Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Nov 1.
Published in final edited form as: Biochim Biophys Acta. 2013 Mar 15;1833(11):2410–2424. doi: 10.1016/j.bbamcr.2013.03.007

Orchestration of secretory protein folding by ER chaperones

Tali Gidalevitz a, Fred Stevens b, Yair Argon c
PMCID: PMC3729627  NIHMSID: NIHMS468940  PMID: 23507200

Abstract

The endoplasmic reticulum is a major compartment of protein biogenesis in the cell, dedicated to production of secretory, membrane and organelle proteins. The secretome has distinct structural and post-translational characteristics, since folding in the ER occurs in an environment that is distinct in terms of its ionic composition, dynamics and requirements for quality contol. The folding machinery in the ER therefore includes chaperones and folding enzymes that introduce, monitor and react to disulfide bonds, glycans, and fluctuations of luminal calcium. We describe the major chaperone networks in the lumen and discuss how they have distinct modes of operation that enable cells to accomplish highly efficient production of the secretome.

Keywords: endoplasmic reticulum, chaperones, protein folding, secretome


The great majority of extracellular, surface-expressed, secretory and endosomal eukaryotic proteins, totaling over 18,000 in the secreted protein database [1], fold and assemble in the endoplasmic reticulum (ER). Thus, in addition to being a major site for lipid biosynthesis and a major calcium store (reviewed elsewhere in this issue), folding of proteins is one of the main functions of the ER. Precise control over the rate of production and the “quality” of their structure is essential for every aspect of metazoan organisms, from morphogenesis during the embryonic development, to the complex endocrine regulation in aging. In this review, we will discuss how the complex processes of protein folding and assembly are supported by the specialized conditions in the ER.

1. Restricting an infinite folding landscape

Anfinsen’s axiom that the necessary information for folding of a protein is encoded within the amino acid sequence itself was derived from spontaneous refolding of pure protein samples after denaturation [2, 3]. Though undoubtedly true, this principle does not suffice to explain protein folding in the cell, because the theoretical “folding space” is too vast [4], and therefore needs to be restricted to be compatible with biological constraints. Most importantly, the wide span of time-scales observed for protein folding suggests a paradigm distinct from random sampling of all possible conformations. Various mechanisms of limiting the conformational space have been proposed, including a bias towards native-like interactions [5], formation of intermediate states on the folding pathway [6], funneling into a particular folding pathway [7] with optional errors [8], and folding down the energy landscape through multiple pathways [9].

1.1. Chaperones limit the available conformational space

Molecular chaperones, by selectively interacting with certain sequence and structural elements, can both contribute to limiting the conformational space and favor particular folding pathways. For example, we traditionally consider that binding of hydrophobic sequences by HSP70 chaperones serves to protect nascent chains and folding intermediates from aggregation, by shielding these recognition regions from intermolecular hydrophobic interactions [10]. However, another consequence of HSP70 binding may be retarding the collapse of these hydrophobic regions into the core of the forming protein. Shielding of HSP70-target peptides from intramolecular hydrophobic interactions early in the folding process would favor a different final spatial arrangement of this sequence, compared to the arrangement that could result if it was allowed to participate in an early hydrophobic collapse. In fact, the hydrophobic residues that are involved in misfolding and aggregation are often the same residues that form chaperone recognition sites. For example, the ER HSP70 chaperone, BiP, binds to immunoglobulin (Ig) light chain (LC) mainly through two peptides on the LC variable domain [11, 12]. One of these two peptides is directly involved in the formation of amyloid fibrils by the variable domain, as shown in [13]. Interestingly, these two dominant BiP-binding sites are located in the center of each of the two β sheets, and the BiP-binding residues face the core of the β sandwich in the folded protein [12]. The release of BiP from the variable domain is coupled to the formation of the stabilizing disulfide bond [11]; BiP binding then serves to delay the closing of the sandwich. Thus, by retarding the incorporation of certain hydrophobic regions into the hydrophobic core, chaperone binding favors a different folding path leading to specific contacts. Finally, ATP-driven cycles of chaperone binding, and enzyme-like actions of protein disulfide isomerases (PDIs) and protein prolyl isomerases (PPIs), also contribute to conformational remodeling during the folding process, by allowing proteins to escape unproductive intermediate states stabilized by non-native interactions. So far, we lack good understanding of how chaperones recognize such stabilized states.

1.2. Consequences of vectorial synthesis

The vectorial synthesis of polypeptides provides yet another means of restricting the conformational space, by controlling the order of folding steps. Because the emergent nascent chain starts folding concomitantly with synthesis, the folding opportunities for α helices and β sheets are vastly different: helix-forming amino acids are translocated sequentially and therefore are available immediately to form all the interactions that stabilize helices. Formation of β sheets, on the other hand, must be severely delayed by the translocation process, since each peptide that would assume an extended strand conformation must ‘wait’ until the other strands are synthesized before hydrogen bonds can be satisfied. This difference implies that if such secondary structure elements form the initial folding intermediates, then α helical proteins have a kinetic advantage in the ER. This difference also explains why the peptide binding chaperones of the HSP70 family evolved to bind extended β strand peptides. In addition to structural elements like β sheets, the organization of certain domains may require slowing down or pausing their folding until the entire domain has emerged from the translocon. For example, a C-type lectin fold, present in multiple metazoan extracellular proteins, has a loop-like form, connected at its base by a disulfide bond and an antiparallel β-sheet formed by N- and C-terminal β strands (β1, β5) coming close together [14]. In such arrangements, not only does the N-terminal β strand need to be prevented from inappropriately participating in the hydrophobic collapse, but the formation of the disulfide bond by the cysteine in the N-terminus has to be delayed as well. Other examples of such delay are the folding of influenza hemagglutinin [1517] and chorionic gonadotropin β chain [18], where binding of molecular chaperones is used to direct the folding pathways in vivo.

Vectorial synthesis has another global effect on the folding options available for large proteins. The organization of large proteins into a series of independently folding domains restricts the conformation space: as each domain folds, the number of options for the rest of the sequence are reduced. However, the complexity of the proteome and the oligomeric structure of many proteins mean that even this simple rule may not be sufficient, as discussed in the examples below.

2. Unique aspects of protein folding in the ER

The conditions for folding within the ER differ significantly from those in other cellular compartments like the cytosol, nucleus or mitochondria. The molecular crowding in the ER is 3–6X higher than in the cytosol [19], the redox potential is 1000X more oxidizing [20], free Ca++ concentration oscillates and can reach 1mM [21], and carbohydrates and a glycosylation machinery are unique to the ER. These conditions impact the energetics and kinetics of protein folding and provide the evolutionary pressure for the presence of specialized enzymes and chaperones to facilitate folding in the ER. Indeed, the most abundant luminal proteins (e.g. BiP, GRP94, PDI, HSP47 and calreticulin, see Table I) all function in protein folding. We review the impact of these unique conditions first, and then discuss how the folding machinery ‘responds’ to them. Our main focus is folding of proteins in the ER lumen and the folding of ER membrane proteins is considered elsewhere.

Table I.

Relative of abundance of selected ER proteins.

Protein Function 3T31 HeLa2 Dendritic
Cells3
Caco23 Pancreatic
RER4
Chaperones
BiP=GRP78 HSP70 family chaperone 100.00 100.00 100.00 100.00 100.00
GRP170=Hyou1 BiP exchange factor 11.14 12.22 20.62 18.18 12.00
Sil1=BAP BiP exchange factor 0.96 0.34 0.11 0.55 0.10
ERdj3=DnaJB11 J domain co-chaperone 4.21 7.65 5.38 6.09 5.80
ERdj5=DnaJC10 J domain; reductase 1.17 0.89 1.01 0.95 40.00
P58ipk=DnaJC3 J domain co-chaperone 1.35 3.20 1.94 3.69
Sec63 Translocon J domain 0.37 9.98 0.61 6.88 39.60
ERdj1=MTJ1=DnaJC1 J domain co-chaperone 0.09 0.47 N.D. 1.15 7.20
HSP47=Serpm H1 Collagen specific 204.29 62.09 N.D. 81.47
GRP94 HSP90 family chaperone 63.44 90.96 70.65 123.48
CNPY3 Client-restricted GRP94 co-chaperone 0.94 2.91 2.85 4.44

Prolyl isomerases
Cyclophilin B CsA-sensitive folding/ERAD facilitator 204.00 165.00 34.00 208.00
FKBP13 FK506/rapamycin binder 10.02 9.57 7.00 10.96
FKBP23 FK506/rapamycin binder 1.74 1.05 N.D. 0.00
FKBP65 FK506/rapamycin binder 5.25 11.06 3.67

Glycoprotein quality control
Calnexin Membrane lectin chaperone 19.76 56.75 17.84 97.65
Calreticulin Lumenal lectin chaperone 39.45 129.61 35.40 164.94
UGT1 Glucosyl-transferase 2.91 3.98 65.11 3.72
Ribophorin I Oligosaccharyl transferase complex 17.96 N.D. 14.87 42.70
Ribophorin II Oligosaccharyl transferase
complex
15.23 24.64 8.54 16.66
Glucosidase IIα Glycan trimming 14.33 33.94 28.24
Glucosidase IIβ Glycan trimming 9.50 32.74 4.76 33.61
OS-9 ERAD lectin N.D. 0.31 0.71
Erlectin=XTP3-B ERAD lectin 0.66 0.00 0.44
Mannosidase Man1b1 Glycan trimming 0.18 0.40 0.31
EDEM2 ERAD lectin N.D. 0.04 0.06
EDEM3 ERAD mannosidase N.D. 0.28 0.11

Redox proteins
Erol1 Redox generation 5.28 0.02 0.40 15.47
ERp44 5.30 11.06 5.10 19.42
ERp29 9.37 31.96 18.75 47.32
PDIA1 = PDI Isomerase and chaperone 78.66 109.76 27.69 91.24
PDIA3 = ERp57 Glycoprotein preference 82.51 70.26 64.81 91.58
PDIA4 = Erp72 16.09 28.58 15.01 43.87
PDIA5 = Pdir 1.16 1.34 0.13 3.92
PDIA6 = P5 30.95 21.35 30.34 36.83
Prdx4 Redox generation 18.74 31.90 2.35 6.88

Data derived from published quantitative proteomic methods. The abundance of BiP/GRP78 in each dataset was set to 100 and the expression values of the other proteins were normalized to it.

1

3T3 proteomic data from [108].

3

D4-positive dendritic cells, peak intensity counts reported in [231].

2

eLa cell and Caco2 cell proteomics from [232], using peptide spectral counts. Similar results were reported by [233], based on SILAC quantitation.

4

Values from [234] and are based on purification of proteins.

2.1. Proximity to the membranes

One consequence of the vectorial translocation of the nascent chain into the ER lumen is that it is positioned in proximity to the membrane, which can have detrimental effects on folding. Negatively charged membrane surfaces have been shown to influence conformational transitions and induce aggregation and fibrilogenesis of Ig LCs [22], lysozyme [23], and a proteolytic collagen fragment endostatin [24], as well as a variety of non-secretory proteins [25]. As there is no ER equivalent to the cytosolic ribosome-associated chaperones or to chaperonins, the protection of the emerging nascent chain from interaction with membrane surfaces and its channeling into folding pathways is mediated by the close spatial proximity of the BiP molecules which gate the translocon [26] and the oligosaccharide transferase [27].

2.2. Folding in the viscous luminal environment

One constraint on folding in the ER lumen is its high macromolecular density, which was measured to be 3–6 fold higher than in the cytosol and 9–18X higher than in typical aqueous buffers [19]. The viscosity of the ER lumen is high not only because of the presence of many proteins in various states of folding, including the high concentration of ER-resident chaperones, but also because of other macromolecules, estimated to occupy up to 30% of the volume [28]. Nonetheless, the diffusional mobility of GFP-tagged chaperones is only moderately slowed in the ER compared to the cytosol (E.L. Snapp, personal communication). The viscous environment is one evolutionary reason for molecular chaperones, whose activity imposes a timing sequence on the folding pathway that is not emulated in monodisperse in vitro systems. Molecular crowding clearly also affects the activity of the folding machinery, for example the enzymes that trim glycans [29].

2.3. Consequences of glycosylation

Carbohydrates serve several roles in protein folding. First, because of their hydrophilic nature, carbohydrates increase the solubility of the glycoprotein. Second, they generally mark the surface of folding modules and are not buried within them. Third, they make the process of translocation across the ER membrane less reversible by increasing the energy barrier to back-translocation. Apart from these general roles, however, carbohydrates also affect protein folding, albeit in individualized and idiosyncratic ways. Detailed studies with glycoproteins whose Asn-linked glycosylation sites were mutated systematically showed that in many cases glycosylation is needed for proper folding: under-glycosylated proteins form intracellular aggregates and are retained in the ER [3032]. Somewhat paradoxically, however, almost no individual glycan is necessary for folding of VSV G protein or influenza hemagglutinin [33, 34]. Only one of seven hemagglutinin glycans, attached to Asn81, created a kinetic barrier to folding [32], and even this glycan effect could not be detected in another set of experiments [34]. Among the highly homologous MHC class I molecules, some are sensitive to the presence of glycans [35], whereas others fold equivalently in the presence or absence of the carbohydrates [36]. Furthermore, even proteins that are normally not glycosylated can sometimes benefit from the inclusion of ectopic glycans [37]. Thus, the ‘rules’ that govern the interplay between glycosylation and folding are at present still poorly understood. In part, the explanation lies in the observation that aromatic and hydrophobic amino acids are over-represented near those N-glycosylation sites that are important for proper folding [38]. This correlation was shown experimentally to be important because in such ‘aromatic sequons’ the aromatic side chain interacts with the first N-acetylglucosamine of the glycan, and grafting such sequons onto non-glycosylated proteins increases their stability [39]. Table II provides comparative examples from several proteins where some N-glycosylation sites are known to impact, but others are dispensable for folding. This compilation further reinforces the potential relevance of neighboring hydrophobic residues. We suggest that for certain sites, an additional role for attachment of glycans and the subsequent association with chaperones is to overcome the thermodynamic tendency of hydrophobic patches in close proximity to these sites to avoid display on the protein surface, thus nucleating particular folding pathways.

Table II.

Correlation between N-glycosylation sites that are important for folding and hydrophobic residues neighboring them.

Protein Sequence Foldinga Comment Ref.
Mouse Tyrosinase VFYN86RT Necessary [38]
Mouse Tyrosinase GDEN230FT Dispensable
Mouse Tyrosinase RTAN337FS Dispensable Necessary in human tyrosinase [235, 236]
Mouse Tyrosinase IFMN371GT Necessary
Human MHC class Ib GYYN86QS Necessary [237]
Human α3 integrin AIYVFMN349QS Impedes Gain-of-function mutant [238]
Rat Adenylate cyclase 6 RQIN791YS Dispensable But needed for function [239]
Rat Adenylate cyclase 6 ASSN876ET Dispensable But needed for function
Human PDIA2 TLKFFRNGN127RT Dispensable [240]
Human PDIA2 ILNHLLLFVN248OT Important
Human PDIA2 AAPFPEPPAN516ST Dispensable
Rat ICAM-5 RGGSLWLN54CS Necessary [241]
Rat ICAM-5 ETSLRRN74GT Dispensable
Human Chymotrypsin C SLQYLKN52DT Necessary Not for function [242]
Human Chymotrypsin C LNCQLEN226GS Dispensable

Sequences upstream of N-glycosylation sites are shown with aromatic and aliphtic residues underlined. Where information is available, sites that affect folding and sites that do not were chosen.

a

Designation of the glycosylation site as necessary for proper folding, important but not essential, impeding or dispensible.

b

The sole site in human HLA; rodent class I are functional without their N-glycosylation. Note that all sites are further than 50 amino acids from the N-terminus, satisfying the prediction made in [166].

2.4. Folding in an oxidizing environment

A general property distinguishing proteins that fold in the ER from cytosolic proteins is the preponderance of disulfide bonds. Cys residues in newly synthesized secretory polypeptides tend to oxidize in the ER lumen because of its high oxidative redox potential. In the past decade there has been a major shift in the understanding of the redox buffer in the ER, from an emphasis on a glutathione-buffering system [20] to the realization that the eukaryotic ER uses a mechanism similar to the bacterial periplasm [40] to maintain its redox potential: a protein-based relay of oxidation/reduction reactions (Reviewed in [41, 42]). The relay involves Erol, a conserved FAD-dependent enzyme, which is oxidized by molecular oxygen and in turn acts as a specific oxidant of protein disulfide isomerase (PDI), which then directly oxidizes disulfide bonds in folding proteins. In addition to Erol, PDIs can be oxidized by peroxiredoxin IV, which metabolizes the H2O2 formed by the Erol reaction, couples this oxidation to disulphide formation [4345], and provides a parallel pathway for oxidative folding in the ER [46].

The use of molecular oxygen as the terminal electron acceptor can lead to oxidative stress through the production of reactive oxygen species and oxidized glutathione. That cellular oxidative stress affects protein folding is shown, for example, by the slower maturation of LDL receptor under high H2O2 concentrations [47].

