Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2017 Jul 24;591(17):2616–2635. doi: 10.1002/1873-3468.12750

Repair or destruction—an intimate liaison between ubiquitin ligases and molecular chaperones in proteostasis

Éva Kevei 1,, Wojciech Pokrzywa 2,, Thorsten Hoppe 3,
PMCID: PMC5601288  PMID: 28699655

Abstract

Cellular differentiation, developmental processes, and environmental factors challenge the integrity of the proteome in every eukaryotic cell. The maintenance of protein homeostasis, or proteostasis, involves folding and degradation of damaged proteins, and is essential for cellular function, organismal growth, and viability 1, 2. Misfolded proteins that cannot be refolded by chaperone machineries are degraded by specialized proteolytic systems. A major degradation pathway regulating cellular proteostasis is the ubiquitin (Ub)/proteasome system (UPS), which regulates turnover of damaged proteins that accumulate upon stress and during aging. Despite a large number of structurally unrelated substrates, Ub conjugation is remarkably selective. Substrate selectivity is mainly provided by the group of E3 enzymes. Several observations indicate that numerous E3 Ub ligases intimately collaborate with molecular chaperones to maintain the cellular proteome. In this review, we provide an overview of specialized quality control E3 ligases playing a critical role in the degradation of damaged proteins. The process of substrate recognition and turnover, the type of chaperones they team up with, and the potential pathogeneses associated with their malfunction will be further discussed.

Keywords: aging, chaperone, CHIP, E3 ligase, proteostasis, quality control, ubiquitin

Abbreviation

AD, Alzheimer's disease

ALS, amyotrophic lateral sclerosis

ARTs, arrestin‐related trafficking adaptors

BMP, bone morphogenetic proteins

CAP, chaperone‐assisted proteasomal degradation

CASA, chaperone‐assisted selective autophagy

CFTR, cystic fibrosis transmembrane conductance regulator

CHIP, C terminus of Hsc70‐interacting protein

CMA, chaperone‐mediated autophagy

CP, core particle

DUBs, deubiquitylating enzymes

ER, endoplasmic reticulum

ERBB2, erythroblastic leukemia viral oncogene homolog 2

GR, glucocorticoid receptor

H3, Histone3

HD, Huntington's disease

HSF1, heat shock factor 1

HSP, heat shock proteins

Hul5, HECT ubiquitin ligase 5

IIS, insulin/IGF signaling

INM, inner nuclear membrane

LRR, leucine‐rich repeat

MLPs, mislocalized membrane proteins

NAC, nascent polypeptide‐associated complex

ONM, outer nuclear membrane

PD, Parkinson's disease

PML, promyelocytic leukemia

polyQ, polyglutamine

PRMT5, protein arginine methyltransferase 5

Psh1, Pob3/Spt16/histone 1

RING, really interesting new gene

RP, regulatory particle

SCF, Skp1/Cullin/F‐box

Smad1, Sma‐mother against decapentaplegic 1

SOD1, superoxide dismutase1

TPR, tetratricopeptide repeat

Ub, ubiquitin

UPS, ubiquitin/proteasome system

Concept of proteostasis

The three‐dimensional organization and conformation of a polypeptide chain is important for its cellular function. Maintaining the correct folding state of a protein is challenging particularly due to kinetic effects of a crowded cellular environment 3. The high concentration of macromolecules within most intracellular compartments strongly increases the tendency of misfolding of non‐native and structurally flexible proteins, often causing their polymerization and aggregate formation 3. Certain proteins only obtain an ordered, native structure and adopt folded conformations upon binding to the appropriate partner molecule or chaperone 4, 5. Cells are frequently exposed to proteotoxic conditions, including heat or oxidative stress, which makes it even more difficult to establish and preserve native protein structures 6. Genetic mutations might also increase the tendency of protein aggregation resulting in significant pressure on the cellular protein quality control systems 7. Intracellular pathways involved in the maintenance of the proteome build a complex proteostasis network. The term proteostasis refers to the preservation of the proper concentration, distribution, and function of proteins. The intricate balance is mainly achieved by vigilant control and safeguarding of protein synthesis, protein maturation, and folding, protein transport, as well as the timely disposal of unwanted and damaged proteins by the main proteolytic routes: the ubiquitin/proteasome system (UPS) or the lysosome‐autophagy pathway 8, 9, 10.

With age, the ability of postmitotic cells to keep a balanced proteome is gradually compromised particularly by downregulation of molecular chaperones and reduced efficiency of protein degradation 7. As such, impairment of proteostasis is seen as one major hallmark of aging, correlated with dementia and neurodegeneration, type 2 diabetes, cystic fibrosis, cancer, and cardiovascular diseases 11, 12. Notably, longevity‐promoting pathways, such as dietary restriction, insulin/IGF signaling (IIS), mitochondrial respiration, or germ line immortality provide increased stability to the proteome, delaying the onset of age‐related diseases 1, 9, 13, 14. One of the central nodes in the eukaryotic proteostasis network is the interaction between molecular chaperones and proteolytic machineries. To maintain the cellular proteome molecular chaperones and ubiquitin (Ub)‐dependent degradation pathways coordinate protein refolding and removal of terminally damaged proteins. Irreversibly affected proteins are particularly recognized by chaperone‐assisted E3 Ub ligases, which target them for degradation by the UPS or autophagy.

Protein degradation machineries

Ubiquitin/proteasome system (UPS)

Selective degradation of misfolded or aggregated proteins is crucial to maintain functionality of the cell. A fundamental proteolytic module of the cellular proteostasis network is the UPS 9. Substrates of the UPS are earmarked by covalent attachment of Ub to internal lysine residues through the concerted action of E1 Ub‐activating enzymes, E2 Ub‐conjugating enzymes, and E3 Ub ligases 15, 16, 17. Ubiquitylation is a dynamic and reversible process as deubiquitylating enzymes (DUBs) modulate the size and topology of poly Ub chains and thereby influence the fate of the conjugated substrate. DUBs that bind to the proteasome either remove proteolytic Ub tags 18 or edit the topology of Ub chains to generate efficient degradation signals 19, 20. Moreover, DUBs catalyze processing of inactive Ub precursors and recycling of inhibitory free Ub chains, which otherwise would inhibit polyubiquitin–substrate binding at the proteasome 21, 22. The 26S proteasome is a multicatalytic protease complex composed of a barrel‐shaped 20S proteolytic core particle (CP) and a 19S regulatory particle (RP) translocating substrates into the 20S CP where they are degraded into short peptides 23. Usually polyubiquitin attachment is sufficient for targeting substrate proteins for proteasomal turnover 17. While chains connected through Lys48 of Ub promote proteasomal degradation, Lys63‐linked chains provide regulatory or targeting functions 24.

Despite the large number of structurally unrelated substrates, Ub conjugation is remarkably selective. E3 Ub ligases represent the largest group of enzymes within the UPS, which is linked to their key role in substrate selection. Most E3 enzymes are not essential for cell growth and exhibit only mild loss‐of‐function phenotypes, suggesting the existence of similar functional redundancy and adaptation mechanisms even between degradation pathways of the UPS. The detailed analysis of several classes of E3 ligases led to identification of specific substrates and molecular pathways that they regulate 15, 16. Interestingly, a subgroup of specialized quality control E3 ligases have been identified to team up with a variety of molecular chaperones in order to recognize and target particularly damaged proteins for proteolysis (Fig. 1).

Figure 1.

Figure 1

Illustrative representation of subcellular localization of various PQC Ub ligases. E3 enzymes operating in quality control pathways (e.g., CHIP, Doa10, HUWE1, Listerin, Rsp5, and San1) and associated chaperones (in green and gray) are widely distributed in most subcellular compartments, including nucleus, cytoplasm, ER, and plasma membrane.

Selective autophagy

The other central component of the proteolytic system is the autophagy‐lysosome pathway, which supports proteostasis by turnover of defective and aggregated proteins, multimeric complexes, and even whole organelles that cannot be handled by the proteasome 25, 26, 27. A characteristic hallmark of macroautophagy (hereafter autophagy) is the formation of double‐membrane autophagosomes, which engulf their particular cargo substrate and deliver it to the lysosome for degradation. Although nonselective autophagy is mainly induced to recycle nutrients upon starvation, selective autophagy degrades specific cargos in response to environmental stress conditions and physiological changes including aging 28, 29, 30. Substrate selectivity is ensured by a cargo–ligand–receptor–scaffold interaction 27. As an initial step, the interaction between receptor and scaffold recruits a specific cargo to the phagophore assembly site for subsequent autophagosome formation. Commonly, the receptor proteins bind ATG8/LC3, which couples the cargo directly with the macroautophagy apparatus. Interestingly, some branches of selective autophagy also use Ub as recognition signal on target proteins, which involves Ub‐selective autophagy receptors and subsequent degradation of targets in the lysosome 31.

In higher organisms, selective degradation of single proteins is also arranged via chaperone‐mediated autophagy (CMA) and chaperone‐assisted selective autophagy (CASA), initiating cargo uptake directly at the lysosomal or endosomal membrane through a specific protein translocation complex 32. In both variants of selective autophagy, chaperones, and quality control E3 ligases play a key role. For instance, CMA‐based degradation of the transcription factor HIF1A requires the concerted action of HspA8/Hsc70 (heat shock 70 kDa protein 8) and the E3 ligase C terminus of Hsc70‐interacting protein (CHIP) 33.

Upon impaired capacity of the UPS and CMA, substrates are degraded by CASA 34. The CASA complex contains the molecular chaperones Hsc70 and HspB8, the cochaperone BAG3, and the E3 ligase CHIP 35. Under normal growth conditions, CASA is the main route for chaperone‐mediated lysosomal degradation, whereas CMA is induced by proteotoxic stress. CASA (like CMA) is also necessary for protein quality control in aged cells, which is reflected by elevated BAG3 levels and increased targeting of oxidized and ubiquitylated proteins to the lysosome in aged neurons 36. Substrates to be degraded by this complex include polyglutamine (polyQ)‐expanded huntingtin 37 and superoxide dismutase (SOD1) 38. Furthermore, CASA is essential for muscle maintenance especially during aging 39.

The chaperone network

With increasing genome size, the amount of proteostasis guarding factors expressed in eukaryotic cells has been adjusted to the growing complexity of the proteome during evolution 40. The coevolvement of proteostasis networks helps to cope with the higher burden of protein folding allocated to different subcellular organelles, more complex developmental reorganization of cell type‐specific metabolism, and unique stresses faced by multicellular, complex organisms. Thus, it is not surprising that the human chaperome consisting of chaperones, cochaperones, and adaptors possess more than 300 different members 41. Molecular chaperones are folding machines that assist client proteins in acquiring and keeping their active conformation, directing their folding, unfolding, and refolding. In case of irreversible damage chaperones also direct misfolded proteins toward specialized proteolytic systems of the cell 42. Chaperones coordinate lysosome‐dependent degradation pathways like CASA and CMA, and also target misfolded proteins to the 26S proteasome in chaperone‐assisted proteasomal degradation (CAP) 34, 43, 44. Chaperones usually recognize exposed hydrophobic protein surface and help client proteins to acquire and keep their active conformation by directing (re)folding 45. The biggest classes of chaperones are named according to their molecular weight [Hsp100, Hsp90, Hsp70, Hsp60, Hsp40, and small heat shock proteins (HSPs)] 6 and form different subgroups based on their mechanistic function.

The major de novo protein folding chaperones in eukaryotes are Hsp90 and Hsp70. These ATP‐dependent chaperones appear as constitutively expressed and stress‐induced forms and team up with various cochaperones for substrate recognition, binding, and activation 46, 47, 48, 49. Hsp70 and Hsp90 chaperone systems play a role in different organelles of the eukaryotic cell and regulate a wide range of events including folding of de novo synthesized polypeptides, refolding, or degradation of misfolded proteins. Furthermore, they also show disaggregation activity, facilitate protein translocation through membranes and they are involved in remodeling of multimeric protein complexes 50, 51. Hsp90 is also involved in regulation of receptor‐ligand binding or assembly of protein complexes, and has been implicated in regulatory pathways such as DNA repair or immune response 52. The broader role of this chaperone is well reflected by its various client proteins. In fact, it has been shown that Hsp90 associates with more than 10% of the total proteome 53.

