradation of non-glycosylated proteins via ERAD [150,151].

3.2. Retrotranslocation

Energy-dependent retrotranslocation of substrates from the ER back into the cytoplasm is essential for ERAD, since the ER does not possess the components of UPS such as the E1 activating enzyme, E2 conjugating enzyme, and the proteasome [152]. Importantly, the processes of retrotranslocation, ubiquitination, and proteasomal degradation of substrates during ERAD should be tightly coupled, thereby preventing protein aggregation in an aqueous environment. p97/valosin-containing protein (VCP) is a member of the type II AAA+ protein family of ATPases and is composed of two AAA domains, D1 and D2, which are assembled in a head-to-tail manner, an N-terminal domain that functions in substrate recognition, and a C-terminal domain that associates with a large variety of adaptors [153,154,155,156]. p97/VCP plays a key role in the retrotranslocation process of nearly all ERAD substrates via coupling of ATP hydrolysis to the unfolding of ERAD substrates with the assistance of cofactors recruited through p97/VCP-binding domains, including SHP (BS1, binding segment 1) box, p97/VCP-binding region (VBR), and p97/VCP-interacting motif (VIM) [154]. Most of the p97/VCP cofactors also contain ubiquitin-binding domains (UBDs), thereby leading to the direct association with ubiquitinated substrates. Interestingly, p97/VCP in cooperation with nuclear protein localization protein 4 (Npl4) and ubiquitin fusion degradation protein 1 (Ufd1) can generate a driving force for the retrotranslocation of substrates for ERAD [157]. To summarize, the ERAD substrate is slightly exposed to the ER surface through the retrotranslocon, subject to E3 ubiquitin ligase-mediated polyubiquitination and subsequently further retrotranslocated by the p97/Npl4/Ufd1 complex, highlighting that the two processes of substrate retrotranslocation and polyubiquitination by ERAD E3 ubiquitin ligases should be tightly coupled and that p97/VCP functions as a scaffold to recruit the various factors involved in ERAD and to regulate ubiquitination at sites of retrotranslocation [129,158,159,160,161,162].

3.3. Ubiquitination and E3 Ubiquitin Ligases

Protein quality control E3 ubiquitin ligases involved in ERAD determine the specificity, plasticity, and efficiency of ERAD and subsequently fine-tune proteostasis. As unrestrained degradation would destroy cellular homeostasis, it is crucial that E3 ligases ensure only the specificity required to eliminate misfolded proteins, while leaving their correctly folded counteracts intact. Moreover, E3 ligases usually need to target similarly misfolded and potentially cytotoxic factors, suggesting the necessity of a delicate balance between specificity and plasticity. Approximately 40 ERAD E3 ubiquitin ligases have been identified in mammals and three in yeast to date [163]. However, most of the ERAD E3 ubiquitin ligases are poorly characterized and only a few substrates have been identified (Table 1).
Several ERAD E3 ubiquitin ligases, including glycoprotein 78 (gp78), also known as autocrine motility factor receptor (AMFR)), hydroxymethylglutaryl reductase degradation protein 1 (HRD1), membrane-associated RING C3HC4 finger 6 (MARCHF6), and RING finger protein 5 (RNF5), are transmembrane proteins [129,164,165,166,167,168,169,170], while another set of E3 ubiquitin ligases, including CHIP, SMAD ubiquitination regulatory factor 1 (Smurf1), neuregulin receptor degradation pathway protein 1 (Nrdp1)/fetal lever ring finger (FLRF), PARKIN, and the Skp, Cullin and F-box containing complex (SCF) with the F-box proteins Fbx2, Fbx6, or beta-transducin repeat-containing proteins 1 and 2 (β-TrCP1/2), are localized in the cytoplasm and involved in ERAD [96,171,172,173,174,175,176]. ERAD E3 ubiquitin ligases may accomplish polyubiquitination of ERAD substrates not only in cooperation with other E3 ubiquitin ligase-mediated multiple monoubiquitinations or E4 ubiquitin ligase-mediated extensions after initial monoubiquitination but also by the aid of sequential rounds of ubiquitination and deubiquitination, raising the possibility that various strategies have evolved for the optimal efficiency of ERAD [15,177,178,179,180].
Interestingly, for the degradation of ERAD substrates, mixed K48/K11-linked polyubiquitin chains are appended, facilitating the association with downstream components of ERAD machinery [181]. Further, DUBs fine-tune the timing of ERAD by trimming polyubiquitin chains on ERAD substrates [182]. Notably, ubiquitin can be targeted to thiol or hydroxyl side chains in cysteine, serine, and threonine residues on the substrate proteins, suggesting that the diversity in amino acids modified by ERAD-specific E2 conjugating enzyme/E3 ligase pairs provides the flexibility to ubiquitinate ERAD substrates [37,183].

3.3.1. HRD1

HRD1-mediated ERAD is the most conserved and best characterized branch of mammalian ERAD and it is indispensable for versatile homeostatic processes, including energy metabolism, food intake, systemic water balance, and immune cell development [184,185,186,187,188,189,190,191,192]. HRD1 is involved in the ubiquitination of not only glycosylated but also non-glycosylated ERAD substrates. HRD1 commonly targets substrates for degradation in a suppressor/enhancer of Lin12-like (SEL1L)-dependent manner. SEL1L is crucial not only for the transfer of substrates to HRD1 but also for the stabilization of HRD1, indicating the key role of SEL1L in the recruitment, retrotranslocation, and ubiquitination of ERAD substrates [146,193,194,195,196]. SEL1L also recruits recognition factors, including EDEMs, ERdj5, OS-9, PDI, and XTP3-B, to the retrotranslocon for ERAD substrates [197]. Further, SEL1L functions as a scaffold to generate a complex with ancient ubiquitous protein 1 (AUP1), degradation in endoplasmic reticulum protein (Derlin)-1, Derlin-2, ubiquitin regulatory X (UBX) domain containing protein 8 (UBXD8), and VCP-interacting membrane protein (VIMP) [198,199,200,201,202], which recruits the p97/VCP and leads to substrate retrotranslocation. Derlin-1, Derlin-2, or Derlin-3 interact with ERAD substrates and target ERAD substrates to HRD1 and p97/VCP, indicating that Derlins work together as a part of the retrotranslocon channel [203,204,205], although the role of Derlins needs further characterization in mammalian systems. Interestingly, auto-ubiquitination of HRD1 facilitates retrotranslocation of misfolded protein substrates, suggesting that HRD1 might be a protein-conducting channel opened by ubiquitin moiety [206].
B cell development-specific pre-B cell receptor (pre-BCR), B-lymphocyte-induced maturation protein 1 (BLIMP1), Fas, inositol-requiring enzyme 1 (IRE1), NF-E2-related factor 2 (NRF2), p53, peroxisome proliferator activated receptor γ coactivator-1 β (PGC1β), and pro-opiomelanocortin (POMC) have been identified as endogenous substrates for SEL1L/HRD1-mediated ERAD [186,188,189,196,207,208,209]. Notably, HRD1-mediated ERAD functions to ensure optimal levels of substrate proteins not only qualitatively but also quantitatively. The HRD1-mediated degradation of IRE1 constitutively fine-tunes IRE1 activity under normal physiological conditions, thereby limiting the amplitude and duration of IRE1 activity and maintaining cellular homeostasis [189]. Interestingly, deletion of either SEL1L or HRD1 in various tissues and cell types leads to a profound increase in the level of IRE1 protein, whereas its mRNA levels are unchanged [189]. However, it remains elusive as to whether IRE1 is subjected to ERAD-mediated quality control to eliminate misfolded IRE1 or to ERAD-mediated quantity control to ensure adequate the protein level of IRE1. Additionally, HRD1-mediated constitutive degradation of pre-BCR in developing B cells regulates the abundance of surface pre-BCR, thereby restraining pre-BCR-mediated signaling during the transition from the large to small pre-B cell stage [188,210]. Moreover, HRD1-mediated constitutive degradation of misfolded pro-arginine vasopressin (AVP) maintains systemic water balance [186]. Mice depleted of SEL1L in AVP neurons develop polyuria and polydipsia, suggesting the pathophysiological importance of the constitutive ERAD in modulating the maturation and abundance of specific ER proteins and fine-tuning their activities. Recently, SEL1L/HRD1-mediated ERAD has been reported to target transforming growth factor beta (TGF-β) receptor I for proteasomal degradation, resulting in the control of β-cell identity via TGF-β signaling [211].
A variety of studies have suggested HRD1 as a critical immune regulator. HRD1-mediated ERAD eliminates major histocompatibility complex (MHC) class I heavy chains that fail to acquire their native conformation in complex with β2-microglobulin and peptide, suggesting the role of HRD1 as a positive regulator for the peptide antigen presentation by MHC I molecules [212,213,214]. Additionally, HRD1 directly recognizes, ubiquitinates, and mediates proteasomal degradation of BLIMP1, a transcription co-repressor that inhibits the transcriptional activity MHC II transcription activator (CIITA) [215,216]. Further, HRD1 attaches ubiquitin to the serine side chains on T-cell receptor α (TCRα) [183] and to serine, threonine, and lysine residues on immunoglobulin nonsecreted Ig light chain (NS1LC) [37], suggesting the importance of HRD1 for the flexibility of ubiquitination during ERAD to cope with cellular stresses.
HRD1 is involved in anti-apoptosis [217]. HRD1-mediated degradation of IRE1 via ERAD prevents ER stress-induced cell death [189,218]. Additionally, HRD1 facilitates cell cycle progression and anti-apoptosis by mediating p53 ubiquitination [208]. HRD1 has also been reported to mediate ubiquitination and subsequent degradation of Fas, thereby resulting in the prevention of B-cell apoptosis [209].
Although the underlying mechanisms by which HRD1 recognizes nuclear substrates remain elusive, HRD1 may directly mediate the degradation of nuclear transcription factors, including BLIMP1, NRF2, and PGC1β [192,207,215,219], revealing intriguing regulatory cascades from protein degradation at the ER membrane to gene transcription in the nucleus.

3.3.2. RNF45/gp78/AMFR

Gp78 is another major ERAD E3 ubiquitin ligase in mammals [166]. Five transmembrane domains reside at the N-terminus of gp78 and the RING finger domain is localized at the C-terminal cytoplasmic tail [220,221]. Additionally, the E2 ubiquitin-conjugating enzyme, UBE2G2 binding region (G2BR) of gp78 associates with UBE2G2, which leads to conformational alterations in UBE2G2 and an increase in the affinity of UBE2G2 for gp78 [222,223]. In a similar manner to HRD1, gp78 recruits UBXD2 or UBXD8, thereby functioning as a bridge to connect p97/VCP to the ER membrane. Alternatively, gp78 directly interacts with p97/VCP via its VIM domain, which is localized at the C-terminus [160]. Interestingly, USP13 associates with gp78 to attenuate polyubiquitination of ubiquitin-like protein 4A (UBL4A), a part of the Bag6 complex that helps chaperone retrotranslocated ERAD substrates to the proteasome, thereby protecting the Bag6 complex from proteasomal degradation and maintaining an adequate ERAD pathway, highlighting the importance of the associated DUB activity during ERAD [224].
Gp78 plays an essential role in fine-tuning the levels of proteins quantitatively in response to specific signals. Sterol accumulation alters transmembrane regions of 3-hydroxy-3-methyl-glutaryl-coenzyme A (CoA) reductase (HMGCR), which is an ER membrane-anchored protein, and catalyzes the rate-limiting step of sterol biosynthesis, in which HMG-CoA is reduced to mevalonate [225,226] and facilitates the binding of HMGCR to ER membrane-embedded insulin-induced genes 1/2 (INSIG-1/2) proteins, thereby resulting in the association of HMGCR with gp78 and subsequent ubiquitination and proteasomal degradation of HMGCR [221]. Interestingly, viral or microbial DNA promotes the association of gp78 with INSIG-1 and stimulator of interferon genes (STING) and facilitates K27-linked polyubiquitination of STING, which results in the association of STING with TANK-binding kinase 1 (TBK1) and the translocation of STING to perinuclear microsome, thereby leading to innate immune response [227].
It has been suggested that Gp78 has neurobiological functions. Gp78 mediates ubiquitination and subsequent degradation of disease-associated aggregative proteins, such as superoxide dismutase-1 (SOD1), ataxin-3, and the mutant form of huntingtin [228,229]. Further, gp78 specifically associates with the C-terminal region of unglycosylated prion proteins that cause bovine spongiform encephalopathy and Creutzfeltd–Jacob diseases and facilitates ubiquitination and subsequent proteasomal degradation of prion proteins [230].
Interestingly, gp78 also functions as an E4 ubiquitin ligase in a cooperative manner with other ERAD E3 ubiquitin ligases [177]. For example, gp78 serves as an E4 ubiquitin ligase for RNF5-mediated ubiquitination of cystic fibrosis transmembrane conductance regulator (CFTR)∆F508 in ERAD [168,177].

3.3.3. MARCHF6/TEB4/RNF176

MARCHF6, akin to Doa10 in yeast, is a large multi-pass E3 ubiquitin ligase embedded in the membranes of the ER and plays a central role in sterol and lipid metabolism [167,231,232,233]. MARCHF6 is characterized by two functional domains: the catalytic N-terminal RING domain and the C-terminal element (CTE) that is involved in its self-regulation [167,234]. MARCHF6 cooperates with the E2 ubiquitin-conjugating enzyme, UBE2G1, and mediates ubiquitination and proteasomal degradation of ERAD substrates with the aid of HSC70, DNAJB12, HDJ2, and UBXD8 [235,236,237,238,239].
HMGCR and squalene epoxidase (SQLE, also known as squalene monooxygenase) are controlled by MARCHF6-mediated ERAD, although SQLE is widely considered as the canonical substrate for MARCHF6 [240,241]. SQLE is subjected to ERAD when additional cholesterol synthesis is unnecessary. MARCHF6 and the E2 conjugating enzyme Ube2J2 facilitate degradation of SQLE via ERAD [242]. Sterol-dependent ERAD is mediated by an N-terminal SQLE degron containing an amphipathic helix, which allows sterol-dependent recognition of SQLE by MARCHF6 [242,243,244]. Additionally, Perilipin-2 (PLIN2), mutant versions of Niemann–Pick disease type C intracellular cholesterol transporter 1 (NPC1), and bile salt export pump (BSEP) are identified as target substrates for MARCHF6 [245,246,247]. Interestingly, MARCHF6 preferentially targets N-terminally acetylated residues in PLIN2, suggesting that MARCHF6 could degrade substrates via the Ac/N-end rule pathway, in which N-terminally acetylated residues lead to destabilization of substrates [245,248]. Further, MARCHF6 indirectly affects sterol regulatory element binding proteins (SREBP) and liver X receptor (LXR), both of which are master transcription factors for the control of cholesterol levels [249].
MARCHF6 itself has a short half-life due to its autoubiquitination. Interestingly, USP19, an ER-resident DUB, and cholesterol enhance the stability of MARCHF6 [250,251].

3.3.4. RNF5/RMA1

RNF5 is a membrane-anchored E3 ubiquitin ligase localized not only in the ER but also in mitochondrial membranes [168,169,170]. RNF5 functions sequentially with CHIP to degrade misfolded CFTR [168]. RNF5 recognizes folding defects in CFTR∆F508 coincident with translation, while CHIP acts on CFTR∆F508 post-translationally. Further, RNF5-mediated ubiquitination-dependent relocalization of the proteasome adaptor protein JNK-associated membrane protein (JAMP) contributes to the recruitment of the proteasomes in the vicinity of the ER [252].

3.4. Delivery and Degradation

p97/VCP functions as a bridge to connect retrotranslocated substrates to cytoplasmic cofactors for further processing and subsequent degradation of ERAD substrates, suggesting that p97/VCP is closely associated with the proteasome-mediated degradation of substrates [154]. Cytoplasmic N-glycanase 1 (NGly1), a deglycosylating enzyme, is recruited to retrotranslocon complexes via direct association with p97/VCP and cleaves N-linked glycans from retrotranslcated ERAD substrates [253,254]. DUBs, including Ataxin-3, ovarian tumor family of deubiquitinating enzyme 2 (OTUD2, also known as YOD1), USP13, and valosin-containing protein p97/p47 complex-interacting protein p135 (VCIP135) interact with p97/VCP either in a direct or indirect manner and subsequently deubiquitinate ERAD substrates [255,256,257]. OTUD2 in cooperation with p97/VCP may facilitate the efficient retrotranslocation of ubiquitinated ERAD substrates. Notably, the impairment of p97/VCP-associated deubiquitination or expression of dominant-negative OTUD2 impedes retrotranslocation and subsequent degradation of ERAD substrates [255], indicating that the sequential rounds of ubiquitination and deubiquitination are indispensable for an efficient ERAD process. USP13 interacts with p97/VCP, Ufd1, Npl4, and UBXD8 [256]. Interestingly, the depletion of USP13 leads to the accumulation of ERAD substrates. The role of ataxin-3 in ERAD remains to be elucidated. Ataxin-3 has been suggested to act after retrotranslocation of ERAD substrates [257]. Alternatively, ataxin-3 may cleave off ubiquitin from polyubiquitinated ERAD substrates, which increases the half-life of protein and provides time for folding of ERAD substrates [258].
Retrotranslocated ERAD substrates should be rapidly eliminated to prevent the formation of protein aggregates in the cytoplasm. Bag6, ubiquitin-like protein 4A (Ubl4A), and the transmembrane domain recognition complex 35 (Trc35) co-chaperone small glutamine rich TPR-containing protein α (SGTA) form a chaperone complex, which prevents the formation of protein aggregates [259]. Bag6 oligomerizes through its proline-rich domain. Interestingly, the proline-rich domain is sufficient for the association of Bag6 with the hydrophobic segments of misfolded proteins, thereby leading to the maintenance of misfolded proteins in a soluble state [260]. Moreover, the holdase activity of Bag6 is indispensable for maintaining retrotranslocated substrates in a competent state for proteasome-mediated degradation [261]. Interestingly, Bag6 also interacts with the proteasome and adaptor proteins of the proteasome, resulting in the transfer of retrotranslocated substrates to proteasome for ERAD.