Disulfide bonds are very important in dictating folding pathways, because they form covalent interactions that stabilize folding intermediates [48] and severely restrict the landscape of available conformations. Formation of disulfide bonds begins very early in the life of the protein, sometimes as soon as the required Cys residues are available in the lumen [49, 50]. On the other hand, some disulfide bonds do not form until much later in the folding sequence, even within the same domain [16, 51]. Such dichotomy suggests that in vivo there are mechanisms that selectively delay some oxidation steps. As a further apparent contradiction, proteins can undergo post-translational oxidation and achieve the same native structure as with co-translational oxidation. Addition of reducing agents to the medium of live cells prevented disulfide bond formation in newly synthesized influenza hemagglutinin or asialoglycoprotein receptor and reduced the already oxidized glycoproteins inside the ER. When the reductant was washed out, the reduced proteins rapidly oxidized, folded correctly and assembled [52, 53]. These examples show that apparently oxidation follows the same pathway when it occurs post- or co-translationally and underscore the concept that in most cases disulfide bonds serve to stabilize a local fold, achieved by multiple cooperative interactions, rather than to initiate the folding. In support of this view, mutagenesis experiments show that lysozyme can be reasonably well folded even without some disulfide bonds [54] and that if an Ig is made stable enough via genetic engineering, it can withstand the removal of its disulfide bonds [55].

All the above observations also indicate the need for precise control over protein oxidation in the ER. As discussed in section 4.4, the PDIs are capable of shuffling disulfide bonds, stabilizing proper intermediates, and resolving aberrant disulfide bonds. One indication of the importance of the control over disulfide bonds is that the level of expression of PDIs in pancreatic β cells is proportional to insulin expression, consistent with oxidative folding being an important aspect of this hormone’s biogenesis [56].

2.5. Influence of calcium on folding

The ER is a major calcium store in the cell and therefore has to respond to the metabolic needs of the cell by rapid changes in free luminal Ca++ levels. This must impact protein folding in the lumen, but the information available is surprisingly superficial. In principle, Ca++ can influence folding directly through binding to amino acids side chains. Perturbation of Ca++ levels also affects folding of luminal proteins indirectly, in a manner more pertinent for this review: several members of the folding machinery in the ER, including GRP94, BiP, calreticulin, ERp72, PDIA1, and PDIA6, are Ca++ binders (e.g. [57]) and their interactions can be modulated by Ca++, as suggested by interaction experiments [5860]. The refolding activity of ERp72 and PDIA6 [61] and the peptide-binding activity of GRP94 [62] are modulated by Ca++. Whether folding of a given protein is affected by Ca++ directly or indirectly, is not easily predictable. Clearly, different secretory proteins respond differently to Ca++ changes: depletion of Ca++ has no effect on folding or secretion of albumin, but severely inhibits asialoglycoprotein receptor and α1-antitrypsin maturation [63, 64]; The initial assembly step of large and small subunits of Heymann nephritis antigenic complex is calcium-dependent, but a later step is not [65]. Under low luminal Ca++, thyroglobulin is retained in the ER, but interestingly, it exits the calnexin/calreticulin cycle prematurely, while its interactions with the other chaperone axis of BiP and GRP94 is stabilized and prolonged [66, 67]. Conversely, for other proteins, like TCRα, chelation of Ca++ causes release from BiP [60]. Thus, the roles of chaperones can change depending on the Ca++ concentration in the ER, but the resulting effects depend on the substrate protein.

2.6. A distinct spectrum of folds in secreted proteins

All the constraints on folding in the ER that are discussed here presumably provided evolutionary pressure that favored certain types of folds in secreted proteins and cell-surface receptors. There are several examples of types of domains that are over-represented in secretory proteins and under-represented in other types of proteins. One example is the immunoglobulin fold, which consists of a sandwich of two β-sheets with a Greek key topology and is found in hundreds of proteins of different functions, including integrins, adhesion molecules, interleukin receptors, receptor tyrosine kinases and histocompatibility proteins. The immunoglobulin fold is, however, under-represented among cytoplasmic proteins, and when it is found there, it is one subtype of the immunoglobulin fold (e.g. [68]). A second example is the ligand-binding domain of the low-density lipoprotein receptor. It is ~40 residues long with little recognizable secondary structure organized around a calcium ion [69], stabilized by Cys residues that form three disulfide bonds [69, 70]. Variations of this ligand-binding domain are found in a many kinds of cell surface receptors and conversely are underrepresented in cytosolic proteins, presumably because of the requirement for cysteines oxidation. Yet a third example is the EGF-like domains, present on a large number of membrane or extracellular proteins, but not in cytosolic proteins. These 30–45 residue-long domains consist of a small β sheet and a flexible loop, with three disulfide bonds stabilizing the structure [71].

3. Folding intermediates in vitro and in vivo

In vitro folding experiments often describe proteins with folding intermediates at equilibrium, which can be trapped by non-physiological conditions, like drastic pH shifts. In contrast, folding in vivo is often made non-reversible by post-translational modifications, such as proline or disulfide isomerization, which are driven by enzymes. Tyrosinase, a glycoprotein where the carbohydrates are essential for the native fold, exhibits several inactive intermediates, at least two of which are recognized by calnexin. If the association with calnexin is prevented, folding is actually more rapid, but the resulting protein fails to bind copper and is inactive [72]. In the large family of proteinase inhibitors, some intermediates are kinetically trapped and previously acquired disulfide-linked structures need to unfold in order to progress to the native state [73].

Whether through chaperone action or via the biophysical conditions that exist in the ER, restriction and simplification of the folding pathway is sometimes inherent in the protein: while exponential functions are required to describe the folding kinetics of phosphoglycerate kinase in vitro, when this cytosolic protein is engineered to fold in the ER, the measured folding kinetics are more two-state-like, reflecting the modifications imposed by the intra-luminal conditions [74]

The in vivo conditions sometimes dictate a folding pathway that is not only simpler but actually different from that observed in vitro. For example, although each of the two domains of immunoglobulin light chains can fold autonomously when expressed alone, during immunoglobulin biosynthesis in vivo, the variable domain, which is N-terminal and emerges first, folds after the constant domain [75]. The reason is that BiP continues to engage its V domain peptide binding sites and, if this interaction persists, the light chain folding intermediate is targeted for ER-Associated Degradation (ERAD) [12, 13]. Most light chains variants that form amyloids are mutated in the variable domains and do not complete folding in the cell. The partner heavy chain also has an important in vivo intermediate where the first constant domain is slow to fold [76]. Folding of this domain is coupled to assembly with the light chain, illustrating the importance of folding intermediates in vivo. Neither of the above folding intermediates would have been predicted by in vitro techniques and they illustrate how the complex luminal environment dictates the path to productive folding.

The insulin fold family provides an example for a related concept: intermediate structures necessary for attaining the native state, but not for activity of the protein. The N-terminal 8 amino acids of the B chain dictate the foldability of proinsulin, but once the native state is achieved, this peptide is dispensable for native structure, activity, or stability of mature insulin [77, 78]. Thus, the folding process could, in itself, provide an evolutionary pressure for sequence conservation.

3.1. Oligomeric protein assembly

Once domains and polypeptides fold, many proteins require assembly of subunits before they are biologically active and are competent for export from the ER. Native viral glycoproteins often form non-covalent trimers that assemble in the ER [79] in ATP-dependent fashion [80]. B and T cell receptors or MHC proteins assemble from 2 to 7 subunits into the biologically active entity before they exit the ER. In either case, there is at least one subunit whose folding is a pre-requisite for assembly: in the case of B cell receptor, the folding of the heavy chain is completed only upon subunit assembly [81, 82] and in the case of the T cell receptor - only upon folding of the CD3 epsilon subunit [[83]. A similar situation is observed for nicotinic acetylcholine receptor and likely many other receptors [84]: by stabilizing and sequestering subunits during assembly, chaperones like calnexin, BiP and ERp57 regulate the levels of assembled functional receptors. In general, unassembled subunits either aggregate or are targeted to degradation via the ERAD or autophagy pathways, but are not transported to the Golgi complex.

Intermediates do not only define the folding of protein domains or polypeptide chains, but are also characteristic of oligomeric assembly of many proteins. For example, the Heymann nephritis antigenic complex assembles in two discrete stages. First, a large glycoprotein (gp330) associates with a 44kD subunit early after synthesis, and then, >60 min after synthesis, a larger hetero-oligomer forms, still before acquisition of Endo H resistance [65].

What features enable the cell to discriminate between free subunits and assembled oligomers? No global answer can yet be provided, but in individual cases there are examples of features that may operate. As mentioned above, during Ig assembly there is a mechanism for coupling the folding of a ‘sentinel’ CH1 domain whose folding is retarded [85], to assembly of heavy and light chains by coordinating the formation an intra-domain disulfide bond with a disulfide linking the two chains [51, 81], thus releasing chaperones and folding enzymes ([85, 86], and see below). T cell receptor assembly depends heavily on hetero-pairing of subunits via transmembrane domains that contain charged amino acids [8790]. If left unpaired, these subunits do not assemble and are dislocated from the ER membrane. The presence of unsatisfied intramembrane charge prevents stabilization of the unassembled subunits via homo-oligomerization and maintains them in a retrotranslocation-competent state [91]. The unassembled α chains can be retained in the ER by BiP, in Ca++-dependent fashion [60].

In glutamate transporters, an evolutionarily conserved arginine-based motif acts as an ER retention signal, while a luminal leucine motif is required for suppression of this signal and allows traffic to the Golgi complex [92]. This motif is likely subject to conformation modulation that could occur upon subunit assembly. Another case of conformational control in the ER is the maturation of AMPA receptors to acquire ligand-binding activity [93]. The extracellular region that distinguishes isoforms of AMPA receptor can interact with binding proteins such as stargazin, or with other luminal isoform-specific proteins, to determine whether a transport-incompetent subunit has assembled sufficiently to traffic to the cell surface [94].

4. Chaperone networks in the ER

The complex demands of orchestrating the folding of secreted and membrane proteins require at least 10 ubiquitously expressed chaperones and some client-specific ones (Table I, and other reviews in this issue). These chaperones do not generally work alone, but rather fulfill each task by forming transient networks of chaperones, cofactors and folding enzymes that operate in concert. The composition of these networks appears to adapt to the needs of the folding clients.

4.1. The BiP network

4.1.1. Client recognition

The most abundant chaperone network and the one involved with most client proteins is the BiP/GRP78 network (Table I). BiP is a multifunctional chaperone, involved in nascent chain translocation, folding of polypeptides, resolving and removal of misfolded proteins and monitoring the ER stress transducers (entries in [95]). Consistent with these multiple functions and the variety of known client proteins, BiP is an essential protein. In the cell, it is usually not possible to reduce BiP expression below 40% of its normal level [96] (Argon, unpublished data), and in the whole organism, BiP deletion leads to very early embryonic lethality at the pre-implantation stage [96]. The primary biochemical activity of BiP/GRP78, which unifies all these functions, is binding to 7–11 residue peptides that in the folded state form β strands and whose sequences tend to have alternate hydrophobic residues [97, 98]. These residues are buried in the interior of the native fold but are exposed in early intermediates or in misfolded proteins [12, 99]. The inherent affinity of BiP for peptides is in the range of 1 to 100 µM, as appropriate for a chaperone that had evolved for transient interactions. Higher affinity interactions would have inhibited folding altogether. It is estimated that a BiP-binding site would be created in a random sequence once every 36 amino acids [98]. This frequency would explain why BiP has a large number of client proteins. Yet, domains that have predicted BiP binding sites do not always employ them to a significant degree in the cell, and this accounts for a chaperone-directed folding sequence in which BiP dictates the path of folding [13, 100]. This can be exemplified by BiP binding to Ig light chain molecules: although light chain is predicted to have >5 binding sites in each of the variable (VL) or constant (VC) domains [12], BiP only binds unoxidized VL domain in cells, with no detectable binding to the constant domain [86, 101](Argon, unpublished data). Furthermore, two out of the potential five binding sites in VL are sufficient to explain the BiP-light chain interaction [12]. A combination of in vivo and in vitro folding studies revealed that the constant domain folds rapidly and stably even in the absence of an intradomain disulfide bond [75, 86]. Thus, the simple presence of a BiP-binding site on a nascent chain does not ensure that BiP will bind and play a role in its folding. Instead, it appears that the rate and stability of protein folding determines whether or not a particular site is recognized, with BiP preferentially binding to proteins and domains that fold slowly or are less stable.

4.1.2. Regulation of ATPase cycle

The action of BiP in folding reactions, like its activity in other processes, requires nucleotides. Since there is no in situ ATP generator, the ER lumen depends on mitochondrial ATP that is imported into the ER via a 56 kDa permease [102, 103]. In the lumen, ATP is needed for the action of various proteins. Binding of BiP to clients is very fast in the ATP-bound state, but hydrolysis must occur to stabilize the binding. Exchange of ADP to ATP is then needed to accelerate the release of the client protein. The peptide-binding and release activity of BiP thus requires its ATPase cycle [104], but, because BiP's inherent ATPase activity is low and the spontaneous transition between the two states is extremely slow, its ATPase cycle depends on cofactors.

At least 7 proteins can modulate BiP’s ATPase cycle: Four HSP40-family proteins containing a J domain, two nucleotide exchange factors (Sill/BAP and GRP170/ORP150), and ERdj5, a PDI that also possesses a J domain. The J domain proteins stimulate BiP’s ATP hydrolysis and thus enhance BiP’s interaction with misfolded proteins, while Sill and GRP170 exchange ADP for ATP so that BiP can dissociate from them more readily. The combined action of J proteins and nucleotide exchange factors creates a cycle of binding and release that is key to the mode of action of the BiP machine (reviewed in [105]).

GRP170 can form a stable complex with BiP in the absence of ATP, mediated through contacts between their nucleotide binding domains [106]. This association decreases the affinity of nucleotide binding to BiP but does not stimulate its ATPase activity [106]. It is noteworthy that there is a 9-fold excess of BiP over GRP170 and an even larger excess over Sill or each of the J domain proteins [107, 108] (Table I), underscoring that these cofactors may be limiting components and must operate catalytically. Alternatively the cofactors may be only needed for a very limited set of substrates, but this seems unlikely, since the BiP cycle and not stable association seem to be the norm. Although GRP170 is an HSP70 family protein and has domains that can bind peptides [109, 110], no direct role has yet been demonstrated for GRP170 in folding client proteins in vivo, though it is associated with immunoglobulins in B cells [111] and with misfolded mutants in other cells [112]. These interactions of GRP170 with both BiP and misfolded species may serve to target BiP to a specific pool of client proteins.

J domain proteins may also adapt the BiP complex to the demands of the specific task. Thus, ERdj3, can bind directly to several nascent, unfolded and mutant secretory proteins, even without BiP [113]. After ERdj3-client binding, BiP joins the complex, which leads to dissociation of ERdj3 and the client polypeptide and concomitant stimulation of BiP’s ATPase activity before folding is completed [113]. The higher ATPase activity converts BiP to its high-affinity state for clients, so ERdj3 seems to present the client for BiP-dependent folding [114]. Similarly, P58(IPK)/DnaJC3 presents other clients for BiP-dependent folding [115]. On the other hand, binding of the J protein ERdj5 to misfolded αl-antitrypsin, with its subsequent recruitment of BiP and activation of its ATPase, leads to reduction of the client’s disulfide bond by the PDI activity of ERdj5 and its subsequent targeting by BiP and EDEM1 to disposal via ERAD [116]. How different J domain proteins target BiP is not understood. Although the various J proteins can bind alternatively to the same surface on BiP, the exact modes of binding are not identical, since a mutation in this binding region of BiP abolishes interactions with some J proteins but not with others [117].

4.1.3. Functional consequences of BiP binding

The usual activity of BiP involves ATP-dependent cycles of binding and release of the substrates. The cycling is mediated by the continual exposure of hydrophobic BiP binding sites on the immature folding intermediates. One of the exceptions to the ‘canonical’ binding and release action cycle is the BiP-substrate interaction that is sufficiently prolonged so as to be stable under physiological conditions. A prime example is BiP interaction with the CH1 domain of Ig heavy chains [85]. This domain is intrinsically disordered in vitro, which sets it apart from other Ig domains. CH1 folds only upon interaction with the CL domain of the Ig light-chain. Structure formation proceeds via a trapped intermediate and can be accelerated in vitro by the ER-specific peptidyl-prolyl isomerase cyclophilin B [85]. BiP recognizes incompletely folded states of the CH1 domain and competes for binding to the CL domain. Unlike BiP binding to many other sites in the Ig molecule [12, 13], its binding to the CH1 domain is persistent, ensuring that CH1 remains unoxidized long after other domains have folded [81]. The ‘purpose’ of this built-in delay is to coordinate the formation of the intradomain disulfide with that of the inter-subunit disulfide bond [51, 82], a key step that links completion of the heavy chain subunit folding to the assembly of heavy and light chains [51, 85]. In vivo experiments demonstrate that these steps, including association with a folded CL domain and isomerization of a conserved proline residue, are essential for antibody assembly and secretion in the cell [85]. When chain assembly cannot be completed because the right partner is absent, or when a protein cannot fold due to mutation, the persistent BiP association leads to retention and/or ERAD [51, 76, 81, 118]. The retention of incompletely folded proteins is therefore a consequence of the normal action of the BiP system. Nonetheless, what determines persistent BiP binding to the sites in CH1 vs. short-lived interactions with other sites on the same molecule is not yet known.