The HSP60 chaperones are also known as chaperonins 54. The mitochondria localized HSP60–HSP10 and the cytosolic TriC/CCT complex chaperonins act as multimeric ring‐shaped folding chambers that encapsulate client proteins for folding 55, 56. Cochaperones, like HSP40/J‐proteins, and BAG family cochaperones regulate the activity of their cognate chaperones through the modulation of their ATPase cycle or via binding substrate proteins or other cochaperones 50, 57, 58, 59.

Age‐dependent and disease‐linked changes of the chaperome

Molecular chaperones are vital for protein quality assurance recognizing non‐native proteins and diminishing their toxicity. In the presence of proteotoxic conditions including heat stress, oxidative changes, and aging, all cells induce a highly conserved gene expression program, the heat stress response. This stress response is tightly regulated in eukaryotes, inducing expression of Hsp genes by the evolutionarily conserved transcription factor heat shock factor 1 (HSF1) 9, 60, 61.

Studying the Caenorhabditis elegans and human chaperome, Brehme et al. has shown that a conserved subnetwork of chaperones safeguards the aging process 41. The expression of the major cytosolic chaperones, including Hsc70, Hsp90s, CCT/TRiC complex, Hsp40, and tetratricopeptide motif repeat (TPR)‐domain cochaperones, is repressed in old worms and the adult human brain. This age‐related chaperome repression leads to accumulation of misfolded proteins and increases the risk of proteotoxic diseases, such as Parkinson's disease (PD), Alzheimer's disease (AD), and Huntington's disease (HD). On the contrary, increased HSP expression have been attributed to many human cancer types, providing the fast dividing cells with ultimate protein folding capacity, thus promoting tumor cell survival, proliferation, and invasion 62, 63. A recent study sheds light on the underlying chaperome reorganization event in the cell, which could trigger the survival of tumors. Examining the chaperone complexes of different tumor species, Rodina and coworkers identified the formation of tumor‐specific chaperone subnetworks that are distinct to physiological chaperone interactions 64. This so‐called cancer epi‐chaperome is based on enhanced physical integration of the two major cellular chaperone networks of Hsc70 and Hsp90. The epi‐chaperome of these cancer cells is nucleated by stable, high molecular weight complexes of Hsp90 and Hsc70, and supports the survival of fast dividing cells 64.

E3–chaperone interaction in proteostasis maintenance

As a result of cellular and environmental stresses, the eukaryotic cell is continuously confronted with handling defective proteins. Compartmentalized degradation pathways are specifically involved in the removal of misfolded proteins of the endoplasmic reticulum (ER), mitochondrion, cytoplasm, and nucleus. In order to target damaged proteins for degradation, specialized E3 Ub ligases are recruited by targeting chaperones. So far only a few E3–chaperone complexes have been described mechanistically (Fig. 2 and Table 1). Quantitative protein interaction analysis suggests that more than 30% of all human E3 ligases interact with Hsp90 65. This observation indicates a highly complex and intricate PQC network guided by diverse E3–chaperone teams. In the following section, we will discuss PQC systems based on the coordination of protein folding and degradation. We will focus on the recent discoveries on cytosolic, nuclear, and membrane‐directed PQC; ER‐ and mitochondrion‐specific PQC regulation was previously described 66, 67, 68, 69, 70.

Figure 2.

Figure 2

E3 Ub ligases with different roles in PQC. (A) CHIP is a major PQC ligase of the cytosol. (left panel) In cooperation with Hsp70/90, CHIP ameliorates proteotoxicity in various proteinopathies by referring aggregates of beta‐amyloid, mutant SOD1, polyQ protein, or alpha‐synuclein for degradation. (right panel) In addition to misfolded proteins of the cytoplasm, CHIP promotes degradation of a broad array of substrates when bound to Hsps, such as CFTR, GR, ERBB2, or HIF1A. (B) (left panel) Rsp5/Nedd4 participates in the removal of cytosolic misfolded proteins. Upon heat stress Rsp5/Nedd4 associates with cochaperone Hsp40 (Ydj1), which supports recognition and degradation of misfolded proteins. (right panel) Beyond its role in the removal of cytosolic substrates, Rsp5/Nedd4 also targets misfolded proteins at the plasma membrane. ARTs enable Rsp5 to selectively target a wide range of plasma membrane proteins and initiate their endocytosis and lysosomal degradation. (C) The E3 ligase Listerin directly associates with the 60s ribosomal subunit to specifically target newly synthesized aberrant polypeptides expressing a translated polyA tail. Listerin collaborates with three cofactors for ribosomal binding and substrate processing: NEMF, TCF25, and the Ub‐selective chaperone p97. (D) The role of chaperone‐directed E3 ligases in nuclear PQC. (left panel) San1 cooperates with Hsp70 chaperones to recruit misfolded proteins from the cytosol for proteasomal degradation in the nucleus, whereas Doa10 targets substrates independent of Hsp70/Hsp40. In addition, both Doa10 and San1 interact with Cdc48/p97 to facilitate proteasomal degradation of a subset of their substrates. (right panel) The yeast E3 ligases Hel1, Hel2, Snt2 together with the histone chaperones Pep5 and Asf1 trigger ubiquitylation and subsequent proteasomal turnover of surplus histones.

Table 1.

List of quality control E3 ligases and their chaperone partners reviewed in this paper

E3 ligase Chaperone Target References
Cytosol
CHIP Hsp70; Hsp90 Misfolded proteins; Hsp90 clients 72, 73
Ubr1/UBR1 Hsp110 (Sse1); Hsp70 (SSa1/2); Hsp40 (Ydj1, Sis1) Misfolded proteins 101, 102, 103, 108, 114
Ubr2 Hsp110; Hsp70 Misfolded proteins 101, 103
Tom1/HUWE1 Hsp27; CDC48 Unassembled proteins 122, 123, 124
E6‐AP Hsp70/Hsc70 Misfolded, aggregated proteins 98
RNF126 BAG6 complex Mislocalized ER proteins 116, 117
Rsp5/NEDD4 Hsp40 Heat‐induced misfolded proteins; plasma membrane proteins 133, 143, 144, 145, 146, 147
Cullin5 Hsp90 Hsp90 clients 99, 100
Hul5/UBE3C Heat shock‐induced misfolded proteins [132, 134, 135, 136, 139, 140, 219]
SCF Hsp90‐Sgt1
Hsp70‐Sgt1
LRR domain proteins; kinetochore 125, 126, 127, 128, 129, 130, 131, 220
Nuclei
Pep5, Snt2, Hel1, Hel2 Asf1 Surplus histones 186
Rtt101Mms1/
Cul4ADDB1
Asf1
Asf1a/Asf1b
H3 192
Psh1 FACT CENP‐A (H3) 191
Doa10/MARCH6 Hsp70 (Ssa1/Ssa2), Hsp40 (Ydj1, Sis1) ER and INM 109, 177
Asi ligase complex (Asi1‐Asi3) Mislocalized proteins at INM 175, 176
Cytosol/Nuclei
San1 Hsp70; CDC‐48/p97 Nuclear misfolded proteins 101, 105, 107, 108, 221
Ribosomes
Ltn1p/Listerin Cdc48/p97 Aberrant nonstop polypeptides 148, 152, 161

CHIP

CHIP has been characterized as the first QC E3 Ub ligase important for the cellular proteostasis network. CHIP binds the molecular chaperones Hsp70/Hsp90 to coordinate the cellular balance between protein folding and degradation 71, 72, 73, 74, 75. It was first identified as a chaperone binding protein based on its N‐terminal tandem TPR 76. The evolutionarily conserved U‐box domain of CHIP, responsible for its E3 ligase activity, is a modified form of the more frequently found really interesting new gene (RING) domain 77, 78. Notably, the chaperone‐directed recruitment of CHIP involves a mutual allosteric interaction between the TPR and U‐box domains 79. Structural diversity and dynamics within CHIP are well described in the recent review by VanPelt and Page 80.

Various studies provide evidence that CHIP tightly regulates chaperone function. Upon acute stress, CHIP facilitates nuclear translocation and activation of HSF1 to protect against stress‐induced apoptosis 81, 82. CHIP also modulates the proteotoxic stress response by reducing the level of Hsc70/Hsp70 chaperone after heat shock 72, 80, 83, 84. Furthermore, CHIP modulates the activity of certain chaperones. For instance, this E3 ligase stimulates the release of the Hsp90 ATPase activity modulating cofactor p23 from the Hsp90 complex, thereby suppressing the affinity and folding activity of Hsp90, which results in Ub‐dependent degradation of substrate proteins 72, 85. On the other hand, CHIP competes for Hsp70 binding with Hsp40, which attenuates Hsp40 ATPase activity and suppresses protein folding by Hsp70 76.

Importantly, CHIP orchestrates regulation of cellular proteins from folding to degradation, including a coordinated degradation of substrates, which are beyond the refolding range. In addition to misfolded proteins of the cytoplasm, CHIP promotes degradation of a broad array of substrates when bound to Hsps, like cystic fibrosis transmembrane conductance regulator (CFTR), glucocorticoid receptor (GR), androgen receptor, estrogen receptor, erythroblastic leukemia viral oncogene homolog 2 (ERBB2), or protein arginine methyltransferase 5 (PRMT5) 86, 87, 88, 89, 90 (Fig. 2A). Hypothetically, all clients of HSP70 or HSP90 are potential targets of CHIP. However, not all CHIP substrates are recruited via interaction with Hsps. For example, Sma‐mother against decapentaplegic 1 (Smad1) level is regulated by CHIP, which subsequently influences bone morphogenetic proteins (BMP) signal transduction 91. Identified substrates of the CHIP/Hsp complex are detailed in the recent reviews by Paul and Gosh, and Joshi et al. 92, 93.

Aside from ameliorating proteotoxicity, CHIP plays a role in developmental regulation and aging. For instance, CHIP is involved in osteoblast differentiation by regulating the protein level of Runx2 94. In agreement with its role in protein quality control, CHIP knockout mice show reduced lifespan associated with age‐related pathophysiological defects 95. However, CHIP deletion mice exhibit normal embryonic development and unaffected turnover of many known CHIP substrates, suggesting functional redundancy among quality control Ub ligases 95, 96. In contrast, CHIP deficiency induces accelerated aging, which suggests the existence of at least one critical CHIP‐specific substrate that controls longevity. We have recently revealed an important function of CHIP‐mediated proteolysis in insulin/IGF‐like signaling (IIS). CHIP triggers degradation of the insulin receptor (INSR), which regulates metabolic changes and determines lifespan in metazoan organisms. Upon proteotoxic stress and during aging, CHIP preferentially functions in PQC, causing a stabilization of the INSR. Accordingly, proteotoxic accumulation of damaged proteins or aberrant CHIP function attenuates INSR degradation and affects metabolism and longevity through increased IIS 97. Our observation suggests an evolutionarily conserved coordination of proteostasis and aging regulated by CHIP‐assisted protein degradation.

Different strategies for targeting misfolded proteins in the cytosol

The cytosol of the eukaryotic cell melds protein synthesis, folding, and transport, all of which are continuously defining cellular proteostasis. The key insights into how E3 ligases and chaperones work together have been uncovered by research directed towards understanding the role of CHIP in degradation of misfolded proteins. As described above CHIP is the most extensively studied E3 Ub ligase associated with molecular chaperones 83, but not the only one maintaining proteostasis. PQC pathways deploy a variety of E3 ligases linked to various degradation routes to cope with the constant protein folding stress applied by physiological or stress‐related processes. Similarly to CHIP, E6‐AP—a HECT‐domain Ub ligase found in higher eukaryotes—interacts with Hsp70/Hsc70 chaperones and ubiquitylates their client proteins, such as aggregated proteins 98. Multisubunit Cullin‐based E3 ligases have also been implicated in PQC within the cytoplasm. The mammalian Cullin5–RING E3 Ub ligase interacts with the Hsp90 chaperone and mediates Ub‐dependent degradation of Hsp90 client proteins, including protein kinases (such as ERBB2) or transcription factors (like HIF1α) 99, 100.