4. “Protein Quality Control” at the ER: ER-Phagy

While most of the misfolded proteins in the ER are efficiently eliminated via ERAD, the pore size of the retrotranslocon may preclude misfolded oligomeric or aggregated proteins from being exported from the ER, which leads to the development of another mechanism, selective autophagic degradation of the ER, referred to as ER-phagy, which eliminates these toxic substrates, thereby maintaining proteostasis. The term ER-phagy was first used upon the discovery of ER whorls, which occur after unfolded protein response (UPR)-associated lipid synthesis in yeast [262,263]. ER-phagy is then initiated to utilize the core autophagy machinery and eliminate ERAD-resistant misfolded proteins [17,264].
The identification of versatile ER-phagy receptors responsible for substrate recognition in mammals, including family with sequence similarity 134 member B (FAM134B), reticulon domain containing family member 3 long (RTN3L), secretory 62 homologue (SEC62), cell-cycle progression gene 1 (CCPG1), atlastin 3 (ATL3), and testis-expressed protein 264 (TEX264), has advanced our understanding of ER-phagy [265,266,267,268,269,270,271]. By definition, the ER-phagy receptors can directly mediate the recognition of the ER membrane by phagophores or lysosomes. The ER-phagy receptors are either directly ER membrane-anchored, through insertion of a part of the polypeptide into the ER membrane from the cytosolic side, or bona fide transmembrane proteins [272,273]. ATL3, FAB134B, and RTN3L possess an intramembrane region (IM) lacking an ER luminal domain. In contrast, CCPG1, SEC62, and TEX264 are transmembrane proteins with a cytosolic, ER membrane, and ER luminal domains. The ER tightly binds to the outer and inner membranes of autophagosomes to form ER–autophagosome contacts, through which the ER can supply lipids and other components for autophagosome expansion [274,275,276]. ER-phagy can be classified into three distinct pathways on the basis of topology (Figure 3) [277]. During macro-ER-phagy, autophagosomes enclose fragments of the ER and fuse with lysosomes to eliminate the internal material harboring the ER [278,279]. During micro-ER-phagy, lysosomal membranes invaginate and pinch off parts of the ER into the lysosomal lumen [272,280]. Lastly, during vesicular delivery, lysosomes can directly fuse with ER-derived vesicles for degradation [17,281].

4.1. Macro-ER-Phagy

4.1.1. A Conceptual Overview of Macro-ER-Phagy

Since the discovery of sequestration of the ER by autophagosomes over 50 years ago [282], it has remained uncertain whether this process involves selective or random sequestration of the ER. However, the identification of ER-phagy receptors indicates that ER-phagy involves the selective recognition of the ER to ensure selective autophagy. Importantly, ER-phagy is a type of selective autophagy for the removal of misfolded protein aggregates to maintain proteostasis. To date, six ER-phagy receptors, FAM134B, RTL3L, CCPG1, SEC62, TEX264, and ATL3, have been identified in mammals [265,266,267,268,269,270]. These ER-phagy receptors are ER membrane-anchored proteins and possess at least one microtubule-associated protein 1A/1B-light chain 3 (LC3)-interacting motif (LIR), which leads to the association with autophagosomal LC3 proteins. Additionally, all ER-phagy receptors except for ATL3 possess long intrinsically disordered regions (IDRs) between the LIR and the ER transmembrane domain. IDR functions as a bridge connecting the ER and autophagosomal membranes [277].
Several functional characteristics of the ER-phagy receptors have been reported [277]. The ER-phagy receptors can mediate the degradation of different subdomains of the ER, including nuclear membranes, the sheet ER, the tubular ER, and the ER exit site (ERES). Additionally, individual ER-phagy receptors can facilitate specific stimuli-triggered ER-phagy. ER-phagy is induced upon ER stress or starvation conditions. Further, some ER-phagy receptors contain ER-fragmentation activity, which cleaves a portion of the ER network before or during sequestration into autophagosomes. Reticulon homology domains (RHDs) consist of two short hairpin transmembrane domains, which can generate membrane curvature by partially inserting into the membrane, thereby leading to an increase in the surface area of the cytosolic leaflet [283]. Moreover, association of ER-phagy receptors with membrane-tethered LC3 is indispensable not only for linkage of the ER to the phagophore but also for concomitant membrane reshaping to pack into autophagosomes [265]. Lastly, ER-phagy receptors are distributed differently in a tissue-specific manner. FAM134B, TEX264, SEC62, and ATL3 show ubiquitous expression, whereas CCPG1 shows predominant expression in the pancreas, kidney, and liver [268,271,284].

4.1.2. Macro-ER-Phagy and Protein Quality Control

FAM134B, also known as reticulophagy regulator 1 (RETREG1), is classified as an IM ER-resident protein characterized by the presence of LIR and RHD [265]. RHD mediates post-translational insertion of FAB134B into the lipid bilayer. N- and C-termini of FAM134B are cytosolic and LIR is at the C-terminus. FAM134B localizes at the edge of ER sheets via RHDs and promotes membrane curvature during autophagosome formation, thereby mediating starvation-induced ER-phagy [265]. In addition to RHDs, two amphipathic helices of FAM134B are involved in membrane curvature [285,286]. The LIR motif within FAM134B recruits LC3 and gamma-aminobutyric acid receptor-associated protein (GABARAP) into the growing phagophores. Intriguingly, a group of misfolded proteins has been identified as FAM134B cargoes. FAM134B-mediated ER-phagy eliminates ERAD-resistant protein aggregates such as Niemman–Pick type C disease protein (NPC), mutant forms of type I procollagen (PC1), and the collagen chaperone HSP47 [246,281,287,288,289]. The folding process of PC is complicated and approximately 15% of newly synthesized procollagens is degraded [290]. Calnexin not only recognizes luminal misfolded PCs but also interacts with FAM13B, thereby leading to the degradation of misfolded PCs via ER-phagy [288]. Given that FAM134B does not have an ER luminal domain, calnexin is essential for the selective interaction of FAM134B with PCs. Further, the TM domain of calnexin associates with the RHD of FAB134B, indicating the role of calnexin as a bridge connecting ER luminal PCs to FAM134B [288].
RTN3L, a long splice isoform of RTN3, is an ER tubules resident protein. RTN3L possesses C-terminal RHD for anchorage of RTN3L to the ER from the cytosolic face of the membrane. RTN3L possesses an extended N-terminus containing six distinct LIR motifs spaced at uneven intervals and oligomerizes to engage and cluster LC3, thereby leading to the formation of autophagosomes at the ER tubules [267,291]. RTN3L is upregulated during starvation and facilitates starvation-induced ER-phagy. RTN3L is not involved in either LC3 lipidation or p62 degradation, thus indicating its selective role in ER degradation rather than in general autophagy. Interestingly, RTN3L is involved in the degradation of misfolded proinsulin protein aggregates via ER-phagy and the enhanced expression of RTN3L is involved in the relief of ER aggregates and in the export of wild-type proinsulin [292].
CCPG1 is a type II single-pass transmembrane protein and possesses both LIR- and focal adhesion kinase family interacting protein of 200 kDa (FIP200)-interacting regions, which associate with LC3 and FIP200, respectively [268]. In contrast to FAM134B and RTN3L, CCPG1 possesses an extensive C-terminal ER luminal region of undefined structure and an IDR at the N-terminus. Although CCPG1 does not have intrinsic ER fragmentation activity, it is likely that accumulation of IDR-containing CCPG1 induces molecular crowding at the site facing autophagosomes, resulting in the membrane curvature and scission [293,294]. Even though it is not clear whether CCPG1 selectively recognizes aggregated or misfolded proteins, upregulation of CCPG1 upon ER stress is known to be involved in the removal of the portions of the peripheral ER, resulting in the reduction of protein aggregation. Particularly, CCPG1 deficiency in the ER leads to numerous luminal protein inclusions and subsequent elevation of UPR [268].
TEX264 is a single-pass transmembrane protein and possesses a negligible N-terminal luminal region of approximately five amino acids and a C-terminal cytosolic LIR [269,271]. TEX264 is involved in approximately 50% of starvation-induced autophagic flux from the ER.
In addition to the transmembrane ER-phagy receptors, p62 could also mediate ER-phagy. In mouse liver, xenobiotics-induced excess ER is eliminated by p62-mediated autophagy [295]. IRE1 associates with p62 and cytosolic ubiquitin-binding receptors, optineurin and neighbor of BRCA1 gene 1 (NBR1), suggesting the role of ER-phagy in the sequestration of active IRE1-enriched ER subdomains to terminate UPR. Further, the association of p62 with ubiquitinated TRIM13 facilitates ER-phagy [296].

4.2. Micro-ER-Phagy

Micro-ER-phagy was first discovered as piecemeal micronucleophagy in yeast [297]. Additionally, ER-targeted microautophagy is also observed in yeast. Dithiothreitol-induced ER stress mediates formation of ER whorls, which are selectively delivered to the vacuole by microautophagy [263,280]. Microautophagy in yeast requires the endosomal sorting complexes required for transport (ESCRT) machinery for scission of vacuolar membrane, whereas this process does not require core autophagy factors and macro-ER-phagy receptors [298].
In mammals, micro-ER-phagy is involved during recovery from cyclopiazonic acid-induced ER stress, in which ER-derived vesicles are directly engulfed by endolysosomes [299]. Compared to micro-ER-phagy in yeast, micro-ER-phagy in mammals requires SEC62 and the LC3 lipidation system components, including autophagy-related protein 4B (ATG4B), ATG7, and AYG16L1. However, in the process of SEC62-dependent micro-ER-phagy, LC3 is present on SEC62-decorated ER rather than on lysosomes. Therefore, further studies are warranted to investigate the incorporation of selectivity during micro-ER-phagy. Additionally, mammalian micro-ER-phagy requires CHMP4B, an ESCRT-III component. Importantly, a subset of proteins can be eliminated via micro-ER-phagy. A disease-causing mutant form of PCs (G610C) accumulates at the ERES, a representative of very specialized ER zones for the transport of cargo proteins from the ER to the Golgi apparatus [300] and in which ubiquitin, p62, LC3, ATG14, ATG4, and COPII also accumulate, and leaves the ER through the ERES via COPII-coated vesicles, thereby resulting in the direct engulfment of the mutant form of PCs by lysosomes via micro-ER-phagy [301]. Notably, this micro-ER-phagy occurs at an early stage of PC trafficking, thereby ensuring that only adequately folded PCs enter the secretory pathway.

4.3. Vesicular Delivery

Vesicular delivery, a pathway for lysosomal degradation of ER-derived single-membrane vesicles, has been recently proposed [17,281]. Calnexin segregates α1-antitrypsin Z (ATZ) polymers into subdomains of the ER and is involved in the formation of a complex with LC3 and FAM134B, resulting in the generation of local membrane curvature and scission and subsequent formation of ER-derived vesicles. Further, ER-resident SNARE syntaxin 17 (STX17) and lysosomal SNARE vesicle-associated membrane protein 8 (VAMP8) mediate fusion of the ER-derived vesicles with endolysosomes. Interaction of FAM134B with LC3 at vesicle-lysosome contact sites drives fusion, indicating the role of LC3 in the recognition of vesicular receptor. The vesicular delivery does not require autophagy initiation complex components, including ATG13, ATG9, FIP200, and unc-51 like autophagy activating kinase (ULK1/2), but requires LC3-conjugation machinery, which could be important for the docking of ER-derived vesicles into endolysosomes [302]. However, ATZ has also been reported to be degraded by lysosomes in an ATG conjugation-dependent manner, suggesting the requirement in some ATGs for this novel pathway [303,304,305].

5. Brief Linking of “Protein Quality Control” at the ER to Neurodegenerations

In the last decade, multiple studies have implicated the pivotal roles of the ER in the progression of neurodegenerative diseases, including Alzheimer’s disease (AD), amyotrophic lateral sclerosis (ALS), Parkinson’s disease (PD), and Gaucher disease (GD) [306,307,308,309,310]. Although it is still controversial whether ER stress and the subsequent protein quality control at the ER are closely linked to neurodegenerative diseases, misfolded proteins and their aggregates are implicated in the pathogenesis of neurodegenerative diseases.
PD is a progressive neurodegenerative disease and is characterized by the aggregates accumulation of misfolded α-synuclein (α-SYN) fibrils into proteinaceous inclusions in Lewy bodies (LB) or Lewy neurites [311]. Toxicity resulting from the overexpression of α-SYN is associated with ER stress and activation of the UPR [312,313,314,315,316,317], which subsequently facilitates ER stress-induced dopaminergic neuronal death [312,318]. Interestingly, α-SYN aggregates associate with BiP, thereby leading to the activation of the UPR [314,319]. Additionally, α-SYN has been reported to be abundant in ER/microsome fractions of human and mice PD brain tissues, resulting in the ER stress-associated accumulation of polyubiquitin chains and the activation of caspase-12 [315]. Moreover, accumulation of aggregates of α-SYN inhibits the ER–Golgi trafficking by directly binding to ras-associated binding 1 (RAB1) GTPase, which leads to ER stress [313,320,321]. Intriguingly, accumulation of aggregates of α-SYN destroys ER Ca2+ homeostasis, thereby leading to the induction of ER stress. α-SYN aggregates activate sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) in neurons, which results in alteration of ROS production and calcium metabolism and cell death [322,323].
PARKIN, an E3 ubiquitin ligase, is involved in a large variety of cellular processes related to PD [324,325]. Interestingly, ER stress-induced upregulation of PARKIN plays a neuroprotective role against ER stress [325]. Additionally, the Pael receptor, which is involved in ER stress-induced cell death, is a substrate of PARKIN [307,326]. Further, PARKIN overexpression in drosophila models results in an increase in K48-linked polyubiquitination and a decrease in protein aggregation [327].
HRD1 attenuates 6-hydroxydopamine (6OHDA)-induced cell death in dopaminergic neuroblastoma cells [328].
GD is the most common lysosomal storage disorder (LSD) and is an autosomal recessive sphingolipidosis caused by mutations in a lysosomal enzyme, glucocerebrosidase (GCase), or its activator protein, saposin C, which functions in hydrolysis of glucosylceramide (GlcCer) to ceramide and glucose [308,329]. Therefore, GD is characterized by accumulation of GlcCer. Interestingly, accumulation of GlcCer in neurons leads to an increase in ER Ca2+ release through ryanodine receptors (RyRs) and subsequent neuronal cell death [330,331]. Interestingly, RyR antagonist dantrolene corrected deregulated Ca2+ signaling and autophagy defects in a GD mouse model, suggesting the relation between Ca2+ and autophagy in GD [332,333].
Misfolded mutants of GCase are retained in the ER and are not efficiently transported from the ER to the lysosomes, which are subject to ERAD [334,335]. Interestingly, mutants of GCase show variable degrees of ER retention and subsequent ERAD, which is correlated with GD severity [336].

6. Concluding Remarks and Future Perspectives

The ER is an interconnected organelle that plays fundamental roles in the biosynthesis, folding, stabilization, maturation, and trafficking of secretory and transmembrane proteins. It is the largest organelle and critically modulates nearly all aspects of life, including food intake, water balance, growth, metabolism, and immunity. Therefore, the ER undergoes continuous reconstruction to maintain its function and integrity. Unfortunately, even with an elaborate network to ensure adequate protein folding and assembly, misfolding and subsequent formation of protein aggregates can occur quite frequently, which could be a primary pathogenic mechanism in multiple human disorders. Therefore, mammals have developed two sophisticated guardian pathways, ERAD and ER-phagy, for the maintenance of proteostasis and organismal homeostasis. However, the relative contributions of these two guardian pathways remain still elusive. Many proteins can be eliminated by both pathways, suggesting that ERAD and ER-phagy are not mutually exclusive [16,246,337]. Given that more than 8000 proteins pass through the ER in a single human cell [338], the diversity as well as the flexibility of the two guardian pathways must be getting more complex and elaborate.
The field of ubiquitin has been extensively studied in the last several decades and a large number of studies aim to improve our understanding of the complex nature of the ubiquitin network. Given that E3 ligases are dysregulated in various processes of pathogenesis, E3 ligases have emerged as promising therapeutic targets for the treatment of diseases. Compared to ERAD, the involvement of specific E3 ubiquitin ligase and ubiquitination during the process of ER-phagy remains largely unknown. p62 has been shown to engage ubiquitinated cargos with autophagosomes during ER-phagy [339]. More recently, the association of p62 with TRIM13 has been reported to mediate ER-phagy [296]. Addressing outstanding mechanistic questions regarding E3 ubiquitin ligases and the subsequent ubiquitination of ER proteins during ER-phagy is important as the recruitment of ubiquitin-binding proteins for recognition by the phagophore could expand fundamental knowledge about this fascinating process.
Deregulation of protein quality control as well as protein quantity control at the ER and the subsequent failure in the re-establishment of proteostasis are closely associated with various human diseases, including cancer, cardiovascular diseases, immune diseases and neurodegenerative diseases [119,120]. Therefore, more emphasis must be placed on interrogating how cells manage the complexity of the ubiquitin code and how the cellular ubiquitin dynamics are fine-tuned in the ER, as this could lead to the identification of specific biomarkers, such as up- or downregulated UPR markers and dysregulated E3 ubiquitin ligases, for multiple human diseases and the characterization of new targets for drug development.

Author Contributions

Writing—original draft preparation, J.A.K. and Y.J.J.; writing—review and editing, Y.J.J.; supervision, Y.J.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by grant from the National Research Foundation of Korea to Y.J.J. (NRF2019R1A2C2002324). J.A.K. was the recipient of the BK21 Plus fellowship.

Acknowledgments

This work was supported by a grant from the National Research Foundation of Korea to Y.J.J. (NRF-2019R1A2C2002324). J.A. Kang was the recipient of the BK21 Plus fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bedford, L.; Lowe, J.; Dick, L.R.; Mayer, R.J.; Brownell, J.E. Ubiquitin-like protein conjugation and the ubiquitin–proteasome system as drug targets. Nat. Rev. Drug Discov. 2011, 10, 29–46. [Google Scholar] [CrossRef] [PubMed]
  2. Komander, D.; Rape, M. The ubiquitin code. Annu. Rev. Biochem. 2012, 81, 203–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Clague, M.J.; Liu, H.; Urbé, S. Governance of endocytic trafficking and signaling by reversible ubiquitylation. Dev. Cell 2012, 23, 457–467. [Google Scholar] [CrossRef] [Green Version]
  4. Wright, J.D.; Mace, P.D.; Day, C.L. Noncovalent ubiquitin interactions regulate the catalytic activity of ubiquitin writers. Trends Biochem. Sci. 2016, 41, 924–937. [Google Scholar] [CrossRef] [PubMed]
  5. Zheng, N.; Shabek, N. Ubiquitin ligases: Structure, function, and regulation. Annu. Rev. Biochem. 2017, 86, 129–157. [Google Scholar] [CrossRef]
  6. Mattern, M.; Sutherland, J.; Kadimisetty, K.; Barrio, R.; Rodriguez, M.S. Using ubiquitin binders to decipher the ubiquitin code. Trends Biochem. Sci. 2019, 44, 599–615. [Google Scholar] [CrossRef]
  7. Oh, E.; Akopian, D.; Rape, M. Principles of ubiquitin-dependent signaling. Annu. Rev. Cell Dev. Biol. 2018, 34, 137–162. [Google Scholar] [CrossRef]
  8. Hershko, A.; Ciechanover, A. The ubiquitin system. Annu. Rev. Biochem. 1998, 67, 425–479. [Google Scholar] [CrossRef]
  9. Song, L.; Luo, Z.-Q. Post-translational regulation of ubiquitin signaling. J. Cell Biol. 2019, 218, 1776–1786. [Google Scholar] [CrossRef]
  10. Popovic, D.; Vucic, D.; Dikic, I. Ubiquitination in disease pathogenesis and treatment. Nat. Med. 2014, 20, 1242–1253. [Google Scholar] [CrossRef]
  11. Heaton, S.M.; Borg, N.A.; Dixit, V.M. Ubiquitin in the activation and attenuation of innate antiviral immunity. J. Exp. Med. 2016, 213, 1–13. [Google Scholar] [CrossRef] [Green Version]
  12. Gilberto, S.; Peter, M. Dynamic ubiquitin signaling in cell cycle regulation. J. Cell Biol. 2017, 216, 2259–2271. [Google Scholar] [CrossRef] [PubMed]
  13. Han, H.G.; Moon, H.W.; Jeon, Y.J. ISG15 in cancer: Beyond ubiquitin-like protein. Cancer Lett. 2018, 438, 52–62. [Google Scholar] [CrossRef]
  14. Kang, J.A.; Jeon, Y.J. Emerging roles of USP18: From biology to pathophysiology. Int. J. Mol. Sci. 2020, 21, 6825. [Google Scholar] [CrossRef] [PubMed]
  15. Jeon, Y.J.; Yoo, H.M.; Chung, C.H. ISG15 and immune diseases. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2010, 1802, 485–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sun, Z.; Brodsky, J.L. Protein quality control in the secretory pathway. J. Cell Biol. 2019, 218, 3171–3187. [Google Scholar] [CrossRef] [Green Version]
  17. Fregno, I.; Molinari, M. Proteasomal and lysosomal clearance of faulty secretory proteins: ER-associated degradation (ERAD) and ER-to-lysosome-associated degradation (ERLAD) pathways. Crit. Rev. Biochem. Mol. Biol. 2019, 54, 153–163. [Google Scholar] [CrossRef] [PubMed]
  18. Sloper-Mould, K.E.; Jemc, J.C.; Pickart, C.M.; Hicke, L. Distinct functional surface regions on ubiquitin. J. Biol. Chem. 2001, 276, 30483–30489. [Google Scholar] [CrossRef] [Green Version]
  19. Özkaynak, E.; Finley, D.; Varshavsky, A. The yeast ubiquitin gene: Head-to-tail repeats encoding a polyubiquitin precursor protein. Nature 1984, 312, 663–666. [Google Scholar] [CrossRef]
  20. Finley, D.; Bartel, B.; Varshavsky, A. The tails of ubiquitin precursors are ribosomal proteins whose fusion to ubiquitin facilitates ribosome biogenesis. Nature 1989, 338, 394–401. [Google Scholar] [CrossRef] [PubMed]
  21. Rape, M. Ubiquitylation at the crossroads of development and disease. Nat. Rev. Mol. Cell Biol. 2018, 19, 59–70. [Google Scholar] [CrossRef]
  22. Kliza, K.; Husnjak, K. Resolving the complexity of ubiquitin networks. Front. Mol. Biosci. 2020, 7, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lake, M.W.; Wuebbens, M.M.; Rajagopalan, K.; Schindelin, H. Mechanism of ubiquitin activation revealed by the structure of a bacterial MoeB–MoaD complex. Nature 2001, 414, 325–329. [Google Scholar] [CrossRef] [PubMed]
  24. Walden, H.; Podgorski, M.S.; Schulman, B.A. Insights into the ubiquitin transfer cascade from the structure of the activating enzyme for NEDD8. Nature 2003, 422, 330–334. [Google Scholar] [CrossRef]
  25. Lois, L.M.; Lima, C.D. Structures of the SUMO E1 provide mechanistic insights into SUMO activation and E2 recruitment to E1. EMBO J. 2005, 24, 439–451. [Google Scholar] [CrossRef] [Green Version]
  26. Lee, I.; Schindelin, H. Structural insights into E1-catalyzed ubiquitin activation and transfer to conjugating enzymes. Cell 2008, 134, 268–278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Huang, D.T.; Paydar, A.; Zhuang, M.; Waddell, M.B.; Holton, J.M.; Schulman, B.A. Structural basis for recruitment of Ubc12 by an E2 binding domain in NEDD8′s E1. Mol. Cell 2005, 17, 341–350. [Google Scholar] [CrossRef] [PubMed]
  28. Walden, H.; Podgorski, M.S.; Huang, D.T.; Miller, D.W.; Howard, R.J.; Minor, D.L., Jr.; Holton, J.M.; Schulman, B.A. The structure of the APPBP1-UBA3-NEDD8-ATP complex reveals the basis for selective ubiquitin-like protein activation by an E1. Mol. Cell 2003, 12, 1427–1437. [Google Scholar] [CrossRef]
  29. Schulman, B.A.; Harper, J.W. Ubiquitin-like protein activation by E1 enzymes: The apex for downstream signalling pathways. Nat. Rev. Mol. Cell Biol. 2009, 10, 319–331. [Google Scholar] [CrossRef] [Green Version]
  30. Ye, Y.; Rape, M. Building ubiquitin chains: E2 enzymes at work. Nat. Rev. Mol. Cell Biol. 2009, 10, 755–764. [Google Scholar] [CrossRef] [Green Version]
  31. Burroughs, A.M.; Jaffee, M.; Iyer, L.M.; Aravind, L. Anatomy of the E2 ligase fold: Implications for enzymology and evolution of ubiquitin/Ub-like protein conjugation. J. Struct. Biol. 2008, 162, 205–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. van Wijk, S.J.; Timmers, H.M. The family of ubiquitin-conjugating enzymes (E2s): Deciding between life and death of proteins. FASEB J. 2010, 24, 981–993. [Google Scholar] [CrossRef]
  33. Buetow, L.; Huang, D.T. Structural insights into the catalysis and regulation of E3 ubiquitin ligases. Nat. Rev. Mol. Cell Biol. 2016, 17, 626. [Google Scholar] [CrossRef] [Green Version]
  34. Kravtsova-Ivantsiv, Y.; Ciechanover, A. Non-canonical ubiquitin-based signals for proteasomal degradation. J. Cell Sci. 2012, 125, 539–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Pao, K.-C.; Wood, N.T.; Knebel, A.; Rafie, K.; Stanley, M.; Mabbitt, P.D.; Sundaramoorthy, R.; Hofmann, K.; van Aalten, D.M.; Virdee, S. Activity-based E3 ligase profiling uncovers an E3 ligase with esterification activity. Nature 2018, 556, 381–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Cadwell, K.; Coscoy, L. Ubiquitination on nonlysine residues by a viral E3 ubiquitin ligase. Science 2005, 309, 127–130. [Google Scholar] [CrossRef] [Green Version]
  37. Shimizu, Y.; Okuda-Shimizu, Y.; Hendershot, L.M. Ubiquitylation of an ERAD substrate occurs on multiple types of amino acids. Mol. Cell 2010, 40, 917–926. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, X.; Herr, R.A.; Hansen, T.H. Ubiquitination of substrates by esterification. Traffic 2012, 13, 19–24. [Google Scholar] [CrossRef] [Green Version]
  39. Bhogaraju, S.; Kalayil, S.; Liu, Y.; Bonn, F.; Colby, T.; Matic, I.; Dikic, I. Phosphoribosylation of ubiquitin promotes serine ubiquitination and impairs conventional ubiquitination. Cell 2016, 167, 1636–1649.e1613. [Google Scholar] [CrossRef]
  40. Qiu, J.; Sheedlo, M.J.; Yu, K.; Tan, Y.; Nakayasu, E.S.; Das, C.; Liu, X.; Luo, Z.-Q. Ubiquitination independent of E1 and E2 enzymes by bacterial effectors. Nature 2016, 533, 120–124. [Google Scholar] [CrossRef] [Green Version]
  41. Mevissen, T.E.; Komander, D. Mechanisms of deubiquitinase specificity and regulation. Annu. Rev. Biochem. 2017, 86, 159–192. [Google Scholar] [CrossRef] [Green Version]
  42. Clague, M.J.; Barsukov, I.; Coulson, J.M.; Liu, H.; Rigden, D.J.; Urbé, S. Deubiquitylases from genes to organism. Physiol. Rev. 2013, 93, 1289–1315. [Google Scholar] [CrossRef]
  43. Rehman, S.A.A.; Kristariyanto, Y.A.; Choi, S.-Y.; Nkosi, P.J.; Weidlich, S.; Labib, K.; Hofmann, K.; Kulathu, Y. MINDY-1 is a member of an evolutionarily conserved and structurally distinct new family of deubiquitinating enzymes. Mol. Cell 2016, 63, 146–155. [Google Scholar]
  44. Yau, R.; Rape, M. The increasing complexity of the ubiquitin code. Nat. Cell Biol. 2016, 18, 579–586. [Google Scholar] [CrossRef] [PubMed]
  45. Deshaies, R.J.; Joazeiro, C.A. RING domain E3 ubiquitin ligases. Annu. Rev. Biochem. 2009, 78, 339–434. [Google Scholar] [CrossRef] [PubMed]
  46. Ohi, M.D.; Vander Kooi, C.W.; Rosenberg, J.A.; Chazin, W.J.; Gould, K.L. Structural insights into the U-box, a domain associated with multi-ubiquitination. Nat. Struct. Mol. Biol. 2003, 10, 250–255. [Google Scholar] [CrossRef]
  47. Dou, H.; Buetow, L.; Sibbet, G.J.; Cameron, K.; Huang, D.T. BIRC7–E2 ubiquitin conjugate structure reveals the mechanism of ubiquitin transfer by a RING dimer. Nat. Struct. Mol. Biol. 2012, 19, 876. [Google Scholar] [CrossRef] [Green Version]
  48. Plechanovová, A.; Jaffray, E.G.; Tatham, M.H.; Naismith, J.H.; Hay, R.T. Structure of a RING E3 ligase and ubiquitin-loaded E2 primed for catalysis. Nature 2012, 489, 115–120. [Google Scholar] [CrossRef] [Green Version]
  49. Pruneda, J.N.; Littlefield, P.J.; Soss, S.E.; Nordquist, K.A.; Chazin, W.J.; Brzovic, P.S.; Klevit, R.E. Structure of an E3: E2∼ Ub complex reveals an allosteric mechanism shared among RING/U-box ligases. Mol. Cell 2012, 47, 933–942. [Google Scholar] [CrossRef] [Green Version]
  50. Zheng, N.; Wang, P.; Jeffrey, P.D.; Pavletich, N.P. Structure of a c-Cbl–UbcH7 complex: RING domain function in ubiquitin-protein ligases. Cell 2000, 102, 533–539. [Google Scholar] [CrossRef] [Green Version]
  51. Dominguez, C.; Bonvin, A.M.; Winkler, G.S.; van Schaik, F.M.; Timmers, H.T.M.; Boelens, R. Structural model of the UbcH5B/CNOT4 complex revealed by combining NMR, mutagenesis, and docking approaches. Structure 2004, 12, 633–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Buetow, L.; Gabrielsen, M.; Anthony, N.G.; Dou, H.; Patel, A.; Aitkenhead, H.; Sibbet, G.J.; Smith, B.O.; Huang, D.T. Activation of a primed RING E3-E2–ubiquitin complex by non-covalent ubiquitin. Mol. Cell 2015, 58, 297–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Mace, P.D.; Linke, K.; Feltham, R.; Schumacher, F.-R.; Smith, C.A.; Vaux, D.L.; Silke, J.; Day, C.L. Structures of the cIAP2 RING domain reveal conformational changes associated with ubiquitin-conjugating enzyme (E2) recruitment. J. Biol. Chem. 2008, 283, 31633–31640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yin, Q.; Lin, S.-C.; Lamothe, B.; Lu, M.; Lo, Y.-C.; Hura, G.; Zheng, L.; Rich, R.L.; Campos, A.D.; Myszka, D.G. E2 interaction and dimerization in the crystal structure of TRAF6. Nat. Struct. Mol. Biol. 2009, 16, 658. [Google Scholar] [CrossRef] [Green Version]
  55. Plechanovová, A.; Jaffray, E.G.; McMahon, S.A.; Johnson, K.A.; Navrátilová, I.; Naismith, J.H.; Hay, R.T. Mechanism of ubiquitylation by dimeric RING ligase RNF4. Nat. Struct. Mol. Biol. 2011, 18, 1052. [Google Scholar] [CrossRef]
  56. Brzovic, P.S.; Rajagopal, P.; Hoyt, D.W.; King, M.-C.; Klevit, R.E. Structure of a BRCA1–BARD1 heterodimeric RING–RING complex. Nat. Struct. Biol. 2001, 8, 833–837. [Google Scholar] [CrossRef] [PubMed]
  57. Buchwald, G.; van der Stoop, P.; Weichenrieder, O.; Perrakis, A.; van Lohuizen, M.; Sixma, T.K. Structure and E3-ligase activity of the Ring–Ring complex of Polycomb proteins Bmi1 and Ring1b. EMBO J. 2006, 25, 2465–2474. [Google Scholar] [CrossRef]
  58. Badciong, J.C.; Haas, A.L. MdmX is a RING finger ubiquitin ligase capable of synergistically enhancing Mdm2 ubiquitination. J. Biol. Chem. 2002, 277, 49668–49675. [Google Scholar] [CrossRef] [Green Version]
  59. Linares, L.K.; Hengstermann, A.; Ciechanover, A.; Müller, S.; Scheffner, M. HdmX stimulates Hdm2-mediated ubiquitination and degradation of p53. Proc. Natl. Acad. Sci. USA 2003, 100, 12009–12014. [Google Scholar] [CrossRef] [Green Version]
  60. Li, Z.; Cao, R.; Wang, M.; Myers, M.P.; Zhang, Y.; Xu, R.-M. Structure of a Bmi-1-Ring1B polycomb group ubiquitin ligase complex. J. Biol. Chem. 2006, 281, 20643–20649. [Google Scholar] [CrossRef] [Green Version]
  61. Micale, L.; Chaignat, E.; Fusco, C.; Reymond, A.; Merla, G. The tripartite motif: Structure and function. Adv. Exp. Med. Biol. 2012, 770, 11. [Google Scholar] [PubMed]
  62. Lydeard, J.R.; Schulman, B.A.; Harper, J.W. Building and remodelling Cullin–RING E3 ubiquitin ligases. EMBO Rep. 2013, 14, 1050–1061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Petroski, M.D.; Deshaies, R.J. Function and regulation of cullin–RING ubiquitin ligases. Nat. Rev. Mol. Cell Biol. 2005, 6, 9–20. [Google Scholar] [CrossRef] [Green Version]
  64. Chang, L.; Barford, D. Insights into the anaphase-promoting complex: A molecular machine that regulates mitosis. Curr. Opin. Struct. Biol. 2014, 29, 1–9. [Google Scholar] [CrossRef] [PubMed]
  65. Scheffner, M.; Kumar, S. Mammalian HECT ubiquitin-protein ligases: Biological and pathophysiological aspects. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Res. 2014, 1843, 61–74. [Google Scholar] [CrossRef]
  66. Rotin, D.; Kumar, S. Physiological functions of the HECT family of ubiquitin ligases. Nat. Rev. Mol. Cell Biol. 2009, 10, 398–409. [Google Scholar] [CrossRef] [PubMed]
  67. Wenzel, D.M.; Lissounov, A.; Brzovic, P.S.; Klevit, R.E. UBCH7 reactivity profile reveals parkin and HHARI to be RING/HECT hybrids. Nature 2011, 474, 105–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Smit, J.J.; Monteferrario, D.; Noordermeer, S.M.; Van Dijk, W.J.; Van Der Reijden, B.A.; Sixma, T.K. The E3 ligase HOIP specifies linear ubiquitin chain assembly through its RING-IBR-RING domain and the unique LDD extension. EMBO J. 2012, 31, 3833–3844. [Google Scholar] [CrossRef] [Green Version]
  69. Stieglitz, B.; Morris-Davies, A.C.; Koliopoulos, M.G.; Christodoulou, E.; Rittinger, K. LUBAC synthesizes linear ubiquitin chains via a thioester intermediate. EMBO Rep. 2012, 13, 840–846. [Google Scholar] [CrossRef]
  70. Spratt, D.E.; Walden, H.; Shaw, G.S. RBR E3 ubiquitin ligases: New structures, new insights, new questions. Biochem. J. 2014, 458, 421–437. [Google Scholar] [CrossRef] [Green Version]
  71. Chaugule, V.K.; Burchell, L.; Barber, K.R.; Sidhu, A.; Leslie, S.J.; Shaw, G.S.; Walden, H. Autoregulation of Parkin activity through its ubiquitin-like domain. EMBO J. 2011, 30, 2853–2867. [Google Scholar] [CrossRef] [PubMed]
  72. Kitada, T.; Asakawa, S.; Hattori, N.; Matsumine, H.; Yamamura, Y.; Minoshima, S.; Yokochi, M.; Mizuno, Y.; Shimizu, N. Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature 1998, 392, 605–608. [Google Scholar] [CrossRef] [PubMed]
  73. Kirisako, T.; Kamei, K.; Murata, S.; Kato, M.; Fukumoto, H.; Kanie, M.; Sano, S.; Tokunaga, F.; Tanaka, K.; Iwai, K. A ubiquitin ligase complex assembles linear polyubiquitin chains. EMBO J. 2006, 25, 4877–4887. [Google Scholar] [CrossRef]
  74. Aguilera, M.; Oliveros, M.; Martínez-Padrón, M.; Barbas, J.A.; Ferrús, A. Ariadne-1: A vital Drosophila gene is required in development and defines a new conserved family of ring-finger proteins. Genetics 2000, 155, 1231–1244. [Google Scholar] [PubMed]
  75. Ikeda, F.; Deribe, Y.L.; Skånland, S.S.; Stieglitz, B.; Grabbe, C.; Franz-Wachtel, M.; Van Wijk, S.J.; Goswami, P.; Nagy, V.; Terzic, J. SHARPIN forms a linear ubiquitin ligase complex regulating NF-κB activity and apoptosis. Nature 2011, 471, 637–641. [Google Scholar] [CrossRef] [PubMed]
  76. Tokunaga, F.; Nakagawa, T.; Nakahara, M.; Saeki, Y.; Taniguchi, M.; Sakata, S.-I.; Tanaka, K.; Nakano, H.; Iwai, K. SHARPIN is a component of the NF-κB-activating linear ubiquitin chain assembly complex. Nature 2011, 471, 633–636. [Google Scholar] [CrossRef]
  77. Walden, H.; Rittinger, K. RBR ligase–mediated ubiquitin transfer: A tale with many twists and turns. Nat. Struct. Mol. Biol. 2018, 25, 440–445. [Google Scholar] [CrossRef]
  78. Stewart, M.D.; Ritterhoff, T.; Klevit, R.E.; Brzovic, P.S. E2 enzymes: More than just middle men. Cell Res. 2016, 26, 423–440. [Google Scholar] [CrossRef] [Green Version]
  79. Koegl, M.; Hoppe, T.; Schlenker, S.; Ulrich, H.D.; Mayer, T.U.; Jentsch, S. A novel ubiquitination factor, E4, is involved in multiubiquitin chain assembly. Cell 1999, 96, 635–644. [Google Scholar] [CrossRef] [Green Version]
  80. Hoppe, T. Multiubiquitylation by E4 enzymes:‘one size’doesn’t fit all. Trends Biochem. Sci. 2005, 30, 183–187. [Google Scholar] [CrossRef] [PubMed]
  81. Lopata, A.; Kniss, A.; Löhr, F.; Rogov, V.V.; Dötsch, V. Ubiquitination in the ERAD Process. Int. J. Mol. Sci. 2020, 21, 5369. [Google Scholar] [CrossRef]
  82. Antoniou, N.; Lagopati, N.; Balourdas, D.I.; Nikolaou, M.; Papalampros, A.; Vasileiou, P.V.; Myrianthopoulos, V.; Kotsinas, A.; Shiloh, Y.; Liontos, M. The role of E3, E4 ubiquitin ligase (UBE4B) in human pathologies. Cancers 2020, 12, 62. [Google Scholar] [CrossRef] [Green Version]
  83. Pant, V.; Lozano, G. Limiting the power of p53 through the ubiquitin proteasome pathway. Genes Dev. 2014, 28, 1739–1751. [Google Scholar] [CrossRef] [Green Version]
  84. Shi, D.; Pop, M.S.; Kulikov, R.; Love, I.M.; Kung, A.L.; Grossman, S.R. CBP and p300 are cytoplasmic E4 polyubiquitin ligases for p53. Proc. Natl. Acad. Sci. USA 2009, 106, 16275–16280. [Google Scholar] [CrossRef] [Green Version]
  85. Hatakeyama, S.; Yada, M.; Matsumoto, M.; Ishida, N.; Nakayama, K.-I. U box proteins as a new family of ubiquitin-protein ligases. J. Biol. Chem. 2001, 276, 33111–33120. [Google Scholar] [CrossRef] [Green Version]
  86. Johnson, E.S.; Ma, P.C.; Ota, I.M.; Varshavsky, A. A proteolytic pathway that recognizes ubiquitin as a degradation signal. J. Biol. Chem. 1995, 270, 17442–17456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Zhang, Y.; Lv, Y.; Zhang, Y.; Gao, H. Regulation of p53 level by UBE4B in breast cancer. PLoS ONE 2014, 9, e90154. [Google Scholar]
  88. Contino, G.; Amati, F.; Pucci, S.; Pontieri, E.; Pichiorri, F.; Novelli, A.; Botta, A.; Mango, R.; Nardone, A.M.; Sangiuolo, F.C. Expression analysis of the gene encoding for the U-box-type ubiquitin ligase UBE4A in human tissues. Gene 2004, 328, 69–74. [Google Scholar] [CrossRef] [PubMed]
  89. Hosoda, M.; Ozaki, T.; Miyazaki, K.; Hayashi, S.; Furuya, K.; Watanabe, K.-i.; Nakagawa, T.; Hanamoto, T.; Todo, S.; Nakagawara, A. UFD2a mediates the proteasomal turnover of p73 without promoting p73 ubiquitination. Oncogene 2005, 24, 7156–7169. [Google Scholar] [CrossRef] [Green Version]
  90. Carén, H.; Holmstrand, A.; Sjöberg, R.-M.; Martinsson, T. The two human homologues of yeast UFD2 ubiquitination factor, UBE4A and UBE4B, are located in common neuroblastoma deletion regions and are subject to mutations in tumours. Eur. J. Cancer 2006, 42, 381–387. [Google Scholar] [CrossRef] [PubMed]
  91. Morreale, F.E.; Walden, H. Types of ubiquitin ligases. Cell 2016, 165, 248–248.e241. [Google Scholar] [CrossRef]
  92. Grossman, S.R.; Deato, M.E.; Brignone, C.; Chan, H.M.; Kung, A.L.; Tagami, H.; Nakatani, Y.; Livingston, D.M. Polyubiquitination of p53 by a ubiquitin ligase activity of p300. Science 2003, 300, 342–344. [Google Scholar] [CrossRef]
  93. Arany, Z.n.; Sellers, W.R.; Livingston, D.M.; Eckner, R. E1A-associated p300 and CREB-associated CBP belong to a conserved family of coactivators. Cell 1994, 77, 799–800. [Google Scholar] [CrossRef]
  94. Lundblad, J.R.; Kwok, R.P.; Laurance, M.E.; Harter, M.L.; Goodman, R.H. Adenoviral ElA-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature 1995, 374, 85–88. [Google Scholar] [CrossRef]
  95. Ballinger, C.A.; Connell, P.; Wu, Y.; Hu, Z.; Thompson, L.J.; Yin, L.-Y.; Patterson, C. Identification of CHIP, a novel tetratricopeptide repeat-containing protein that interacts with heat shock proteins and negatively regulates chaperone functions. Mol. Cell. Biol. 1999, 19, 4535–4545. [Google Scholar] [CrossRef] [Green Version]
  96. Imai, Y.; Soda, M.; Hatakeyama, S.; Akagi, T.; Hashikawa, T.; Nakayama, K.-I.; Takahashi, R. CHIP is associated with Parkin, a gene responsible for familial Parkinson’s disease, and enhances its ubiquitin ligase activity. Mol. Cell 2002, 10, 55–67. [Google Scholar] [CrossRef]
  97. Sui, G.; Affar, E.B.; Shi, Y.; Brignone, C.; Wall, N.R.; Yin, P.; Donohoe, M.; Luke, M.P.; Calvo, D.; Grossman, S.R. Yin Yang 1 is a negative regulator of p53. Cell 2004, 117, 859–872. [Google Scholar] [CrossRef] [Green Version]
  98. Lee, J.-S.; Galvin, K.M.; See, R.H.; Eckner, R.; Livingston, D.; Moran, E.; Shi, Y. Relief of YY1 transcriptional repression by adenovirus E1A is mediated by E1A-associated protein p300. Genes Dev. 1995, 9, 1188–1198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Chen, B.B.; Mallampalli, R.K. Masking of a nuclear signal motif by monoubiquitination leads to mislocalization and degradation of the regulatory enzyme cytidylyltransferase. Mol. Cell. Biol. 2009, 29, 3062–3075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Braten, O.; Livneh, I.; Ziv, T.; Admon, A.; Kehat, I.; Caspi, L.H.; Gonen, H.; Bercovich, B.; Godzik, A.; Jahandideh, S. Numerous proteins with unique characteristics are degraded by the 26S proteasome following monoubiquitination. Proc. Natl. Acad. Sci. USA 2016, 113, E4639–E4647. [Google Scholar] [CrossRef] [Green Version]
  101. Ohtake, F.; Tsuchiya, H. The emerging complexity of ubiquitin architecture. J. Biochem. 2017, 161, 125–133. [Google Scholar] [CrossRef] [Green Version]
  102. Dwane, L.; Gallagher, W.M.; Chonghaile, T.N.; O’Connor, D.P. The emerging role of non-traditional ubiquitination in oncogenic pathways. J. Biol. Chem. 2017, 292, 3543–3551. [Google Scholar] [CrossRef] [Green Version]
  103. Akutsu, M.; Dikic, I.; Bremm, A. Ubiquitin chain diversity at a glance. J. Cell Sci. 2016, 129, 875–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Kwon, Y.T.; Ciechanover, A. The ubiquitin code in the ubiquitin-proteasome system and autophagy. Trends Biochem. Sci. 2017, 42, 873–886. [Google Scholar] [CrossRef] [PubMed]
  105. Swatek, K.N.; Komander, D. Ubiquitin modifications. Cell Res. 2016, 26, 399–422. [Google Scholar] [CrossRef] [Green Version]
  106. Alfano, C.; Faggiano, S.; Pastore, A. The ball and chain of polyubiquitin structures. Trends Biochem. Sci. 2016, 41, 371–385. [Google Scholar] [CrossRef] [Green Version]
  107. Thrower, J.S.; Hoffman, L.; Rechsteiner, M.; Pickart, C.M. Recognition of the polyubiquitin proteolytic signal. EMBO J. 2000, 19, 94–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chan, N.-L.; Hill, C.P. Defining polyubiquitin chain topology. Nat. Struct. Biol. 2001, 8, 650–652. [Google Scholar] [CrossRef] [PubMed]
  109. Hicke, L.; Dunn, R. Regulation of membrane protein transport by ubiquitin and ubiquitin-binding proteins. Annu. Rev. Cell Dev. Biol. 2003, 19, 141–172. [Google Scholar] [CrossRef]
  110. Huang, T.T.; D’Andrea, A.D. Regulation of DNA repair by ubiquitylation. Nat. Rev. Mol. Cell Biol. 2006, 7, 323–334. [Google Scholar] [CrossRef] [PubMed]
  111. Kolas, N.K.; Chapman, J.R.; Nakada, S.; Ylanko, J.; Chahwan, R.; Sweeney, F.D.; Panier, S.; Mendez, M.; Wildenhain, J.; Thomson, T.M. Orchestration of the DNA-damage response by the RNF8 ubiquitin ligase. Science 2007, 318, 1637–1640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Meyer, H.-J.; Rape, M. Enhanced protein degradation by branched ubiquitin chains. Cell 2014, 157, 910–921. [Google Scholar] [CrossRef] [Green Version]
  113. Kristariyanto, Y.A.; Rehman, S.A.A.; Campbell, D.G.; Morrice, N.A.; Johnson, C.; Toth, R.; Kulathu, Y. K29-selective ubiquitin binding domain reveals structural basis of specificity and heterotypic nature of k29 polyubiquitin. Mol. Cell 2015, 58, 83–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Michel, M.A.; Swatek, K.N.; Hospenthal, M.K.; Komander, D. Ubiquitin linkage-specific affimers reveal insights into K6-linked ubiquitin signaling. Mol. Cell 2017, 68, 233–246. e235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Yau, R.G.; Doerner, K.; Castellanos, E.R.; Haakonsen, D.L.; Werner, A.; Wang, N.; Yang, X.W.; Martinez-Martin, N.; Matsumoto, M.L.; Dixit, V.M. Assembly and function of heterotypic ubiquitin chains in cell-cycle and protein quality control. Cell 2017, 171, 918–933.e920. [Google Scholar] [CrossRef] [Green Version]
  116. Emmerich, C.H.; Bakshi, S.; Kelsall, I.R.; Ortiz-Guerrero, J.; Shpiro, N.; Cohen, P. Lys63/Met1-hybrid ubiquitin chains are commonly formed during the activation of innate immune signalling. Biochem. Biophys. Res. Commun. 2016, 474, 452–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Emmerich, C.H.; Ordureau, A.; Strickson, S.; Arthur, J.S.C.; Pedrioli, P.G.; Komander, D.; Cohen, P. Activation of the canonical IKK complex by K63/M1-linked hybrid ubiquitin chains. Proc. Natl. Acad. Sci. USA 2013, 110, 15247–15252. [Google Scholar] [CrossRef] [Green Version]
  118. Wertz, I.E.; Newton, K.; Seshasayee, D.; Kusam, S.; Lam, C.; Zhang, J.; Popovych, N.; Helgason, E.; Schoeffler, A.; Jeet, S. Phosphorylation and linear ubiquitin direct A20 inhibition of inflammation. Nature 2015, 528, 370–375. [Google Scholar] [CrossRef]
  119. Chiti, F.; Dobson, C.M. Protein misfolding, functional amyloid, and human disease. Annu. Rev. Biochem. 2006, 75, 333–366. [Google Scholar] [CrossRef] [Green Version]
  120. Wang, M.; Kaufman, R.J. Protein misfolding in the endoplasmic reticulum as a conduit to human disease. Nature 2016, 529, 326–335. [Google Scholar] [CrossRef]
  121. Nam, S.M.; Jeon, Y.J. Proteostasis in the endoplasmic reticulum: Road to cure. Cancers 2019, 11, 1793. [Google Scholar] [CrossRef] [Green Version]
  122. Moon, H.W.; Han, H.G.; Jeon, Y.J. Protein quality control in the endoplasmic reticulum and cancer. Int. J. Mol. Sci. 2018, 19, 3020. [Google Scholar] [CrossRef] [Green Version]
  123. Yoo, Y.S.; Han, H.G.; Jeon, Y.J. Unfolded protein response of the endoplasmic reticulum in tumor progression and immunogenicity. Oxidative Med. Cell. Longev. 2017, 2017, 2969271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Brodsky, J.L.; Wojcikiewicz, R.J. Substrate-specific mediators of ER associated degradation (ERAD). Curr. Opin. Cell Biol. 2009, 21, 516–521. [Google Scholar] [CrossRef] [Green Version]
  125. Hampton, R.Y. ER-associated degradation in protein quality control and cellular regulation. Curr. Opin. Cell Biol. 2002, 14, 476–482. [Google Scholar] [CrossRef]
  126. Vembar, S.S.; Brodsky, J.L. One step at a time: Endoplasmic reticulum-associated degradation. Nat. Rev. Mol. Cell Biol. 2008, 9, 944–957. [Google Scholar] [CrossRef]
  127. Adle, D.J.; Wei, W.; Smith, N.; Bies, J.J.; Lee, J. Cadmium-mediated rescue from ER-associated degradation induces expression of its exporter. Proc. Natl. Acad. Sci. USA 2009, 106, 10189–10194. [Google Scholar] [CrossRef] [Green Version]
  128. Morito, D.; Nagata, K. Pathogenic hijacking of ER-associated degradation: Is ERAD flexible? Mol. Cell 2015, 59, 335–344. [Google Scholar] [CrossRef] [Green Version]
  129. Christianson, J.C.; Ye, Y. Cleaning up in the endoplasmic reticulum: Ubiquitin in charge. Nat. Struct. Mol. Biol. 2014, 21, 325. [Google Scholar] [CrossRef] [PubMed]
  130. Olzmann, J.A.; Kopito, R.R.; Christianson, J.C. The mammalian endoplasmic reticulum-associated degradation system. Cold Spring Harb. Perspect. Biol. 2013, 5, a013185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Xu, C.; Ng, D.T. Glycosylation-directed quality control of protein folding. Nat. Rev. Mol. Cell Biol. 2015, 16, 742–752. [Google Scholar] [CrossRef]
  132. Dougan, S.K.; Hu, C.-C.A.; Paquet, M.-E.; Greenblatt, M.B.; Kim, J.; Lilley, B.N.; Watson, N.; Ploegh, H.L. Derlin-2-deficient mice reveal an essential role for protein dislocation in chondrocytes. Mol. Cell. Biol. 2011, 31, 1145–1159. [Google Scholar] [CrossRef] [Green Version]
  133. Francisco, A.B.; Singh, R.; Li, S.; Vani, A.K.; Yang, L.; Munroe, R.J.; Diaferia, G.; Cardano, M.; Biunno, I.; Qi, L. Deficiency of suppressor enhancer Lin12 1 like (SEL1L) in mice leads to systemic endoplasmic reticulum stress and embryonic lethality. J. Biol. Chem. 2010, 285, 13694–13703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Ruggiano, A.; Foresti, O.; Carvalho, P. ER-associated degradation: Protein quality control and beyond. J. Cell Biol. 2014, 204, 869–879. [Google Scholar] [CrossRef] [Green Version]
  135. Helenius, A.; Aebi, M. Roles of N-linked glycans in the endoplasmic reticulum. Annu. Rev. Biochem. 2004, 73, 1019–1049. [Google Scholar] [CrossRef]
  136. Gonzalez, D.S.; Karaveg, K.; Vandersall-Nairn, A.S.; Lal, A.; Moremen, K.W. Identification, expression, and characterization of a cDNA encoding human endoplasmic reticulum mannosidase I, the enzyme that catalyzes the first mannose trimming step in mammalian Asn-linked oligosaccharide biosynthesis. J. Biol. Chem. 1999, 274, 21375–21386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Tremblay, L.O.; Herscovics, A. Cloning and expression of a specific human α1, 2-mannosidase that trims Man9GlcNAc2 to Man8GlcNAc2 isomer B during N-glycan biosynthesis. Glycobiology 1999, 9, 1073–1078. [Google Scholar] [CrossRef] [PubMed]
  138. Olivari, S.; Cali, T.; Salo, K.E.; Paganetti, P.; Ruddock, L.W.; Molinari, M. EDEM1 regulates ER-associated degradation by accelerating de-mannosylation of folding-defective polypeptides and by inhibiting their covalent aggregation. Biochem. Biophys. Res. Commun. 2006, 349, 1278–1284. [Google Scholar] [CrossRef] [PubMed]