In addition to the folding trajectories of their client proteins, chaperones like BiP can influence their aggregation pathways [119]. A mechanistic insight into this process is provided by considering the role of BiP in the folding and aggregation of the amyloidogenic mutant immunoglobulin light chain [120]. Of the two dominant BiP-binding sites important for the folding of VL, one is directly involved in the formation of amyloid fibrils, as the synthetic peptide containing this site specifically inhibits VL amyloidogenesis [12, 13]. Thus, binding to the same binding site mediates BiP’s function in promoting the efficient folding of VL in the cell and in the prevention of amyloid formation by its mutant forms. This conclusion is echoed by the reactivity of the anti-oligomer antibody A11, which, in addition to recognizing aggregated forms of several dissimilar proteins, also recognizes a subset of molecular chaperones, including members of the HSP70 family [121]. A11 binding to these chaperones can interfere with the suppression of aggregation or the refolding of their substrates [121]. Therefore, increased competition for such a chaperone from other substrates, particularly if uncompensated for by a stress response, may favor a shift from productive folding toward amyloidogenesis and the formation of toxic species by the disease-associated protein.

4.1.4. Organismal roles of the BiP network

While BiP is essential in all cells, none of its associated J proteins seem to be essential in metazoa, and as far as we know none is a generalist J protein. Though some of them can bind misfolded polypeptides directly (e.g. ERdj3 [113] or MTJ1 [122]), others function without client binding, and yet others, like Sec63, are specific not for protein folding, but rather for translocation across the ER membrane. Considering the important role of the J domain proteins in regulating BiP’s ATPase cycle, deletions of two J proteins in the mouse have surprisingly specific phenotypes. Deletion of ERdj5 leads to a pronounced salivary gland phenotype, consistent with failure of folding of amylase that induces persistent ER stress response [123]. This phenotype of ERdj5 −/− mice is very tissue-specific, since plasma cells or pancreatic β cells, which are also professional secretory cells, are not affected significantly. On the other hand, deletion of mouse P58(IPK) has a β cell-specific phenotype [124]. The absence of phenotype in unaffected cells may be due to compensation for the loss of specific J domain protein by an unknown component, potentially pointing to a redundancy in the J domain proteins. Alternatively, the cell-specific dependence on a given J domain protein may be dictated by the specific protein clients expressed in these cells, pointing to much more specialized roles of these proteins.

Mutations in the Sill nucleotide exchange factor [125] cause Marinesco-Sjogren syndrome (MSS) in humans [126, 127] and ataxia and neurodegeneration in woozy mouse (wz) [128]. Since Sill, like BiP and the J proteins, is expressed ubiquitously, this pathological presentation suggests selective vulnerability of certain cell types and tissues, most notably the cerebellum. In wz mice, loss of Sill causes ER stress and accumulation of ubiquitinated proteins in cerebellar Purkinje cells, leading to degeneration and apoptosis [128]. Furthermore, even within Purkinje cells there is differential sensitivity to loss of this chaperone, as Purkinje cells in the vestibulocerebellum of wz mice are spared. Since Sill functions to regulate the ADP to ATP exchange, and thus the chaperone cycle of BiP, the reasons for such selective vulnerability are again not immediately obvious. One potential explanation is that another BiP nucleotide exchange factor, GRP170/ORP150 (Lhs1) may be redundant with Sil1 and thus compensates for its absence in protected cells [129]. However, ORP150 is upregulated in affected Purkinje cells of the wz mouse, so this upregulation is apparently insufficient to protect the cells [130]. On the other hand, early ectopic over-expression of ORP150 rescues ER stress, protein aggregation, and neurodegeneration, while its down-regulation exacerbates these phenotypes [130].

A fascinating result is that inactivation of one of the DnaJ proteins, P58(IPK)/DnaJC3, which promotes the ATPase activity and substrate loading of BiP [115, 131], also attenuated the phenotypes of Sil1−/− mice [130]. This result supports the view that disturbance in the BiP chaperone cycle is a cause of the neurodegeneration and other phenotypes in the wz mouse (and by extension, in MSS patients). The simultaneous absence of both the ATPase stimulating activity of P58(IPK) and the nucleotide exchange activity of Sil1 is predicted to decrease cycling of BiP, at least with respect to the specific set of substrates that the two co-chaperones may recognize. Thus, rescue of Sil1 deficiency by deletion of P58(IPK) points to the imbalance between the two opposing regulators of the BiP chaperone cycle as the potential etiological factor, rather then the deficiency of BiP chaperone activity per se. Since the ERdj5 knockout was not tested for genetic interaction with Sil1, it is not yet known whether the rescue property of the P58(IPK)-deficient strain is unique to this J protein, or is shared by all.

Another non-obvious result is that deletion of P58(IPK) by itself, unlike deletion of Sil1, does not cause brain pathology or ataxia, suggesting that sensitivity to the imbalance between J proteins and nucleotide exchange factors may be one-sided and cell-specific. The deletion of P58(IPK) does cause pancreatic β cell failure, due to persistent UPR [124], but the ability of Sil1 deficiency to mitigate the β cell phenotype was not reported [130]. Thus, the sensitivity to the imbalance between J proteins and exchange factors may be one-sided.

In addition to the differential sensitivity of Purkinje cells to Sil1 depletion, there is evidence that they are also generally intolerant of protein misfolding in the ER. A missense mutation (Ser658 to Pro) in a ubiquitously expressed ERAD protein, SEL1L, in dogs leads to selective cerebellum-restricted neurodegeneration with marked loss of Purkinje cells [132]. As expected based on the role of SEL1L in disposal of misfolded proteins from ER, the affected Purkinje cells activated strongly the ER stress response [132]. Yet a third example of selective sensitivity of these cells can be seen in a mouse carrying the sticky (sti) mutation - a missense mutation in the editing domain of the alanyl-tRNA synthetase gene, which leads to mistranslation and possibly misfolding of many cellular proteins [133]. For unknown reasons, this mutation affected Purkinje cells selectively, causing accumulated ubiquitinated inclusions, increased expression of cytoplasmic chaperones and also activated the ER stress response. Persistent CHOP expression, which is thought to be associated with pathological levels of ER stress, was seen through the onset of degeneration, suggesting that perhaps the low but chronic levels of protein misfolding were not tolerated in these susceptible cells [133]. Therefore, the conspicuous sensitivity of Purkinje cells to slowing the BiP cycle may be indicative of their increased sensitivity to altered proteostasis, as in the examples of mistranslation and defective ERAD.

4.2. GRP94

4.2.1. Client recognition

A chaperone that is often associated with the BiP complex is GRP94, the single HSP90 family representative in the ER [134]. When working on a folding pathway together with BiP, GRP94 usually is the second chaperone in the sequence, thought to engage late folding intermediates (e.g., Ig [11, 135, 136], Tg mutant [137, 138]), and this distinction seems to be due to different structural cues to which GRP94 is sensitive. Unlike extended peptides with specific sequence features that are recognized by BiP, no sequence or structural motives have emerged for GRP94. This may simply reflect the relatively restricted number of known GRP94 client proteins; alternatively, GRP94 may have a more complex recognition mode, similar to the HSP90’s ability to recognize metastable and near-native domains. The rules of engaging GRP94 with clients are also still unknown; the chaperone activity of GRP94 in vivo requires ATP [139], but unlike the BiP complex, no accelerator of ATPase nor an exchange factor has been discovered for GRP94. The only identified co-factor for GRP94 so far is CNPY3, which is widely expressed (Table I), but is a client-specific co-chaperone that does not modulate the ATPase cycle of GRP94 [140].

The order of BiP-GRP94 interaction mirrors the HSP70-HSP90 axis in folding of cytoplasmic clients [141], and the position of GRP94 as the second in the sequence likely accounts for its restricted clientele. BiP’s sequence-dependent recognition of slow folding or unstable domains may account for most of the chaperone-dependent secretory proteins that are not engaged with the calnexin-calreticulin cycle; only a subset of these proteins is subsequently engaged by GRP94, guided by the yet unknown thermodynamic or structural features. Such sequential action of chaperones provides a way to restrict the folding options of a client and direct it to a preferred pathway. While it is not clear whether GRP94 interacts physically with BiP, there is obviously a strong functional interaction between them, as demonstrated by the upregulation of expression of either chaperone when the other is silenced genetically or inhibited pharmacologically [142].

4.2.2. Organismal roles of GRP94

The GRP94−/− mouse provides a remarkable example of specific, client-restricted phenotype due to the lack of a chaperone. The embryonic lethality at E6.5–7.5 coincides with the developmental time when mesoderm induction occurs and when IGF-II, a client which is totally dependent on GRP94 activity is first expressed [143]. Another client, Ig, is not so strictly dependent on GRP94, and indeed targeted deletion of the chaperone in B lineage cells does not appreciably depress the circulating levels of antibodies [144]. The phenotypes of tissue-specific deletion of GRP94 also reflect its client interactions: deletion in skeletal muscle mostly affects the muscular growth pathway that depends on IGFs [145], and thrombocytopenia due to deletion in hematopoietic cells is caused by the effect on platelet glycoprotein Ib-IX-V complex [146]. Lack of GRP94 in a B cell line leads to a failure to respond to bacterial endotoxins, due to a selective defect in maturation of Toll-like receptors and integrins [144, 147]. In all of these studies, GRP94 was not required for the cellular viability, which again is consistent with its selectivity.

Like GRP94, at least one other ER chaperone recognizes advanced folding intermediates. HSP47 preferentially recognizes Gly-X-Y repeats in the triple helices of various collagens, in a rather folded conformation [148150]. Unlike GRP94, HSP47 does not require ATP and may dissociate from collagens in a pH-dependent manner [148], though it is not clear how this would be regulated in vivo. Ablation of the mouse gene for HSP47 leads to embryonic lethality at E11.5, accompanied by defective collagen biosynthesis [151], which is needed for mesenchymal tissues. These and other data are consistent with HSP47 being a collagen-specific chaperone.

4.3. The Calnexin/Calreticulin network

The ER has one unique type of chaperone - lectins that bind to glycoproteins. Two such chaperones are known: the lumenal protein calreticulin (Crt) and the membrane-spanning chaperone calnexin (Cnx) [152]. Both prefer a glycoprotein whose glycan has a mono-glucosylated intermediate, as a result of trimming after the initial glycosylation [153155], although they also display distinct sensitivities to folding context of individual glycans [32, 156158]). There is also evidence that calnexin and calreticulin recognize features of proteins other than glycans. Dissecting intermediates in hemagglutinin, Hebert et al. observed that in contrast to Cnx, Crt bound primarily to early folding intermediates. Though the two chaperones share the same carbohydrate specificity, Crt binding depends on the oligosaccharides in the more rapidly folding top domain of HA, whereas calnexin is less discriminating [32]. Distinct epitopes also trigger calnexin and calreticulin binding in MHC class I heavy chains [159]. Another major difference between Crt and Cnx is the topology of their clients. Calreticulin selectively interacts with nascent luminal secretory proteins, such as transferring; however, when calreticulin was made artificially membrane-anchored, the spectrum of proteins it recognized became remarkably similar to that observed with calnexin [156]. Conversely, calnexin’s binding to its membrane-spanning client hemagglutinin was reduced when the latter was expressed as a soluble anchor-free protein [32].

4.3.1. Recruitment into the cycle

Calnexin and calreticulin participate in the dynamic sorting of immature proteins towards attainment of the native state vs. elimination by ERAD, in the so-called “calnexin/calreticulin cycle”. The sorting process consists of series of decision and commitment events. The first commitment event in the cycle is represented by the sequential action of oligosaccharyl-transferase complex (OST) and glucosidases I and II, which generates a mono-glucosylated protein intermediate [153, 156, 160163]. Because OST and glucosidase I are localized to the translocon [164, 165], the presence of a glycosylation site within N-terminal 50 residues ensures attachment of the glycan moiety early in the translocation process, thus directing the nascent protein into the Cnx/Crt cycle [166]. If the first glycosylation site is more C-terminal, the nascent protein may associate with BiP instead and proceed along a different chaperone-assisted folding path. The decision between the two chaperone complexes appears to be due to competition, since removing the glycosylation sites allows the protein that is normally a Cnx/Crt substrate to be recognized by BiP [166]. Cnx/Crt recruit into the complex the disulfide isomerase ERp57 [167170], which facilitates formation of the disulfide bonds in the immature glycoproteins.

4.3.2. Return to the cycle

Once the protein is released from the calnexin/calreticulin complex and further glucose-trimmed by glucosidase II to a high-mannose intermediate, it is subject to a second triage decision. If it is in a mature, secretion-competent conformation, it is targeted to the ER exit sites and proceeds to its destination. If, however, it is in a non-native conformation, it can be recognized by the folding sensor UDP-glucose:glycoprotein glucosyl transferase (UGT1) which re-glucosylates the high-mannose moiety, thus re-committing the incompletely folded polypeptide to the calnexin/calreticulin cycle for another round [153]. What proportion of glycoproteins is recommitted to repeated engagement with Cnx/Crt is unknown, and in three out of four cell types studied, UGT1 is quite sub-stoichiometric to both Cnx/Crt and glucosidase II (Table I).

In addition to UGT1, the high-mannose intermediate can be recognized by α-mannosidase I, resulting in the trimming of the mannose moiety [171173]. This step increases the chances of the polypeptide’s exit from the Cnx/Crt cycle, but does not commit it to exit. Further mannose trimming, for example by EDEM1 and EDEM3 [174, 175], produces a protein that can no longer be re-glucosylated by UGT1, thus essentially removing it from the further refolding attempts, and instead committing it to the degradation process [173, 176]. The relatively slow kinetics of a-mannosidase I enzymatic activity is thought to provide a timer [171] that allows for a certain residence time in the refolding part of the Cnx/Crt cycle, thus eliminating futile cycling of an intermediate that is unable to fold. Recognition of both - the trimmed mannose residues and the non-native polypeptide conformation by ERAD lectins such as OS9 and XTP3-B, is thought to be the final signal for degradation [177].

4.3.3. Client recognition

While we have a good idea of the sequence of events and identities of proteins in the calnexin cycle, there is much to be understood about both the recognition of the polypeptide by the different chaperones involved, and the consequences of their binding for the conformational maturation of the polypeptide. The first triage step – the attachment of the glycan to nascent protein emerging from the translocation channel – is unlikely to be driven by global conformational information. Local conformation, on the other hand, could affect the recognition of the glycosylation site by OST, since proline residues are not present at position +1 following the Asn, and are very rare in position +3 [178]. The binding by Cnx/Crt in vivo may be mediated by the specific glycan configuration, even though in vitro Cnx is able to bind unfolded, unglycosylated proteins [158]. For example, when glucose trimming is slowed down, the mature, folded forms of HA protein can be found bound to calnexin [163], suggesting that only the sugar residues were mediating the binding in this case. On the other hand, complete inhibition of glucose trimming has been shown to induce a prolonged association of a normally Cnx/Crt substrate (p62 viral glycoprotein) with BiP [166]. Given that BiP preferentially binds immature folding intermediates, selection into Cnx/Crt cycle is not likely to be based on the folding state of the polypeptide.

The conformational maturation of polypeptides released from Cnx/Crt is accessed by UGT1. The structural basis for this triage step is still not well understood. Different modes of recognition have been reported for UGT1 in vitro, from near-native conformation of chymotrypsin inhibitor-2 [179], to non-native conformations of several proteins with exposed hydrophobic residues on the surface [180], to small, local deviations from the native fold [181]. In plants, inactivating mutation of UGT1 results in less stringent ER retention of a biochemically active brassinosteroid receptor mutant harboring localized structural distortion [182]. It is possible that the degree of non-nativeness recognized by this folding sensor is dependent on, or influenced by, other, perhaps protein-specific, structural or conformational information. Clearly, a better characterization of the substrate recognition by UGT1 will be necessary for understanding the principles of triage of folding intermediates in the ER.

When the above triage mechanism is not sufficient, for example under ER stress conditions, an additional lectin-based mechanism exists. The ER membrane protein malectin specifically binds the Glc2Man9GlcNAc2 intermediate glycan, presumably formed during the re-glucosylation process, through a novel carbohydrate-binding site in its luminal domain [183]. Glycoprotein recognition by malectin does not affect their entry to the Cnx/Crt chaperone system and malectin binds after Cnx [184]. Malectin is induced by ER stress and associates preferentially with malfolded conformers of glycoproteins, which is consistent with a role as a backup to the lectin chaperone system that would be activated when misfolded forms accumulate in the ER [184].

4.3.4. Functional consequences of lectin chaperone binding

What is the contribution of Cnx/Crt binding to the folding of their substrates? One straightforward explanation, common to all chaperones interacting with early intermediates, is preventing aggregation of immature proteins [185]. The charged nature of glycans may in itself serve an anti-aggregation function. Second, Cnx/Crt binding mediates the recruitment of ERp57, which monitors the proper disulfide bond structure of the glycoprotein (reviewed in [186]).

Another possibility is that attachment of the glycan residue and binding of the lectin chaperones anchor the surrounding amino acids to the surface of the protein. In addition to reducing the conformational freedom of peptide backbone around the glycosylation site, such anchoring may allow conformations where amino acids that are normally buried in the core remain on the surface, without compromising the protein stability. Examination of 506 glycoprotein crystals showed enrichment for surface-exposed aromatic and hydrophobic residues in close spatial proximity to the N-glycosylation sites [38]. Petrescu et al. [187] propose that without attached glycans, these residues will seek the hydrophobic core of the molecule, thus favoring a different conformation. For example, among four used glycosylation sites of mouse tyrosinase, two are dispensable and two are necessary for efficient folding/function. The dispensable sites are GDEN203FT and RTAN337FS (see Table II), while the two whose occupancy is important for folding efficiency are VFYN86RT and IFMN371GT, with neighboring aromatic side chains. In fact, other studies suggest that glycans at certain positions can directly protect hydrophobic residues. In recombinant erythropoietin, the inner regions of highly branched glycans appear to stabilize the mature protein conformation by cliging to the hydrophobic protein surface areas [188], while in an unassembled alpha subunit of human chorionic gonadotropin, a glycan residue directly shields the hydrophobic core region that is normally protected by interaction with the beta subunit [189].