Although CHIP is thought to be the major PQC E3 ligase in the cytosol of higher eukaryotes, budding yeast lacks this enzyme. Instead, Ubr1, Ubr2, and San1 are major PQC E3 ligases in Saccharomyces cerevisiae which ubiquitylate misfolded proteins to maintain proteostasis 101, 102, 103. Interestingly, these enzymes evolved two different strategies to safeguard the proteome. San1 has been first described as nuclear PQC E3 ligase with intrinsic capacity to bind aberrant proteins in the nucleus 104. As Rosenbaum and coworkers have shown, San1 can directly bind to its substrates through its disordered N‐terminal and C‐terminal domains, which provide conformational flexibility and serve as substrate recognition sites for misfolded proteins 105. In yeast, where proteasomal degradation capacity of the cell is highly concentrated in the nucleus 106, numerous cytosolic PQC substrates are degraded on San1‐dependent clearance mechanisms. Misfolded proteins in the cytosol are delivered to the nucleus by Hsp70 where San‐1‐driven ubiquitylation initiates proteasomal degradation 101, 107, 108. This suggests a major role of San1 in proteostasis by removing a wide range of cytoplasmic and nuclear misfolded proteins. So far no mammalian San1 homolog has been identified, but recent bioinformatic analysis suggests the existence of several mammalian E3 ligases that bear related, disordered regions and thus might function in a similar way 109.

The E3 ligases Ubr1 and Ubr2 have been first characterized in the N‐end rule pathway regulating degradation of short‐lived proteins presenting N‐terminal destabilizing amino acids 110, 111, 112, 113. The Ubr1/Ubr2‐dependent yeast PQC pathway operates in the cytosol, where Ubr1 employs the Hsp70 chaperones Ssa1/Ssa2, the Hsp40 cochaperone Ydj1 or Sis1, and the Hsp110 chaperone Sse1 for target recognition 101, 103, 114. Similar quality control function has been recently attributed to the mammalian UBR1 (N‐recognin 1) E3 ligase targeting HSP90 client proteins 115.

While the CHIP–Hsp70/Hsp90 complex directs the degradation of a multitude of different misfolded proteins, other ligase–chaperone complexes adopted more specialized strategies to bind target proteins, revealing a set of E3 ligases dedicated to distinct PQC pathways. RNF126 is an interesting example of a specialized PQC E3 ligase, which cooperates with the Bag6 chaperone 116. Eukaryotic cells have extensive endomembrane systems, hosting a significant portion of the cellular proteome. Utilizing specific signal sequences, the newly synthesized membrane proteins are rapidly integrated into the ER membrane. However, those that fail to target to the ER must be removed from the cytosol to avoid protein aggregation. Rodrigo‐Brenni et al. identified RNF126 as the key component of Bag6‐dependent degradation of mislocalized membrane proteins (MLPs) in the cytosol 117. The Bag6 chaperone preferentially binds to misfolded proteins with extensive hydrophobic domains 116, while the typical client proteins of Hsp70/Hsp90 characteristically expose shorter hydrophobic stretches 118, 119. This suggests that the Hsp70/Hsp90 chaperone system provides a different role in PQC compared to the Bag6‐driven pathway. RNF126 specifically ubiquitylates lysine residues located directly next to the hydrophobic segment of the MLPs 117. Interestingly, positively charged residues, such as lysines often flank chaperone‐recognized hydrophobic regions in membrane proteins 120, 121. This observation supports the idea that the RNF126/Bag6 complex implements a degree of specialization in substrate recognition and binding.

Quality control of multiprotein complexes

Proteins destined to work in multimeric protein complexes often expose unshielded segments of hydrophobic residues on their surface that mediate their assembly into higher order molecular machineries. Excess of free subunits of unassembled protein complexes either cause protein aggregation or interfere with their normal function. Using a model substrate to study the degradation of unassembled soluble polypeptides of multisubunit complexes HUWE1 has been identified as novel PQC Ub ligase 122. HUWE1 targets both cytosolic and nuclear subunits in human cells, providing a constant quality control and removal of incomplete protein assemblies. Similarly, the yeast HUWE1 homolog Tom1 facilitates ubiquitylation and proteasomal degradation of unassembled ribosomal proteins 123. Although HUWE1‐dependent degradation of unassembled proteins is linked to the Ub‐selective chaperone p97, the possible involvement of other chaperones engaged in recognizing and presenting misfolded HUWE1 targets is not addressed yet. A recent proteomic study identified HUWE1–Hsp27 interaction, which might link HUWE1 to chaperone‐dependent degradation of hydrophobic polypeptides 124.

As discussed above, chaperones are not only important for folding of proteins but also play vital roles in supporting accurate assembly of multiprotein complexes. The presence of available components in proper stoichiometric ratios is critical to facilitate the build‐up of functional protein complexes. As such, the Hsp90–Sgt1 chaperone has a critical role in the assembly of kinetochores, the multivalent microtubule binding sites in the cells 125. Sgt1 acts as an adaptor and cochaperone for Hsp90 and Hsp70 to connect to multiple client proteins during their folding and assembly into protein complexes 126. Sgt1 also links Hsp90 to the Skp1/Cullin/F‐box (SCF) E3 ligase via direct binding to Skp1, thereby regulating assembly and activity of the SCF complex 127, 128, 129. The client proteins of Sgt1 and the Skp1 component of the SCF ligase share similar sequence feature, the leucine‐rich repeat (LRR) domain that supports recognition and interaction with the cochaperone Sgt1 130. Recent studies also suggested that Sgt1 client proteins are often ubiquitylated by SCF to facilitate their removal, although this regulatory function needs further experimental evidence 125, 131.

Heat stress E3 ligases

Heat stress increases protein unfolding and acutely overloads the cell with misfolded proteins. It has been shown that heat shock primarily increases the ubiquitylation of cytosolic proteins 132. HECT ubiquitin ligase 5 (Hul5) and Rsp5 have been identified in yeast as major E3 ligases that regulate ubiquitylation and proteasomal degradation of heat‐induced misfolded proteins 132, 133. Although degradation of many unfolded yeast proteins depends on the Hsp70 (SSA1‐5) chaperones, Hul5 recognizes target proteins without the help of these chaperones. It is a critical challenge to discriminate between terminally or temporarily misfolded proteins and only target those for degradation that cannot be refolded and used anymore. Hul5‐dependent ubiquitylation of terminally misfolded proteins occurs when mono‐ubiquitylated proteins are not refolded for a longer time window. Hul5 directly associates with the 19S RP of the proteasome where it acts as an E4 enzyme elongating Ub chains on proteasome‐bound substrates 134, 135, 136, 137, 138. UBE3C, the human homolog of Hul5, is also a proteasome‐associated E3 ligase which further ubiquitylates proteins that are difficult to degrade thereby assisting proteasomal degradation 135, 139, 140. The potential involvement of the mammalian Hul5 homologs in heat stress‐induced PQC is not verified yet. In addition to Hul5, the yeast Rsp5 and its mammalian homolog Nedd4 have major roles in the removal of cytosolic misfolded proteins upon heat stress 133 (Fig. 2B). Overexpression of Rsp5 increases thermotolerance in yeast 141, which is in agreement with its important role in response to heat‐induced damage. In contrast to Hul5, Rsp5/Nedd4 uses a bipartite mechanism for recognition of its cytosolic misfolded substrates. Upon heat stress, Rsp5 associates with Hsp40 (Ydj1) cochaperone promoting ubiquitylation and degradation of misfolded proteins. On the other hand, Rsp5 can bind some of its targets directly. These substrates typically contain short stretches of amino acids, which are proline‐rich motifs (called the PY or PY‐like) that confer binding to the WW‐domains of Rsp5 133. These motifs act as degrons promoting heat stress‐induced substrate–Rsp5 interaction.

Degradation of plasma membrane proteins

In contrast to cytosolic or ER proteins, plasma membrane anchored or integral proteins are mainly degraded by the endolysosomal degradation pathway 142. Proteotoxic stress dramatically changes the landscape of membrane proteins as a result of the endocytic removal of damaged proteins. Beyond its role in removal of cytosolic heat‐induced misfolded proteins, Rsp5/Nedd4 has also been reported to target misfolded plasma membrane proteins for lysosomal degradation 143, 144, 145 (Fig. 2B). Furthermore, Rsp5 is also involved in a range of cargo‐sorting events within the endosomal and Golgi transport pathway. In yeast, arrestin‐related trafficking adaptors (ARTs) enable Rsp5 to specifically target a wide range of plasma membrane proteins and initiate their endocytosis and lysosomal degradation 144, 146, 147. Defects of this PQC pathway result in severe loss of plasma membrane integrity.

Ribosome‐associated quality control

Protein synthesis is a highly error‐prone process. In eukaryotic cells, a fraction of the newly synthesized proteins is immediately degraded by the 26S proteasome, indicating the existence of a strictly cotranslational PQC to ensure elimination of aberrant proteins 148. The ribosomal Ltn1/Rkr1 and Hel2 E3 ligases together with the nonribosomal Ub ligases Doa10, Hrd1 and Hul5 mediate ubiquitylation and proteasomal degradation of nascent proteins, which escaped cotranslational folding control of newly synthesized proteins directed by the ribosome‐bound nascent polypeptide‐associated complex (NAC) chaperone 148, 149, 150, 151. The recently described conserved PQC pathway requires the yeast Ltn1 Ub ligase, or its mammalian homolog Listerin, which directly associates with ribosomes to specifically target newly synthesized aberrant polypeptides expressing a translated polyA tail 152. Ltn1‐dependent polyubiquitylation and subsequent proteasomal degradation of nonstop proteins is triggered by stalling them at the translation machinery 153. Ltn1/Listerin utilizes three cofactors for binding to the ribosome and for processing the targets: Tae2/NEMF, Rqc1/TCF25, and Cdc48/p97 (Fig. 2C). Tae2 (NEMF) recognizes the stalled ribosomes and recruits Ltn1 to the 60S–peptidyl–tRNA complex, which together with Rqc1 enables binding of the Cdc48/p97 Ub‐selective chaperone 154, 155. Cdc48/p97 mediates segregation/unfolding of ubiquitylated substrates from the ribosomal complex and their proteasomal degradation 156, 157, 158. Tae2 (Rqc2) also recruits an enzyme that generates chloramphenicol acetyltransferase tail on aberrant nascent peptides, which is crucial for induction of translational folding stress response 159. When Cdc48 is not recruited, the Ltn1–Rcq1–Cdc48 PQC pathway fails to initiate the degradation of aberrant translation products arising from ribosomes. Consequently, the chloramphenicol acetyltransferase‐tailed ubiquitylated peptides localize to aggregates, which are specifically associated with Sis1, Sgt2, Ssa1/2, and Hsp82 chaperones 160. Hence, Ltn1 Ub ligase‐driven PQC of ribosomal translation is also essential in prevention of cytosolic aggregate formation 161. A recent study using a large set of model substrates in yeast revealed that besides its role in ribosomal PQC, Ltn1 also mediates degradation of substrates bearing different degradation signals (degrons) fused to their C terminus, suggesting a broader role for Ltn1 in cellular PQC 162.

E3–chaperone complexes in the nuclear protein quality control

Although it is spared from folding burden of nascent polypeptides, when it comes to the nucleus, the PQC pathways face unique folding problems. PQC of the nucleus should have rigorous control over the identity and folding of nuclear membrane proteins as well as chromatin‐associated proteins 163. The nuclear PQC is exceptionally important because failure to repair or remove misfolded nuclear proteins can lead to a deterioration of the nuclear genome and mRNA integrity. In addition, the nucleus is especially enriched in proteins possessing low complexity and intrinsically disordered regions 164. Compared to regulation in the cytosol, nuclear PQC is also governed by the cooperative action of HSPs, molecular chaperones, associated E3 ligases, and proteasomal degradation. In addition, increasing evidence suggests that nuclear envelope components are also degraded by autophagy 165, 166.

Exposed hydrophobic protein stretches are key determinants of nuclear quality control degradation pathways. San1 is a central PQC E3 ligase of the yeast nucleus, involved in ubiquitylation and proteasomal degradation of a wide range of misfolded nuclear and imported cytosolic proteins 104 (Fig. 2D). Recognition of misfolded proteins by San1 is triggered via surface exposure of a few contiguous hydrophobic residues 167. Lacking San1, mammalian cells employ other nuclear PQC E3 ligases, such as UHRF2, which associates with and ubiquitylates nuclear polyQ aggregates 168.

Asi protein ligase preserves the identity of the inner nuclear membrane

The double‐membrane‐based nuclear envelope has crucial function in providing compartmentalization for the genomic DNA. As the outer nuclear membrane (ONM) is contiguous with the ER, the quality control of proteins localized in this membrane layer is generally performed by the ER‐associated PQC systems. As such, E3 ligase complexes based on Hrd1 and Doa10 drive ubiquitylation and proteasomal degradation of the majority of misfolded ER proteins through the yeast ERAD pathway 68, 169, 170, 171. The protein content of the inner nuclear membrane (INM) is distinct from that of the outer layer and it is thought that nuclear quality control mechanisms are in charge to maintain its integrity. The INM is connected to the outer membrane–ER system at the nuclear pores which, in addition to restricting protein exchange between the cytosol and nuclei, also regulates protein transport from the ER membrane to the INM 172. In yeast, the Ub ligase Asi complex consisting of Asi1, Asi2, and As3 is involved in the process that controls promoter access of two transcription factors Spt1 and Spt2 173, 174. Recently, Foresti et al. and Khmelinskii et al. found that the INM‐localized Asi E3 ligases regulate degradation of mislocalized proteins that are not destined to INM, defining a novel PQC pathway of the eukaryotic cell that maintains and safeguards the identity of the INM 175, 176. Notably, ER membrane bound Doa10 has been linked to ubiquitylation of soluble and INM‐associated nuclear proteins, by recognizing hydrophobic patches of proteins exposed to the nucleoplasm 177. Doa10 targets proteins in an Hsp70/Hsp40‐dependent manner 178, and teams up with Cdc48/p97 for proteasomal targeting of a subset of its substrates 179, 180.

Aggregation‐prone proteins, such as the pathogenic polyQ‐exposing proteins, represent another major threat for the nucleus. Guo and coworkers recently described a dedicated nuclear team responsible for recognition and removal of polyQ aggregates 181. This interesting mechanism is based on the promyelocytic leukemia protein (PML) that selectively recognizes and interacts with the nuclear, misfolded polyQ proteins and sumoylates them. In turn, the Ub ligase RNF4 attaches polyubiquitin chains to the aggregates, which targets them for proteasomal degradation. The role of E3 ligase‐associated chaperones in nuclear protein quality control is not as well established as in the ER or cytoplasmic PQC. While San1 might use Hsp70 chaperones for delivering cytosolic misfolded proteins for nuclear degradation by the proteasome 107, chaperone partners of the Asi complex have not been described yet.

Histone chaperone–E3 complexes safeguard genome stability

In eukaryotic cells, the genomic DNA is packed by histones, building a compact chromatin structure. Histone complexes act as spools as DNA winds up around them to form the basic structural elements of the chromatin, the nucleosomes. Histone proteins dynamically regulate chromatin structure to adapt its activity status to the cellular demand 182. Chromatin‐bound histones appear to be very stable 183; however, nonchromatin‐bound histones are rapidly degraded with a short half‐life 184. Degradation of excess histones is essential because they interfere with cellular viability and promote toxic effect leading to genomic instability 185. The yeast E3 ligases Tom1, Pep5, Snt2, Hel1, and Hel2 are involved in ubiquitylation and subsequent proteasomal degradation of surplus histones, where Hel1, Hel2, Snt2, and Pep5 works together with the Asf1 histone chaperone 186 (Fig. 2D). Other histone chaperones such as FACT, NAP1, HIRA, and DAXX are involved in histone shuttling between the cytoplasm and the nuclei, as well as histone deposition into and extraction from chromatin 187, 188, 189, 190. Notably, the histone chaperone FACT cooperates with Pob3/Spt16/histone 1 (Psh1) E3 ligase that targets ectopically localized Histone3 (H3) variant CENP‐A (Cse4) for degradation, to maintain centromere identity, and to support proper chromatin segregation, and genomic stability 191.

Recently, Han and coworkers described a novel player in nucleosome maintenance, the mammalian E3 Ub ligase Cul4ADDB1 and its yeast homolog Rtt101Mms1 192. Nucleosomes are dynamically formed and disassembled in order to allow the gene transcription and DNA replication machinery to access distinct regions of the genomic DNA on a temporally regulated fashion 182. During nucleosome assembly, the Cul4ADDB1 (Rtt101Mms1) E3 ligase preferentially binds to and ubiquitylates newly synthesized Lys56‐acetylated H3. This promotes H3 dissociation from the Asf1a or Asf1b (Asf1) chaperone and facilitates its binding to downstream processing histone chaperones, such as the mammalian CAF‐1 for H3.1 or Daxx and HIRA for H3.3, to support nucleosome formation. The Cul4 E3 ligase‐driven histone hand‐off between chaperones does not lead to histone degradation, but it stands as an interesting example of the nonproteolytic E3 ligase–chaperone role 192.

Conclusions and perspectives

Selective degradation of misfolded or aggregated proteins is crucial for maintaining functionality of the cell. PQC mechanisms are present at all steps of a protein's lifetime and specialized enzyme complexes safeguard distinct steps of proteostasis processes. Molecular chaperones are vital in the protein quality assurance pathways recognizing the non‐native proteins and preventing their interference with the cellular functions. To deliver damaged proteins to the cellular degradation pathways, dedicated E3 ligases cooperate with a variety of chaperones (Table 1). Environmental threats, endogenous stress, and aging constantly challenge the cellular proteome, and ultimately affect organismal viability. As we described above, eukaryotic cells adopted various mechanisms to cope with proteotoxic insults, which put pressure on their limited protein folding capacity. The implications of the eukaryotic proteostasis pathways in human disease are far reaching, as failure of any components of the PQC pathways could lead to disease 7.

It is commonly thought that an age‐related impairment of protein degradation affects general proteostasis networks, causing enhanced accumulation of damaged proteins that can be cytotoxic and shortens lifespan 28, 193, 194, 195, 196, 197, 198, 199. During aging, the cellular proteostasis network shows significant changes in expression, mainly causing an overall reduction in protein synthesis, which reflects age‐dependent remodeling of an imbalanced proteome 200. Progressive decline of proteostasis can lead to the development of various diseases 201. An apparent consequence of PQC downregulation is the appearance of various forms of neurodegeneration. The formation of protein aggregates is universally observed in about 30 different human diseases 197, 202, 203, 204. Accordingly, the age‐dependent deposition of protein aggregates linked to disturbed proteasomal degradation of misfolded proteins is a major hallmark of neurodegenerative proteinopathies such as AD, HD, or Parkinsons's disease 11, 205. Along with molecular chaperones, PQC E3 Ub ligases associate with dysfunctional proteins in different neurodegenerative disorders. CHIP for example marks alpha‐synuclein in PD 206, beta‐amyloid in AD 207, phosphorylated tau, mutant SOD1 aggregates in amyotrophic lateral sclerosis (ALS) 75, 208, 209, 210, and polyQ aggregates in polyQ diseases 211, 212, for their proteasomal degradation. Therefore, failure of chaperone‐assisted degradation of these misfolded proteins might aggravate various neurodegenerative disorders. Chaperone upregulation is widely observed in different cancer types providing a cell with increased protein folding capacity 213, 214, 215. Such deregulation in chaperone level may also lead to impaired or to excessive recruitment of E3 ligases that can significantly change the selectivity and pace of substrate degradation. Hsp90 client proteins, including various kinases, have been implicated in malignant transformation 52. Therefore, Hsp inhibition has emerged as a central strategy in cancer treatment 216, 217.

Although special E3 ligase–chaperone partners involved in different PQC pathways have been identified, it is still far from being understood how much these pathways overlap in target recognition and processing, and how substrate selectivity is driven by the chaperones or the E3 ligases involved. It is especially interesting regarding the diverse cytosolic PQC pathways, where numerous E3–chaperone complexes work in parallel. For instance, interaction of chaperones with cochaperones could provide another layer of regulating the activity of chaperone–E3 ligase complexes. While E3 ligases are continuously being identified, the role of DUBs counteracting the activity of PQC E3 ligases is fairly unknown. Although their role in fine tuning the ubiquitylation processes is well established, only few DUBs have been assigned to PQC pathways so far 218. It would be also interesting to examine the potential consequences of disturbed proteostasis on the function of PQC E3 ligases and determine how the imbalance in protein folding alters substrate processing. Recent identification of the human chaperone network supports the idea that tight temporal and spatial regulation of the activity and abundance of chaperone groups, and potentially of their cofactors, armors the cell against specific challenges during aging. Therefore, future characterization of tissue‐, age‐, or disease‐specific chaperomes might reveal important mechanistic insights into how E3 ligases team up with chaperones to safeguard the proteome especially in multicellular organisms. Future research on tissue‐specific PQC pathways and their role in tissue functionality will enable therapeutic intervention strategies against age‐related protein aggregation diseases.

Disclosures

No conflicts of interest, financial or otherwise, are declared by the authors.

Acknowledgements

We thank members of our laboratory for crucial discussion and helpful advice on the manuscript. This work is supported by the European Research Council (consolidator grant 616499) to TH. In addition, this article is based upon work from COST Action (PROTEOSTASIS BM1307). We apologize for not having cited valuable contributions due to size limitation.

Edited by Wilhelm Just

References

  • 1. Douglas PM and Dillin A (2010) Protein homeostasis and aging in neurodegeneration. J Cell Biol 190, 719–729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Morimoto RI and Cuervo A (2014) Proteostasis and the aging proteome in health and disease. J Gerontol A Biol Sci Med Sci 69, S33–S38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Ellis RJ and Minton AP (2006) Protein aggregation in crowded environments. Biol Chem 387, 485–497. [DOI] [PubMed] [Google Scholar]
  • 4. Dunker AK, Silman I, Uversky VN and Sussman JL (2008) Function and structure of inherently disordered proteins. Curr Opin Struct Biol 18, 756–764. [DOI] [PubMed] [Google Scholar]
  • 5. Ciryam P, Kundra R, Morimoto RI, Dobson CM and Vendruscolo M (2015) Supersaturation is a major driving force for protein aggregation in neurodegenerative diseases. Trends Pharmacol Sci 36, 72–77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Hartl FU, Bracher A and Hayer‐Hartl M (2011) Molecular chaperones in protein folding and proteostasis. Nature 475, 324–332. [DOI] [PubMed] [Google Scholar]
  • 7. Labbadia J and Morimoto RI (2015) The biology of proteostasis in aging and disease. Annu Rev Biochem 84, 435–464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Douglas PM and Cyr DM (2010) Interplay between protein homeostasis networks in protein aggregation and proteotoxicity. Biopolymers 93, 229–236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Morimoto RI (2008) Proteotoxic stress and inducible chaperone networks in neurodegenerative disease and aging. Genes Dev 22, 1427–1438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Powers ET, Morimoto RI, Dillin A, Kelly JW and Balch WE (2009) Biological and chemical approaches to diseases of proteostasis deficiency. Annu Rev Biochem 78, 959–991. [DOI] [PubMed] [Google Scholar]
  • 11. Saez I and Vilchez D (2014) The mechanistic links between proteasome activity, aging and agerelated diseases. Curr Genomics 15, 38–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Lopez‐Otin C, Blasco MA, Partridge L, Serrano M and Kroemer G (2013) The hallmarks of aging. Cell 153, 1194–1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Balch WE, Morimoto RI, Dillin A and Kelly JW (2008) Adapting proteostasis for disease intervention. Science 319, 916–919. [DOI] [PubMed] [Google Scholar]
  • 14. Kenyon CJ (2010) The genetics of ageing. Nature 464, 504–512. [DOI] [PubMed] [Google Scholar]
  • 15. Kuhlbrodt K, Mouysset J and Hoppe T (2005) Orchestra for assembly and fate of polyubiquitin chains. Essays Biochem 41, 1–14. [DOI] [PubMed] [Google Scholar]
  • 16. Kerscher O, Felberbaum R and Hochstrasser M (2006) Modification of proteins by ubiquitin and ubiquitin‐like proteins. Annu Rev Cell Dev Biol 22, 159–180. [DOI] [PubMed] [Google Scholar]
  • 17. Komander D and Rape M (2012) The ubiquitin code. Annu Rev Biochem 81, 203–229. [DOI] [PubMed] [Google Scholar]
  • 18. Lee BH, Lee MJ, Park S, Oh DC, Elsasser S, Chen PC, Gartner C, Dimova N, Hanna J, Gygi SP et al (2010) Enhancement of proteasome activity by a small‐molecule inhibitor of USP14. Nature 467, 179–184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Kuhlbrodt K, Janiesch PC, Kevei E, Segref A, Barikbin R and Hoppe T (2011) The Machado‐Joseph disease deubiquitylase ATX‐3 couples longevity and proteostasis. Nat Cell Biol 13, 273–281. [DOI] [PubMed] [Google Scholar]
  • 20. Kim HT, Kim KP, Uchiki T, Gygi SP and Goldberg AL (2009) S5a promotes protein degradation by blocking synthesis of nondegradable forked ubiquitin chains. EMBO J 28, 1867–1877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Sowa ME, Bennett EJ, Gygi SP and Harper JW (2009) Defining the human deubiquitinating enzyme interaction landscape. Cell 138, 389–403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Amerik A, Swaminathan S, Krantz BA, Wilkinson KD and Hochstrasser M (1997) In vivo disassembly of free polyubiquitin chains by yeast Ubp14 modulates rates of protein degradation by the proteasome. EMBO J 16, 4826–4838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Finley D (2009) Recognition and processing of ubiquitin‐protein conjugates by the proteasome. Annu Rev Biochem 78, 477–513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Xu P, Duong DM, Seyfried NT, Cheng D, Xie Y, Robert J, Rush J, Hochstrasser M, Finley D and Peng J (2009) Quantitative proteomics reveals the function of unconventional ubiquitin chains in proteasomal degradation. Cell 137, 133–145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Behrends C, Sowa ME, Gygi SP and Harper JW (2010) Network organization of the human autophagy system. Nature 466, 68–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Xie Z and Klionsky DJ (2007) Autophagosome formation: core machinery and adaptations. Nat Cell Biol 9, 1102–1109. [DOI] [PubMed] [Google Scholar]
  • 27. Jin M, Liu X and Klionsky DJ (2013) SnapShot: selective autophagy. Cell 152, 368e1–368.e2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Salminen A and Kaarniranta K (2009) Regulation of the aging process by autophagy. Trends Mol Med 15, 217–224. [DOI] [PubMed] [Google Scholar]
  • 29. Levine B and Kroemer G (2008) Autophagy in the pathogenesis of disease. Cell 132, 27–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Mizushima N, Levine B, Cuervo AM and Klionsky DJ (2008) Autophagy fights disease through cellular self‐digestion. Nature 451, 1069–1075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Kraft C, Peter M and Hofmann K (2010) Selective autophagy: ubiquitin‐mediated recognition and beyond. Nat Cell Biol 12, 836–841. [DOI] [PubMed] [Google Scholar]
  • 32. Cuervo AM and Wong E (2014) Chaperone‐mediated autophagy: roles in disease and aging. Cell Res 24, 92–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Ferreira JV, Fôfo H, Bejarano E, Bento CF, Ramalho JS, Girão H and Pereira P (2013) STUB1/CHIP is required for HIF1A degradation by chaperone‐mediated autophagy. Autophagy 9, 1349–1366. [DOI] [PubMed] [Google Scholar]
  • 34. Kettern N, Dreiseidler M, Tawo R and Hohfeld J (2010) Chaperone‐assisted degradation: multiple paths to destruction. Biol Chem 391, 481–489. [DOI] [PubMed] [Google Scholar]
  • 35. Arndt V, Dick N, Tawo R, Dreiseidler M, Wenzel D, Hesse M, Fürst DO, Saftig P, Saint R, Fleischmann BK et al (2010) Chaperone‐assisted selective autophagy is essential for muscle maintenance. Curr Biol 20, 143–148. [DOI] [PubMed] [Google Scholar]
  • 36. Gamerdinger M, Hajieva P, Kaya AM, Wolfrum U, Hartl FU and Behl C (2009) Protein quality control during aging involves recruitment of the macroautophagy pathway by BAG3. EMBO J 28, 889–901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Carra S, Seguin SJ, Lambert H and Landry J (2007) HspB8 chaperone activity toward poly(Q)‐containing proteins depends on its association with Bag3, a stimulator of macroautophagy. J Biol Chem 283, 1437–1444. [DOI] [PubMed] [Google Scholar]
  • 38. Crippa V, Sau D, Rusmini P, Boncoraglio A, Onesto E, Bolzoni E, Galbiati M, Fontana E, Marino M, Carra S et al (2010) The small heat shock protein B8 (HspB8) promotes autophagic removal of misfolded proteins involved in amyotrophic lateral sclerosis (ALS). Hum Mol Genet 19, 3440–3456. [DOI] [PubMed] [Google Scholar]
  • 39. Ulbricht A, Eppler Felix J, Tapia Victor E, van der Ven Peter FM, Hampe N, Hersch N, Vakeel P, Stadel D, Haas A, Saftig P et al (2013) Cellular mechanotransduction relies on tension‐induced and chaperone‐assisted autophagy. Curr Biol 23, 430–435. [DOI] [PubMed] [Google Scholar]
  • 40. Powers ET and Balch WE (2013) Diversity in the origins of proteostasis networks–a driver for protein function in evolution. Nat Rev Mol Cell Biol 14, 237–248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Brehme M, Voisine C, Rolland T, Wachi S, Soper JH, Zhu Y, Orton K, Villella A, Garza D, Vidal M et al (2014) A chaperome subnetwork safeguards proteostasis in aging and neurodegenerative disease. Cell Rep 9, 1135–1150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Saibil H (2013) Chaperone machines for protein folding, unfolding and disaggregation. Nat Rev Mol Cell Biol 14, 630–642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Park C and Cuervo AM (2013) Selective autophagy: talking with the UPS. Cell Biochem Biophys 67, 3–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Kaushik S and Cuervo AM (2012) Chaperones in autophagy. Pharmacol Res 66, 484–493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Buchner J (1996) Supervising the fold: functional principles of molecular chaperones. FASEB J 10, 10–19. [PubMed] [Google Scholar]
  • 46. Zuehlke A and Johnson JL (2010) Hsp90 and co‐chaperones twist the functions of diverse client proteins. Biopolymers 93, 211–217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Tzankov S, Wong MJH, Shi K, Nassif C and Young JC (2008) Functional divergence between co‐chaperones of Hsc70. J Biol Chem 283, 27100–27109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Nillegoda NB, Kirstein J, Szlachcic A, Berynskyy M, Stank A, Stengel F, Arnsburg K, Gao X, Scior A, Aebersold R et al (2015) Crucial HSP70 co‐chaperone complex unlocks metazoan protein disaggregation. Nature 524, 247–251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Nillegoda NB, Stank A, Malinverni D, Alberts N, Szlachcic A, Barducci A, De Rios Los P, Wade RC and Bukau B (2017) Evolution of an intricate J‐protein network driving protein disaggregation in eukaryotes. Elife 6, e24560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Craig EA and Marszalek J (2017) How do J‐proteins get Hsp70 to do so many different things? Trends Biochem Sci 42, 355–368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Zuiderweg ER, Hightower LE and Gestwicki JE (2017) The remarkable multivalency of the Hsp70 chaperones. Cell Stress Chaperones 22, 173–189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. Schopf FH, Biebl MM and Buchner J (2017) The HSP90 chaperone machinery. Nat Rev Mol Cell Biol 18, 345–360. [DOI] [PubMed] [Google Scholar]
  • 53. Zhao R, Davey M, Hsu YC, Kaplanek P, Tong A, Parsons AB, Krogan N, Cagney G, Mai D, Greenblatt J et al (2005) Navigating the chaperone network: an integrative map of physical and genetic interactions mediated by the hsp90 chaperone. Cell 120, 715–727. [DOI] [PubMed] [Google Scholar]
  • 54. Hemmingsen SM, Woolford C, van der Vies SM, Tilly K, Dennis DT, Georgopoulos CP, Hendrix RW and Ellis RJ (1988) Homologous plant and bacterial proteins chaperone oligomeric protein assembly. Nature 333, 330–334. [DOI] [PubMed] [Google Scholar]
  • 55. Ranford JC, Coates AR and Henderson B (2000) Chaperonins are cell‐signalling proteins: the unfolding biology of molecular chaperones. Expert Rev Mol Med 2, 1–17. [DOI] [PubMed] [Google Scholar]
  • 56. Lopez T, Dalton K and Frydman J (2015) The mechanism and function of group II chaperonins. J Mol Biol 427, 2919–2930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Kampinga HH and Craig EA (2010) The HSP70 chaperone machinery: J proteins as drivers of functional specificity. Nat Rev Mol Cell Biol 11, 579–592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. Bar‐Lavan Y, Shemesh N and Ben‐Zvi A (2016) Chaperone families and interactions in metazoa. Essays Biochem 60, 237–253. [DOI] [PubMed] [Google Scholar]
  • 59. Takayama S and Reed JC (2001) Molecular chaperone targeting and regulation by BAG family proteins. Nat Cell Biol 3, E237–E241. [DOI] [PubMed] [Google Scholar]
  • 60. Morano KA and Thiele D (1999) Heat shock factor function and regulation in response to cellular stress, growth, and differentiation signals. Gene Expr 7, 271–282. [PMC free article] [PubMed] [Google Scholar]
  • 61. Anckar J and Sistonen L (2007) Heat shock factor 1 as a coordinator of stress and developmental pathways In Advances in Experimental Medicine and Biology (Csermely P. and Vígh L, eds.), pp. 78–88. Springer, New York. [DOI] [PubMed] [Google Scholar]
  • 62. Ciocca DR and Calderwood SK (2005) Heat shock proteins in cancer: diagnostic, prognostic, predictive, and treatment implications. Cell Stress Chaperones 10, 86–103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63. Calderwood SK and Gong J (2016) Heat shock proteins promote cancer: it's a protection racket. Trends Biochem Sci 41, 311–323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Rodina A, Wang T, Yan P, Gomes ED, Dunphy MP, Pillarsetty N, Koren J, Gerecitano JF, Taldone T, Zong H et al (2016) The epichaperome is an integrated chaperome network that facilitates tumour survival. Nature 538, 397–401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Taipale M, Tucker G, Peng J, Krykbaeva I, Lin ZY, Larsen B, Choi H, Berger B, Gingras AC and Lindquist S (2014) A quantitative chaperone interaction network reveals the architecture of cellular protein homeostasis pathways. Cell 158, 434–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66. Mehnert M, Sommer T and Jarosch E (2010) ERAD ubiquitin ligases: multifunctional tools for protein quality control and waste disposal in the endoplasmic reticulum. BioEssays 32, 905–913. [DOI] [PubMed] [Google Scholar]
  • 67. Christianson JC and Ye Y (2014) Cleaning up in the endoplasmic reticulum: ubiquitin in charge. Nat Struct Mol Biol 21, 325–335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68. Ruggiano A, Foresti O and Carvalho P (2014) Quality control: ER‐associated degradation: protein quality control and beyond. J Cell Biol 204, 869–879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69. Roberts RF, Tang MY, Fon EA and Durcan TM (2016) Defending the mitochondria: the pathways of mitophagy and mitochondrial‐derived vesicles. Int J Biochem Cell Biol 79, 427–436. [DOI] [PubMed] [Google Scholar]
  • 70. Rüb C, Wilkening A and Voos W (2016) Mitochondrial quality control by the Pink1/Parkin system. Cell Tissue Res 367, 111–123. [DOI] [PubMed] [Google Scholar]
  • 71. Arndt V, Rogon C and Höhfeld J (2007) To be, or not to be — molecular chaperones in protein degradation. Cell Mol Life Sci 64, 2525–2541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Connell P, Ballinger CA, Jiang J, Wu Y, Thompson LJ, Hohfeld J and Patterson C (2001) The co‐chaperone CHIP regulates protein triage decisions mediated by heat‐shock proteins. Nat Cell Biol 3, 93–96. [DOI] [PubMed] [Google Scholar]
  • 73. Cyr DM, Hohfeld J and Patterson C (2002) Protein quality control: U‐box‐containing E3 ubiquitin ligases join the fold. Trends Biochem Sci 27, 368–375. [DOI] [PubMed] [Google Scholar]
  • 74. Murata S, Chiba T and Tanaka K (2003) CHIP: a quality‐control E3 ligase collaborating with molecular chaperones. Int J Biochem Cell Biol 35, 572–578. [DOI] [PubMed] [Google Scholar]
  • 75. Petrucelli L (2004) CHIP and Hsp70 regulate tau ubiquitination, degradation and aggregation. Hum Mol Genet 13, 703–714. [DOI] [PubMed] [Google Scholar]
  • 76. Ballinger CA, Connell P, Wu Y, Hu Z, Thompson LJ, Yin L‐Y and Patterson C (1999) Identification of CHIP, a novel tetratricopeptide repeat‐containing protein that interacts with heat shock proteins and negatively regulates chaperone functions. Mol Cell Biol 19, 4535–4545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Hatakeyama S, Yada M, Matsumoto M, Ishida N and Nakayama KI (2001) U box proteins as a new family of ubiquitin‐protein ligases. J Biol Chem 276, 33111–33120. [DOI] [PubMed] [Google Scholar]
  • 78. Jiang J (2001) CHIP Is a U‐box‐dependent E3 Ubiquitin Ligase. Identification of Hsc70 as a target for ubiquitylation. J Biol Chem 276, 42938–42944. [DOI] [PubMed] [Google Scholar]
  • 79. Matsumura Y, Sakai J and Skach WR (2013) Endoplasmic reticulum protein quality control is determined by cooperative interactions between Hsp/c70 protein and the CHIP E3 ligase. J Biol Chem 288, 31069–31079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80. VanPelt J and Page RC (2017) Unraveling the CHIP:Hsp70 complex as an information processor for protein quality control. Biochim Biophys Acta 1865, 133–141. [DOI] [PubMed] [Google Scholar]
  • 81. Dai Q (2003) CHIP activates HSF1 and confers protection against apoptosis and cellular stress. EMBO J 22, 5446–5458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82. Yu Y, Liu M, Zhang L, Cao Q, Zhang P, Jiang H, Zou Y and Ge J (2012) Heat shock transcription factor 1 inhibits H2O2‐induced cardiomyocyte death through suppression of high‐mobility group box 1. Mol Cell Biochem 364, 263–269. [DOI] [PubMed] [Google Scholar]
  • 83. Murata S, Minami Y, Minami M, Chiba T and Tanaka K (2001) CHIP is a chaperone‐dependent E3 ligase that ubiquitylates unfolded protein. EMBO Rep 2, 1133–1138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84. Soss SE, Rose KL, Hill S, Jouan S and Chazin WJ (2015) Biochemical and proteomic analysis of ubiquitination of Hsc70 and Hsp70 by the E3 ligase CHIP. PLoS ONE 10, e0128240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. McDonough H and Patterson C (2003) CHIP: a link between the chaperone and proteasome systems. Cell Stress Chaperones 8, 303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Meacham GC, Patterson C, Zhang W, Younger JM and Cyr DM (2001) The Hsc70 co‐chaperone CHIP targets immature CFTR for proteasomal degradation. Nat Cell Biol 3, 100–105. [DOI] [PubMed] [Google Scholar]
  • 87. Rees I, Lee S, Kim H and Tsai FTF (2006) The E3 ubiquitin ligase CHIP binds the androgen receptor in a phosphorylation‐dependent manner. Biochim Biophys Acta 1764, 1073–1079. [DOI] [PubMed] [Google Scholar]
  • 88. Fan M, Park A and Nephew KP (2005) CHIP (carboxyl terminus of Hsc70‐interacting protein) promotes basal and geldanamycin‐induced degradation of estrogen receptor‐α. Mol Endocrinol 19, 2901–2914. [DOI] [PubMed] [Google Scholar]
  • 89. Zhang H‐T, Zeng L‐F, He Q‐Y, Tao WA, Zha Z‐G and Hu C‐D (2016) The E3 ubiquitin ligase CHIP mediates ubiquitination and proteasomal degradation of PRMT5. Biochim Biophys Acta 1863, 335–346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Zhou P, Fernandes N, Dodge IL, Reddi AL, Rao N, Safran H, DiPetrillo TA, Wazer DE, Band V and Band H (2003) ErbB2 degradation mediated by the co‐chaperone protein CHIP. J Biol Chem 278, 13829–13837. [DOI] [PubMed] [Google Scholar]
  • 91. Li R‐F, Shang Y, Liu D, Ren Z‐S, Chang Z and Sui S‐F (2007) Differential ubiquitination of smad1 mediated by CHIP: implications in the regulation of the bone morphogenetic protein signaling pathway. J Mol Biol 374, 777–790. [DOI] [PubMed] [Google Scholar]
  • 92. Paul I and Ghosh MK (2015) A CHIPotle in physiology and disease. Int J Biochem Cell Biol 58, 37–52. [DOI] [PubMed] [Google Scholar]
  • 93. Joshi V, Amanullah A, Upadhyay A, Mishra R, Kumar A and Mishra A (2016) A decade of boon or burden: what has the chip ever done for cellular protein quality control mechanism implicated in neurodegeneration and aging? Front Mol Neurosci 9, 93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94. Li X, Huang M, Zheng H, Wang Y, Ren F, Shang Y, Zhai Y, Irwin DM, Shi Y, Chen D et al (2008) CHIP promotes Runx2 degradation and negatively regulates osteoblast differentiation. J Cell Biol 181, 959–972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Min JN, Whaley RA, Sharpless NE, Lockyer P, Portbury AL and Patterson C (2008) CHIP deficiency decreases longevity, with accelerated aging phenotypes accompanied by altered protein quality control. Mol Cell Biol 28, 4018–4025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Morishima Y, Wang AM, Yu Z, Pratt WB, Osawa Y and Lieberman AP (2008) CHIP deletion reveals functional redundancy of E3 ligases in promoting degradation of both signaling proteins and expanded glutamine proteins. Hum Mol Genet 17, 3942–3952. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Tawo R, Pokrzywa W, Kevei É, Akyuz ME, Balaji V, Adrian S, Höhfeld J and Hoppe T (2017) The ubiquitin ligase CHIP integrates proteostasis and aging by regulation of insulin receptor turnover. Cell 169, 470–482.e13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Mishra A, Godavarthi SK, Maheshwari M, Goswami A and Jana NR (2009) The ubiquitin ligase E6‐AP is induced and recruited to aggresomes in response to proteasome inhibition and may be involved in the ubiquitination of Hsp70‐bound misfolded proteins. J Biol Chem 284, 10537–10545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Ehrlich ES, Wang T, Luo K, Xiao Z, Niewiadomska AM, Martinez T, Xu W, Neckers L and Yu XF (2009) Regulation of Hsp90 client proteins by a Cullin5‐RING E3 ubiquitin ligase. Proc Natl Acad Sci USA 106, 20330–20335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Samant RS, Clarke PA and Workman P (2014) E3 ubiquitin ligase Cullin‐5 modulates multiple molecular and cellular responses to heat shock protein 90 inhibition in human cancer cells. Proc Natl Acad Sci USA 111, 6834–6839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Heck JW, Cheung SK and Hampton RY (2010) Cytoplasmic protein quality control degradation mediated by parallel actions of the E3 ubiquitin ligases Ubr1 and San1. Proc Natl Acad Sci USA 107, 1106–1111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102. Eisele F and Wolf DH (2008) Degradation of misfolded protein in the cytoplasm is mediated by the ubiquitin ligase Ubr1. FEBS Lett 582, 4143–4146. [DOI] [PubMed] [Google Scholar]
  • 103. Nillegoda NB, Theodoraki MA, Mandal AK, Mayo KJ, Ren HY, Sultana R, Wu K, Johnson J, Cyr DM and Caplan AJ (2010) Ubr1 and Ubr2 function in a quality control pathway for degradation of unfolded cytosolic proteins. Mol Biol Cell 21, 2102–2116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104. Gardner RG, Nelson ZW and Gottschling DE (2005) Degradation‐mediated protein quality control in the nucleus. Cell 120, 803–815. [DOI] [PubMed] [Google Scholar]
  • 105. Rosenbaum JC, Fredrickson EK, Oeser ML, Garrett‐Engele CM, Locke MN, Richardson LA, Nelson ZW, Hetrick ED, Milac TI, Gottschling DE et al (2011) Disorder targets misorder in nuclear quality control degradation: a disordered ubiquitin ligase directly recognizes its misfolded substrates. Mol Cell 41, 93–106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106. Russell SJ, Steger KA and Johnston SA (1999) Subcellular localization, stoichiometry, and protein levels of 26 S proteasome subunits in yeast. J Biol Chem 274, 21943–21952. [DOI] [PubMed] [Google Scholar]
  • 107. Guerriero CJ, Weiberth KF and Brodsky JL (2013) Hsp70 targets a cytoplasmic quality control substrate to the San1p ubiquitin ligase. J Biol Chem 288, 18506–18520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Prasad R, Kawaguchi S and Ng DT (2010) A nucleus‐based quality control mechanism for cytosolic proteins. Mol Biol Cell 21, 2117–2127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109. Boomsma W, Nielsen SV, Lindorff‐Larsen K, Hartmann‐Petersen R and Ellgaard L (2016) Bioinformatics analysis identifies several intrinsically disordered human E3 ubiquitin‐protein ligases. PeerJ 4, e1725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Bartel B, Wunning I and Varshavsky A (1990) The recognition component of the N‐end rule pathway. EMBO J 9, 3179–3189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Kwon YT, Xia Z, An JY, Tasaki T, Davydov IV, Seo JW, Sheng J, Xie Y and Varshavsky A (2003) Female lethality and apoptosis of spermatocytes in mice lacking the UBR2 ubiquitin ligase of the N‐end rule pathway. Mol Cell Biol 23, 8255–8271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Tasaki T, Mulder LCF, Iwamatsu A, Lee MJ, Davydov IV, Varshavsky A, Muesing M and Kwon YT (2005) A family of mammalian E3 ubiquitin ligases that contain the UBR box motif and recognize N‐degrons. Mol Cell Biol 25, 7120–7136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Tasaki T and Kwon YT (2007) The mammalian N‐end rule pathway: new insights into its components and physiological roles. Trends Biochem Sci 32, 520–528. [DOI] [PubMed] [Google Scholar]
  • 114. Summers DW, Wolfe KJ, Ren HY and Cyr DM (2013) The type II Hsp40 Sis1 cooperates with Hsp70 and the E3 ligase Ubr1 to promote degradation of terminally misfolded cytosolic protein. PLoS ONE 8, e52099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115. Sultana R, Theodoraki MA and Caplan AJ (2012) UBR1 promotes protein kinase quality control and sensitizes cells to Hsp90 inhibition. Exp Cell Res 318, 53–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Hessa T, Sharma A, Mariappan M, Eshleman HD, Gutierrez E and Hegde RS (2011) Protein targeting and degradation are coupled for elimination of mislocalized proteins. Nature 475, 394–397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Rodrigo‐Brenni MC, Gutierrez E and Hegde RS (2014) Cytosolic quality control of mislocalized proteins requires RNF126 recruitment to Bag6. Mol Cell 55, 227–237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118. Fourie AM, Sambrook JF and Gething MJ (1994) Common and divergent peptide binding specificities of hsp70 molecular chaperones. J Biol Chem 269, 30470–30478. [PubMed] [Google Scholar]
  • 119. Rüdiger S, Buchberger A and Bukau B (1997) Interaction of Hsp70 chaperones with substrates. Nat Struct Biol 4, 342–349. [DOI] [PubMed] [Google Scholar]
  • 120. Higy M, Junne T and Spiess M (2004) Topogenesis of membrane proteins at the endoplasmic reticulum. Biochemistry 43, 12716–12722. [DOI] [PubMed] [Google Scholar]
  • 121. Lerch‐Bader M, Lundin C, Kim H, Nilsson I and von Heijne G (2008) Contribution of positively charged flanking residues to the insertion of transmembrane helices into the endoplasmic reticulum. Proc Natl Acad Sci USA 105, 4127–4132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Xu Y, Anderson DE and Ye Y (2016) The HECT domain ubiquitin ligase HUWE1 targets unassembled soluble proteins for degradation. Cell Discov 2, 16040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123. Sung MK, Porras‐Yakushi TR, Reitsma JM, Huber FM, Sweredoski MJ, Hoelz A, Hess S and Deshaies RJ (2016) A conserved quality‐control pathway that mediates degradation of unassembled ribosomal proteins. Elife 5, e19105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Katsogiannou M, Andrieu C, Baylot V, Baudot A, Dusetti NJ, Gayet O, Finetti P, Garrido C, Birnbaum D, Bertucci F et al (2014) The functional landscape of Hsp27 reveals new cellular processes such as DNA repair and alternative splicing and proposes novel anticancer targets. Mol Cell Proteomics 13, 3585–3601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Davies AE and Kaplan KB (2010) Hsp90‐Sgt1 and Skp1 target human Mis12 complexes to ensure efficient formation of kinetochore‐microtubule binding sites. J Cell Biol 189, 261–274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126. Spiechowicz M, Zylicz A, Bieganowski P, Kuznicki J and Filipek A (2007) Hsp70 is a new target of Sgt1–an interaction modulated by S100A6. Biochem Biophys Res Comm 357, 1148–1153. [DOI] [PubMed] [Google Scholar]
  • 127. Kaplan KB, Hyman AA and Sorger PK (1997) Regulating the yeast kinetochore by ubiquitin‐dependent degradation and Skp1p‐mediated phosphorylation. Cell 91, 491–500. [DOI] [PubMed] [Google Scholar]
  • 128. Catlett MG and Kaplan KB (2006) Sgt1p is a unique co‐chaperone that acts as a client adaptor to link Hsp90 to Skp1p. J Biol Chem 281, 33739–33748. [DOI] [PubMed] [Google Scholar]
  • 129. Kitagawa K, Skowyra D, Elledge SJ, Harper JW and Hieter P (1999) SGT1 encodes an essential component of the yeast kinetochore assembly pathway and a novel subunit of the SCF ubiquitin ligase complex. Mol Cell 4, 21–33. [DOI] [PubMed] [Google Scholar]
  • 130. Stuttmann J, Parker JE and Noel LD (2008) Staying in the fold: the SGT1/chaperone machinery in maintenance and evolution of leucine‐rich repeat proteins. Plant Signal Behav 3, 283–285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131. Willhoft O, Kerr R, Patel D, Zhang W, Al‐Jassar C, Daviter T, Millson SH, Thalassinos K and Vaughan CK (2017) The crystal structure of the Sgt1‐Skp1 complex: the link between Hsp90 and both SCF E3 ubiquitin ligases and kinetochores. Sci Rep 7, 41626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132. Fang NN, Ng AH, Measday V and Mayor T (2011) Hul5 HECT ubiquitin ligase plays a major role in the ubiquitylation and turnover of cytosolic misfolded proteins. Nat Cell Biol 13, 1344–1352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133. Fang NN, Chan GT, Zhu M, Comyn SA, Persaud A, Deshaies RJ, Rotin D, Gsponer J and Mayor T (2014) Rsp5/Nedd4 is the main ubiquitin ligase that targets cytosolic misfolded proteins following heat stress. Nat Cell Biol 16, 1227–1237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134. Crosas B, Hanna J, Kirkpatrick DS, Zhang DP, Tone Y, Hathaway NA, Buecker C, Leggett DS, Schmidt M, King RW et al (2006) Ubiquitin chains are remodeled at the proteasome by opposing ubiquitin ligase and deubiquitinating activities. Cell 127, 1401–1413. [DOI] [PubMed] [Google Scholar]
  • 135. Aviram S and Kornitzer D (2010) The ubiquitin ligase Hul5 promotes proteasomal processivity. Mol Cell Biol 30, 985–994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136. Fang NN and Mayor T (2012) Hul5 ubiquitin ligase: good riddance to bad proteins. Prion 6, 240–244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137. Koegl M, Hoppe T, Schlenker S, Ulrich HD, Mayer TU and Jentsch S (1999) A novel ubiquitination factor, E4, is involved in multiubiquitin chain assembly. Cell 96, 635–644. [DOI] [PubMed] [Google Scholar]
  • 138. Hoppe T (2010) Life and destruction: ubiquitin‐mediated proteolysis in aging and longevity. F1000 Biol Rep, 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139. Chu BW, Kovary KM, Guillaume J, Chen LC, Teruel MN and Wandless TJ (2013) The E3 ubiquitin ligase UBE3C enhances proteasome processivity by ubiquitinating partially proteolyzed substrates. J Biol Chem 288, 34575–34587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140. Kuo C‐L and Goldberg AL (2017) Ubiquitinated proteins promote the association of proteasomes with the deubiquitinating enzyme Usp14 and the ubiquitin ligase Ube3c. Proc Natl Acad Sci 114, E3404–E3413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141. Shahsavarani H, Sugiyama M, Kaneko Y, Chuenchit B and Harashima S (2012) Superior thermotolerance of Saccharomyces cerevisiae for efficient bioethanol fermentation can be achieved by overexpression of RSP5 ubiquitin ligase. Biotechnol Adv 30, 1289–1300. [DOI] [PubMed] [Google Scholar]
  • 142. Apaja PM, Xu H and Lukacs GL (2010) Quality control for unfolded proteins at the plasma membrane. J Cell Biol 191, 553–570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143. Hettema EH, Valdez‐Taubas J and Pelham HR (2004) Bsd2 binds the ubiquitin ligase Rsp5 and mediates the ubiquitination of transmembrane proteins. EMBO J 23, 1279–1288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144. Zhao Y, Macgurn JA, Liu M and Emr S (2013) The ART‐Rsp5 ubiquitin ligase network comprises a plasma membrane quality control system that protects yeast cells from proteotoxic stress. Elife 2, e00459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145. Shiga T, Yoshida N, Shimizu Y, Suzuki E, Sasaki T, Watanabe D and Takagi H (2014) Quality control of plasma membrane proteins by Saccharomyces cerevisiae Nedd4‐like ubiquitin ligase Rsp5p under environmental stress conditions. Eukaryot Cell 13, 1191–1199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146. Becuwe M, Herrador A, Haguenauer‐Tsapis R, Vincent O and Leon S (2012) Ubiquitin‐mediated regulation of endocytosis by proteins of the arrestin family. Biochem Res Int 2012, 242764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147. MacGurn JA, Hsu PC and Emr SD (2012) Ubiquitin and membrane protein turnover: from cradle to grave. Annu Rev Biochem 81, 231–259. [DOI] [PubMed] [Google Scholar]
  • 148. Duttler S, Pechmann S and Frydman J (2013) Principles of cotranslational ubiquitination and quality control at the ribosome. Mol Cell 50, 379–393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Maier T, Ferbitz L, Deuerling E and Ban N (2005) A cradle for new proteins: trigger factor at the ribosome. Curr Opin Struct Biol 15, 204–212. [DOI] [PubMed] [Google Scholar]
  • 150. Merz F, Boehringer D, Schaffitzel C, Preissler S, Hoffmann A, Maier T, Rutkowska A, Lozza J, Ban N, Bukau B et al (2008) Molecular mechanism and structure of Trigger Factor bound to the translating ribosome. EMBO J 27, 1622–1632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151. Lakshmipathy SK, Gupta R, Pinkert S, Etchells SA and Hartl FU (2010) Versatility of trigger factor interactions with ribosome‐nascent chain complexes. J Biol Chem 285, 27911–27923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152. Bengtson MH and Joazeiro CA (2010) Role of a ribosome‐associated E3 ubiquitin ligase in protein quality control. Nature 467, 470–473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153. Wang F, Canadeo LA and Huibregtse JM (2015) Ubiquitination of newly synthesized proteins at the ribosome. Biochimie 114, 127–133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154. Shao S, Brown A, Santhanam B and Hegde RS (2015) Structure and assembly pathway of the ribosome quality control complex. Mol Cell 57, 433–444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155. Yonashiro R, Tahara EB, Bengtson MH, Khokhrina M, Lorenz H, Chen KC, Kigoshi‐Tansho Y, Savas JN, Yates JR, Kay SA et al (2016) The Rqc2/Tae2 subunit of the ribosome‐associated quality control (RQC) complex marks ribosome‐stalled nascent polypeptide chains for aggregation. Elife 5, e11794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156. Verma R, Oania RS, Kolawa NJ and Deshaies RJ (2013) Cdc48/p97 promotes degradation of aberrant nascent polypeptides bound to the ribosome. Elife 2, e00308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157. Rape M, Hoppe T, Gorr I, Kalocay M, Richly H and Jentsch S (2001) Mobilization of processed, membrane‐tethered SPT23 transcription factor by CDC48 UFD1/NPL4, a ubiquitin‐selective chaperone. Cell 107, 667–677. [DOI] [PubMed] [Google Scholar]
  • 158. Bodnar NO and Rapoport TA (2017) Molecular mechanism of substrate processing by the Cdc48 ATPase complex. Cell 169, 722–735.e9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159. Brandman O, Stewart‐Ornstein J, Wong D, Larson A, Williams CC, Li GW, Zhou S, King D, Shen PS, Weibezahn J et al (2012) A ribosome‐bound quality control complex triggers degradation of nascent peptides and signals translation stress. Cell 151, 1042–1054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160. Defenouillere Q, Zhang E, Namane A, Mouaikel J, Jacquier A and Fromont‐Racine M (2016) Rqc1 and Ltn1 prevent c‐terminal alanine‐threonine tail (CAT‐tail)‐induced protein aggregation by efficient recruitment of Cdc48 on stalled 60S subunits. J Biol Chem 291, 12245–12253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161. Choe YJ, Park SH, Hassemer T, Korner R, Vincenz‐Donnelly L, Hayer‐Hartl M and Hartl FU (2016) Failure of RQC machinery causes protein aggregation and proteotoxic stress. Nature 531, 191–195. [DOI] [PubMed] [Google Scholar]
  • 162. Maurer MJ, Spear ED, Yu AT, Lee EJ, Shahzad S and Michaelis S (2016) Degradation signals for ubiquitin‐proteasome dependent cytosolic protein quality control (CytoQC) in yeast. G3 6, 1853–1866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163. Boban M and Foisner R (2016) Degradation‐mediated protein quality control at the inner nuclear membrane. Nucleus 7, 41–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164. Meng F, Na I, Kurgan L and Uversky V (2015) Compartmentalization and functionality of nuclear disorder: intrinsic disorder and protein‐protein interactions in intra‐nuclear compartments. Int J Mol Sci 17, 24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165. Roberts P (2002) Piecemeal microautophagy of nucleus in Saccharomyces cerevisiae . Mol Biol Cell 14, 129–141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166. Park Y‐E, Hayashi YK, Bonne G, Arimura T, Noguchi S, Nonaka I and Nishino I (2009) Autophagic degradation of nuclear components in mammalian cells. Autophagy 5, 795–804. [DOI] [PubMed] [Google Scholar]
  • 167. Fredrickson EK, Rosenbaum JC, Locke MN, Milac TI and Gardner RG (2011) Exposed hydrophobicity is a key determinant of nuclear quality control degradation. Mol Biol Cell 22, 2384–2395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168. Iwata A, Nagashima Y, Matsumoto L, Suzuki T, Yamanaka T, Date H, Deoka K, Nukina N and Tsuji S (2009) Intranuclear degradation of polyglutamine aggregates by the ubiquitin‐proteasome system. J Biol Chem 284, 9796–9803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169. Swanson R (2001) A conserved ubiquitin ligase of the nuclear envelope/endoplasmic reticulum that functions in both ER‐associated and Matalpha 2 repressor degradation. Genes Dev 15, 2660–2674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170. Bordallo J, Plemper RK, Finger A and Wolf DH (1998) Der3p/Hrd1p is required for endoplasmic reticulum‐associated degradation of misfolded lumenal and integral membrane proteins. Mol Biol Cell 9, 209–222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171. Bays NW, Gardner RG, Seelig LP, Joazeiro CA and Hampton RY (2001) Hrd1p/Der3p is a membrane‐anchored ubiquitin ligase required for ER‐associated degradation. Nat Cell Biol 3, 24–29. [DOI] [PubMed] [Google Scholar]
  • 172. Laba J, Steen A, Popken P, Chernova A, Poolman B and Veenhoff L (2015) Active nuclear import of membrane proteins revisited. Cells 4, 653–673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173. Boban M, Zargari A, Andreasson C, Heessen S, Thyberg J and Ljungdahl PO (2006) Asi1 is an inner nuclear membrane protein that restricts promoter access of two latent transcription factors. J Cell Biol 173, 695–707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174. Zargari A, Boban M, Heessen S, Andreasson C, Thyberg J and Ljungdahl PO (2007) Inner nuclear membrane proteins Asi1, Asi2, and Asi3 function in concert to maintain the latent properties of transcription factors Stp1 and Stp2. J Biol Chem 282, 594–605. [DOI] [PubMed] [Google Scholar]
  • 175. Foresti O, Rodriguez‐Vaello V, Funaya C and Carvalho P (2014) Quality control of inner nuclear membrane proteins by the Asi complex. Science 346, 751–755. [DOI] [PubMed] [Google Scholar]
  • 176. Khmelinskii A, Blaszczak E, Pantazopoulou M, Fischer B, Omnus DJ, Le Dez G, Brossard A, Gunnarsson A, Barry JD, Meurer M et al (2014) Protein quality control at the inner nuclear membrane. Nature 516, 410–413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Deng M and Hochstrasser M (2006) Spatially regulated ubiquitin ligation by an ER/nuclear membrane ligase. Nature 443, 827–831. [DOI] [PubMed] [Google Scholar]
  • 178. Shiber A, Breuer W, Brandeis M and Ravid T (2013) Ubiquitin conjugation triggers misfolded protein sequestration into quality control foci when Hsp70 chaperone levels are limiting. Mol Biol Cell 24, 2076–2087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179. Ravid T, Kreft SG and Hochstrasser M (2006) Membrane and soluble substrates of the Doa10 ubiquitin ligase are degraded by distinct pathways. EMBO J 25, 533–543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180. Metzger MB, Maurer MJ, Dancy BM and Michaelis S (2008) Degradation of a cytosolic protein requires endoplasmic reticulum‐associated degradation machinery. J Biol Chem 283, 32302–32316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181. Guo L, Giasson Benoit I, Glavis‐Bloom A, Brewer Michael D, Shorter J, Gitler Aaron D and Yang X (2014) A cellular system that degrades misfolded proteins and protects against neurodegeneration. Mol Cell 55, 15–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182. Venkatesh S and Workman JL (2015) Histone exchange, chromatin structure and the regulation of transcription. Nat Rev Mol Cell Biol 16, 178–189. [DOI] [PubMed] [Google Scholar]
  • 183. Commerford S, Carsten A and Cronkite E (1982) Histone turnover within nonproliferating cells. Proc Natl Acad Sci 79, 1163–1165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184. Singh RK, Kabbaj M‐HM, Paik J and Gunjan A (2009) Histone levels are regulated by phosphorylation and ubiquitylation‐dependent proteolysis. Nat Cell Biol 11, 925–933. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185. Singh RK, Liang D, Gajjalaiahvari UR, Kabbaj M‐HM, Paik J and Gunjan A (2010) Excess histone levels mediate cytotoxicity via multiple mechanisms. Cell Cycle 9, 4236–4244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186. Singh RK, Gonzalez M, Kabbaj M‐HM and Gunjan A (2012) Novel E3 ubiquitin ligases that regulate histone protein levels in the budding yeast Saccharomyces cerevisiae . PLoS ONE 7, e36295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187. Goldberg AD, Banaszynski LA, Noh K‐M, Lewis PW, Elsaesser SJ, Stadler S, Dewell S, Law M, Guo X, Li X et al (2010) Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell 140, 678–691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188. Hondele M and Ladurner AG (2011) The chaperone–histone partnership: for the greater good of histone traffic and chromatin plasticity. Curr Opin Struct Biol 21, 698–708. [DOI] [PubMed] [Google Scholar]
  • 189. Hondele M, Stuwe T, Hassler M, Halbach F, Bowman A, Zhang ET, Nijmeijer B, Kotthoff C, Rybin V, Amlacher S et al (2013) Structural basis of histone H2A–H2B recognition by the essential chaperone FACT. Nature 499, 111–114. [DOI] [PubMed] [Google Scholar]
  • 190. Mattiroli F, D'Arcy S and Luger K (2015) The right place at the right time: chaperoning core histone variants. EMBO Rep 16, 1454–1466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191. Deyter GMR and Biggins S (2014) The FACT complex interacts with the E3 ubiquitin ligase Psh1 to prevent ectopic localization of CENP‐A. Genes Dev 28, 1815–1826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192. Han J, Zhang H, Zhang H, Wang Z, Zhou H and Zhang Z (2013) A Cul4 E3 ubiquitin ligase regulates histone hand‐off during nucleosome assembly. Cell 155, 817–829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193. Keller JN, Huang FF and Markesbery WR (2000) Decreased levels of proteasome activity and proteasome expression in aging spinal cord. Neuroscience 98, 149–156. [DOI] [PubMed] [Google Scholar]
  • 194. Cuervo AM (2004) Impaired degradation of mutant ‐synuclein by chaperone‐mediated autophagy. Science 305, 1292–1295. [DOI] [PubMed] [Google Scholar]
  • 195. Vilchez D, Morantte I, Liu Z, Douglas PM, Merkwirth C, Rodrigues AP, Manning G and Dillin A (2012) RPN‐6 determines C. elegans longevity under proteotoxic stress conditions. Nature 489, 263. [DOI] [PubMed] [Google Scholar]
  • 196. Kahn NW, Rea SL, Moyle S, Kell A and Johnson TE (2008) Proteasomal dysfunction activates the transcription factor SKN‐1 and produces a selective oxidative‐stress response in Caenorhabditis elegans . Biochem J 409, 205–213. [DOI] [PubMed] [Google Scholar]
  • 197. Ciechanover A and Kwon YT (2017) Protein quality control by molecular chaperones in neurodegeneration. Front Neurosci 11, 185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198. Terman A (1995) The effect of age on formation and elimination of autophagic vacuoles in mouse hepatocytes. Gerontology 41, 319–326. [DOI] [PubMed] [Google Scholar]
  • 199. Donati A, Cavallini G, Paradiso C, Vittorini S, Pollera M, Gori Z and Bergamini E (2001) Age‐related changes in the regulation of autophagic proteolysis in rat isolated hepatocytes. J Gerontol A Biol Sci Med Sci 56, B288–B293. [DOI] [PubMed] [Google Scholar]
  • 200. Walther Dirk M, Kasturi P, Zheng M, Pinkert S, Vecchi G, Ciryam P, Morimoto Richard I, Dobson Christopher M, Vendruscolo M, Mann M et al (2015) Widespread proteome remodeling and aggregation in aging C. elegans . Cell 161, 919–932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201. Kaushik S and Cuervo AM (2015) Proteostasis and aging. Nat Med 21, 1406–1415. [DOI] [PubMed] [Google Scholar]
  • 202. Morimoto RI and Santoro MG (1998) Stress‐inducible responses and heat shock proteins: new pharmacologic targets for cytoprotection. Nat Biotechnol 16, 833–838. [DOI] [PubMed] [Google Scholar]
  • 203. Taylor JP, Hardy J and Fischbeck KH (2002) Toxic proteins in neurodegenerative disease. Science 296, 1991–1995. [DOI] [PubMed] [Google Scholar]
  • 204. Broersen K, Rousseau F and Schymkowitz J (2010) The culprit behind amyloid beta peptide related neurotoxicity in Alzheimer's disease: oligomer size or conformation? Alzheimer's Res Ther 2, 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205. Sulistio YA and Heese K (2016) The ubiquitin‐proteasome system and molecular chaperone deregulation in Alzheimer's disease. Mol Neurobiol 53, 905–931. [DOI] [PubMed] [Google Scholar]
  • 206. Shin Y, Klucken J, Patterson C, Hyman BT and McLean PJ (2005) The co‐chaperone carboxyl terminus of Hsp70‐interacting protein (CHIP) mediates α‐synuclein degradation decisions between proteasomal and lysosomal pathways. J Biol Chem 280, 23727–23734. [DOI] [PubMed] [Google Scholar]
  • 207. Kumar P, Pradhan K, Karunya R, Ambasta RK and Querfurth HW (2011) Cross‐functional E3 ligases Parkin and C‐terminus Hsp70‐interacting protein in neurodegenerative disorders. J Neurochem 120, 350–370. [DOI] [PubMed] [Google Scholar]
  • 208. Qian S‐B, McDonough H, Boellmann F, Cyr DM and Patterson C (2006) CHIP‐mediated stress recovery by sequential ubiquitination of substrates and Hsp70. Nature 440, 551–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209. Koren J 3rd, Jinwal UK, Lee DC, Jones JR, Shults CL, Johnson AG, Anderson LJ and Dickey CA (2009) Chaperone signalling complexes in Alzheimer's disease. J Cell Mol Med 13, 619–630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210. Urushitani M, Kurisu J, Tateno M, Hatakeyama S, Nakayama K‐I, Kato S and Takahashi R (2004) CHIP promotes proteasomal degradation of familial ALS‐linked mutant SOD1 by ubiquitinating Hsp/Hsc70. J Neurochem 90, 231–244. [DOI] [PubMed] [Google Scholar]
  • 211. Jana NR, Dikshit P, Goswami A, Kotliarova S, Murata S, Tanaka K and Nukina N (2005) Co‐chaperone CHIP associates with expanded polyglutamine protein and promotes their degradation by proteasomes. J Biol Chem 280, 11635–11640. [DOI] [PubMed] [Google Scholar]
  • 212. Williams AJ, Knutson TM, Colomer Gould VF and Paulson HL (2009) In vivo suppression of polyglutamine neurotoxicity by C‐terminus of Hsp70‐interacting protein (CHIP) supports an aggregation model of pathogenesis. Neurobiol Dis 33, 342–353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213. Garrido C, Brunet M, Didelot C, Zermati Y, Schmitt E and Kroemer G (2006) Heat shock proteins 27 and 70: anti‐apoptotic proteins with tumorigenic properties. Cell Cycle 5, 2592–2601. [DOI] [PubMed] [Google Scholar]
  • 214. Jego G, Hazoumé A, Seigneuric R and Garrido C (2013) Targeting heat shock proteins in cancer. Cancer Lett 332, 275–285. [DOI] [PubMed] [Google Scholar]
  • 215. Barrott JJ and Haystead TA (2013) Hsp90, an unlikely ally in the war on cancer. FEBS J 280, 1381–1396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216. Nahleh Z, Tfayli A, Najm A, El Sayed A and Nahle Z (2012) Heat shock proteins in cancer: targeting the ‘chaperones’. Future Med Chem 4, 927–935. [DOI] [PubMed] [Google Scholar]
  • 217. Wang X, Chen M, Zhou J and Zhang X (2014) HSP27, 70 and 90, anti‐apoptotic proteins, in clinical cancer therapy (Review). Int J Oncol 45, 18–30. [DOI] [PubMed] [Google Scholar]
  • 218. Fang NN, Zhu M, Rose A, Wu KP and Mayor T (2016) Deubiquitinase activity is required for the proteasomal degradation of misfolded cytosolic proteins upon heat‐stress. Nat Commun 7, 12907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219. Finley D (2011) Misfolded proteins driven to destruction by Hul5. Nat Cell Biol 13, 1290–1292. [DOI] [PubMed] [Google Scholar]
  • 220. Zhang M, Boter M, Li K, Kadota Y, Panaretou B, Prodromou C, Shirasu K and Pearl LH (2008) Structural and functional coupling of Hsp90‐ and Sgt1‐centred multi‐protein complexes. EMBO J 27, 2789–2798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221. Gallagher PS, Clowes Candadai SV and Gardner RG (2014) The requirement for Cdc48/p97 in nuclear protein quality control degradation depends on the substrate and correlates with substrate insolubility. J Cell Sci 127, 1980–1991. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Febs Letters are provided here courtesy of Wiley

RESOURCES