  139. Hosokawa, N.; Tremblay, L.O.; Sleno, B.; Kamiya, Y.; Wada, I.; Nagata, K.; Kato, K.; Herscovics, A. EDEM1 accelerates the trimming of α1, 2-linked mannose on the C branch of N-glycans. Glycobiology 2010, 20, 567–575. [Google Scholar] [CrossRef] [Green Version]
  140. Hirao, K.; Natsuka, Y.; Tamura, T.; Wada, I.; Morito, D.; Natsuka, S.; Romero, P.; Sleno, B.; Tremblay, L.O.; Herscovics, A. EDEM3, a soluble EDEM homolog, enhances glycoprotein endoplasmic reticulum-associated degradation and mannose trimming. J. Biol. Chem. 2006, 281, 9650–9658. [Google Scholar] [CrossRef] [Green Version]
  141. Hosokawa, N.; Kamiya, Y.; Kamiya, D.; Kato, K.; Nagata, K. Human OS-9, a lectin required for glycoprotein endoplasmic reticulum-associated degradation, recognizes mannose-trimmed N-glycans. J. Biol. Chem. 2009, 284, 17061–17068. [Google Scholar] [CrossRef] [Green Version]
  142. Hosokawa, N.; You, Z.; Tremblay, L.O.; Nagata, K.; Herscovics, A. Stimulation of ERAD of misfolded null Hong Kong α1-antitrypsin by Golgi α1, 2-mannosidases. Biochem. Biophys. Res. Commun. 2007, 362, 626–632. [Google Scholar] [CrossRef]
  143. Lederkremer, G.Z. Glycoprotein folding, quality control and ER-associated degradation. Curr. Opin. Struct. Biol. 2009, 19, 515–523. [Google Scholar] [CrossRef] [PubMed]
  144. Ninagawa, S.; Okada, T.; Sumitomo, Y.; Kamiya, Y.; Kato, K.; Horimoto, S.; Ishikawa, T.; Takeda, S.; Sakuma, T.; Yamamoto, T. EDEM2 initiates mammalian glycoprotein ERAD by catalyzing the first mannose trimming step. J. Cell Biol. 2014, 206, 347–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Bernasconi, R.; Pertel, T.; Luban, J.; Molinari, M. A dual task for the Xbp1-responsive OS-9 variants in the mammalian endoplasmic reticulum: Inhibiting secretion of misfolded protein conformers and enhancing their disposal. J. Biol. Chem. 2008, 283, 16446–16454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Christianson, J.C.; Shaler, T.A.; Tyler, R.E.; Kopito, R.R. OS-9 and GRP94 deliver mutant α1-antitrypsin to the Hrd1–SEL1L ubiquitin ligase complex for ERAD. Nat. Cell Biol. 2008, 10, 272–282. [Google Scholar] [CrossRef] [Green Version]
  147. Hosokawa, N.; Wada, I.; Nagasawa, K.; Moriyama, T.; Okawa, K.; Nagata, K. Human XTP3-B forms an endoplasmic reticulum quality control scaffold with the HRD1-SEL1L ubiquitin ligase complex and BiP. J. Biol. Chem. 2008, 283, 20914–20924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Plemper, R.K.; Böhmler, S.; Bordallo, J.; Sommer, T.; Wolf, D.H. Mutant analysis links the translocon and BiP to retrograde protein transport for ER degradation. Nature 1997, 388, 891–895. [Google Scholar] [CrossRef] [PubMed]
  149. Ushioda, R.; Hoseki, J.; Nagata, K. Glycosylation-independent ERAD pathway serves as a backup system under ER stress. Mol. Biol. Cell 2013, 24, 3155–3163. [Google Scholar] [CrossRef] [PubMed]
  150. Shenkman, M.; Groisman, B.; Ron, E.; Avezov, E.; Hendershot, L.M.; Lederkremer, G.Z. A shared endoplasmic reticulum-associated degradation pathway involving the EDEM1 protein for glycosylated and nonglycosylated proteins. J. Biol. Chem. 2013, 288, 2167–2178. [Google Scholar] [CrossRef] [Green Version]
  151. Grubb, S.; Guo, L.; Fisher, E.A.; Brodsky, J.L. Protein disulfide isomerases contribute differentially to the endoplasmic reticulum–associated degradation of apolipoprotein B and other substrates. Mol. Biol. Cell 2012, 23, 520–532. [Google Scholar] [CrossRef]
  152. Hampton, R.Y.; Sommer, T. Finding the will and the way of ERAD substrate retrotranslocation. Curr. Opin. Cell Biol. 2012, 24, 460–466. [Google Scholar] [CrossRef] [PubMed]
  153. DeLaBarre, B.; Brunger, A.T. Complete structure of p97/valosin-containing protein reveals communication between nucleotide domains. Nat. Struct. Mol. Biol. 2003, 10, 856–863. [Google Scholar] [CrossRef]
  154. Meyer, H.; Bug, M.; Bremer, S. Emerging functions of the VCP/p97 AAA-ATPase in the ubiquitin system. Nat. Cell Biol. 2012, 14, 117–123. [Google Scholar] [CrossRef] [PubMed]
  155. Huyton, T.; Pye, V.E.; Briggs, L.C.; Flynn, T.C.; Beuron, F.; Kondo, H.; Ma, J.; Zhang, X.; Freemont, P.S. The crystal structure of murine p97/VCP at 3.6 Å. J. Struct. Biol. 2003, 144, 337–348. [Google Scholar] [CrossRef]
  156. Dreveny, I.; Kondo, H.; Uchiyama, K.; Shaw, A.; Zhang, X.; Freemont, P.S. Structural basis of the interaction between the AAA ATPase p97/VCP and its adaptor protein p47. EMBO J. 2004, 23, 1030–1039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Rabinovich, E.; Kerem, A.; Fröhlich, K.-U.; Diamant, N.; Bar-Nun, S. AAA-ATPase p97/Cdc48p, a cytosolic chaperone required for endoplasmic reticulum-associated protein degradation. Mol. Cell. Biol. 2002, 22, 626–634. [Google Scholar] [CrossRef] [Green Version]
  158. Liang, J.; Yin, C.; Doong, H.; Fang, S.; Peterhoff, C.; Nixon, R.A.; Monteiro, M.J. Characterization of erasin (UBXD2): A new ER protein that promotes ER-associated protein degradation. J. Cell Sci. 2006, 119, 4011–4024. [Google Scholar] [CrossRef] [Green Version]
  159. Suzuki, M.; Otsuka, T.; Ohsaki, Y.; Cheng, J.; Taniguchi, T.; Hashimoto, H.; Taniguchi, H.; Fujimoto, T. Derlin-1 and UBXD8 are engaged in dislocation and degradation of lipidated ApoB-100 at lipid droplets. Mol. Biol. Cell 2012, 23, 800–810. [Google Scholar] [CrossRef]
  160. Ballar, P.; Shen, Y.; Yang, H.; Fang, S. The role of a novel p97/valosin-containing protein-interacting motif of gp78 in endoplasmic reticulum-associated degradation. J. Biol. Chem. 2006, 281, 35359–35368. [Google Scholar] [CrossRef] [Green Version]
  161. Ye, Y.; Shibata, Y.; Yun, C.; Ron, D.; Rapoport, T.A. A membrane protein complex mediates retro-translocation from the ER lumen into the cytosol. Nature 2004, 429, 841–847. [Google Scholar] [CrossRef] [PubMed]
  162. Greenblatt, E.J.; Olzmann, J.A.; Kopito, R.R. Derlin-1 is a rhomboid pseudoprotease required for the dislocation of mutant α-1 antitrypsin from the endoplasmic reticulum. Nat. Struct. Mol. Biol. 2011, 18, 1147. [Google Scholar] [CrossRef] [Green Version]
  163. Kaneko, M.; Iwase, I.; Yamasaki, Y.; Takai, T.; Wu, Y.; Kanemoto, S.; Matsuhisa, K.; Asada, R.; Okuma, Y.; Watanabe, T. Genome-wide identification and gene expression profiling of ubiquitin ligases for endoplasmic reticulum protein degradation. Sci. Rep. 2016, 6, 30955. [Google Scholar] [CrossRef] [PubMed]
  164. Nadav, E.; Shmueli, A.; Barr, H.; Gonen, H.; Ciechanover, A.; Reiss, Y. A novel mammalian endoplasmic reticulum ubiquitin ligase homologous to the yeast Hrd1. Biochem. Biophys. Res. Commun. 2003, 303, 91–97. [Google Scholar] [CrossRef]
  165. Kikkert, M.; Doolman, R.; Dai, M.; Avner, R.; Hassink, G.; Van Voorden, S.; Thanedar, S.; Roitelman, J.; Chau, V.; Wiertz, E. Human HRD1 is an E3 ubiquitin ligase involved in degradation of proteins from the endoplasmic reticulum. J. Biol. Chem. 2004, 279, 3525–3534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Fang, S.; Ferrone, M.; Yang, C.; Jensen, J.P.; Tiwari, S.; Weissman, A.M. The tumor autocrine motility factor receptor, gp78, is a ubiquitin protein ligase implicated in degradation from the endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 2001, 98, 14422–14427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Hassink, G.; Kikkert, M.; Voorden, S.v.; Lee, S.-J.; Spaapen, R.; LAAR, T.v.; Coleman, C.S.; Bartee, E.; Früh, K.; Chau, V. TEB4 is a C4HC3 RING finger-containing ubiquitin ligase of the endoplasmic reticulum. Biochem. J. 2005, 388, 647–655. [Google Scholar] [CrossRef] [Green Version]
  168. Younger, J.M.; Chen, L.; Ren, H.-Y.; Rosser, M.F.; Turnbull, E.L.; Fan, C.-Y.; Patterson, C.; Cyr, D.M. Sequential quality-control checkpoints triage misfolded cystic fibrosis transmembrane conductance regulator. Cell 2006, 126, 571–582. [Google Scholar] [CrossRef] [Green Version]
  169. Jeon, Y.J.; Khelifa, S.; Ratnikov, B.; Scott, D.A.; Feng, Y.; Parisi, F.; Ruller, C.; Lau, E.; Kim, H.; Brill, L.M. Regulation of glutamine carrier proteins by RNF5 determines breast cancer response to ER stress-inducing chemotherapies. Cancer Cell 2015, 27, 354–369. [Google Scholar] [CrossRef] [Green Version]
  170. Tomati, V.; Sondo, E.; Armirotti, A.; Caci, E.; Pesce, E.; Marini, M.; Gianotti, A.; Jeon, Y.J.; Cilli, M.; Pistorio, A. Genetic inhibition of the ubiquitin ligase Rnf5 attenuates phenotypes associated to F508del cystic fibrosis mutation. Sci. Rep. 2015, 5, 12138. [Google Scholar] [CrossRef]
  171. Meacham, G.C.; Patterson, C.; Zhang, W.; Younger, J.M.; Cyr, D.M. The Hsc70 co-chaperone CHIP targets immature CFTR for proteasomal degradation. Nat. Cell Biol. 2001, 3, 100–105. [Google Scholar] [CrossRef] [PubMed]
  172. Yoshida, Y.; Chiba, T.; Tokunaga, F.; Kawasaki, H.; Iwai, K.; Suzuki, T.; Ito, Y.; Matsuoka, K.; Yoshida, M.; Tanaka, K. E3 ubiquitin ligase that recognizes sugar chains. Nature 2002, 418, 438–442. [Google Scholar] [CrossRef]
  173. Yoshida, Y.; Tokunaga, F.; Chiba, T.; Iwai, K.; Tanaka, K.; Tai, T. Fbs2 is a new member of the E3 ubiquitin ligase family that recognizes sugar chains. J. Biol. Chem. 2003, 278, 43877–43884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Magadan, J.G.; Perez-Victoria, F.J.; Sougrat, R.; Ye, Y.; Strebel, K.; Bonifacino, J.S. Multilayered mechanism of CD4 downregulation by HIV-1 Vpu involving distinct ER retention and ERAD targeting steps. PLoS Pathog 2010, 6, e1000869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Guo, X.; Shen, S.; Song, S.; He, S.; Cui, Y.; Xing, G.; Wang, J.; Yin, Y.; Fan, L.; He, F. The E3 ligase Smurf1 regulates Wolfram syndrome protein stability at the endoplasmic reticulum. J. Biol. Chem. 2011, 286, 18037–18047. [Google Scholar] [CrossRef] [Green Version]
  176. Fry, W.H.; Simion, C.; Sweeney, C.; Carraway, K.L. Quantity control of the ErbB3 receptor tyrosine kinase at the endoplasmic reticulum. Mol. Cell. Biol. 2011, 31, 3009–3018. [Google Scholar] [CrossRef] [Green Version]
  177. Morito, D.; Hirao, K.; Oda, Y.; Hosokawa, N.; Tokunaga, F.; Cyr, D.M.; Tanaka, K.; Iwai, K.; Nagata, K. Gp78 cooperates with RMA1 in endoplasmic reticulum-associated degradation of CFTRΔF508. Mol. Biol. Cell 2008, 19, 1328–1336. [Google Scholar] [CrossRef] [Green Version]
  178. Jo, Y.; Lee, P.C.; Sguigna, P.V.; DeBose-Boyd, R.A. Sterol-induced degradation of HMG CoA reductase depends on interplay of two Insigs and two ubiquitin ligases, gp78 and Trc8. Proc. Natl. Acad. Sci. USA 2011, 108, 20503–20508. [Google Scholar] [CrossRef] [Green Version]
  179. Zhang, T.; Xu, Y.; Liu, Y.; Ye, Y. gp78 functions downstream of Hrd1 to promote degradation of misfolded proteins of the endoplasmic reticulum. Mol. Biol. Cell 2015, 26, 4438–4450. [Google Scholar] [CrossRef]
  180. Jeon, Y.J.; Park, J.H.; Chung, C.H. Interferon-stimulated gene 15 in the control of cellular responses to genotoxic stress. Mol. Cells 2017, 40, 83. [Google Scholar] [CrossRef] [Green Version]
  181. Leto, D.E.; Morgens, D.W.; Zhang, L.; Walczak, C.P.; Elias, J.E.; Bassik, M.C.; Kopito, R.R. Genome-wide CRISPR analysis identifies substrate-specific conjugation modules in ER-associated degradation. Mol. Cell 2019, 73, 377–389.e311. [Google Scholar] [CrossRef] [Green Version]
  182. Blount, J.R.; Burr, A.A.; Denuc, A.; Marfany, G.; Todi, S.V. Ubiquitin-specific protease 25 functions in Endoplasmic Reticulum-associated degradation. PLoS ONE 2012, 7, e36542. [Google Scholar] [CrossRef] [Green Version]
  183. Ishikura, S.; Weissman, A.M.; Bonifacino, J.S. Serine residues in the cytosolic tail of the T-cell antigen receptor α-chain mediate ubiquitination and endoplasmic reticulum-associated degradation of the unassembled protein. J. Biol. Chem. 2010, 285, 23916–23924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Qi, L.; Tsai, B.; Arvan, P. New insights into the physiological role of endoplasmic reticulum-associated degradation. Trends Cell Biol. 2017, 27, 430–440. [Google Scholar] [CrossRef] [PubMed]
  185. Hwang, J.; Qi, L. Quality control in the endoplasmic reticulum: Crosstalk between ERAD and UPR pathways. Trends Biochem. Sci. 2018, 43, 593–605. [Google Scholar] [CrossRef] [PubMed]
  186. Shi, G.; Somlo, D.R.; Kim, G.H.; Prescianotto-Baschong, C.; Sun, S.; Beuret, N.; Long, Q.; Rutishauser, J.; Arvan, P.; Spiess, M. ER-associated degradation is required for vasopressin prohormone processing and systemic water homeostasis. J. Clin. Investig. 2017, 127, 3897–3912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Kim, G.H.; Shi, G.; Somlo, D.R.; Haataja, L.; Song, S.; Long, Q.; Nillni, E.A.; Low, M.J.; Arvan, P.; Myers, M.G. Hypothalamic ER–associated degradation regulates POMC maturation, feeding, and age-associated obesity. J. Clin. Investig. 2018, 128, 1125–1140. [Google Scholar] [CrossRef]
  188. Ji, Y.; Kim, H.; Yang, L.; Sha, H.; Roman, C.A.; Long, Q.; Qi, L. The Sel1L-Hrd1 endoplasmic reticulum-associated degradation complex manages a key checkpoint in B cell development. Cell Rep. 2016, 16, 2630–2640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Sun, S.; Shi, G.; Sha, H.; Ji, Y.; Han, X.; Shu, X.; Ma, H.; Inoue, T.; Gao, B.; Kim, H. IRE1α is an endogenous substrate of endoplasmic-reticulum-associated degradation. Nat. Cell Biol. 2015, 17, 1546–1555. [Google Scholar] [CrossRef] [Green Version]
  190. Sun, S.; Lourie, R.; Cohen, S.B.; Ji, Y.; Goodrich, J.K.; Poole, A.C.; Ley, R.E.; Denkers, E.Y.; McGuckin, M.A.; Long, Q. Epithelial Sel1L is required for the maintenance of intestinal homeostasis. Mol. Biol. Cell 2016, 27, 483–490. [Google Scholar] [CrossRef] [Green Version]
  191. Bhattacharya, A.; Sun, S.; Wang, H.; Liu, M.; Long, Q.; Yin, L.; Kersten, S.; Zhang, K.; Qi, L. Hepatic Sel1L-Hrd1 ER-associated degradation (ERAD) manages FGF21 levels and systemic metabolism via CREBH. EMBO J. 2018, 37, e99277. [Google Scholar] [CrossRef]
  192. Bhattacharya, A.; Qi, L. ER-associated degradation in health and disease–from substrate to organism. J. Cell Sci. 2019, 132, jcs232850. [Google Scholar] [CrossRef] [PubMed]
  193. Gardner, R.G.; Swarbrick, G.M.; Bays, N.W.; Cronin, S.R.; Wilhovsky, S.; Seelig, L.; Kim, C.; Hampton, R.Y. Endoplasmic reticulum degradation requires lumen to cytosol signaling: Transmembrane control of Hrd1p by Hrd3p. J. Cell Biol. 2000, 151, 69–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Carvalho, P.; Goder, V.; Rapoport, T.A. Distinct ubiquitin-ligase complexes define convergent pathways for the degradation of ER proteins. Cell 2006, 126, 361–373. [Google Scholar] [CrossRef] [Green Version]
  195. Gauss, R.; Jarosch, E.; Sommer, T.; Hirsch, C. A complex of Yos9p and the HRD ligase integrates endoplasmic reticulum quality control into the degradation machinery. Nat. Cell Biol. 2006, 8, 849–854. [Google Scholar] [CrossRef] [PubMed]
  196. Sun, S.; Shi, G.; Han, X.; Francisco, A.B.; Ji, Y.; Mendonça, N.; Liu, X.; Locasale, J.W.; Simpson, K.W.; Duhamel, G.E. Sel1L is indispensable for mammalian endoplasmic reticulum-associated degradation, endoplasmic reticulum homeostasis, and survival. Proc. Natl. Acad. Sci. USA 2014, 111, E582–E591. [Google Scholar] [CrossRef] [Green Version]
  197. Williams, J.M.; Inoue, T.; Banks, L.; Tsai, B. The ERdj5-Sel1L complex facilitates cholera toxin retrotranslocation. Mol. Biol. Cell 2013, 24, 785–795. [Google Scholar] [CrossRef]
  198. Mueller, B.; Lilley, B.N.; Ploegh, H.L. SEL1L, the homologue of yeast Hrd3p, is involved in protein dislocation from the mammalian ER. J. Cell Biol. 2006, 175, 261–270. [Google Scholar] [CrossRef] [Green Version]
  199. Mueller, B.; Klemm, E.J.; Spooner, E.; Claessen, J.H.; Ploegh, H.L. SEL1L nucleates a protein complex required for dislocation of misfolded glycoproteins. Proc. Natl. Acad. Sci. USA 2008, 105, 12325–12330. [Google Scholar] [CrossRef] [Green Version]
  200. Iida, Y.; Fujimori, T.; Okawa, K.; Nagata, K.; Wada, I.; Hosokawa, N. SEL1L protein critically determines the stability of the HRD1-SEL1L endoplasmic reticulum-associated degradation (ERAD) complex to optimize the degradation kinetics of ERAD substrates. J. Biol. Chem. 2011, 286, 16929–16939. [Google Scholar] [CrossRef] [Green Version]
  201. Klemm, E.J.; Spooner, E.; Ploegh, H.L. Dual role of ancient ubiquitous protein 1 (AUP1) in lipid droplet accumulation and endoplasmic reticulum (ER) protein quality control. J. Biol. Chem. 2011, 286, 37602–37614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Christianson, J.C.; Olzmann, J.A.; Shaler, T.A.; Sowa, M.E.; Bennett, E.J.; Richter, C.M.; Tyler, R.E.; Greenblatt, E.J.; Harper, J.W.; Kopito, R.R. Defining human ERAD networks through an integrative mapping strategy. Nat. Cell Biol. 2012, 14, 93–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Mehnert, M.; Sommer, T.; Jarosch, E. Der1 promotes movement of misfolded proteins through the endoplasmic reticulum membrane. Nat. Cell Biol. 2014, 16, 77–86. [Google Scholar] [CrossRef] [Green Version]
  204. Wahlman, J.; DeMartino, G.N.; Skach, W.R.; Bulleid, N.J.; Brodsky, J.L.; Johnson, A.E. Real-time fluorescence detection of ERAD substrate retrotranslocation in a mammalian in vitro system. Cell 2007, 129, 943–955. [Google Scholar] [CrossRef]
  205. Brodsky, J.L. Cleaning up: ER-associated degradation to the rescue. Cell 2012, 151, 1163–1167. [Google Scholar] [CrossRef] [Green Version]
  206. Baldridge, R.D.; Rapoport, T.A. Autoubiquitination of the Hrd1 ligase triggers protein retrotranslocation in ERAD. Cell 2016, 166, 394–407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Fujita, H.; Yagishita, N.; Aratani, S.; Saito-Fujita, T.; Morota, S.; Yamano, Y.; Hansson, M.J.; Inazu, M.; Kokuba, H.; Sudo, K. The E3 ligase synoviolin controls body weight and mitochondrial biogenesis through negative regulation of PGC-1β. EMBO J. 2015, 34, 1042–1055. [Google Scholar] [CrossRef] [Green Version]
  208. Yamasaki, S.; Yagishita, N.; Sasaki, T.; Nakazawa, M.; Kato, Y.; Yamadera, T.; Bae, E.; Toriyama, S.; Ikeda, R.; Zhang, L. Cytoplasmic destruction of p53 by the endoplasmic reticulum-resident ubiquitin ligase ‘Synoviolin’. EMBO J. 2007, 26, 113–122. [Google Scholar] [CrossRef]
  209. Kong, S.; Yang, Y.; Xu, Y.; Wang, Y.; Zhang, Y.; Melo-Cardenas, J.; Xu, X.; Gao, B.; Thorp, E.B.; Zhang, D.D. Endoplasmic reticulum-resident E3 ubiquitin ligase Hrd1 controls B-cell immunity through degradation of the death receptor CD95/Fas. Proc. Natl. Acad. Sci. USA 2016, 113, 10394–10399. [Google Scholar] [CrossRef] [Green Version]
  210. Yang, Y.; Kong, S.; Zhang, Y.; Melo-Cardenas, J.; Gao, B.; Zhang, Y.; Zhang, D.D.; Zhang, B.; Song, J.; Thorp, E. The endoplasmic reticulum–resident E3 ubiquitin ligase Hrd1 controls a critical checkpoint in B cell development in mice. J. Biol. Chem. 2018, 293, 12934–12944. [Google Scholar] [CrossRef] [Green Version]
  211. Shrestha, N.; Liu, T.; Ji, Y.; Reinert, R.B.; Torres, M.; Li, X.; Zhang, M.; Tang, C.-H.A.; Hu, C.-C.A.; Liu, C. Sel1L-Hrd1 ER-associated degradation maintains β cell identity via TGF-β signaling. J. Clin. Investig. 2020, 130, 3499–3510. [Google Scholar] [CrossRef] [PubMed]
  212. Huang, L.; Marvin, J.M.; Tatsis, N.; Eisenlohr, L.C. Cutting Edge: Selective role of ubiquitin in MHC class I antigen presentation. J. Immunol. 2011, 186, 1904–1908. [Google Scholar] [CrossRef]
  213. Burr, M.L.; Cano, F.; Svobodova, S.; Boyle, L.H.; Boname, J.M.; Lehner, P.J. HRD1 and UBE2J1 target misfolded MHC class I heavy chains for endoplasmic reticulum-associated degradation. Proc. Natl. Acad. Sci. USA 2011, 108, 2034–2039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Burr, M.L.; van den Boomen, D.J.; Bye, H.; Antrobus, R.; Wiertz, E.J.; Lehner, P.J. MHC class I molecules are preferentially ubiquitinated on endoplasmic reticulum luminal residues during HRD1 ubiquitin E3 ligase-mediated dislocation. Proc. Natl. Acad. Sci. USA 2013, 110, 14290–14295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Yang, H.; Qiu, Q.; Gao, B.; Kong, S.; Lin, Z.; Fang, D. Hrd1-mediated BLIMP-1 ubiquitination promotes dendritic cell MHCII expression for CD4 T cell priming during inflammation. J. Exp. Med. 2014, 211, 2467–2479. [Google Scholar] [CrossRef] [Green Version]
  216. Piskurich, J.F.; Lin, K.-I.; Lin, Y.; Wang, Y.; Ting, J.P.-Y.; Calame, K. BLIMP-1 mediates extinction of major histocompatibility class II transactivator expression in plasma cells. Nat. Immunol. 2000, 1, 526–532. [Google Scholar] [CrossRef]
  217. Xu, Y.; Fang, D. Endoplasmic reticulum-associated degradation and beyond: The multitasking roles for HRD1 in immune regulation and autoimmunity. J. Autoimmun. 2020, 109, 102423. [Google Scholar] [CrossRef]
  218. Gao, B.; Lee, S.M.; Chen, A.; Zhang, J.; Zhang, D.D.; Kannan, K.; Ortmann, R.A.; Fang, D. Synoviolin promotes IRE1 ubiquitination and degradation in synovial fibroblasts from mice with collagen-induced arthritis. EMBO Rep. 2008, 9, 480–485. [Google Scholar] [CrossRef] [Green Version]
  219. Wu, T.; Zhao, F.; Gao, B.; Tan, C.; Yagishita, N.; Nakajima, T.; Wong, P.K.; Chapman, E.; Fang, D.; Zhang, D.D. Hrd1 suppresses Nrf2-mediated cellular protection during liver cirrhosis. Genes Dev. 2014, 28, 708–722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. PONTING, C.P. Proteins of the endoplasmic-reticulum-associated degradation pathway: Domain detection and function prediction. Biochem. J. 2000, 351, 527–535. [Google Scholar] [CrossRef]
  221. Song, B.-L.; Sever, N.; DeBose-Boyd, R.A. Gp78, a membrane-anchored ubiquitin ligase, associates with Insig-1 and couples sterol-regulated ubiquitination to degradation of HMG CoA reductase. Mol. Cell 2005, 19, 829–840. [Google Scholar] [CrossRef]
  222. Chen, B.; Mariano, J.; Tsai, Y.C.; Chan, A.H.; Cohen, M.; Weissman, A.M. The activity of a human endoplasmic reticulum-associated degradation E3, gp78, requires its Cue domain, RING finger, and an E2-binding site. Proc. Natl. Acad. Sci. USA 2006, 103, 341–346. [Google Scholar] [CrossRef] [Green Version]
  223. Das, R.; Mariano, J.; Tsai, Y.C.; Kalathur, R.C.; Kostova, Z.; Li, J.; Tarasov, S.G.; McFeeters, R.L.; Altieri, A.S.; Ji, X. Allosteric activation of E2-RING finger-mediated ubiquitylation by a structurally defined specific E2-binding region of gp78. Mol. Cell 2009, 34, 674–685. [Google Scholar] [CrossRef] [Green Version]
  224. Liu, Y.; Soetandyo, N.; Lee, J.-g.; Liu, L.; Xu, Y.; Clemons Jr, W.M.; Ye, Y. USP13 antagonizes gp78 to maintain functionality of a chaperone in ER-associated degradation. eLife 2014, 3, e01369. [Google Scholar] [CrossRef]
  225. Jo, Y.; DeBose-Boyd, R.A. Control of cholesterol synthesis through regulated ER-associated degradation of HMG CoA reductase. Crit. Rev. Biochem. Mol. Biol. 2010, 45, 185–198. [Google Scholar] [CrossRef] [Green Version]
  226. Wangeline, M.A.; Vashistha, N.; Hampton, R.Y. Proteostatic tactics in the strategy of sterol regulation. Annu. Rev. Cell Dev. Biol. 2017, 33, 467–489. [Google Scholar] [CrossRef] [PubMed]
  227. Wang, Q.; Liu, X.; Cui, Y.; Tang, Y.; Chen, W.; Li, S.; Yu, H.; Pan, Y.; Wang, C. The E3 ubiquitin ligase AMFR and INSIG1 bridge the activation of TBK1 kinase by modifying the adaptor STING. Immunity 2014, 41, 919–933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Ying, Z.; Wang, H.; Fan, H.; Zhu, X.; Zhou, J.; Fei, E.; Wang, G. Gp78, an ER associated E3, promotes SOD1 and ataxin-3 degradation. Hum. Mol. Genet. 2009, 18, 4268–4281. [Google Scholar] [CrossRef] [Green Version]
  229. Yang, H.; Liu, C.; Zhong, Y.; Luo, S.; Monteiro, M.J.; Fang, S. Huntingtin interacts with the cue domain of gp78 and inhibits gp78 binding to ubiquitin and p97/VCP. PLoS ONE 2010, 5, e8905. [Google Scholar] [CrossRef] [PubMed]
  230. Shao, J.; Choe, V.; Cheng, H.; Tsai, Y.C.; Weissman, A.M.; Luo, S.; Rao, H. Ubiquitin ligase gp78 targets unglycosylated prion protein PrP for ubiquitylation and degradation. PLoS ONE 2014, 9, e92290. [Google Scholar]
  231. Foresti, O.; Ruggiano, A.; Hannibal-Bach, H.K.; Ejsing, C.S.; Carvalho, P. Sterol homeostasis requires regulated degradation of squalene monooxygenase by the ubiquitin ligase Doa10/Teb4. eLife 2013, 2, e00953. [Google Scholar] [CrossRef] [PubMed]
  232. Zavacki, A.M.; e Drigo, R.A.; Freitas, B.C.; Chung, M.; Harney, J.W.; Egri, P.; Wittmann, G.; Fekete, C.; Gereben, B.; Bianco, A.C. The E3 ubiquitin ligase TEB4 mediates degradation of type 2 iodothyronine deiodinase. Mol. Cell. Biol. 2009, 29, 5339–5347. [Google Scholar] [CrossRef] [Green Version]
  233. Zelcer, N.; Sharpe, L.J.; Loregger, A.; Kristiana, I.; Cook, E.C.; Phan, L.; Stevenson, J.; Brown, A.J. The E3 ubiquitin ligase MARCH6 degrades squalene monooxygenase and affects 3-hydroxy-3-methyl-glutaryl coenzyme A reductase and the cholesterol synthesis pathway. Mol. Cell. Biol. 2014, 34, 1262–1270. [Google Scholar] [CrossRef] [Green Version]
  234. Zattas, D.; Berk, J.M.; Kreft, S.G.; Hochstrasser, M. A conserved C-terminal element in the yeast Doa10 and human MARCH6 ubiquitin ligases required for selective substrate degradation. J. Biol. Chem. 2016, 291, 12105–12118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Meacham, G.C.; Lu, Z.; King, S.; Sorscher, E.; Tousson, A.; Cyr, D.M. The Hdj-2/Hsc70 chaperone pair facilitates early steps in CFTR biogenesis. EMBO J. 1999, 18, 1492–1505. [Google Scholar] [CrossRef]
  236. Fisher, E.A.; Zhou, M.; Mitchell, D.M.; Wu, X.; Omura, S.; Wang, H.; Goldberg, A.L.; Ginsberg, H.N. The degradation of apolipoprotein B100 is mediated by the ubiquitin-proteasome pathway and involves heat shock protein 70. J. Biol. Chem. 1997, 272, 20427–20434. [Google Scholar] [CrossRef] [Green Version]
  237. Younger, J.M.; Ren, H.-Y.; Chen, L.; Fan, C.-Y.; Fields, A.; Patterson, C.; Cyr, D.M. A foldable CFTRΔF508 biogenic intermediate accumulates upon inhibition of the Hsc70–CHIP E3 ubiquitin ligase. J. Cell Biol. 2004, 167, 1075–1085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Yamamoto, Y.-h.; Kimura, T.; Momohara, S.; Takeuchi, M.; Tani, T.; Kimata, Y.; Kadokura, H.; Kohno, K. A novel ER J-protein DNAJB12 accelerates ER-associated degradation of membrane proteins including CFTR. Cell Struct. Funct. 2010, 35, 107–116. [Google Scholar] [CrossRef] [Green Version]
  239. Grove, D.E.; Fan, C.-Y.; Ren, H.Y.; Cyr, D.M. The endoplasmic reticulum–associated Hsp40 DNAJB12 and Hsc70 cooperate to facilitate RMA1 E3–dependent degradation of nascent CFTRΔF508. Mol. Biol. Cell 2011, 22, 301–314. [Google Scholar] [CrossRef] [PubMed]
  240. van den Boomen, D.J.; Volkmar, N.; Lehner, P.J. Ubiquitin-mediated regulation of sterol homeostasis. Curr. Opin. Cell Biol. 2020, 65, 103–111. [Google Scholar] [CrossRef]
  241. Scott, N.A.; Sharpe, L.J.; Brown, A.J. The E3 ubiquitin ligase MARCHF6 as a metabolic integrator in cholesterol synthesis and beyond. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Biol. Lipids 2020, 1866, 158837. [Google Scholar] [CrossRef]
  242. Chua, N.K.; Hart-Smith, G.; Brown, A.J. Non-canonical ubiquitination of the cholesterol-regulated degron of squalene monooxygenase. J. Biol. Chem. 2019, 294, 8134–8147. [Google Scholar] [CrossRef] [Green Version]
  243. Chua, N.K.; Howe, V.; Jatana, N.; Thukral, L.; Brown, A.J. A conserved degron containing an amphipathic helix regulates the cholesterol-mediated turnover of human squalene monooxygenase, a rate-limiting enzyme in cholesterol synthesis. J. Biol. Chem. 2017, 292, 19959–19973. [Google Scholar] [CrossRef] [Green Version]
  244. Howe, V.; Chua, N.K.; Stevenson, J.; Brown, A.J. The regulatory domain of squalene monooxygenase contains a re-entrant loop and senses cholesterol via a conformational change. J. Biol. Chem. 2015, 290, 27533–27544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Nguyen, K.T.; Lee, C.-S.; Mun, S.-H.; Truong, N.T.; Park, S.K.; Hwang, C.-S. N-terminal acetylation and the N-end rule pathway control degradation of the lipid droplet protein PLIN2. J. Biol. Chem. 2019, 294, 379–388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Schultz, M.L.; Krus, K.L.; Kaushik, S.; Dang, D.; Chopra, R.; Qi, L.; Shakkottai, V.G.; Cuervo, A.M.; Lieberman, A.P. Coordinate regulation of mutant NPC1 degradation by selective ER autophagy and MARCH6-dependent ERAD. Nat. Commun. 2018, 9, 3671. [Google Scholar] [CrossRef]
  247. Wang, L.; Dong, H.; Soroka, C.J.; Wei, N.; Boyer, J.L.; Hochstrasser, M. Degradation of the bile salt export pump at endoplasmic reticulum in progressive familial intrahepatic cholestasis type II. Hepatology 2008, 48, 1558–1569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Nguyen, K.T.; Mun, S.-H.; Lee, C.-S.; Hwang, C.-S. Control of protein degradation by N-terminal acetylation and the N-end rule pathway. Exp. Mol. Med. 2018, 50, 1–8. [Google Scholar] [CrossRef]
  249. Loregger, A.; Cook, E.C.L.; Nelson, J.K.; Moeton, M.; Sharpe, L.J.; Engberg, S.; Karimova, M.; Lambert, G.; Brown, A.J.; Zelcer, N. A MARCH6 and IDOL E3 ubiquitin ligase circuit uncouples cholesterol synthesis from lipoprotein uptake in hepatocytes. Mol. Cell. Biol. 2016, 36, 285–294. [Google Scholar] [CrossRef] [Green Version]
  250. Sharpe, L.J.; Howe, V.; Scott, N.A.; Luu, W.; Phan, L.; Berk, J.M.; Hochstrasser, M.; Brown, A.J. Cholesterol increases protein levels of the E3 ligase MARCH6 and thereby stimulates protein degradation. J. Biol. Chem. 2019, 294, 2436–2448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Nakamura, N.; Harada, K.; Kato, M.; Hirose, S. Ubiquitin-specific protease 19 regulates the stability of the E3 ubiquitin ligase MARCH6. Exp. Cell Res. 2014, 328, 207–216. [Google Scholar] [CrossRef]
  252. Tcherpakov, M.; Delaunay, A.; Toth, J.; Kadoya, T.; Petroski, M.D.; Ze’ev, A.R. Regulation of endoplasmic reticulum-associated degradation by RNF5-dependent ubiquitination of JNK-associated membrane protein (JAMP). J. Biol. Chem. 2009, 284, 12099–12109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Kim, I.; Ahn, J.; Liu, C.; Tanabe, K.; Apodaca, J.; Suzuki, T.; Rao, H. The Png1–Rad23 complex regulates glycoprotein turnover. J. Cell Biol. 2006, 172, 211–219. [Google Scholar] [CrossRef]
  254. Li, G.; Zhao, G.; Zhou, X.; Schindelin, H.; Lennarz, W.J. The AAA ATPase p97 links peptide N-glycanase to the endoplasmic reticulum-associated E3 ligase autocrine motility factor receptor. Proc. Natl. Acad. Sci. USA 2006, 103, 8348–8353. [Google Scholar] [CrossRef] [Green Version]
  255. Ernst, R.; Mueller, B.; Ploegh, H.L.; Schlieker, C. The otubain YOD1 is a deubiquitinating enzyme that associates with p97 to facilitate protein dislocation from the ER. Mol. Cell 2009, 36, 28–38. [Google Scholar] [CrossRef] [Green Version]
  256. Sowa, M.E.; Bennett, E.J.; Gygi, S.P.; Harper, J.W. Defining the human deubiquitinating enzyme interaction landscape. Cell 2009, 138, 389–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Wang, Q.; Li, L.; Ye, Y. Regulation of retrotranslocation by p97-associated deubiquitinating enzyme ataxin-3. J. Cell Biol. 2006, 174, 963–971. [Google Scholar] [CrossRef]
  258. Zhong, X.; Pittman, R.N. Ataxin-3 binds VCP/p97 and regulates retrotranslocation of ERAD substrates. Hum. Mol. Genet. 2006, 15, 2409–2420. [Google Scholar] [CrossRef]
  259. Xu, Y.; Cai, M.; Yang, Y.; Huang, L.; Ye, Y. SGTA recognizes a noncanonical ubiquitin-like domain in the Bag6-Ubl4A-Trc35 complex to promote endoplasmic reticulum-associated degradation. Cell Rep. 2012, 2, 1633–1644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Xu, Y.; Liu, Y.; Lee, J.-G.; Ye, Y. A ubiquitin-like domain recruits an oligomeric chaperone to a retrotranslocation complex in endoplasmic reticulum-associated degradation. J. Biol. Chem. 2013, 288, 18068–18076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Wang, Q.; Liu, Y.; Soetandyo, N.; Baek, K.; Hegde, R.; Ye, Y. A ubiquitin ligase-associated chaperone holdase maintains polypeptides in soluble states for proteasome degradation. Mol. Cell 2011, 42, 758–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Bernales, S.; McDonald, K.L.; Walter, P. Autophagy counterbalances endoplasmic reticulum expansion during the unfolded protein response. PLoS Biol. 2006, 4, e423. [Google Scholar] [CrossRef] [Green Version]
  263. Bernales, S.; Schuck, S.; Walter, P. ER-phagy: Selective autophagy of the endoplasmic reticulum. Autophagy 2007, 3, 285–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Grumati, P.; Dikic, I.; Stolz, A. ER-phagy at a glance. J. Cell Sci. 2018, 131, jcs217364. [Google Scholar] [CrossRef] [Green Version]
  265. Khaminets, A.; Heinrich, T.; Mari, M.; Grumati, P.; Huebner, A.K.; Akutsu, M.; Liebmann, L.; Stolz, A.; Nietzsche, S.; Koch, N. Regulation of endoplasmic reticulum turnover by selective autophagy. Nature 2015, 522, 354–358. [Google Scholar] [CrossRef]
  266. Fumagalli, F.; Noack, J.; Bergmann, T.J.; Cebollero, E.; Pisoni, G.B.; Fasana, E.; Fregno, I.; Galli, C.; Loi, M.; Soldà, T. Translocon component Sec62 acts in endoplasmic reticulum turnover during stress recovery. Nat. Cell Biol. 2016, 18, 1173–1184. [Google Scholar] [CrossRef] [Green Version]
  267. Grumati, P.; Morozzi, G.; Hölper, S.; Mari, M.; Harwardt, M.-L.I.; Yan, R.; Müller, S.; Reggiori, F.; Heilemann, M.; Dikic, I. Full length RTN3 regulates turnover of tubular endoplasmic reticulum via selective autophagy. eLife 2017, 6, e25555. [Google Scholar] [CrossRef] [PubMed]
  268. Smith, M.D.; Harley, M.E.; Kemp, A.J.; Wills, J.; Lee, M.; Arends, M.; von Kriegsheim, A.; Behrends, C.; Wilkinson, S. CCPG1 is a non-canonical autophagy cargo receptor essential for ER-phagy and pancreatic ER proteostasis. Dev. Cell 2018, 44, 217–232.e211. [Google Scholar] [CrossRef]
  269. An, H.; Ordureau, A.; Paulo, J.A.; Shoemaker, C.J.; Denic, V.; Harper, J.W. TEX264 is an endoplasmic reticulum-resident ATG8-interacting protein critical for ER remodeling during nutrient stress. Mol. Cell 2019, 74, 891–908.e810. [Google Scholar] [CrossRef]
  270. Chen, Q.; Xiao, Y.; Chai, P.; Zheng, P.; Teng, J.; Chen, J. ATL3 is a tubular ER-phagy receptor for GABARAP-mediated selective autophagy. Curr. Biol. 2019, 29, 846–855.e846. [Google Scholar] [CrossRef] [Green Version]
  271. Chino, H.; Hatta, T.; Natsume, T.; Mizushima, N. Intrinsically disordered protein TEX264 mediates ER-phagy. Mol. Cell 2019, 74, 909–921.e906. [Google Scholar] [CrossRef]
  272. Wilkinson, S. ER-phagy: Shaping up and destressing the endoplasmic reticulum. FEBS J. 2019, 286, 2645–2663. [Google Scholar] [CrossRef] [Green Version]
  273. Wilkinson, S. Emerging principles of selective ER autophagy. J. Mol. Biol. 2020, 432, 185–205. [Google Scholar] [CrossRef]
  274. Axe, E.L.; Walker, S.A.; Manifava, M.; Chandra, P.; Roderick, H.L.; Habermann, A.; Griffiths, G.; Ktistakis, N.T. Autophagosome formation from membrane compartments enriched in phosphatidylinositol 3-phosphate and dynamically connected to the endoplasmic reticulum. J. Cell Biol. 2008, 182, 685–701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Hayashi-Nishino, M.; Fujita, N.; Noda, T.; Yamaguchi, A.; Yoshimori, T.; Yamamoto, A. A subdomain of the endoplasmic reticulum forms a cradle for autophagosome formation. Nat. Cell Biol. 2009, 11, 1433–1437. [Google Scholar] [CrossRef] [PubMed]
  276. Kishi-Itakura, C.; Koyama-Honda, I.; Itakura, E.; Mizushima, N. Ultrastructural analysis of autophagosome organization using mammalian autophagy-deficient cells. J. Cell Sci. 2014, 127, 4089–4102. [Google Scholar] [CrossRef] [Green Version]
  277. Chino, H.; Mizushima, N. ER-Phagy: Quality control and turnover of endoplasmic reticulum. Trends Cell Biol. 2020, 30, 384–398. [Google Scholar] [CrossRef] [PubMed]
  278. De Leonibus, C.; Cinque, L.; Settembre, C. Emerging lysosomal pathways for quality control at the endoplasmic reticulum. FEBS Lett. 2019, 593, 2319–2329. [Google Scholar] [CrossRef] [Green Version]
  279. Stolz, A.; Grumati, P. The various shades of ER-phagy. FEBS J. 2019, 286, 4642–4649. [Google Scholar] [CrossRef]
  280. Schuck, S.; Gallagher, C.M.; Walter, P. ER-phagy mediates selective degradation of endoplasmic reticulum independently of the core autophagy machinery. J. Cell Sci. 2014, 127, 4078–4088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Fregno, I.; Fasana, E.; Bergmann, T.J.; Raimondi, A.; Loi, M.; Soldà, T.; Galli, C.; D’Antuono, R.; Morone, D.; Danieli, A. ER-to-lysosome-associated degradation of proteasome-resistant ATZ polymers occurs via receptor-mediated vesicular transport. EMBO J. 2018, 37, e99259. [Google Scholar] [CrossRef]
  282. Bolender, R.P.; Weibel, E.R. A morphometric study of the removal of phenobarbital-induced membranes from hepatocytes after cessation of treatment. J. Cell Biol. 1973, 56, 746–761. [Google Scholar] [CrossRef] [PubMed]
  283. Zurek, N.; Sparks, L.; Voeltz, G. Reticulon short hairpin transmembrane domains are used to shape ER tubules. Traffic 2011, 12, 28–41. [Google Scholar] [CrossRef] [Green Version]
  284. Rismanchi, N.; Soderblom, C.; Stadler, J.; Zhu, P.-P.; Blackstone, C. Atlastin GTPases are required for Golgi apparatus and ER morphogenesis. Hum. Mol. Genet. 2008, 17, 1591–1604. [Google Scholar] [CrossRef]
  285. Bhaskara, R.M.; Grumati, P.; Garcia-Pardo, J.; Kalayil, S.; Covarrubias-Pinto, A.; Chen, W.; Kudryashev, M.; Dikic, I.; Hummer, G. Curvature induction and membrane remodeling by FAM134B reticulon homology domain assist selective ER-phagy. Nat. Commun. 2019, 10, 2370. [Google Scholar] [CrossRef] [Green Version]
  286. Brady, J.P.; Claridge, J.K.; Smith, P.G.; Schnell, J.R. A conserved amphipathic helix is required for membrane tubule formation by Yop1p. Proc. Natl. Acad. Sci. USA 2015, 112, E639–E648. [Google Scholar] [CrossRef] [Green Version]
  287. Ishida, Y.; Kubota, H.; Yamamoto, A.; Kitamura, A.; Bächinger, H.P.; Nagata, K. Type I collagen in Hsp47-null cells is aggregated in endoplasmic reticulum and deficient in N-propeptide processing and fibrillogenesis. Mol. Biol. Cell 2006, 17, 2346–2355. [Google Scholar] [CrossRef]
  288. Forrester, A.; De Leonibus, C.; Grumati, P.; Fasana, E.; Piemontese, M.; Staiano, L.; Fregno, I.; Raimondi, A.; Marazza, A.; Bruno, G. A selective ER-phagy exerts procollagen quality control via a Calnexin-FAM 134B complex. EMBO J. 2019, 38, e99847. [Google Scholar] [CrossRef]
  289. Gelsthorpe, M.E.; Baumann, N.; Millard, E.; Gale, S.E.; Langmade, S.J.; Schaffer, J.E.; Ory, D.S. Niemann-Pick type C1 I1061T mutant encodes a functional protein that is selected for endoplasmic reticulum-associated degradation due to protein misfolding. J. Biol. Chem. 2008, 283, 8229–8236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Bienkowski, R.S.; Gotkin, M.G. Control of collagen deposition in mammalian lung. Proc. Soc. Exp. Biol. Med. 1995, 209, 118–140. [Google Scholar] [CrossRef] [PubMed]
  291. Yang, Y.S.; Strittmatter, S.M. The reticulons: A family of proteins with diverse functions. Genome Biol. 2007, 8, 234. [Google Scholar] [CrossRef] [Green Version]
  292. Cunningham, C.N.; Williams, J.M.; Knupp, J.; Arunagiri, A.; Arvan, P.; Tsai, B. Cells deploy a two-pronged strategy to rectify misfolded proinsulin aggregates. Mol. Cell 2019, 75, 442–456.e444. [Google Scholar] [CrossRef]
  293. Busch, D.J.; Houser, J.R.; Hayden, C.C.; Sherman, M.B.; Lafer, E.M.; Stachowiak, J.C. Intrinsically disordered proteins drive membrane curvature. Nat. Commun. 2015, 6, 7875. [Google Scholar] [CrossRef] [Green Version]
  294. Snead, W.T.; Hayden, C.C.; Gadok, A.K.; Zhao, C.; Lafer, E.M.; Rangamani, P.; Stachowiak, J.C. Membrane fission by protein crowding. Proc. Natl. Acad. Sci. USA 2017, 114, E3258–E3267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Yang, H.; Ni, H.-M.; Guo, F.; Ding, Y.; Shi, Y.-H.; Lahiri, P.; Fröhlich, L.F.; Rülicke, T.; Smole, C.; Schmidt, V.C. Sequestosome 1/p62 protein is associated with autophagic removal of excess hepatic endoplasmic reticulum in mice. J. Biol. Chem. 2016, 291, 18663–18674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Ji, C.H.; Kim, H.Y.; Heo, A.J.; Lee, S.H.; Lee, M.J.; Kim, S.B.; Srinivasrao, G.; Mun, S.R.; Cha-Molstad, H.; Ciechanover, A. The N-Degron pathway mediates ER-phagy. Mol. Cell 2019, 75, 1058–1072.e1059. [Google Scholar] [CrossRef]
  297. Roberts, P.; Moshitch-Moshkovitz, S.; Kvam, E.; O’Toole, E.; Winey, M.; Goldfarb, D.S. Piecemeal microautophagy of nucleus in Saccharomyces cerevisiae. Mol. Biol. Cell 2003, 14, 129–141. [Google Scholar] [CrossRef] [Green Version]
  298. Schäfer, J.A.; Schessner, J.P.; Bircham, P.W.; Tsuji, T.; Funaya, C.; Pajonk, O.; Schaeff, K.; Ruffini, G.; Papagiannidis, D.; Knop, M. ESCRT machinery mediates selective microautophagy of endoplasmic reticulum in yeast. EMBO J. 2020, 39, e102586. [Google Scholar] [CrossRef]
  299. Loi, M.; Raimondi, A.; Morone, D.; Molinari, M. ESCRT-III-driven piecemeal micro-ER-phagy remodels the ER during recovery from ER stress. Nat. Commun. 2019, 10, 5058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Kurokawa, K.; Nakano, A. The ER exit sites are specialized ER zones for the transport of cargo proteins from the ER to the Golgi apparatus. J. Biochem. 2019, 165, 109–114. [Google Scholar] [CrossRef] [Green Version]
  301. Omari, S.; Makareeva, E.; Roberts-Pilgrim, A.; Mirigian, L.; Jarnik, M.; Ott, C.; Lippincott-Schwartz, J.; Leikin, S. Noncanonical autophagy at ER exit sites regulates procollagen turnover. Proc. Natl. Acad. Sci. USA 2018, 115, E10099–E10108. [Google Scholar] [CrossRef] [Green Version]
  302. Itakura, E.; Kishi-Itakura, C.; Mizushima, N. The hairpin-type tail-anchored SNARE syntaxin 17 targets to autophagosomes for fusion with endosomes/lysosomes. Cell 2012, 151, 1256–1269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Kamimoto, T.; Shoji, S.; Hidvegi, T.; Mizushima, N.; Umebayashi, K.; Perlmutter, D.H.; Yoshimori, T. Intracellular inclusions containing mutant α1-antitrypsin Z are propagated in the absence of autophagic activity. J. Biol. Chem. 2006, 281, 4467–4476. [Google Scholar] [CrossRef] [Green Version]
  304. Kroeger, H.; Miranda, E.; MacLeod, I.; Pérez, J.; Crowther, D.C.; Marciniak, S.J.; Lomas, D.A. Endoplasmic reticulum-associated degradation (ERAD) and autophagy cooperate to degrade polymerogenic mutant serpins. J. Biol. Chem. 2009, 284, 22793–22802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Hidvegi, T.; Ewing, M.; Hale, P.; Dippold, C.; Beckett, C.; Kemp, C.; Maurice, N.; Mukherjee, A.; Goldbach, C.; Watkins, S. An autophagy-enhancing drug promotes degradation of mutant α1-antitrypsin Z and reduces hepatic fibrosis. Science 2010, 329, 229–232. [Google Scholar] [CrossRef]
  306. Colla, E. Linking the endoplasmic reticulum to Parkinson’s disease and alpha-synucleinopathy. Front. Neurosci. 2019, 13, 560. [Google Scholar] [CrossRef] [Green Version]
  307. Costa, C.A.d.; Manaa, W.E.; Duplan, E.; Checler, F. The Endoplasmic Reticulum Stress/Unfolded Protein Response and Their Contributions to Parkinson’s Disease Physiopathology. Cells 2020, 9, 2495. [Google Scholar] [CrossRef] [PubMed]
  308. Liu, E.A.; Lieberman, A.P. The intersection of lysosomal and endoplasmic reticulum calcium with autophagy defects in lysosomal diseases. Neurosci. Lett. 2019, 697, 10–16. [Google Scholar] [CrossRef]
  309. Behl, T.; Kaur, G.; Fratila, O.; Buhas, C.; Judea-Pusta, C.T.; Negrut, N.; Bustea, C.; Bungau, S. Cross-talks among GBA mutations, glucocerebrosidase, and α-synuclein in GBA-associated Parkinson’s disease and their targeted therapeutic approaches: A comprehensive review. Transl. Neurodegener. 2021, 10, 4. [Google Scholar] [CrossRef] [PubMed]
  310. Ghemrawi, R.; Khair, M. Endoplasmic reticulum stress and unfolded protein response in neurodegenerative diseases. Int. J. Mol. Sci. 2020, 21, 6127. [Google Scholar] [CrossRef] [PubMed]
  311. Goedert, M.; Spillantini, M.G.; Del Tredici, K.; Braak, H. 100 years of Lewy pathology. Nat. Rev. Neurol. 2013, 9, 13. [Google Scholar] [CrossRef]
  312. Smith, W.W.; Jiang, H.; Pei, Z.; Tanaka, Y.; Morita, H.; Sawa, A.; Dawson, V.L.; Dawson, T.M.; Ross, C.A. Endoplasmic reticulum stress and mitochondrial cell death pathways mediate A53T mutant alpha-synuclein-induced toxicity. Hum. Mol. Genet. 2005, 14, 3801–3811. [Google Scholar] [CrossRef] [PubMed]
  313. Cooper, A.A.; Gitler, A.D.; Cashikar, A.; Haynes, C.M.; Hill, K.J.; Bhullar, B.; Liu, K.; Xu, K.; Strathearn, K.E.; Liu, F. α-Synuclein blocks ER-Golgi traffic and Rab1 rescues neuron loss in Parkinson’s models. Science 2006, 313, 324–328. [Google Scholar] [CrossRef] [Green Version]
  314. Bellucci, A.; Navarria, L.; Zaltieri, M.; Falarti, E.; Bodei, S.; Sigala, S.; Battistin, L.; Spillantini, M.; Missale, C.; Spano, P. Induction of the unfolded protein response by α-synuclein in experimental models of Parkinson’s disease. J. Neurochem. 2011, 116, 588–605. [Google Scholar] [CrossRef]
  315. Colla, E.; Coune, P.; Liu, Y.; Pletnikova, O.; Troncoso, J.C.; Iwatsubo, T.; Schneider, B.L.; Lee, M.K. Endoplasmic reticulum stress is important for the manifestations of α-synucleinopathy in vivo. J. Neurosci. 2012, 32, 3306–3320. [Google Scholar] [CrossRef] [PubMed]
  316. Chung, C.Y.; Khurana, V.; Auluck, P.K.; Tardiff, D.F.; Mazzulli, J.R.; Soldner, F.; Baru, V.; Lou, Y.; Freyzon, Y.; Cho, S. Identification and rescue of α-synuclein toxicity in Parkinson patient–derived neurons. Science 2013, 342, 983–987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Heman-Ackah, S.M.; Manzano, R.; Hoozemans, J.J.; Scheper, W.; Flynn, R.; Haerty, W.; Cowley, S.A.; Bassett, A.R.; Wood, M.J. Alpha-synuclein induces the unfolded protein response in Parkinson’s disease SNCA triplication iPSC-derived neurons. Hum. Mol. Genet. 2017, 26, 4441–4450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Song, J.; Kim, B.C.; Nguyen, D.-T.T.; Samidurai, M.; Choi, S.-M. Levodopa (L-DOPA) attenuates endoplasmic reticulum stress response and cell death signaling through DRD2 in SH-SY5Y neuronal cells under α-synuclein-induced toxicity. Neuroscience 2017, 358, 336–348. [Google Scholar] [CrossRef]
  319. Bellani, S.; Mescola, A.; Ronzitti, G.; Tsushima, H.; Tilve, S.; Canale, C.; Valtorta, F.; Chieregatti, E. GRP78 clustering at the cell surface of neurons transduces the action of exogenous alpha-synuclein. Cell Death Differ. 2014, 21, 1971–1983. [Google Scholar] [CrossRef]
  320. Gitler, A.D.; Bevis, B.J.; Shorter, J.; Strathearn, K.E.; Hamamichi, S.; Su, L.J.; Caldwell, K.A.; Caldwell, G.A.; Rochet, J.-C.; McCaffery, J.M. The Parkinson’s disease protein α-synuclein disrupts cellular Rab homeostasis. Proc. Natl. Acad. Sci. USA 2008, 105, 145–150. [Google Scholar] [CrossRef] [Green Version]
  321. Mazzulli, J.R.; Zunke, F.; Isacson, O.; Studer, L.; Krainc, D. α-Synuclein–induced lysosomal dysfunction occurs through disruptions in protein trafficking in human midbrain synucleinopathy models. Proc. Natl. Acad. Sci. USA 2016, 113, 1931–1936. [Google Scholar] [CrossRef] [Green Version]
  322. Betzer, C.; Lassen, L.B.; Olsen, A.; Kofoed, R.H.; Reimer, L.; Gregersen, E.; Zheng, J.; Calì, T.; Gai, W.P.; Chen, T. Alpha-synuclein aggregates activate calcium pump SERCA leading to calcium dysregulation. EMBO Rep. 2018, 19, e44617. [Google Scholar] [CrossRef]
  323. Jung, E.-M.; Yoo, Y.-M.; Park, S.Y.; Ahn, C.; Jeon, B.-H.; Hong, E.-J.; Kim, W.-Y.; Jeung, E.-B. Calbindin-D 9k Is a Novel Risk Gene for Neurodegenerative Disease. Cell. Physiol. Biochem. Int. J. Exp. Cell. Physiol. Biochem. Pharmacol. 2020, 54, 438–456. [Google Scholar]
  324. Shimura, H.; Hattori, N.; Kubo, S.-i.; Mizuno, Y.; Asakawa, S.; Minoshima, S.; Shimizu, N.; Iwai, K.; Chiba, T.; Tanaka, K. Familial Parkinson disease gene product, parkin, is a ubiquitin-protein ligase. Nat. Genet. 2000, 25, 302–305. [Google Scholar] [CrossRef] [PubMed]
  325. Imai, Y.; Soda, M.; Takahashi, R. Parkin suppresses unfolded protein stress-induced cell death through its E3 ubiquitin-protein ligase activity. J. Biol. Chem. 2000, 275, 35661–35664. [Google Scholar] [CrossRef] [Green Version]
  326. Imai, Y.; Soda, M.; Inoue, H.; Hattori, N.; Mizuno, Y.; Takahashi, R. An unfolded putative transmembrane polypeptide, which can lead to endoplasmic reticulum stress, is a substrate of Parkin. Cell 2001, 105, 891–902. [Google Scholar] [CrossRef] [Green Version]
  327. Rana, A.; Rera, M.; Walker, D.W. Parkin overexpression during aging reduces proteotoxicity, alters mitochondrial dynamics, and extends lifespan. Proc. Natl. Acad. Sci. USA 2013, 110, 8638–8643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  328. Omura, T.; Kaneko, M.; Okuma, Y.; Orba, Y.; Nagashima, K.; Takahashi, R.; Fujitani, N.; Matsumura, S.; Hata, A.; Kubota, K. A ubiquitin ligase HRD1 promotes the degradation of Pael receptor, a substrate of Parkin. J. Neurochem. 2006, 99, 1456–1469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  329. Meikle, P.J.; Hopwood, J.J.; Clague, A.E.; Carey, W.F. Prevalence of lysosomal storage disorders. JAMA 1999, 281, 249–254. [Google Scholar] [CrossRef] [PubMed]
  330. Korkotian, E.; Schwarz, A.; Pelled, D.; Schwarzmann, G.; Segal, M.; Futerman, A.H. Elevation of intracellular glucosylceramide levels results in an increase in endoplasmic reticulum density and in functional calcium stores in cultured neurons. J. Biol. Chem. 1999, 274, 21673–21678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  331. Lloyd-Evans, E.; Pelled, D.; Riebeling, C.; Bodennec, J.; de-Morgan, A.; Waller, H.; Schiffmann, R.; Futerman, A.H. Glucosylceramide and glucosylsphingosine modulate calcium mobilization from brain microsomes via different mechanisms. J. Biol. Chem. 2003, 278, 23594–23599. [Google Scholar] [CrossRef] [Green Version]
  332. Sun, Y.; Liou, B.; Ran, H.; Skelton, M.R.; Williams, M.T.; Vorhees, C.V.; Kitatani, K.; Hannun, Y.A.; Witte, D.P.; Xu, Y.-H. Neuronopathic Gaucher disease in the mouse: Viable combined selective saposin C deficiency and mutant glucocerebrosidase (V394L) mice with glucosylsphingosine and glucosylceramide accumulation and progressive neurological deficits. Hum. Mol. Genet. 2010, 19, 1088–1097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Liou, B.; Peng, Y.; Li, R.; Inskeep, V.; Zhang, W.; Quinn, B.; Dasgupta, N.; Blackwood, R.; Setchell, K.D.; Fleming, S. Modulating ryanodine receptors with dantrolene attenuates neuronopathic phenotype in Gaucher disease mice. Hum. Mol. Genet. 2016, 25, 5126–5141. [Google Scholar] [CrossRef]
  334. Zimmer, K.P.; le Coutre, P.; Aerts, H.M.; Harzer, K.; Fukuda, M.; O’Brien, J.S.; Naim, H.Y. Intracellular transport of acid β-glucosidase and lysosome-associated membrane proteins is affected in Gaucher’s disease (G202R mutation). J. Pathol. 1999, 188, 407–414. [Google Scholar] [CrossRef]
  335. Sawkar, A.R.; Cheng, W.-C.; Beutler, E.; Wong, C.-H.; Balch, W.E.; Kelly, J.W. Chemical chaperones increase the cellular activity of N370S β-glucosidase: A therapeutic strategy for Gaucher disease. Proc. Natl. Acad. Sci. USA 2002, 99, 15428–15433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Ron, I.; Horowitz, M. ER retention and degradation as the molecular basis underlying Gaucher disease heterogeneity. Hum. Mol. Genet. 2005, 14, 2387–2398. [Google Scholar] [CrossRef] [Green Version]
  337. Chu, C.T.; Ji, J.; Dagda, R.K.; Jiang, J.F.; Tyurina, Y.Y.; Kapralov, A.A.; Tyurin, V.A.; Yanamala, N.; Shrivastava, I.H.; Mohammadyani, D. Cardiolipin externalization to the outer mitochondrial membrane acts as an elimination signal for mitophagy in neuronal cells. Nat. Cell Biol. 2013, 15, 1197–1205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  338. Preston, G.M.; Brodsky, J.L. The evolving role of ubiquitin modification in endoplasmic reticulum-associated degradation. Biochem. J. 2017, 474, 445–469. [Google Scholar] [CrossRef] [Green Version]
  339. Anding, A.L.; Baehrecke, E.H. Cleaning house: Selective autophagy of organelles. Dev. Cell 2017, 41, 10–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The ubiquitination machinery. Ubiquitination utilizes a three-step enzymatic cascade. Ubiquitin is activated by the E1 ubiquitin-activating enzyme at the expense of ATP and subsequently conjugated to the E1 ubiquitin-activating enzyme via thioester bond. The E1 ubiquitin-activating enzyme then associates with the E2 ubiquitin-conjugating enzyme and transfers ubiquitin from its catalytic cysteine to the active-site cysteine of the E2 ubiquitin-conjugating enzyme to form a thioester bond. Depending on the E3 ubiquitin ligase involved, ubiquitin linked to the E2 ubiquitin-conjugating enzyme can be transferred and subsequently conjugated to the substrate protein via at least two mechanisms. Homologous to the E6-associated protein (E6AP) carboxyl terminus (HECT)-type and really interesting new gene (RING)-in-between-RINGs (IBR)-RING (RBR)-type E3 ubiquitin ligases mediate ubiquitin transfer to the substrate proteins via a two-step reaction, in which ubiquitin is transferred to an active-site cysteine on the E3 ligase, resulting in the formation of a thioester with ubiquitin, and then to the substrate. RING-type E3 ligases catalyze a direct transfer of ubiquitin from the ubiquitin-charged E2 conjugating enzyme to the substrate proteins.
Figure 1. The ubiquitination machinery. Ubiquitination utilizes a three-step enzymatic cascade. Ubiquitin is activated by the E1 ubiquitin-activating enzyme at the expense of ATP and subsequently conjugated to the E1 ubiquitin-activating enzyme via thioester bond. The E1 ubiquitin-activating enzyme then associates with the E2 ubiquitin-conjugating enzyme and transfers ubiquitin from its catalytic cysteine to the active-site cysteine of the E2 ubiquitin-conjugating enzyme to form a thioester bond. Depending on the E3 ubiquitin ligase involved, ubiquitin linked to the E2 ubiquitin-conjugating enzyme can be transferred and subsequently conjugated to the substrate protein via at least two mechanisms. Homologous to the E6-associated protein (E6AP) carboxyl terminus (HECT)-type and really interesting new gene (RING)-in-between-RINGs (IBR)-RING (RBR)-type E3 ubiquitin ligases mediate ubiquitin transfer to the substrate proteins via a two-step reaction, in which ubiquitin is transferred to an active-site cysteine on the E3 ligase, resulting in the formation of a thioester with ubiquitin, and then to the substrate. RING-type E3 ligases catalyze a direct transfer of ubiquitin from the ubiquitin-charged E2 conjugating enzyme to the substrate proteins.
Ijms 22 02078 g001
Figure 2. ER-associated degradation (ERAD). suppressor/enhancer of Lin12-like (SEL1L)/hydroxymethylglutaryl reductase degradation protein 1 (HRD1)-mediated ERAD is the most conserved ERAD complex in mammals. Substrate recognition: misfolded proteins are recognized by substrate recognition factors, including osteosarcoma 9 (OS-9), binding immunoglobulin protein (BiP), ER degradation-enhancing α-mannosidase like protein (EDEM), and XTP3-transactivated gene B precursor (XTP3-B). Retrotranslocation and ubiquitination: once the misfolded protein is recognized, the misfolded protein is subject to retrotranslocation and ubiquitination. Following retrotranslocation and ubiquitination, the misfolded substrate is eventually guided to the proteasome with the aid of p97/valosin-containing protein (VCP), thereby leading to its proteasomal degradation.
Figure 2. ER-associated degradation (ERAD). suppressor/enhancer of Lin12-like (SEL1L)/hydroxymethylglutaryl reductase degradation protein 1 (HRD1)-mediated ERAD is the most conserved ERAD complex in mammals. Substrate recognition: misfolded proteins are recognized by substrate recognition factors, including osteosarcoma 9 (OS-9), binding immunoglobulin protein (BiP), ER degradation-enhancing α-mannosidase like protein (EDEM), and XTP3-transactivated gene B precursor (XTP3-B). Retrotranslocation and ubiquitination: once the misfolded protein is recognized, the misfolded protein is subject to retrotranslocation and ubiquitination. Following retrotranslocation and ubiquitination, the misfolded substrate is eventually guided to the proteasome with the aid of p97/valosin-containing protein (VCP), thereby leading to its proteasomal degradation.
Ijms 22 02078 g002
Figure 3. Different types of ER-phagy. In macro-ER-phagy, fragments of the ER together with cytoplasmic components are enclosed by autophagosome, which fuses with lysosomes and is subsequently degraded. In micro-ER-phagy, a small portion of the ER is invaginated by lysosomes and subsequently degraded. In vesicular delivery, ER-derived vesicles containing misfolded proteins bud off from the ER and fuse directly with lysosomes.
Figure 3. Different types of ER-phagy. In macro-ER-phagy, fragments of the ER together with cytoplasmic components are enclosed by autophagosome, which fuses with lysosomes and is subsequently degraded. In micro-ER-phagy, a small portion of the ER is invaginated by lysosomes and subsequently degraded. In vesicular delivery, ER-derived vesicles containing misfolded proteins bud off from the ER and fuse directly with lysosomes.
Ijms 22 02078 g003
Table 1. List of ERAD E3 ubiquitin ligases and their identified substrates.
Table 1. List of ERAD E3 ubiquitin ligases and their identified substrates.
E3 Ubiquitin Ligases (Mammalian)SubstratesNotes
HRD1pre-BCR, BLIMP1, Fas, IRE1, NRF2, p53, PGC1β, AVP, TGF-β receptor I, TCR-α, MHC IHomolog of yeast Hrd1
MARCHF6SQLE, HMGCR, PLIN2, BSEPHomolog of yeast Doa10
gp78CFTR, HMGCR, SOD1, ataxin-3, huntingtin, prionHomolog of yeast Hrd1
RNF5CFTR, JAMPIntegral ER membrane-resident
CHIPCFTR, Pael-receptorCytoplasmic
PARKINPael-receptor, GCaseCytoplasmic
Smurf1WFS1Cytoplasmic/nuclear
Nrdp1ErbB3Cytoplasmic
SCFFbx2TCR-α, CFTRCytoplasmic/ER membrane-associated
SCFFbx6TCR-αCytoplasmic/ER membrane-associated
SCFβ-TrCPCD4, TetherinCytoplasmic
Hydroxymethylglutaryl reductase degradation protein 1 (HRD1); membrane-associated RING C3HC4 finger 6 (MARCHF6), glycoprotein 78 (gp78); RING finger protein 5 (RNF5); C-terminus of the Hsc70-interacting protein (CHIP); SMAD ubiquitination regulatory factor 1 (Smurf1); neuregulin receptor degradation pathway protein 1 (Nrdp1); Skp, Cullin and F-box containing complex (SCF); pre-B cell receptor (pre-BCR); B-lymphocyte-induced maturation protein 1 (BLIMP1); inositol-requiring enzyme 1 (IRE1); NF-E2-related factor 2 (NRF2);peroxisome proliferator activated receptor γ coactivator-1 β (PGC1β); pro-arginine vasopressin (AVP); transforming growth factor beta (TGF-β); T-cell receptor alpha (TCR-α); major histocompatibility complex I (MHC I); squalene epoxidase (SQLE); 3-hydroxy-3-methyl-glutaryl-coenzyme A (CoA) reductase (HMGCR); perilipin-2 (PLIN2); bile salt export pump (BSEP); cystic fibrosis transmembrane conductance regulator (CFTR); superoxide dismutase-1 (SOD1); JNK-associated membrane protein (JAMP); glucocerebrosidase (GCase); Erb-B2 receptor tyrosine kinase 3 (ErbB3).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kang, J.A.; Jeon, Y.J. How Is the Fidelity of Proteins Ensured in Terms of Both Quality and Quantity at the Endoplasmic Reticulum? Mechanistic Insights into E3 Ubiquitin Ligases. Int. J. Mol. Sci. 2021, 22, 2078. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22042078

AMA Style

Kang JA, Jeon YJ. How Is the Fidelity of Proteins Ensured in Terms of Both Quality and Quantity at the Endoplasmic Reticulum? Mechanistic Insights into E3 Ubiquitin Ligases. International Journal of Molecular Sciences. 2021; 22(4):2078. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22042078

Chicago/Turabian Style

Kang, Ji An, and Young Joo Jeon. 2021. "How Is the Fidelity of Proteins Ensured in Terms of Both Quality and Quantity at the Endoplasmic Reticulum? Mechanistic Insights into E3 Ubiquitin Ligases" International Journal of Molecular Sciences 22, no. 4: 2078. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22042078

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop