specificity and regulation of the endoplasmic reticulum-associated degradation machinery

23
Specificity and Regulation of the Endoplasmic Reticulum-Associated Degradation Machinery Jessica Merulla 1,2 , Elisa Fasana 1 , Tatiana Soldà 1 and Maurizio Molinari 1,3* 1 Institute for Research in Biomedicine, Protein Folding and Quality Control, CH-6500 Bellinzona, Switzerland 2 Graduate School for Cellular and Biomedical Sciences, University of Bern, Switzerland 3 Ecole Polytechnique Fédérale de Lausanne, School of Life Sciences, CH-1015 Lausanne, Switzerland Correspondence to: [email protected] Keywords Chemical and pharmacological chaperones Conformational diseases Defective ribosomal products (DRiPS) Endoplasmic reticulum ER-associated degradation (ERAD) ERAD tuning ER stress Hijacking by pathogens Misfolded proteins Proteostasis Ubiquitin proteasome system Unfolded protein response (UPR) This article has been accepted for publication and undergone full peer review but has not been through the copyediting, typesetting, pagination and proofreading process, which may lead to differences between this version and the Version of Record. Please cite this article as doi: 10.1111/tra.12068 © 2013 John Wiley & Sons A/S

Upload: maurizio

Post on 08-Dec-2016

215 views

Category:

Documents


3 download

TRANSCRIPT

Specificity and Regulation of the Endoplasmic Reticulum-Associated Degradation Machinery

Jessica Merulla1,2, Elisa Fasana1, Tatiana Soldà1 and Maurizio Molinari1,3*

1Institute for Research in Biomedicine, Protein Folding and Quality Control,

CH-6500 Bellinzona, Switzerland 2Graduate School for Cellular and Biomedical Sciences, University of Bern, Switzerland

3Ecole Polytechnique Fédérale de Lausanne, School of Life Sciences, CH-1015 Lausanne,

Switzerland

Correspondence to: [email protected]

Keywords Chemical and pharmacological chaperones Conformational diseases Defective ribosomal products (DRiPS) Endoplasmic reticulum ER-associated degradation (ERAD) ERAD tuning ER stress Hijacking by pathogens Misfolded proteins Proteostasis Ubiquitin proteasome system Unfolded protein response (UPR)

This article has been accepted for publication and undergone full peer review but has not been through the copyediting, typesetting, pagination and proofreading process, which may lead to differences between this version and the Version of Record. Please cite this article as doi: 10.1111/tra.12068

© 2013 John Wiley & Sons A/S

Abstarct

The endoplasmic reticulum-associated degradation (ERAD) machinery selects native and misfolded polypeptides

for dislocation across the ER membrane and proteasomal degradation. Regulated degradation of native proteins

is an important aspect of cell physiology. For example, it contributes to the control of lipid biosynthesis, calcium

homeostasis and ERAD capacity by setting the turnover rate of crucial regulators of these pathways. In contrast,

degradation of native proteins has pathologic relevance when caused by viral or bacterial infections, or when it

occurs as a consequence of dysregulated ERAD activity. The efficient disposal of misfolded proteins prevents

toxic depositions and persistent sequestration of molecular chaperones that could induce cellular stress and

perturb maintenance of cellular proteostasis.

In the first section of this review, we survey the available literature on mechanisms of selection of native and non-

native proteins for degradation from the ER and on how pathogens hijack them. In the second section, we

highlight the mechanisms of ERAD activity adaptation to changes in the ER environment with a particular

emphasis on the post-translational regulatory mechanisms collectively defined as ERAD tuning.

Introduction The cellular proteome is mostly synthesized by

cytosolic ribosomes to operate in the cytosol and,

upon appropriate targeting, in various intracellular

organelles or in the extracellular space. Cellular

compartments where protein folding occurs (e.g.,

the cytosol, mitochondria, the endoplasmic

reticulum (ER)) contain two classes of non-native

polypeptides: i) newly synthesized polypeptide

chains that must be assisted by folding chaperones

and enzymes to attain the native mono- or

oligomeric structure; ii) terminally misfolded

conformers that must efficiently be degraded to

prevent the formation of toxic deposits and the

persistent sequestration of chaperones that could

eventually inhibit the cellular protein folding capacity

and elicit stress (1-3) (Fig. 1A). The distinction

between the two classes of non-native chains, one

to be preserved, the second to be cleared from the

folding compartment is not an easy task for the

cellular quality control machineries. Selection for

disposal might be a stochastic process: the longer

the persistency of structural defects in the ER, the

greater is the probability to be selected for

destruction. Therefore, mutations that delay folding

may channel the polypeptide into destructive

pathways, even if they do not compromise the

function of the mutated protein.

In the ER, non-native, but also native proteins are

selected for degradation by components of the

ERAD machinery that deliver them at dislocation

sites embedded in the ER membrane. Dislocation

sites consist of a multitude of luminal and

membrane-bound specialized ER-resident proteins

as well as a number of cytosolic factors insuring

dislocation across the ER membrane, poly-

ubiquitylation and disposal of ERAD substrates by

26S proteasomes (Section A, Specificity of the

ERAD machinery). Dysregulated ERAD activity

caused by excessive luminal content of ERAD

factors may result in premature interruption of

ongoing folding attempts where proteins that would

have attained functional structures are

inappropriately destroyed thus causing loss-of-

function phenotypes or disorders (Fig. 1B). All in all,

balanced activity of the degradation machinery

crucially contributes to maintenance of protein

homeostasis (proteostasis) (Section B, Regulation

of the ERAD activity).

A. Specificity of the ERAD machinery

ERAD of native proteins in healthy cells: select

examples

The regulated degradation of functional proteins is a

crucial aspect of cellular physiology and contributes,

for example, to set the level of lipid biosynthesis, to

adjust calcium homeostasis, to determine the

constitutive level of the ERAD activity and to adapt

it to variations in misfolded proteins load (4, 5).

The ERAD machinery controls the intracellular

content of the 3-hydroxy-3-methyl-glutaryl-CoA

reductase (HMG-CoA reductase), a rate-limiting

enzyme in the synthesis of sterols from acetyl-CoA.

The HMG-CoA reductase turnover is enhanced

when the cellular sterol level is high, a classical

example of feedback regulation (6-8). Likewise,

ERAD factors may select the lipid carrier ApoB (9,

10) or the IP3P receptor (11) for destruction when

lipid levels are low or when cytosolic levels of IP3

and calcium are high, respectively.

Poorly studied, but crucial to maintain cellular

proteostasis, are the constitutive and the regulated

degradation of ERAD factors (5). ERAD factors

turnover may directly be affected by the presence of

misfolded polypeptides in the ER lumen and

contributes in determining the rapid adaptation of

the overall ERAD activity to changes in the ER

folding environment as described in Section B.

Degradation of native proteins triggered by

pathogens: select examples

Viruses and bacteria may dysregulate ERAD

pathways of infected cells to enhance their chances

to survive in the host, to generate their progeny and

to escape immunosurveillance (12, 13). This

highlights the impact that ERAD modulation may

have on cellular physiology and on protein

production.

The human cytomegalovirus (HCMV): One of the

most revealing examples of manipulation of the

equilibrium existing in eukaryotic cells between

maturation and disposal of newly synthesized

polypeptides relates to cells infected with HCMV.

These cells produce the viral proteins US2 and

US11 that select for disposal newly synthesized

class I or class II MHC (14-16). The intervention of

US2 and US11 anticipates the assembly of

complexes that bind viral epitopes in the ER lumen

and transport them at the cell surface for activation

of the host immune response. In infected cells,

class I and class II molecules are cleared from the

ER with intervention of conventional ERAD factors

(thoroughly described below) such as SEL1L, derlin

1, VIMP, p97/VCP and the E3 ubiquitin ligase TRC8

(17-19). Infected cells may also express the viral

protein Pp65 to inhibit proteasomal activity (20), a

strategy that HMCV shares with several other

viruses (Epstein-Barr virus (21), human

immunodeficiency virus (HIV) (22), hepatitis B virus

(HBV) (23), Kaposi sarcoma herpes virus (24)) to

elude immunosurveillance.

HIV: Cells infected with HIV express the viral

protein Vpu causing the proteasome-mediated

degradation of the viral receptor CD4 thereby

preventing superinfection (25-28).

The murine γ-Herpes virus 68 and the rodent

herpes virus Peru (RHVP): The expression of viral

E3 ubiquitin ligases is yet another strategy used by

pathogens to specifically clear host cell proteins that

could activate the immune response. The viral E3

ligase mK3 expressed in cells infected with murine

γ-Herpes virus 68 binds to and uses the

TAP/Tapasin complex as an adaptor to specifically

polyubiquitylate and destroy the class I MHC heavy

chain (29-32). Interestingly, the ubiquitin ligase

activity of mK3 is directed to serine and threonine in

addition to the conventional lysine residues (33).

The viral pK3 expressed in cells infected with RHVP

promotes proteasome-mediated degradation of the

class I MHC heavy chain (34).

The hepatitis C (HCV) and HBV viruses: Another

example of manipulation of the host cell ERAD

activity during the viral life cycle is reported for cells

infected with HCV (35) or HBV (36). ERAD

enhancement resulting from the ER stress-induced

up-regulation of proteins of the EDEM family (37)

reduces the fraction of newly synthesized viral

glycoproteins able to attain the native structure in

due time. Consistent with data showing that ERAD

enhancement interferes with completion of ongoing

folding programs (38-41), folding and assembly of

viral spikes becomes inefficient. This is proposed to

maintain low production of viral particles thus

contributing to infection persistency.

Degradation of misfolded proteins: specificity of

the ERAD machinery

Genomic defects, transcriptional and translational

errors may alter the amino acid composition or the

length of the polypeptide sequence. This may

prevent attainment of the native structure or delay

the folding process such that selection for disposal

of the mutated chains precedes completion of the

folding program. Intra- or extracellular accumulation

of folding-defective polypeptides causes gain-of-

toxic-function diseases, for example hereditary lung

emphysema and liver disease caused by mutant

forms of α1-antitrypsin or many types of

neurodegenerative disorders. On the other hand,

clearance of the aberrant gene products can result

in loss-of-function diseases, for example cystic

fibrosis or lysosomal storage diseases (thoroughly

reviewed elsewhere (42)). The development of

appropriate therapeutic approaches to contrast the

onset and the course of conformational (or protein

misfolding) diseases is of great interest. Such

diseases can individually be treated with

pharmacologic chaperones that specifically

enhance the folding-efficiency of a mutated gene

product (43). Alternatively, loss-of-function defects

could be rescued by targeting with therapeutic

agents the ER quality control machinery so that, as

one example, mutated polypeptides are offered

wider time windows to eventually attain the native

structure. Such approaches would greatly benefit of

the capacity to predict for each disease-causing

polypeptide, which chaperones and pathways will

be engaged for ER-retention, quality control and/or

disposal.

The presence of N-linked oligosaccharides

The oligosaccharyl transferase complex located at

the entry site of the ER transfers pre-assembled

glucose3-mannose9-N-acetyl glucosamine2-

oligosaccharides onto nascent polypeptide chains

entering the ER lumen (44). Normally, two terminal

glucose residues are rapidly removed by the

sequential intervention of the α-glucosidases I and

II. This generates a mono-glucosylated protein-

bound oligosaccharide that engages the lectin

chaperones calnexin and calreticulin and exposes

newly synthesized polypeptides to a folding

environment comprising the oxidoreductase ERp57

and the peptidyl-prolyl isomerase CypB. ERp57

promotes formation of the native set of disulfide

bonds, CypB the appropriate structuring of peptidyl-

prolyl bonds in the cis or in the trans configuration

(44).

Some nascent polypeptides rapidly enter off-

pathways of the folding program. This may inhibit

the action of the α-glucosidase II and results in the

inappropriate persistence of the di-glucosylated

form of protein-bound oligosaccharides. Aberrantly

structured newly synthesized polypeptides

displaying di-glucosylated oligosaccharides are

sequestered by a recently characterized membrane

anchored ER lectin, malectin (45). Malectin plays a

crucial role in retention of folding-defective

glycoproteins in the ER and acts in concert with

ribophorin I, a subunit of the ER glycosylation

machinery with proposed molecular chaperone

function (46, 47).

Persistent ER retention of misfolded conformers

elicits intervention of ER-resident mannosidases of

the glycosyl hydrolase 47 family (48) comprising the

ERManI, EDEM1, EDEM2, EDEM3. These directly

or indirectly enhance removal of α1,2-bonded

mannose residues (49-57) to extract misfolded

polypeptides from the calnexin chaperone system

(58) and to generate disposal signals decoded by

mannose-binding ERAD lectins such as OS-9 and

XTP3-B variants (50, 51, 59-62). The ERAD lectins,

which are interchangeable in function at least for

soluble ERAD substrates (ERAD-LS) (63), shuttle

misfolded polypeptides from the ER lumen to

dislocation sites in the ER membrane (Fig. 2). Both

the mannose-cleaving and the mannose-binding

ERAD factors also bind non-glycosylated proteins

(60, 61, 64-66) possibly as components of

supramolecular complexes containing general

chaperones such as BiP and GRP94 that direct

their action towards terminally misfolded proteins

(67).

The position of the folding defect and the

topology of the misfolded protein

E3 ubiquitin ligases embedded in the ER membrane

participate in supramolecular complexes containing

membrane-anchored and peripherally associated

factors both at the luminal and at the cytosolic side

of the bilayer. These multimeric complexes, the

dislocons, deliver misfolded proteins from the ER

lumen into the cytosol for proteasome-mediated

degradation (Fig. 2).

In S. cerevisiae, dislocons are built around two E3

ubiquitin ligases, Hrd1p and Doa10p. The luminal

(ERAD-L), transmembrane (ERAD-M) or cytosolic

(ERAD-C) position of the folding defect may

determine the engagement of either one of them.

Hrd1p preferentially assists ERAD-L and ERAD-M

substrates (68-72); Doa10p assists ERAD-C

substrates (68, 69, 73). The two yeast E3 ligases

have redundant function in assisting polytopic,

folding-defective polypeptides such as the ΔF508

cystic fibrosis transmembrane regulator (CFTR)

(74).

The mammalian ER membrane contains at least 24

RING-finger E3 ubiquitin ligases (75), some of

which (e.g., HRD1, gp78, RNF5, TEB4, RFP2,

TRC8) have a documented role in ERAD (76). The

ERAD system in Metazoa is therefore much more

complex and redundant than in Fungi. Only very

few attempts have been performed to systematically

establish how changes in the biophysical features of

misfolded polypeptides determine the engagement

of specific dislocation factors. For example,

misfolded luminal modules (ERAD-LS substrates)

strongly depend on the supramolecular complex

built around HRD1, comprising the associated

adaptor SEL1L and the two interchangeable ERAD

lectins OS-9 and XTP3-B that deliver the misfolded

polypeptides from the ER lumen to the dislocation

site. The same misfolded modules tethered at the

ER membrane (ERAD-LM substrates) are efficiently

cleared from the ER even in cells with non-

functional HRD1 dislocons, possibly because lateral

diffusion in the lipid bilayer facilitates delivery to

alternative machineries (Fig. 2) (63, 77).

The disposal of orphan/misfolded membrane

proteins: select examples

The study of the fate of orphan subunits of

oligomeric complexes is of historical importance.

Klausner and colleagues crucially reported that

unassembled subunits of the T cell receptor (TCR)

are degraded in a pre-lysosomal compartment

being part or closely related to the ER (78). Since

then, several TCR subunits, namely TCRα, TCRβ

and CD3δ have become model substrates to

investigate the mechanisms of protein disposal from

the ER (79-86). More recently, CD147 a component

of several membrane transporters, has been

identified as an endogenous ERAD substrate

whose turnover may regulate expression of the

corresponding functional complexes (87). The

peculiarity of all these proteins is the presence of

charged residues in the intramembrane space.

These residues promote the assembly of functional

complexes by pairing with residues of opposite

charge present in the transmembrane domains of

partner polypeptides in oligomeric complexes.

When oligomerization fails, the intramembrane

domains displaying charged residues might

facilitate the rapid degradation of the orphan

subunits by maintaining them in a monomeric state

and by recruiting specific ERAD factors (86, 88-90).

They may also serve, in some specific case, as

intramembrane cleavage sites that may facilitate

extraction from the ER membrane of the orphan

subunit to be degraded (91). TCRα and CD3δ are

both type I glycopolypeptides with charged residues

in the intramembrane space. Yet, their requirements

for efficient degradation differ quite extensively, thus

showing how difficult it may be to predict factors

and pathways engaged during specific polypeptides

quality control. For instance, only degradation of

CD3δ requires extensive N-glycans de-

mannosylation and only the cytosolic dislocation of

CD3δ is coupled with proteasomal activity (82).

Significantly, the cytosolic domain of TCRα is poly-

ubiquitylated on serine residues (85), rather than on

the more conventional lysine residues. Serine,

cysteine and threonine residues as well as the

polypeptide N-terminus are alternative ubiquitin

acceptor sites (33, 92-94).

The evolution of transmembrane domains

displaying destabilizing charged residues is one of

the strategies developed by viruses to co-opt the

mammalian dislocation machinery during the

infection cycle. This has been reported for simian

virus 40, where a charged residue in the

membrane-inserted domain of the VP2 protein,

which is conserved in several other polyomaviruses,

engages the ERAD factor BAP31 to possibly

promote dislocation of the entire virion across the

ER membrane (90).

The multispanning CFTR protein has also

extensively been described. It is characterized by

low folding-efficiency, about 25%, which is further

decreased upon gene mutations (95-97). The

ΔF508CFTR mutant form was the first reported

example of involvement of cytosolic programs in the

clearance of folding-defective polypeptides from the

mammalian ER (98, 99). This cystic fibrosis-causing

defective gene product is a paradigm for all mutant

proteins that maintain their function (100), yet are

rapidly destroyed from the ER (98, 99) or from the

cell surface (101) because structurally unstable.

This inevitably results in a loss-of-function protein

folding disease. Such mutants, or the quality control

pathways that they engage, are or may become

targets of therapeutic intervention based on the use

of chemical or pharmacologic chaperones as well

as proteostasis modifiers (95, 102). This alone

justifies the importance of deciphering the

mechanisms regulating protein quality control at the

molecular level.

The needs for unfolding

Many models (and at least as many candidates for

a channel in the ER membrane (103-105)) have

been proposed to explain how misfolded

polypeptides, tightly folded domains, catalytic

subunits of AB toxins (e.g., ricin, cholera, Shiga-

like) and even entire viral particles are dislocated

across the ER membrane. This remains poorly

understood but it relies on the intervention of

several luminal, cytosolic and membrane bound

factors including BiP, BAP31 and BAP29 (90, 106),

EDEM1 (107, 108), SEL1L (109), derlins (110-112),

HRD1 (113, 114), gp78 (114), p97/VCP (115, 116)

and many others.

Seemingly contrasting data are available on the

requirement for polypeptides unfolding to facilitate

transport across the lipid bilayer. For example, it

has been shown that the tightly folded DHFR

protein does not affect dislocation when appended

to class I MHC in US2 and US11-expressing cells

(117, 118). Moreover, compelling data do exist

showing that virions can dislocate across the ER

membrane and that misfolded glycopolypeptides,

which display bulky, highly hydrophilic

oligosaccharides bound to the polypeptide

backbone can efficiently pass the lipid bilayer to be

exported into the cytosol. However, other data imply

that a certain degree of unfolding is required for

efficient dislocation. For example, in contrast to

reference (117), DHFR appended to ERAD

substrates in non-infected cells may substantially

impair dislocation (119). Moreover, several

evidence highlight the involvement of members of

the protein disulfide isomerase (PDI) family in the

preparation of misfolded polypeptides for dislocation

across the ER membrane and in transport across

the bilayer of virions and of catalytic subunits of

toxins. For example, intra- (120) and inter-molecular

disulfide bonds (121) in ERAD substrates as well as

interchain disulfides linking the regulatory and

catalytic subunits of bacterial toxins are reduced

with the intervention of PDI (122-126) and/or PDI

family members (e.g., ERdj5 (127-129), ERp57

(106), ERp72, ERp29 (130), TMX1 (131), ERFAD-

ERp90 (132) and ERO1 (133)).

Recent work from our lab showed that the

conversion of cis into trans peptidyl-prolyl bonds by

the peptidyl-prolyl cis/trans isomerase CyPB may

facilitate transport of misfolded polypeptides

through the elusive transmembrane dislocation

channel by unfolding the polypeptide chain (134).

Finally, a role of rhomboid proteases and

pseudoproteases in ERAD substrates extraction

from the ER has been proposed and relates to their

capacity to cleave intramembrane anchors or to

cause bilayer thinning, respectively (91, 135, 136).

B. Regulation of the ERAD activity

Physiologic and pathologic destruction of

nascent chains

Contrasting data are available on the overall

fraction of nascent chains degraded before

attainment of the native structure, i.e., of defective

ribosomal products (DRiPS). Values range from

more than 30% (137) to substantially less (138).

DRiPs include polypeptides that have failed the

folding program. In this context, individual

mammalian polypeptides are characterized by

specific folding efficiencies (as low as 25% for the

CFTR, approaching the 90% for α1AT (60)). DRiPs

may also derive from defective mRNA or may be

characterized by errors in the primary sequence

that occur when the genetic information is

converted into proteins and most frequently result

from tRNA-aminoacid mis-acylation. If strictly

monitored by quality control programs, the relative

inefficiency of the protein biogenesis pathways is

fully compatible with life as it warrants production of

sufficient amount of functional proteins that keep

cells, tissues and organisms running. It probably

offers evolutionary advantages; DRiPs have been

proposed to provide ligands for class I MHC

presentation, thus enabling the immune system to

rapidly detect alterations in cellular gene expression

possibly caused by pathogen infection or

tumorogenic deviance (139).

A dysregulated ERAD activity may substantially

contribute to shift the physiological inefficiency of

protein folding to a pathological condition. A

constitutive ERAD activity surveys the normal

production of by-products of protein biogenesis

thereby contrasting their toxic accumulation. If the

constitutive ERAD activity is insufficient, folding-

defective polypeptides may remain in the ER lumen

thereby eventually interfering with protein

biogenesis and secretion upon, as one example,

persistent sequestration of folding chaperones (41,

140, 141). In contrast, excessive ERAD activity

would drastically reduce the time allocated to

nascent chains for maturation, thus causing loss-of-

function phenotypes and diseases (19, 34, 35, 38-

40, 142-149).

Understanding ERAD activity regulation is one of

the fascinating questions addressed in several

laboratories working on cellular proteostasis. The

transcriptional/translational induction of ERAD

factors upon activation of unfolded protein response

(UPR) programs is the best studied mechanism

operating in eukaryotic cells to adapt ERAD (and

folding) activity to changes in cellular needs (150).

UPR activation may optimally respond to long-

lasting (chronic) ER stresses. It seems however

inappropriate to respond to transient, acute and/or

periodic fluctuations in ER cargo load. In fact, onset

of these programs has a latency of several hours (2,

151). Moreover, the recovery from a stress requires

activation of elusive mechanisms that must reduce

the size of the ER after the UPR-triggered

expansion and must remove the excess of

chaperones produced during the ER stress phase.

These might be very slow processes because ER

stress-induced chaperones and enzymes such as

PDI family members, calnexin, calreticulin, BiP have

half-lives above the 24 h (152).

We propose that in healthy eukaryotic cells, the

constitutive ERAD that must cope with the

physiological production of by-products of protein

biogenesis without interfering with ongoing folding

programs is controlled by mechanisms collectively

defined as ERAD tuning. These mechanisms

survey the turn-over, the segregation from the ER

and the engagement of ERAD factors in

supramolecular functional complexes (5).

The concept of ERAD tuning

The concept of ERAD tuning is based on three

observations: i) unlike conventional ER-resident

chaperones and enzymes, several ERAD factors

(e.g., ERManI (153, 154), EDEM1 (38, 155-157),

OS-9 (149), XTP3-B (59), HERP (158, 159), SEL1L

(160, 161), the E3 ligases SMURF1 (162, 163) and

gp78 (164, 165), JAMP (166), ataxin-3 (167, 168))

are subjected to fast turnover, which may involve

the ubiquitin proteasome system, autophagy or

autophagy-like pathways; ii) select ERAD factors

are constitutively segregated from the ER (149,

169); iii) luminal expression of misfolded

polypeptides may delay the turnover of ERAD

factors, may retain them in the ER by interfering

with their vesicle-mediated segregation from the

compartment and may directly affect the

composition of supramolecular ERAD complexes

thereby setting their activity (149, 170, 171) (Fig. 3).

The concept of ERAD tuning proposes that

interactions with misfolded proteins synthesized in

the ER directly regulate (tune) the abundance, the

localization and the activity of components of the

ERAD machinery. Misfolded polypeptides may

support the assembly (or inhibit the disassembly) of

active supramolecular ERAD complexes as

reported, for example, for the association of the

adaptor protein SEL1L with the ER membrane-

embedded HRD1 dislocation machinery (149).

Moreover, since E3 ligases may ubiquitylate

themselves (164, 165, 172-174) and other

components of the ERAD machinery that are

engaged with them in supramolecular complexes

(158-160, 164, 165, 168), a possible scenario is that

folding-defective polypeptides serve as preferred

acceptors for the ubiquitylating activity of the E3

ligases thereby protecting the ERAD machinery

from self-destruction (Fig. 3).

The concept of ERAD tuning implies that misfolded

proteins may directly enhance the activity of the

ERAD machinery by a sort of autocrine regulatory

mechanism(s) that controls turnover, localization

and assembly of ERAD components (Fig. 3). In this

scenario, rapid adaptation, i.e., enhancement and

inhibition of ERAD activity, does not require signal

transduction from the ER to the nucleus, as it is the

case when the UPR must be activated. All in all, the

ERAD tuning pathways may offer rapid and readily

reversible adaptation responses to deal with

transient problems that may arise in the folding

compartment.

Interestingly, pathogens such as the mouse

hepatitis virus have been shown to hijack ERAD

tuning pathways to co-opt ER-derived vesicles

containing segregated ERAD factors as platforms

for replication of the viral genome (149, 156).

Conclusion The capacity to distinguish between non-native

polypeptides to be retained in the folding

compartment to complete the maturation program

and non-native polypeptides that must be removed

from the ER to prevent their toxic accumulation is

crucial to maintain cellular proteostasis.

Dysregulated ERAD may lead to inappropriate,

premature selection of folding-competent

conformers for destruction (too high activity) or to

an equally inappropriate and harmful tolerance of

misfolded conformers in the folding-compartment

(too weak activity). As these conditions inevitably

affect maintenance of cellular and organisms

proteostasis, the molecular characterization of the

processes covered by this review is of daunting

importance. What are the mechanisms that

maintain the fragile equilibrium between the activity

of the folding and the degradation machineries

operating in the ER and in other cellular

compartments such as the cytosol and the

mitochondria where not-yet-native proteins must

attain their functional conformations, yet misfolded

conformers must efficiently be cleared? How is this

equilibrium maintained upon the frequent

modifications of the environmental conditions cells

are subjected to? How do cells revert situations of

disequilibrium that may intervene upon pathogen

attack, nutrient deprivation, variations in protein

synthesis levels, transient failure of folding

programs? How do pathogens actually exploit the

host protein folding and degradation machineries to

enhance their survival and replication chances?

How can we intervene to contrast their infection

cycles? These are few of the challenging questions

remaining for future research.

Acknowledgments Members of the Molinari’s lab are acknowledged for

discussions and suggestions. M.M. is supported by

grants from the Foundation for Research on

Neurodegenerative Diseases, the Swiss National

Science Foundation, the Association Française

contre les Myopathies, the Gabriele Foundation.

1.  Balch WE, Morimoto RI, Dillin A, Kelly  JW.  Adapting  proteostasis  for disease  intervention.  Science 2008;319(5865):916‐919. 2.  Walter  P,  Ron  D.  The  unfolded protein  response:  from  stress  pathway to  homeostatic  regulation.  Science 2011;334(6059):1081‐1086. 3.  Molinari  M.  N‐glycan  structure dictates  extension  of  protein  folding  or onset  of  disposal.  Nat  Chem  Biol 2007;3(6):313‐320. 4.  Hampton  RY,  Garza  RM.  Protein quality control as a strategy  for cellular regulation:  lessons  from  ubiquitin‐mediated  regulation  of  the  sterol pathway.  Chem Rev  2009;109(4):1561‐1574. 5.  Bernasconi  R,  Molinari  M.  ERAD and ERAD tuning: disposal of cargo and of  ERAD  regulators  from  the mammalian  ER.  Curr  Opin  Cell  Biol 2011;23(2):176‐183. 6.  Gardner  RG,  Hampton  RY.  A highly  conserved  signal  controls degradation  of  3‐hydroxy‐3‐methylglutaryl‐coenzyme  A  (HMG‐CoA) reductase  in  eukaryotes.  The  Journal  of biological  chemistry 1999;274(44):31671‐31678. 7.  Hampton RY, Gardner RG, Rine J. Role of 26S proteasome and HRD genes in  the  degradation  of  3‐hydroxy‐3‐ methylglutaryl‐CoA  reductase,  an integral  endoplasmic  reticulum membrane  protein.  Mol  Biol  Cell 1996;7(12):2029‐2044. 

8.  Sever  N,  Song  BL,  Yabe  D, Goldstein  JL,  Brown  MS,  DeBose‐Boyd RA.  Insig‐dependent  ubiquitination  and degradation  of  mammalian  3‐hydroxy‐3‐methylglutaryl‐CoA  reductase stimulated  by  sterols  and geranylgeraniol.  The  Journal  of biological  chemistry 2003;278(52):52479‐52490. 9.  Dixon  JL,  Furukawa  S,  Ginsberg HN.  Oleate  stimulates  secretion  of apolipoprotein  B‐containing lipoproteins  from  Hep  G2  cells  by inhibiting  early  intracellular degradation  of  apolipoprotein  B.  The Journal  of  biological  chemistry 1991;266(8):5080‐5086. 10.  Zhou M, Fisher EA, Ginsberg HN. Regulated  Co‐translational ubiquitination of apolipoprotein B100. A new  paradigm  for  proteasomal degradation of  a  secretory protein.  The Journal  of  biological  chemistry 1998;273(38):24649‐24653. 11.  Alzayady KJ, Panning MM, Kelley GG, Wojcikiewicz RJ. Involvement of the p97‐Ufd1‐Npl4 complex in the regulated endoplasmic  reticulum‐associated degradation  of  inositol  1,4,5‐trisphosphate  receptors.  The  Journal  of biological  chemistry 2005;280(41):34530‐34537. 12.  Loureiro  J,  Ploegh  HL.  Antigen presentation  and  the  ubiquitin‐proteasome  system  in  host‐pathogen interactions.  Adv  Immunol 2006;92:225‐305. 13.  Randow  F,  Lehner  PJ.  Viral avoidance  and  exploitation  of  the ubiquitin  system.  Nat  Cell  Biol 2009;11(5):527‐534. 14.  Wiertz EJ, Jones TR, Sun L, Bogyo M,  Geuze  HJ,  Ploegh  HL.  The  human cytomegalovirus  US11  gene  product dislocates  MHC  class  I  heavy  chains from  the  endoplasmic  reticulum  to  the cytosol. Cell 1996;84(5):769‐779. 15.  Wiertz EJ, Tortorella D, Bogyo M, Yu J, Mothes W,  Jones TR, Rapoport TA, 

Ploegh HL. Sec61‐mediated transfer of a membrane  protein  from  the endoplasmic  reticulum  to  the proteasome  for  destruction.  Nature 1996;384(6608):432‐438. 16.  Tomazin R, Boname J, Hegde NR, Lewinsohn  DM,  Altschuler  Y,  Jones  TR, Cresswell  P,  Nelson  JA,  Riddell  SR, Johnson  DC.  Cytomegalovirus  US2 destroys  two  components  of  the  MHC class II pathway, preventing recognition by  CD4+  T  cells.  Nat  Med 1999;5(9):1039‐1043. 17.  Lilley  BN,  Ploegh  HL.  A membrane  protein  required  for dislocation  of  misfolded  proteins  from the  ER.  Nature  2004;429(6994):834‐840. 18.  Ye  Y,  Shibata  Y,  Yun  C,  Ron  D, Rapoport  TA.  A  membrane  protein complex  mediates  retro‐translocation from  the  ER  lumen  into  the  cytosol. Nature 2004;429(6994):841‐847. 19.  Stagg  HR,  Thomas  M,  van  den Boomen  D,  Wiertz  EJ,  Drabkin  HA, Gemmill  RM,  Lehner  PJ.  The  TRC8  E3 ligase  ubiquitinates  MHC  class  I molecules  before  dislocation  from  the ER. J Cell Biol 2009;186(5):685‐692. 20.  Gilbert MJ, Riddell SR, Plachter B, Greenberg  PD.  Cytomegalovirus selectively  blocks  antigen  processing and presentation of  its  immediate‐early gene  product.  Nature 1996;383(6602):720‐722. 21.  Masucci  MG.  Epstein‐Barr  virus oncogenesis  and  the  ubiquitin‐proteasome  system.  Oncogene 2004;23(11):2107‐2115. 22.  Remoli  AL,  Marsili  G,  Perrotti  E, Gallerani  E,  Ilari  R,  Nappi  F,  Cafaro  A, Ensoli  B,  Gavioli  R,  Battistini  A. Intracellular  HIV‐1  Tat  protein represses  constitutive  LMP2 transcription  increasing  proteasome activity  by  interfering with  the  binding of  IRF‐1  to  STAT1.  Biochem  J 2006;396(2):371‐380. 

23.  Zhang Z, Protzer U, Hu Z, Jacob J, Liang  TJ.  Inhibition  of  cellular proteasome  activities  enhances hepadnavirus  replication  in  an  HBX‐dependent  manner.  J  Virol 2004;78(9):4566‐4572. 24.  Zaldumbide  A,  Ossevoort  M, Wiertz EJ, Hoeben RC. In cis inhibition of antigen  processing  by  the  latency‐associated  nuclear  antigen  I  of  Kaposi sarcoma  herpes  virus.  Mol  Immunol 2007;44(6):1352‐1360. 25.  Schubert  U,  Anton  LC,  Bacik  I, Cox JH, Bour S, Bennink JR, Orlowski M, Strebel K, Yewdell JW. CD4 glycoprotein degradation  induced  by  human immunodeficiency  virus  type  1  Vpu protein  requires  the  function  of proteasomes  and  the  ubiquitin‐conjugating  pathway.  J  Virol 1998;72(3):2280‐2288. 26.  Willey  RL,  Maldarelli  F,  Martin MA,  Strebel  K.  Human immunodeficiency  virus  type  1  Vpu protein  induces  rapid  degradation  of CD4. J Virol 1992;66(12):7193‐7200. 27.  Fujita K, Omura S, Silver  J. Rapid degradation  of  CD4  in  cells  expressing human  immunodeficiency  virus  type  1 Env  and Vpu  is  blocked  by  proteasome inhibitors.  J  Gen  Virol  1997;78  (  Pt 3):619‐625. 28.  Margottin  F,  Bour  SP,  Durand H, Selig  L,  Benichou  S,  Richard  V,  Thomas D, Strebel K, Benarous R. A novel human WD protein,  h‐beta  TrCp,  that  interacts with HIV‐1 Vpu connects CD4 to the ER degradation  pathway  through  an  F‐box motif. Mol Cell 1998;1(4):565‐574. 29.  Lybarger  L,  Wang  X,  Harris  MR, Virgin  HWt,  Hansen  TH.  Virus subversion  of  the MHC  class  I  peptide‐loading  complex.  Immunity 2003;18(1):121‐130. 30.  Boname  JM,  Stevenson  PG.  MHC class  I  ubiquitination  by  a  viral PHD/LAP  finger  protein.  Immunity 2001;15(4):627‐636. 

31.  Yu  YY,  Harris  MR,  Lybarger  L, Kimpler  LA,  Myers  NB,  Virgin  HWt, Hansen  TH.  Physical  association  of  the K3  protein  of  gamma‐2  herpesvirus  68 with  major  histocompatibility  complex class I molecules with impaired peptide and  beta(2)‐microglobulin  assembly.  J Virol 2002;76(6):2796‐2803. 32.  Boname  JM,  de  Lima BD,  Lehner PJ,  Stevenson  PG.  Viral  degradation  of the  MHC  class  I  peptide  loading complex.  Immunity  2004;20(3):305‐317. 33.  Wang  X,  Herr  RA,  Chua  WJ, Lybarger  L,  Wiertz  EJ,  Hansen  TH. Ubiquitination  of  serine,  threonine,  or lysine  residues  on  the  cytoplasmic  tail can  induce  ERAD  of  MHC‐I  by  viral  E3 ligase  mK3.  The  Journal  of  cell  biology 2007;177(4):613‐624. 34.  Herr  RA,  Wang  X,  Loh  J,  Virgin HW, Hansen TH. Newly Discovered Viral E3  Ligase  pK3  Induces  Endoplasmic Reticulum‐associated  Degradation  of Class  I  Major  Histocompatibility Proteins  and  Their  Membrane‐bound Chaperones.  The  Journal  of  biological chemistry 2012;287(18):14467‐14479. 35.  Saeed M,  Suzuki  R, Watanabe  N, Masaki  T,  Tomonaga  M,  Muhammad  A, Kato T, Matsuura Y, Watanabe H, Wakita T,  Suzuki  T.  Role  of  the  endoplasmic reticulum‐associated  degradation (ERAD)  pathway  in  degradation  of hepatitis  C  virus  envelope proteins  and production  of  virus  particles.  The Journal  of  biological  chemistry 2011;286(43):37264‐37273. 36.  Lazar  C,  Macovei  A,  Petrescu  S, Branza‐Nichita  N.  Activation  of  ERAD pathway  by  human  hepatitis  B  virus modulates  viral  and  subviral  particle production.  PLoS  One 2012;7(3):e34169. 37.  Olivari  S,  Molinari  M. Glycoprotein  folding  and  the  role  of EDEM1,  EDEM2  and  EDEM3  in degradation  of  folding‐defective 

glycoproteins.  FEBS  Lett 2007;581(19):3658‐3664. 38.  Cali T, Galli C, Olivari S, Molinari M.  Segregation  and  rapid  turnover  of EDEM1  by  an  autophagy‐like mechanism  modulates  standard  ERAD and  folding  activities.  Biochem Biophys Res Commun 2008;371(3):405‐410. 39.  Wu Y, Swulius MT, Moremen KW, Sifers  RN.  Elucidation  of  the  molecular logic  by  which  misfolded  alpha  1‐antitrypsin  is preferentially selected  for degradation.  Proc  Natl  Acad  Sci  U  S  A 2003;100(14):8229‐8234. 40.  Younger  JM,  Chen  L,  Ren  HY, Rosser  MF,  Turnbull  EL,  Fan  CY, Patterson C, Cyr DM. Sequential quality‐control  checkpoints  triage  misfolded cystic  fibrosis  transmembrane conductance  regulator.  Cell 2006;126(3):571‐582. 41.  Eriksson  KK,  Vago  R,  Calanca  V, Galli  C,  Paganetti  P,  Molinari  M.  EDEM contributes  to  maintenance  of  protein folding  efficiency  and  secretory capacity.  J  Biol  Chem 2004;279(43):44600‐44605. 42.  Powers ET, Morimoto RI, Dillin A, Kelly  JW,  Balch  WE.  Biological  and chemical  approaches  to  diseases  of proteostasis  deficiency.  Annu  Rev Biochem 2009;78:959‐991. 43.  Bernier  V,  Lagace  M,  Bichet  DG, Bouvier  M.  Pharmacological chaperones:  potential  treatment  for conformational  diseases.  Trends Endocrinol Metab 2004;15(5):222‐228. 44.  Aebi  M,  Bernasconi  R,  Clerc  S, Molinari  M.  N‐glycan  structures: recognition  and  processing  in  the  ER. Trends Biochem Sci 2010;35(2):74‐82. 45.  Galli  C,  Bernasconi  R,  Solda  T, Calanca  V,  Molinari  M.  Malectin participates  in  a  backup  glycoprotein quality  control  pathway  in  the mammalian  ER.  PLoS  One 2011;6(1):e16304. 46.  Qin  SY,  Hu  D,  Matsumoto  K, Takeda  K,  Matsumoto  N,  Yamaguchi  Y, 

Yamamoto K. Malectin  forms a complex with  ribophorin  I  for  enhanced association  with  misfolded glycoproteins.  The  Journal  of  biological chemistry 2012;287(45):38080‐38089. 47.  Wilson  CM,  Kraft  C,  Duggan  C, Ismail  N,  Crawshaw  SG,  High  S. Ribophorin I associates with a subset of membrane  proteins  after  their integration  at  the  sec61  translocon.  J Biol Chem 2005;280(6):4195‐4206. 48.  Moremen  KW,  Molinari  M.  N‐linked  glycan  recognition  and processing:  the  molecular  basis  of endoplasmic  reticulum  quality  control. Curr  Opin  Struct  Biol  2006(16):592‐599. 49.  Olivari  S,  Cali  T,  Salo  KE, Paganetti  P,  Ruddock  LW,  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(4):1278‐1284. 50.  Clerc  S,  Hirsch  C,  Oggier  DM, Deprez  P,  Jakob  C,  Sommer  T,  Aebi  M. Htm1  protein  generates  the  N‐glycan signal  for  glycoprotein  degradation  in the  endoplasmic  reticulum.  J  Cell  Biol 2009;184:159‐172. 51.  Quan  EM,  Kamiya  Y,  Kamiya  D, Denic V, Weibezahn J, Kato K, Weissman JS. Defining the glycan destruction signal for  endoplasmic  reticulum‐associated degradation.  Mol  Cell  2008;32(6):870‐877. 52.  Kitzmuller  C,  Caprini  A,  Moore SE,  Frenoy  JP,  Schwaiger E, Kellermann O, Ivessa NE, Ermonval M. Processing of N‐linked  glycans  during  endoplasmic‐reticulum‐associated  degradation  of  a short‐lived  variant  of  ribophorin  I. Biochem J 2003;376(Pt 3):687‐696. 53.  Hosokawa  N,  Tremblay  LO,  You Z,  Herscovics  A,  Wada  I,  Nagata  K. Enhancement of  endoplasmic  reticulum (ER)  degradation  of  misfolded  Null Hong  Kong  alpha1‐antitrypsin  by 

human  ER  mannosidase  I.  J  Biol  Chem 2003;278(28):26287‐26294. 54.  Frenkel Z, Gregory W, Kornfeld S, Lederkremer  GZ.  Endoplasmic reticulum‐associated  degradation  of mammalian  glycoproteins  involves sugar  chain  trimming  to  Man6‐5GlcNAc2.  J  Biol  Chem 2003;278(36):34119‐34124. 55.  Foulquier F, Duvet S, Klein A, Mir AM,  Chirat  F,  Cacan  R.  Endoplasmic reticulum‐associated  degradation  of glycoproteins  bearing  Man5GlcNAc2 and Man9GlcNAc2  species  in  the MI8‐5 CHO  cell  line.  Eur  J  Biochem 2004;271(2):398‐404. 56.  Foulquier  F,  Harduin‐Lepers  A, Duvet S, Marchal I, Mir AM, Delannoy P, Chirat F, Cacan R. The unfolded protein response  in  a  dolichyl  phosphate mannose‐deficient  Chinese  hamster ovary cell line points out the key role of a  demannosylation  step  in  the  quality‐control  mechanism  of  N‐glycoproteins. Biochem J 2002;362(Pt 2):491‐498. 57.  Ermonval  M,  Kitzmuller  C,  Mir AM,  Cacan  R,  Ivessa  NE.  N‐glycan structure  of  a  short‐lived  variant  of ribophorin I expressed in the MadIA214 glycosylation‐defective  cell  line  reveals the role of a mannosidase that is not ER mannosidase  I  in  the  process  of glycoprotein  degradation.  Glycobiology 2001;11(7):565‐576. 58.  Molinari  M,  Calanca  V,  Galli  C, Lucca  P,  Paganetti  P.  Role  of  EDEM  in the  release  of  misfolded  glycoproteins from  the  calnexin  cycle.  Science 2003;299(5611):1397‐1400. 59.  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(30):20914‐20924. 60.  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(24):16446‐16454. 61.  Christianson  JC,  Shaler  TA,  Tyler RE, Kopito RR. OS‐9 and GRP94 deliver mutant  alpha1‐antitrypsin  to  the Hrd1/SEL1L  ubiquitin  ligase  complex for ERAD. Nat Cell Biol 2008;10(3):272‐282. 62.  Alcock F, Swanton E. Mammalian OS‐9  is  upregulated  in  response  to endoplasmic  reticulum  stress  and facilitates  ubiquitination  of  misfolded glycoproteins.  J  Mol  Biol 2009;385(4):1032‐1042. 63.  Bernasconi  R,  Galli  C,  Calanca  V, Nakajima  T,  Molinari  M.  Stringent requirement  for  HRD1,  SEL1L,  and  OS‐9/XTP3‐B  for  disposal  of  ERAD‐LS substrates.  J Cell Biol 2010;188(2):223‐235. 64.  Cormier JH, Tamura T, Sunryd JC, Hebert  DN.  EDEM1  recognition  and delivery  of  misfolded  proteins  to  the SEL1L‐containing  ERAD  complex.  Mol Cell 2009;34(5):627‐633. 65.  Ron E, Shenkman M, Groisman B, Izenshtein  Y,  Leitman  J,  Lederkremer GZ.  Bypass  of  glycan‐dependent glycoprotein  delivery  to  ERAD  by  up‐regulated  EDEM1.  Mol  Biol  Cell 2011;22(21):3945‐3954. 66.  Shenkman M, Groisman B, Ron E, Avezov E, Hendershot LM, Lederkremer GZ.  A  Shared  Endoplasmic  Reticulum‐associated  Degradation  Pathway Involving  the  EDEM1  Protein  for Glycosylated  and  Nonglycosylated Proteins.  The  Journal  of  biological chemistry 2013;288(4):2167‐2178. 67.  Hebert  DN, Molinari  M.  Flagging and docking: dual roles for N‐glycans in protein  quality  control  and  cellular proteostasis.  Trends  Biochem  Sci 2012;37(10):404‐410. 68.  Vashist  S,  Ng  DT.  Misfolded proteins  are  sorted  by  a  sequential 

checkpoint  mechanism  of  ER  quality control. J Cell Biol 2004;165(1):41‐52. 69.  Huyer  G,  Piluek  WF,  Fansler  Z, Kreft  SG,  Hochstrasser  M,  Brodsky  JL, Michaelis  S.  Distinct  machinery  is required  in  Saccharomyces  cerevisiae for  the  endoplasmic  reticulum‐associated  degradation  of  a multispanning membrane protein and a soluble  luminal  protein.  J  Biol  Chem 2004;279(37):38369‐38378. 70.  Taxis C, Hitt R, Park SH, Deak PM, Kostova  Z,  Wolf  DH.  Use  of  modular substrates  demonstrates  mechanistic diversity  and  reveals  differences  in chaperone  requirement  of  ERAD.  J  Biol Chem 2003;278(38):35903‐35913. 71.  Willer  M,  Forte  GM,  Stirling  CJ. Sec61p  is  required  for  ERAD‐L:  genetic dissection  of  the  translocation  and ERAD‐L functions of Sec61P using novel derivatives  of  CPY.  J  Biol  Chem 2008;283(49):33883‐33888. 72.  Carvalho  P,  Goder  V,  Rapoport TA.  Distinct  ubiquitin‐ligase  complexes define  convergent  pathways  for  the degradation  of  ER  proteins.  Cell 2006;126(2):361‐373. 73.  Metzger MB, Michaelis S. Analysis of  quality  control  substrates  in  distinct cellular  compartments  reveals a unique role  for  Rpn4p  in  tolerating  misfolded membrane  proteins.  Mol  Biol  Cell 2009;20(3):1006‐1019. 74.  Gnann  A,  Riordan  JR,  Wolf  DH. Cystic  fibrosis  transmembrane conductance  regulator  degradation depends  on  the  lectins  Htm1p/EDEM and the Cdc48 protein complex in yeast. Mol Biol Cell 2004;15(9):4125‐4135. 75.  Neutzner  A,  Neutzner  M, Benischke AS, Ryu SW, Frank S, Youle RJ, Karbowski  M.  A  systematic  search  for endoplasmic reticulum (ER) membrane‐associated  RING  finger  proteins identifies Nixin/ZNRF4 as a regulator of calnexin  stability  and  ER  homeostasis. The  Journal  of  biological  chemistry 2011;286(10):8633‐8643. 

76.  Claessen  JH,  Kundrat  L,  Ploegh HL.  Protein  quality  control  in  the  ER: balancing  the  ubiquitin  checkbook. Trends Cell Biol 2012;22(1):22‐32. 77.  Ninagawa  S,  Okada  T,  Takeda  S, Mori  K.  SEL1L  is  required  for endoplasmic  reticulum‐associated degradation  of  misfolded  luminal proteins  but  not  transmembrane proteins  in  chicken  DT40  cell  line.  Cell Struct Funct 2011;36(2):187‐195. 78.  Lippincott‐Schwartz J, Bonifacino JS,  Yuan  LC,  Klausner  RD.  Degradation from  the  endoplasmic  reticulum: disposing of newly synthesized proteins. Cell 1988;54(2):209‐220. 79.  Huppa  JB,  Ploegh  HL.  The  alpha chain  of  the  T  cell  antigen  receptor  is degraded  in  the  cytosol.  Immunity 1997;7(1):113‐122. 80.  DeLaBarre  B,  Christianson  JC, Kopito  RR,  Brunger  AT.  Central  pore residues  mediate  the  p97/VCP  activity required  for  ERAD.  Mol  Cell 2006;22(4):451‐462. 81.  Wang Q, Li L, Ye Y. Regulation of retrotranslocation  by  p97‐associated deubiquitinating enzyme ataxin‐3.  J Cell Biol 2006;174(7):963‐971. 82.  Yang M,  Omura  S,  Bonifacino  JS, Weissman  AM.  Novel  aspects  of degradation  of  T  cell  receptor  subunits from  the  ER  in  T  cells:  importance  of oligosaccharide  processing, ubiquitination,  and  proteasome‐dependent  removal  from  ER membranes.  J  Exp  Med 1998;187(6):835‐846. 83.  Yu  H,  Kaung  G,  Kobayashi  S, Kopito  RR.  Cytosolic  degradation  of  T‐cell  receptor  alpha  chains  by  the proteasome.  J  Biol  Chem 1997;272(33):20800‐20804. 84.  Young  J,  Kane  LP,  Exley  M, Wileman  T.  Regulation  of  selective protein degradation  in  the endoplasmic reticulum  by  redox  potential.  J  Biol Chem 1993;268(26):19810‐19818. 

85.  Ishikura  S,  Weissman  AM, Bonifacino  JS.  Serine  residues  in  the cytosolic  tail  of  the  T‐cell  antigen receptor  alpha‐chain  mediate ubiquitination  and  endoplasmic reticulum‐associated degradation of  the unassembled  protein.  The  Journal  of biological  chemistry 2010;285(31):23916‐23924. 86.  Soetandyo N, Wang Q, Ye Y, Li L. Role  of  intramembrane  charged residues  in  the  quality  control  of unassembled  T‐cell  receptor  alpha‐chains  at  the  endoplasmic  reticulum.  J Cell Sci 2010;123(Pt 7):1031‐1038. 87.  Tyler RE,  Pearce MM,  Shaler  TA, Olzmann  JA,  Greenblatt  EJ,  Kopito  RR. Unassembled  CD147  is  an  endogenous ER‐associated  degradation  (ERAD) substrate.  Mol  Biol  Cell 2012;23(24):4668‐4678. 88.  Bonifacino JS, Cosson P, Klausner RD.  Colocalized  transmembrane determinants  for  ER  degradation  and subunit  assembly  explain  the intracellular  fate  of  TCR  chains.  Cell 1990;63(3):503‐513. 89.  Bonifacino  JS,  Cosson  P,  Shah  N, Klausner RD. Role of potentially charged transmembrane  residues  in  targeting proteins  for  retention  and  degradation within the endoplasmic reticulum. Embo J 1991;10(10):2783‐2793. 90.  Geiger R, Andritschke D, Friebe S, Herzog F, Luisoni S, Heger T, Helenius A. BAP31  and  BiP  are  essential  for dislocation  of  SV40  from  the endoplasmic  reticulum  to  the  cytosol. Nat Cell Biol 2011;13(11):1305‐1314. 91.  Fleig  L,  Bergbold  N, Sahasrabudhe  P,  Geiger  B,  Kaltak  L, Lemberg  MK.  Ubiquitin‐dependent intramembrane  rhomboid  protease promotes  ERAD of membrane  proteins. Mol Cell 2012;47(4):558‐569. 92.  Cadwell  K,  Coscoy  L. Ubiquitination on nonlysine residues by a  viral  E3  ubiquitin  ligase.  Science 2005;309(5731):127‐130. 

93.  Wang  X,  Herr  RA,  Rabelink  M, Hoeben  RC,  Wiertz  EJ,  Hansen  TH. Ube2j2  ubiquitinates  hydroxylated amino  acids  on  ER‐associated degradation  substrates.  The  Journal  of cell biology 2009;187(5):655‐668. 94.  Shimizu  Y,  Okuda‐Shimizu  Y, Hendershot  LM.  Ubiquitylation  of  an ERAD  substrate  occurs  on  multiple types  of  amino  acids.  Mol  Cell 2010;40(6):917‐926. 95.  Ledford  H.  Drug  bests  cystic‐fibrosis  mutation.  Nature 2012;482(7384):145. 96.  Kopito  RR.  Biosynthesis  and degradation  of  CFTR.  Physiol  Rev 1999;79(1 Suppl):S167‐173. 97.  Ward CL, Kopito RR. Intracellular turnover  of  cystic  fibrosis transmembrane  conductance  regulator. Inefficient  processing  and  rapid degradation  of  wild‐type  and  mutant proteins.  J  Biol  Chem 1994;269(41):25710‐25718. 98.  Ward  CL,  Omura  S,  Kopito  RR. Degradation  of  CFTR  by  the  ubiquitin‐proteasome  pathway.  Cell 1995;83(1):121‐127. 99.  Jensen  TJ,  Loo  MA,  Pind  S, Williams  DB,  Goldberg  AL,  Riordan  JR. Multiple  proteolytic  systems,  including the  proteasome,  contribute  to  CFTR processing. Cell 1995;83(1):129‐135. 100.  Drumm  ML,  Wilkinson  DJ,  Smit LS,  Worrell  RT,  Strong  TV,  Frizzell  RA, Dawson  DC,  Collins  FS.  Chloride conductance  expressed  by  delta  F508 and  other  mutant  CFTRs  in  Xenopus oocytes.  Science  1991;254(5039):1797‐1799. 101.  Okiyoneda  T,  Barriere  H, Bagdany M, Rabeh WM, Du K, Hohfeld J, Young JC, Lukacs GL. Peripheral Protein Quality Control Removes Unfolded CFTR from  the  Plasma  Membrane.  Science 2010. 102.  Silva MC, Fox S, Beam M, Thakkar H,  Amaral  MD,  Morimoto  RI.  A  genetic screening  strategy  identifies  novel 

regulators  of  the  proteostasis  network. PLoS Genet 2011;7(12):e1002438. 103.  Hebert  DN,  Bernasconi  R, Molinari M. ERAD substrates: which way out?  Semin  Cell  Dev  Biol 2010;21(5):526‐532. 104.  Hampton RY, Sommer T. Finding the will and the way of ERAD substrate retrotranslocation.  Curr  Opin  Cell  Biol 2012;24(4):460‐466. 105.  Smith MH, Ploegh HL, Weissman JS.  Road  to  ruin:  targeting  proteins  for degradation  in  the  endoplasmic reticulum.  Science 2011;334(6059):1086‐1090. 106.  Schelhaas  M,  Malmstrom  J, Pelkmans  L,  Haugstetter  J,  Ellgaard  L, Grunewald  K,  Helenius  A.  Simian  Virus 40  depends  on  ER  protein  folding  and quality  control  factors  for  entry  into host cells. Cell 2007;131(3):516‐529. 107.  Slominska‐Wojewodzka  M, Gregers TF, Walchli S, Sandvig K. EDEM is involved in retrotranslocation of ricin from  the  endoplasmic  reticulum  to  the cytosol. Mol  Biol  Cell  2006;17(4):1664‐1675. 108.  Sokolowska I, Walchli S, Wegrzyn G, Sandvig K, Slominska‐Wojewodzka M. A single point mutation  in ricin A‐chain increases toxin degradation and inhibits EDEM1‐dependent  ER retrotranslocation.  Biochem  J 2011;436(2):371‐385. 109.  Redmann  V,  Oresic  K,  Tortorella LL,  Cook  JP,  Lord  M,  Tortorella  D. Dislocation  of  ricin  toxin  A  chains  in human  cells  utilizes  selective  cellular factors.  The  Journal  of  biological chemistry 2011;286(24):21231‐21238. 110.  Bernardi KM, Forster ML, Lencer WI, Tsai B. Derlin‐1 facilitates the retro‐translocation  of  cholera  toxin.  Mol  Biol Cell 2008;19(3):877‐884. 111.  Dixit  G,  Mikoryak  C,  Hayslett  T, Bhat  A,  Draper  RK.  Cholera  toxin  up‐regulates  endoplasmic  reticulum proteins  that  correlate  with  sensitivity 

to  the  toxin.  Exp  Biol  Med  (Maywood) 2008;233(2):163‐175. 112.  Lilley  BN,  Gilbert  JM,  Ploegh HL, Benjamin  TL.  Murine  polyomavirus requires  the  endoplasmic  reticulum protein  Derlin‐2  to  initiate  infection.  J Virol 2006;80(17):8739‐8744. 113.  Li  S,  Spooner  RA,  Hampton  RY, Lord  JM, Roberts LM. Cytosolic entry of Shiga‐like  toxin  a  chain  from  the  yeast endoplasmic  reticulum  requires catalytically  active  Hrd1p.  PLoS  One 2012;7(7):e41119. 114.  Bernardi  KM,  Williams  JM, Kikkert M, van Voorden S, Wiertz EJ, Ye Y, Tsai B. The E3 ubiquitin  ligases Hrd1 and  gp78  bind  to  and  promote  cholera toxin  retro‐translocation.  Mol  Biol  Cell 2010;21(1):140‐151. 115.  Kothe  M,  Ye  Y,  Wagner  JS,  De Luca  HE,  Kern  E,  Rapoport  TA,  Lencer WI.  Role  of  p97  AAA‐ATPase  in  the retrotranslocation  of  the  cholera  toxin A1 chain, a non‐ubiquitinated substrate. The  Journal  of  biological  chemistry 2005;280(30):28127‐28132. 116.  Wolf  DH,  Stolz  A.  The  Cdc48 machine  in  endoplasmic  reticulum associated protein degradation. Biochim Biophys Acta 2012;1823(1):117‐124. 117.  Tirosh B, Furman MH, Tortorella D, Ploegh HL. Protein unfolding  is not a prerequisite for endoplasmic reticulum‐to‐cytosol  dislocation.  J  Biol  Chem 2003;278(9):6664‐6672. 118.  Fiebiger  E,  Story  C,  Ploegh  HL, Tortorella D. Visualization of  the ER‐to‐cytosol  dislocation  reaction  of  a  type  I membrane  protein.  Embo  J 2002;21(5):1041‐1053. 119.  Bhamidipati A, Denic V, Quan EM, Weissman  JS.  Exploration  of  the topological  requirements  of  ERAD identifies  Yos9p  as  a  lectin  sensor  of misfolded  glycoproteins  in  the  ER lumen. Mol Cell 2005;19(6):741‐751. 120.  Fagioli  C,  Mezghrani  A,  Sitia  R. Reduction of  interchain  disulfide bonds precedes the dislocation of Ig‐mu chains 

from  the  endoplasmic  reticulum  to  the cytosol  for  proteasomal  degradation.  J Biol Chem 2001;276(44):40962‐40967. 121.  Molinari  M,  Galli  C,  Piccaluga  V, Pieren  M,  Paganetti  P.  Sequential assistance of molecular chaperones and transient  formation  of  covalent complexes  during  protein  degradation from  the  ER.  J  Cell  Biol 2002;158(2):247‐257. 122.  Majoul  I,  Ferrari  D,  Soling  HD. Reduction  of  protein  disulfide  bonds  in an oxidizing environment. The disulfide bridge  of  cholera  toxin  A‐subunit  is reduced  in  the  endoplasmic  reticulum. FEBS Lett 1997;401(2‐3):104‐108. 123.  Tsai  B,  Rodighiero  C,  Lencer WI, Rapoport  TA.  Protein  disulfide isomerase  acts  as  a  redox‐dependent chaperone  to  unfold  cholera  toxin.  Cell 2001;104(6):937‐948. 124.  Spooner  RA,  Watson  PD, Marsden  CJ,  Smith DC, Moore KA,  Cook JP,  Lord  JM,  Roberts  LM.  Protein disulphide‐isomerase  reduces  ricin  to its  A  and  B  chains  in  the  endoplasmic reticulum.  Biochem  J  2004;383(Pt 2):285‐293. 125.  Forster  ML,  Sivick  K,  Park  YN, Arvan  P,  Lencer  WI,  Tsai  B.  Protein disulfide  isomerase‐like  proteins  play opposing  roles  during retrotranslocation.  J  Cell  Biol 2006;173(6):853‐859. 126.  Moore MS. Protein translocation: Nuclear export‐out of  the dark. Current Biol 1996;6:137‐140. 127.  Ushioda  R,  Hoseki  J,  Araki  K, Jansen G, Thomas DY, Nagata K. ERdj5 is required  as  a  disulfide  reductase  for degradation of misfolded proteins in the ER. Science 2008;321(5888):569‐572. 128.  Hosoda  A,  Kimata  Y,  Tsuru  A, Kohno  K.  JPDI,  a  novel  endoplasmic reticulum‐resident  protein  containing both  a  BiP‐interacting  J‐domain  and thioredoxin‐like  motifs.  J  Biol  Chem 2003;278(4):2669‐2676. 

129.  Dong M, Bridges JP, Apsley K, Xu Y,  Weaver  TE.  ERdj4  and  ERdj5  are required  for  endoplasmic  reticulum‐associated  protein  degradation  of misfolded surfactant protein C. Mol Biol Cell 2008;19(6):2620‐2630. 130.  Walczak CP, Tsai B. A PDI  family network acts distinctly and coordinately with  ERp29  to  facilitate  polyomavirus infection. J Virol 2011;85(5):2386‐2396. 131.  Pasetto M, Barison E, Castagna M, Della  Cristina  P,  Anselmi  C,  Colombatti M.  Reductive  activation  of  type  2 ribosome‐inactivating  proteins  is promoted  by  transmembrane thioredoxin‐related protein. The Journal of  biological  chemistry 2012;287(10):7367‐7373. 132.  Riemer  J,  Hansen  HG, Appenzeller‐Herzog  C,  Johansson  L, Ellgaard  L.  Identification  of  the  PDI‐family member ERp90 as an interaction partner  of  ERFAD.  PLoS  One 2011;6(2):e17037. 133.  Moore  P,  Bernardi  KM,  Tsai  B. The  Ero1alpha‐PDI  redox  cycle regulates  retro‐translocation  of  cholera toxin.  Mol  Biol  Cell  2010;21(7):1305‐1313. 134.  Bernasconi  R,  Soldà  T,  Galli  C, Pertel  T,  Luban  J,  Molinari  M. CyclosporineA‐sensitive,  cyclophillinB‐dependent  endoplasmic  reticulum‐associated  protein  degradation.  PLoS One 2010;5(9):e13008. 135.  Greenblatt EJ, Olzmann JA, Kopito RR.  Derlin‐1  is  a  rhomboid pseudoprotease  required  for  the dislocation  of  mutant  alpha‐1 antitrypsin  from  the  endoplasmic reticulum.  Nat  Struct  Mol  Biol 2011;18(10):1147‐1152. 136.  Zettl  M,  Adrain  C,  Strisovsky  K, Lastun V, Freeman M. Rhomboid  family pseudoproteases  use  the  ER  quality control  machinery  to  regulate intercellular  signaling.  Cell 2011;145(1):79‐91. 

137.  Schubert  U,  Anton  LC,  Gibbs  J, Norbury  CC,  Yewdell  JW,  Bennink  JR. Rapid degradation of a  large  fraction of newly  synthesized  proteins  by proteasomes.  Nature 2000;404(6779):770‐774. 138.  Vabulas  RM,  Hartl  FU.  Protein synthesis  upon  acute  nutrient restriction  relies  on  proteasome function. Science 2005;310(5756):1960‐1963. 139.  Yewdell  JW.  DRiPs  solidify: progress  in  understanding  endogenous MHC  class  I  antigen  processing.  Trends Immunol 2011;32(11):548‐558. 140.  Gidalevitz  T,  Ben‐Zvi  A,  Ho  KH, Brignull  HR,  Morimoto  RI.  Progressive disruption  of  cellular  protein  folding  in models  of  polyglutamine  diseases. Science 2006;311(5766):1471‐1474. 141.  Gidalevitz  T,  Krupinski  T,  Garcia S,  Morimoto  RI.  Destabilizing  protein polymorphisms  in  the  genetic background  direct  phenotypic expression  of  mutant  SOD1  toxicity. PLoS Genet 2009;5(3):e1000399. 142.  Tsai YC, Mendoza A, Mariano JM, Zhou M, Kostova Z, Chen B, Veenstra T, Hewitt  SM,  Helman  LJ,  Khanna  C, Weissman  AM.  The  ubiquitin  ligase gp78  promotes  sarcoma  metastasis  by targeting KAI1 for degradation. Nat Med 2007;13(12):1504‐1509. 143.  Liang  JS,  Kim  T,  Fang  S, Yamaguchi  J, Weissman  AM,  Fisher  EA, Ginsberg  HN.  Overexpression  of  the tumor autocrine motility factor receptor Gp78, a ubiquitin protein  ligase,  results in  increased  ubiquitinylation  and decreased  secretion  of  apolipoprotein B100  in  HepG2  cells.  J  Biol  Chem 2003;278(26):23984‐23988. 144.  Chen  X,  Tukachinsky  H,  Huang CH,  Jao  C,  Chu  YR,  Tang HY, Mueller  B, Schulman  S,  Rapoport  TA,  Salic  A. Processing  and  turnover  of  the Hedgehog  protein  in  the  endoplasmic reticulum.  J  Cell  Biol  2011;192(5):825‐838. 

145.  Nunziante  M,  Ackermann  K, Dietrich  K,  Wolf  H,  Gadtke  L,  Gilch  S, Vorberg  I,  Groschup  M,  Schatzl  HM. Proteasomal  dysfunction  and endoplasmic  reticulum  stress  enhance trafficking  of  prion  protein  aggregates through  the  secretory  pathway  and increase  accumulation  of  pathologic prion  protein.  The  Journal  of  biological chemistry 2011;286(39):33942‐33953. 146.  Yamasaki  S,  Yagishita  N, Tsuchimochi K, Nishioka K, Nakajima T. Rheumatoid  arthritis  as  a  hyper‐endoplasmic‐reticulum‐associated degradation  disease.  Arthritis  Res  Ther 2005;7(5):181‐186. 147.  Joshi  B,  Li  L,  Nabi  IR.  A  role  for KAI1  in  promotion  of  cell  proliferation and mammary gland hyperplasia by the gp78  ubiquitin  ligase.  The  Journal  of biological  chemistry 2010;285(12):8830‐8839. 148.  Wang  F,  Song  W,  Brancati  G, Segatori  L.  Inhibition  of  ER‐associated degradation  rescues  native  folding  in loss  of  function  protein  misfolding diseases.  The  Journal  of  biological chemistry 2011;286(50):43454‐43464. 149.  Bernasconi  R,  Galli  C,  Noack  J, Bianchi  S,  de  Haan  CA,  Reggiori  F, Molinari  M.  Role  of  the  SEL1L:LC3‐I Complex as an ERAD Tuning Receptor in the  Mammalian  ER.  Mol  Cell 2012;46(6):809‐819. 150.  Kincaid  MM,  Cooper  AA. ERADicate  ER  stress  or  die  trying. Antioxid Redox Signal 2007;9(12):2373‐2387. 151.  Pincus D, Walter P. A first line of defense against ER stress. The Journal of cell biology 2012;198(3):277‐279. 152.  Cambridge SB, Gnad F, Nguyen C, Bermejo JL, Kruger M, Mann M. Systems‐wide  proteomic  analysis  in mammalian cells  reveals  conserved,  functional protein  turnover.  J  Proteome  Res 2011;10(12):5275‐5284. 153.  Wu  Y,  Termine  DJ,  Swulius  MT, Moremen  KW,  Sifers  RN.  Human 

endoplasmic reticulum mannosidase I is subject  to  regulated  proteolysis.  J  Biol Chem 2007;282(7):4841‐4849. 154.  Termine DJ, Moremen KW, Sifers RN.  The  mammalian  UPR  boosts glycoprotein  ERAD  by  suppressing  the proteolytic  downregulation  of  ER mannosidase  I.  J  Cell  Sci  2009;122(Pt 7):976‐984. 155.  Le  Fourn  V,  Gaplovska‐Kysela  K, Guhl  B,  Santimaria  R,  Zuber  C,  Roth  J. Basal  autophagy  is  involved  in  the degradation  of  the  ERAD  component EDEM1.  Cell  Mol  Life  Sci 2009;66(8):1434‐1445. 156.  Reggiori  F,  Monastyrska  I, Verheije MH, Cali T, Ulasli M, Bianchi S, Bernasconi  R,  de  Haan  CA,  Molinari  M. Coronaviruses Hijack  the LC3‐I‐positive EDEMosomes,  ER‐derived  vesicles exporting  short‐lived  ERAD  regulators, for  replication.  Cell  Host  Microbe 2010;7(6):500‐508. 157.  Gauss R, Kanehara K, Carvalho P, Ng DT, Aebi M. A complex of Pdi1p and the  mannosidase  Htm1p  initiates clearance  of  unfolded  glycoproteins from  the  endoplasmic  reticulum.  Mol Cell 2011;42(6):782‐793. 158.  Hori O,  Ichinoda F, Yamaguchi A, Tamatani  T,  Taniguchi  M,  Koyama  Y, Katayama  T,  Tohyama  M,  Stern  DM, Ozawa K, Kitao Y, Ogawa S. Role of Herp in  the  endoplasmic  reticulum  stress response.  Genes  Cells  2004;9(5):457‐469. 159.  Miura H, Hashida K, Sudo H, Awa Y,  Takarada‐Iemata  M,  Kokame  K, Takahashi  T,  Matsumoto  M,  Kitao  Y, Hori  O.  Deletion  of  Herp  facilitates degradation of cytosolic proteins. Genes Cells 2010;15(8):843‐853. 160.  Mueller  B,  Lilley  BN,  Ploegh  HL. SEL1L, the homologue of yeast Hrd3p, is involved in protein dislocation from the mammalian  ER.  J  Cell  Biol 2006;175(2):261‐270. 161.  Cattaneo  M,  Lotti  LV,  Martino  S, Alessio  M,  Conti  A,  Bachi  A,  Mariani‐

Costantini  R,  Biunno  I.  Secretion  of novel  SEL1L  endogenous  variants  is promoted  by  ER  stress/UPR  via endosomes and  shed vesicles  in human cancer  cells.  PLoS  One 2011;6(2):e17206. 162.  Xie  Y,  Avello  M,  Schirle  M, Whinnie  EM,  Feng  Y,  Bric‐Furlong  E, Wilson C, Nathans R, Zhang J, Kirschner MW, Huang SM, Cong F. Deubiquitinase FAM/USP9X  interacts  with  the  E3 ubiquitin ligase SMURF1 and protects it from  ligase  activity‐dependent  self‐degradation.  The  Journal  of  biological chemistry 2012;288:2976‐2985. 163.  Guo X, Shen S, Song S, He S, Cui Y, Xing G, Wang J, Yin Y, Fan L, He F, Zhang L.  The  E3  ligase  Smurf1  regulates Wolfram  syndrome  protein  stability  at the  endoplasmic  reticulum. The  Journal of  biological  chemistry 2011;286(20):18037‐18047. 164.  Shmueli  A,  Tsai  YC,  Yang  M, Braun  MA, Weissman  AM.  Targeting  of gp78  for  ubiquitin‐mediated proteasomal  degradation  by  Hrd1: cross‐talk  between  E3s  in  the endoplasmic  reticulum.  Biochem Biophys Res Commun 2009;390(3):758‐762. 165.  Ballar P, Ors AU, Yang H, Fang S. Differential  regulation  of CFTRDeltaF508  degradation  by ubiquitin  ligases  gp78  and  Hrd1.  Int  J Biochem Cell Biol 2010;42(1):167‐173. 166.  Tcherpakov  M,  Broday  L, Delaunay  A,  Kadoya  T,  Khurana  A, Erdjument‐Bromage  H,  Tempst  P,  Qiu XB,  DeMartino  GN,  Ronai  Z.  JAMP optimizes  ERAD  to  protect  cells  from unfolded  proteins.  Mol  Biol  Cell 2008;19(11):5019‐5028. 167.  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(22):4268‐4281. 168.  Durcan  TM,  Kontogiannea  M, Bedard  N,  Wing  SS,  Fon  EA.  Ataxin‐3 

Deubiquitination  Is  Coupled  to  Parkin Ubiquitination  via  E2  Ubiquitin‐conjugating  Enzyme.  The  Journal  of biological  chemistry  2012;287(1):531‐541. 169.  Leitman J, Ron E, Ogen‐Shtern N, Lederkremer GZ.  Compartmentalization of  Endoplasmic  Reticulum  Quality Control  and  ER‐Associated Degradation Factors. DNA Cell Biol 2013;32(1):2‐7. 170.  Olzmann  JA,  Kopito  RR, Christianson  JC.  The  Mammalian Endoplasmic  Reticulum‐Associated Degradation  System.  Cold  Spring  Harb Perspect Biol 2012. 171.  Kny  M,  Standera  S,  Hartmann‐Petersen R, Kloetzel PM, Seeger M. Herp regulates  Hrd1‐mediated  ubiquitylation in  a  ubiquitin‐like  domain‐dependent manner.  The  Journal  of  biological chemistry 2011;286(7):5151‐5156. 172.  Laney  JD,  Hochstrasser  M. Analysis  of  protein  ubiquitination.  Curr Protoc Protein Sci 2002;Chapter 14:Unit 14 15. 173.  Morito  D,  Hirao  K,  Oda  Y, Hosokawa  N,  Tokunaga  F,  Cyr  DM, Tanaka  K,  Iwai  K,  Nagata  K.  Gp78 cooperates  with  RMA1  in  endoplasmic reticulum‐associated  degradation  of CFTRDeltaF508.  Mol  Biol  Cell 2008;19(4):1328‐1336. 174.  Jo Y, Lee PC, Sguigna PV, DeBose‐Boyd RA.  Sterol‐induced degradation of HMG  CoA  reductase  depends  on interplay  of  two  Insigs  and  two ubiquitin  ligases,  gp78  and  Trc8. Proceedings of the National Academy of Sciences of the United States of America 2011;108(51):20503‐20508.  

Figure Legends

Fig. 1 Role of ERAD in cellular proteostasis. A Nascent polypeptide chains are co-translationally translocated

into the ER. Most nascent chains attain the native structure (step 1) and are exported from the folding

compartment (step 2). A variable percentage of the newly synthesized polypeptides enter off pathways of the

folding programs (step 3) and is removed from the ER by constitutive ERAD (step 4). In healthy cells, folding

intermediates are protected from recognition by ERAD factors and have sufficient time to complete the folding

process. B Dysregulated ERAD as a consequence of elevated luminal levels of ERAD factors may cause

inappropriate disposal of intermediates of ongoing folding programs (step 5) thereby reducing production of

native polypeptides (steps 1 and 2).

Fig. 2 Substrate-specific ERAD pathways. The same misfolded module shows different requirements for

efficient clearance from the ER when anchored to (ERAD-LM substrate) and when detached from the membrane

(ERAD-LS substrate), the latter showing stronger dependency on ERAD lectins and HRD1 dislocons.

Consistently, OS-9 and XTP3-B are depicted in lighter colors for ERAD-LM substrates and red arrows show that

ERAD-LM substrates can diffuse in the lipid bilayer to enter HRD1 or alternative dislocation machineries.

Fig. 3 ERAD regulation by misfolded proteins. A In healthy cells, unused dislocation machineries are

disassembled. E3 ubiquitin ligases may for example poly-ubiquitylate themselves or components of the

supramolecular complexes thereby resulting in their degradation. Orphan components may also be segregated

from the ER lumen as shown by the vesicle-mediated release of SEL1L, EDEM1 and OS-9 (149). B Misfolded

proteins engage ERAD factors thereby preserving ERAD complexes, delaying ERAD factors turnover and

segregation from the ER.