4.3.5. Organismal roles of the chaperones of the calnexin/calreticulin cycle

The fundamental role of monitoring the quality of glycoproteins would predict that the calnexin/calreticulin cycle is essential, but perhaps the overlapping specificity of calnexin and calreticulin would render each one of them dispensable This expectation is born out in tissue culture cells [190] and, surprisingly, also in knock-out animals. Two different Cnx−/− mouse models had normal embryonic development, but exhibited defective postnatal growth and neurological and behavioral deficits, accompanied by myelination-related defects [176, 191]. On the other hand, although calreticulin−/− mice [192] showed variably early embryonic lethality, with notable defects in heart development and function and a failure to absorb the umbilical hernia, they had no other gross morphological changes. At least in part, the heart function defects could be attributed to the calcium-buffering function of calreticulin rather than to its chaperone activity, since Crt−/− embryos had inhibited bradykinin-induced Ca2+ release by the InsP3-dependent pathway [192].

The restricted phenotypes and normal development of most tissues in these knock-out animals suggest, despite the distinctions in client recognition between Crt and Cnx, as discussed above, that for the majority of clients, the lectin chaperone requirement for their efficient folding is relatively relaxed. Alternatively, there may be sufficient ability in the non-lectin chaperone machinery to support the folding of many, but not all clients. The specific neurological phenotypes, myelination defects, and heart defects in knock-out animals suggest specific, non-redundant requirements for either calnexin or calreticulin function. At present, we don’t know whether this requirement reflects a heightened chaperone dependency of the secretory proteome in the susceptible cells (for example the presence of substrates that are highly dependent on calnexin or calreticulin, or the absence of compensating chaperones), or perhaps an increased sensitivity of these cells to protein misfolding (or decreased calcium buffering) in the ER.

4.4. Protein Disulfide and Prolyl Isomerases

In addition to chaperones, which do not directly alter the structure of itinerant proteins in the ER, the lumen is rich with enzymes that can change protein structure. Two of the important types of enzymes are protein disulfide isomerases (PDI) and prolyl isomerases (PPI), and each type is represented by multiple family members (Table I). As discussed above, the importance of dealing with the constraint of an oxidizing luminal environment provides obvious roles to several PDIs. However, what is the significance of such a large number of related proteins with thioredoxin domains? Do they have unique properties, or are many of them redundant? Similarly, among the PPIs, do all of them function as prolyl isomerases and do they have distinct substrates and functions?

4.4.1. The ER PDIs

At last count, there are 21 ER proteins with one or more thioredoxin-like domains, the defining feature of the PDI family [193]. Each thioredoxin domain is a platform for the CXXC active site that mediates the electron transfer involved in forming disulfide bonds (see chapter 6 and [42]). The PDIs are oxidoreductases, which can oxidize, reduce or isomerase disulfides, depending on the redox potential. Some PDIs, like ERdj5, are reductases in vivo and some, like ERp29, are catalytically inactive. In addition, PDIs are also involved in other activities: the highly abundant PDIA1 is the β subunit of prolyl 4-hydroxylase [194] and ERp44 binds to IP3 receptor [195]. The most abundant PDIs in a number of distinct cell types are the 60 kDa protein disulfide isomerase PDIA1 and two structurally-related proteins, P5 and ERp57 (Table I).

The relevance of these PDIs to the fate of secretory proteins has been established using pulse-chase experiments in the presence or absence of individual PDIs. For example α fetoprotein can still form disulfide bonds in the absence of PDI or ERp57, but its traffic to the Golgi is delayed, consistent with the presence of non-native folding intermediates that may require a disulfide isomerization reaction [196]. Yet, given the large number of ER PDIs, an important question is to what extent are they functionally redundant.

Genetic insight into the redundancy question is provided by the phenotypes of null alleles of the three PDIs in C. elegans: two of them are not essential, but the third is non-redundant and is required for prolyl 4-hydroxylase activity in collagen biosynthesis [197]. Williams et al. used an RNAi knockdown approach to address the same question in mammalian cells [196]. They showed that some PDIs clearly have defined tasks. PDIA1 depletion impacted oxidative folding of each one of several well-characterized secretory liver proteins. However, the phenotype was surprisingly modest, suggesting that other PDIs can compensate for PDIA1 depletion, albeit with lower efficacy. In contrast, depletion of ERp72 or P5, either alone or in combination with PDI or ERp57, had minimal impact on oxidative folding [196]. The RNAi approach also showed that one PDI, ERp57, also has broad specificity, but with a clear preference for glycoproteins. This specificity is explained because ERp57 must be physically associated with the calnexin cycle components to catayze isomerization reactions with most of its substrates [198]. Yet, while for some glycoproteins, like influenza virus hemagglutinin, the action of ERp57 is important for post-translational, later phases of oxidative folding, for many other glycoproteins association with ERp72 could replace ERp57 and maintain folding competence [199]. In perhaps similar fashion, P5 is a co-factor associated with BiP and interacts with (at least some) BiP substrates [200].

In a similar approach, Rajpal et al. [56] examined the effect of PDIA1 knockdown in β cells and showed that with reduced PDI activity, oxidation of proinsulin is unimpaired and, in fact, enhanced. This is accompanied by improved proinsulin exit from the ER and increased total insulin secretion, with no evidence of ER stress [56]. Thus, in this professional secretory cell, PDIA1 slows down folding.

If the calnexin cycle has evolved with a specialized oxidoreductase to facilitate native disulfide formation in complex glycoproteins, the process of ER-associated degradation involves one PDI with a unique activity: ERdj5 acts as a reductase, not an oxidase, on misfolded proteins and it is the only PDI with a J domain that enables it to bind to BiP and thus couple the peptide binding activity to the reductase activity in order to triage proteins for ERAD [116]. ERp72 has also been shown to interact with mutant thyroglobulin [201] and to participate with ERp29, ERp57 and PDIA1 in the unfolding of polyoma virus proteins during viral infection [202]. This unfolding is initiated by ERp29, a PDI-family member without oxidoreductase activity, which coordinates ERp57 and PDIA1 to unfold the C-terminus of the capsid protein [202]. A final example of a unique role for a PDI is pERpl, which is specialized for supporting antibody secretion by plasma cells [203, 204]. All these studies now enable the conclusion that although there is redundancy in supporting disulfide bond formation, a number of PDIs perform unique functions due to specific protein interactions mediated by their non-thioredoxin domains.

4.4.2. The ER immunophilins

The immunophilins/PPIs family is characterized by peptidyl-prolyl cis/trans isomerase activity (at least in vitro) and sensitivity to one or more of the immunosuppressive compounds cyclosporine A (CsA), FK506 and rapamycin. Of the three branches of the family, two are represented in the ER - the cyclosporine A (CsA)-binding cyclophilin B (CypB), and at least five FK506-binding proteins, FKBP13, FKBP19, FKBP22, FKBP23, and FKBP65. All of these immunophilins can catalyze isomerization of prolines, often a rate-limiting step in folding [205], but their mode of substrate recognition, and indeed the importance of their PPI activity in vivo is still poorly understood. Some of the ER immunophilins are induced when misfolded proteins accumulate, and even associate with misfolded substrates [206]. Yet, there is little detailed understanding for how the ER immunophilins participate in folding in the secretory pathway. Transferrin biosynthesis is sensitive to cyclosporine A [207], so since CypB is the main target of CsA in the ER [208], this immunophilin is implicated as a folding catalyst. CypB also accelerates the resolution of a kinetically trapped, on-pathway intermediate in immunoglobulin heavy chain folding, explaining the requirement for isomerization of Pro32 in cells [85].

Most of the data about the other immunophilins documents association and genetic necessity, but with relatively little mechanistic understanding. The large FKBP65 associates with the extracellular matrix proteins tropoelastin and collagen during its residence in the ER [209]. FKBP19 associates with CD81 in osteoblasts, which interacts with CD9 and a prostaglandin receptor regulator, and these associations cumulatively result in expression of interferon-inducible genes [210]. No specific substrate proteins for FKBP13 have been documented.

Human mutations in the FKBP14 gene are a cause of a type of Ehlers-Danlos syndrome (EDS) with progressive kyphoscoliosis, myopathy, and hearing loss [211]. Consistent with this phenotype, FKBP14-deficient fibroblasts exhibit altered assembly of the extracellular matrix in culture [211]. Both humans and mice deficient for CypB have Osteogenesis imperfecta and other bone abnormalities during development, as well as molecular defects in collagen biosynthesis [212, 213]. Nonetheless, the null phenotypes of these immunophilins do not exhibit growth, development, fertility, immunity or other defects aside from bone and connective tissues, as would have been inferred from cell biological and proteomic studies. One would have expected, for example, a defect in humoral immunity, given the interactors of FKBP19 and CypB [85, 210, 214]. Similarly, FKBP13 knockout mice do not have an obvious phenotype that would be consistent with effects on folding of clients (Argon, unpublished data).

C. elegans, an important model genetic organism, has three secretory pathway FKBPs, expressed in hypodermal cells [215]. Neither one is an essential protein, but as a functional group, all three are essential for normal nematode development, collagen biogenesis, and the formation of an intact exoskeleton under adverse physiological conditions [215]. Thus, the genetic evidence indicates extensive functional redundancy among the ER immunophilins in relation to protein biosynthesis.

While studies implicating immunophilins in folding of specific substrates are few, there are studies about the association of immunophilins with the ER folding machinery. One functional interaction is between FKBP23 and BiP. The Neurospora FKBP23 homologue binds physically to BiP, which enhances the chaperone activity of the FKBP [216]. Mouse FKBP23 also binds BiP [217]. This binding is Ca++-dependent, can suppress the ATPase activity of BiP through the PPIase activity of FKBP23 and is mediated by catalyzing the cis/trans isomerization of Pro 117 in the ATPase domain of BiP [218].

An intriguing recent finding has been that ER immunophilins interact with six of the protein disulfide isomerases [214], suggesting that PPIs and PDIs can modulate each other’s enzymatic activity in vivo, as demonstrated in some in vitro cases [219, 220]. CypB, in particular, interacts with three PDIs (PDIA1, PDIA4 and PDIA6) as well as with three major chaperone networks in the ER (BiP, Cnx/Crt and GRP94). The interaction of CypB with the lectin chaperones is mediated by their proline-rich P domain, through the same surface that binds ERp57, and is not sensitive to cyclosporine A [221]. Therefore, the protein interaction module is distinct from the enzymatic activity of CypB [221]. While the dynamics of these interactions in vivo are yet to be explored, a number of cells in fact contain a sufficient copy number of CypB to accommodate all these interactions (Table I).

Recently, a role for CypB has been reported in ERAD. Either pharmacological inhibition of CypB with cyclosporine A, or its depletion with siRNA inhibits the degradation of a subset of misfolded soluble luminal proteins, those that contain cis proline residues [208]. CypB is apparently the only ER-resident target of cyclosporine A and its catalytic activity likely enhances disposal from the ER by resolving local structures that are retro-transclocated inefficiently. Functional roles in vivo for the other PPIs are yet to be elucidated.

4.5. Coordination of chaperone actions

Though we present each of the major chaperone systems separately, their various modes of recognition and activities are all needed, in parallel or in sequence, to fold proteins. In addition to the aforementioned sequential interaction of BiP and GRP94 with some clients, there are also other examples of coordinated chaperone action. Many of the folding factors that process non-glycoproteins interact sufficiently frequently to be isolated as a multi-chaperone complex from antibody-producing cells [222]. Intermediates of apolipoprotein B-100 were associated with GRP94, ERp72, BiP, calreticulin, and cyclophilin B [223]. Remarkably, as judged by subcellular fractionation, this array of chaperones remain associated with apolipoprotein B-100 during subsequent processing in the Golgi complex [223]. Presumably, the association of chaperones and lectins and/or folding enzymes creates a task-specific functional complex that works ad hoc to fold a particular client.

When a large glycoprotein like thyroglobulin is bound by BiP, GRP94, calreticulin, ERp29 and others [67, 224], the binding is not necessarily simultaneous; different domains of thyroglobulin likely require different chaperones [225], and the cycles of binding and release may be temporally programmed [226]. Likewise, when a disulfide-bonded glycoprotein is targeted to ERAD, it is released from the lectin chaperones, and its disulfides are reduced by ERdj5, which uses its J domain to couple the activity to BiP[116] in preparation for escort by OS-9 to the SEL1L/Hrd1 retrotranslocon [116]. These examples illustrate the dynamic nature of the chaperone networks, which are transient, come together based on individual protein’s ‘needs’, as well as on the stage of folding that is being chaperoned. The dynamic association of chaperones, enzymes and their substrates and clients is also evident by the use of live cell imaging of their fusions with fluoresent protein [227229].

5. Conclusions

The folding machinery in the lumen of the ER has evolved to accommodate the unique features of the ER environment as well as common signatures of secretory proteins, such as disulfide bonds and glycans. Thus, while some of the ER chaperones and enzymes work in similar fashion to homologous cytosolic chaperones (e.g. BiP and HSP70, FKBP13 and FKBP12), other components (lectin chaperones and adaptors, PDIs) are unique to this folding compartment. Even GRP94, which is structurally highly similar to the two cytosolic HSP90s, has evolved a distinct mode of chaperone action. To recognize a particular substrate, the various ER chaperone systems use distinct structural features, and for the most part engage in cycles of binding and release, during which distinct folding reactions are executed. The dynamics is even more complicated, since different chaperones may work on different domains of a given protein molecule, and may collaborate with an isomerase or a glycosylation enzyme, as required ad hoc by the substrate. The identity of the substrate dictates which chaperone acts as a hub, that in turn forms transient complexes with one or more other folding factors, in characteristic order.

A common biophysical role that the ER folding machinery fulfills is restriction of the folding options early on, at a time that many pathways can be followed. Perhaps the unifying mechanism for this action is “pinning” a patch of residues to either the core or the surface as a means of favoring a particular folding pathway (e.g. [230]). However, the BiP and the lectin chaperone networks achieve this goal by recognizing distinctly different structures (peptides vs. glycans). The BiP network marks peptides that are destined to be shielded from solvent, and thereby allowing other segments of the polypeptide to fold around these peptides. The attachment of glycans and the subsequent association with the lectin chaperone network overcomes the thermodynamic tendency of nearby patches enriched in hydrophobic and aromatic residues to avoid display on the protein surface, thus initiating particular folding pathways. The action of PDIs and PPIs obviously restricts folding options by introducing irreversible steps. The GRP94 network may also restrict conformational space, but in a way that is currently not understood because the structural cues of GRP94 engagement with folding intermediates are yet to be deciphered

Despite the impressive progress of our knowledge, many questions are still not resolved, including some that are common to several chaperone systems. How is the ‘productive’ folding intermediate distinguished from the ‘unproductive’ one? Are the rules intrinsic to the intermediate states, or are they different for different chaperones? When a substrate binds cyclically to a chaperone, how is the number of repetitive interactions determined? If a substrate presents several chaperone binding sites, when does its folding depend on just one dominant site (e.g. one glycan), and when are multiple sites required? When does the abundance of limiting co-chaperones, like GRP170, or sensors, like UGT1, become an impediment for efficient folding?

Future studies should also explain how unassembled subunits stay occupied with the ER chaperone networks, rather than sent for degradation. Our knowledge of substrate recognition by components like OS-9, which participate in such triage decisions, is still incomplete and it will be important to understand when and how such components engage with the folding machinery, and when they are committed to the substrate disposal via ERAD or autophagy.

ER review highlights.

  • -

    Unique conditions for protein folding in the ER

  • -

    Unique properties of secretory and membrane proteins

  • -

    The major chaperone systems in the lumen

  • -

    Specialization of ER chaperones to the folding needs in the lumen

  • -

    Folding enzymes interact with the major ER chaperones

Acknowledgements

We thank Drs. J. Brodsky (Univ. Pitt.), E. Snapp (Albert Einstein College of Medicine), M. Molinari (Institute for Research in Biomedicine, Bellinzona, Switserland), P. Cresswell (Yale Univ.) and D. Hebert (Univ. Mass.) for helpful comments and personal communications. Work in the authors’ lab was supported by NIH grants AG 18001 and GM07748.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Chen Y, Zhang Y, Yin Y, Gao G, Li S, Jiang Y, Gu X, Luo J. SPD--a web-based secreted protein database. Nucleic Acids Res. 2005;33:D169–D173. doi: 10.1093/nar/gki093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Anfinsen CB. Principles that govern the folding of protein chains. Science. 1973;181:223–230. doi: 10.1126/science.181.4096.223. [DOI] [PubMed] [Google Scholar]
  • 3.Anfinsen CB, Haber E, Sela M, White FHJ. The kinetics of formation of native ribonuclease during oxidation of the reduced polypeptide chain. Proc. Natl. Acad. Sci. USA. 1961;47:1309–1314. doi: 10.1073/pnas.47.9.1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Levinthal C. Are there pathways for protein folding? J Chim Physique. 1968;65:44–45. [Google Scholar]
  • 5.Karplus M. The Levinthal paradox: yesterday and today. Fold Des. 1997;2:S69–S75. doi: 10.1016/s1359-0278(97)00067-9. [DOI] [PubMed] [Google Scholar]
  • 6.Brockwell DJ, Radford SE. Intermediates: ubiquitous species on folding energy landscapes? Curr Opin Struct Biol. 2007;17:30–37. doi: 10.1016/j.sbi.2007.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Fersht AR, Itzhaki LS, El-Masry NF, Matthews JM, Otzen DE. Single versus parallel pathways of protein folding and fractional formation of structure in the transition state. Proc Natl Acad Sci U S A. 1994;91:10426–10429. doi: 10.1073/pnas.91.22.10426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Krishna MM, Englander SW. A unified mechanism for protein folding: predetermined pathways with optional errors. Protein Sci. 2007;16:449–464. doi: 10.1110/ps.062655907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Onuchic JN, Wolynes PG. Theory of protein folding. Curr Opin Struct Biol. 2004;14:70–75. doi: 10.1016/j.sbi.2004.01.009. [DOI] [PubMed] [Google Scholar]
  • 10.Fink AL. Chaperone-mediated protein folding. Physiol Rev. 1999;79:425–449. doi: 10.1152/physrev.1999.79.2.425. [DOI] [PubMed] [Google Scholar]
  • 11.Melnick J, Dul JL, Argon Y. Sequential interaction of the chaperones BiP and GRP94 with immunoglobulin chains in the endoplasmic reticulum. Nature. 1994;370:373–375. doi: 10.1038/370373a0. [DOI] [PubMed] [Google Scholar]
  • 12.Davis DP, Khurana R, Meredith S, Stevens FJ, Argon Y. Mapping the major interaction between BiP and immunoglobulin light chains to sites within the variable domain. J. Immunol. 1999;163:3842–3850. [PubMed] [Google Scholar]
  • 13.Davis DP, Raffen R, Dul JL, Vogen SM, Williamson EK, Stevens FJ, Argon Y. Inhibition of amyloid fiber assembly by both BiP and its target peptide. Immunity. 2000;13:433–442. doi: 10.1016/s1074-7613(00)00043-1. [DOI] [PubMed] [Google Scholar]
  • 14.Zelensky AN, Gready JE. The C-type lectin-like domain superfamily. FEBS J. 2005;272:6179–6217. doi: 10.1111/j.1742-4658.2005.05031.x. [DOI] [PubMed] [Google Scholar]
  • 15.Hurtley SM, Bole DG, Hoover-Litty H, Helenius A, Copeland CS. Interactions of misfolded influenza virus hemagglutinin with binding protein (BiP) J. Cell Biol. 1989;108:2117–2126. doi: 10.1083/jcb.108.6.2117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Segal MS, Bye JM, Sambrook JF, Gething MJ. Disulfide bond formation during the folding of influenza virus hemagglutinin. J. Cell Biol. 1992;118:227–244. doi: 10.1083/jcb.118.2.227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Chen W, Helenius A. Role of ribosome and translocon complex during folding of influenza hemagglutinin in the endoplasmic reticulum of living cells. Mol. Biol. Cell. 2000;11:765–772. doi: 10.1091/mbc.11.2.765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Ruddon RW, Sherman SA, Bedows E. Protein folding in the endoplasmic reticulum: lessons from the human chorionic gonadotropin beta subunit. Protein Sci. 1996;5:1443–1452. doi: 10.1002/pro.5560050801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Dayel MJ, Horn EF, Verkman AS. Diffusion of green fluorescent protein in the aqueous-phase lumen of endoplasmic reticulum. Biophys J. 1999;76:2843–2851. doi: 10.1016/S0006-3495(99)77438-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Hwang C, Sinskey AJ, Lodish HF. Oxidized redox state of glutathione in the endoplasmic reticulum. Science. 1992;257:1496–1502. doi: 10.1126/science.1523409. [DOI] [PubMed] [Google Scholar]
  • 21.Montero M, Brini M, Marsault R, Alvarez J, Sitia R, Pozzan T, Rizzuto R. Monitoring dynamic changes in free Ca2+ concentration in the endoplasmic reticulum of intact cells. EMBO J. 1995;14:5467–5475. doi: 10.1002/j.1460-2075.1995.tb00233.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Zhu M, Souillac PO, Ionescu-Zanetti C, Carter SA, Fink AL. Surface-catalyzed amyloid fibril formation. J Biol Chem. 2002;277:50914–50922. doi: 10.1074/jbc.M207225200. [DOI] [PubMed] [Google Scholar]
  • 23.Gorbenko GP, Ioffe VM, Kinnunen PK. Binding of lysozyme to phospholipid bilayers: evidence for protein aggregation upon membrane association. Biophys J. 2007;93:140–153. doi: 10.1529/biophysj.106.102749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Zhao H, Jutila A, Nurminen T, Wickstrom SA, Keski-Oja J, Kinnunen PK. Binding of endostatin to phosphatidylserine-containing membranes and formation of amyloid-like fibers. Biochemistry. 2005;44:2857–2863. doi: 10.1021/bi048510j. [DOI] [PubMed] [Google Scholar]
  • 25.Zhao H, Tuominen EK, Kinnunen PK. Formation of amyloid fibers triggered by phosphatidylserine-containing membranes. Biochemistry. 2004;43:10302–10307. doi: 10.1021/bi049002c. [DOI] [PubMed] [Google Scholar]
  • 26.Hamman BD, Hendershot LM, Johnson AE. BiP maintains the permeability barrier of the ER membrane by sealing the lumenal end of the translocon pore before and early in translocation. Cell. 1998;92:747–758. doi: 10.1016/s0092-8674(00)81403-8. [DOI] [PubMed] [Google Scholar]
  • 27.Silberstein S, Gilmore R. Biochemistry, molecular biology, and genetics of the oligosaccharyltransferase. FASEB J. 1996;10:849–858. [PubMed] [Google Scholar]
  • 28.van den Berg B, Ellis RJ, Dobson CM. Effects of macromolecular crowding on protein folding and aggregation. EMBO J. 1999;18:6927–6933. doi: 10.1093/emboj/18.24.6927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Totani K, Ihara Y, Matsuo I, Ito Y. Effects of macromolecular crowding on glycoprotein processing enzymes. J Am Chem Soc. 2008;130:2101–2107. doi: 10.1021/ja077570k. [DOI] [PubMed] [Google Scholar]
  • 30.Parham P. Functions for MHC class I carbohydrates inside and outside the cell. Trends Biochem. Sci. 1996;21:427–433. doi: 10.1016/s0968-0004(96)10053-0. [DOI] [PubMed] [Google Scholar]
  • 31.Machamer CE, Rose JK. Vesicular stomatitis virus G proteins with altered glycosylation sites display temperature-sensitive intracellular transport and are subject to aberrant intermolecular disulfide bonding. J. Biol. Chem. 1988;263:5955–5960. [PubMed] [Google Scholar]
  • 32.Hebert DN, Zhang JX, Chen W, Foellmer B, Helenius A. The number and location of glycans on influenza hemagglutinin determine folding and association with calnexin and calreticulin. J Cell Biol. 1997;139:613–623. doi: 10.1083/jcb.139.3.613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Machamer CE, Florkiewicz RZ, Rose JK. A single N-linked oligosaccharide at either of the two normal sites is sufficient for transport of vesicular stomatitis virus G protein to the cell surface. Mol. Cell. Biol. 1985;5:3074–3083. doi: 10.1128/mcb.5.11.3074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Gallagher PJ, Henneberry JM, Sambrook JF, Gething MJ. Glycosylation requirements for intracellular transport and function of the hemagglutinin of influenza virus. J. Virol. 1992;66:7136–7145. doi: 10.1128/jvi.66.12.7136-7145.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Martayan A, Sibilio L, Setini A, Lo Monaco E, Tremante E, Fruci D, Colonna M, Giacomini P. N-linked glycosylation selectively regulates the generic folding of HLA-Cw1. J Biol Chem. 2008;283:16469–16476. doi: 10.1074/jbc.M709175200. [DOI] [PubMed] [Google Scholar]
  • 36.Zhang Q, Salter RD. Distinct patterns of folding and interactions with calnexin and calreticulin in human class I MHC proteins with altered N-glycosylation. J Immunol. 1998;160:831–837. [PubMed] [Google Scholar]
  • 37.Jost CR, Kurucz I, Jacobus CM, Titus JA, George AJ, Segal DM. Mammalian expression and secretion of functional single-chain Fv molecules. J. Biol. Chem. 1994;269:26267–26273. [PubMed] [Google Scholar]
  • 38.Petrescu AJ, Milac AL, Petrescu SM, Dwek RA, Wormald MR. Statistical analysis of the protein environment of N-glycosylation sites: implications for occupancy, structure, and folding. Glycobiology. 2004;14:103–114. doi: 10.1093/glycob/cwh008. [DOI] [PubMed] [Google Scholar]
  • 39.Culyba EK, Price JL, Hanson SR, Dhar A, Wong CH, Gruebele M, Powers ET, Kelly JW. Protein native-state stabilization by placing aromatic side chains in N-glycosylated reverse turns. Science. 331:571–575. doi: 10.1126/science.1198461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Bardwell JC, McGovern K, Beckwith J. Identification of a protein required for disulfide bond formation in vivo. Cell. 1991;67:581–589. doi: 10.1016/0092-8674(91)90532-4. [DOI] [PubMed] [Google Scholar]
  • 41.Tu BP, Weissman JS. Oxidative protein folding in eukaryotes: mechanisms and consequences. J Cell Biol. 2004;164:341–346. doi: 10.1083/jcb.200311055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Sevier CS, Kaiser CA. Ero1 and redox homeostasis in the endoplasmic reticulum. Biochim Biophys Acta. 2008;1783:549–556. doi: 10.1016/j.bbamcr.2007.12.011. [DOI] [PubMed] [Google Scholar]
  • 43.Tavender TJ, Sheppard AM, Bulleid NJ. Peroxiredoxin IV is an endoplasmic reticulum-localized enzyme forming oligomeric complexes in human cells. Biochem J. 2008;411:191–199. doi: 10.1042/BJ20071428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Tavender TJ, Springate JJ, Bulleid NJ. Recycling of peroxiredoxin IV provides a novel pathway for disulphide formation in the endoplasmic reticulum. EMBO J. 2010;29:4185–4197. doi: 10.1038/emboj.2010.273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Tavender TJ, Bulleid NJ. Peroxiredoxin IV protects cells from oxidative stress by removing H2O2 produced during disulphide formation. J Cell Sci. 2010;123:2672–2679. doi: 10.1242/jcs.067843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Zito E, Melo EP, Yang Y, Wahlander A, Neubert TA, Ron D. Oxidative protein folding by an endoplasmic reticulum-localized peroxiredoxin. Mol Cell. 2010;40:787–797. doi: 10.1016/j.molcel.2010.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.van der Vlies D, Makkinje M, Jansens A, Braakman I, Verkleij AJ, Wirtz KW, Post JA. Oxidation of ER resident proteins upon oxidative stress: effects of altering cellular redox/antioxidant status and implications for protein maturation. Antioxid Redox Signal. 2003;5:381–387. doi: 10.1089/152308603768295113. [DOI] [PubMed] [Google Scholar]
  • 48.Creighton TE. Protein folding coupled to disulphide bond formation. Biol. Chem. 1997;378:731–744. doi: 10.1515/bchm.1997.378.8.731. [DOI] [PubMed] [Google Scholar]
  • 49.Bergman LW, Kuehl WM. Formation of intermolecular disulfide bonds on nascent immunoglobulin polypeptides. J. Biol. Chem. 1979;254:5690–5694. [PubMed] [Google Scholar]
  • 50.Marquardt T, Helenius A. Misfolding and aggregation of newly synthesized proteins in the endoplasmic reticulum. J. Cell Biol. 1992;117:505–513. doi: 10.1083/jcb.117.3.505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Elkabetz Y, Argon Y, Bar-Nun S. Cysteines in CH1 underlie retention of unassembled Ig heavy chains. J Biol Chem. 2005;280:14402–14412. doi: 10.1074/jbc.M500161200. [DOI] [PubMed] [Google Scholar]
  • 52.Braakman I, Helenius J, Helenius A. Manipulating disulfide bond formation and protein folding in the endoplasmic reticulum. EMBO J. 1992;11:1717–1722. doi: 10.1002/j.1460-2075.1992.tb05223.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Mezghrani A, Fassio A, Benham A, Simmen T, Braakman I, Sitia R. Manipulation of oxidative protein folding and PDI redox state in mammalian cells. EMBO J. 2001;20:6288–6296. doi: 10.1093/emboj/20.22.6288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Taniyama Y, Ogasahara K, Yutani K, Kikuchi M. Folding mechanism of mutant human lysozyme C77/95A with increased secretion efficiency in yeast. J. Biol. Chem. 1992;267:4619–4624. [PubMed] [Google Scholar]
  • 55.Worn A, Pluckthun A. An intrinsically stable antibody scFv fragment can tolerate the loss of both disulfide bonds and fold correctly. FEBS Lett. 1998;427:357–361. doi: 10.1016/s0014-5793(98)00463-3. [DOI] [PubMed] [Google Scholar]
  • 56.Rajpal G, Schuiki I, Liu M, Volchuk A, Arvan P. Action of protein disulfide isomerase on proinsulin exit from endoplasmic reticulum of pancreatic beta-cells. J Biol Chem. 2012;287:43–47. doi: 10.1074/jbc.C111.279927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Macer DPJ, Koch GLE. Identification of a set of calcium-binding protein in reticuloplasm, the luminal content of the endoplasmic reticulum. J. Cell Sci. 1988;91:61–70. doi: 10.1242/jcs.91.1.61. [DOI] [PubMed] [Google Scholar]
  • 58.Li LJ, Li X, Ferrario A, Rucker N, Liu ES, Wong S, Gomer CJ, Lee AS. Establishment of a Chinese hamster ovary cell line that expresses grp78 antisense transcripts and suppresses A23187 induction of both GRP78 and GRP94. J. Cell. Physiol. 1992;153:575–582. doi: 10.1002/jcp.1041530319. [DOI] [PubMed] [Google Scholar]
  • 59.Corbett EF, Oikawa K, Francois P, Tessier DC, Kay C, Bergeron JJ, Thomas DY, Krause KH, Michalak M. Ca2+ Regulation of Interactions between Endoplasmic Reticulum Chaperones. J. Biol. Chem. 1999;274:6203–6211. doi: 10.1074/jbc.274.10.6203. [DOI] [PubMed] [Google Scholar]
  • 60.Suzuki CK, Bonifacino JS, Lin AY, Davis MM, Klausner RD. Regulating the retention of T-cell receptor zeta chain variants within the endoplasmic reticulum: Ca2+-dependent association with BiP. J. Cell Biol. 1991;114:189–205. doi: 10.1083/jcb.114.2.189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Rupp K, Birnbach U, Lundstrom J, Van PN, Soling HD. Effects of CaBP2, the rat analog of ERp72, and of CaBPl on the refolding of denatured reduced proteins. Comparison with protein disulfide isomerase. J Biol Chem. 1994;269:2501–2507. [PubMed] [Google Scholar]
  • 62.Biswas C, Ostrovsky O, Makarewich CA, Wanderling S, Gidalevitz T, Argon Y. The peptide binding activity of GRP94 is regulated by Calciun. Biochem. J. 2007;405:233–241. doi: 10.1042/BJ20061867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Lodish HF, Kong N, Wikstrom L. Calcium is required for folding of newly made subunits of the asialoglycoprotein receptor within the endoplasmic reticulum. J. Biol. Chem. 1992;267:12753–12760. [PubMed] [Google Scholar]
  • 64.Lodish HF, Kong N. Perturbation of cellular calcium blocks exit of secretory proteins from the rough endoplasmic reticulum. J. Biol. Chem. 1990;265:10893–10899. [PubMed] [Google Scholar]
  • 65.Biemesderfer D, Dekan G, Aronson PS, Farquhar MG. Biosynthesis of the gp330/44-kDa Heymann nephritis antigenic complex: assembly takes place in the ER. Am J Physiol. 1993;264:F1011–F1020. doi: 10.1152/ajprenal.1993.264.6.F1011. [DOI] [PubMed] [Google Scholar]
  • 66.Kuznetsov G, Chen LB, Nigam SK. Multiple molecular chaperones complex with misfolded large oligomeric glycoproteins in the endoplasmic reticulum. J. Biol. Chem. 1997;272:3057–3063. doi: 10.1074/jbc.272.5.3057. [DOI] [PubMed] [Google Scholar]
  • 67.Di Jeso B, Ulianich L, Pacifico F, Leonardi A, Vito P, Consiglio E, Formisano S, Arvan P. Folding of thyroglobulin in the calnexin/calreticulin pathway and its alteration by loss of Ca2+ from the endoplasmic reticulum. Biochem J. 2003;370:449–458. doi: 10.1042/BJ20021257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Geierhaas CD, Paci E, Vendruscolo M, Clarke J. Comparison of the transition states for folding of two Ig-like proteins from different superfamilies. J Mol Biol. 2004;343:1111–1123. doi: 10.1016/j.jmb.2004.08.100. [DOI] [PubMed] [Google Scholar]
  • 69.North CL, Blacklow SC. Solution structure of the sixth LDL-A module of the LDL receptor. Biochemistry. 2000;39:2564–2571. doi: 10.1021/bi992087a. [DOI] [PubMed] [Google Scholar]
  • 70.Fass D, Blacklow S, Kim PS, Berger JM. Molecular basis of familial hypercholesterolaemia from structure of LDL receptor module. Nature. 1997;388:691–693. doi: 10.1038/41798. [DOI] [PubMed] [Google Scholar]
  • 71.Campbell ID, Bork P. Epidermal growth factor-like modules. Current Opinion in Structural Biology. 1993;3:385–392. [Google Scholar]
  • 72.Branza-Nichita N, Petrescu AJ, Dwek RA, Wormald MR, Platt FM, Petrescu SM. Tyrosinase folding and copper loading in vivo: a crucial role for calnexin and alpha-glucosidase II. Biochem Biophys Res Commun. 1999;261:720–725. doi: 10.1006/bbrc.1999.1030. [DOI] [PubMed] [Google Scholar]
  • 73.Weissman JS, Kim PS. A kinetic explanation for the rearrangement pathway of BPTI folding. Nat. Struct. Biol. 1995;2:1123–1130. doi: 10.1038/nsb1295-1123. [DOI] [PubMed] [Google Scholar]
  • 74.Ebbinghaus S, Dhar A, McDonald JD, Gruebele M. Protein folding stability and dynamics imaged in a living cell. Nat Methods. 7:319–323. doi: 10.1038/nmeth.1435. [DOI] [PubMed] [Google Scholar]
  • 75.Skowronek MH, Hendershot LM, Haas IG. The variable domain of nonassembled Ig light chains determines both their half-life and binding to the chaperone BiP. Proc. Natl. Acad. Sci. USA. 1998;95:1574–1578. doi: 10.1073/pnas.95.4.1574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Vanhove M, Usherwood YK, Hendershot LM. Unassembled Ig heavy chains do not cycle from BiP in vivo but require light chains to trigger their release. Immunity. 2001;15:105–114. doi: 10.1016/s1074-7613(01)00163-7. [DOI] [PubMed] [Google Scholar]
  • 77.Hua QX, Liu M, Hu SQ, Jia W, Arvan P, Weiss MA. A conserved histidine in insulin is required for the foldability of human proinsulin: structure and function of an ALAB5 analog. J Biol Chem. 2006;281:24889–24899. doi: 10.1074/jbc.M602617200. [DOI] [PubMed] [Google Scholar]
  • 78.Liu M, Hodish I, Haataja L, Lara-Lemus R, Rajpal G, Wright J, Arvan P. Proinsulin misfolding and diabetes: mutant INS gene-induced diabetes of youth. Trends Endocrinol Metab. 2010;21:652–659. doi: 10.1016/j.tem.2010.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Copeland CS, Doms RW, Bolzau EM, Webster RG, Helenius A. Assembly of influenza hemagglutinin trimers and its role in intracellular transport. J. Cell Biol. 1986;103:1179–1191. doi: 10.1083/jcb.103.4.1179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Doms RW, Keller DS, Helenius A, Balch WE. Role for adenosine triphosphate in regulating the assembly and transport of vesicular stomatitis virus G protein trimers. J. Cell Biol. 1987;105:1957–1969. doi: 10.1083/jcb.105.5.1957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Lee YK, Brewer JW, Hellman R, Hendershot LM. BiP and immunoglobulin light chain cooperate to control the folding of heavy chain and ensure the fidelity of immunoglobulin assembly. Mol. Biol. Cell. 1999;10:2209–2219. doi: 10.1091/mbc.10.7.2209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Elkabetz Y, Ofir A, Argon Y, Bar-Nun S. Alternative pathways of disulfide bond formation yield secretion-competent, stable and functional immunoglobulins. Mol Immunol. 2008 doi: 10.1016/j.molimm.2008.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Huppa JB, Ploegh HL. In vitro translation and assembly of a complete T cell receptor-CD3 complex. J. Exp. Med. 1997;186:393–403. doi: 10.1084/jem.186.3.393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Wanamaker CP, Green WN. Endoplasmic reticulum chaperones stabilize nicotinic receptor subunits and regulate receptor assembly. J Biol Chem. 2007;282:31113–31123. doi: 10.1074/jbc.M705369200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Feige MJ, Groscurth S, Marcinowski M, Shimizu Y, Kessler H, Hendershot LM, Buchner J. An unfolded CHI domain controls the assembly and secretion of IgG antibodies. Mol Cell. 2009;34:569–579. doi: 10.1016/j.molcel.2009.04.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hellman R, Vanhove M, Lejeune A, Stevens FJ, Hendershot LM. The in vivo association of BiP with newly synthesized proteins is dependent on the rate and stability of folding and not simply on the presence of sequences that can bind to BiP. J. Cell Biol. 1999;144:21–30. doi: 10.1083/jcb.144.1.21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Bonifacino JS, Cosson P, Klausner RD. Colocalized transmembrane determinants for ER degradation and subunit assembly explain the intracellular fate of TCR chains. Cell. 1990;63:503–513. doi: 10.1016/0092-8674(90)90447-m. [DOI] [PubMed] [Google Scholar]
  • 88.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;10:2783–2793. doi: 10.1002/j.1460-2075.1991.tb07827.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Bonifacino JS, Suzuki CK, Klausner RD. A peptide sequence confers retention and rapid degradation in the endoplasmic reticulum. Science. 1990;247:79–82. doi: 10.1126/science.2294595. [DOI] [PubMed] [Google Scholar]
  • 90.Rutledge T, Cosson P, Manolios N, Bonifacino JS, Klausner RD. Transmembrane helical interactions: zeta chain dimerization and functional association with the T cell antigen receptor. EMBO J. 1992;11:3245–3254. doi: 10.1002/j.1460-2075.1992.tb05402.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.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;123:1031–1038. doi: 10.1242/jcs.059758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kalandadze A, Wu Y, Fournier K, Robinson MB. Identification of motifs involved in endoplasmic reticulum retention-forward trafficking of the GLT-1 subtype of glutamate transporter. J Neurosci. 2004;24:5183–5192. doi: 10.1523/JNEUROSCI.0839-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Coleman SK, Moykkynen T, Jouppila A, Koskelainen S, Rivera C, Korpi ER, Keinanen K. Agonist occupancy is essential for forward trafficking of AMPA receptors. J Neurosci. 2009;29:303–312. doi: 10.1523/JNEUROSCI.3953-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Coleman SK, Moykkynen T, Cai C, von Ossowski L, Kuismanen E, Korpi ER, Keinanen K. Isoform-specific early trafficking of AMPA receptor flip and flop variants. J Neurosci. 2006;26:11220–11229. doi: 10.1523/JNEUROSCI.2301-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Gething MJ. Guidebook to Molecular Chaperones and Protein-Folding Catalysts. 1998 [Google Scholar]
  • 96.Luo S, Mao C, Lee B, Lee AS. GRP78/BiP is required for cell proliferation and protecting the inner cell mass from apoptosis during early mouse embryonic development. Mol Cell Biol. 2006;26:5688–5697. doi: 10.1128/MCB.00779-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Flynn GC, Pohl J, Flocco MT, Rothman JE. Peptide-binding specificity of the molecular chaperone BiP. Nature. 1991;353:726–730. doi: 10.1038/353726a0. [DOI] [PubMed] [Google Scholar]
  • 98.Blond-Elguindi S, Cwirla SE, Dower WJ, Lipshutz RJ, Sprang SR, Sambrook JF, Gething MJ. Affinity panning of a library of peptides displayed on bacteriophages reveals the binding specificity of BiP. Cell. 1993;75:717–728. doi: 10.1016/0092-8674(93)90492-9. [DOI] [PubMed] [Google Scholar]
  • 99.Machamer CE, Doms RW, Bole DG, Helenius A, Rose JK. Heavy chain binding protein recognizes incompletely disulfide-bonded forms of vesiciular stomatitis virus G protein. J. Biol. Chem. 1990;265:6879–6883. [PubMed] [Google Scholar]
  • 100.Dul JL, Davis DP, Williamson EK, Stevens FJ, Argon Y. Hsp70 and antifibrillogenic peptides promote degradation and inhibit intracellular aggregation of amyloidogenic light chains. J. Cell Biol. 2001;152:705–716. doi: 10.1083/jcb.152.4.705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Hendershot L, Wei J, Gaut J, Melnick J, Aviel S, Argon Y. Inhibition of immunoglobulin folding and secretion by dominant negative BiP ATPase mutant. Proc. Natl. Acad. Sci. USA. 1996;93:5269–5274. doi: 10.1073/pnas.93.11.5269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Clairmont CA, De Maio A, Hirschberg CB. Translocation of ATP into the lumen of rough endoplasmic reticulum-derived vesicles and its binding to luminal proteins including BiP (GRP 78) and GRP 94. J. Biol. Chem. 1992;267:3983–3990. [PubMed] [Google Scholar]
  • 103.Dierks T, Volkmer J, Schlenstedt G, Jung C, Sandholzer U, Zachmann K, Schlotterhose P, Neifer K, Schmidt B, Zimmermann R. A microsomal ATP-binding protein involved in efficient protein transport into the mammalian endoplasmic reticulum. EMBO J. 1996;15:6931–6942. [PMC free article] [PubMed] [Google Scholar]
  • 104.Blond-Elguindi S, Fourie AM, Sambrook JF, Gething MJ. Peptide-dependent stimulation of the ATPase activity of the molecular chaperone BiP is the result of conversion of oligomers to active monomers. J. Biol. Chem. 1993;268:12730–12735. [PubMed] [Google Scholar]
  • 105.Kampinga HH, Craig EA. The HSP70 chaperone machinery: J proteins as drivers of functional specificity. Nat Rev Mol Cell Biol. 2010;11:579–592. doi: 10.1038/nrm2941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Andreasson C, Rampelt H, Fiaux J, Druffel-Augustin S, Bukau B. The endoplasmic reticulum Grp170 acts as a nucleotide exchange factor of Hsp70 via a mechanism similar to that of the cytosolic Hsp110. J Biol Chem. 2010;285:12445–12453. doi: 10.1074/jbc.M109.096735. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Weitzmann A, Baldes C, Dudek J, Zimmermann R. The heat shock protein 70 molecular chaperone network in the pancreatic endoplasmic reticulum - a quantitative approach. FEBS J. 2007;274:5175–5187. doi: 10.1111/j.1742-4658.2007.06039.x. [DOI] [PubMed] [Google Scholar]
  • 108.Schwanhausser B, Busse D, Li N, Dittmar G, Schuchhardt J, Wolf J, Chen W, Selbach M. Global quantification of mammalian gene expression control. Nature. 2011;473:337–342. doi: 10.1038/nature10098. [DOI] [PubMed] [Google Scholar]
  • 109.Spee P, Subjeck J, Neefjes J. Identification of novel peptide binding proteins in the endoplasmic reticulum: ERp72, calnexin, and grpl70. Biochemistry. 1999;38:10559–10566. doi: 10.1021/bi990321r. [DOI] [PubMed] [Google Scholar]
  • 110.Park J, Easton DP, Chen X, MacDonald IJ, Wang XY, Subjeck JR. The chaperoning properties of mouse grp170, a member of the third family of hsp70 related proteins. Biochemistry. 2003;42:14893–14902. doi: 10.1021/bi030122e. [DOI] [PubMed] [Google Scholar]
  • 111.Lin HY, Masso-Welch P, Di YP, Cai JW, Shen JW, Subjeck JR. The 170-kDa glucose-regulated stress protein is an endoplasmic reticulum protein that binds immunoglobulin. Mol Biol Cell. 1993;4:1109–1119. doi: 10.1091/mbc.4.11.1109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Schmidt BZ, Perlmutter DH. Grp78, Grp94, and Grp170 interact with alpha 1-antitrypsin mutants that are retained in the endoplasmic reticulum. Am J Physiol Gastrointest Liver Physiol. 2005;289:G444–G455. doi: 10.1152/ajpgi.00237.2004. [DOI] [PubMed] [Google Scholar]
  • 113.Shen Y, Hendershot LM. ERdj3, a stress-inducible endoplasmic reticulum DnaJ homologue, serves as a cofactor for BiP's interactions with unfolded substrates. Mol Biol Cell. 2005;16:40–50. doi: 10.1091/mbc.E04-05-0434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Jin Y, Awad W, Petrova K, Hendershot LM. Regulated release of ERdj3 from unfolded proteins by BiP. EMBO J. 2008;27:2873–2882. doi: 10.1038/emboj.2008.207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Petrova K, Oyadomari S, Hendershot LM, Ron D. Regulated association of misfolded endoplasmic reticulum lumenal proteins with P58/DNAJc3. EMBO J. 2008;27:2862–2872. doi: 10.1038/emboj.2008.199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.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;321:569–572. doi: 10.1126/science.1159293. [DOI] [PubMed] [Google Scholar]
  • 117.Vembar JSS, Jonikas MC, Hendershot LM, Weissman JS, Brodsky JL. J domain co-chaperone specificity defines the role of BiP during protein translocation. J Biol Chem. 2010;285:22484–22494. doi: 10.1074/jbc.M110.102186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Molinari M, Galli C, Vanoni O, Arnold SM, Kaufman RJ. Persistent glycoprotein misfolding activates the glucosidase II/UGT1-driven calnexin cycle to delay aggregation and loss of folding competence. Mol Cell. 2005;20:503–512. doi: 10.1016/j.molcel.2005.09.027. [DOI] [PubMed] [Google Scholar]
  • 119.Chan HY, Warrick JM, Gray-Board GL, Paulson HL, Bonini NM. Mechanisms of chaperone suppression of polyglutamine disease: selectivity, synergy and modulation of protein solubility in Drosophila. Hum Mol Genet. 2000;9:2811–2820. doi: 10.1093/hmg/9.19.2811. [DOI] [PubMed] [Google Scholar]
  • 120.Stevens FJ, Argon Y. Pathogenic light chains and the B-cell repertoire. Immunol. Today. 1999;20:451–457. doi: 10.1016/s0167-5699(99)01502-9. [DOI] [PubMed] [Google Scholar]
  • 121.Yoshiike Y, Minai R, Matsuo Y, Chen YR, Kimura T, Takashima A. Amyloid oligomer conformation in a group of natively folded proteins. PLoS One. 2008;3:e3235. doi: 10.1371/journal.pone.0003235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Kroczynska B, King-Simmons L, Alloza L, Alava MA, Elguindi EC, Blond SY. BIP co-chaperone MTJ1/ERDJ1 interacts with inter-alpha-trypsin inhibitor heavy chain 4. Biochem Biophys Res Commun. 2005;338:1467–1477. doi: 10.1016/j.bbrc.2005.10.101. [DOI] [PubMed] [Google Scholar]
  • 123.Hosoda A, Tokuda M, Akai R, Kohno K, Iwawaki T. Positive contribution of ERdj5/JPDI to endoplasmic reticulum protein quality control in the salivary gland. Biochem J. 2009;425:117–125. doi: 10.1042/BJ20091269. [DOI] [PubMed] [Google Scholar]
  • 124.Ladiges WC, Knoblaugh SE, Morton JF, Korth MJ, Sopher BL, Baskin CR, MacAuley A, Goodman AG, LeBoeuf RC, Katze MG. Pancreatic beta-cell failure and diabetes in mice with a deletion mutation of the endoplasmic reticulum molecular chaperone gene P58IPK. Diabetes. 2005;54:1074–1081. doi: 10.2337/diabetes.54.4.1074. [DOI] [PubMed] [Google Scholar]
  • 125.Kabani M, Beckerich JM, Brodsky JL. Nucleotide exchange factor for the yeast Hsp70 molecular chaperone Ssa1p. Mol Cell Biol. 2002;22:4677–4689. doi: 10.1128/MCB.22.13.4677-4689.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Anttonen AK, Mahjneh I, Hamalainen RH, Lagier-Tourenne C, Kopra O, Waris L, Anttonen M, Joensuu T, Kalimo H, Paetau A, Tranebjaerg L, Chaigne D, Koenig M, Eeg-Olofsson O, Udd B, Somer M, Somer H, Lehesjoki AE. The gene disrupted in Marinesco-Sjogren syndrome encodes SIL1, an HSPA5 cochaperone. Nat Genet. 2005;37:1309–1311. doi: 10.1038/ng1677. [DOI] [PubMed] [Google Scholar]
  • 127.Senderek J, Krieger M, Stendel C, Bergmann C, Moser M, Breitbach-Faller N, Rudnik-Schoneborn S, Blaschek A, Wolf NI, Harting I, North K, Smith J, Muntoni F, Brockington M, Quijano-Roy S, Renault F, Herrmann R, Hendershot LM, Schroder JM, Lochmuller H, Topaloglu H, Voit T, Weis J, Ebinger F, Zerres K. Mutations in SIL1 cause Marinesco-Sjogren syndrome, a cerebellar ataxia with cataract and myopathy. Nat Genet. 2005;37:1312–1314. doi: 10.1038/ng1678. [DOI] [PubMed] [Google Scholar]
  • 128.Zhao L, Longo-Guess C, Harris BS, Lee JW, Ackerman SL. Protein accumulation and neurodegeneration in the woozy mutant mouse is caused by disruption of SIL1, a cochaperone of BiP. Nat Genet. 2005;37:974–979. doi: 10.1038/ng1620. [DOI] [PubMed] [Google Scholar]
  • 129.Steel GJ, Fullerton DM, Tyson JR, Stirling CJ. Coordinated activation of Hsp70 chaperones. Science. 2004;303:98–101. doi: 10.1126/science.1092287. [DOI] [PubMed] [Google Scholar]
  • 130.Zhao L, Rosales C, Seburn K, Ron D, Ackerman SL. Alteration of the unfolded protein response modifies neurodegeneration in a mouse model of Marinesco-Sjogren syndrome. Hum Mol Genet. 2009;19:25–35. doi: 10.1093/hmg/ddp464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Rutkowski DT, Kang SW, Goodman AG, Garrison JL, Taunton J, Katze MG, Kaufman RJ, Hegde RS. The role of p58IPK in protecting the stressed endoplasmic reticulum. Mol Biol Cell. 2007;18:3681–3691. doi: 10.1091/mbc.E07-03-0272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Kyostila K, Cizinauskas S, Seppala EH, Suhonen E, Jeserevics J, Sukura A, Syrja P, Lohi H. A SEL1L mutation links a canine progressive early-onset cerebellar ataxia to the endoplasmic reticulum-associated protein degradation (ERAD) machinery. PLoS Genet. 8:e1002759. doi: 10.1371/journal.pgen.1002759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Lee JW, Beebe K, Nangle LA, Jang J, Longo-Guess CM, Cook SA, Davisson MT, Sundberg JP, Schimmel P, Ackerman SL. Editing-defective tRNA synthetase causes protein misfolding and neurodegeneration. Nature. 2006;443:50–55. doi: 10.1038/nature05096. [DOI] [PubMed] [Google Scholar]
  • 134.Eletto D, Dersh D, Argon Y. GRP94 in ER quality control and stress responses. Semin. Cell Dev. Biol. 2010;21:479–485. doi: 10.1016/j.semcdb.2010.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Melnick J, Aviel S, Argon Y. The endoplasmic reticulum stress protein GRP94, in addition to BiP, associates with unassembled immunoglobulin chains. J. Biol. Chem. 1992;267:21303–21306. [PubMed] [Google Scholar]
  • 136.Melnick J, Argon Y. Molecular chaperones and the biosynthesis of antigen receptors. Immunol. Today. 1995;16:243–250. doi: 10.1016/0167-5699(95)80167-7. [DOI] [PubMed] [Google Scholar]
  • 137.Kim PS, Bole D, Arvan P. Transient aggregation of nascent thyroglobulin in the endoplasmic reticulum: relationship to the molecular chaperone, BiP. J. Cell Biol. 1992;118:541–549. doi: 10.1083/jcb.118.3.541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Muresan Z, Arvan P. Thyroglobulin transport along the secretory pathway. Investigation of the role of molecular chaperone, GRP94, in protein export from the endoplasmic reticulum [published erratum appears in J Biol Chem 1997 Nov 28;272(48):30590] J. Biol. Chem. 1997;272:26095–26102. doi: 10.1074/jbc.272.42.26095. [DOI] [PubMed] [Google Scholar]
  • 139.Ostrovsky O, Makarewich C, Snapp EL, Argon Y. An essential role for ATP binding and hydrolysis in the chaperone activity of GRP94 in cells. Proc Natl Acad Sci U S A. 2009;106:11600–11605. doi: 10.1073/pnas.0902626106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Liu B, Yang Y, Qiu Z, Staron M, Hong F, Li Y, Wu S, Hao B, Bona R, Han D, Li Z. Folding of Toll-like receptors by the HSP90 paralogue gp96 requires a substrate-specific cochaperone. Nat Commun. 2010;1:79. doi: 10.1038/ncomms1070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Wegele H, Wandinger SK, Schmid AB, Reinstein J, Buchner J. Substrate transfer from the chaperone Hsp70 to Hsp90. J Mol Biol. 2006;356:802–811. doi: 10.1016/j.jmb.2005.12.008. [DOI] [PubMed] [Google Scholar]
  • 142.Eletto D, Maganty A, Eletto D, Dersh D, Makarewich C, Biswas C, Doroudgar S, Glembotski CC, Argon Y. Limitation of Individual Folding Resources in the ER Leads to Outcomes Distinct from the Unfolded Protein Response. Journal of Cell Science. 2012 doi: 10.1242/jcs.108928. Epub August 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Wanderling S, Simen BB, Ostrovsky O, Ahmed NT, Vogen S, Gidalevitz T, Argon Y. GRP94 is essential for mesoderm induction and muscle development because it regulates IGF secretion. Mol Biol Cell. 2007;18:3764–3775. doi: 10.1091/mbc.E07-03-0275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Liu B, Li Z. Endoplasmic reticulum HSP90M (gp96, grp94) optimizes B-cell function via chaperoning integrin and TLR but not immunoglobulin. Blood. 2008;112:1223–1230. doi: 10.1182/blood-2008-03-143107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Barton E, Park S, James JK, Makarewich CA, Eletto D, Philippou A, Lei H, Brisson B, Ostrovsky O, Li Z, Argon Y. Muscle-specific deletion of GRP94 leads to small muscles and impaired organismal growth due to inhibition of muscle-derived IGF production. FASEB J. 2012 doi: 10.1096/fj.11-203026. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Staron M, Wu S, Hong F, Stojanovic A, Du X, Bona R, Liu B, Li Z. Heat-shock protein gp96/grp94 is an essential chaperone for the platelet glycoprotein Ib-IX-V complex. Blood. 2011;117:7136–7144. doi: 10.1182/blood-2011-01-330464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Randow F, Seed B. Endoplasmic reticulum chaperone gp96 is required for innate immunity but not cell viability. Nat Cell Biol. 2001;3:891–896. doi: 10.1038/ncb1001-891. [DOI] [PubMed] [Google Scholar]
  • 148.Widmer C, Gebauer JM, Brunstein E, Rosenbaum S, Zaucke F, Drogemuller C, Leeb T, Baumann U. Molecular basis for the action of the collagen-specific chaperone Hsp47/SERPINH1 and its structure-specific client recognition. Proc Natl Acad Sci U S A. 2012;109:13243–13247. doi: 10.1073/pnas.1208072109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Koide T, Nishikawa Y, Asada S, Yamazaki CM, Takahara Y, Homma DL, Otaka A, Ohtani K, Wakamiya N, Nagata K, Kitagawa K. Specific recognition of the collagen triple helix by chaperone HSP47. II. The HSP47-binding structural motif in collagens and related proteins. J Biol Chem. 2006;281:11177–11185. doi: 10.1074/jbc.M601369200. [DOI] [PubMed] [Google Scholar]
  • 150.Tasab M, Jenkinson L, Bulleid NJ. Sequence-specific recognition of collagen triple helices by the collagen-specific molecular chaperone HSP47. J Biol Chem. 2002;277:35007–35012. doi: 10.1074/jbc.M202782200. [DOI] [PubMed] [Google Scholar]
  • 151.Nagai N, Hosokawa M, Itohara S, Adachi E, Matsushita T, Hosokawa N, Nagata K. Embryonic lethality of molecular chaperone hsp47 knockout mice is associated with defects in collagen biosynthesis. J Cell Biol. 2000;150:1499–1506. doi: 10.1083/jcb.150.6.1499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Maattanen P, Gehring K, Bergeron JJ, Thomas DY. Protein quality control in the ER: the recognition of misfolded proteins. Semin Cell Dev Biol. 2010;21:500–511. doi: 10.1016/j.semcdb.2010.03.006. [DOI] [PubMed] [Google Scholar]
  • 153.Hebert DN, Foellmer B, Helenius A. Glucose trimming and reglucosylation determine glycoprotein association with calnexin in the endoplasmic reticulum. Cell. 1995;81:425–433. doi: 10.1016/0092-8674(95)90395-x. [DOI] [PubMed] [Google Scholar]
  • 154.Ruddock LW, Molinari M. N-glycan processing in ER quality control. J Cell Sci. 2006;119:4373–4380. doi: 10.1242/jcs.03225. [DOI] [PubMed] [Google Scholar]
  • 155.Vassilakos A, Michalak M, Lehrman MA, Williams DB. Oligosaccharide binding characteristics of the molecular chaperones calnexin and calreticulin. Biochemistry. 1998;37:3480–3490. doi: 10.1021/bi972465g. [DOI] [PubMed] [Google Scholar]
  • 156.Wada I, Imai S, Kai M, Sakane F, Kanoh H. Chaperone function of calreticulin when expressed in the endoplasmic reticulum as the membrane-anchored and soluble forms. J. Biol. Chem. 1995;270:20298–20304. doi: 10.1074/jbc.270.35.20298. [DOI] [PubMed] [Google Scholar]
  • 157.Harris MR, Yu YY, Kindle CS, Hansen TH, Solheim JC. Calreticulin and calnexin interact with different protein and glycan determinants during the assembly of MHC class I. J Immunol. 1998;160:5404–5409. [PubMed] [Google Scholar]
  • 158.Ihara Y, Cohen-Doyle MF, Saito Y, Williams DB. Calnexin discriminates between protein conformational states and functions as a molecular chaperone in vitro. Mol Cell. 1999;4:331–341. doi: 10.1016/s1097-2765(00)80335-4. [DOI] [PubMed] [Google Scholar]
  • 159.Harris MR, Yu YY, Kindle CS, Hansen TH, Solheim JC. Calreticulin and calnexin interact with different protein and glycan determinants during the assembly of MHC class I. J Immunol. 1998;160:5404–5409. [PubMed] [Google Scholar]
  • 160.Kornfeld R, Kornfeld S. Assembly of asparagine-linked oligosaccharides. Annu. Rev. Biochem. 1985;54:631–664. doi: 10.1146/annurev.bi.54.070185.003215. [DOI] [PubMed] [Google Scholar]
  • 161.David V, Hochstenbach F, Rajagopalan S, Brenner MB. Interaction with newly synthesized and retained proteins in the endoplasmic reticulum suggests a chaperone function for human integral membrane protein IP90 (calnexin) J. Biol. Chem. 1993;268:9585–9592. [PubMed] [Google Scholar]
  • 162.Ou W-J, Cameron PH, Thomas DY, Bergeron JJM. Association of folding intermediates of glycoproteins with calnexin during protein maturation. Nature. 1993;364:771–776. doi: 10.1038/364771a0. [DOI] [PubMed] [Google Scholar]
  • 163.Hammond C, Braakman I, Helenius A. Role of N-linked oligosaccharide recognition, glucose trimming, and calnexin in glycoprotein folding and quality control. Proc. Natl. Acad. Sci. USA. 1994;91:913–917. doi: 10.1073/pnas.91.3.913. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Chavan M, Yan A, Lennarz WJ. Subunits of the translocon interact with components of the oligosaccharyl transferase complex. J Biol Chem. 2005;280:22917–22924. doi: 10.1074/jbc.M502858200. [DOI] [PubMed] [Google Scholar]
  • 165.Dejgaard K, Theberge JF, Heath-Engel H, Chevet E, Tremblay ML, Thomas DY. Organization of the Sec61 translocon, studied by high resolution native electrophoresis. J Proteome Res. 2010;9:1763–1771. doi: 10.1021/pr900900x. [DOI] [PubMed] [Google Scholar]
  • 166.Molinari M, Helenius A. Chaperone selection during glycoprotein translocation into the endoplasmic reticulum. Science. 2000;288:331–333. doi: 10.1126/science.288.5464.331. [DOI] [PubMed] [Google Scholar]
  • 167.Zapun A, Darby NJ, Tessier DC, Michalak M, Bergeron JJ, Thomas DY. Enhanced catalysis of ribonuclease B folding by the internacion of calnexin or calreticulin with ERp57. J Biol Chem. 1998;273:6009–6012. doi: 10.1074/jbc.273.11.6009. [DOI] [PubMed] [Google Scholar]
  • 168.Oliver JD, Roderick HL, Llewellyn DH, High S. ERp57 functions as a subunit of specific complexes formed with the ER lectins calreticulin and calnexin. Mol Biol Cell. 1999;10:2573–2582. doi: 10.1091/mbc.10.8.2573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Molinari M, Helenius A. Glycoproteins form mixed disulphides with oxidoreductases during folding in living cells. Nature. 1999;402:90–93. doi: 10.1038/47062. [DOI] [PubMed] [Google Scholar]
  • 170.Frickel EM, Riek R, Jelesarov I, Helenius A, Wuthrich K, Ellgaard L. TROSY-NMR reveals interaction between ERp57 and the tip of the calreticulin P-domain. Proc Natl Acad Sci USA99. 2002:1954–1959. doi: 10.1073/pnas.042699099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Helenius A. How N-linked oligosaccharides affect glycoprotein folding in the endoplasmic reticulum. Mol. Biol. Cell. 1994;5:253–265. doi: 10.1091/mbc.5.3.253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Liu H, Bowes RCr, van de Water B, Sillence C, Nagelkerke JF, Stevens JL. Endoplasmic reticulum chaperones GRP78 and calreticulin prevent oxidative stress, Ca2+ disturbances, and cell death in renal epithelial cells. J. Biol. Chem. 1997;272:21751–21759. doi: 10.1074/jbc.272.35.21751. [DOI] [PubMed] [Google Scholar]
  • 173.de Virgilio M, Kitzmuller C, Schwaiger E, Klein M, Kreibich G, Ivessa NE. Degradation of a short-lived glycoprotein from the lumen of the endoplasmic reticulum: the role of N-linked glycans and the unfolded protein response. Mol Biol Cell. 1999;10:4059–4073. doi: 10.1091/mbc.10.12.4059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.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;349:1278–1284. doi: 10.1016/j.bbrc.2006.08.186. [DOI] [PubMed] [Google Scholar]
  • 175.Hirao K, Natsuka Y, Tamura T, Wada I, Morito D, Natsuka S, Romero P, Sleno B, Tremblay LO, Herscovics A, Nagata K, Hosokawa N. EDEM3, a soluble EDEM homolog, enhances glycoprotein endoplasmic reticulum-associated degradation and mannose trimming. J Biol Chem. 2006;281:9650–9658. doi: 10.1074/jbc.M512191200. [DOI] [PubMed] [Google Scholar]
  • 176.Denzel A, Molinari M, Trigueros C, Martin JE, Velmurgan S, Brown S, Stamp G, Owen MJ. Early postnatal death and motor disorders in mice congenitally deficient in calnexin expression. Mol Cell Biol. 2002;22:7398–7404. doi: 10.1128/MCB.22.21.7398-7404.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Hebert DN, Bernasconi R, Molinari M. ERAD substrates: which way out? Semin Cell Dev Biol. 2010;21:526–532. doi: 10.1016/j.semcdb.2009.12.007. [DOI] [PubMed] [Google Scholar]
  • 178.Bause E. Structural requirements of N-glycosylation of proteins. Studies with proline peptides as conformational probes. Biochem J. 1983;209:331–336. doi: 10.1042/bj2090331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Caramelo JJ, Castro OA, de Prat-Gay G, Parodi AJ. The endoplasmic reticulum glucosyltransferase recognizes nearly native glycoprotein folding intermediates. J Biol Chem. 2004;279:46280–46285. doi: 10.1074/jbc.M408404200. [DOI] [PubMed] [Google Scholar]
  • 180.Sousa M, Parodi AJ. The molecular basis for the recognition of misfolded glycoproteins by the UDP-Glc:glycoprotein glucosyltransferase. EMBO J. 1995;14:4196–4203. doi: 10.1002/j.1460-2075.1995.tb00093.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Ritter C, Quirin K, Kowarik M, Helenius A. Minor folding defects trigger local modification of glycoproteins by the ER folding sensor GT. EMBO J. 2005;24:1730–1738. doi: 10.1038/sj.emboj.7600645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Jin H, Yan Z, Nam KH, Li J. Allele-specific suppression of a defective brassinosteroid receptor reveals a physiological role of UGGT in ER quality control. Mol Cell. 2007;26:821–830. doi: 10.1016/j.molcel.2007.05.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Schallus T, Jaeckh C, Feher K, Palma AS, Liu Y, Simpson JC, Mackeen M, Stier G, Gibson TJ, Feizi T, Pieler T, Muhle-Goll C. Malectin: a novel carbohydrate-binding protein of the endoplasmic reticulum and a candidate player in the early steps of protein N-glycosylation. Mol Biol Cell. 2008;19:3404–3414. doi: 10.1091/mbc.E08-04-0354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.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:e16304. doi: 10.1371/journal.pone.0016304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Hebert DN, Foellmer B, Helenius A. Calnexin and calreticulin promote folding, delay oligomerization and suppress degradation of influenza hemagglutinin in microsomes. EMBO J. 1996;15:2961–2968. [PMC free article] [PubMed] [Google Scholar]
  • 186.Hebert DN, Molinari M. In and out of the ER: protein folding, quality control, degradation, and related human diseases. Physiol Rev. 2007;87:1377–1408. doi: 10.1152/physrev.00050.2006. [DOI] [PubMed] [Google Scholar]
  • 187.Branza-Nichita N, Negroiu G, Petrescu AJ, Garman EF, Piatt FM, Wormald MR, Dwek RA, Petrescu SM. Mutations at critical N-glycosylation sites reduce tyrosinase activity by altering folding and quality control. J Biol Chem. 2000;275:8169–8175. doi: 10.1074/jbc.275.11.8169. [DOI] [PubMed] [Google Scholar]
  • 188.Toyoda T, Itai T, Arakawa T, Aoki KH, Yamaguchi H. Stabilization of human recombinanterythropoietin through interactions with the highly branched N-glycans. J Biochem. 2000;128:731–737. doi: 10.1093/oxfordjournals.jbchem.a022809. [DOI] [PubMed] [Google Scholar]
  • 189.Erbel PJ, Karimi-Nejad Y, De Beer T, Boelens R, Kamerling JP, Vliegenthart JF. Solution structure of the alpha-subunit of human chorionic gonadotropin. Eur J Biochem. 1999;260:490–498. doi: 10.1046/j.1432-1327.1999.00188.x. [DOI] [PubMed] [Google Scholar]
  • 190.Scott JE, Dawson JR. MHC class I expression and transport in a calnexin-deficient cell line. J. Immunol. 1995;155:143–148. [PubMed] [Google Scholar]
  • 191.Kraus A, Groenendyk J, Bedard K, Baldwin TA, Krause KH, Dubois-Dauphin M, Dyck J, Rosenbaum EE, Korngut L, Colley NJ, Gosgnach S, Zochodne D, Todd K, Agellon LB, Michalak M. Calnexin deficiency leads to dysmyelination. J Biol Chem. 2010;285:18928–18938. doi: 10.1074/jbc.M110.107201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Mesaeli N, Nakamura K, Zvaritch E, Dickie P, Dziak E, Krause KH, Opas M, MacLennan DH, Michalak M. Calreticulin is essential for cardiac development. J Cell Biol. 1999;144:857–868. doi: 10.1083/jcb.144.5.857. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Galligan JJ, Petersen DR. The human protein disulfide isomerase gene family. Hum Genomics. 2012;6:6. doi: 10.1186/1479-7364-6-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Pihlajaniemi T, Helaakoski T, Tasanen K, Myllyla R, Huhtala ML, Koivu J, Kivirikko KI. Molecular cloning of the beta-subunit of human prolyl 4-hydroxylase. This subunit and protein disulphide isomerase are products of the same gene. EMBO J. 1987;6:643–649. doi: 10.1002/j.1460-2075.1987.tb04803.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Cortini M, Sitia R. From antibodies to adiponectin: role of ERp44 in sizing and timing protein secretion. Diabetes Obes Metab. 2010;2(12 Suppl):39–47. doi: 10.1111/j.1463-1326.2010.01272.x. [DOI] [PubMed] [Google Scholar]
  • 196.Rutkevich LA, Cohen-Doyle MF, Brockmeier U, Williams DB. Functional relationship between protein disulfide isomerase family members during the oxidative folding of human secretory proteins. Mol Biol Cell. 2010;21:3093–3105. doi: 10.1091/mbc.E10-04-0356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Winter AD, McCormack G, Page AP. Protein disulfide isomerase activity is essential for viability and extracellular matrix formation in the nematode Caenorhabditis elegans. Dev Biol. 2007;308:449–461. doi: 10.1016/j.ydbio.2007.05.041. [DOI] [PubMed] [Google Scholar]
  • 198.Jessop CE, Tavender TJ, Watkins RH, Chambers JE, Bulleid NJ. Substrate specificity of the oxidoreductase ERp57 is determined primarily by its interaction with calnexin and calreticulin. J Biol Chem. 2009;284:2194–2202. doi: 10.1074/jbc.M808054200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Solda T, Garbi N, Hammerling GJ, Molinari M. Consequences of ERp57 deletion on oxidative folding of obligate and facultative clients of the calnexin cycle. J Biol Chem. 2006;281:6219–6226. doi: 10.1074/jbc.M513595200. [DOI] [PubMed] [Google Scholar]
  • 200.Jessop CE, Watkins RH, Simmons JJ, Tasab M, Bulleid NJ. Protein disulphide isomerase family members show distinct substrate specificity: P5 is targeted to BiP client proteins. J Cell Sci. 2009;122:4287–4295. doi: 10.1242/jcs.059154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Menon S, Lee J, Abplanalp WA, Yoo SE, Agui T, Furudate S, Kim PS, Arvan P. Oxidoreductase interactions include a role for ERp72 engagement with mutant thyroglobulin from the rdw/rdw rat dwarf. J Biol Chem. 2007;282:6183–6191. doi: 10.1074/jbc.M608863200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Walczak CP, Tsai B. A PDI family network acts distinctly and coordinately with ERp29 to facilitate polyomavirus infection. J Virol. 2010;85:2386–2396. doi: 10.1128/JVI.01855-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Shimizu Y, Meunier L, Hendershot LM. pERpl is significantly up-regulated during plasma cell differentiation and contributes to the oxidative folding of immunoglobulin. Proc Natl Acad Sci U S A. 2009;106:17013–17018. doi: 10.1073/pnas.0811591106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.van Anken E, Pena F, Hafkemeijer N, Christis C, Romijn EP, Grauschopf U, Oorschot VM, Pertel T, Engels S, Ora A, Lastun V, Glockshuber R, Klumperman J, Heck AJ, Luban J, Braakman I. Efficient IgM assembly and secretion require the plasma cell induced endoplasmic reticulum protein pERpl. Proc Natl Acad Sci U S A. 2009;106:17019–17024. doi: 10.1073/pnas.0903036106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Schmid FX, Baldwin RL. Acid catalysis of the formation of the slow-folding species of RNase A: evidence that the reaction is proline isomerization. Proc Natl Acad Sci U S A. 1978;75:4764–4768. doi: 10.1073/pnas.75.10.4764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Nigam SK, Jin YJ, Jin MJ, Bush KT, Bierer BE, Burakoff SJ. Localization of the FK506-binding protein, FKBP 13, to the lumen of the endoplasmic reticulum. Biochem. J. 1993;294:511–515. doi: 10.1042/bj2940511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Lodish HF, Kong N. Cyclosporin A inhibits an initial step in folding of transferrin within the endoplasmic reticulum. J. Biol. Chem. 1991;266:14835–14838. [PubMed] [Google Scholar]
  • 208.Bernasconi R, Solda T, Galli C, Pertel T, Luban J, Molinari M. Cyclosporine A-sensitive, cyclophilin B-dependent endoplasmic reticulum-associated degradation. PLoS One. 2010;5:e13008. doi: 10.1371/journal.pone.0013008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Davis EC, Broekelmann TJ, Ozawa Y, Mecham RP. Identification of tropoelastin as a ligand for the 65-kD FK506-binding protein, FKBP65, in the secretory pathway. J. Cell Biol. 1998;140:295–303. doi: 10.1083/jcb.140.2.295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Hanagata N, Li X. Osteoblast-enriched membrane protein IFITM5 regulates the association of CD9 with an FKBP11-CD81-FPRP complex and stimulates expression of interferon-induced genes. Biochem Biophys Res Commun. 2011;409:378–384. doi: 10.1016/j.bbrc.2011.04.136. [DOI] [PubMed] [Google Scholar]
  • 211.Baumann M, Giunta C, Krabichler B, Ruschendorf F, Zoppi N, Colombi M, Bittner RE, Quijano-Roy S, Muntoni F, Cirak S, Schreiber G, Zou Y, Hu Y, Romero NB, Carlier RY, Amberger A, Deutschmann A, Straub V, Rohrbach M, Steinmann B, Rostasy K, Karall D, Bonnemann CG, Zschocke J, Fauth C. Mutations in FKBP14 cause a variant of Ehlers-Danlos syndrome with progressive kyphoscoliosis, myopathy, and hearing loss. Am J Hum Genet. 2012;90:201–216. doi: 10.1016/j.ajhg.2011.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Choi JW, Sutor SL, Lindquist L, Evans GL, Madden BJ, Bergen HR, 3rd, Hefferan TE, Yaszemski MJ, Bram RJ. Severe osteogenesis imperfecta in cyclophilin B-deficient mice. PLoS Genet. 2009;5:e1000750. doi: 10.1371/journal.pgen.1000750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Barnes AM, Cabral WA, Weis M, Makareeva E, Mertz EL, Leikin S, Eyre D, Trujillo C, Marini JC. Absence of FKBP10 in recessive type XI osteogenesis imperfecta leads to diminished collagen cross-linking and reduced collagen deposition in extracellular matrix. Hum Mutat. 2012;33:1589–1598. doi: 10.1002/humu.22139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Jansen G, Maattanen P, Denisov AY, Scarffe L, Schade B, Balghi H, Dejgaard K, Chen LY, Muller WJ, Gehring K, Thomas DY. An interaction map of endoplasmic reticulum chaperones and foldases. Mol Cell Proteomics. 2012;11:710–723. doi: 10.1074/mcp.M111.016550. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Winter AD, Eschenlauer SC, McCormack G, Page AP. Loss of secretory pathway FK506-binding proteins results in cold-sensitive lethality and associate extracellular matrix defects in the nematode Caenorhabditis elegans. J Biol Chem. 2007;282:12813–12821. doi: 10.1074/jbc.M700274200. [DOI] [PubMed] [Google Scholar]
  • 216.Tremmel D, Tropschug M. Neurospora crassa FKBP22 is a novel ER chaperone and functionally cooperates with BiP. J Mol Biol. 2007;369:55–68. doi: 10.1016/j.jmb.2007.01.092. [DOI] [PubMed] [Google Scholar]
  • 217.Wang Y, Han R, Wu D, Li J, Chen C, Ma H, Mi H. The binding of FKBP23 to BiP modulates BiP's ATPase activity with its PPIase activity. Biochem Biophys Res Commun. 2007;354:315–320. doi: 10.1016/j.bbrc.2006.12.209. [DOI] [PubMed] [Google Scholar]
  • 218.Feng M, Gu C, Ma S, Wang Y, Liu H, Han R, Gao J, Long Y, Mi H. Mouse FKBP23 mediates conformer-specific functions of BiP by catalyzing Prol 17 cis/trans isomerization. Biochem Biophys Res Commun. 2011;408:537–540. doi: 10.1016/j.bbrc.2011.04.050. [DOI] [PubMed] [Google Scholar]
  • 219.Horibe T, Yosho C, Okada S, Tsukamoto M, Nagai H, Hagiwara Y, Tujimoto Y, Kikuchi M. The chaperone activity of protein disulfide isomerase is affected by cyclophilin B and cyclosporin A in vitro. J Biochem. 2002;132:401–407. doi: 10.1093/oxfordjournals.jbchem.a003236. [DOI] [PubMed] [Google Scholar]
  • 220.Motohashi K, Koyama F, Nakanishi Y, Ueoka-Nakanishi H, Hisabori T. Chloroplast cyclophilin is a target protein of thioredoxin. Thiol modulation of the peptidyl-prolyl cis-trans isomerase activity. J Biol Chem. 2003;278:31848–31852. doi: 10.1074/jbc.M304258200. [DOI] [PubMed] [Google Scholar]
  • 221.Kozlov G, Bastos-Aristizabal S, Maattanen P, Rosenauer A, Zheng F, Killikelly A, Trempe JF, Thomas DY, Gehring K. Structural basis of cyclophilin B binding by the calnexin/calreticulin P-domain. J Biol Chem. 2010;285:35551–35557. doi: 10.1074/jbc.M110.160101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Meunier L, Usherwood YK, Chung KT, Hendershot LM. A subset of chaperones and folding enzymes form multiprotein complexes in endoplasmic reticulum to bind nascent proteins. Mol Biol Cell. 2002;13:4456–4469. doi: 10.1091/mbc.E02-05-0311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Zhang J, Herscovitz H. Nascent lipidated apolipoprotein B is transported to the Golgi as an incompletely folded intermediate as probed by its association with network of endoplasmic reticulum molecular chaperones, GRP94, ERp72, BiP, calreticulin, and cyclophilin B. J Biol Chem. 2003;278:7459–7468. doi: 10.1074/jbc.M207976200. [DOI] [PubMed] [Google Scholar]
  • 224.Di Jeso B, Park YN, Ulianich L, Treglia AS, Urbanas ML, High S, Arvan P. Mixed-disulfide folding intermediates between thyroglobulin and endoplasmic reticulum resident oxidoreductases ERp57 and protein disulfide isomerase. Mol Cell Biol. 2005;25:9793–9805. doi: 10.1128/MCB.25.22.9793-9805.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Lee J, Di Jeso B, Arvan P. Maturation of thyroglobulin region-I. J Biol Chem. 2011 doi: 10.1074/jbc.M111.281337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Kim PS, Arvan P. Calnexin and BiP act as sequential molecular chaperones during thyroglobulin folding in the endoplasmic reticulum. J. Cell Biol. 1995:29–38. doi: 10.1083/jcb.128.1.29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Nehls S, Snapp EL, Cole NB, Zaal KJ, Kenworthy AK, Roberts TH, Ellenberg J, Presley JF, Siggia E, Lippincott-Schwartz J. Dynamics and retention of misfolded proteins in native ER membranes. Nat Cell Biol. 2000;2:288–295. doi: 10.1038/35010558. [DOI] [PubMed] [Google Scholar]
  • 228.Snapp EL, Sharma A, Lippincott-Schwartz J, Hegde RS. Monitoring chaperone engagement of substrates in the endoplasmic reticulum of live cells. Proc Natl Acad Sci U S A. 2006;103:6536–6541. doi: 10.1073/pnas.0510657103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Lai CW, Aronson DE, Snapp EL. BiP availability distinguishes states of homeostasis and stress in the endoplasmic reticulum of living cells. Mol Biol Cell. 2010;21:1909–1921. doi: 10.1091/mbc.E09-12-1066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.England JL. Allostery in protein domains reflects a balance of steric and hydrophobic effects. Structure. 2011;19:967–975. doi: 10.1016/j.str.2011.04.009. [DOI] [PubMed] [Google Scholar]
  • 231.Luber CA, Cox J, Lauterbach H, Fancke B, Selbach M, Tschopp J, Akira S, Wiegand M, Hochrein H, O'Keeffe M, Mann M. Quantitative proteomics reveals subset-specific viral recognition in dendritic cells. Immunity. 2010;32:279–289. doi: 10.1016/j.immuni.2010.01.013. [DOI] [PubMed] [Google Scholar]
  • 232.Wisniewski JR, Ostasiewicz P, Dus K, Zielinska DF, Gnad F, Mann M. Extensive quantitative remodeling of the proteome between normal colon tissue and adenocarcinoma. Mol Syst Biol. 2012;8:611. doi: 10.1038/msb.2012.44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Boisvert FM, Ahmad Y, Gierlinski M, Charriere F, Lamont D, Scott M, Barton G, Lamond AI. A quantitative spatial proteomics analysis of proteome turnover in human cells. Mol Cell Proteomics. 2012;11 doi: 10.1074/mcp.M111.011429. M111 011429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Dudek J, Benedix J, Cappel S, Greiner M, Jalal C, Muller L, Zimmermann R. Functions and pathologies of BiP and its interaction partners. Cell Mol Life Sci. 2009;66:1556–1569. doi: 10.1007/s00018-009-8745-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Wang N, Hebert DN. Tyrosinase maturation through the mammalian secretory pathway: bringing color to life. Pigment Cell Res. 2006;19:3–18. doi: 10.1111/j.1600-0749.2005.00288.x. [DOI] [PubMed] [Google Scholar]
  • 236.Cioaca D, Ghenea S, Spiridon LN, Marin M, Petrescu AJ, Petrescu SM. C-terminus glycans with critical functional role in the maturation of secretory glycoproteins. PLoS One. 2011;6:e19979. doi: 10.1371/journal.pone.0019979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Parham P, Alpert BN, Orr HT, Strominger JL. Carbohydrate moiety of HLA antigens. Antigenic properties and amino acid sequences around the site of glycosylation. J Biol Chem. 1977;252:7555–7567. [PubMed] [Google Scholar]
  • 238.Nicolaou N, Margadant C, Kevelam SH, Lilien MR, Oosterveld MJ, Kreft M, van Eerde AM, Pfundt R, Terhal PA, van der Zwaag B, Nikkels PG, Sachs N, Goldschmeding R, Knoers NV, Renkema KY, Sonnenberg A. Gain of glycosylation in integrin alpha3 causes lung disease and nephrotic syndrome. J Clin Invest. 2012;122:4375–4387. doi: 10.1172/JCI64100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Wu GC, Lai HL, Lin YW, Chu YT, Chern Y. N-glycosylation and residues Asn805 andAsn890 are involved in the functional properties of type VI adenylyl cyclase. J Biol Chem. 2001;276:35450–35457. doi: 10.1074/jbc.M009704200. [DOI] [PubMed] [Google Scholar]
  • 240.Walker AK, Soo KY, Levina V, Talbo GH, Atkin JD. N-linked Glycosylation Modulates Dimerization of Protein Disulfide Isomerase Family A Member 2 (PDIA2) FEBS J. 2012 doi: 10.1111/febs.12063. [DOI] [PubMed] [Google Scholar]
  • 241.Ohgomori T, Nanao T, Morita A, Ikekita M. Asn54-linked glycan is critical for functional folding of intercellular adhesion molecule-5. Glycoconj J. 2011;29:47–55. doi: 10.1007/s10719-011-9363-0. [DOI] [PubMed] [Google Scholar]
  • 242.Bence M, Sahin-Toth M. Asparagine-linked glycosylation of human chymotrypsin C is required for folding and secretion but not for enzyme activity. FEBS J. 2011;278:4338–4350. doi: 10.1111/j.1742-4658.2011.08351.x. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES